Download as pdf or txt
Download as pdf or txt
You are on page 1of 639

Handbook of 

Aggregation-­Induced Emission
Handbook of Aggregation-­Induced Emission

Volume 1 Tutorial Lectures and Mechanism Studies

Edited by

Youhong Tang
Flinders University
Adelaide, Australia

Ben Zhong Tang


The Chinese University of Hong Kong
Shenzhen, China
This edition first published 2022
© 2022 John Wiley & Sons Ltd. Published 2022 by John Wiley & Sons Ltd.

All rights reserved. No part of this publication may be reproduced, stored in a retrieval system, or transmitted, in
any form or by any means, electronic, mechanical, photocopying, recording or otherwise, except as permitted by
law. Advice on how to obtain permission to reuse material from this title is available at http://www.wiley.com/go/
permissions.

The right of Youhong Tang and Ben Zhong Tang to be identified as the author(s) of the editorial material in this
work has been asserted in accordance with law.

Registered Office(s)
John Wiley & Sons, Inc., 111 River Street, Hoboken, NJ 07030, USA
John Wiley & Sons Ltd, The Atrium, Southern Gate, Chichester, West Sussex, PO19 8SQ, UK

Editorial Office
The Atrium, Southern Gate, Chichester, West Sussex, PO19 8SQ, UK

For details of our global editorial offices, customer services, and more information about Wiley products visit us at
www.wiley.com.

Wiley also publishes its books in a variety of electronic formats and by print-­on-­demand. Some content that appears
in standard print versions of this book may not be available in other formats.

Limit of Liability/Disclaimer of Warranty


In view of ongoing research, equipment modifications, changes in governmental regulations, and the constant flow
of information relating to the use of experimental reagents, equipment, and devices, the reader is urged to review
and evaluate the information provided in the package insert or instructions for each chemical, piece of equipment,
reagent, or device for, among other things, any changes in the instructions or indication of usage and for added
warnings and precautions. While the publisher and authors have used their best efforts in preparing this work, they
make no representations or warranties with respect to the accuracy or completeness of the contents of this work
and specifically disclaim all warranties, including without limitation any implied warranties of merchantability
or fitness for a particular purpose. No warranty may be created or extended by sales representatives, written sales
materials or promotional statements for this work. The fact that an organization, website, or product is referred
to in this work as a citation and/or potential source of further information does not mean that the publisher and
authors endorse the information or services the organization, website, or product may provide or recommendations
it may make. This work is sold with the understanding that the publisher is not engaged in rendering professional
services. The advice and strategies contained herein may not be suitable for your situation. You should consult with
a specialist where appropriate. Further, readers should be aware that websites listed in this work may have changed
or disappeared between when this work was written and when it is read. Neither the publisher nor authors shall
be liable for any loss of profit or any other commercial damages, including but not limited to special, incidental,
consequential, or other damages.

Library of Congress Cataloging-­in-­Publication Data applied for


Hardback (Volume 1): 9781119642916

Cover Design: Wiley


Cover Image: © nadla/iStock/Getty Images Plus

Set in 9.5/12.5pt STIXTwoText by Straive, Pondicherry, India

10  9  8  7  6  5  4  3  2  1
Pay our great respects to all the researchers who have contributed and are contributing to the study
of Aggregation Induced Emission (AIE).
vii

Contents

List of Contributors  xv
Preface to Handbook of Aggregation-Induced Emission  xxi
Preface to Volume 1: Fundamentals  xxiii

1 The Mechanistic Understanding of the Importance of Molecular Motions


to Aggregation-­induced Emission  1
Junkai Liu and Ben Zhong Tang
1.1 ­Introduction  1
1.2 ­Restriction of Intramolecular Motion  2
1.2.1 Restriction of Intramolecular Rotation  3
1.2.2 Restriction of Intramolecular Vibration  4
1.2.3 Ultrafast Insights into Tetraphenylethylene Derivatives  6
1.2.4 Theoretical Insights into Restriction of Intramolecular Motion  8
1.3 ­Restricted Access to Conical Intersection  12
1.4 ­Restriction of Access to the Dark State  14
1.5 ­Suppression of Kasha’s Rule  15
1.6 ­Through Space Conjugation  17
1.6.1 Clusterization-­Triggered Emission  18
1.6.2 Polymerization-­induced Emission  19
1.6.3 Excited-­state Through-­space Conjugation  19
1.7 ­Perspective  21
­References  23

2 Understanding the AIE Mechanism at the Molecular Level  27


Xiaoyan Zheng and Qian Peng
2.1 ­Introduction  27
2.2 ­Theoretical Methods  28
2.2.1 Radiative and Nonradiative Rate Constants  28
2.2.2 Computational Details  29
2.3 ­Revealed AIE Mechanism  31
2.3.1 Rotating Vibrations of Intramolecular Aromatic Ring  31
2.3.2 Stretching Vibrations of Bonds  33
2.3.3 Bending Vibration of Bonds  34
2.3.4 Flipping Vibrations of Molecular Skeletons  35
2.3.5 Twisting Vibration of Molecular Skeletons  36
2.4 ­Visualize Calculated Parameters in Experiments  37
2.4.1 Stokes Shift vs Reorganization Energy  37
viii Contents

2.4.2 Resonance Raman Spectroscopy (RSS) vs Reorganization Energy  38


2.4.3 Isotope Effect vs DRE  40
2.4.4 Linear Relationship between Fluorescence Intensity and Amorphous
Aggregate Size  42
2.4.5 Pressure-­induced Enhanced Emission (PIEE)  44
2.5 ­Molecular Design Based on AIE Mechanism  45
2.6 ­Summary and Outlook  46
Acknowledgments  48
­References  48

3 Aggregation-­induced Emission from the Restriction of Double Bond Rotation at


the Excited State  55
Ming Hu and Yan-­Song Zheng
3.1 ­Introduction  55
3.2 ­AIE Phenomena and Applications from RDBR Mechanism  58
3.2.1 Evolvement and Development of AIE Mechanisms  58
3.2.2 Investigation of RDBR AIE Mechanism by E/Z isomerization  64
3.2.3 Investigating of RDBR AIE Mechanism by Immobilization of TPE Propeller-­like
Conformation  69
3.2.4 Research of Theoretical Calculation on RDBR  78
3.2.5 Other AIEgens Involving RBDR Process  84
3.3 ­Conclusions  93
­References  94

4 The Expansion of AIE Thought: From Single Molecule to Molecular Uniting  99


Qiuyan Liao, Qianqian Li, and Zhen Li
4.1 ­Aggregation-­Induced Emission  99
4.2 ­Photoluminescence Materials Based on Molecular Set  101
4.3 ­Mechanoluminescence Materials Based on Molecular Set  106
4.3.1 Mechanoluminescence Materials with Fluorescence Emission  106
4.3.2 Mechanoluminescence Materials with Mechanical Induced Dual-­ or Tri-­color
Emission  115
4.3.3 Quantitative Research of Mechanoluminescence Property  121
4.4 ­Mechanochromism Materials  122
4.4.1 Mechanochromism Materials Based on Polymorphs  122
4.4.2 Mechanochromism Materials Based on Excimer Emission  125
4.4.3 Other Kinds of Mechanochromism Materials  128
4.5 ­Room Temperature Phosphorescence Materials Based on Molecular Uniting  131
4.5.1 Room Temperature Phosphorescence Materials with Aromatics  131
4.5.2 Room Temperature Phosphorescence Materials with Simple or Nonaromatic
Structure  140
4.5.3 Room Temperature Phosphorescence Materials with Multiple Emission  142
4.5.4 Photoinduced Room Temperature Phosphorescence Materials  144
4.6 ­Conclusion and Perspectives  147
­References  147

5 Clusterization-­Triggered Emission  153


Haoke Zhang and Ben Zhong Tang
5.1 ­Introduction  153
5.2 ­Pure n-­Electron Systems  156
5.3 ­Pure π-­Electron Systems  160
Contents ix

5.4 (­ n, π)-­Electrons Systems  164


5.5 ­Other Systems  166
5.6 ­Summary  167
­References  168

6 Crystallization-­induced Emission Enhancement  177


Yong Qiang Dong, Yingying Liu, Mengyang Liu, Qian Wang, and Kang Wang
6.1 ­Introduction  177
6.2 ­Tetraphenylethylene Derivatives  178
6.3 ­CIEE Active Luminogens with Bulky Conjugation Core  183
6.3.1 Dibenzofulvene (DBF) Derivatives (Chart 6.2)  183
6.3.2 9-­([1,1′-­Biphenyl]-­4-­ylphenylmethylene)-­9H-­xanthene  185
6.3.3 Dicyanomethylenated Acridones  186
6.3.4 Bis(diarylmethylene)dihydroanthracene [31]  187
6.4 ­Other High-­contrast CIEE Luminogens  190
6.4.1 4-­Dimethylamino-­2-­Benzylidene Malonic Acid Dimethyl Ester  190
6.4.2 Diphenyl Maleimide Derivatives [33]  191
6.4.3 3,4-­Bisthienylmaleic Anhydride [34]  192
6.4.4 Boron-­containing CIEE Luminogens  193
6.5 ­Potential Applications  196
6.5.1 Volatile Organic Compounds (VOCs) Sensor  196
6.5.2 OLED  196
6.5.3 High-­density Data Storage  197
6.5.4 Mechanochromic (MC) Luminescent Sensor  198
6.6 ­Summary and Perspective  198
­References  198

7 Surface-­fixation Induced Emission  203


Yohei Ishida and Shinsuke Takagi
7.1 ­Introduction  203
7.2 ­What Happened to the Characteristics of Molecules
on the Clay Mineral Nanosheets  205
7.3 ­Clay–Molecular Complexes  206
7.4 ­Absorption Spectra of Clay–Molecular Complexes  207
7.5 ­Emission Enhancement Phenomenon in Clay–Molecular Complexes: S-­FIE  208
7.6 ­Mechanism of Surface-­Fixation Induced Emission  211
7.7 ­Summary and Outlook  214
Acknowledgment  215
­References  215

8 Aggregation-­induced Delayed Fluorescence  221


Yan Fu, Hao Chen, Zujin Zhao, and Ben Zhong Tang
8.1 ­Introduction  221
8.2 ­Novel Aggregation-­induced Delayed Fluorescence Luminogens  222
8.3 ­Conclusion and Outlook  247
­References  247

9 Homogeneous Systems to Induce Emission of AIEgens  251


Kenta Kokado and Kazuki Sada
9.1 ­Introduction  251
9.2 ­Homogeneous Solution  252
x Contents

9.2.1 Complexation with Anions  253


9.2.2 Complexation with Cations  254
9.2.3 Inclusion Complexes  256
9.2.4 Adhesion on Macromolecules  257
9.2.5 Steric Hindrance  258
9.2.6 Covalent Linkage  259
9.3 ­Liquid  260
9.4 ­Gels and Network Polymers  261
9.4.1 Chemically Crosslinked Gels  261
9.4.2 Physically Crosslinked Gels  262
9.5 ­Crystalline Materials  264
9.6 ­Outlook and Future Perspectives  266
­References  266

10 Hetero-­aggregation-­induced Tunable Emission (HAITE) Through Cocrystal


Strategy  273
Yinjuan Huang and Qichun Zhang
10.1 ­Introduction  273
10.2 ­Interactions Within Organic Cocrystals  274
10.3 ­Preparation of Organic Cocrystals  275
10.4 ­Molecular Stacking Modes Within Organic Cocrystals  276
10.5 ­Characterization of Organic Cocrystals  277
10.6 ­HAITE Through Cocrystal Strategy  277
10.6.1 HAITE with Tunable Color and Enhanced Emission  278
10.6.1.1 Insignificant Changed Intensity but Tuned Color  278
10.6.1.2 Enhanced Emission and Tuned Color  287
10.6.2 HAITE with Increased PLQY but Intrinsic Color  291
10.6.3 HAITE: Thermally Activated Delayed Fluorescence  297
10.6.4 HAITE-­phosphorescence  300
10.7 ­Summary and Outlook  302
­References  304

11 Anti-­Kasha Emission from Organic Aggregates  311


Wenbin Huang and Zikai He
11.1 ­Introduction  311
11.2 ­Anti-­Kasha Emission from Aromatic Carbonyl
Compounds in Aggregates  312
11.3 ­Anti-­Kasha Emission from Azulene Compounds in Aggregate  322
11.4 ­Anti-­Kasha Emission from Other Unconventional
Aromatic Compounds in Aggregates  324
11.5 ­Conclusions  327
­References  327

12 Aggregation-enhanced Emission: From Flexible to Rigid Cores  333


Harnimarta Deol, Gurpreet Singh, Vandana Bhalla, and Manoj Kumar
12.1 ­Introduction  333
12.2 ­Freely Moving Rotors-­induced Emission Enhancement  334
Contents xi

12.3 ­ uest-­induced Emission Enhancement  344


G
12.4 ­Conclusion  366
Acknowledgment   367
­References  367

13 Room-­temperature Phosphorescence of Pure Organics  371


Tianwen Zhu, Zihao Zhao, Tianjia Yang, and Wang Zhang Yuan
13.1 ­Introduction  371
13.2 ­Fundamental Mechanism in Organic Phosphorescence  372
13.2.1 Photophysical Process for Phosphorescence  372
13.2.2 Theoretical Study on Phosphorescent Process  373
13.3 ­Recent Progress in Organic RTP Materials  375
13.3.1 Crystallization-­induced RTP  375
13.3.1.1 Heavy Atom Effect  376
13.3.1.2 Molecular Interaction  380
13.3.1.3 H-­aggregation  380
13.3.2 Doping in Rigid Matrix-­induced RTP  382
13.3.2.1 Host–Guest System  385
13.3.2.2 Doping in Polymer Matrix  387
13.3.3 Clustering-­triggered RTP  389
13.3.3.1 Natural Products  389
13.3.3.2 Synthetic Compounds  394
13.3.4 Other Systems  399
13.3.4.1 Amorphous Organics  399
13.3.4.2 Organic Framework  399
13.3.4.3 Supramolecular Organics  402
13.3.4.4 Hybrid Perovskites  403
13.3.5 Applications  405
13.4 ­Conclusions and Perspectives  405
­References  407

14 A Global Potential Energy Surface Approach to the Photophysics of AIEgens: The Role


of Conical Intersections  411
Rachel Crespo-­Otero and Lluís Blancafort
14.1 ­Introduction  411
14.2 ­Methodological Aspects  412
14.2.1 Intramolecular Restriction Models and the FGR-­based Approach  412
14.2.2 A PES-­based Description of Photochemical Mechanisms  412
14.2.3 Computational Approaches for Excited States  416
14.2.3.1 Electronic Structure Methods for Excited States  416
14.2.3.2 Dynamics Simulations in the Context of AIE  420
14.2.4 Methods for Large Systems  420
14.3 ­CI-­centered Global PES for AIEgens  424
14.3.1 Double-­bond Torsion  424
14.3.2 Double-­bond Torsion vs Cyclization in TPE Derivatives  428
14.3.3 Excited-­state Intramolecular Proton Transfer (ESIPT) Compounds  431
14.3.4 Ring Puckering  432
14.3.5 Bond Stretching  435
xii Contents

14.3.6 A View of AIE Based on the RACI Model and the Global PES  436
14.4 ­Crystallization-­induced Phosphorescence  436
14.5 ­Effect of Intermolecular and Intramolecular Interactions
on the Photophysics of AIEgens  437
14.5.1 Excitonic Effects in AIE  437
14.5.2 Effect of Intramolecular and Intermolecular Interactions on Emission Color  439
14.6 ­New Challenges  439
14.6.1 The Role of Dark States in AIE  439
14.6.2 Pressure-­induced Emission Enhancement  440
14.6.3 AIE in Transition Metal (TM) Compounds  442
14.7 ­Conclusions and Outlook  443
­References  444

15 Multicomponent Reactions as Synthetic Design Tools of AIE and Emission


Solvatochromic Quinoxalines  455
Lukas Biesen and Thomas J. J. Müller
15.1 ­Introduction  455
15.2 ­Synthetic Approaches to Quinoxalines via Multicomponent Reactions and One-­Pot
Processes  456
15.3 ­Photophysical Properties and Emission Solvatochromicity of Quinoxalines  462
15.4 ­AIE Characteristics and Effects of Quinoxalines  468
15.5 ­Conclusion  476
­Acknowledgments  476
­References  476

16 Aggregation-­induced Emission Luminogens with Both High-­luminescence Efficiency


and Charge Mobility  485
Ying Yu, Zheng Zhao, and Ben Zhong Tang
16.1 ­Introduction  485
16.2 ­p-­Type OSCs  487
16.3 ­n-­Type OSCs  495
16.4 ­Ambipolar OSCs  500
16.5 ­Conclusion and Perspective  505
­References  505

17 Morphology Modulation of Aggregation-­induced Emission: From Thermodynamic


Self-­assembly to Kinetic Controlling  509
Kaizhi Gu, Chenxu Yan, Zhiqian Guo, and Wei-­Hong Zhu
17.1 ­Introduction  509
17.2 ­Aggregation Modulation of AIE Bioprobes via Hydrophilicity Improvement  511
17.2.1 Molecular Modification  511
17.2.2 Polymerization with Hydrophilic Matrix  515
17.3 ­Thermodynamic Self-­assembly of AIE Materials  519
17.4 ­Morphology Tuning of AIE Nanoaggregates  519
17.5 ­Kinetic-­driven Preparation of AIE NPs  523
17.6 ­Conclusion and Outlook  527
­References  527
Contents xiii

18 AIE-­active Polymer  531


Rong Hu, Anjun Qin, and Ben Zhong Tang
18.1 ­Introduction  531
18.2 ­Photophysical Properties  532
18.2.1 Quantum Yield  532
18.2.2 Photosensitization  536
18.2.3 Two-­photon Absorption and Emission  538
18.2.4 Circularly Polarized Luminescence  540
18.3 ­Applications  541
18.3.1 Chem-­sensor  541
18.3.2 Bioimaging  543
18.3.3 Therapy Applications  546
18.4 ­Conclusion and Perspective  549
Acknowledgments  550
­References  550

19 Liquid-­crystalline AIEgens: Materials and Applications  555


Kyohei Hisano, Supattra Panthai, and Osamu Tsutsumi
19.1 ­Introduction  555
19.2 ­Materials: Molecular Design  556
19.2.1 Discotic LC AIEgen  556
19.2.2 Calamitic LC AIEgens  561
19.2.3 Polymeric LC AIEgens  566
19.3 ­Applications of LC AIEgens  567
19.3.1 Linearly Polarized Luminescence  567
19.3.2 Circularly Polarized Luminescence  568
19.4 ­Conclusion  571
­References  571

20 Push–Pull AIEgens  575


Andrea Nitti and Dario Pasini
20.1 ­Introduction  575
20.2 ­Basic Concept of Molecular Design  576
20.2.1 Photophysical Excited States in Aggregates  576
20.2.2 Fundamental Molecular Design to Achieve Push–Pull AIEgens  579
20.3 ­Push–Pull AIEgens from Rotor Structure  581
20.3.1 Double Bond Stator  582
20.3.2 Point-­restricted Rotors from Atoms or Functional Groups  584
20.3.3 Aromatic Rotors  587
20.4 ­Push–Pull AIEgens from ACQ Chromophores  589
20.4.1 BT-­based AIEgens  589
20.4.2 Cyanine and DCM-­based AIEgens  594
20.4.3 QM-­based AIEgens  595
20.4.4 DPP-­based AIEgens  597
20.4.5 Rylene-­based AIEgens  599
20.5 ­Concluding Remarks  602
­References  602

Index  609
xv

List of Contributors

Vandana Bhalla Yong Qiang Dong


Department of Chemistry Beijing Key Laboratory of Energy Conversion
UGC Sponsored Centre for Advanced and Storage Materials
Studies-­II College of Chemistry
Guru Nanak Dev University, Amritsar Beijing Normal University
PB, India Beijing, China

Lukas Biesen Yan Fu


Institut für Organische Chemie und State Key Laboratory of Luminescent
Makromolekulare Chemie Materials and Devices
Heinrich-­Heine-­Universität Düsseldorf Guangdong Provincial Key Laboratory of
Düsseldorf, Germany Luminescence from Molecular Aggregates
South China University of Technology
Lluís Blancafort Guangzhou, China
Institut de Química Computacional i Catàlisi
and Departament de Química Kaizhi Gu
Facultat de Ciències, Universitat de Shanghai Key Laboratory of Functional
Girona, Spain Materials Chemistry
Key Laboratory for Advanced Materials and
Hao Chen Institute of Fine Chemicals
State Key Laboratory of Luminescent Joint International Research Laboratory
Materials and Devices of Precision Chemistry and Molecular
Guangdong Provincial Key Laboratory of Engineering
Luminescence from Molecular Aggregates Feringa Nobel Prize Scientist Joint
South China University of Technology Research Center
Guangzhou, China School of Chemistry and Molecular
Engineering, East China University of Science
Rachel Crespo-­Otero & Technology
Department of Chemistry Shanghai, China
Queen Mary University of London
London, United Kingdom Zhiqian Guo
Shanghai Key Laboratory of Functional
Harnimarta Deol Materials Chemistry
Department of Chemistry Key Laboratory for Advanced Materials and
UGC Sponsored Centre for Advanced Institute of Fine Chemicals
Studies-­II Joint International Research Laboratory
Guru Nanak Dev University, Amritsar of Precision Chemistry and Molecular
PB, India Engineering
xvi List of Contributors

Feringa Nobel Prize Scientist Joint Yohei Ishida


Research Center Division of Materials Science and Engineering
School of Chemistry and Molecular Faculty of Engineering
Engineering Hokkaido University
East China University of Science & Sapporo, Hokkaido, Japan
Technology
Shanghai, China Kenta Kokado
Research Institute for Electronic Science
Zikai He Hokkaido University
School of Science Sapporo, Hokkaido, Japan
Harbin Institute of Technology
Shenzhen, HIT Campus of University Town Graduate School of Environmental Science
Shenzhen, China Hokkaido University
Sapporo, Hokkaido, Japan
Kyohei Hisano PRESTO, JST, Chiyodaku, Tokyo, Japan
Department of Applied Chemistry
College of Life Sciences
Ritsumeikan University Manoj Kumar
Kusatsu, Japan Department of Chemistry, UGC Sponsored
Centre for Advanced Studies-­II
Ming Hu Guru Nanak Dev University, Amritsar
Key Laboratory of Material Chemistry for PB, India
Energy Conversion and Storage
Ministry of Education Qianqian Li
School of Chemistry and Chemical Department of Chemistry
Engineering Wuhan University, Wuhan, China
Huazhong University of Science and
Technology Zhen Li
Wuhan, China Department of Chemistry
Wuhan University, Wuhan, China
Rong Hu
Institute of Molecular Aggregation Science
State Key Laboratory of Luminescent
Tianjin University, Tianjin, China
Materials and Devices, Guangdong Provincial
Key Laboratory of Luminescence from
Molecular Qiuyan Liao
Aggregates, Center for Aggregation-Induced Department of Chemistry
Emission, South China University of Wuhan University
Technology, Guangzhou, China Wuhan, China

Wenbin Huang Junkai Liu


School of Science Department of Chemistry
Harbin Institute of Technology Hong Kong Branch of Chinese National
Shenzhen, HIT Campus of University Town Engineering Research Center for Tissue
Shenzhen, China Restoration and Reconstruction and Institute
for Advanced Study
Yinjuan Huang The Hong Kong University of Science and
School of Materials Science and Engineering Technology
Nanyang Technological University Clear Water Bay, Kowloon
Singapore, Singapore Hong Kong, China
List of Contributors xvii

Mengyang Liu Kazuki Sada


Beijing Key Laboratory of Energy Conversion Faculty of Science, Hokkaido University
and Storage Materials Sapporo, Hokkaido, Japan
College of Chemistry
Graduate School of Chemical Sciences and
Beijing Normal University
Engineering
Beijing, China
Hokkaido University
Sapporo, Hokkaido, Japans
Yingying Liu
Beijing Key Laboratory of Energy Conversion Gurpreet Singh
and Storage Materials Department of Chemistry
College of Chemistry UGC Sponsored Centre for Advanced Studies-­II
Beijing Normal University Guru Nanak Dev University, Amritsar
Beijing, China PB, India

Thomas J. J. Müller Shinsuke Takagi


Institut für Organische Chemie und Department of Applied Chemistry
Makromolekulare Chemie Graduate Course of Urban Environmental
Heinrich-­Heine-­Universität Düsseldorf Sciences
Düsseldorf, Germany Tokyo Metropolitan University
Minami-­Oshawa, Hachiohji, Tokyo, Japan
Andrea Nitti
Department of Chemistry Ben Zhong Tang
University of Pavia and INSTM Research Unit State Key Laboratory of Luminescent
Pavia, Italy Materials and Devices
Guangdong Provincial Key Laboratory of
Supattra Panthai Luminescence from Molecular Aggregates
Department of Applied Chemistry Center for Aggregation-­Induced Emission
College of Life Sciences South China University of Technology
Ritsumeikan University Guangzhou, China
Kusatsu, Japan
Department of Chemistry, Hong Kong Branch
of Chinese National Engineering Research
Dario Pasini Centre for Tissue Restoration and
Department of Chemistry Reconstruction, Institute for Advanced Study,
University of Pavia and INSTM Research Unit and Department of Chemical and Biological
Pavia, Italy Engineering, The Hong Kong University of
Science & Technology, Clear Water Bay,
Qian Peng Kowloon, Hong Kong, China
School of Chemical Sciences
University of Chinese Academy of Sciences Shenzhen Institute of Aggregate Science
Beijing, China and Technology
School of Science and Engineering
Anjun Qin The Chinese University of Hong Kong
State Key Laboratory of Luminescent Shenzhen, Guangdong, China
Materials and Devices
Guangdong Provincial Key Laboratory of Osamu Tsutsumi
Luminescence from Molecular Aggregates Department of Applied Chemistry
Center for Aggregation-­Induced Emission College of Life Sciences
South China University of Technology Ritsumeikan University
Guangzhou, China Kusatsu, Japan
xviii List of Contributors

Kang Wang Haoke Zhang


Beijing Key Laboratory of Energy Conversion Department of Chemistry, The Hong Kong
and Storage Materials University of Science and Technology, Clear
College of Chemistry Water Bay, Kowloon, Hong Kong, China
Beijing Normal University
Beijing, China Qichun Zhang
Department of Chemistry, Department of
Qian Wang Materials Science and Engineering, City
Beijing Key Laboratory of Energy Conversion University of Hong Kong, Hong Kong, China
and Storage Materials
College of Chemistry Zheng Zhao
Beijing Normal University Shenzhen Institute of Aggregate Science and
Beijing, China Technology
School of Science and Engineering
Chenxu Yan The Chinese University of Hong Kong
Shanghai Key Laboratory of Functional Shenzhen, Guangdong, China
Materials Chemistry
Key Laboratory for Advanced Materials and Zihao Zhao
Institute of Fine Chemicals School of Chemistry and Chemical
Joint International Research Laboratory Engineering
of Precision Chemistry and Molecular Shanghai Jiao Tong University
Engineering Shanghai, China
Feringa Nobel Prize Scientist Joint
Research Center Zujin Zhao
School of Chemistry and Molecular State Key Laboratory of Luminescent
Engineering Materials and Devices
East China University of Science & Guangdong Provincial Key Laboratory of
Technology, Shanghai, China Luminescence from Molecular Aggregates
South China University of Technology
Tianjia Yang Guangzhou, China
School of Chemistry and Chemical
Engineering Xiaoyan Zheng
Shanghai Jiao Tong University Beijing Key Laboratory of Photoelectronic/
Shanghai, China Electrophotonic Conversion Materials
Key Laboratory of Cluster Science of Ministry
Ying Yu of Education
Shenzhen Institute of Aggregate Science School of Chemistry and Chemical
and Technology Engineering
School of Science and Engineering Beijing Institute of Technology, Beijing, China
The Chinese University of Hong Kong
Shenzhen, Guangdong, China Yan-­Song Zheng
Key Laboratory of Material Chemistry for
Energy Conversion and Storage
Wang Zhang Yuan Ministry of Education
School of Chemistry and Chemical School of Chemistry and Chemical Engineering
Engineering Huazhong University of Science and
Shanghai Jiao Tong University Technology
Shanghai, China Wuhan, China
List of Contributors xix

Tianwen Zhu Joint International Research Laboratory


School of Chemistry and Chemical of Precision Chemistry and Molecular
Engineering Engineering
Shanghai Jiao Tong University Feringa Nobel Prize Scientist Joint
Shanghai, China Research Center
School of Chemistry and Molecular
Wei-­Hong Zhu Engineering
Shanghai Key Laboratory of Functional East China University of Science &
Materials Chemistry Technology, Shanghai, China
Key Laboratory for Advanced Materials and
Institute of Fine Chemicals
xxi

Preface to Handbook of Aggregation-Induced Emission

Aggregation-­induced emission (AIE) describes a photophysical phenomenon in which molecular


aggregate exhibits stronger emission than its single-­molecule counterpart. AIE research provides
an excellent platform to explore aggregate behaviours since it emphasizes the new properties
endowed by molecular aggregates beyond the microscopic molecular level. The 21-­year develop-
ment of AIE research has witnessed the great achievements in this area, including mechanistic
understanding, structure-­property relationship, and high-­tech applications. The achievements of
AIE research demonstrate that many behaviours and functions that are absent in molecular spe-
cies can be founded in molecular aggregates. Since the properties of molecular ensembles are
closer to those of macroscopic materials, AIE research is of both scientific value and technological
importance for real applications.
Due to its great value and huge potential, AIE research is attracting more and more attention
from chemistry, physics, materials science, medical and biomedical science and shows promising
applications in these fields. Noteworthy, as a general platform to study aggregate science, AIE
keeps integrating with other research fields involving materials, biology, medicine, energy and
environment, and injects new vitality into these fields. When more researchers are working on this
field, more breakthroughs in both fundamental research and application are envisioned in
the future.
This handbook is an essential reading for scientists and engineers who are designing optoelec-
tronic materials and chemical/biomedical sensors. It is also a valuable reference book to academic
researchers in materials science, physical and synthetic organic chemistry as well as physicists and
biological chemists.
xxiii

Preface to Volume 1: Fundamentals

Volume 1 surveys the breakthrough of aggregation-­induced emission (AIE) research area, focusing
mainly on the fundamentals of various branched areas. In particularly, this volume presents the
new properties that molecular ensembles bring to molecules and highlight the role of molecular
aggregates in endowing or improving the performance of organic materials. The branches of AIE
include crystallization-­induced emission (CIE), room temperature phosphorescence (RTP),
aggregation-­induced delayed fluorescence (AIDF), anti-­Kasha transition (AKT), clusterization-­
triggered emission (CTE), through space interaction (TSI), mechanoluminescence (ML), circularly
polarized and others. We specifically focus on the new properties of materials endowed by molecu-
lar aggregates beyond the microscopic molecular level. We hope this volume will inspire more
research into molecular ensembles at/beyond meso level and lead to the significant progresses in
material ­science, biological science, etc.

Youhong Tang
Flinders University, Australia

Ben Zhong Tang


The Chinese University of Hong Kong, Shenzhen, China
1

The Mechanistic Understanding of the Importance of Molecular


Motions to Aggregation-­induced Emission
Junkai Liu1 and Ben Zhong Tang1,2,3
1
 Department of Chemistry, Hong Kong Branch of Chinese National Engineering Research Center for Tissue Restoration and
Reconstruction and Institute for Advanced Study, The Hong Kong University of Science and Technology, Clear Water Bay, Kowloon, Hong
Kong, China
2
 Shenzhen Institute of Aggregate Science and Technology, School of Science and Engineering, The Chinese University of Hong Kong,
Shenzhen, 2001 Longxiang Boulevard, Longgang District, Shenzhen City, Guangdong 518172, China
3
 State Key Laboratory of Luminescent Materials and Devices, Guangdong Provincial Key Laboratory of Luminescence from Molecular
Aggregates, Center for Aggregation-Induced Emission, South China University of Technology, Guangzhou, China

1.1 ­Introduction

Molecular motions drive most natural processes, ranging from the formation and annihilation of
astronomical objects to the metabolism of microbes. All kinds of intra-­or intermolecular interac-
tions induce molecular motions in specific forms, through which the energy is generated and
transformed. Mechanistic understanding and manipulation of molecular motions can lead to the
effective design of smart molecular systems to achieve programmed tasks, as well as dynamic
­control of multiple processes. Indeed, by employing the light as a stimulus, researchers have
­developed a variety of functional molecular machines (e.g. molecular motors, molecular switches,
molecular shuttles, and supramolecular assemblies) to achieve effective control of catalysis activ-
ity and chirality transfer in various practical applications, in which the dynamic manipulation of
diverse molecular motions is critical [1].
For organic luminescence processes, molecular motions also extensively affect the photophysical
behaviors for organic luminophores and govern the excited-­state decay pathways, including vibra-
tional relaxation, internal conversion, intersystem crossing, kinetic quenching, etc.[2]. Upon photo-
excitation, the coupled nuclear and electronic motions will drive the excited molecules to evolve
through radiative and nonradiative pathways [3]. However, traditional photophysical research usu-
ally focuses on the highly rigid and conjugated molecules and investigates the targets in the gas
phase or solution, in which the importance of intramolecular motions to the luminescence is less
considered, and the variation of molecular motions in the solid state is often ignored [4]. In 2001,
the discovery and proposal of aggregation-­induced emission (AIE) triggered the research on molec-
ular motions and photophysical studies in the aggregate or the solid state [5]. The luminogens with
the AIE property (AIEgens) often show different scales of molecular motions in different phases.
Vigorous intramolecular motions in the solution can nonradiatively dissipate excited-­state energy
and always result in weak light emission, whereas such motions can be restricted in the solid state
due to the environmental constraints so that the emission can be strongly enhanced [6].

Handbook of Aggregation-Induced Emission: Volume 1 Tutorial Lectures and Mechanism Studies, First Edition.
Edited by Youhong Tang and Ben Zhong Tang.
© 2022 John Wiley & Sons Ltd. Published 2022 by John Wiley & Sons Ltd.
2 1  The Mechanistic Understanding of the Importance of Molecular Motions to Aggregation-­induced Emission

The restriction of intramolecular motion (RIM) has been concluded as the most widely
accepted and applicable mechanism for the AIE phenomenon through numerous experimen-
tal and theoretical exploration, and the strong electron-­vibration coupling (EVC) between the
excited and the ground state has been revealed to be the quantum origin of the weak emission
of AIEgens in the solution [3, 7]. Recently, the connotation of RIM has been further clarified
with deeper mechanistic insights into various novel AIEgens. The nonadiabatic excited-­state
molecular dynamics  [8] and theoretical model based on multiple electronic states  [9] have
been employed to elucidate the dynamic evolution of excited states through molecular motions
of tetraphenylethylene (TPE) derivatives and heteroatom-­containing AIEgens. Meanwhile,
the aromaticity reversal in the excited state  [10] and the anti-­Kasha emission from higher
excited states  [11] have provided another proof for the enhanced emission efficiency in the
solid state for AIEgens. Apart from the critical effect on the nonradiative decay, molecular
motions have also been verified to promote the electronic transition process  [12, 13]. The
excited-­state through-­space conjugation facilitated by excited-­state molecular motions has
been proved to contribute to the enhanced visible light emission of a series of nonconjugated
AIEgens [13].
In this chapter, we will take a journey of mechanistic studies for AIE from the general RIM to
mechanisms developed recently and propose the perspective on the further exploration in
the future.

1.2 ­Restriction of Intramolecular Motion

When we look into the AIE phenomenon, two essential questions arise: why do the AIEgens show
none or weak emission in the solution? Why does the aggregation brighten the light emission of
AIEgens? Enthusiastic efforts have been devoted to deciphering the AIE mechanism. Several pos-
sible mechanisms have been proposed, including intramolecular conformation planarization,
J-­aggregate formation, E/Z isomerization, and excited-­state intramolecular proton transfer, but
they are only applicable for specific molecular systems [6c]. A general working mechanism for AIE
is highly desired for fundamental understanding and material designing.
Upon absorption of photons, the excited molecule will decay through radiative and nonradiative
paths or the photochemical process [2]. Hence, AIEgens in the solution may mainly undergo non-
radiative decay or photochemical reaction to dissipate the majority of energy. Once aggregation
occurs, the nonradiative decay paths can be blocked, or the radiative paths can be facilitated so that
the emission can be enhanced. As the structure determines the property, typical AIEgens own flex-
ible structures, containing multiple rotor or vibrator substituents, like hexaphenylsilole (HPS) and
TPE, which endow them with high flexibility and potential to consume energy through intensive
intramolecular motion, whereas multiple intermolecular interactions exist in the aggregate state,
which can serve as constraints to molecular motions detrimental to the emission [6]. Continuous
experimental exploration has been made through manipulation of the molecular motions by
adjusting environmental factors, including hydrophobic interaction, temperature, viscosity, pres-
sure, host–guest interaction, and intramolecular constraints such as substituent steric hindrance,
ring closing, conjugation effect, and so on [14]. With persistent effort, Tang and coworkers have
proposed the restriction of intramolecular rotation (RIR), the restriction of intramolecular vibra-
tion (RIV), and, finally, concluded that the restriction of intramolecular motion (RIM) was the
general working principle for AIEgens [15].
1.2  ­Restriction of Intramolecular Motio 3

1.2.1  Restriction of Intramolecular Rotation


Investigation of the effect from molecular motions on luminescence processes has drawn increas-
ing interests. The less conjugated structures and intrinsic steric congestions are responsible for the
high conformational flexibility of AIEgens.
Taking HPS as the model AIEgen [14a] as shown in Figure 1.1, when HPS is fully dissolved in
the THF solution to form the isolated molecular species, it shows almost no emission. With increas-
ing water gradually into the dilute solution and keeping the same concentration, the HPS mole-
cules tend to aggregate due to the hydrophobicity. When the water fraction reaches 60%, notable
aggregate forms and the quantum yield of the whole system starts to increase. Further increasing
the water fraction, the light emission is continuously enhanced due to the further aggregation.
According to general physics, any form of motion can consume energy. When we look into the
structure of HPS analyzed from the X-­ray diffraction, it can be found that six phenyl rings are
attached to the central silole ring through single bonds with large dihedral angles, which allow free
torsion or twisting motions, indicating that motions of phenyl rings are responsible for the energy
dissipation in the solution state. With increasing the solvent viscosity by adding the glycerol into
the solution or decreasing the temperature of the solution from room temperature to −196 °C, the
emission of HPS-­2 can be gradually enhanced due to the constraints coming from the highly vis-
cous solvent or rigidified solvent environment at the low temperature. Furthermore, introducing

(a) (b) (c)


25 10 000
10 000

20
PL peak intensity (a.u.)

PL peak intensity (a.u.)


Quantum yield (%)

1 000 Si
15 Si
1 000

10 HPS HPS-2
100
5 100

0 10
0 30 60 90 0 30 60 90 50 0 –50 –100 –150 –200
Water fraction (vol%) Glycerol fraction (%) Temperature (°C)

(d) (e) 250

ΦF (%)
Si Si Si 200
HPS-3 83
PL intensity (a.u.)

HPS-4 3.2
150 HPS-5 0.55
HPS-3 HPS-4 HPS-5
100

50

0
400 500 600 700
Wavelength (nm)

Figure 1.1  (a) Fluorescence quantum yield of HPS in the acetone/water mixtures with different water
fraction. (b) Photoluminescence peak intensity of HPS in the glycerol/methanol mixtures with different
glycerol fraction [HPS] = 10−5 M. (c) Photoluminescence peak intensity of HPS-­2 in the 1,4-­dioxane solution
at different temperatures. [HPS-­2] = 10−5 M. Source: Adapted from Ref. [14a] with permission from American
Chemical Society. (d) Chemical structures and fluorescence photos. (e) Photoluminescence spectra of HPS
3−5 in the acetone solution (10−5 M). Source: Adapted from Ref. [14c] with permission from American
Chemical Society.
4 1  The Mechanistic Understanding of the Importance of Molecular Motions to Aggregation-­induced Emission

internal steric hindrance shows a similar effect by attaching the isopropyl groups onto the
2,3,4,5-­positions of the central silole as presented in Figure 1.1d and e [14c]. Crowdedness induced
by the bulky groups hampers the torsion of the phenyl rings, then the nonradiative decay process
can be restricted, and the emission can be restored.
TPE is another prototype of AIEgens with a similar intrinsic structure as HPS. Four phenyl rings
are connected with the central double bond in a highly twisted conformation. As presented in
Figure  1.3a, upon excitation, four phenyl rings can drastically twist against the double bond.
Meanwhile, the C=C double bond will be weakened due to the photoexcitation, which endows it
with strong twisting ability to reorganize the conformation. All these intramolecular motions
result in the undetectable emission for TPE in the dilute solution. However, its emission can be
dramatically boosted once the aggregation occurs. Further locking the adjacent phenyl rings
through chemical bonds or attaching bulky groups onto the phenyl rings of TPE, its emission can
be enhanced even in the solution [16]. All these experiments of controlling the internal or external
constraints to the intramolecular motions have proved that the RIR mechanism is responsible for
the AIE effect of the rotor-­based AIEgens.

1.2.2  Restriction of Intramolecular Vibration


Apart from the intramolecular rotation, the intramolecular vibration in the luminogens-­lacking
rotors can also cause emission quenching in the dilute solution. Bu et al. have found that two
coumarin derivatives fused with five-­membered and seven-­membered alkyl rings, respectively,
show totally different photophysical behaviors, despite their similar conjugation degrees in the
ground state, as indicated by the similar ultraviolet (UV)–vis absorption spectra presented in
Figure 1.2 [10a].
The compound CD-­5 with a five-­membered ring is much more rigid and can show strong light
emission in the dilute solution, but its emission is weakened in the solid, whereas the CD-­7 is
almost nonemissive in the dilute solution, but it can show enhanced light emission in the solid
state. The calculated conformations of these two derivatives have revealed that the CD-­7  with
seven-­membered ring owns much higher vibrational flexibility in the excited state. The central π-­
conjugated plane of CD-­7 will suffer the vibrational motion with a dihedral angle of around 18°, so
its emission can be easily quenched in the dilute solution. But such intramolecular vibration can
be efficiently restricted in the solid state, which finally leads to the AIE phenomenon. The mecha-
nism for such AIEgens-­lacking rotors has been concluded as the RIV.
However, it remains unclear what the driving force is for the vigorous molecular motions on the
excited state. Zhao et al. have recently systematically investigated the structure–property relation-
ship of a series of annulene-­based AIEgens and figured out the driving force of the strong vibra-
tions in the excited state  [10]. The parent cyclooctatetrathiophene (COTh) contains an
eight-­membered ring fused by four thiophene rings and takes a noncoplanar and saddle-­like con-
formation as presented in Figure 1.3b. Such a highly twisted structure makes COTh nonaromatic
in the ground state. The COTh shows typical AIE behavior that it can only emit negligible light in
the dilute solution but shows enhanced green emission in the aggregate and solid film. The opti-
mized structure of COTh in the ground state is similar to its single crystal structure, whereas it can
relax to a transition state with a planar conformation, and finally decay through fast nonradiative
pathways to the ground state. Through the calculation of the nucleus-­independent chemical shift
(NICS) and anisotropy of the induced current density (ACID) to characterizing the aromaticity of
the structures, it has been revealed that the aromaticity reversal from the ground state to the excited
1.2  ­Restriction of Intramolecular Motio 5

(a) (b)

CD-7 CD-7
CD-5 CD-5
Absorbance

300 350 400 450 500 PL intensity 400 450 500 550 600
Wavelength (nm) Wavelength (nm)
(c)

S1

~ 18°
N
CD-7 N
3
4 CD-5
N5 3 4
N5
AIE 12 ACQ
76
O 12
6 O
7
O8 O
O 8 O

S0

Gas phase Solid phase Gas phase

Figure 1.2  (a) UV–vis absorption and (b) photoluminescence (PL) spectra of coumarin derivative-­7 (CD-­7)
and coumarin derivative-­5 (CD-­5) in CH2Cl2. Concentration: 10−5 M; excitation wavelength: 411 nm.
(c) Optimized structures at the S0,min and S1,min of CD-­7 in the gas and solid phases and CD-­5 in the gas
phase at time dependent (TD) M062X/6-­31G(d) level. Source: Adapted from Ref. [10a] with permission from
John Wiley and Sons.

state drives the conformational change through intense up–down vibration. But when the aggrega-
tion occurs, the vibration can be effectively restricted due to the environmental hindrance.
Meanwhile, the energy of the planar transition state has been elevated, and the pathways to the
transition state are not energetically accessible, which leads to the major radiative decay and pro-
moted luminescence efficiency.
From the abovementioned cases, it can be concluded that RIR and RIV are the main cause of the
AIE properties of propeller-­shaped and shell-­shaped AIEgens. Hence, the RIM can be unified as
the general working principle for the AIE effect, and RIM can serve as the basic design principle
for the development of AIE materials.
6 1  The Mechanistic Understanding of the Importance of Molecular Motions to Aggregation-­induced Emission

(a)

Aggregation

Restriction of
intramolecular
rotation (RIR)

TPE

Active rotation Restriction of


RIR in aggregate
in solution intramolecular motion
Highly emissive
(RIM)
Non emissive RIM = RIR + RIV

(b)
Aggregation
S S
Restriction of
S intramolecular
S vibration (RIV)
COTh

Dynamic vibrations
in solution RIV in aggregate
Highly emissive
Non emissive

Figure 1.3  Schematic illustration of RIM mechanism. (a) Restriction of intramolecular rotation.


(b) Restriction of intramolecular vibration. Source: Adapted from Ref. [6a] with permission from American
Chemical Society and Ref. [10b] with permission from Springer Nature.

1.2.3  Ultrafast Insights into Tetraphenylethylene Derivatives


Molecular motions significantly affect the decay behaviors for AIEgens. Certain excited-­state
dynamics require further clarification. Tang and coworkers have developed detailed understanding
of excited-­state molecular motions of TPE derivatives and further elucidated the RIM mechanism
through the lens of ultrafast spectroscopy [16]. As presented in Figure 1.4, six TPE derivatives with
different internal constraints and structural rigidity have been prepared and studied through the
first-­principle calculation based on (time-­dependent) density function ­theory ((TD)DFT) methods.
Different structural rigidities may lead to different freedom of motion and structural variation
during the excited-­state decay. Optimized conformations in the S0 state and the S1 state show that
all the studied TPE derivatives undergo the elongation of the ethylenic double bond and form the
quasi-­double bond after excitation, but the single bonds connecting the peripheral phenyl rings
with the central double bond have been shortened. These two types of bond change originate from
the weakening of the double bond and the electron density flowing from the double bond to the
surrounding single bonds. Meanwhile, the bond variation leads to the twisting of the quasi-­double
bond and the phenyl rotors.
Comparison between the S1 and S0 conformations shows that the parent TPE has encountered
the largest structural reorganization with the torsion angle of the quasi-­double bond of 56.80° and
1.2  ­Restriction of Intramolecular Motio 7

0
(a) (b) 140

Quasi C = C bond twisting (°)


10

kcal/mol
S1 min FC*
120 20
FC* 100
30
S0 min 80
40
60
50
40
60
20 S1 min
TPE TPE-2 0
70
80
(60,24,60)
90
Excited state 0 10 20 30 40 50 60 70 80 90

Energy
Phenyl torsion (°)
0
S0 min 10

C = C bond twisting (°)


90 20
TPE-3 TPE-4
(Q 80 0
30
ua 70 Ground state
si) 10 40
C 60 20 50
=C 50 (8,50,0) 30
bo 40 40 60
n °)
d
tw 30 50 n( 70
ist 20 60 rsio
in 70 yl to 80
g en
(°) 10 80 Ph
0 90 90
0 10 20 30 40 50 60 70 80 90
TPE-5 TPE-6 Phenyl torsion (°)

(c)
Sn τ1 τ2 τ3 τ4
τ1 < 0.4 ps
S*
c a b a H H
τ2 τ2 τ2
11.2 ps 1.2 ps < 0.4 ps

hv
FL
b H H
τ3
18.9 ps
IM S0
τ4
S0 > 1 ms
c H3C H

TPE derivatives PC + R20R22 (S0)

Figure 1.4  (a) Chemical structures of TPE derivatives with increasing rigidity. (b) Potential energy surfaces
(PES) and projections of PES of the S1 and the S0 state of TPE. The MEPs on the S1 and the S0 state are
marked as black and white dash line, respectively. (c) Schematic illustration of ultrafast excited-­state decay
processes of TPE derivatives and dominant molecules with different timescales. Source: Adapted from
Ref. [16] with permission from The Royal Society of Chemistry.

the torsion angle of the phenyl ring of 29.04°, whereas with decreasing the motion freedom, deriv-
atives show decreased structural changes. The TPE-­6 with double vicinal locking of the phenyl
rings only shows the double bond elongation of 0.03  Å and less than 9° twisting of the quasi-­
double bond. Hence, the twisting of the quasi-­double bond and the torsion of the phenyl rings have
been chosen as the motion coordinates to scan the potential energy surfaces (PES) of the S1 and S0
states. The calculated minimum energy path (MEP) in the ground state of the parent TPE is cou-
pled by the phenyl-­ring torsion from 50° to 90° and the double bond twisting of less than 9° with
energy change of less than 7 kcal/mol. There is a transition state connecting two minimum points
on the ground-­state PES with the highly twisted quasi double bond. However, the MEP on the
excited-­state PES is dominantly associated with the twisting of the quasi-­double bond of 50° due to
the significant elongation of the original double bond of 0.12 Å, accompanied by the phenyl ring’s
torsion of less than 25°. The PES around the S1,min is shallow with the energy of 60 kcal/mol, which
8 1  The Mechanistic Understanding of the Importance of Molecular Motions to Aggregation-­induced Emission

is rather close to the transition state on the ground-­state PES, which will lead to fast decay from
S1,min to the ground state.
The ultrafast transient absorption spectra have further revealed the excited-­state decay dynamics
through the coupling of the double bond twisting and phenyl ring’s torsion. For the parent TPE,
upon absorption of pump light, the excited species will decay from Sn states to the S1 state, through
internal conversion within sub-­picosecond, mainly associated with elongation and twisting of the
C=C double bond, which can be signalized by the stimulated emission at 1.3 ps. Then, from 1.3 to
3.79 ps, the excitons populated on the emissive state revert back through twisting the quasi-­double
bond. After 3.79 ps, the photocyclization intermediate forms associated with the phenyl ring’s tor-
sion with an ultralong lifetime of 159 seconds. With increasing the intramolecular steric congests,
TPE-­2 shows similar excited-­state dynamics with the parent TPE, whereas TPE-­3 emits light effi-
ciently within sub-­picosecond even in the dilute solution. The transient emission at 480  nm of
TPE-­3 within 1.19 ps has confirmed the ultrafast population on the emissive S1,min, which is con-
sistent with the decrease of the peak at 477 nm on the ultrafast transient absorption spectra due to
the stimulated emission between 0.63 and 1.2  ps. The population period takes a longer time as
compared to the parent TPE, which indicates that the steric hindrance on the phenyl rings can
slow down the twisting of the quasi-­double bond and contribute to the stabilization of the emissive
state. Further attaching the alkyl tethers between the geminal or vicinal phenyl rings will largely
reduce freedom of motion of the phenyl rings and accelerate the formation of the photocyclization
intermediate.
Hence, it can be concluded that TPE derivatives will undergo an ultrafast vibrational relaxation
and populate on the S1,min after excitation through the coupled molecular motions of elongation
and twisting of the C=C double bond as well as torsion of phenyl rings within the picosecond
timescale, after which the photocyclization intermediate will be formed. The dominant time com-
ponent in this relaxation varies, depending on the conformational rigidity. The increased steric
hindrance or structural locking on the parent TPE will hamper the twisting of the quasi-­double
bond and stabilize the emissive state to generate efficient light, as in the case of TPE-­3, or directly
minimize the motions involving the double bond and the phenyl rings and make the photocycliza-
tion intermediate formed on the sub-­picosecond timescale as in the case of TPE-­6. Thus, motions
involving the C=C double bond and the formation of the photocyclization intermediate are the
dominant pathways for the excited-­state deactivation of TPE derivatives in the solution.

1.2.4  Theoretical Insights into Restriction of Intramolecular Motion


The experimental investigation has proved that the intramolecular motions are responsible for the
weak emission of AIEgens in the solution, and the aggregate environment can restrict such molec-
ular motions and enhance the light emission. What kind of molecular motions can govern the
excited-­state decay pathways for AIEgens? How do the molecular motions dissipate the excited-­
state energy? Theoretical researchers have developed the proper modes for radiative decay rates
and nonradiative decay rates based on the thermal vibration correlation function (TVCF) formal-
ism, and precisely calculated the quantum yields in both solution state and aggregate state and
finally revealed the quantum origins of the RIM mechanism [7].
Studies of the transition process between two electronic states relate to the quantum dynamics.
The fluorescence and phosphorescence processes usually occur in the timescale from nanoseconds
to milliseconds, while the internal conversion and intersystem crossing often occur in picoseconds
or a shorter timescale. However, the current excited-­state dynamic simulation can only deal with
the process occurring within picoseconds, which is much faster than the luminescence process.
1.2  ­Restriction of Intramolecular Motio 9

Hence, theoreticians usually evaluate the luminescence quantum efficiency by employing the
decay rate formalism based on the Fermi golden rule [3]. The luminescence quantum yield ΦF can
be defined as ΦF = kr + knr, in which kr stands for the radiative decay rate and knr stands for the
nonradiative decay rate. The kr can be calculated with integration of multiple theories even for
polyatomic molecular systems. The problem to evaluate the quantum yield is to precisely calculate
the nonradiative decay rates, which contain more factors related to complicated molecular motion
modes [3].
Traditional target molecules for the photophysical research always own highly rigid π-­electron
structure with large conjugation, which makes the effect of molecular motions less notable,
whereas most AIEgens possess highly flexible molecular structures, allowing the high freedom of
motion in the isolated state, which facilitates the nonradiative pathways to a large extent through
diverse decay modes. Peng et al. have proved that calculation of internal conversion rates consider-
ing the Duschinsky rotation effect (DRE) leads to more accurate quantitative evaluation of the
nonradiative decay process for the 1,2,3,4-­tetraphenylbutadiene (TPBD) with AIE property since
DRE describes the coupling among the multiple modes of molecular motions  [7e]. Such mode
mixing also significantly contributes to the nonradiative dissipation of excited-­state energy as well
as the intrinsic multiple molecular motions. On the other hand, the evaluation of the temperature-­
dependent luminescence behaviors of TPBD considering DRE produces more reliable results that
are much closer to the experimental observation [7e].
Furthermore, for the fluorescence process, the nonradiative decay pathways are governed by the
nonadiabatic coupling (NAC), which can be divided into nonadiabatic electronic coupling (NAEC)
and EVC [7e]. The NAEC deciphers the electronic part that contributes to the nonradiative decay,
and the EVC relates to the interaction between the electronic and the nuclear motions. The reor-
ganization energy (RE) can be applied to evaluate the intensity of EVC and can offer detailed
information on the structure–property relationship [17]. Through investigating NAEC and EVC,
researchers have revealed the quantum origins of the energy-­consuming mechanism from molecu-
lar motions. Taking HPS as the research model, its fluorescence quantum yield is only 0.30% in the
dilute solution, whereas it can reach 78% in the solid film, which is 260 times as high as that in the
solution. Zhang et al. have calculated the electronic structures of HPS in both isolated state and
crystal state, using the combined quantum mechanics and molecular mechanics (QM/MM)
method to explore the detailed working mechanism of HPS from quantum view [7a].
During the vibrational relaxation in the excited state, large conformational reorganization
occurs, which mainly relates to the torsion of phenyl rotors on the 2,5-­positions of the silole ring
due to their high mobility in the isolated state. However, the optimized HPS in the crystal shows
only minor conformational reorganization during the relaxation with almost no change in the
dihedral angles of phenyl rings, which proves the RIM features in the solid state. Further, NAC
analysis reveals that the distribution and total values of NAEC have no notable change before and
after the aggregation occurs (Figure 1.5), indicating that the NAEC is insensitive to the aggregate
environment. Nevertheless, the EVC analysis shows that many low-­frequency motion modes
(ω < 200 cm−1) of HPS exist in the isolated state, and they are mainly recognized as the torsion of
phenyl rotors. The RE of these low-­frequency modes is 197 meV, which occupies 40% of the total
RE of the whole system. Afterwards, the RE of low-­frequency modes significantly decreases to
84 meV in the crystal, occupying 21% of the total value. Due to the large suppression of EVC, the
calculated knr decreases with four orders in magnitude from the isolated state to the crystal state.
Hence, the emission efficiency of HPS can be boosted in the crystal state due to the restricted EVC.
Theoretically, the large EVC of most AIEgens is the main cause for their low emission efficiency
in the isolated state, which originates from the nature of the conformational flexibility and high
10 1  The Mechanistic Understanding of the Importance of Molecular Motions to Aggregation-­induced Emission

(a) (b)
Gas phase
6
Solid phase

Si 4

Rkk (cm–1)
3
HPS
2 2
2
15° 6°
1
15° 1°
5 5 0
0 20 40 60 80 100 120 140 160 180 200

Gas Solid Normal mode index

(c) (d)
100 100
Gas phase Solid phase
90 90
Reorganization energy (meV)

Reorganization energy (meV)

80 80
70 70
60 60
50 50
40 40
30 30
20 20
10 10
0 0
0 500 1000 1500 2000 2500 3000 3500 0 500 1000 1500 2000 2500 3000 3500
Normal mode (cm–1) Normal mode (cm–1)

Figure 1.5  (a) Chemical structures and the overlap of the S1,min conformation and the S0,min conformation
in the gas and the solid state of HPS. (b) Diagonal elements Rkk of the nonadiabatic electronic coupling
matrix versus normal mode index in the gas and the solid state. (c) Reorganization energy versus normal
mode wave numbers in the gas state and (d) the solid state. Source: Adapted from Ref. [7a] with permission
from American Chemical Society.

freedom of molecular motions of the AIEgens. Tang and coworkers have studied the substituent
effect on the AIE behaviors of a series of stilbene derivatives, as presented in Figure 1.6, and veri-
fied the influence from freedom of molecular motions on the excited-­state properties [17b]. Three
methyl groups can be attached to 2,4,5-­positions or 2,4,6-­positions of each phenyl rings of the
stilbene, which leads to opposite photophysical behaviors. The 2,4,5-­TMe-­DPE shows aggregation-­
caused quenching effect, whereas the 2,4,6-­TMe-­DPE shows the typical AIE behavior. The first
question that arises is which factors make the 2,4,5-­substituted derivative highly emissive with ΦF
of 13.4% but the 2,4,6-­substituted counterpart almost nonemissive with ΦF as 0.6% in the dilute
solution. The quantum simulations in the ground state show that 2,4,5-­TMe-­DPE owns a more
planar conjugation plane, while 2,4,6-­TMe-­DPE is more twisted. The calculated rotation barrier
shows that 2,4,6-­TMe-­DPE undergoes lower barrier to rotate along the coordinate of twisting, but
it suffers higher barrier to move along the coordinate of planarization, which reveals that methyl
groups in 4,6-­positions can produce strong steric hindrance to prevent the coplanarization between
phenyl rings and the double bond, but such steric hindrance can cause preliminary twisting of
phenyl rings and reduce the conjugation degree, which leads to the easier torsional motions of the
1.2  ­Restriction of Intramolecular Motio 11

1500 50
(a) αʹ (b) 16.7% (c)
αʹ 2.5%
α α

Reorganization energy (cm–1)


1200 40
80.8%

Normal mode index


0.00
S1 S1 900 30
S0 S0 Bond length
0.125
Bond angle
600 20
2,4,5-TMe-DPE 2,4,6-TMe-DPE Dihedral angle
Total λ =12 384 cm–1 0.250
300 10
∣α∣ 21.6° 41.6°
0.375
ΦF,soln 13.4% 0.6%
0 0
0 500 1000 1500 2000 0 10 20 30 40 50
Normal mode wavenumber (cm–1) Normal mode index
80 50 0.500
1000
(d) (e) (f)
26.3%
2,4,5-TMe-DPE 4.5% 0.625
Rotation barrier (kJ/mol)

40
Reorganization energy (cm–1)

60 800 69.2%
2,4,6-TMe-DPE

Normal mode index


600 30 0.750
40 Bond length
Bond angle
400 20 0.875
Dihedral angle
20
Total λ = 2883 cm–1
200 10 1.00

0
0 0
0 10 20 30 40 50 60 70 0 500 1000 1500 2000 0 10 20 30 40 50
α(°) Normal mode wavenumber (cm–1)
Normal mode index

Figure 1.6  (a) Chemical structures, the overlaps of the S1,min conformation and the S0,min conformation in
the gas phase, and fluorescence quantum yield in the dilute solution of 2,4,5-­TMe-­DPE and 2,4,6-­TMe-­
DPE. Reorganization energy versus normal mode wave numbers in (b) the gas state and (e) the solid state
of 2,4,6-­TMe-­DPE. Duschinsky rotation matrix of normal modes with index of 1–50 in (c) the gas state and
(f) the solid state of 2,4,6-­TMe-­DPE. (d) Rotation barriers calculated through rotating the dihedral angle α.
Source: Adapted from Ref. [17b] with permission from The Royal Society of Chemistry.

phenyl rings in 2,4,6-­TMe-­DPE. From the ultrafast spectroscopy measurement, the lifetime of
excited-­state species of 2,4,6-­TMe-­DPE is only 2.1 ps, much shorter than that of 2,4,5-­TMe-­DPE
(263 ps), which indicates that the 2,4,6-­TMe-­DPE is less conjugated and possesses higher motion
ability, so it can undergo much faster excited-­state decay process.
Further EVC analysis is also consistent with the abovementioned result. The low-­frequency
molecular motions dominate in isolated 2,4,6-­TMe-­DPE, which are mainly associated with torsion
of the phenyl rings and twisting of the double bond in the excited state. Projection of RE onto the
internal coordinate shows that over 80% of the total RE comes from the change in dihedral angles,
which further verifies the higher twisting mobility of 2,4,6-­TMe-­DPE. The calculation of the
Duschinsky rotation matrix (DRM) shows that the mixing of low-­frequency molecular motion
modes of 2,4,6-­TMe-­DPE is much stronger than that of 2,4,5-­TMe-­DPE. The large area of
­nondiagonal elements on DRM indicates that 2,4,6-­TMe-­DPE owns multiple coupled decay path-
ways, connecting the excited-­state state with the ground state, through molecular motions, that
can further enhance the nonradiative decay rate. On the contrary, the highly emissive 2,4,5-­TMe-­DPE
shows much less RE and less mixed motion modes. In the crystal state, the RE of 2,4,6-­TMe-­DPE
decreases to a large extent and the light emission is brightened. The abovementioned analysis
draws a clear mechanistic picture of the AIE effect that high freedom of the isolated AIE molecule
generates dominant low-­frequency molecular motions, rendering the excited-­state energy mainly
dissipated through the nonradiative pathways; meanwhile, the intensive low-­frequency molecular
motions can cause mixing among the motion modes, making the multiple decay pathways coupled
together so that the nonradiative decay is further facilitated. These two factors cause the weak
emission of AIEgens in the dilute solution. The low-­frequency molecular motions can be drasti-
cally restricted in the solid state due to their higher sensitivity to the environmental constraints, so
12 1  The Mechanistic Understanding of the Importance of Molecular Motions to Aggregation-­induced Emission

the EVC can decrease to a large extent, leading to the decrease in the nonradiative decay rate, and
then the luminescence quantum efficiency can be significantly boosted.
With persistent effort, researchers have revealed the importance of molecular motions to the
luminescence process and triggered the experimental and theoretical studies on the molecular
motions, and then, finally, they have concluded RIM, or theoretically the restriction of EVC, as the
general working principle for AIE effect. RIM works well for explaining AIE behaviors for most
AIEgens and serves as an effective design principle for the development of AIE materials. Recently,
theoreticians have further explored specific features of PES in the excited states of AIEgens, con-
sidering the crossing between the excited state and the ground state and the coupling among the
excited states.

1.3 ­Restricted Access to Conical Intersection

Previous experimental studies of molecular motions and theoretical analysis of EVC have eluci-
dated the decay pathways of AIEgens, considering the starting and final equilibrium points on the
first excited state and the ground state, respectively. Recently, theoreticians have investigated the
dynamic features for both the excited state and the ground state of typical AIEgens by scanning the
PES along specific molecular motion coordinates or by employing excited-­state molecular dynam-
ics method, which has presented a clear mechanism picture of the photoinduced dynamic evolu-
tion for AIEgens [18].
When we look into the NAC between the two equilibrium points on the excited state and the
ground state, based on the displacement harmonic approximation, it is the intensive low-­
frequency modes that contribute to the energy conversion from the electronic form to the vibra-
tional form. However, the massive molecular motions of a series of AIEgens can lead the
excited-­state structures to evolve beyond the assumption of harmonic approximation to form
highly twisted structures. In such cases, the conical intersections (CIs) between the excited state
and the ground state will become the main cause for the fast nonradiative decay or channels for
photochemical reactions [18a].
Taking the dimethyl tetraphenylsilole (DMTPS) as the model compound [18b], the central silole
ring can undergo the alternation of the C=C double bond after absorption of photons, which fur-
ther promotes its motion ability in the excited state. The potential energy profiles of the DMTPS in
the excited state optimized by CASPT2//CASSCF method (Figure 1.7b) show that the vibrational
relaxation leads to a minimum point in the S1 state with 3.1 eV, at which the central silole ring still
keeps planar, whereas along with the twisting of the silole ring, the potential energy profile experi-
ences an energy barrier at 3.54 eV and, finally, leads to a CI point at 3.0 eV between the S1 and the
S0 state, at which the silole ring turns into a highly twisted conformation, similar to the structural
reorganization of the cis-­butadiene. Since the energy barrier separating the CI and the Frank–
Condon region is lower than the vertical excitation energy, the CI is energetically accessible. Once
the excited DMTPS species reach the CI, the NAC becomes infinitely large and finally results in the
ultrafast nonradiative decay from the S1 state to the S0 state to quench the emission in the solution.
But in the crystal, the QM/MM calculation of the potential energy profile shows that the CI is
located at 4.91 eV, which is much higher than the vertical excitation energy of 3.65 eV, making the
CI energetically inaccessible. Hence, the nonradiative decay through CI can be effectively blocked
to recover the strong light emission.
When taking the dynamic evolution along the PES into consideration, it can be found that
the TPE derivatives also have abundant decay pathways in the excited state, including both
1.3  ­Restricted Access to Conical Intersectio 13

(a) (b)
4.91

3.72
3.54 3.01 3.65 3.12
3.10

Si hv

TPE-4mM TPE-4oM
DMTPS
ΦF,soln ~0.1% ΦF,soln = 64.3% Solution Crystal

(c) (d)
120 120
TPE-4mM Barrier : 6.2 kcal/mol TPE-4oM Barrier : 8.4 kcal/mol
100 Barrier : 1.8 kcal/mol
100
Energy (kcal/mol)

80
CIcyc CIpyr 80
CIcyc CIpyr
60 60
FC FC
40 40

20 20
S1 -cyc S1 -pyr S1 -cyc S1 -pyr
0 S0 -cyc S0 -pyr S0 -cyc S0 -pyr 0

10 8 6 4 2 2 4 6 8 10 10 8 6 4 2 2 4 6 8 10

LIIC path LIIC path

Figure 1.7  (a) Chemical structures and fluorescence quantum yield in the dilute solution of DMTPS,
TPE-­4mM, and TPE-­4oM. (b) Schematic illustration of the potential energy profiles of DMTPS in the
cyclohexane and the crystal. The potential energy profile in the cyclohexane is optimized at the
CASSCF/6-­31G (d) level with the PCM model. The vertical excitation energy for each point is
recalculated using CASPT2 with the ANO-­S basis set contracted to Si[4s3p1d]/C[3s2p1d]/H[2s1p]. The
calculation in the crystal is performed by using ONIOM(CASSCF/6-­31G(d):UFF) model with
recalculating the QM energy by CASPT2. Source: Adapted from Ref. [18b] with permission from The
Royal Society of Chemistry. Linear interpolation in internal coordinates paths calculated by using the
OM2/MRCI method of (c) TPE-­4mM and (d) TPE-­4oM. Source: Adapted from Ref. [8a] with permission
from American Chemical Society.

photophysical processes and photochemical reactions [8]. The excited-­state molecular dynam-


ics simulation for 1.5 ps based on the trajectory surface hopping (TSH) model shows that 75%
of the simulated trajectories decay to the ground state through a CI and form the cyclization
product with a new C–C single bond between the two adjacent phenyl rings. The distance
between the two carbons is only 2.0 Å at the cyclization CI. There is 5% of the trajectory decay
through the twisting of the C=C double bond. Similar decay features have been found in two
TPE derivatives substituted with methyl groups at meta and ortho positions. The scanning of
the linear interpolation in internal coordinates (LIIC) path using the OM2/multireference con-
figuration interaction (MRCI) method is presented in Figures 1.6d and 1.7c.
Upon photoexcitation, TPE-­4 mM can decay through the CI of cyclization with a negligible
barrier, which is consistent with the TSH simulation that 88% of the total 558 trajectories
drops into the ground state along the cyclization direction. On the other hand, there is only a
relatively small barrier of 1.8 kcal/mol between the FC region and the CI for E/Z isomeriza-
tion, which indicates TPE-­4mM will undergo ultrafast nonradiative decay so that it is almost
nonemissive in the solution. On the contrary, for the TPE-­4oM with the methyl groups substi-
tuted at the ortho position of the phenyl rings, there are two notable barriers on both the
cyclization and isomerization pathways, connecting the FC region with the cyclization CI and
14 1  The Mechanistic Understanding of the Importance of Molecular Motions to Aggregation-­induced Emission

the isomerization CI, respectively. Hence, there is no decay for TPE-­4oM to the ground state of
the total 568 trajectories in the TSH simulation for 1 ps, which accounts for its higher emission
efficiency in the solution.
With scanning the PES or simulating by the TSH method, we can find that vigorous molecular
motions of the highly flexible AIEgens in the excited state can lead to the CIs between the excited
state and the ground state. These CIs will result in enormous nonradiative decay rates or generate
another photoinduced product. Furthermore, the crossing between the excited states can also
strongly affect the photophysical behaviors of luminogens that will generate the transition forbid-
den dark states.

1.4 ­Restriction of Access to the Dark State

RIM serves as the most effective guideline for the design of AIE molecules. However, RIM requires
a more detailed elaboration when it is applied to the heteroatom-­containing AIEgens. On the one
hand, the introduction of heteroatoms endows the luminogens versatile functions, whereas, on the
other hand, it makes the excited-­state features of the luminogens more complicated. First, intro-
duction of heteroatoms, often, produces the electron-­donating or -­withdrawing effect, and then
facilitates the mixing of the overlap-­forbidden charge-­transfer (CT) state with the local-­excited
(LE) state. Second, heteroatoms with lone-­pair electrons can import the crossing between the
overlap-­forbidden (n,π*) state with the (π,π*) state. What is more, (n,π*) state will facilitate the
intersystem crossing from the singlet states to the triplet states, which are spin forbidden and can
be easily quenched. All these overlap-­forbidden and spin-­forbidden states have been defined as
dark states that are detrimental to the fluorescence process [19].
Taking the (9-­anthrylmethyl) bis(2-­pyridylmethyl) amine (APA) as an example (Figure 1.8) [9],
the APA only contains a large π-­conjugation plane consisting of the anthracene and a triple-­pyridine
motif, but it shows the unique AIE property, which differs from typical AIEgens, usually, with
multiple rotors. Although the APA contains the anthracene with large conjugation, it shows weak
UV emission with a ΦF of 0.6% in the dilute tetrahydrofuran (THF) solution with vibronic peaks.
After chelation with the zinc ion, it shows enhanced emission with the ΦF of about 100% with
the similar vibronic peaks. Meanwhile, with more than 90% water added into the pure solution
of APA, the resulting APA aggregate shows a broad and red-­shifted emission band with a peak at
461 nm. The APA crystal also shows a similar emission band at 492 nm. The large spectral variation
between the dilute solution and the solid state indicates the mixing of excited states with different
characteristics.
Indeed, with optimization of the first two excited states of the isolated APA involving the transi-
tion of the lone pair electrons, two kinds of excited-­state features have been figured out. According
to the transition characteristics, the two close-­lying excited states are assigned as (π,π*) state and
(n,π*) state, which are vibronically coupled by the molecular motions. The bright (π,π*) min owns
large oscillator strength and is responsible for the light emission, whereas the lower-­lying (n,π*)
min has the oscillator strength nearly close to 0 and crosses the (π,π*) state through the CI, which
causes the quenching of the light emission. The key molecular motion connecting these two steady
excited-­state points is the rotating of the nitrogen-­containing group. Hence, when chelating with
the zinc ion, the lone pair electrons of the nitrogen atoms and the motion of the whole chelating
group are blocked, which results in the elevation of the (n,π*) state and recovering of the light
emission from the bright (π,π*) state. Similarly, the APA molecules in the crystal can pack tightly
1.5  ­Suppression of Kasha’s Rul 15

(a)
Molecular
motion RADS
RADS

Bright state (π,π*) Dark state (n,π*)


f = 0.1114 f = 0.0000
Bright
(b) Dark (c) Bright (d) state
Dark Dark
state state
state state
Bright
state Non rad.
Ex. Non rad.
Ex. Fluo.
Ex. Fluo.

Nuclear coordination Nuclear coordination Nuclear coordination

Figure 1.8  (a) Illustration of the coupling between the bright state and the dark state by the molecular
motion, and the structures at the minimum points of the bright state and the dark state, the emissive
chelation complex, and the dimer in the crystal. Schematic potential energy surfaces of RADS related to
(b) the chelation with the zinc ion, (c) the dilute solution state, and (d) the dimers in the crystal.
Source: Adapted from Ref. [9] with permission from John Wiley and Sons.

to form dimers and stabilized by the intermolecular π–π interaction and multiple C–H interactions,
which effectively hinder the rotation of the chelating groups. The lock of lone-­pair electrons and
motions of the nitrogen-­containing group lead to the elevation of the dark (n,π*) state and finally
boost the fluorescence from the bright (π,π*) state.
This mechanism can be concluded as the restriction of access to the dark state (RADS), which
further elucidates the connotation of the RIM mechanism. The previous investigation mainly
focuses on the NAC between the first excited state and the ground state. As an extreme scenario of
NAC, the CI can cause ultrafast deactivation. In fact, the PES of multiple excited states can be cou-
pled by molecular motions and arranged in a complicated manner, especially for the heteroatom-­
containing molecular systems. The accessible dark states like the CT state, (n,π*) state, and the
triplet state will cause the fluorescence quenching. Once the molecular motions that lead to the
dark states undergo the intramolecular or environmental constraints, the fluorescence can be
restored. The multistate model has been proved effective to evaluate the excited-­state deactivation
of the heteroatom-­containing luminogens.

1.5 ­Suppression of Kasha’s Rule

For classical luminophores, the internal conversions from higher excited states to the lowest
excited states are much faster than the luminescence processes due to the rigid structures and large
conjugation even in the gas phase or dilute solution, so the light emission always comes from the
lowest excited state with a given spin multiplicity [20].
However, most AIEgens possess highly flexible structures with substantial low-­frequency molec-
ular motions that are sensitive to the environmental constraints. The vigorous molecular motions
will contribute to the coupling and mixing of multiple excited states. Once the relevant molecular
16 1  The Mechanistic Understanding of the Importance of Molecular Motions to Aggregation-­induced Emission

motions are restricted in the aggregate or solid state, it is possible that the excitons can be stabilized
in higher excited states and generate anti-­Kasha light emissions.
Qian et al., first, proposed the suppression of Kasha’s rule (SOKR) as the mechanism for the
AIE behaviors of molecular rotors based on the boron-­difluorohydrazone (BODIHY) [11a]. The
luminescence properties related to higher excited states of the BODIHY derivatives in the solu-
tion have been studied through the spectroscopy by changing the viscosity of the solution and
varying the excitation wavelength. These derivatives show viscosity-­dependent emission
enhancement but nearly no response to the solution polarity due to weak partial charge trans-
fer. According to the calculation at the TDA-­PBE level, the first excited state is designated as a
dark state. Instead, the S3 state is a bright state, which is more populated due to the relatively
large energy gap between the S3 and lower excited states. It is demonstrated that once the vis-
cosity increases, the rotation of the phenyl rings can be hindered and make the excitons stabi-
lized in the higher excited state to generate enhanced anti-­Kasha emission. However, these
results still remain controversial. Zhou et al. have recently employed more DFT functions to
recheck the excited-­state properties of the BODIHY derivatives and have challenged the SOKR
mechanism  [11b]. They have found that the TDA-­PBE method used in Ref.  [11a] may not
describe the correct order of the excited states, and the energy gaps between S3 and S2 states
obtained from this method are small enough to generate efficient internal conversion from S3
and S2 states. Hence, the emission of BODIHY derivatives in higher viscosity may not be
induced by the SOKR. Instead, they proposed that the restriction of access to the CI caused by
the flip-­flop motion is responsible for the AIE behaviors of BODIHY derivatives.
Guo et al. has recently reported another luminogen named DMF-­BP-­PXZ with highly efficient
light emission from the S2 state (Figure 1.9a, b) [11c]. According to the calculation, the S1 state is a
transition-­forbidden dark state, and the internal conversion dominates the decay process in the
solution due to the severe molecular motions, so the rapid internal conversion from S2 to S0 state,
through the intermediate S1 state, quenches the light emission. However, its kIC from S2 to S0 can
be suppressed by four orders of magnitude, and the fluorescence radiation rate can be enhanced in
the solid state, which leads to the efficient light emission from S2 state to the ground state. In this
regard, further experimental evidence for the stable population of higher excited states is still
highly desired to solidify the claim on the anti-­Kasha emission, but the mechanism demonstrated
earlier also belongs to the category of RIM, and it validates the wide reliability of the RIM
mechanism.
Exciton population on the higher excited states can also occur in the triplet states. Taking the
ClBDBT as an example (Figure 1.9c, d), it exhibits white-­light emission under UV light and persis-
tent yellow afterglow in the room temperature [11d]. According to the calculated energy levels, T1
and T2 states are all lower than the S1 state in energy, which makes both T1 and T2 accessible for the
exciton population coming from the S1 state. Furthermore, the T2 state mainly contains the (n,π*)
transition character, which leads to a larger spin–orbit coupling (SOC) between T2 and the S0 and
a higher radiative decay rate, whereas the T1 state contains more (π,π*) transition character. At
room temperature, the small energy gap between T2 and T1 can promote the thermal population
from T1 to T2. According to Boltzmann distribution, T2 has a smaller population than T1, but the
faster radiative decay from T2 results in a balanced emission intensity from both T2 and T1 states.
Thus, the combined anti-­Kasha blue light from T2 and the yellow light from T1 generate the effi-
cient white-­light emission at room temperature.
In fact, it is also the RIM process in the crystal state that stabilizes the specific electronic struc-
tures of T2 and T1 and restricts the nonradiative decay from triplet states to the ground state, and
then the balanced dual emission can be restored.
1.6 ­Through Space Conjugatio 17

(a) (b)
3.5 3.5
S2 O
S2
3.0 3.0

N 2.5

Excitation energy (ev)


2.5
Excitation energy (ev)

O S1
S1 DMF-BP-PXZ
2.0 2.0
kF(S2–S0)
kF kIC kF(S2–S0) kF kIC
1.5 1.5 = 1.3 × 107 S–1
= 1.2 × 106 S–1
1.0 1.0
kIC(S2–S1–S0)
kIC(S2–S1–S0)
0.5 0.5 = 8.8 × 1010 S–1
= 3.1 × 1010 S–1

0.0 S0 0.0 S0
Solution Solid
(c) (d)
S1
T2 0–1 0–1
0.88 eV Exp
T2
0.27 eV T1
Relative intensity

T1

0–1
3.36 eV

0–0
O 0–2
2.75 eV

CI S 0–3
CIBDBT
S0
350 450 550 650 750
Wavelength (nm)

Figure 1.9  Molecular structure, calculated energy levels, fluorescence, and internal conversion rate
constants of DMF-­BP-­PXZ in the (a) solution and (b) solid state. Source: Adapted from Ref. [11c] with
permission from John Wiley and Sons. (c) Molecular structure, calculated energy levels, and (d) emission
spectra of ClBDBT in the solid state. Source: Adapted from Ref. [11d] with permission from Springer Nature.

1.6 ­Through Space Conjugation

Most of the classical AIEgens are constructed by chromophores with through-­bond conjugation
(TBC), and their emission can be enhanced through restricting the nonradiative decay driven by
molecular motions, whereas the light emission can also be boosted by promoting the radiative
rates through the through-­space conjugation (TSC)  [13]. The TSC plays a key role in radiative
decay processes of molecular systems with clusterization-­triggered emission (CTE) property [21].
Moreover, a certain degree of molecular motions in the solid state will facilitate the intra-­or inter-
molecular excited-­state TSC for the nonconjugated molecules and stabilize the radiative channels
and, thus, promote the emission intensity [13].
18 1  The Mechanistic Understanding of the Importance of Molecular Motions to Aggregation-­induced Emission

1.6.1  Clusterization-­Triggered Emission


Traditional luminogens, generally, consist of aromatic groups or other conjugated building blocks
and take the two-­dimensional conjugated structure connected by chemical bonds. In such molecu-
lar systems with TBC, the EVC between emissive states and the ground state, the access to CIs, or
the dark states driven by molecular motions are usually the detrimental causes for emission
quenching. However, researchers have discovered and developed a nonconventional type of lumi-
nogens without any long-­range conjugated structures but containing only isolated units such as
heteroatoms with lone-­pair electrons, unsaturated C=C, C=O, and C≡N groups as presented in
Figure 1.10. Although this kind of luminogens lacks luminescent centers intuitively, and they are
almost nonemissive in the dilute solution, they can emit visible light efficiently in the aggregate or
solid state, showing the typical AIE behaviors. Because of the lack of largely conjugated lumines-
cent centers and much higher motion ability than the traditional molecules with long-­range con-
jugation, such nonconjugated molecules are nonemissive in the isolated state. However, it can be
noticed that these heteroatom-­containing groups usually exist as amide, imide, ketone, anhydride,
or ester subunits connected by the saturated bonds in the macromolecule backbones, and they are
usually rich in electrons. Hence, once the aggregation or cross-­linking occurs, electron-­rich moie-
ties can cluster together and form a relatively stable long-­range through-­space electron ocean and
energy bands by the electron overlapping for the electronic transition, which is similar to the
energy band structures in the inorganic semiconductors. Upon excitation, electrons can jump into
the excited state based on the stabilized through-­space energy bands and light up the whole clus-
ter system.
Such nonconventional phenomenon has been coined as CTE; it deciphers that the severe molec-
ular motions and nonconjugated structures make the CTE luminogens nonemissive in the isolated
state or the dilute solution, whereas diverse intra-­and intermolecular interactions in the cluster or
cross-­linking can hamper the molecular motions and stabilize the conformations with large elec-
tron overlapping and thus boost the light emission of CTE luminogens, in which the packing
density and the cluster size can affect the emission efficiency  [21]. Taking the polyacrylonitrile

(a) (b) (c)


Isolated Crosslinked Clustered
Eδ+ Eδ– : dipole or transient dipole n
PAN
N
Small Large Eδ+ Eδ– Eδ+ Eδ–
Sn/Tn
Sn/Tn Eδ– Eδ+ Eδ– Eδ+
Sn/Tn
Sn/Tn Eδ+ Eδ– Eδ+ Eδ– n
O CH3
O O
E1 > E2 > E3 > E4 Eδ– Eδ+ Eδ– Eδ+ N
PNHSMA
O
S0
Through-Space Conjugation (TSC)
S0
TSC: n σ* interaction,
S0 n π* interaction,
S0 Non-conjugated linkers π π* interaction,
n electrons: N, O, P, S.../π electrons: aromatic rings, multiple bonds. H-bond interaction....

Figure 1.10  (a) Schematic illustration of the cluster formation, energy variation, and (b) the formation of
through-­space conjugation involved in the CTE. Source: Adapted from Ref. [21a] with permission from
Elsevier. (c) Examples of CTEgens. Source: Adapted from Ref. [21b] with permission from John Wiley
and Sons.
1.6 ­Through Space Conjugatio 19

(PAN)  [21b] and poly(N-­hydroxysuccinimidyl methacrylate) (PNHSMA)  [21c] as the example,


they contain no aromatic groups and are nonconjugated in the isolated polymer backbone
(Figure 1.10c). Therefore, these two molecules are nearly nonemissive in the dilute solution. But
they can emit strong blue light in the concentrated solution or solid powder. CTE phenomenon
also exists in the natural products such as starch, cellulose, and protein; they can show notable
blue light in the solid state under UV excitation [21d].

1.6.2  Polymerization-­induced Emission


Recently, Tang and coworkers have proposed polymerization-­induced emission (PIE), which is
another conceptual innovation related to CTE [22]. It describes the process where the nonemissive
monomers can be converted into luminescent polymers through polymerization. AIE process
occurs mainly by physically manipulating the molecular motions, whereas PIE is achieved through
the chemical ways accompanied by the CTE process. As versatile polymerization methods can be
utilized to construct the PIE polymers, and these unusual luminescent polymers own good pro-
cessability, the PIE acts as a promising solution in developing novel soft luminescence materials.
The working principle underneath is similar with CTE, nonconjugated subunits with rich elec-
trons, such as phenyl, hydroxyl, and carbonyl groups, ether, and amide, can be connected into
polymer chains by chain polymerization or step polymerization, and such polymer chains can be
entangled and form multilevel structures through diverse intra-­ or interchain motions. Then, the
electron-­rich moieties will aggregate into a cluster with electron overlapping in multiple micro-
structures and finally generate visible light. The emission intensity of the PIE polymers will
increase with promoting the polymerization degree and the molecular weight. The intrinsic diverse
structures endow the PIE polymers the potential to create diverse luminescence performance.

1.6.3  Excited-­state Through-­space Conjugation


From the abovementioned discussion, molecular motions are mainly detrimental to lumines-
cence. Tang and coworkers have recently revealed a new role of molecular motions in the
excited-­state deactivation process of the tetraphenylethane (s-­TPE) derivatives as presented in
Figure 1.11a and b [13a].
The s-­TPE only contains four phenyl rings connected by the saturated single bond but can
intensely emit visible light with the peak at 467 nm in the solid state. The dilute solution of s-­TPE
shows ultraviolet emission that basically stems from the isolated phenyl rings, but with adding
more than 70% water into the dilute solution, s-­TPE molecules can aggregate accompanied with a
notable emission peak at 460 nm emerging, showing the typical AIE property. Why do the noncon-
jugated systems emit the visible light? Due to the highly twisted and flexible structure of s-­TPE, it
shows no obvious intermolecular π–π interaction in the single crystal. The theoretical simulation
of the exciton coupling also shows that there is no notable intermolecular coupling in the excited
state, so it should be the intramolecular interaction that affects the emission.
Further optimization of the excited-­state structures shows that s-­TPE will encounter substantial
conformational reorganization, and during the relaxation, the distance between the geminal phe-
nyl rings can gradually decrease. When it decays to the S1,min, two geminal phenyl rings reach a
close-­contacting conformation with large through-­space overlapping of π orbitals. Indeed, the
transition energy gaps decrease along with the phenyl rings getting close, so the emission wave-
lengths can be redshifted with such intramolecular motions. Finally, the calculated emission at
S1,min becomes 460  nm, which is well consistent with the experimental data. The EVC analysis
20 1  The Mechanistic Understanding of the Importance of Molecular Motions to Aggregation-­induced Emission

(a) (b)

LUMO
R H
H –1.406
H

Egap = 3.439
H
s-TPE
–4.845
R

ΦF = 69%
HOMO
R
λem = 467 nm

Through-space conjugation AIE


Excited state
(c) (d)
Clusteroluminescence

LUMO
UV light LUMO
S0 → S1: 5.32 eV S0 → S1: 4.33 eV
Removal
HOMO
HOMO

No/weak Excited-state
intermolecular through-space
interaction complex (ESTSC) Ground state Excited state

Figure 1.11  Schematic illustration of the excited-­state (a) intramolecular through-­space conjugation of


s-­TPE and (c) intermolecular through-­space complex of s-­DPE. Molecular orbitals involved in the transition
between the excited state and the ground state of (b) s-­TPE and (d) s-­DPE. Source: Adapted from Refs. [13a,
b] with permission from American Chemical Society.

shows that s-­TPE molecules own much higher excited-­state motion ability in the dilute solution, so
the conformation with intramolecular TSC can be easily disturbed by the vigorous molecular
motions. Once the s-­TPE molecules aggregate, the molecular motions will be restricted to a certain
degree, as indicated by the suppressed structural reorganization, which, thus, stabilizes the
through-­space conjugated conformation and reduces the nonradiative decay rates, and so the
solid-­state s-­TPE can emit enhanced visible light.
A certain degree of molecular motions in the solid state will facilitate the formation and stabili-
zation of emissive states with TSC. The diphenylethane (s-­DPE) is another nonconventional
AIEgen as presented in Figure 1.11c and d [13b]. It contains no long-­range through-­bond conju-
gated structure but two isolated phenyl rings. The absorption and photoluminescence spectra in
the dilute THF solution are ascribed to the electronic transition of the phenyl rings, with peaks at
270 and 285 nm, respectively. However, with addition of water into its dilute solution, a new emis-
sion peak at 355 nm gradually emerges, showing the AIE phenomenon. The emission peak in the
1.7 ­Perspectiv 21

solid state of s-­DPE is also redshifted to 355 nm with a small fraction of the emission at 285 nm
coming from the isolated state. Different from s-­TPE, the isolated s-­DPE molecule cannot form the
intramolecular TSC, so the only way to redshift the light emission is to form the intermolecular
TSC. The optimization of the dimer of s-­DPE based on two adjacent s-­DPE molecules in the single
crystal shows that the two s-­DPE molecules gradually get closer and turn the orientation of two
phenyl rings from face-­to-­edge at S0,min to face-­to-­face at S1,min and finally form an excited-­state
through-­space complex (ESTSC) with significant orbit overlapping between the two respective
phenyl rings of the two s-­DPE molecules. Meanwhile, the energy gap decreases from 5.32 to 4.33 eV
along with the light-­driven solid-­state intermolecular motions, which is the luminescence origin of
s-­DPE in the solid state.
As the characteristic ultrafast transient absorption spectra reveal the most probable conforma-
tions in the excited-­state timescale, the femtosecond transient absorption (fs-­TA) spectroscopy
has been utilized to prove the existence of ESTSC [13b]. The steady-­state absorption spectra of
­s-­DPE in dilute solution and crystal film have been first measured, respectively. It has been found
that the absorption spectra in these two conditions are matched with each other, indicating there
are no species involving intermolecular electronic coupling in the ground-­state crystal film. Then,
the fs-­TA spectra of s-­DPE in the dilute solution show that there are no notable absorption peaks
in the long-­wavelength region in the excited state but only peaks located in the short-­wavelength
region. However, when it comes to the crystal state, the fs-­TA spectra of s-­DPE crystal film show
that the peaks from 340 to 450  nm almost disappear, but new peaks at 590  nm emerge as the
dominant characteristic absorption peaks with the lifetime of 525 ps, which is similar to a char-
acteristic transient absorption peak at 570 nm for the literature-­reported naphthalene excimer.
According to the theoretical simulation, this long-­wavelength characteristic peaks can be assigned
to the ESTSC. Since the ground-­state s-­DPE crystal will not form complexes, the ESTSC can only
be formed in the excited state along with the intermolecular motions.
What is the driving force for the molecular motions that lead to the formation of ESTSC? The
Mulliken charges have been mapped onto molecular structures of the isolated s-­DPE and the
ESTSC [13b]. It shows that partial molecular charges of two carbon atoms of the isolated s-­DPE
connecting two phenyl rings and the central ethane moiety are negative, whereas the other carbon
atoms show positive partial charges. Hence, it is proposed that such opposite partial charge distri-
bution will induce the transient dipole to attract the adjacent two molecules approaching to each
other and finally form a stable ESTSC. For the ESTSC, the partial charges of the two connection
carbon atoms become even more positive.
CTE can be categorized as a kind of static clusteroluminescence, in which the aggregation
restricts the molecular motions and stabilizes the emissive cluster. In contrast, ESTSC can be
ascribed as dynamic clusteroluminescence, in which the excited-­state molecular motions in solid
state facilitate the formation of the key emissive cluster. The ESTSC opens a broad way to utilize
the excited-­state solid-­state molecular motions for luminescence and offers a new design strategy
for dynamic luminescence materials.

1.7 ­Perspective

Traditional photophysical research usually targets highly rigid and conjugated chromophores.
However, investigation of the relationship between molecular motions with the excited-­state decay
process as well as manipulation of molecular motions in different manners and multiple states has
been a consistent topic in the AIE mechanistic research [23]. From tremendous experimental and
22 1  The Mechanistic Understanding of the Importance of Molecular Motions to Aggregation-­induced Emission

theoretical contributions, RIM has been concluded as the general working principle, or the restric-
tion of EVC in the quantum view [6, 7]. AIEgens own high performance in the aggregate or solid
state that can hamper the internal conversion between excited states, in which the Kasha’s rule
may be invalid [11]. In light of this, the multistate mechanistic model, involving emissive excited
states and transition-­forbidden states, has offered perspectives in exploring the nature of higher
excited states in the AIE process. In fact, multiple excited states can be involved in the photophysi-
cal process of AIEgens, especially for the heteroatom-­containing systems. The proximity effect
between the excited states with different property tends to induce the mixing of the states and
finally leads to the crossing between dark states (CT, (n, n*), triplet states, etc.) and bright states [19].
The aggregate effect can render repopulation of the excited species, which may make the photo-
physical processes more complicated than those predicted by the classical luminescence theo-
ries [9]. Furthermore, the current RIM mechanistic model mainly focuses on the single-­molecule
property, whereas the surrounding molecules serve as the environmental constraint. Hence, more
accurate and general theoretical models for the aggregate are still highly desired that consider
adequate multistate properties and diverse intermolecular interactions.
Recently, research on the CTE has substantially expanded the mechanistic scope into the aggregate
science, that the efficient light emission can be achieved through the clusterization of nonconjugated
subunits [21]. Although the power source of CTE is mainly ascribed to the electronic overlapping
among electron-­rich moieties, the specific cluster structures remain unclear and only at the simula-
tion stage as well as the specific interaction involving in the electronic transition of CTEgens, so reli-
able experimental characterization of the cluster structures and more accurate molecular dynamics
simulation are highly demanded. In the sense that the clusterization in the ground state or the excited
state will lead to electronic transition and boost radiative decay rates, CTE has opened a new broad
way for luminescence research in the mesoscale. It is also likely to lead to the fusion of the organic
luminescence theory with the inorganic band theory in CTE research.
From the traditional rigid chromophores to flexible AIEgens, and from classical AIEgens with
TBC to nonconventional CTEgens with TSC, the mechanistic studies on luminescence have been
evolving forward step by step. The study on the excited-­state TSC has unveiled the mysteries of
dynamic emissive states, driven by the excited-­state molecular motions in the solid state  [13].
Indeed, the intermolecular motions can also influence the luminescence process. On the one hand,
the light-­driven solid-­state intermolecular motions contribute to the formation of ESTSC for non-
conjugated luminogens. On the other hand, they can also be detrimental to the intermolecular
emissive species and cause kinetic quenching. In this sense, the research of molecular motions can
be expanded into multiple intermolecular motions.
From the abovementioned cases, a certain degree of solid-­state molecular motions can do advan-
tageous work. This effect can be further demonstrated in the Herzberg–Teller (HT) effect [12], in
which the vibronic coupling will facilitate the electronic transition, as well as the spin-­vibronic
coupling (SVC) for intersystem crossing process [24]. The HT effect describes that the electronic
transition will be forbidden if the lowest excited states are dark states. However, with specific
motion degrees, the interaction between electronic and vibrational motions can help to compen-
sate for the transition dipole moment through specific vibration modes and facilitate the radiative
transition. Such radiative enhancement has not been investigated for solid-­state AIEgens, which
may help to open another mechanistic research topic based on aggregate science. For most organic
room-­temperature phosphorescence (RTP) systems, increasing the SOC and restricting the nonra-
diative decay are the two main ways to enhance RTP efficiency. In this regime, the internal conver-
sion driven by NAC associated with the molecular motions and the intersystem crossing driven by
the SOC are always discussed separately due to the largely different timescale for these processes
to take effect. In fact, the vigorous molecular motions can cause strong mixing among excited
 ­Reference 23

states, in which the NAC and the SOC will present simultaneously. Hence, for evaluating the inter-
system crossing process, the electron, spin, and vibration should all be considered in spin–vibronic
levels [24]. The discussion on the HT effect shows that a specific vibrational degree of freedom is
required to activate the HT transition. Similarly, the SOC is not always limited to a specific configu-
ration. The higher-­order terms depending on the molecular vibrations in the SOC expression will
have significant contributions to the total SOC and finally facilitate the intersystem crossing. In
this sense, representative molecular models can be designed to explore the SVC mechanism for
organic RTP systems and finally make efficient use of the molecular motions.

­References

1 (a) Kassem S, Leeuwen T, Lubbe AS, et al. Artificial molecular motors. Chem Soc Rev 2017; 46:
2592–2621; (b) Leeuwen T, Lubbe AS, Štacko P, et al. Dynamic control of function by light-­driven
molecular motors. Nat Rev Chem 2017; 1: 0096; (c) Erbas-­Cakmak S, Leigh DA, McTernan CT, et al.
Artificial molecular machines. Chem Rev 2015; 115: 10081−10206.
2 Turro NJ, Scaiano JC and Ramamurthy V. Modern molecular photochemistry of organic molecules.
University Science Books, 2010.
3 (a) Shuai Z and Peng Q. Organic light-­emitting diodes: theoretical understanding of highly efficient
materials and development of computational methodology. Natl Sci Rev 2017; 4: 224−239; (b) Shuai
Z and Peng Q. Excited states structure and processes: understanding organic light-­emitting diodes at
the molecular level. Phys Rep 2014; 537: 123−156.
4 Forster T and Kasper K. Ein Konzentrationsumschlag der Fluoreszenz. Zeitschrift für Physikalische
Chemie 1954; 1: 275−277.
5 Luo J, Xie Z, Lam JWY, et al. Aggregation-­induced emission of 1-­methyl-­1,2,3,4,5-­pentaphenylsilole.
Chem Commun 2001; 18: 1740−1741.
6 (a) Mei J, Leung NLC, Kwok RTK, et al. Aggregation-­induced emission: together we shine, united we
soar! Chem Rev 2015; 115: 11718−11940; (b) Mei J, Hong Y, Lam JWY, et al. Aggregation-­induced
emission: the whole is more brilliant than the parts. Adv Mater 2014; 26: 5429−5479; (c) He Z, Ke C and
Tang BZ. Journey of aggregation-­induced emission research. ACS Omega 2018; 3: 3267−3277; (d) Chen
Y, Lam JWY, Kwok RTK, et al. Aggregation-­induced emission: fundamental understanding and future
developments. Mater Horiz 2019; 6: 428−433; (e) Zhao Z, Zhang H, Lam JWY, et al. Aggregation-­
induced emission: new vistas at the aggregate level. Angewandte Chemie Int Edn 2020; 59: 2−22.
7 (a) Zhang T, Jiang Y, Niu Y, et al. Aggregation effects on the optical emission of
1,1,2,3,4,5-­Hexaphenylsilole (HPS): a QM/MM study. J Phys Chem A 2014; 118: 9094−9104; (b)
Zhang T, Peng Q, Quan C, et al. Using the isotope effect to probe an aggregation induced emission
mechanism: theoretical prediction and experimental validation. Chem Sci 2016; 7: 5573−5580; (c)
Zhang T, Ma H, Niu Y, et al. Spectroscopic signature of the aggregation-­induced emission
phenomena caused by restricted nonradiative decay: a theoretical proposal. J Phys Chem C 2015;
119: 5040−5047; (d) Wu Q, Deng C, Peng Q, et al. Quantum chemical insights into the aggregation
induced emission phenomena: a QM/MM study for Pyrazine derivatives. J Comput Chem 2012; 33:
1862−1869; (e) Peng Q, Yi Y, Shuai Z, et al. Toward quantitative prediction of molecular
fluorescence quantum efficiency: role of Duschinsky rotation. J Am Chem Soc 2007; 129: 9333−9339;
(f) Niu Y, Li W, Peng Q, et al. Molecular materials property prediction package (MOMAP) 1.0: a
software package for predicting the luminescent properties and mobility of organic functional
materials. Mol Phys 2018; 116: 1078−1090; (g) Niu Y, Peng Q, Shuai Z, et al. Promoting-­mode free
formalism for excited state radiationless decay process with Duschinsky rotation effect. Sci China
Ser B Chem 2008; 51: 1153−1158.
24 1  The Mechanistic Understanding of the Importance of Molecular Motions to Aggregation-­induced Emission

8 (a) Gao Y, Chang X, Liu X, et al. Excited-­state decay paths in tetraphenylethene derivatives. J Phys
Chem A 2017; 121: 2572–2579; (b) Prlj A, Došlić N, Corminboeuf C, et al. How does
tetraphenylethylene relax from its excited states? Phys Chem Chem Phys 2016; 18: 11606–11609.
9 Tu Y, Liu J, Zhang H, et al. Restriction of access to the dark state: a new mechanistic model for
heteroatom-­containing AIE systems. Angewandte Chemie Int Edn 2019; 58: 14911–14914.
10 (a) Bu F, Duan R, Xie Y, et al. Unusual aggregation-­induced emission of a coumarin derivative as a
result of the restriction of an intramolecular twisting motion. Angewandte Chemie Int Edn 2015;
54: 14492–14497; (b) Zhao Z, Zhen X, Du L, et al. Non-­aromatic annulene-­based aggregation-­
induced emission system via aromaticity reversal process. Nat Commun 2019; 10: 2952.
11 (a) Qian H, Cousins ME, Horak EH, et al. Suppression of Kasha’s rule as a mechanism for
fluorescent molecular rotors and aggregation-­induced emission. Nat Chem 2017; 9: 83–87; (b)
Zhou P, Li P, Zhao Y, et al. Restriction of flip-­flop motion as a mechanism for aggregation-­induced
emission. J Phys Chem Lett 2019; 10: 6929−6935; (c) Guo J, Fan J, Lin L, et al. Mechanical insights
into aggregation-­induced delayed fluorescence materials with anti-­Kasha behavior. Adv Sci 2019; 6:
1801629; (d) He Z, Zhao W, Lam JWY, et al. White light emission from a single organic molecule
with dual phosphorescence at room temperature. Nat Commun 2017; 8: 416.
12 Herzberg G and Teller E. Schwingungsstruktur der Elektronenübergänge bei mehratomigen
Molekülen. Zeitschrift für Physikalische Chemie 1933; 21: 410.
13 (a) Zhang H, Zheng X, Xie N, et al. Why do simple molecules with “isolated” phenyl rings emit
visible light? J Am Chem Soc 2017; 139: 16264–16272; (b) Zhang H, Du L, Wang L, et al.
Visualization and manipulation of molecular motion in the solid state through photoinduced
clusteroluminescence. J Phys Chem Lett 2019; 10: 7077−7085; (c) Sturala J, Etherington MK,
Bismillah AN, et al. Excited-­state aromatic interactions in the aggregation-­induced emission of
molecular rotors. J Am Chem Soc 2017; 139: 17882–17889.
14 (a) Chen J, Lam CCW, Lam JWY, et al. Synthesis, light emission, nanoaggregation, and restricted
intramolecular rotation of 1,1-­substituted 2,3,4,5-­tetraphenylsiloles. Chem Mater 2003; 15:
1535–1546; (b) Fan X, Sun J, Wang F, et al. Photoluminescence and electroluminescence of
hexaphenylsilole are enhanced by pressurization in the solid state. Chem Commun 2008; 26:
2989−2991; (c) Li Z, Dong Y, Mi B, et al. Structural control of the photoluminescence of silole
regioisomers and their utility as sensitive regiodiscriminating chemosensors and efficient
electroluminescent materials. J Phys Chem B 2005; 109: 10061–10066; (d) Zhao E, Lam JWY, Hong
Y, et al. How do substituents affect silole emission? J Mater Chem C 2013; 1: 5661−5668; (e) Liang
GD, Lam JWY, Qin W, et al. Molecular luminogens based on restriction of intramolecular motions
through host–guest inclusion for cell imaging. Chem Commun 2014; 50: 1725−1727; (f) Qin A,
Lam JWY, Mahtab F, et al. Pyrazine luminogens with “free” and “locked” phenyl rings:
understanding of restriction of intramolecular rotation as a cause for aggregation-­induced
emission. Appl Phys Lett 2009; 94: 253308.
15 Leung NLC, Xie N, Yuan W, et al. Restriction of intramolecular motions: the general mechanism
behind aggregation-­induced emission. Chem A Eur J 2014; 20: 15349–15353.
16 Cai Y, Du L, Samedov K, et al. Deciphering the working mechanism of aggregation-­induced
emission of tetraphenylethylene derivatives by ultrafast spectroscopy Chem Sci 2018; 9: 4662.
17 (a) Liu J, Pan L, Peng Q, et al. Tetraphenylpyrimidine-­based AIEgens: facile preparation,
theoretical investigation and practical application. Molecules 2017; 22: 1679; (b) Zhang H, Liu J, Du
L, et al. Drawing a clear mechanistic picture for the aggregation-­induced emission process. Mater
Chem Front 2019; 3: 1143–1150; (c) Chen M, Hu X, Liu J, et al. Rational design of red AIEgens with
a new core structure from non-­emissive heteroaromatics. Chem Sci 2018; 9: 7829–7834; (d) Chen
 ­Reference 25

M, Zhang X, Liu J, et al. Evoking photothermy by capturing intramolecular bond stretching


vibration-­induced dark-­state energy. ACS Nano 2020; 14: 4265–4275.
18 (a) Crespo-­Otero R, Li Q and Blancafort L. Exploring potential energy surfaces for aggregation-­
induced emission—­from solution to crystal. Chem Asian J 2019; 14: 700–714; (b) Peng X, Ruiz-­
Barragan S, Li Z, et al. Restricted access to a conical intersection to explain aggregation induced
emission in dimethyl tetraphenylsilole. J Mater Chem C 2016; 4: 2802–2810; (c) Li Q and Blancafort
L. A conical intersection model to explain aggregation induced emission in diphenyl
dibenzofulvene. Chem Commun 2013; 49: 5966–5968; (d) Sasaki S, Suzuki S, Sameera WMC, et al.
Highly twisted N,N-­dialkylamines as a design strategy to tune simple aromatic hydrocarbons as
steric environment-­sensitive fluorophores. J Am Chem Soc 2016; 138: 8194–8206.
19 (a) Liese D and Haberhauer G. Rotations in excited ICT states—­fluorescence and its
microenvironmental sensitivity. Isr J Chem 2018; 58: 813–826; (b) Grabowski ZR, Rotkiewicz K and
Rettig W. Structural changes accompanying intramolecular electron transfer: focus on twisted
intramolecular charge-­transfer states and structures. Chem Rev 2003; 103: 3899–4031; (c) Lower S
and El-­Sayed M. The triplet state and molecular electronic processes in organic molecules. Chem
Rev 1966; 66: 199–241; (d) Zgierski MZ, Fujiwara T and Lim EC. Role of the πσ* state in molecular
photophysics. Acc Chem Res 2010; 43: 506–517.
20 Kasha, M. Characterization of electronic transitions in complex molecules. Discuss Faraday Soc
1950; 9: 14–19.
21 (a) Zhang H, Zhao Z, McGonigal PR, et al. Clusterization-­triggered emission: uncommon
luminescence from common materials. Mater Today 2020; 32: 275–292; (b) Zhou Q, Cao B, Zhu C,
et al. Clustering-­triggered emission of nonconjugated polyacrylonitrile. Small 2016; 12: 6586–6592;
(c) Bin X, Luo W, Yuan W, et al. Clustering-­triggered emission of poly (N-­hydroxysuccinimide
methacrylate). Acta Chim Sin 2016; 74: 935–941; (d) Gong Y, Tan Y, Mei J, et al. Room temperature
phosphorescence from natural products: crystallization matters. Sci China Chem 2013; 56:
1178–1182; (e) Yuan W and Zhang Y. Nonconventional macromolecular luminogens with
aggregation-­induced emission characteristics. Polym Chem 2017; 55: 560–574.
22 Liu B, Zhang H, Liu S, et al. Polymerization-­induced emission. Mater Horiz 2020; 7: 987–998.
23 Liu S, Li Y, Zhang H, et al. Molecular motion in the solid state. ACS Mater Lett 2019; 1: 425–431.
24 Penfold TJ, Gindensperger E, Daniel C, et al. Spin-­vibronic mechanism for intersystem crossing.
Chem Rev 2018; 118: 6975–7025.
27

Understanding the AIE Mechanism at the Molecular Level


Xiaoyan Zheng1 and Qian Peng2
1
Beijing Key Laboratory of Photoelectronic/Electrophotonic Conversion Materials, Key Laboratory of Cluster Science of Ministry
of Education, School of Chemistry and Chemical Engineering, Beijing Institute of Technology, Beijing, China
2
School of Chemical Sciences, University of Chinese Academy of Sciences, 100049 Beijing, China

2.1 ­Introduction

Traditionally, molecular aggregation tends to reduce fluorescence efficiency because of


aggregation-­caused quenching (ACQ): the formation of nonemissive excimer, intermolecular
charge transfer, or energy transfer to quenching sites through strong π–π stacking, which is a
large obstacle to utilize a variety of efficient luminogens in the real world [1]. The opposite phe-
nomenon, where the compounds are non-­or low-­emissive in dilute solution but highly emissive
in the aggregated phase (aggregates in solution, film, or crystalline phase), is termed as
aggregation-­induced emission (AIE) in 2001 by Tang et  al., which paves an efficient way for
highly efficient luminescent material design  [2–4]. The compounds with AIE characteristics
(AIEgens) have attracted much attention for their wide applications in solid-­state lighting, flat
panel display, chemical sensor, cell imaging, and so on in the last 20 years [2]. The luminescent
properties in the solid state are often significantly different from the isolated component at the
molecular level. It is urgent to reveal the intrinsic AIE mechanism to realize the precise design
of more efficient emitters; however, it is a challenging task.
In this rapid expanding field, many endeavors have been devoted to unravel the AIE mecha-
nism through both theory and experimental means. The J-­aggregation was demonstrated to play
an important role for the enhanced solid-­state emission of the strong conjugated rigid systems in
earlier works [5]. For flexible systems, the restriction of intramolecular rotations (RIR) [6–8] was
proposed firstly by comparatively analyzing the molecular steric hindrance-­, temperature-­, and
viscosity-­dependent fluorescence intensities through all kinds of experimental measurements.
Then, the restriction of intramolecular vibration (RIV) [3] and the restriction of intramolecular
motion (RIM) [9] were put forward to explain more universal AIE mechanism through combin-
ing experiments and theoretical calculations. In theory, Peng and Shuai et al. developed a thermal
vibration correlation function (TVCF) formalism  [10–12] based on Fermi’s golden rule (FGR),
which could quantitatively evaluate the nonradiative decay rate constants for isolated molecules,
nanoaggregates, and crystals. And they proposed the AIE mechanism that the trip-­out of

Handbook of Aggregation-Induced Emission: Volume 1 Tutorial Lectures and Mechanism Studies, First Edition.
Edited by Youhong Tang and Ben Zhong Tang.
© 2022 John Wiley & Sons Ltd. Published 2022 by John Wiley & Sons Ltd.
28 2  Understanding the AIE Mechanism at the Molecular Level

electron–vibration coupling and the demix of vibration modes block the excited-­state nonradia-
tive decay channels in various rigid environments, sharply decreasing the nonradiative rate con-
stants and turning fluorescence on [12–14]. Blancafort et al. proposed the restricted access to a
conical interaction mechanism based on the potential energy surface (PES) topology analysis [15,
16] (see Chapter  9). In addition, many other scenarios have been declared, such as hydrogen-­
bond-­induced excimer emission  [17], vibration-­induced emission  [18], the restriction of E/Z
isomerization process [19], the blockage of access to dark state via isomerization mechanism [20,
21], and the reversal from dark 1(n,π*) or 1(n + σ,π*) to bright 1(π,π*) excited states for generation
of the solid-­state emission [22]. In this chapter, we will focus on introducing the revealed AIE
mechanism in which the photophysical behaviors are supposed to occur in the Franck–Condon
region without any photoisomerizations or conical intersections in the PESs. The chapter is
organized in the following manner. In the second section, we show a short overview of the theo-
retical methodology and procedure. In the third section, we carefully elaborate on the revealed
AIE mechanism at the level of first principle by several representative AIEgens. In the fourth
section, we further validate our unraveled AIE mechanism by establishing the relationship
between the calculated key photophysical parameters with measurable signals in experiments.
Finally, we give a brief conclusion of the currently revealed AIE mechanism and outlook of the
development direction theoretically in the future.

2.2 ­Theoretical Methods

2.2.1  Radiative and Nonradiative Rate Constants


The radiative decay rate constant (kr) and nonradiative decay rate constant (knr = kic + kisc) are
decisive parameters for the luminescence quantum efficiency, where kic and kisc are nonradiative
internal conversion (IC) and intersystem crossing (ISC) rate constants, respectively. In most
organic molecules, kisc can be neglected owing to the small spin−orbit coupling for the π → π*
electronic transition; therefore, knr ≈ kic.
The kr was computed by integrating over the whole emission spectrum:

k r T em ,T d (2.1)

3
4  2
em ,T Piv T fu fi i i ,fu (2.2)
3c 3 u,

where Piν is the Boltzmann distribution function of the initial state at a certain temperature, Θ is
 
the nuclear vibrational wave function, and fi f i is the electric transition dipole moment
between two electronic states.
Based on the FGR, the nonradiative IC constant can be written as [23]:

2
2  fu,iv
k IC u,
Pi H Eiv Efu , (2.3)


where Eiv(Efu) reflects the electronic and vibrational energies of the initial (final) state, and H
represents the non-­Born–Oppenheimer coupling.
2.2 ­Theoretical Method 29

Based on the Franck–Condon principle, applying Fourier transform of the δ-­function,


Equation 2.3 can be written as [24, 25]:

1
kIC Rkl dtei if t
Zi 1 IC t, T , (2.4)
kl
2 z

where Rkl f Pˆfk i i Pˆfl f is the nonadiabatic electronic coupling and ρIC(t, T) is the ­thermal

vibration correlation function (TVCF) [10–12]

t, T Tr Pˆfk e i f Hˆ f
Pˆfl e i i Hˆ i (2.5)
IC ,kl

As shown in Equation 2.4, kIC is mainly ruled by three factors: nonadiabatic electronic coupling,
transition energy, and the vibronic relaxation energy. Based on the electronic structure informa-
tion obtained from first-­principle calculations, the radiative and nonradiative rate constants were
calculated by solving Equations 2.1 and 2.3 through the TVCF rate theory in the MOMAP pro-
gram [26]. The difference between the PESs of two electronic states is considered by Qi = SQf + D,
where S is the Duschinsky rotation matrix and D is the displacement vector, that is, Duschinsky
rotation effect (DRE) [13, 14].

2.2.2  Computational Details


To unravel the AIE mechanism, the photophysical properties of compounds in both the isolated
(dilute solution) and the aggregated states (in amorphous aggregate and crystal) need to be consid-
ered. Different models were built according to the specific environment of the studied systems.
For the dispersed molecule in dilute solution, the polarizable continuum model (PCM) model is
chosen to include bulk solvent effect implicitly (see Figure 2.1a). The polarities of different sol-
vents are represented by dielectric constants.
In crystal, the hybrid quantum mechanics/molecular mechanics (QM/MM) model can be set up
based on the experimental crystal structure to consider the environment effect. First, a large clus-
ter is extracted from the crystal. To balance the accuracy and computational cost, the cluster is
spatially divided into a region where the QM description is essential, with the remainder repre-
sented by MM classically (Figure 2.1b). The general Amber force field (GAFF) [27] or the united
force field (UFF) was used for the MM treatment. In the QM part, density functional theory (DFT)/
time-­dependent DFT (TDDFT) is chosen to deal with the properties of the ground and excited
states, respectively. During the calculation, the QM molecule is active, and all molecules in the
MM region are frozen. The electrostatic embedding scheme was adopted to consider the interac-
tions between QM and MM regions [28, 29].
For AIEgens in amorphous aggregate, the molecular conformation is not available in the experi-
ment. Therefore, the large time-­scale molecular dynamics (MD) simulations need to be performed
in solution to obtain the conformations of amorphous aggregates of AIEgens. The topology of the
AIEgens could be set up by GAFF. The initial conformations were generated by randomly placing
tens or hundreds of AIEgens in cubic simulation boxes with enough edge length. Then, the ran-
domly dispersed AIEgen system was solvated by specific solvents reported in the experiment. The
energy minimization was firstly performed using the steepest descent algorithm, followed by a 5-­ns
30 2  Understanding the AIE Mechanism at the Molecular Level

(b)
(a)

PCM

Solution Crystal

(c)

MD conformation Aggregate

Figure 2.1  (a) Setup of the PCM model for an isolated molecule in solution; setup of the QM/MM models
for a molecule in (b) crystal and (c) amorphous aggregate. The molecule in green is the QM region and the
others in silver are the MM region (taking HPS as an example).

pre-­equilibrated MD simulation under the NPT ensemble. Finally, several long time-­scale (more
than 20 ns) production MD simulations were performed with different initial velocities coupled by
the velocity rescaling thermostat [30] and Parrinello–Rahman barostat [31]. The time step was 2 fs.
The configurations were stored at a specific time interval (e.g. 20 ps) for data analysis and the fol-
lowing calculation of photophysical properties. Then, we randomly extracted equilibrated aggre-
gates from MD trajectories and set up QM/MM models accordingly to calculate the optical properties.
Different from crystal, the molecular packing of amorphous HPS aggregates is irregular, resulting in
distinct local environments for each molecule in amorphous aggregate. For example, the molecule
embedded in aggregate is much different from the exposed one. In addition, conformations of dif-
ferent frames or trajectories are also different. Thus, several MD conformations need to be extracted
from MD trajectories to set up QM/MM models to consider various molecular packings in amor-
phous aggregate. In the QM/MM model, the solvents could be removed from the extracted MD
conformations to reduce the computational cost. In each QM/MM model for amorphous aggregate,
one molecule is chosen as the QM region, and the others are treated as the MM region (see
Figure 2.1c). The other details of QM/MM calculations are similar to that in crystal.
2.3  ­Revealed AIE Mechanis 31

The QM calculation in the PCM model can perform in Gaussian 09 [32], Turbomole 6.5 [33],
NW-­Chem [34], and so on. QM/MM calculations could be carried out by the ONIOM model in
Gaussian 16  [32] or through the ChemShell interface  [35]. MD simulation can be done by the
GROMACS package. The DRE, reorganization energy (λ), kr, and kic can be calculated by the
MOMAP program [26, 36].

2.3 ­Revealed AIE Mechanism

AIEgens are non-­ or low-­emissive in dilute solution but highly emissive in the aggregate phase.
The direct characterization of the luminescent property is the fluorescence quantum efficiency
(ΦF), which is related to kr and kic by the following equation: ΦF = kr/(kr + kic). The competition
between kr and kic determines the emission behavior of AIEgens. As shown in Figure 2.2a, both
PESs of the ground state (S0) and first excited state (S1) in the solution are very flat; many vibra-
tional modes with high quantum numbers are activated at high temperatures. These vibrational
modes are strongly coupled with the electron, generating large reorganization energy, and differ-
ent vibrational modes are mixed seriously, the so-­called dramatic DRE. Both the behaviors produce
more nonradiative decay channels from the excited state to the ground state and highly speed up
the nonradiative decay rate, while in the solid state, the compact molecular packing and electro-
static interactions restrict the intramolecular vibrational motions, which cause more steep PESs.
Accordingly, the active vibrational modes and their corresponding mixtures become less favor, and
the couplings between the electron and vibration become weak. Consequently, the nonradiative
decay channels are blocked or even are removed and thus the rate is decreased significantly. A lot
of calculated results demonstrate that aggregation can restrict the couplings between the electron
and a great variety of molecular vibration modes, that is, various kinds of nonradiative decay chan-
nels can be hindered, such as rotating (see Figure 2.2b) [37–40], stretching [41], bending [42], flip-
ping [43], and twisting vibrations [9, 44]. In the following, we overview the revealed AIE mechanism
related to each kind of vibrational normal mode.

2.3.1  Rotating Vibrations of Intramolecular Aromatic Ring


Siloles (silacyclopentadines) are the first kind of AIEgens  [4, 45, 46]. The first silole molecule
where the ΦF in both gas phase and crystal are quantitatively investigated is HPS (Figure 2.3) [37].
HPS is poorly luminescent in cyclohexane at room temperature with a ΦF as low as 0.30%, while
in thin film, the ΦF reaches 78% [45]. Combining QM and QM/MM calculations, the structural
change analysis of optimized structures between S0 and S1 reflects that the torsional angles
between the central silacycle and phenyl groups at the 2,5-­positions show larger modifications
(14.56° and 14.63°) in the isolated state than those (5.85° and 1.05°) in solid phase, suggesting that
the geometric relaxations of the phenyl rings at the 2,5-­positions are largely hindered in the solid
phase. Moreover, the dihedral angle between the central silacycle and the phenyl ring at the
5-­position decreases, which strengthens the intramolecular conjugation, resulting in the increase
of the electric transition dipole moment (μ) from 5.20 to 5.85 Debye (Table  2.1). Hence, upon
aggregation, the calculated kr increases 6 times, while kic decreases about 4 orders of magnitude
(Table 2.1). The nonradiative decay channels are generated by the coupling between the electrons
and thermal vibrations that can be quantified by the molecular reorganization energy
2 2 N
i 1 / 2 i Di for the ith vibrational mode and total i i
, which is very sensitive to the sur-
rounding environment. As seen, λtotal for HPS in the solid phase (403 meV) is obviously smaller
than that in the gas phase (492 meV). The decrease in λtotal mainly stems from the low-­frequency
32 2  Understanding the AIE Mechanism at the Molecular Level

(a)

S1
In solution In solid

IC IC
v=3
E v=1
v=2
v=1 v=0
S0 Restriction of
v=0
non-radiative
decay channel
A F
A F

Nuclear coordinates

(b) (c) (d)

Rotating Stretching Bending


(e) (f) (g)

Flipping Twisting Twisting

Figure 2.2  (a) PES scheme of AIEgens in solution and solid states. (b–g) Overview of vibrational modes
involved in the excited-­state nonradiative decay channels, which contribute to AIE.

vibrational modes (see Figure 2.4a), which are assigned to the rotating vibrations of the phenyl
rings at the 2,5-­position. Thus, the excited-­state energy dissipation pathways are easily restricted
via the decoupling between the electron and low-­frequency out-­of-­plane rotating modes in aggre-
gates (see Figure 2.2b). The corresponding theoretically calculated ΦF in the gas phase and solid
phases is 0.003 and 76% (see Table 2.1), respectively, consistent with experimental results [45].
This well explains the bright emission of HPS in aggregate phase.
A similar AIE mechanism is followed by the other siloles, with the degree of π-­conjugation of the
2,5-­substituents increasing from TPS  [47], BrTPS  [48], HPS  [45], BTPES  [48] to BFTPS  [49]
(Figure 2.3). As shown in Table 2.1, as the conjugation increases, ∆Eg in both gas phase and solid
state decreases regularly. Simultaneously, the μ value increases drastically due to the electron delo-
calization. The balance of excitation energy and μ makes kr to first increase and then levels off as
2.3  ­Revealed AIE Mechanis 33

4 3
5 2 Si
Br Si Si
Si Si Br Si Si

HPS TPS BrTPS BFTPS BTPES

N N S S
3 I II
O O
4 N N N
V
HO OH 1 2 1 1
2 O 2 O
Cu IV III
Cl O O S S
O O
TPA (CAACAd)CuCl CD-7 CD-5 THBDBA COTh

N CN
N CN
H
H N CN
N CN

DSA DCDPP TPBD DCPP HPDMCb

Si Si
S S
S

BtTPS DPTDTP BPS Perylene DSB

N+ N+ N+

S S
N N

Anthracene TTPy TTVP

Figure 2.3  Overview of molecular structures discussed in Sections 2.3 and 2.4.

the extension of conjugation from TPS to BFTPS. After aggregation, λtotal for all AIEgens decreases
obviously and kic decreases by more than 2 orders of magnitude from 9.30 × 105 s−1 for TPS to
1.22 × 108 s−1 for BFTPS. Overall, with the degree of π-­conjugation of the 2,5-­substituent increases,
ΦF in the solid state first increases sharply, then levels off, and finally starts to decrease slightly, as
the competition between kr and kic. Similar to the above discussed HPS, the dramatic decrease of
kic upon aggregation turns the fluorescence on, which is mainly caused by the decoupling of the
electron and low-­frequency rotational normal modes (Figure 2.4a) [38].

2.3.2  Stretching Vibrations of Bonds


Terephthalic acid (TPA) is a representative AIEgen, which could emit both fluorescence and phos-
phorescence simultaneously  [41]. Through a combined QM/MM approach, the photophysical
properties of TPA in both gas phase and crystal were investigated to unravel the effect of crystalli-
zation on the nature of the molecular excited states. It is found that the nature of S1 changes from
(n,π*) in the gas phase to (π,π*) in crystal because the excitation energy of (n,π*) is increased up to
5.05 eV from 4.81 eV, whereas the (π,π*) state is reduced to 4.76 eV from 4.99 eV due to the strong
34 2  Understanding the AIE Mechanism at the Molecular Level

Table 2.1  Calculated HOMO–LUMO energy gap (∆Eg), electric transition dipole moment (μ), total
reorganization energy (λtotal), kr, kic, and ΦF in gas phase and aggregate phase at room temperature for HPS,
TPS, BrTPS, BTPES, and BFTPS, respectively.

λtotal
∆Eg (eV) μ (Debye) (meV) kr (s−1) kic (s−1) ΦF (%)

In gas phase
HPS [37] 3.59 5.20   492 1.05 × 107 3.76 × 1011   0.003
TPS [38] 3.21 0.58 1120 9.30 × 105 1.62 × 1010   0.01
6 9
BrTPS [38] 2.90 1.71 1161 5.55 × 10 5.72 × 10   0.09
HPS [38] 2.62 5.26   891 4.98 × 107 2.53 × 1010   0.20
7 10
BTPES [38] 2.48 6.03   667 6.76 × 10 2.66 × 10   0.25
BFTPS [38] 2.36 9.54   821 1.22 × 108 1.66 × 109   6.86
In solid phase
HPS [37] 3.48 5.85   403 6.56 × 107 2.06 × 107 76.0
6 6
TPS [38] 3.26 0.60   932 1.15 × 10 3.32 × 10 25.8
BrTPS [38] 2.91 1.80   890 8.12 × 106 1.50 × 106 82.3
7 6
HPS [38] 2.70 5.90   753 7.43 × 10 1.57 × 10 97.9
BTPES [38] 2.49 5.75   592 6.57 × 107 1.93 × 106 97.1
8 7
BFTPS [38] 2.30 9.10   607 1.14 × 10 1.07 × 10 91.4

electrostatic interaction upon aggregation. Accordingly, the oscillator strength of S1 is largely


enhanced to 3.49 × 10−2 in the crystalline phase from 3.33 × 10−5 in the gas phase, which recovers
the radiative decay from S1 to S0. And the resultant kr is largely increased by 3 orders of magnitude
from 3.34 × 104 s−1 in the gas phase to 3.43×107 s−1 in the solid phase.
The reorganization energy analysis from S1 → S0 process for each normal mode in the gas phase
reflects that the C=O stretching vibration contributes the most (∼1974.37 cm−1), consistent with the
corresponding large structural modification (0.08 Å between S0 and S1). However, in the solid phase,
this value is reduced to 212.96 cm−1 and the corresponding structural change is only 0.01 Å. Such a
remarkable reduction is due to the alternation to π → π* electronic transition, which is decoupled
with the C=O stretching vibration. As a result, from the gas to solid phases, kic decreases about 1
order of magnitude from 4.97  ×  107 to 5.15  ×  106 s−1. That is, the nonradiative decay process is
blocked by the decoupling between the high-­frequency C=O stretching vibration and transition
electrons (Figures 2.2c and 2.4b). Overall, the largely accelerated radiative decay rate and slowed
nonradiative decay rate induce the observed strong fluorescence in the solid phase [41].

2.3.3  Bending Vibration of Bonds


(CAACAd)CuCl (see Figure 2.3) is a two-­coordinate Cu(I) complex, which exhibits highly efficient
fluorescence in aggregate [50]. The excited-­state decay dynamics for (CAACAd)CuCl in both solu-
tion and solid states is studied by the hybrid QM/MM approach, coupled with the TVCF rate for-
malism [42]. Analyzing the geometrical changes between S0 and S1 upon excitation, it is found that
the complex is more flexible in solution. The largest changes appear in the coordination bond
angles between copper and two ligands with ∠C1−Cu−Cl, ∠Cu−C1−N, and ∠Cu−C1−C4 decreas-
ing from 9.67°, 12.82°, and 10.73° in solution to 3.66°, 4.46°, and 6.75° in the solid phase,
2.3  ­Revealed AIE Mechanis 35

(a) (b)
100 Gas 2000 Solution
Reorganization energy (meV)

Reorganization energy (cm–1)


80
1500
60
1000
40 96 cm–1
20 500
0 0
100 Soild 2000 Soild
80
1500
60
1000
40
20 193 cm–1 500
0 0
0 500 1000 1500 2000 0 500 1000 1500 2000
Normal mode (cm–1) Normal mode (cm–1)

(c) (d)
1200 Solution 400 Gas
Reorganization energy (cm–1)

Reorganization energy (meV) 300


900
600 57.07 cm–1 200
49.85 cm–1
300 100 42 cm–1
1432.20 cm–1
0 0
1200 Solid 400 Soild
300
900
600 200
1449.31 cm–1
120.35 cm–1 100 85 cm–1
300
0 0
0 500 1000 1500 2000 0 500 1000 1500 2000

Normal mode (cm–1) Normal mode (cm–1)

Figure 2.4  Calculated reorganization energies versus normal mode of (a) HPS, (b) TPA, (c) (CAACAd)CuCl,
and (d) COTh in both gas/solution and solid phases, respectively.

respectively. And the transition character changes from metal-­to-­ligand charge transfer (MLCT) to
hybrid MLCT and halogen-­to-­ligand charge transfer (XLCT) upon aggregation. The kr are close in
both phases, while kic decreases by about 3 orders of magnitude from 8.38 × 107 to 1.47 × 104 s−1
(see Table 2.2). The λtotal decreases a lot from solution (7121 cm−1) to solid state (2274 cm−1). It is
found that the largest contributions to λtotal come from several low-­frequency bending modes and
one high-­frequency stretching mode in solution. These low-­frequency modes are assigned to be the
bending vibrations associated with coordination bonds C1−Cu and Cu−Cl, and the high-­frequency AQ4
mode belongs to the stretching vibration of the C−N bond in carbene ligand. Upon aggregation,
the vibrations of angles C1−Cu−Cl and Cu–C1–N are restricted largely, while the stretching vibra-
tions of the C1−N bond are insensitive to the environment (Figure 2.4c). Overall, the strong solid-­
state fluorescence of (CAACAd)CuCl is induced by removing the nonradiative decay channels
owing to the decoupling between transition electron and the bending vibrations of coordination
bonds (see Figure 2.2d).

2.3.4  Flipping Vibrations of Molecular Skeletons


Cyclooctotetraene (COT) is a prototype molecule with nonaromatic annulenes [51–53]. The COT
derivative cyclooctatetrathiophene (COTh) was found to be AIE active and its AIE mechanism was
further investigated theoretically and experimentally [43]. The theoretically optimized geometries
in the S0 and S1 states show that the isolated COTh is much more flexible than that in a cluster. For
36 2  Understanding the AIE Mechanism at the Molecular Level

Table 2.2  Calculated kr, kic, and ΦF in both solution and aggregate phases at room temperature for TPA,
(CAACAd)CuCl, and COTh, respectively.

kr (s−1) kic (s−1) ΦF (%) kr (s−1) kic (s−1) ΦF (%)

In isolated state In solid phase


4 7
TPA [41] 3.34 × 10 4.97 × 10 0.067 3.43 × 107 5.15 × 106 86.95
(CAACAd)CuCl [42] 6.26 × 105 8.38 × 107 0.44 7.83 × 105 1.47 × 104 98.0
4 9 5 7
COTh [43] 9.80 × 10 4.13 × 10 0.002 6.05 × 10 1.87 × 10   3.13

example, from the S0 to S1 state, the modifications of the dihedral angles ΘI–II and ΘII–III (I, II, and
III denote the aromatic rings marked in Figure 2.3) are 22.82° and 24.22° in the gas phase, much
larger than those in the solid state of 12.78° and 12.96°. Upon excitation, the dihedral angles
between the neighboring thienyl rings are decreased from S0 to S1. As shown in Table 2.3, for exam-
ple, ΘI–II decreases from 44.93° to 22.11° in the gas phase, while it reduces from 43.84° to 31.06° in
the solid phase, indicating a better planarity of the central large eight-­membrane ring. Therefore,
the conjugation of the central eight-­membrane improves and the oscillator strength of S1 enhances
accordingly. In addition, the excitation energy at the S1 geometry increases from 2.42 eV in the gas
phase to 2.69 eV in cluster. Therefore, kr increases from 9.80 × 104 to 6.05 × 105 s−1, owing to the
enhancement of oscillator strength and excitation energy; see Table 2.2. The kic is closely related to
the reorganization energy. For COTh, the λtotal decreases to 636 meV in crystal from 860 meV in the
gas phase and then the kic decreases by 2 orders of magnitude. Thus, the ΦF in the crystal is 3 orders
of magnitude larger than that in the isolated state. As shown in Figure 2.4d, the major differences
in the reorganization energies between the gas phase and the solid state are related to the low-­
frequency eight-­member annulene flipping vibrational mode. Thus, the fluorescence of COTh
upon aggregation is induced by the suppression of the nonradiative decay channel contributed by
the electron–vibration coupling between the transition electron and low-­frequency flipping vibra-
tion (see Figure 2.2e).

2.3.5  Twisting Vibration of Molecular Skeletons


Coumarin (chromen-­2-­one) is a natural organic substance that is abundant in many plants. CD-­7
(Figure  2.3) is a coumarin derivative with a seven-­membered aliphatic ring; it has typical AIE
characteristics, while its analogue with a five-­membered aliphatic ring (CD-­5; Figure 2.3) exhib-
ited an opposite ACQ effect  [44]. The electronic structures in S0 and S1 at the level of (TD)
M062X/6-­31G* were performed. The structural changes between S0 and S1 of CD-­7 are obvious.

Table 2.3  Calculation structural parameters of COTh [43] in both gas phase and the crystal (in degree).

S0 S1 ∣∆(S0 − S1)∣ S0 S1 ∣∆(S0 − S1)∣ Crystal

In gas phase In solid phase


ΘI–II 44.93 22.11 22.82 43.84 31.06 12.78 42.85
ΘII–III −48.58 −24.36 24.22 −48.67 −35.71 12.96 −48.83
ΘIII–IV 44.92 22.11 22.81 45.24 32.60 12.64 45.16
ΘI–IV −48.59 −24.36 24.23 −48.94 −35.72 13.22 −46.99
2.4  ­Visualize Calculated Parameters in Experiment 37

During the transition process from S0 to S1, the mean torsion angle between the pyridinone and
chromen-­2-­one fragments changes about 18°, indicating that CD-­7 undergoes out-­of-­plane twist-
ing motions along the C1–C2 bond under photoexcitation, while the corresponding geometric
changes in the solid phase are only 5.8°, which suggests that the intramolecular twisting motion is
largely restricted. In sharp contrast, CD-­5  has a more planar conformation of the conjugated
framework in both the S0 and S1 and changes little in the gas phase, indicating its intrinsic struc-
tural rigidity. In addition, λtotal of CD-­7 reduced from 4028 cm−1 in solution to 3438 cm−1 in aggre-
gates, owing to the restriction of twisting motion in aggregates. Thus, the restriction of twisting
motion in aggregates blocks the nonradiative decay channels and enables CD-­7 to fluoresce
strongly (see Figure 2.2f) [44].
THBDBA contains two pairs of phenyl rings that are covalently linked through a singly bonded
ethyl chain (Figure  2.3); it is almost nonemissive when fully solvated in the THF solution and
starts to emit with the addition of water [9]. In THBDBA crystal, each set of phenyl rings locked by
the ethyl chain adopts a V-­shaped or “boat” conformation. Theoretical calculation indicates that
the single molecule has a λtotal of 6898.6  cm−1, whereas the molecule in a cluster only has
5901.5 cm−1. The normal mode analysis shows that low-­frequency vibrational modes mainly con-
tribute to λtotal for both single molecule and in crystal, but λtotal in the crystal is largely reduced.
Upon aggregation, the restriction of low-­frequency twisting motion blocks the nonradiative decay
channel and turns the fluorescence on (Figure 2.2g) [9].

2.4  ­Visualize Calculated Parameters in Experiments

As shown in Section  8.3, the theoretical calculation is an effective tool for unraveling the deep
insights into experimental phenomenon; we know that the reorganization energy and DRE play
important roles in the AIE mechanism; however, how to validate the theoretically proposed AIE
mechanism and visualize the calculated parameters in experiments? In this section, we will bridge
the calculated parameters and the experimental measurable parameters to validate our above pro-
posed AIE mechanisms.

2.4.1  Stokes Shift vs Reorganization Energy


Three representative AIEgens, 9,10-­distyrylanthracene (DSA) [54], 2,3-­dicyano-­5,6-­diphenylpy
razine (DCDPP) [55], and cis,cis-­1,2,3,4-­tetraphenyl-­1,3-­butadiene (TPBD) [56] in both solution
phase (PCM model) and crystal (QM/MM model), were investigated at the PBE0-­D3(BJ)/6-­31G(d)
level [57]. According to the Franck–Condon principle, the maximum peak of the optical spec-
trum usually appears at the vertical transition point (Figure 2.5a); the Stokes shift (κ) between
the absorption and emission maxima can be regarded as λtotal, which is the sum of the reorgani-
zation energy of the ground (λgs) and excited state (λes). Then, the emission energy can be writ-
ten as Eem = Eabs − λtotal. Upon aggregate from solution, the shift of the emission spectrum can
be regarded as:

Eem
agg sol agg sol agg sol agg sol (6)
Eem Eem Eabs tot

The common red-­shift spectra of the ACQ molecule originate from ΔEagg − sol < 0 for both absorp-
agg sol
tion and emission and the tiny change in reorganization energy ( tot 0 ). However, for
AIEgens, the abnormally blue-­shifted emission occurs from the significant difference
38 2  Understanding the AIE Mechanism at the Molecular Level

(a) (b)

aggr aggr
Eem Eab

S1 λaggr
b
E
λes Aggregate

d
E solu solu
Eab
Eabs Eem
Eem
λsolu
S0
E
c Solution
λgs
aggr–solu = ΔEaggr–solu + λaggr–solu
ΔEem
a ab total
Qg Qe Q

Figure 2.5  (a) The schematic representation of the adiabatic potential energy surface of the S0 and S1, as
well as the corresponding vertical excitation energies. (b) The relative wavelength of AIEgen in both
solution and aggregated state.

environments between the isolated molecule and aggregated phase. In AIEgens, the λtotal in the
agg sol
aggregate is usually much smaller than that in the solution phase, that is, tot 0 , while in
agg sol the absorption spectra of AIEgens in two phases are close to zero. As a result, the emis-
Eabs ,
agg sol
sion spectrum blue-­shifts with Eem 0; see Figure 2.5.
The optical properties and reorganization energies are summarized in Table 2.4. It is clear that,
upon aggregation, the change of absorption spectra is tiny, while the emission spectra blue-­shifted
obviously. Thus, the κsol is much larger than κagg. According to the approximately equivalent relation
sol agg
between κ and λtotal, as shown in Table 2.4, total is larger than total . Thus, the blue-­shift emission is
caused by the larger λtotal in solution than that in aggregate. Structural analysis and reorganization
energy analysis reflect the key role of dihedral angles, and the restriction of low-­frequency rota-
tional or twisting vibrations blocks the nonradiative decay channel and makes the light to emit.

2.4.2  Resonance Raman Spectroscopy (RSS) vs Reorganization Energy


Resonance Raman spectroscopy (RRS) is a spectroscopic technique in which the incident laser
frequency is close to the electronic transition of chromophore. RRS is able to provide the informa-
tion of excited-­state properties. Under the Franck−Condon approximation and resonance condi-
tion, the RRS intensity σ(ωj) from the jth normal mode is proportional to reorganization energy λj
times frequency ωj: σ ∝ ωjλj → σ/ωj ∝ λj [58, 59]. As discussed above, the reorganization energy is
one key factor to determine kic, which plays an important role in the AIE mechanism. In this part,
we use RRS to detect the aggregation effect on the reorganization energy during the nonradiative
process and validates the above-­proposed AIE mechanism.
The AIEgen HPDMCb [60] and AIE-­inactive DCPP [55] (see Figure 2.3) were chosen as models
to show the connection between RRS intensity and the reorganization energy [61]. By combining
PCM and QM/MM models, the photophysical properties of these two compounds were studied in
Table 2.4  The calculated spectral properties and reorganization energies corrected by zero-­point energy for DSA, DCDPP, and TPBD in solution
and aggregate phase, respectively.

Absorption Emission Stokes shift Reorganization energy

agg sol
Unit: eV E ab E ab agg-sol
E ab agg
E em sol
E em agg-sol
E em κsol κagg sol
total
agg
total

DSA [57] 2.82 2.75 0.07 2.18 2.01 0.17 0.74 0.64 0.74 0.64
DCDPP [57] 3.36 3.24 0.12 2.53 2.19 0.34 1.05 0.83 1.05 0.83
TPBD [57] 3.72 3.55 0.17 2.98 2.51 0.47 1.04 0.74 1.03 0.73

Source: © 2014 Royal Society of Chemistry.


40 2  Understanding the AIE Mechanism at the Molecular Level

Table 2.5  Calculated kr, kic, and ΦF in both solution and aggregate phases at room temperature for
HPDMCb and DCPP, respectively.

kr (s−1) kic (s−1) ΦF (%) kr (s−1) kic (s−1) ΦF (%)

In isolated state In solid phase


7 11
HPDMCb [61] 8.64 × 10 1.31 × 10 0.07 7.95 × 107 2.29 × 107 78.0
DCPP [61] 7.98 × 106 1.01 × 106 88.8 3.30 × 106 0.61 × 106 84.4

both solution and solid phases. It is found that the kr of HPDMCb is similar in both phases; while
kic decreases by 4 orders of magnitude from 1.31 × 1011 to 2.29 × 107 s−1 upon aggregation, the cor-
responding ΦF increases to 78.0% from 0.07%, well consistent with the experimental results (see
Table  2.5)  [60]. For AIE-­inactive DCPP, the radiative and nonradiative decay rates and ΦF are
insensitive to the aggregation (see Table 2.5). Analyzing the three factors for the nonradiative decay
rate, it is found that both the nonadiabatic coupling and adiabatic energy gap are almost unchanged
for HPDMCb and DCPP. Moreover, the reorganization energy of DCPP has a little change from
solution to solid phase. Thus, the excited-­state decay process of DCPP is independent on the sur-
rounding environment. Much differently, the reorganization energy of HPDMCb is significantly
reduced from solution to solid phase, as shown in Figure 2.6a, which leads to the large decrease of
kic and the obvious increase of ΦF in the solid phase. From Figure 2.6a, it is also seen that the fre-
quencies of the low-­frequency modes (<100 cm−1) become ca. two to threefold larger upon aggre-
gation; however, the reorganization energy of each mode reduces a lot. These suggest that the
coupling between low-­frequency modes and the transition electrons is strongly decoupled upon
aggregation, leading to the decrease of nonradiative decay rate. With the incident wavelength
equal to the adiabatic excitation energy, the RRS plot of HPDMCb in both solution and solid phases
is shown in Figure 2.6b, and the blue shift of the low-­frequency peaks with the relatively decreased
RRS intensity fully reflects the change character of the reorganization energies. In addition, the
high-­frequency normal modes are almost unresponsive to the environment, as shown in both reor-
ganization energies and RRS signals for AIE-­active HPDMCb. Therefore, the above-­proposed AIE
mechanism is successfully confirmed by RRS signals and the RRS can act as a good detection
means of AIE property.

2.4.3  Isotope Effect vs DRE


Isotope effect (IE) has been widely applied to probe the excited-­state decay process [62, 63]. For con-
ventional luminogens, deuteration always causes the remarkable decrease of kic while almost has no
effect on kr, leading to greatly increased ΦF. It is easy to understand because under the displaced
1 Sn
harmonic oscillator model based on FGR, kic is proportional to e S when assuming one aver-
n!
age accepting-­mode approximation, in which S denotes the Huang–Rhys factor and is determined by
normal mode reorganization energy S = λ/ℏ. The reorganization energy can be obtained from the
calculated excitation energies at the equilibrium geometries of S0 and S1
(λ = ΔES1 − geom − ΔES0 − geom), which is independent of isotopic substitution owing to the same equi-
librium geometries and electronic-­state energy after deuteration. Thus, the decrease of frequency
by deuteration implies the increase of Sj, which largely decreases kic owing to the exponentially pro-
portional relation between kic and –S mentioned above. It should be known that the conventional
2.4  ­Visualize Calculated Parameters in Experiment 41

(a)
180 1624
HPDMCb solution
150
120
24
90
60 1097
51 1679
557
30 78 186 1215
λj (meV)

258
423
0
HPDMCb solid
150 1634
120
90
60 158 1101
141 1685
30 70 205 1218
271 573
434
0
0 200 400 600 800 1000 1200 1400 1600 1800
Normal mode ωi (cm–1)

(b)
1.0 HPDMCb solution 1624
1097
0.8 78
Normalized raman intensity σ (ω)

0.6 51
24
0.4 1008 1679
0.2 186 423 557 1215
258
0.0
1634
1.0 HPDMCb solid
0.8 1101

0.6
0.4 141 158 1218
205 1685
0.2 70 573 1027
271 434
0.0
0 200 400 600 800 1000 1200 1400 1600 1800
Roman shift ω (cm–1)

Figure 2.6  (a) Calculated λj versus ωj in both solution and solid phases for HPDMCb. (b) Calculated
resonance Raman spectroscopy in both solution and solid phases for HPDMCb. Source: Reproduced from
Ref. [61]. Copyright 2015 American Chemical Society.

luminogens have little DRE for their planar and rigid geometries and the above approximation is
suitable for them. While for flexible AIEgens with rotational groups in solution, beyond the displaced
harmonic oscillator model, the mixing of different vibrational modes (so-­called DRE) must be con-
sidered in kic. The DRE always occurs among low-­frequency modes, thus becoming more serious for
much more active low-­frequency modes after deuteration, which sharply increases the kic. Therefore,
the deuteration gives rise to two competitive effects on kic: the normal negative effect through increas-
ing Sj and the abnormally positive effect via the enhancement of DRE (Figure 2.7a) [64]. Thus, it can
also be easily understood that AIEgens exhibit normal negative IE because of the reduction of the
electron–vibration coupling and vibration–vibration mixture upon aggregation. On the basis of the
above, it is inferred that for AIEgens, a more remarkable IE exists in solid phase than in the solution
phase, while for non-­AIEgens, high IE is expected to occur in both solid and solution phases. And we
proposed that the IE can be used as a method to probe the AIE property of emitters and confirm the
mixture of vibrational modes in the excited-­state decay process at the microscopic level.
The kr, kic, and IE of five AIEgens (HPS  [37], BtTPS  [65], HPDMCb  [60], TPBD  [56], and
DPTDTP [66]) and four non-­AIEgens (BPS [67], perylene [68], DSB [69], and anthracene [68]) in
42 2  Understanding the AIE Mechanism at the Molecular Level

(a) (b)
–100 Solution
Solid AIEgens
Isotope effect –80
–60
–40
ωj –20

IE(%)
0
HPS BtTPS HPDMCb TPBD DPTDTP
–80 non-AIEgens
Sj DRE –60
Normal Abnormal –40
negative positive
effect effect –20
kic kic
0
BPS Perylene DSB Anthracene

Figure 2.7  (a) Representation of the isotope effect on kic. (b) IE results for AIEgens and non-­AIEgens.
Source: Reproduced from Ref. [64]. Copyright 2016 The Royal Society of Chemistry.

both solution and solid states were calculated by combining the PCM and hybrid QM/MM models,
respectively [64]. As shown in Table 2.6, it is found that the kic decreases sharply by several orders
of magnitude from the solution to solid phase for AIEgens but suffers a tiny change for
non-­AIEgens. As shown in Figure 2.7b, it is found that (i) the IE for all cases is negative; (ii) for
AIEgens, the IE is minor in solution (~10%) but becomes remarkable in solid phases (~65–95%);
(iii) for non-­AIEgens, the IE results in both solution and solid phases are close to each other and
fall in the range of −40 to −90%. The calculated isotopic characteristic of AIEgens is further verify
by experiments for HPS before and after deuteration; the experimental IEs well reproduce the cal-
culated results with very little IE in solution but remarkably in solid phase. Thus, the proposed IE
scheme further verifies the AIE mechanism theoretically and experimentally in which the nonra-
diative decay channels are blocked greatly owing to the decoupling of vibration–vibration and
electron–vibration and then turning the fluorescence on.

2.4.4  Linear Relationship between Fluorescence Intensity and Amorphous


Aggregate Size
As the major existing form of AIEgens in experiments, the AIE mechanism of amorphous aggregate
is rarely studied theoretically and is a great challenge because of their structural heterogeneity. Here,
we proposed a theoretical protocol combining MD simulations and quantum mechanics/molecular
mechanics calculations, and the TVCF spectrum and rate theory, to quantitatively investigate the
AIE property in amorphous state and explore the relationship between molecular packing, optical
spectra, and fluorescence quantum efficiency [70]. Take HPS [37] as an example; amorphous aggre-
gates with five different sizes were set up in aqueous solution to demonstrate clearly the relationship
between molecular packing, optical spectra, and ΦF, as well as their size effect. MD simulations indi-
cate that HPS aggregates are generally near-­spherical and the packing density of aggregates is size-­
independent and smaller than that of crystals (Figure 2.8a, b). Due to the structural heterogeneity of
the aggregates, AIEgen in amorphous aggregate could not be represented by an aggregate of the
solely fixed conformation. Thus, at each size, five embedded and one exposed QM/MM models were
set up for all sizes of aggregates. As shown in Figure 2.8c, the absorption of amorphous aggregates is
Table 2.6  Calculated room-­temperature kic (s−1) for nondeuterated (H-­all) and fully deuterated (D-­all) isotopomers of the AIEgens and nonAIEgens in both
solution and solid phases.

Solution Solid Solution Solid Solution Solid

HPS BtTPS HPDMCb


H-­all 2.44 × 1011 8.58 × 106 2.20 × 1011 2.73 × 107 1.31 × 1011 2.26 × 107
D-­all 2.22 × 1011 2.61 × 106 1.97 × 1011 6.89 × 106 1.27 × 1011 7.11 × 106
IE –9.0% –69.6% –10.5% –74.8% –3.1% –68.5%
TPBD DPTDTP BPS
H-­all 2.16 × 1010 4.21 × 106 1.79 × 109 6.76 × 105 1.13 × 1010 2.19 × 109
D-­all 1.96 × 1010 2.77 × 105 1.61 × 109 2.23 × 105 7.05 × 109 1.44 × 109
IE –9.3% –93.4% –10.1% –67.0% –37.6% –34.2%
Perylene DSB Anthracene
H-­all 1.19 × 103 0.61 × 103 3.84 × 103 5.51 × 103 0.81 × 103 5.25 × 103
3 3 3 3 3
D-­all 0.29 × 10 0.18 × 10 1.04 × 10 0.66 × 10 0.23 × 10 1.05 × 103
IE –75.6% –70.5% –72.9% –88.0% –71.6% –80.0%

Source: © 2016 Royal Society of Chemistry [64].


44 2  Understanding the AIE Mechanism at the Molecular Level

(a) (c)
1200
Agg-60
Packing density (g/L)

Embedded Exposed
1.0 1.0
900 Abs. Em.

Absorption

Emission
600
0.5 0.5
300
Crystal
0 0.0 0.0
300 450 600 750
Ag l
60

Ag 0
30

Ag 0
10
a

2
st

g-

g-

g-

g-

g-
ry

Wavelength (nm)
Ag

Ag
C

(b) (d)

16 9.53 Å Agg-10 100


7.61 Å Agg-20
80 Embedded
12 10.04 Å Agg-30
FQE (%)

Agg-40 60
RDF

8 Agg-60
40
4 16.31 Å 20 Exposed
0
0
0 4 8 12 16 20 24 28
10

20

30

40

60
g-

g-

g-

g-

g-
Si–Si distance (Å)
Ag

Ag

Ag

Ag

Ag
Figure 2.8  (a) The average packing density of amorphous HPS aggregates. (b) The average RDFs as a
function of the intermolecular Si–Si distance. The lattice constants of the HPS crystal and the
intermolecular distance of two HPS molecules in the unit cell are shown by dashed lines for comparison.
(c) The calculated absorption (left) and emission (right) spectra for five embedded and one exposed HPS for
aggregates with size 60. (d) ΦF of five embedded and one exposed HPS molecule in amorphous aggregates
with different sizes. Source: Reproduced from Ref. [70]. Copyright 2016 The Royal Society of Chemistry.

similar to that of crystals, while the emission is clearly red-­shifted from the crystalline one. This can
be explained by the lower-­packing-­density-­induced larger reorganization energies in amorphous
HPS aggregates than that in crystal. Strikingly, ΦF of the embedded HPS is size-­independent and
about 1 or 2 orders of magnitude larger than those of the exposed ones (<7%), see Figure 2.8d. In
addition, kr is insensitive to environments, kic is very sensitive to the environment, and kic of embed-
ded HPS molecules are 2–4 orders of magnitude smaller than those of the exposed ones because of
the restriction of the intramolecular low-­frequency rotational motions of embedded HPS molecules.
More excitingly, the size-­independent ΦF predicts a linear relationship between the fluorescence
intensity and aggregate size of the nanosized HPS aggregates. This indicates that the electron–vibra-
tion and vibration–vibration coupling are dependent on the compactness of molecular packing other
than the size of the aggregate. In addition, the crystallization-­enhanced emission phenomenon in the
experiment can be explained by the calculated blue-­shifted emission and enhanced fluorescence
intensity of HPS from amorphous aggregate to the crystalline phase.

2.4.5  Pressure-­induced Enhanced Emission (PIEE)


As discussed above, the electron–vibration and vibration–vibration couplings are sensitive to molec-
ular packing. If so, they are sure to further decouple to larger extent under external pressure because
of the more densely molecular packing and then enhancing the solid-­phase fluorescent emission,
called pressure-­induced emission enhancement (PIEE) [71–73]. Take HPS [37] as an example; the
2.5 ­Molecular Design Based on AIE Mechanis 45

Figure 2.9  (a) Reorganization energies (a)


λg(e) and excitonic couplings J of HPS in 400
aggregates at different pressures. λg λe J
(b) Calculated kic and ΦF of HPS in 350
aggregates at different pressures. Source: 300
Reproduced from Ref. [74]. Copyright 2018

Energy (meV)
The Royal Society of Chemistry. 250

150

100

50

0
0.0 1.5 3.0 4.5 6.0 7.5 9.0 10.5
Pressure (GPa)

(b)
100 1.8
kic ΦF
5.06 GPa
1.5
Quantum efficiency ΦF (%)

95
1.2

kic (107 s–1)


90 0.9

0.6
85
0.3
80
0.0
0.0 1.5 3.0 4.5 6.0 7.5 9.0 10.5
Pressure (GPa)

mechanism behind PIEE was studied by combining DFT-­D crystalline-­structure simulations, QM/
MM, and TVCF formalism  [74]. It is found that upon compression, the volume is continuously
reduced, and the intermolecular C–H. . .π hydrogen bond, steric hindrance, and π–π interactions are
increased accordingly. The torsional motions of phenyl rings at the 2,5-­positions are firstly hindered
from 5.06 to 10.86 GPa and then slightly active as pressure increases. As expected, with the increase
of the pressure, the reorganization energy λg(e) is gradually decreased until it reaches a saturation
value, while the intermolecular excitonic coupling (J) changes a little, and it is very small and would
have no effect on the optical properties; see Figure 2.9a. More excitedly, the kic is expected to reduce
quickly first when pressure increases from 0 to 5.06 GPa and then levels off beyond 5.06 GPa, as
shown in Figure 2.9b. As expected, the kr of HPS is insensitive to pressure because it is not mainly
controlled by the reorganization energy  [13]. Accordingly, ΦF rises rapidly with pressure up to
5.06 GPa but tends to level off at higher pressures. The calculated results are well consistent with the
experiments [71]. Overall, the PIEE fully supports the origin of the AIE mechanism proposed above.

2.5  ­Molecular Design Based on AIE Mechanism


On the basis of the above proposed AIE mechanism, we further rationally design a series of AIE
probes for specific biological applications. Two representative amphiphilic AIEgens, TTVP  [75]
and TTPy  [76], with similar structures were chosen as examples (Figure  2.3). As shown in
46 2  Understanding the AIE Mechanism at the Molecular Level

Figure 2.3, the subtle difference in the chemical structures of them leads to dramatically different
behavior in cell imaging. TTPy can penetrate the cell membrane and target mitochondria with
excellent fluorescence imaging [76], while TTVP is able to specifically target and “light up” the
lipid membrane [75]. Through a theoretical protocol combining large-­scale MD simulations and
the hybrid QM/MM model, the mechanism behind their strikingly different fluorescence imaging
behaviors was revealed [77]. The free-­energy profiles indicate that crossing the hydrophobic core
of the lipid membrane, the free-­energy barrier for TTVP (30.44 kJ/mol) is much larger than that for
TTPy (19.04 kJ/mol). The calculated permeability coefficient of TTPy (7.685 cm/s) is ca. 213 times
larger than that of TTVP (3.612 × 10−2 cm/s), implying that TTPy could translocate through the
lipid membrane and target mitochondria, while TTVP can target cell membrane.
As shown in Figure 2.10a, the QM/MM model of TTVP in the lipid membrane is set up to study
its photophysical properties and to unravel the behind AIE mechanism. In both solution and lipid
membrane, the quaternary 3-­(trimethylammonio)propyl group does not contribute to natural tran-
sition orbitals and thus no effect on the emission property (Figure 2.10b, c). Upon excitation from
S0 to S1, the largest change of the dihedral angles is decreased from 24.8° to less than 4.7°, and the
largest modification of the bond length is decreased from 0.07 to 0.02 Å (Figure 2.10d, e), indicat-
ing the stronger rigidity of TTVP in the lipid membrane. The calculated kr and kic in the THF solu-
tion are 3.3 × 108 and 2.4 × 1010 s−1, respectively; thus, ΦF is 1.3%, consistent with the nonemissive
phenomenon of TTVP in dilute THF solution. While in lipid membrane, kr of TTVP is about
3.6 × 108 to 3.7×108 s−1, which is a little bit larger than that in solution, in addition, the λtotal of
TTVP in the lipid membrane is much smaller than that in solution; thus, kic in the lipid membrane
should also be small (Figure 2.10f). In addition, the emission of TTVP in the lipid membrane is
blue-­shifted than in the THF solution (Figure 2.10g), consistent with the experimental result [75].
That is to say, the enhancement of kr and the decrease of kic turn the fluorescence on in the lipid
membrane. Thus, the design strategy of AIEgen-­based fluorescent probes proposed that it is
amphiphilic AIEgens, including AIE-­featured tail group and head group with different positive
charges, to target the lipid membrane or mitochondria selectively. According to the design princi-
ple, four amphiphilic AIEgens were designed. Two of the designed AIEgens are composed of the
AIE-­core and pyridine group, having TTPy-­analogue characteristics, while the other two AIEgens
are structurally similar to TTVP, considering the inclusion of a functional group with two positive
charges. The cell image experiments indicate that the newly designed TTPy-­analogues can translo-
cate through the cell membrane and target mitochondria, while the TTVP analogues can selec-
tively light up the cell membrane. The experimental results verified the theoretical prediction.

2.6  ­Summary and Outlook

In this chapter, the detailed AIE mechanism is discussed at the molecular level by combining mul-
tiscale modeling with TVCF formalism based on FGR using the homebuilt MOMAP program
package. The solvent effect and aggregate effect were considered by using PCM and QM/MM mod-
els, respectively. We proposed that the restriction of the nonradiative decay channel by compact
molecular packing and electrostatic interaction in aggregate state turns the fluorescence on. The
restriction of the nonradiative decay channel occurs by the electron–vibration decoupling and the
vibration–vibration demixing among various kinds of normal modes, such as rotating, stretching,
bending, flipping, and twisting vibrations. To validate the proposed AIE mechanism, we build the
relationship between the calculated photophysical parameters and the experimentally visualizable
signals. (i) The reorganization energy is comparable with the Stokes shift between the absorption
(a) (b) (d)
ΔD4 = 24.8°
ΔB7 = 0.07 Å

In THF solution
93% in THF solution (e)
ΔD4 = 4.7°
ΔB7 = 0.02 Å

In lipid membrane
(c)
(f) (g)
750
1.0
600
0.8
450
0.6
95% in lipid membrane 300
0.4
150
0.2
Emission intensity

THF solution
0 Membrane
Reorganization energy (meV)

F d 0.0
TH tion Lipi rane 450 600 750 900
l u mb
so Wavelength (nm)
Me

Figure 2.10  (a) The QM/MM model of TTVP in lipid membrane. (b, c) Calculated NTOs for TTVP at S1 states in both lipid membrane and dilute THF
solution. (d, e) Superposition of optimized structures at both S0 and S1 states for TTVP in both lipid membrane and dilute THF solution. Calculated
(f) total reorganization energies and (g) fluorescence emission spectrum of TTVP in both lipid membrane and dilute THF solution. Source: Reproduced
from Ref. [77]. Copyright 2019 The Royal Society of Chemistry.
48 2  Understanding the AIE Mechanism at the Molecular Level

and emission spectra, the decrease of Stokes shift induced by the decreasing reorganization energy,
which causes blue-­shifted emission upon aggregation. (ii) The reorganization energy of each nor-
mal mode is proportional to the ratio of the RRS intensity and frequency of the corresponding
mode. (iii) The DRE, which plays an important role in the nonradiative decay process in flexible
systems, can be validated by IE. (iv) The fluorescence efficiency is independent of the size of the
aggregate, and the fluorescence intensity can be quantified by the aggregate size because of the
linear relationship between them. (v) The pressure can further weaken the electron–vibration
­coupling until reaching a certain value and gradually enhance the solid-­phase fluorescence to a
maximum. The experiments could be well reproduced by theoretical calculation quantitatively.
Combining the proposed AIE mechanism, the deep insights into the different bioimaging behav-
iors of two representative AIEgens with similar structures are unraveled. Based on this, we
­proposed the rational design principles for AIE probes to selectively target the lipid membrane or
mitochondria, which is confirmed by the cell imaging study in experiments.
Accurately describing the structure of excited state and decay processes is still a long-­term
challenge computationally because the electron–electron correlation and electron–vibration
coupling are required to be involved [78]. The anharmonic effect and extremely strong vibronic
coupling beyond the Franck–Condon region are not considered in the current theoretical
approaches, although it is found that the excitonic coupling has very minor effect on the nonra-
diative decay rate with only 12–33% enhancement for typical AIEgens [79, 80]. The effect of the
intermolecular excitonic coupling should be taken into account, as well as the intermolecular
charge transfer and excimer, for more compact molecular packing with short distance or strong
π–π stacking. In addition, polarizable force field needs to be involved in the QM/MM protocol to
consider the electron-­density change of the QM molecule in the excited state. All these are being
actively pursued.

­Acknowledgments

This work is supported by the National Natural Science Foundation of China (Grants 21973099,
21803007) and Beijing Institute of Technology Research Fund Program for Young Scholars.

­References
1 Briks, J. B. (1970). Photophysics of Aromatic Molecules. Wiley: London.
Mei, J.; Leung, N. L. C.; Kwok, R. T. K.; Lam, J. W. Y.; Tang, B. Z. (2015). Aggregation-­induced
2
emission: together we shine, united we soar! Chem. Rev., 115 (21), 11718–11940.
3 Mei, J.; Hong, Y.; Lam, J. W. Y.; Qin, A.; Tang, Y.; Tang, B. Z. (2014). Aggregation-­induced emission:
the whole is more brilliant than the parts. Adv. Mater. 26 (31), 5429–5479.
4 Luo, J.; Xie, Z.; Lam, J. W. Y.; Cheng, L.; Chen, H.; Qiu, C.; Kwok, H. S.; Zhan, X.; Liu, Y.; Zhu, D.;
Tang, B. Z. (2001). Aggregation-­induced emission of 1-­methyl-­1,2,3,4,5-­pentaphenylsilole. Chem.
Commun. (18), 1740–1741.
5 Oelkrug, D.; Tompert, A.; Gierschner, J.; Egelhaaf, H.-­J.; Hanack, M.; Hohloch, M.; Steinhuber,
E. (1998). Tuning of fluorescence in films and nanoparticles of oligophenylenevinylenes. J. Phys.
Chem. B 102 (11), 1902–1907.
6 Hong, Y.; Lam, J. W. Y.; Tang, B. Z. (2009). Aggregation-­induced emission: phenomenon, mechanism
and applications. Chem. Commun. 4332–4353.
  ­Reference 49

7 Chen, J.; Law, C. C. W.; Lam, J. W. Y.; Dong, Y.; Lo, S. M. F.; Williams, I. D.; Zhu, D.; Tang, B. Z.,
(2003). Synthesis, light emission, nanoaggregation, and restricted intramolecular rotation of
1,1-­substituted 2,3,4,5-­tetraphenylsiloles. Chem. Mater. 15 (7), 1535–1546.
8 Dong, Y.; Lam, J. W. Y.; Qin, A.; Sun, J.; Liu, J.; Li, Z.; Sun, J.; Sung, H. H. Y.; Williams, I. D.; Kwok,
H. S.; Tang, B. Z. (2007). Aggregation-­induced and crystallization-­enhanced emissions of
1,2-­diphenyl-­3,4-­bis(diphenylmethylene)-­1-­cyclobutene. Chem. Commun. 3255–3257.
9 Leung, N. L. C.; Xie, N.; Yuan, W.; Liu, Y.; Wu, Q.; Peng, Q.; Miao, Q.; Lam, J. W. Y.; Tang,
B. Z. (2014). Restriction of intramolecular motions: the general mechanism behind aggregation-­
induced emission. Chem. Eur. J. 20 (47), 15349–15353.
10 Shuai, Z.; Peng, Q. (2014). Excited states structure and processes: understanding organic light-­
emitting diodes at the molecular level. Phys. Rep. 537 (4), 123–156.
11 Peng, Q.; Niu, Y.; Shi, Q.; Gao, X.; Shuai, Z. (2013). Correlation function formalism for triplet
excited state decay: combined spin–orbit and nonadiabatic couplings. J. Chem. Theory Comput. 9
(2), 1132–1143.
12 Niu, Y.; Peng, Q.; Deng, C.; Gao, X.; Shuai, Z. (2010). Theory of excited state decays and optical
spectra: application to polyatomic molecules. J. Phys. Chem. A 114 (30), 7817–7831.
13 Peng, Q.; Yi, Y.; Shuai, Z.; Shao, J. (2007). Toward quantitative prediction of molecular fluorescence
quantum efficiency: role of Duschinsky rotation. J. Am. Chem. Soc. 129 (30), 9333–9339.
14 Peng, Q.; Yi, Y.; Shuai, Z.; Shao, J. (2007). Excited state radiationless decay process with Duschinsky
rotation effect: formalism and implementation. J. Chem. Phys. 126 (11), 114302.
15 Li, Q.; Blancafort, L. (2013). A conical intersection model to explain aggregation induced emission
in diphenyl dibenzofulvene. Chem. Commun. 49 (53), 5966–5968.
16 Crespo-­Otero, R.; Li, Q.; Blancafort, L. (2019). Exploring potential energy surfaces for aggregation-­
induced emission—­from solution to crystal. Chem. Asian J. 14 (6), 700–714.
17 Liu, Y.; Tao, X.; Wang, F.; Shi, J.; Sun, J.; Yu, W.; Ren, Y.; Zou, D.; Jiang, M. (2007). Intermolecular
hydrogen bonds induce highly emissive excimers: enhancement of solid-­state luminescence.
J. Phys. Chem. C 111 (17), 6544–6549.
18 Huang, W.; Sun, L.; Zheng, Z.; Su, J.; Tian, H. (2015). Colour-­tunable fluorescence of single
molecules based on the vibration induced emission of phenazine. Chem. Commun. 51 (21),
4462–4464.
19 Chung, J. W.; Yoon, S.-­J.; An, B.-­K.; Park, S. Y. (2013). High-­contrast on/off fluorescence switching
via reversible E–Z isomerization of diphenylstilbene containing the α-­cyanostilbenic moiety.
J. Phys. Chem. C 117 (21), 11285–11291.
20 Yang, L.; Ye, P.; Li, W.; Zhang, W.; Guan, Q.; Ye, C.; Dong, T.; Wu, X.; Zhao, W.; Gu, X.; Peng, Q.;
Tang, B.; Huang, H. (2018). Uncommon aggregation-­induced emission molecular materials with
highly planar conformations. Adv. Optical Mater. 6 (9), 1701394.
21 Tu, Y.; Liu, J.; Zhang, H.; Peng, Q.; Lam, J. W. Y.; Tang, B. Z. (2019). Restriction of access to the
dark state: a new mechanistic model for heteroatom-­containing AIE systems. Angew. Chem. Int.
Ed. 58 (42), 14911–14914.
22 Yang, S.; Yin, P.-­A.; Li, L.; Peng, Q.; Gu, X.; Gao, G.; You, J.; Tang, B. Z. (2020). Crystallization-­
induced reversal from dark to bright excited states for construction of solid-­emission-­tunable
squaraines. Angew. Chem. Int. Ed. 59(25), 10136–10142. doi: 10.1002/anie.201914437.
23 Lin, S. H.; Chang, C. H.; Liang, K. K.; Chang, R.; Shiu, Y. J.; Zhang, J. M.; Yang, T. S.; Hayashi, M.;
Hsu, F. C. (2002) Ultrafast dynamics and spectroscopy of bacterial photosynthetic reaction centers.
Adv. Chem. Phys. 121, 1–88.
24 Niu, Y.; Peng, Q.; Shuai, Z. (2008). Promoting-­mode free formalism for excited state radiationless
decay process with Duschinsky rotation effect. Sci. China Ser. B Chem., 51 (12), 1153–1158.
50 2  Understanding the AIE Mechanism at the Molecular Level

25 Peng, Q.; Yi, Y.; Shuai, Z.; Shao, J. (2007). Excited state radiationless decay process with Duschinsky
rotation effect: formalism and implementation. J. Chem. Phys. 126 (11), 114302.
26 Niu, Y.; Li, W.; Peng, Q.; Geng, H.; Yi, Y.; Wang, L.; Nan, G.; Wang, D.; Shuai, Z. (2018). MOlecular
MAterials Property Prediction Package (MOMAP) 1.0: a software package for predicting the
luminescent properties and mobility of organic functional materials. Mol. Phys. 116 (7–8),
1078–1090.
27 Wang, J.; Wolf, R. M.; Caldwell, J. W.; Kollman, P. A.; Case, D. A. (2004). Development and testing
of a general amber force field. J. Comput. Chem. 25 (9), 1157–1174.
28 Aaqvist, J.; Warshel, A. (1993). Simulation of enzyme reactions using valence bond force fields and
other hybrid quantum/classical approaches. Chem. Rev. 93 (7), 2523–2544.
29 Lin, H.; Truhlar, D. (2007). QM/MM: what have we learned, where are we, and where do we go
from here? Theor. Chem. Acc. 117 (2), 185–199.
30 Bussi, G.; Donadio, D.; Parrinello, M. (2007). Canonical sampling through velocity rescaling.
J. Chem. Phys. 126 (1), 014101.
31 Parrinello, M.; Rahman, A. (1981). Polymorphic transitions in single crystals: a new molecular
dynamics method. J. Appl. Phys., 52 (12), 7182–7190.
32 Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.;
Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G. A.; Nakatsuji, H.; Caricato, M.; Hratchian,
H. P.; Izmaylov, A. F.; Bloino, J.; Zheng, G.; Sonnenberg, J. L.; Hada, M.; Ehara, M.; Toyota, K.;
Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Vreven, T.;
Montgomery, J., J. A.; Peralta, J. E.; Ogliaro, F.; Bearpark, M.; Heyd, J. J.; Brothers, E.; Kudin, K. N.;
Staroverov, V. N.; Kobayashi, R.; Normand, J.; Raghavachari, K.; Rendell, A.; Burant, J. C.; Iyengar,
S. S.; Tomasi, J.; Cossi, M.; Rega, N.; Millam, N. J.; Klene, M.; Knox, J. E.; Cross, J. B.; Bakken, V.;
Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.; Cammi, R.;
Pomelli, C.; Ochterski, J. W.; Martin, R. L.; Morokuma, K.; Zakrzewski, V. G.; Voth, G. A.; Salvador,
P.; Dannenberg, J. J.; Dapprich, S.; Daniels, A. D.; Farkas, Ö.; Foresman, J. B.; Ortiz, J. V.;
Cioslowski, J.; Fox, D. J. (2009). Gaussian, Inc.: Wallingford, CT.
33 Ahlrichs, R.; Bär, M.; Häser, M.; Horn, H.; Kölmel, C. (1989). Electronic structure calculations on
workstation computers: the program system turbomole. Chem. Phys. Lett. 162 (3), 165–169.
34 Valiev, M.; Bylaska, E. J.; Govind, N.; Kowalski, K.; Straatsma, T. P.; Van Dam, H. J. J.; Wang, D.;
Nieplocha, J.; Apra, E.; Windus, T. L.; de Jong, W. A. (2010). NWChem: a comprehensive and
scalable open-­source solution for large scale molecular simulations. Comput. Phys. Commun. 181
(9), 1477–1489.
35 Metz, S.; Kästner, J.; Sokol, A. A.; Keal, T. W.; Sherwood, P. (2014). ChemShell—­a modular
software package for QM/MM simulations. WIRES Comput. Mol. Sci. 4 (2), 101–110.
36 Shuai, Z.; Peng, Q.; Niu, Y.; Geng, H. (2014). MOMAP, a free and open-­source molecular materials
property prediction package. Revision 0.2.004; available online: http://www.shuaigroup.net/,
Beijing,China.
37 Zhang, T.; Jiang, Y.; Niu, Y.; Wang, D.; Peng, Q.; Shuai, Z. (2014). Aggregation effects on the optical
emission of 1,1,2,3,4,5-­hexaphenylsilole (HPS): a QM/MM study. J. Phys. Chem. A 118 (39),
9094–9104.
38 Xie, Y.; Zhang, T.; Li, Z.; Peng, Q.; Yi, Y.; Shuai, Z. (2015). Influences of conjugation extent on the
aggregation-­induced emission quantum efficiency in silole derivatives: a computational study.
Chem. Asian J. 10 (10), 2154–2161.
39 Wu, Q.; Deng, C.; Peng, Q.; Niu, Y.; Shuai, Z. (2012). Quantum chemical insights into the
aggregation induced emission phenomena: a QM/MM study for pyrazine derivatives. J. Comput.
Chem. 33 (23), 1862–1869.
  ­Reference 51

40 Wu, Q.; Peng, Q.; Niu, Y.; Gao, X.; Shuai, Z. (2012). Theoretical insights into the aggregation-­
induced emission by hydrogen bonding: a QM/MM study. J. Phys. Chem. A 116 (15), 3881–3888.
41 Ma, H.; Shi, W.; Ren, J.; Li, W.; Peng, Q.; Shuai, Z. (2016). Electrostatic interaction-­induced
room-­temperature phosphorescence in pure organic molecules from QM/MM calculations. J. Phys.
Chem. Lett. 7 (15), 2893–2898.
42 Lin, S.; Peng, Q.; Ou, Q.; Shuai, Z. (2019). Strong solid-­state fluorescence induced by restriction of
the coordinate bond bending in two-­coordinate copper(I)–carbene complexes. Inorg. Chem. 58
(21), 14403–14409.
43 Zhao, Z.; Zheng, X.; Du, L.; Xiong, Y.; He, W.; Gao, X.; Li, C.; Liu, Y.; Xu, B.; Zhang, J.; Song, F.; Yu,
Y.; Zhao, X.; Cai, Y.; He, X.; Kwok, R. T. K.; Lam, J. W. Y.; Huang, X.; Phillips, D. L.; Wang, H.;
Tang, B. Z. (2019). Non-­aromatic annulene-­based aggregation-­induced emission system via
aromaticity reversal process. Nat. Commun. 10 (1), 2952.
44 Bu, F.; Duan, R.; Xie, Y.; Yi, Y.; Peng, Q.; Hu, R.; Qin, A.; Zhao, Z.; Tang, B. Z. (2015). Unusual
aggregation-­induced emission of a coumarin derivative as a result of the restriction of an
intramolecular twisting motion. Angew. Chem. Int. Ed. 54 (48), 14492–14497.
45 Yu, G.; Yin, S.; Liu, Y.; Chen, J.; Xu, X.; Sun, X.; Ma, D.; Zhan, X.; Peng, Q.; Shuai, Z.; Tang, B.; Zhu,
D.; Fang, W.; Luo, Y. (2005). Structures, electronic states, photoluminescence, and carrier transport
properties of 1,1-­disubstituted 2,3,4,5-­tetraphenylsiloles. J. Am. Chem. Soc. 127 (17), 6335–6346.
46 Tang, B. Z.; Zhan, X.; Yu, G.; Sze Lee, P. P.; Liu, Y.; Zhu, D. (2001). Efficient blue emission from
siloles. J. Mater. Chem. 11 (12), 2974–2978.
47 Zhao, E.; Lam, J. W. Y.; Hong, Y.; Liu, J.; Peng, Q.; Hao, J.; Sung, H. H. Y.; Williams, I. D.; Tang,
B. Z. (2013). How do substituents affect silole emission? J. Mater. Chem. C 1 (36), 5661–5668.
48 Zhao, Z.; Liu, D.; Mahtab, F.; Xin, L.; Shen, Z.; Yu, Y.; Chan, C. Y. K.; Lu, P.; Lam, J. W. Y.; Sung,
H. H. Y.; Williams, I. D.; Yang, B.; Ma, Y.; Tang, B. Z. (2011). Synthesis, structure, aggregation-­
induced emission, self-­assembly, and electron mobility of 2,5-­bis(triphenylsilylethynyl)-­3,4-­
diphenylsiloles. Chem. Eur. J. 17 (21), 5998–6008.
49 Zhan, X.; Haldi, A.; Risko, C.; Chan, C. K.; Zhao, W.; Timofeeva, T. V.; Korlyukov, A.; Antipin,
M. Y.; Montgomery, S.; Thompson, E.; An, Z.; Domercq, B.; Barlow, S.; Kahn, A.; Kippelen, B.;
Brédas, J.-­L.; Marder, S. R. (2008). Fluorenyl-­substituted silole molecules: geometric, electronic,
optical, and device properties. J. Mater. Chem. 18 (26), 3157–3166.
50 Romanov, A. S.; Di, D.; Yang, L.; Fernandez-­Cestau, J.; Becker, C. R.; James, C. E.; Zhu, B.;
Linnolahti, M.; Credgington, D.; Bochmann, M. (2016). Highly photoluminescent copper carbene
complexes based on prompt rather than delayed fluorescence. Chem. Commun. 52 (38),
6379–6382.
51 Karadakov, P. B. (2008). Aromaticity and antiaromaticity in the low-­lying electronic states of
cyclooctatetraene. J. Phys. Chem. A 112 (49), 12707–12713.
52 Feixas, F.; Vandenbussche, J.; Bultinck, P.; Matito, E.; Solà, M. (2011). Electron delocalization and
aromaticity in low-­lying excited states of archetypal organic compounds. Phys. Chem. Chem. Phys.
13 (46), 20690–20703.
53 Gogonea, V.; Schleyer, P. v. R.; Schreiner, P. R. (1998). Consequences of triplet aromaticity in
4nπ-­electron annulenes: calculation of magnetic shieldings for open-­shell species. Angew. Chem.
Int. Ed. 37 (13-­14), 1945–1948.
54 Dong, Y.; Xu, B.; Zhang, J.; Lu, H.; Wen, S.; Chen, F.; He, J.; Li, B.; Ye, L.; Tian, W. (2012).
Supramolecular interactions induced fluorescent organic nanowires with high quantum yield
based on 9,10-­distyrylanthracene. CrystEngComm 14 (20), 6593–6598.
55 Qin, A.; Lam, J. W. Y.; Mahtab, F.; Jim, C. K. W.; Tang, L.; Sun, J.; Sung, H. H. Y.; Williams, I. D.;
Tang, B. Z. (2009). Pyrazine luminogens with “free” and “locked” phenyl rings: understanding of
52 2  Understanding the AIE Mechanism at the Molecular Level

restriction of intramolecular rotation as a cause for aggregation-­induced emission. Appl. Phys. Lett.
94 (25), 253308.
56 Chen, J.; Xu, B.; Ouyang, X.; Tang, B. Z.; Cao, Y. (2004). Aggregation-­induced emission of
cis,cis-­1,2,3,4-­tetraphenylbutadiene from restricted intramolecular rotation. J. Phys. Chem. A 108
(37), 7522–7526.
57 Wu, Q.; Zhang, T.; Peng, Q.; Wang, D.; Shuai, Z. (2014). Aggregation induced blue-­shifted
emission -­the molecular picture from a QM/MM study. Phys. Chem. Chem. Phys. 16 (12),
5545–5552.
58 Heller, E. J.; Sundberg, R.; Tannor, D. (1982). Simple aspects of Raman scattering. J. Phys. Chem. 86
(10), 1822–1833.
59 Santoro, F.; Cappelli, C.; Barone, V. (2011). Effective time-­independent calculations of vibrational
resonance Raman spectra of isolated and solvated molecules including Duschinsky and Herzberg–
Teller effects. J. Chem. Theory Comput. 7 (6), 1824–1839.
60 Dong, Y.; Lam, J. W. Y.; Qin, A.; Sun, J.; Liu, J.; Li, Z.; Sun, J.; Sung, H. H. Y.; Williams, I. D.; Kwok,
H. S.; Tang, B. Z. (2007). Aggregation-­induced and crystallization-­enhanced emissions of
1,2-­diphenyl-­3,4-­bis(diphenylmethylene)-­1-­cyclobutene. Chem. Commun. (31), 3255–3257.
61 Zhang, T.; Ma, H.; Niu, Y.; Li, W.; Wang, D.; Peng, Q.; Shuai, Z.; Liang, W. (2015). Spectroscopic
signature of the aggregation-­induced emission phenomena caused by restricted nonradiative
decay: a theoretical proposal. J. Phys. Chem. C 119 (9), 5040–5047.
62 Lin, S. H.; Bersohn, R. (1968). Effect of partial deuteration and temperature on triplet-­state
lifetimes. J. Chem. Phys. 48 (6), 2732–2736.
63 Saltiel, J.; Waller, A. S.; Sears, D. F.; Garrett, C. Z. (1993). Fluorescence quantum yields of trans-­
stilbene-­d0 and -­d2 in n-­hexane and n-­tetradecane: medium and deuterium isotope effects on
decay processes. J. Phys. Chem. 97 (11), 2516–2522.
64 Zhang, T.; Peng, Q.; Quan, C.; Nie, H.; Niu, Y.; Xie, Y.; Zhao, Z.; Tang, B. Z.; Shuai, Z. (2016). Using
the isotope effect to probe an aggregation induced emission mechanism: theoretical prediction and
experimental validation. Chem. Sci. 7 (8), 5573–5580.
65 Chen, J.; Xu, B.; Yang, K.; Cao, Y.; Sung, H. H. Y.; Williams, I. D.; Tang, B. Z. (2005).
Photoluminescence spectral reliance on aggregation order of 1,1-­bis(2′-­thienyl)-­2,3,4,5-­
tetraphenylsilole. J. Phys. Chem. B 109 (36), 17086–17093.
66 Zhang, X.; Sørensen, J. K.; Fu, X.; Zhen, Y.; Zhao, G.; Jiang, L.; Dong, H.; Liu, J.; Shuai, Z.; Geng,
H.; Bjørnholm, T.; Hu, W. (2014). Rubrene analogues with the aggregation-­induced emission
enhancement behaviour. J. Mater. Chem. C 2 (5), 884–890.
67 Yoshiharu, N.; Chitoshi, K.; Hiroyuki, K.; Takeshi, K. (2011). Diacenaphtho[1,2-­b;1′,2′-­d]silole and
-­pyrrole. Chem. Lett. 40 (12), 1437–1439.
68 Katoh, R.; Suzuki, K.; Furube, A.; Kotani, M.; Tokumaru, K. (2009). Fluorescence quantum yield of
aromatic hydrocarbon crystals. J. Phys. Chem. C 113 (7), 2961–2965.
69 Varghese, S.; Park, S. K.; Casado, S.; Fischer, R. C.; Resel, R.; Milián-­Medina, B.; Wannemacher, R.;
Park, S. Y.; Gierschner, J. (2013). Stimulated emission properties of sterically modified
distyrylbenzene-­based H-­aggregate single crystals. J. Phys. Chem. Lett. 4 (10), 1597–1602.
70 Zheng, X.; Peng, Q.; Zhu, L.; Xie, Y.; Huang, X.; Shuai, Z. (2016). Unraveling the aggregation effect
on amorphous phase AIE luminogens: a computational study. Nanoscale 8, 15173–15180.
71 Fan, X.; Sun, J.; Wang, F.; Chu, Z.; Wang, P.; Dong, Y.; Hu, R.; Tang, B. Z.; Zou, D. (2008).
Photoluminescence and electroluminescence of hexaphenylsilole are enhanced by pressurization
in the solid state. Chem. Commun. (26), 2989–2991.
72 Gu, Y.; Wang, K.; Dai, Y.; Xiao, G.; Ma, Y.; Qiao, Y.; Zou, B. (2017). Pressure-­induced emission
enhancement of carbazole: the restriction of intramolecular vibration. J. Phys. Chem. Lett. 8 (17),
4191–4196.
  ­Reference 53

73 Yuan, H.; Wang, K.; Yang, K.; Liu, B.; Zou, B. (2014). Luminescence properties of compressed
tetraphenylethene: the role of intermolecular interactions. J. Phys. Chem. Lett. 5 (17), 2968–2973.
74 Zhang, T.; Shi, W.; Wang, D.; Zhuo, S.; Peng, Q.; Shuai, Z. (2018). Pressure-­induced emission
enhancement in hexaphenylsilole: a computational study. J. Mater. Chem. C 7, 1388–1398.
75 Wang, D.; Su, H.; Kwok, R. T. K.; Hu, X.; Zou, H.; Luo, Q.; Lee, M. S.; Xu, W.; Lam, J. W. Y.; Tang,
B. Z. (2018). Rational design of a water-­soluble NIR AIEgen, and its application in ultrafast
wash-­free cellular imaging and photodynamic cancer cell ablation. Chem. Sci. 9, 3685–3693.
76 Wang, D.; Lee, M. M. S.; Shan, G.; Kwok, R. T. K.; Lam, J. W. Y.; Su, H.; Cai, Y.; Tang, B. Z. (2018).
Highly efficient photosensitizers with far-­red/near-­infrared aggregation-­induced emission for
in vitro and in vivo cancer theranostics. Adv. Mater. 30, 1802105.
77 Zheng, X.; Wang, D.; Xu, W.; Cao, S.; Peng, Q.; Tang, B. Z. (2019). Charge control of fluorescent
probes to selectively target the cell membrane or mitochondria: theoretical prediction and
experimental validation. Mater. Horiz. 6, 2016–2023.
78 Shuai, Z.; Xu, W.; Peng, Q.; Geng, H. (2013). From electronic excited state theory to the property
predictions of organic optoelectronic materials. Sci. China Chem. 56 (9), 1277–1284.
79 Li, W.; Zhu, L.; Shi, Q.; Ren, J.; Peng, Q.; Shuai, Z. (2017). Excitonic coupling effect on the
nonradiative decay rate in molecular aggregates: formalism and application. Chem. Phys. Lett. 683,
507–514.
80 Li, W.; Peng, Q.; Xie, Y.; Zhang, T.; Shuai, Z. (2016). Effect of intermolecular excited-­state
interaction on vibrationally resolved optical spectra in organic molecular aggregates. Acta Chim.
Sinica 74 (11), 902–909.
55

Aggregation-­induced Emission from the Restriction of Double


Bond Rotation at the Excited State
Ming Hu and Yan-­Song Zheng
Key Laboratory of Material Chemistry for Energy Conversion and Storage, Ministry of Education, School of Chemistry and Chemical
Engineering, Huazhong University of Science and Technology, Wuhan, China

3.1 ­Introduction

In the whole history of human being, mankind is being dependent on light, from natural sunlight,
flame light, incandescent lamp, to light-­emitting diode (LED). While light has become an integral
part of human civilization, it is far away from fully understanding the light. Luminescence, a class
of light emitted by luminophoric molecules and called as “cold” light unlike sunlight and torch
light, not only can light up macroscopic space but can also help us to see the microscopic species
such as cell and protein. Therefore, luminescence is a cutting-­edge research field in chemistry,
materials, and biology. Countless chemists are committed to the design and preparation of new
luminescence materials and their applicational development and mechanistic disclosure and have
made a huge contribution to the progress of the scientific community.
During the course of luminescence research over the past decades, people have found that there
is a deleterious photophysical phenomenon called concentration quenching (CQ): luminescence
of solution becomes dimming and even quenching when concentration is increasing. It has gradu-
ally been identified that the quenching of luminescence is caused by the aggregation of lumines-
cence molecules, referred to as “aggregation-­caused quenching” (ACQ). Aromatic hydrocarbon
luminophores, due to their high modifiability and diversity, have always been the most studied
luminescence materials. However, the ACQ effect is especially obvious in aromatic hydrocarbons
and their derivatives [1]. Taking fluorescein as an example (see Figure 3.1, upper panel) [2], the
dilute solution of fluorescein in water has a bright green emission but the fluorescence of the solu-
tion gradually diminishes and eventually disappears after acetone fraction is increased and aggre-
gates are formed. This ACQ effect is ascribed to the close π–π stacking interactions between
fluorescein molecules in aggregated state due to planar structure of the polycyclic aromatic hydro-
carbon. While these planar luminophores are beneficial for application in solution, they suffer
from severe limitation in applications that must be carried out in aggregated state, such as biosen-
sor [3] and organic light-­emitting diode (OLED) [4], which are usually used in solid state.

Handbook of Aggregation-Induced Emission: Volume 1 Tutorial Lectures and Mechanism Studies, First Edition.
Edited by Youhong Tang and Ben Zhong Tang.
© 2022 John Wiley & Sons Ltd. Published 2022 by John Wiley & Sons Ltd.
56 3  Aggregation-­induced Emission from the Restriction of Double Bond Rotation at the Excited State

HO O O

COOH

Fluorescein ACQ

Si

Hexaphenylsilole AIE
(HPS)

Figure 3.1  Fluorescence photographs of solutions and suspensions of (upper panel) fluorescein (15 μM) in
water/acetone mixtures with different fractions of acetone (fa) and (lower panel) hexaphenylsilole (HPS;
20 μM) in THF/water mixtures with different fractions of water (fw). Source: Reproduced with permission
from Ref. [2]. Copyright 2014, WILEY-­VCH Verlag GmbH & Co. KGaA.

Figure 3.2  Representative AIEgens, including


pentaphenylsilole, tetraphenylethylene (TPE), aryl-­o-­
Si carborane, 1-­cyano-­1,2-­bis(4′-­methylbiphenyl)ethylene
C C (CN-­MBE), and diphenyl dibenzofulvene (DPDBF).

TPE
Aryl-o-carborane
Pentaphenylsilole

NC

(E)-CN-MBE
DPDBHF

Many efforts have been made to mitigate the effects of the ACQ, but few have succeeded until
2001 [5]. In this year, Tang et al. discovered an uncommon luminophore, 1-­methyl-­1,2,3,4,5-­penta
phenylsilole, which conducted no emission in dilute solution but emitted strong fluorescence in
aggregated state. They coined a new phenomenon as aggregation-­induced emission (AIE), which
is opposite to the ACQ phenomenon. Since then, a wide variety of molecules such as hexaphenyl-
silole (HPS) [6], tetraphenylethylene (TPE) [7], 1-­cyano-­1,2-­bis (4′-­methylbiphenyl)ethylene (CN-­
MBE) [8], and aryl-­o-­carborane [9] (see Figure 3.2) have been discovered to have AIE properties.
These AIE luminogenes are also shortened as AIEgens afterwards. Taking HPS as an example, it is
nonemissive when its molecules are dissolved in a good solvent like THF, but its fluorescence turns
on and becomes very strong after poor solvent like water was added and a suspension appeared
(see Figure 3.1, lower panel). In contrary to the planar structure of the conventional luminophores,
AIEgens often have a nonplanar and twisted structure, bearing crowded and anticoplanar aro-
matic rings. Due to the discovery of this new system and providing a better platform to study
3.1 ­Introductio 57

fluorescence in aggregated states, many research groups have been enthusiastically working on the
practical application and decryption of AIE effect in the past two decades. A large number of mag-
nificent achievements have been reached in various fields, ranging from solid emitter, bioimaging,
chemosensing, optoelectronics to stimuli-­responsive systems [10–23].
To better guide the discovery and practical application of novel AIE molecules, an insight into the
detailed luminescence process is highly needed. Up to date, although there are a large number of AIE
mechanisms that have been put forward, the most popular and the most successful mechanism is the
mechanism of restriction of intramolecular rotation (RIR) [12]. By this mechanism, it is known that
the substituted groups such as aromatic rings of the AIEgens are able to freely rotate in solution so
that the excited energy is released by the nonemissive rotation process and no or weak fluorescence
light is emitted. In the aggregated state, the rotation of the substituted groups is restricted and the
nonemissive rotation process is blocked. Therefore, the luminophore turns on its luminescence light
and even displays strong emission. Later, RIR mechanism incorporates the process of the restriction
of intramolecular vibrations (RIVs) and is developed into a mechanism of restriction of intramolecu-
lar motions (RIMs), which is applicable for wider AIE molecules [12–17].
However, many AIEgens such as TPE, α-­cyano-­stilbene, and their derivatives, which are most
extensively studied as AIE molecules, bear a central double bond [12–17]. Under irradiation, their
double bond may undergo rotation due to the homolytic cleavage of the π-­bond at the excited state
(see Figure 3.3). Direct and indirect evidence as well as theoretical calculation have demonstrated
the restriction of the double bond rotation (RDBR), just like the restriction of the phenyl ring

In solid
In solution Excited state strong
no restrictioin restriction

E-Z isomerization High emission


no emission
UV irradiation

Ground state

Figure 3.3  The rotation of a double bond at the excited state results in no fluorescence of TPE in solution,
but in solid state, the rotation is blocked and strong emission turns on.
58 3  Aggregation-­induced Emission from the Restriction of Double Bond Rotation at the Excited State

rotation, and play a key role on the AIE phenomena. Some researchers even consider that the
RDBR mechanism has a major effect on the AIE effect. Therefore, RDBR is the important
­complement and enrichment of RIR mechanism. It further specifies the factors that affect the AIE
phenomena and will be beneficial for exactly designing and improving novel AIE molecules.
Therefore, we make a review on the RDBR mechanism of AIE phenomena in this chapter.

3.2  ­AIE Phenomena and Applications from RDBR Mechanism

3.2.1  Evolvement and Development of AIE Mechanisms


Since AIE concept is coined for the first time by Tang’s group in 2001, a large number of AIE mol-
ecules are reported. Accompanying the appearance of the AIE molecules in massive quantities,
different kinds of AIE mechanisms have also been suggested. In general, when a planar lumino-
phore forms an excimer after irradiation, the excited energy of one molecule will be transferred to
other one and nonradiative decay occurs. Therefore, the fluorescence of the luminophore will be
quenched in solid state. However, in 2007, Tao et  al.  [24] found that fluorenone derivatives
2,7-­dip-­tolyl-­fluorenone (DTFO) and 2,7-­bis(4-­(tert-­butylthio) phenyl)-­fluorenone (DSFO) could
form excimers but displayed AIE effect. Due to the formation of a dimer through hydrogen bonds
in the crystal state, DSFO could give an excimer without arrangement adjustment after the dimer
was excited. The excimer could decay back to the dimer, and the nonradiative decay pathways that
existed in common excimers were greatly reduced, offering a strongly enhanced luminescence in
the solid state (see Figure 3.4).
In 2004, Park et al. [25] reported that 1-­cyano-­trans-­1,2-­bis-­(3′,5′-­bis-­trifluoromethyl-­biphenyl)
ethylene (CN-­TFMBE) could strongly gelatinate 1,2-­dichloroethane. The resultant gel emitted a

O
DTFOH

S S
O
DSFO

S S
O H

H O
S S

Excimer of DSFO

Figure 3.4  Structure of DTFO, DSFO, and excimer of DSFO.


3.2  ­AIE Phenomena and Applications from RDBR Mechanis 59

strong fluorescence but the solution showed no fluorescence. They ascribed the AIE effect to the
J-­aggregation of CN-­TFMBE molecules in solid state because a stronger red-­shift emission was
observed in a mixed solvent of THF/H2O 1 : 4 compared with the solution in THF. Meanwhile, the
absorption band of the suspension in THF/H2O 1 : 4 also displayed an obvious red shift with that
of solution. But later, Ma et al. [26] found that the similar α-­cyano-­stilbene derivative CN-­DPDSB
did not form compact π–π stacking in the crystal structure due to large cyano and phenyl groups.
Instead, strong intermolecular cyano nitrogen–hydrogen interactions (2.43 and 2.56 Å) restricted
the rotation of a double bond and a phenyl group in the solid state, causing its AIE effect (see
Figure 3.5).
Baumgartner et al. [27] synthesized a series of amphiphilic phosphonium compounds including
DTP, which could form liquid crystalline and soft crystal phases in the solid state, emitting fluores-
cence 40 times stronger than its solution. It was found that DTP formed a folded conformation in
solution in which the electron-­rich trialkoxybenzyl group was close to the electron-­deficient
phospholium-­containing fused scaffold. In this closed structure, the photoinduced electron transfer
(PET) process combined with phenyl rotation led to weak emission. In contrast, the solid DTP pos-
sessed an opened conformation with the trialkoxybenzyl ring pointed away from the main fused scaf-
fold and the PET process disappeared. Therefore, DTP emitted a stronger fluorescence (see Figure 3.6).

CF3
CF3

CF3 NC
CN-TFMBE CF3

NC

CN

CN-DPDSBH

Figure 3.5  Structure of CN-­TFMBE and CN-­DPDSB.

S S S
S
OC12H25
H H
P+ PET
P+ OC12H25 X
OC12H25

OC12H25
OC12H25
DTP DTP
OC12H25
In solution In solid state

Figure 3.6  Structure of DTP and its PET process.


60 3  Aggregation-­induced Emission from the Restriction of Double Bond Rotation at the Excited State

By cross-­dipole stacking instead of H-­aggregation, stilbene derivatives trans-­DPDSB without a


cyano substituent could emit a strong blue fluorescence in the crystal state. It is considered that the
cross-­dipole stacking of the trans-­DPDSB molecules can increase a finite transition dipole moment
between the ground state and the lowest excited state of the clusters, promoting the light-­emission
properties [28]. Tian et al. [29] also found that a butterfly-­like molecule 9,10-­bis(2,2-­diphenylvinyl)
anthracene (BDPVA) formed cross-­dipole stacking in the crystal state and emitted a stronger fluo-
rescence in solid than in solution (see Figure 3.7).
In addition, molecular planarization is also one of possible AIE mechanisms. Park et al. [8] ever
reported that 1-­cyano-­trans-­1,2-­bis-­(4′-­methylbiphenyl)-­ethylene (CN-­MBE) possessed a twisted
structure so that it emitted no fluorescence in solution. However, in solid state, the twisted struc-
ture was stretched out to give a more coplanar structure, which was helpful for its J-­aggregation
and showed a strong emission (see Figure 3.8).
Very specially, Zheng et  al.  [30] recently found that a totally planar molecule
5,7,12,14-­tetraoxapentacene (TOP) also displays the typical AIE effect. In highly polar solvent,
TOP had no emission due to the solvation role but emitted a strong fluorescence in solid state
because of the largely slipped π–π stacking. Although this molecule is planar, oxygen atoms on the
different molecular planes prevented cofacial π–π stacking that can arouse ACQ result because of
the repulsive force between oxygen atoms (see Figure 3.9).
Although these above AIE mechanisms can well explain the emission or emission enhancement
of some special luminophores in solid state, they are not applicable for more extensive and typical
AIEgens such as TPE and phenylsilole derivatives until the RIR mechanism was put forward by
Tang et  al.  [12–17]. The RIR mechanism was first proposed to explain the AIE effect of silole
derivatives such as HPS (see Figure 3.10). HPS is a typical AIE molecule. When HPS molecules are
dissolved in a good solvent at low concentrations, fluorescent quantum yield of the solution is
extremely low (Φf ∼ 0.1%). However, the emission is dramatically enhanced after the fraction (fw)
of poor solvent such that water reaches ∼50 vol % and reaches a maximum of ∼22% at fw = 90 vol%,

d- d-

d+
d+
trans-DPDSB

Cross stacking of DPDSB

BDPVA

Cross stacking of BDPVA

Figure 3.7  Structure of DPDSB and BDPVA and their cross-­dipole stacking.


3.2  ­AIE Phenomena and Applications from RDBR Mechanis 61

In dilute solution

49°
48° 47° Non-radiative decay
No
fluorescence

Twisted structure

Planarization

In nanoparticles

Preventing π-stacking interaction


(in favor of J-aggregation)

High
fluorescence
Effective
conjugation length
Coplanar structure

Figure 3.8  Emission of CN-­MBE in the solid state by molecular coplanarization. Source: Reproduced with
permission from Ref [8]. Copyright 2002, American Chemical Society.

CN CN
F O O F

F O O F
CN TOP CN

(a)

(b)

(c)
2.095

3.334
3.350

Figure 3.9  Molecular structure and crystal structure of TOP viewed (a) perpendicular to the molecular
plane and (b) along the molecular plane; (c) the slipped π–π packing of the molecules.
62 3  Aggregation-­induced Emission from the Restriction of Double Bond Rotation at the Excited State

(a) (b)
0.24
θ1 = –58.57°
θ2 = +96.15°
0.18 θ3 = + 3.95°
θ4 = –39.46°
Si
θ5 = –35.59°
0.12 θ2 θ1
ΦF

θ4 θ6 = +47.41°
θ3

0.06 θ5
θ6

0.00
0 30 60 90
Water fraction (vol%)
(c)

Aggregation

Restriction of
intramolecular
rotation

HPS

Active rotation
in solution RIR in aggregate
Non-emissive Highly emissive

Figure 3.10  (a) Fluorescence quantum yield of HPS vs water fraction in acetone/water. Source: Reproduced
with permission from Ref. [11]. Copyright 2003, American Chemical Society. (b) Molecular conformation of
HPS. Source: Reproduced with permission from Ref. [12]. Copyright 2015, American Chemical Society. (c) HPS
molecule emission. Source: Reproduced with permission from Ref. [15]. Copyright 2015, Royal Society of
Chemistry.

being approximately 220-­fold increase in fluorescence quantum yield. As shown in the molecular
structure of HPS, the central silole is connected with six phenyl rings by rotatable single bonds,
which make HPS to show a great conformational flexibility in isolated phase. Due to the six phenyl
rings around a tiny core, the HPS molecular structure is rather distorted. This torsion makes a great
tension between adjacent phenyl rings, and it is difficult for the entire molecule to form a planar
conformation. It can be seen from the crystal structure that HPS molecules possess a propeller-­like
conformation and each phenyl rings exhibit different degrees of distortion compared with the
silole core. This nonplanar structure makes it difficult for molecules to pack tightly. Therefore, no
π−π stacking interaction is observed for HPS in the solid state, avoiding the ACQ effect. Meanwhile,
the dynamical rotation of six phenyl rings is restricted in the solid state and the nonradiative chan-
nel is blocked. Therefore, in aggregates, high emission is accessible.
3.2  ­AIE Phenomena and Applications from RDBR Mechanis 63

However, it is found that some luminophores lack substituents that can rotate but display typical
AIE phenomenon. For example, annulenylidene THBA (see Figure 3.11a), whose phenyl rings are
fixed by a pair of ethylene tethers, still exhibits strong emission in aggregates but no emission in
dilute solution [12, 31, 32]. Consequently, the RIV mechanism is raised. For this class of lumino-
phores, intramolecular vibration instead of rotation is restricted in the aggregated state, which
turns on the luminescence. Recently, Tang et al. reported cyclooctatetrathiophene (COTh) and its
derivatives that are AIE active [33]. There are no rotatable unit in the molecular structure of COTh
(see Figure  3.11b) in both ground and excited states. However, in solution, the intramolecular
vibration of excited state could cause up-­down conformation inversion and result in a fluorescent
quenching. Because of the RIV mechanism, COTh emits a strong green fluorescence in the
solid state.
If an AIEgen includes both rotating and vibrating units, obviously, neither RIR nor RIV can
explain its AIE effect alone. Therefore, RIM mechanism is used to explain AIEgens with more
complex structures and emission units. Some examples of such AIEgens are shown in Figure 3.12
Those molecules are suitable for the RIM mechanism because of their multiform motions of iso-
lated molecules [34–37]. RIM mechanism advocates that the fluorescent quenching of AIEgens in
dilute solution results from various ways of intramolecular motions including twisting, deforming,
flapping and probably motion of solvating molecules in addition to rotation and vibration. While it

Figure 3.11  (a) RIV mechanism of THBA. Source: Reproduced with permission from Ref. [12]. Copyright
2014, WILEY-­VCH Verlag GmbH & Co. KGaA. (b) RIV mechanism of COTh. Source: Reproduced with
permission from Ref. [33]. Copyright 2019, Springer Nature.
64 3  Aggregation-­induced Emission from the Restriction of Double Bond Rotation at the Excited State

(a)
θN = 142°
N S S N
N N

N Side view
C6H13
PTZ-BZP

(b)
N N
C C

c
b
b a
c
C C
N N

dP-TCAQ

Figure 3.12  Examples of luminogens whose AIE activities are ascribed to the process of restriction of
intramolecular motions (RIMs). (a) Molecular structure of PTZ-­BZP and DFT-­optimized geometry of a
simplified structure with a methyl substituent. Source: Reproduced with permission Ref. [34]. Copyright
2014, WILEY-­VCH Verlag GmbH & Co. KGaA. (b) Molecular structure and single-­molecular structure of
dP-­TCAQ. Source: Reproduced with permission from Ref. [35]. Copyright 2013, Royal Society of Chemistry.

greatly broadens the scope of application of the AIEgens, it also poses new challenges for research-
ers. Because, under real conditions, in order to better guide the optimization of the structure of
AIEgens, people need to know the exact factor of motions, rather than simple and general motions.
Consequently, although the RIM mechanism is the most popular and applicable, a lot of detail
contents need to be embodied and supplied.

3.2.2  Investigation of RDBR AIE Mechanism by E/Z isomerization


It is known in the textbook of organic chemistry that a carbon–carbon double bond is unable to
rotate in general conditions. However, under irradiation or at high temperature, the π-­bond of the
double bond can undergo homolytic cleavage and the double bond can rotate. The most well-­
known example is molecular photoswitches or even molecular motors in which the Z-­isomer and
E-­one can be reversibly transformed into each other (E/Z isomerization, EZI) through double bond
rotation under irradiation or heating [38, 39]. Among AIEgens, most of them such as TPE, cyanos-
tilbene, and their derivatives possess one essential double bond. Whether the double bond under-
goes a rotation under irradiation of ultraviolet (UV) light and arouses nonradiative decay of these
AIEgens in solution is brought to concern even from the moment the AIEgens start to appear.
Because of the relatively simple synthesis method, easy to modify, and stable AIE effect, TPE and
derivatives become the most extensively studied AIEgens. Meanwhile, RIR mechanism is rightly
concluded from the investigation of TPE and its derivatives. Whether the double bond rotation of
the TPE unit is involved in the AIE process is always in controversy. Due to the symmetrical struc-
ture of TPE, it is difficult to observe the obvious effect of double bond rotation without proper
modification. In the beginning, it was thought that the double bond had little effect on the AIE
mechanism.
3.2  ­AIE Phenomena and Applications from RDBR Mechanis 65

In 2012, Tang et al. prepared a dimethylated TPE derivative called 1,2-­diphenyl-­1,2-­di(p-­tolyl)


ethylene (DPDTE)  [40]. To conduct an EZI experiment, a pure E or Z conformation is needed.
Therefore, they used an unconventional way, a Pt-­catalyzed reaction rather than the normal
McMurry coupling reaction, to provide a Z-­rich product. In 1H NMR spectra, E-­isomer of DPDTE
has two classes of protons that appear at δ 6.89 and 2.24 ppm, while the corresponding protons
have signals at δ 6.91 and 2.26 for the Z-­isomer. By comparing the integral areas of these two pro-
tons of the two isomers, it is revealed that the as-­prepared product is Z-­rich and has a Z/E ratio of
93 : 7. Through recrystallization, pure Z-­isomer was obtained. By using a UV lamp of spectrofluo-
rometer as a light source to irradiate a solution of Z-­DPDTE for a long period of time (29 minutes),
they hoped to get EZI result. However, throughout the whole radiation process, no change of 1H
NMR spectra could be observed. But when they used a UV lamp with a much stronger power to
irradiate the DPDTE solution, EZI happened in a much shorter period. The NMR spectra showed
that the peaks of E-­isomer were raised after the strong UV irradiation. In the same year, they
reported similar results of TPE derivative named BPHTATPE [41]. By attaching large and polar
moieties to a TPE core, they prepared BPHTATPE whose difference in the shape and polarity
between E-­and Z-­isomers can allow these two isomers to be separated just by making use of sim-
ple column chromatography. They performed EZI experiments on both the two isomers. It turned
out that BPHTATPE could undergo E/Z isomerization only under a powerful UV light and high
temperature (>200 °C) (see Figure 3.13). Consequently, they deduced that the EZI was not involved
in its AIE process under the normal photoluminescence spectral measurement conditions.
But in 2016, the possible involvement of a double bond rotation in the AIE process of the TPE-­
based luminogen was found by Tang et al. [42] They designed a TPE derivative 1,2-­bis(4-­fluoroph
enyl)-­1,2-­bis(4-­methoxyphenyl) ethylene (TPE-­FM), which possessed more polarity due to the

N
O(CH2)6 N N

E-isomers
Easy to react under strong UV

Hard to react under weak UV

N N (CH3)6O
N
(E)-DPDTE
(E)-BPHTATPE

Z-isomers O(CH2)6 N N N N (CH3)6O


N N

(Z)-DPDTE (Z)-BPHTATPE

Figure 3.13  The molecular structures of DPDTE and BPHTATPE.


66 3  Aggregation-­induced Emission from the Restriction of Double Bond Rotation at the Excited State

strong electron-­attracting fluorine atoms and the electron-­donating methoxyl groups (see
Figure 3.14). Therefore, the E-­ and Z-­isomers are easily separated and single crystals even grow
well due to the large molecular polarity. The obtained crystal structure can make sure the identity
of E-­and Z-­isomers. With the identified isomers in hand, they carried out more careful test on the

(a) (b)
4
TPE-FI λex = 480 nm

3
ε /104 (l/mol/cm)

Fluorescein

PL intensity (a.u.)
TPE
2

1
λex = 330 nm

0
250 350 450 550 450 500 550 600 650
Wavelength (nm) Wavelength (nm)
(c)
100
Rotation
Rotation
+
Isomerization O–
80 O
O– O
O O
Relative PL quantum yield (%)

O
60 O
O O

O O

40

20

0
250 300 350 400 450 500 550 600
Excitation wavelength (nm)

Figure 3.14  Structure of TPE-­FM and TPE-­Fl. (a) Absorption spectra of TPE, TPE-­Fl, and fluorescein.
(b) Emission spectra of TPE-­Fl in solution (λex: excitation wavelength). (c) Relative fluorescence quantum
yields of TPE-­Fl. Source: Reproduced with permission from Ref. [42]. Copyright 2016, Royal Society of
Chemistry.
3.2  ­AIE Phenomena and Applications from RDBR Mechanis 67

F F
F O

O O
O F

(E)-TPE-FM (Z)-TPE-FM

O O
O
O
OH

TPE-FI

Figure 3.14  (Continued)

EZI of TPE-­FM in solution. Due to the difference of 1H NMR spectra of E-­and Z-­isomers, the EZI
process of TPE-­FM was observed in the solution at a high concentration of 10 mM under a UV
radiation of high power (321 nm, ~0.47 mW/cm2) for a 15-­minutes irradiation. But emission spec-
trum was generally measured at a highly diluted solution (~10 μM). The EZI process at high con-
centration did not guarantee its occurrence at highly diluted solution. Considering the low
sensitivity of 1H NMR spectroscopy and unable to measure NMR spectra at very low concentra-
tion, they prepared three cuvettes of solutions (10 μM), which were irradiated by a 312-­nm UV
light from the xenon lamp with a low power (321 nm, ~0.012 mW/cm2) in the spectrofluorometer
for specific periods of time. The solutions from the three cuvettes were combined and evaporated
to dryness in a dark room. The obtained solid was dissolved in a small amount of ­dichloromethane-­d2
for NMR measurement. By this effort, 1H NMR spectra with high quality showed that EZI truly
occurred even at very low concentration under low-­power excitation. Whereas it is difficult to
determine whether the double bond rotation played a major role or minor role on the emission
process of the TPE AIEgens, compared with the phenyl ring rotation.
In order to disclose the different roles of the double bond rotation and phenyl ring rotation on
the AIE process, Tang et al. elaborately designed a TPEgen-­bearing fluorescein unit (TPE-­Fl). This
TPEgen displays two main absorption bands at 340 and 450 nm, which are from TPE and fluores-
cein units, respectively (see Figure 3.15). In solution, TPE emits a weak fluorescence at 490 nm,
which is overlapped with the main absorption band of the fluorescein unit. Therefore, the emis-
sion wavelength of the TPE unit can act as the excitation wavelength of the fluorescein unit
through a fluorescence resonance energy transfer process. Therefore, by using a two-­color excita-
tion, the contribution of the double bond rotation and phenyl ring rotation can be determined.
When TPE-­Fl was excited by a long wavelength of 480 nm, the decrease of the fluorescence inten-
sity of TPE-­Fl compared to that of the fluorescein only came from the phenyl ring rotation because
the double bond rotation would not occur under so long wavelength of light. When TPE-­Fl was
68 3  Aggregation-­induced Emission from the Restriction of Double Bond Rotation at the Excited State

Br
HO OH HO OH
N
OHC CHO
N N
N

TFA/reflux Br
70% OHC CHO
K2CO3/CH3CN
HO OH reflux
HO OH
1 2 43%

O O O O

OHC CHO

NaBH2 HO OH
SOCl2/CH2Cl2
C2H5OH/THF
HO OH
OHC 85%
CHO

O O O
O
3 4

O O OH O O
O

Cl Cl O

Cl Cl OH
O
K2CO3/KI
CH3CN/reflux O
O O 54% O O
5 6

Figure 3.15  The synthetic route of cis-­TPE dicycles 3–5 and TPE tetracycle 6.

excited by a short wavelength of 330 nm, the decrease of the fluorescence intensity of TPE-­Fl com-
pared to that of the fluorescein came not only from the phenyl ring rotation but also from the
double bond rotation because the double bond would break under a high energy of light. By taking
the fluorescence quantum yield (Φf ) of fluorescein to be 100%, the Φf with excitation wavelengths
of 480 and 330 nm were 19.2 and 9.2%, respectively. It suggested that the fluorescence was decreased
by 10% due to double bond rotation but the fluorescence decrease due to the phenyl ring rotation
was 100–19.2% = 80.8%. From this indirect experimental evidence, they concluded that the double
bond rotation played a minor role on the AIE effect.
Indeed, in the ground state, phenyl rotation is much easier than the C═C bond, but during the
fluorescent course, the double bond can be excited and break to generate radical or charged species
3.2  ­AIE Phenomena and Applications from RDBR Mechanis 69

(radical, cation, and/or anion), its twist or torsion become available, which make the true radiation-­
less relaxation pathway become quite complicated.

3.2.3  Investigating of RDBR AIE Mechanism by Immobilization of TPE


Propeller-­like Conformation
In addition to the observation of the EZI process that can disclose RDBR mechanism, immobiliza-
tion of TPE propeller-­like conformation, especially cyclization of TPE at cis-­position, can be used to
explore the RDBR process. After bridging between two phenyl rings of TPE at the cis-­position, the
double bond will be unable to freely rotate due to the restriction of the bridge chain. But the phenyl
rings can still freely rotate. Therefore, the effect of RDBR on the fluorescence will be clearly observed.
In 2016, Zheng’s group [43] found an ideal route for the synthesis of cis-­TPE dicycle in which the
double bond rotation can be blocked at the excited state (see Figure 3.15). By intramolecular nucle-
ophilic substitution of 2 with 1,4-­bis(bromomethyl)benzene, cis-­TPE dicycle tetraldehyde 3 could
be obtained in a 43% yield. The formation of cis-­TPE dicycles between two phenyl groups at the
cis-­position instead of those at the gem-­position was ascribed to the length of 1,4-­benzenedimethyl
tether that was more compatible with the distance between two cis-­phenyl rings than that between
two gem-­phenyl ones. With this key intermediate in hand, even TPE tetracycle 6 whose propeller-­
like conformation was completely immobilized was obtained.
Noticeably, while the TPE derivatives 1 and 2 bearing no intramolecular cycle did not emit fluo-
rescence in solution, cis-­TPE dicycles 3–5 displayed a strong emission in solution. In THF, Φf of 2
was 0% but that of 3 and 4 was 24 and 49%, respectively. Due to the bridge, p-­phenylenedimethyl
units were short and rigid, and the propeller-­like conformation of the TPE dicycle should have been
fixed. However, the enantiomers obtained by chiral high-­pressure liquid chromatography (HPLC)
rapidly racemized in solution at room temperature. This result indicated that the phenyl rings of the
cis-­TPE dicycles was still able to rotate although the rotation had a little restriction. In addition, the
formaldehyde groups of cis-­TPE dicycle 3 reacted with chiral α-­methylbenzylamine to form imine,
and positive circular dichroism (CD) signals could be induced by (R)-­α-­methylbenzylamine but
negative CD signals were aroused by (S)-­α-­methylbenzylamine. The CD signals should result from
a single-­handed propeller-­like conformation induced by the enantiomer of the chiral amine because
neither individual 3 nor amine enantiomers could show such signals. This meant that the propeller-­
like conformation of 3 could be transformed from the left-­handed helical direction to the right-­
handed helical direction, demonstrating that the phenyl could freely rotate. Consequently, it could
be inferred that about 25–50% Φf should come from the RDBR process (see Figure 3.16).
As expected, the resolved enantiomers from TPE tetracycle 6 were stable due to complete immo-
bilization of its propeller-­like conformation. The racemate of 6 had Φf up to 97%, and both of the
two enantiomers had a quantitative Φf up to 100% due to restriction of not only double bond rota-
tion but also phenyl ring rotation. Compared with the corresponding TPE dicycle 4, TPE tetracycle
6 showed a twofold increase in fluorescence intensity, demonstrating that RDBR and RIR play
equal key roles on the AIE effect.
In addition, the two enantiomers of 6, M-­6 (left-­handed helical propeller-­like configuration) and
P-­6 (left-­handed helical one), emitted strong circularly polarized luminescence (CPL) signals in
THF with a dissymmetric factor (glum) of +3.1 × 10−3 for M-­6 and −3.3 × 10−3 for P-­6, indicating
that the propeller-­like conformation of TPE was maintained even at the excited state. Moreover,
the |glum| of CPL was similar with the gabs of absorbance (gabs = 2(Δε/ε)) of 2.4 × 10−3 for M-­6 and
P-­6, indicating little conformational change between the ground state and excited state. This fur-
ther confirmed that no double bond rotation occurred for this TPE tetracycle (see Figure 3.17).
70 3  Aggregation-­induced Emission from the Restriction of Double Bond Rotation at the Excited State

80
NH2 R–7
R–7+3
40 S–7+3
7
CD (mdeg)

–40

–80
250 300 350 400
Wavelength (nm)

Figure 3.16  CD spectra of a mixture of TPE dicycle 3 and enantiomer of α-­methylbenzylamine 7 in the


presence of acetic acid in 1,2-­dichloroethane.

(a) (b)

(c) (d)
30
3 4 6
M-6
15
CPL (mdeg)

–15 P-6

–30
450 500 550 600 650 700
Wavelength (nm)

Figure 3.17  The crystal structures of M-­6 (a) and P-­6 (b); (c) photos of 3, 4, and 6 in THF solution under a
365-­nm UV light and (d) CPL spectra of M-­6 and P-­6 in THF (1.0 × 10−3 M). Source: Reproduced with permission
from Ref. [43]. Copyright 2016, American Chemical Society.
3.2  ­AIE Phenomena and Applications from RDBR Mechanis 71

In order to further disclose the important contribution of the RDBR process to the AIE effect,
gem-­TPE dicycles 7 and 8 and even the typical isomers cis-­ and gem-­TPE dicycles 9 and 10 were
designed and synthesized in Zheng’s group (see Figure 3.18) [44]. Their configuration had been

(b)

253 K

263 K

Lorem Ipsum

273 K

298 K

313 K

7 6 5 4 [ppm]

Figure 3.18  (a) Structures of TPE dicycle isomers 7–10 (left) and photos of their solution in THF
(1.0 × 10−4 M) under a 365-­nm UV light (middle). (b) The change of 1H NMR spectra of 7 in CDCl3 with
temperature. Source: Reproduced with permission from Ref. [44]. Copyright 2018, American Chemical
Society.
72 3  Aggregation-­induced Emission from the Restriction of Double Bond Rotation at the Excited State

confirmed by the crystal structure. By comparing the fluorescence intensity of these isomers, the
effect of groups, atoms, and the bridge chains on the fluorescence could be excluded. Therefore,
more direct and more exact RDBR evidence could be furnished.
As expected, while cis-­TPE dicycles 3, 4, and 9 emitted a strong fluorescence, the gem-­TPE dicy-
cles 7, 8, and 10 had no emission in THF under an irradiation of a 365-­nm UV light. The fluores-
cence quantum yields of the cis-­TPE dicycles 3, 4, and 9 were 24, 49, and 22%, while those of the
gem-­TPE dicycles 7, 8, and 10 were 3.1, 5.7, and 1.9%, respectively, in THF at 25 °C. Compared with
the gem-­isomer 10, the cis-­isomer 9 displayed an increase of fluorescence intensity by 11.6-­fold.
From the 1H NMR of the gem-­TPE dicycle 7 at different temperatures, the phenyl ring of the TPE
unit was fixed and failed to rotate due to linkage of the bridge at the m-­position of the phenyl rings.
However, the bridging unit 1,4-­benzenedioxymethyl was rapidly rotating at 25 °C because the pro-
ton peaks of both the 1,4-­benzenedioxy and methylene unit were very wide and flat. When the
temperature was lowered to 0 °C, these proton resonance signals became sharp and well resolved,
indicating that the free rotation of these bridge units was also restricted. However, the Φf of 7 was
only increased to 9.0% at 0 °C and was still much less than that of cis-­isomers. Compared with the
cis-­isomer, the fluorescent decrease of the gem-­isomer in solution only came from free rotation of
the double bond at the excited state after all other intramolecular rotations had been restricted (see
Figure 3.18 right).
Time-­resolved fluorescence decay of TPE dicycles 3, 4, and 7–10 disclosed that the fluorescence
lifetime was in the range of 6.9–14 ns in THF and 13–19 ns in suspension for these TPE dicycles.
In suspension, their fluorescence lifetime was always larger than that in solution. This demon-
strated that there was further restriction of intramolecular rotation in solid state. By making use of
the fluorescence quantum yields and lifetime values, the radiative (kf ) and nonradiative (knr) rate
constants of 3, 4, and 7–10 (kf, knr (ns−1)) could be calculated. The kf and knr were 0.035 and 0.112
for 3, 0.051 and 0.053 for 4, 0.004 and 0.126 for 7, 0.004 and 0.067 for 8, 0.023 and 0.080 for 9, and
0.002 and 0.103 for 10. And ratios of knr vs. kf for 3, 4, and 7–10 were 3.20, 1.04, 31.5, 16.8, 3.48, and
51.5, respectively. The knr/kf ratio from gem-­isomers was always much larger than that from cis-­
isomers, demonstrating a more nonradiative process from gem-­isomers. This nonradiative process
should be mainly ascribed to the double bond rotation.
If the double bond rotated at the excited state, one intermediate state in which sp2-­hybridized
orbital planes of two carbons are vertical instead of coplanar should exist. This twisted state of the
double bond should be able to be observed by femtosecond transient absorption spectra because of
the decreased conjugation with one another. It was true that two excited-­state absorption (ESA)
bands, which were located at <460 and >600 nm, respectively, were observed. The former should
come from a twisted excited double bond and the latter came from a planar double bond at the
excited state, corroborating the occurrence of the double bond rotation. Surprisingly, the two ESA
bands existed both in gem-­isomers and in cis-­isomers. Two typical time components τ1 and τ2 from
the dynamic decay process were obtained. By the global analysis, the first component τ1 represent-
ing the rise component of ESA and associated with the geometry relaxation from the Franck–
Condon (FC) configuration was the negative amplitude. The second component τ2 represented the
nonradiative internal conversion (IC) decay of the excited state. From time-­resolved spectra, an
obviously growing process and the increase ending up in less than 20 ps were observed for the spe-
cies at short wavelength, suggesting that the planar excited state could be transformed into the
twisted excited state. Therefore, the double bond rotation at the excited state occurred for all the
TPE dicycles.
However, there was an obvious difference of the double bond rotation between the cis-­and gem-­
isomers. As an index of the rotation, the component τ1 was 6 ps for gem-­10 and obviously shorter
3.2  ­AIE Phenomena and Applications from RDBR Mechanis 73

than 15 ps of cis-­9. This should be ascribed to the more freely rotating double bond in gem-­isomers
than in cis-­isomers. At 21 and 14 ps, the rotation was accomplished because the maximum inten-
sity of the absorption spectra of cis-­isomer 9 and gem-­isomer 10 at the excited state was reached,
respectively. It was found that the absorption maximum wavelength of the gem-­isomer 10 was
shortened by 15  nm compared with that of the cis-­isomer 9. Moreover, the area ratio of short-­
wavelength band vs. long-­wavelength band was much larger for the gem-­isomer 10 than that for
the cis-­isomer 9, further corroborating that the double bond of the gem-­isomer rotated more freely
than that of the cis-­isomer. Therefore, the gem-­isomer showed lower fluorescence quantum yield
than the cis-­one because of the freer double bond rotation at the excited state and more nonradia-
tive decay (see Figure 3.19).
Recently, Zheng et al. [45] have exploited the RDBR mechanism to improve the sensitivity of
DNA detection and enhance the chiroptical properties from TPE AIEgens. In this regard, cis-­TPE
macrocycle diquaternary ammoniums 11 were designed and synthesized. As a comparison, gem-­
isomers 12 were also prepared. As soon as the two ammonium arms at the cis-­position

(a)

1.0
21 ps at syn-9
14 ps at gem-10
0.8
Normalized intensity

0.6

0.4

0.2

0.0

–0.2
400 450 500 550 600 650 700
Wavelength (nm)
(b)

hv Difficult

cis cis* cis**

hv Easy

Gem Gem* Gem**

Figure 3.19  (a) Femtosecond transient absorption spectra of 9 and 10 at the respective maximum
intensity. (b) Diagrammatic sketch of the normal excited state (cis* and gem*) and twisted excited state
(cis** and gem**) of the double bond. Source: Reproduced with permission from Ref. [44]. Copyright 2018,
American Chemical Society.
74 3  Aggregation-­induced Emission from the Restriction of Double Bond Rotation at the Excited State

O O +
m N
O O
Cl −
O O O O

m m O
O
Cl−
+
+ X− + X− m N
N m
N
m X
gem-12a 0
cis-11a 0 Br gem-12b 1
cis-11b 1 Cl

Figure 3.20  Structures of cis-­and gem-­TPE macrocycle diquaternary ammoniums 11 and 12.

simultaneously hold on one DNA chain by electrostatic attraction, the formed cycle together with
the original cycle will completely immobilize the double bond rotation at the excited state and
arouse the enhancement of the AIE effect (see Figure 3.20).
Due to the limitation of crown ether cycle, both cis-­11 and gem-­12 display weak emission in solu-
tion. However, while the quantum yield of gem-­isomer 12 was 1.5%, the cis-­TPE ammonium 11 had
a Φf of 3.0%, which was a 2.0-­fold stronger than that of the gem-­one. This should be ascribed to the
partial limitation of the double bond rotation in cis-­isomers but no restriction of the double bond
rotation in the gem-­one. When cis-­TPE ammonium cis-­11a or cis-­11b was added to the solution of
calf thymus DNA (CT-­DNA), strong new CD signals were induced. In addition to CD signals of
DNA itself at short wavelengths, one bisignate band from about 370 nm (+) to 310 nm (−), which
should be ascribed to the single-­handed propeller-­like conformation of the TPE unit, appeared.
Conversely, gem-­isomers gem-­12a and gem-­12b showed weak CD signals from the TPE unit and
did not form a new bisignate band. Probably because of more flexibility of the TPE unit in the gem-­
isomer than in the cis-­one, DNA was unable to induce the stable single-­handed propeller-­like
conformation.
Outstandingly, strong CPL emission was observed in the drop-­cast film from a mixture of cis-­11a
and cis-­11b with CT-­DNA in water, while the gem-­isomer–CT-­DNA film emitted no CPL signals.
The CPL dissymmetric factor (glum) of 0.0028 and 0.016 for cis-­11a and cis-­11b, respectively, was
much larger than that from the mixture of DNA with other AIEgens. Even in solution, strong CPL
light was emitted from a mixture of cis-­isomer with CT-­DNA in water but no CPL signals were
found for the mixture of gem-­isomers with CT-­DNA. With fish sperm DNA (FS-­DNA), a similar
result was obtained. Considering the structure of the cis-­isomers, the CPL enhancement should
result from the more RDBR process of cis-­isomers in which the formed cycle in situ together with
the original cycle in the cis-­position firmly restricts the double bond rotation not only at the ground
state but also at the excited state (see Figure 3.21).
Given that the obvious interaction of the TPE diammoniums with DNA, they should be excellent
sensor for the detection of DNA. It was truly that the fluorescence of TPE macrocycle
3.2  ­AIE Phenomena and Applications from RDBR Mechanis 75

(a)
120
CT-DNA + cis-11a
100 CT-DNA + cis-11b
CT-DNA + gem-12a
CT-DNA + gem-12b
80
CPL (mdeg)

60

40

20

0
500 600 700
Wavelength (nm)

(b)

80 CT-DNA + cis-11a
CT- DNA + cis-11b

60
CPL (mdeg)

40

20

0
500 600 700
Wavelength (nm)

Figure 3.21  CPL spectra of a drop-­cast film (a) and solution (b) from a mixture of cis-­TPE isomers and
gem-­TPE isomers with CT-­DNA in water.

diammoniums in water was increased when FS-­DNA was added into the solution at 1.0 × 10−6 M. But
the solution from a mixture of cis-­isomer with FS-­DNA showed stronger fluorescence than that
from the corresponding gem-­isomer and FS-­DNA. The fluorescence intensity was linearly increased
in the range of DNA concentration less than 1.0 × 10−8 M. As a result, the detection limit for DNA
analysis was obtained. It was found that the detection limits were 123, 74, 496, and 235 pM for 11a,
11b, 12a, and 12b, respectively. The cis-­isomers had always much lower detection limit than the
gem-­ones. And the detection limitation of 74 pM from cis-­isomer is among the best results from
AIE DNA sensors. The higher sensitivity of cis-­isomers than that of gem-­isomers should also come
from the RDBR mechanism. As shown in Figure  3.22, the restriction of a double bond of cis-­
isomers upon binding to DNA chain enhanced the AIE effect. Therefore, the sensitivity was signifi-
cantly increased.
When the cis-­TPE macrocycle diammonium was synthesized using octaethylene glycol as a
bridge, the resultant crown ether cycle is large enough to allow the EZI. As shown in Figure 3.23,
76 3  Aggregation-­induced Emission from the Restriction of Double Bond Rotation at the Excited State

(a) (b)

+
No CPL O P

Strong CPL + –
O
P
O O


O P O
+ P
O –
O
+

Figure 3.22  Schematic diagram of the binding of TPE cycle diammoniums 11 (a) and 12 (b) to the
DNA strand.

the as-­prepared cis-­13 could be converted into trans-­13 under an irradiation of a 365-­nm portable
UV lamp both in organic solvent and in water. Under a 365-­nm light from one fluorophotometer,
the absorption spectrum of cis-­13 in CH2Cl2 had a gradual absorbance increase at 355 and 278 nm
but a constant decrease at 315 nm with irradiation time, and showed an obvious isosbestic point at
336 and 302 nm, indicating the conversion from cis-­isomer to trans-­one. In CDCl3, after irradiation
by a 365-­nm portable UV lamp for one hour, another set of signals appeared beside signals of cis-­13
in the 1H NMR spectrum. Especially, besides the signals of the aromatic proton near to the oxygen
atom (6.71 ppm, double), benzyl methylene (4.82 ppm, single), and the ethylene proton close to the
aromatic ring (4.09 ppm, triple), new peaks that were well separated and had the same shape and
split with that of cis-­isomer appeared at a lower field. The integral area of these new peaks was all
almost equal to that of cis-­13, suggesting a 50% conversion of cis-­isomer to trans-­one. In DMSO, the
converted trans-­isomer (about 20%) by light could completely come back to cis-­isomer under heat-
ing at 180 °C. This result suggested that the double bond of the TPE derivatives was easy to rotate
in the excited state.
The TPE ammonium 13 was tested for the detection of fish sperm DNA. It was found that
cis-­13 had a high sensitivity (139 pM) and high-­intensity ratio Imax/I0 (6.6), whereas a mixture
of cis/trans-­13 about 1 : 1 displayed a low-­sensitivity (326 pM) and low-­intensity ratio Imax/I0
(4.5), indicating that the cis-­isomer was a much better DNA sensor than the trans-­one due to
more restriction of intramolecular motion including the double bond rotation (see
Figure 3.23).
There are more examples in which TPE cycle at the cis-­position could emit fluorescence in solu-
tion due to the restriction of double bond rotation. Rathore et al. synthesized a class of stilbeno-
prismands 14 via an intramolecular McMurry coupling reaction, which contained a TPE
core-­bearing cycle at the cis-­side (see Figure  3.24 left)  [46]. While the parent TPE showed no
detectable emission, the solution of 14 showed significant emission under similar conditions
although there were two substituted phenyl rings that could freely rotate. By the direct coordina-
tion of TPE tetraimidazolium salts with silver or gold, Hahn and Strassert [47] prepared TPE dicy-
cles 15 on the cis-­side. The TPE dicycles 15 showed strong emission in solution but TPE
tetraimidazolium salts emit no fluorescence, demonstrating the important role of the RDBR pro-
cess on the fluorescence (see Figure 3.24 right).
Hu et al. [48] reported one TPE derivative dicycle 17, in which two pairs of the phenyl rings at
the cis-­position were connected by two diacetylenes. While the no cyclized TPE reactant 16 emitted
3.2  ­AIE Phenomena and Applications from RDBR Mechanis 77

(a)

O O O
Cl – +
O O N
O O
O O
O O O O
O
hv O
180 °C
O

O O

– +
+ Cl + N –
N – N Cl
Cl trans-13
cis-13

(b)
0.5
0 minutes
2 minutes
0.4
4 minutes
6 minutes
8 minutes
Absorbance

0.3
10 minutes
15 minutes
0.2 20 minutes
30 minutes
40 minutes
0.1 50 minutes

0.0
250 300 350 400 450 500
Wavelength (nm)
(c)

CHCl3

H2O
No irradiation

H2O

Irradiation

7 6 5 4 3 2 [ppm]

Figure 3.23  (a) The cis-­/trans-­isomerization of 13 under a 365-­nm light irradiation and heating.
(b) Change in UV–vis spectrum of compounds cis-­13 in dichloromethane under an irradiation of 365-­nm
light from fluorophotometer for different periods. [cis-­13] = 1.0 × 10−5 M. (c) The 1H NMR spectra of cis-­13
in CDCl3 before and after irradiation by a 365-­nm portable UV lamp for one hour.
78 3  Aggregation-­induced Emission from the Restriction of Double Bond Rotation at the Excited State

(PF6)2

R N N N N R

Ag Ag

R N N N N R
14 15

Figure 3.24  The structure of AIEgens 14 and 15.

O O O O

CuCl/O2

52%

O O O O
16 17

Figure 3.25  The synthesis of the emissive molecule 17.

very weak fluorescence in solution, the emission of 17 is very strong even in dilute solution. This
fluorescence enhancement after cyclization at the cis-­position should mainly come from the RDBR
process (see Figure 3.25).

3.2.4  Research of Theoretical Calculation on RDBR


The effect of restriction of the double bond rotation can be shown intuitively from the experimen-
tal phenomenon, but to understand in more details the behavior of the double bond in the excited
state and the role it plays, more theoretical studies are needed. In this regard, quantum-­
computational simulation and ultrafast time-­resolved spectroscopy are two major methods. The
former can simulate the changes in molecular energy and structure during the fluorescence pro-
cess, by comparing the energy barriers of different decay routes in the excited state to find where
the nonradiative relaxation takes place. The latter can probe and resolve the excited-­state dynamics
and reaction processes by monitoring the structural changes and the emergence of new species,
finding nonradiative process [49, 50].
Through computational studies, detailed activities of the molecule in the excited state can be
described, especially in single-­molecular behaviors, which is similar to the state of AIEgens in
solution. In this part, an emerging theory called the restricted access to conical intersection (RACI)
mechanism has been used for explaining the AIE effect of many AIEgens. Some conical intersec-
tions (CIs) are blamed for the weak emission of AIEgens in solution, such as π twist, photocycliza-
tion, ring puckering, excited-­state intramolecular proton transfer (ESIPT), bond stretch, and so
on [51–53]. Recent computational studies have contributed to the clarification of the excited-­state
decay process of TPE in the solution state, and the results indicated the presence of a CI in the
ultrafast quenching process of TPE including photocyclization and/or π twist, rather than the
3.2  ­AIE Phenomena and Applications from RDBR Mechanis 79

S1
Relative energy

Cl

S0

0° 90° 180°

The twist angle of double bond

Figure 3.26  A brief illustration of the conical intersection (CI) process through the rotation of double bond
in the excited state for TPE in solution.

propeller-­like rotation of the side phenyl groups. Despite an ongoing debate, there are reports
revealing that the twist of the double bond is a typical CI for the deactivation of TPE. Excitation to
the S1 state (HOMO → LUMO) causes an elongation of the double bond, which initiates the twist-
ing dynamics. This motion stabilizes the first excited state (S1) and destabilizes the ground state
(S0), ultimately causing the two states to become degenerate, which are referred to as CI (see
Figure 3.26). The dynamic process explains subsequent E–Z photoisomerization and weak emis-
sion of isolated TPE molecules.
Zhao et al. [54] report results of the semiclassical simulation study of the excited-­state dynamics
of photoisomerization of TPE. By monitoring the change of the length with time, the stretching
vibrational mode of ethylenic bond in the excited state was examined. When TPE was excited by a
femtosecond laser pulse, the central double bond was excited to stretch from the initial 1.37 to
around 2.20 Å in 300 fs. Then, the twisting motion of the fully extended double bond was activated
by the energy released from the relaxation of the stretching mode, until the central double bond
formed a perpendicular formation and gave an ethylenic bond twisted about 90°. This process was
completed in 600 fs, and this twisted structure remains approximately until about 4800 fs. At 4800 fs,
a nonadiabatic transition to the electronic ground state occurred. The results of the simulation
clearly showed that the ethylenic bond twisting takes place in the subpicosecond scale. This research
first revealed the important influence of twisting of the ethylenic bond on the nonradiative decay of
the photoexcited TPE at molecular levels through the employment of computational studies.
Corminboeuf et  al.  [55] descripted excited-­state dynamics of isolated TPE through trajectory
surface hopping (TSH) simulations using linear response time-­dependent density functional the-
ory (TD-­DFT) within the Tamm–Dancoff approximation (TDA) at the PBE0/def2-­SVP level. By
analyzing motion trajectories, they found that the excited TPE undergoes an ultrafast CI to the
ground state through the rotation of the excited double bond. As shown in Figure 3.27, with the
rotation of the C═C bond, the energy of the initial excited state (S1, red curve) continued to
decrease, but the ground-­state (S0, magenta curve) energy was increasing. After ~1 ps, the S1 state
became nearly degenerate with the ground state and eventually reached the CI between the S1 and
S0 states. To efficiently reach the S0/S1, CIs were responsible for fluorescence quenching after TPE
80 3  Aggregation-­induced Emission from the Restriction of Double Bond Rotation at the Excited State

90°
Ethylenic
twist
Θ
Θ twist


S3
5
S2
4

3 S1
E (eV)

1 S0

0 200 400 600 800 1000


Time (fs)

Figure 3.27  The twist angle of the double bond (upper panel) and electronic-­state potential energies
(lower panel) as a function of time for two representative trajectories showing the ethylenic twist process.
Source: Reproduced with permission from Ref. [55]. Copyright 2016, Royal Society of Chemistry.

Φf.s ≈ 0.1% Φf.s = 64.3%


TPE-4mM TPE-4oM
Figure 3.28  Molecular structures of TPE-­4mM and TPE-­4oM and their fluorescent quantum yields in THF
(Φf.s) are shown below.

photoexcitation in solution. However, there were more trajectories (75%) that decayed to the
ground state through photocyclization. The author also found that the two processes are incompat-
ible. The phenyl rings were initially close to one another and cyclization dominated. As the twist-
ing motion around the central double bond proceeded, the cyclization became inaccessible and
another decay channel (ethylenic twist) opened.
Thiel et al. [56] reported a calculation study of two TPE derivatives, TPE-­4mM and TPE-­4oM
(see Figure 3.28 right) in the isolated gas-­phase state. There is a huge difference of fluorescence
quantum yields between TPE-­4mM (Φf  =  0.1%) and TPE-­4oM (Φf  =  64%) in solution. They
­combined static electronic structural calculations (TD-­DFT, CASSCF, and MS-­CASPT2) and
3.2  ­AIE Phenomena and Applications from RDBR Mechanis 81

OM2/MRCI nonadiabatic dynamics simulations to explore the nonradiative excited-­state decay


mechanisms of them. The computational results showed two pairs of minimum-­energy S1/S0 CI
structures for both TPE-­4mM and TPE-4oM. For TPE-­4mM, there was no barrier to reach the CI
of photocyclization. The energy barrier for CI of the π twist was small (1.8 kcal/mol), indicating
that the rotation of the double bond may also be blamed for the nonemission of TPE-­4mM in
solution. But in contrast, for TPE-­4oM, the ortho-­methyl groups in TPE-­4oM effectively sup-
pressed the rotation of both the phenyl rings and double bond. The energy barriers for the above
two decay paths were non-­negligible barriers (6.2 and 8.4  kcal/mol, respectively), which pre-
vented the nonradiative relaxation of TPE-­4oM. Consequently, the fluorescent quantum yield of
TPE-­4oM was 640-­fold higher than that of TPE-­4mM in solution.
In 2018, Tang et al. [57] studied a series of TPE derivatives with varying structural rigidities and
AIE properties using ultrafast spectroscopy combined with quantum computation. They found
that the stretch and twist of the central double bond in TPE unit upon photoexcitation were two
dominant events that caused nonradiative decay.
Figure 3.29 shows the structures of TPE derivatives 18–23 in the order of increasing structural
rigidity. While 18, 19, and 22 showed typical AIE activity, 20 displayed strong fluorescence in both
solution and solid, but 21 and 23 have very low fluorescence quantum yields in both solutions and
solids. These phenomena do not seem to match the prediction of the RIM mechanism because the
fluorescence quantum yields of their solution should also gradually increase as their rigidity.
However, compounds 22 and 23 with the most rigid structures have very low fluorescence quan-
tum yields in solutions. In contrast, the phenyl rings of compound 20 are not hinged by intramo-
lecular cyclization, but its Φf in solution reached an astonishing 60%. In this case, the exact
mechanism that affects their fluorescent intensity in solution was worth a further study.
Firstly, they explored the geometry changes of 18–20 and 22 and 23 in THF solution from S0 to S1
using DFT calculation. The calculated results revealed that the absolute change of the phenyl tor-
sion and double bond twisting in TPE derivatives decreased as the rigidity of the molecular struc-
ture increases upon excitation. In the excited state, the double bonds of TPE derivatives except for
23 showed a significant extension. Compared with the emission peaks in the film, the fluorescence
emission spectra in dilute solutions displayed extra peaks, which were confirmed to be the emis-
sion peaks of the photocyclization product by experiments. The above results illustrated that both
double bond twisting and phenyl torsion may be responsible for the nonemission of these TPE
derivatives in solutions.
Then, they further constructed the 3D potential energy surface (PES) of 18 in solution (see
Figure 3.30). Along the minimum energy path (MEP) of 18 in the ground state, the phenyl torsion
increased from 50 to 90°, but the change of the twist of double bond (<9°) and potential energy
(PE) (<7  kcal/mol) was slight, indicating that the torsion of the phenyl rings dominated the
ground-­state dynamics in 18. In the excited sate, the stretch and torsion of the double bond resulted
in the FC* geometry changing into minimum energy geometry (S1, minute) along the MEP. In this
course, the twist of double bond was ~50°, which was accompanied by the phenyl torsion with an
amplitude of less than 25°.
The ultrafast time-­resolved spectroscopy was employed to detect the geometry changes and pho-
tocyclized intermediates of 18–23 in excited state. For flexible molecules like 18, 19, and 21, the
measurement demonstrated that the stretch of double bond occurred in the subpicosecond time-
scale (0.6–1.3  ps), and then the stretched double bond began to twist during 1.3–3.79  ps. After
3.79 ps, the photocyclization happened. For 20, due to the steric hindrance from the substituents
at the o-­position, the rotation of the double bond and photocyclization were suppressed and made
the decay of emission band much longer than 18 and 19. For 22 or 23 with a rigid structure, the
82 3  Aggregation-­induced Emission from the Restriction of Double Bond Rotation at the Excited State

Photocyclized QY (%) QY (%)


intermediates solution solid

C3 C21
C1 C22 hν 0.8 24.1
H H
C2 C20
C9 C15

18 18-IM

C22 hν 0.6 30.2


H H
C20

19 19-IM

C22 hν
H3C
H 60 97.6
C20

20 20-IM

C22 hν
H H 1.0 1.3
C20

21 21-IM

C22 hν
H H
C20 0.9 16.5

22 22-IM

C22 hν
C20
H H 0.5 0.7

23 23-IM

Figure 3.29  TPE derivatives 18–23 with increased structural rigidity and their transformation upon UV
irradiation (QY: fluorescence quantum yield). Source: Reproduced with permission from Ref. [57]. Copyright
2018, Royal Society of Chemistry.

formation of the photocyclized intermediate took place directly on the subpicosecond timescale so
that the fluorescence was very weak.
There is a competitive relationship between the two processes of photocyclization and intramo-
lecular rotation. When TPE possessed a flexible structure, the rotations of phenyl rings prohibited
3.2  ­AIE Phenomena and Applications from RDBR Mechanis 83

(a)
0

(Quasi) C=C bond twisting (°)


140
10

kcal/mol
S1 min FC*
120
20
FC* 100
30
S0 min 80
40
60
50
40
60
20 S1 min
70
0
80
(60,24,60)
90
Excited state 0 10 20 30 40 50 60 70 80 90
Energy

Phenyl torsion (°)


(b)
0
S0 min 10
90

C=C bond twisting (°)


80 20
(Q 0 30
ua 70 Ground state 10
si) 60 20 40
C= 50 (8,50,0)
C 30 °) 50
bo 40 40 n(
nd 30 50 r sio 60
lt o
tw 60
ist 20 ny 70
in 10 70
P he
g 80 80
(° 0 90
) 90
0 10 20 30 40 50 60 70 80 90
Phenyl torsion (°)

Figure 3.30  The PES of 18 in the ground state and excited state as a function of the (quasi) C═C bond
twisting and phenyl torsion dihedral angles. (a) Top view of the first excited-­state PES. (b) Top view of the
ground-­state PES. Source: Reproduced with permission from Ref. [57]. Copyright 2018, Royal Society of
Chemistry.

the photocyclization between two adjacent phenyl rings and allowed the excited double bond to
rotate. But when phenyl rings are hinged by ethylene bridge, the short distance of rings promoted
the ultrafast formation of the photocyclization on the subpicosecond timescale, giving no opportu-
nity for the C═C bond to rotate. Just like molecule 20, only after these two nonradiative processes
were blocked at the same time, TPE derivatives could render strong emission.
In 2018, Sada et al. [58] disclosed the RDBR process of disubstituted TPE derivatives TPE-­2OMe
and TPE-­2F through a combination of photochemical experiments and theoretical computations.
As shown in Figure 3.31, E-­or Z-­rich isomers exhibited EZI behavior after the solution was exposed
to UV irradiation. Photostationary state approached in four hours under a deep-­ultraviolet (deep-
­UV) lamp irradiation (6.2 mW/cm2) or in 48 hours under ambient light. Furthermore, the solution
in dark conditions or the solid under a deep-­UV lamp did not show EZI phenomenon, revealing
that the EZI process was triggered by UV irradiation and was suppressed in the aggregated state.
However, no photocyclization was observed in 1H NMR measurements, indicating that isomeriza-
tion was indeed contained in the fluorescence measurement process, rather than the
photocyclization.
The more detailed process was simulated by calculating the steepest-­descent (SD) pathways in
the S1 states for E-­or Z-­isomer, starting from the FC structures. Along the SD pathways, the rota-
tional motion around the central double bond leads to the perpendicular structure. As the change
of the TPE structure, the S1 energy gradually decreased and the S0 energy increased accordingly.
84 3  Aggregation-­induced Emission from the Restriction of Double Bond Rotation at the Excited State

(a) Deep UV lamp to solution (b) Room light to solution


80 80
UV Room light
70 70

60 60
(E) Ratio

(E) Ratio
50 50

40 40
(E)-TPE-2OMe
(Z)-TPE-2OMe
30 30
(E)-TPE-2F
(Z)-TPE-2F
20 20
0 1 2 3 4 5 6 0 6 12 18 24 30 36 42 48
Reaction time (hours) Reaction time (hours)

(c) Solution in dark (d) Deep UV lamp to solid


80 100
70 90
80
60
70
60
(E) Ratio

(E) Ratio

50
50 UV
40
40
Dark
30 30
20
20
10
10 0
0 6 12 18 24 0 6 12 18 24
Reaction time (hours) Reaction time (hours)

Figure 3.31  (a−c) Photoisomerization of TPE-­2OMe and TPE-­2F in chloroform (a) under deep-­UV lamp
irradiation, (b) under ambient-­light irradiation, and (c) in the dark. (d) Photoisomerization of TPE-­2OMe and
TPE-­2F in the solid state. Source: Reproduced with permission from Ref. [58]. Copyright 2017, American
Chemical Society.

Eventually, the S1 and S0 energies came to the closest when the double bond twisted about 90°,
suggesting the existence of a CI near that place (see Figure 3.32).
In addition to free TPE-­2OMe monomer, the behavior in the crystal state was also simulated.
From the crystal computational results, the torsion of the double bond was strictly inhibited by the
other surrounding molecules, leading to only an 8° change of twisting angle. However, the twist of
phenyl rings in the crystal state was identical to that of monomer in the excited state because the
dihedral angle of the phenyl ring showed a similar variation (63° at S0min  →  45° at S1min). This
revealed that the double bond rotation triggered by photoirradiation rather than the phenyl ring
rotation played a key role on the AIE effect.

3.2.5  Other AIEgens Involving RBDR Process


In addition to TPE, there are many other AIEgens with a double bond, in which the RDBR process
is also involved in their luminescence emission.
3.2  ­AIE Phenomena and Applications from RDBR Mechanis 85

4
R
C2
3.5 C3 C1
C4
C5

3 S1
R
Relative energy (eV)

2.5

2
R R
1.5 C2
C3
C2
C3 C1 C1
C4
S0 C4
C5 C5
1 R R
Z E

0.5

0
0 20 40 60 80 100 120 140 160 180
Dihedral angle θ(C2C3C4C5) (°)

Figure 3.32  Energy variations of the S1 and S0 states along the steepest-­descent (SD) pathway in the S1
state of TPE-­2OMe. Source: Reproduced with permission from Ref. [58]. Copyright 2017, American Chemical
Society.

CN CN

NC NC

DCHNT DCHNPH2

Figure 3.33  Molecular structures of dinitriles DCNT and DCNP.

Kobayashi et al. [59] prepared dinitriles DCNT and DCNP (see Figure 3.33) that exhibited AIE
and isomerization properties. When the solution of their E-­or Z-­isomer was exposed to a UV lamp,
the central ethylenic bond substituted by cyano groups could rotate and result in photoisomeriza-
tion and fluorescence quenching. In the packed state, no isomerization of them was observed on
the same experimental conditions due to the locked conformation of the compounds, providing a
bright emission.
The (E)-­CN-­MBE is a typical AIEgen having great photophysical and self-­assembling character-
istics, whose Φf is dramatically enhanced almost 700-­fold from solution to aggregation  [8]. But
(Z)-­CN-­MBE was the opposite, which emitted no fluorescence in both solution and aggregated
states. Park et al. reported that the solid (Z)-­CN-­MBE became intense emissive when it was exposed
86 3  Aggregation-­induced Emission from the Restriction of Double Bond Rotation at the Excited State

to a UV lamp under ambient temperatures due to the EZI process [60]. It was thought that the
bent-­shape structure of (Z)-­CN-­MBE led to loose packing, which was unable to effectively restrict
the rotation of double bond even in the solid state. Therefore, nonradiative photoisomerization
occurred. In contrary, the planar molecular structure of (E)-­CN-­MBE was easier to form tight
packing, effectively blocking the double bond rotation (see Figure 3.34).
This inference was confirmed by Yamamoto’s calculation results [61]. Electronic structural cal-
culations were employed to analyze the mechanisms of AIE and photo/thermal E/Z isomerization
of CN-­MBE. In addition to study the single-­point PE changing based on ethylenic bond rotation
(ϕ) of isolated CN-­MBE, free energy (FE) including thermodynamic influence from the environ-
ment was also considered.
In the PE profile of CN-­MBE, it was revealed that isomerization from E-­ (ϕ = 180°) or Z-­form
(ϕ = 0°) in the S0 state is difficult for CN-­MBE because of the large energy barrier (34 kcal/mol).
However, in the S1 state, the torsional motion of the double bond reduced the energy from the
vertically excited FC points of the E-­or Z-­form to the minimum-­energy point (ϕ = 90°) having no
barrier. And the geometry corresponding to the minimum-­energy point of the conical intersection
(MECI) between the S0 and S1 states of CN-­MBE demonstrated that the C═C bond had a signifi-
cant twisting (ϕ = 75°), indicating that the rotation around the ethylenic C C bond of CN-­MBE
was an important coordinate that led to the S0/S1 CIs.
In the FE profile of CN-­MBE, both the solution state and crystal phase were calculated. In the
THF solution (see Figure 3.35), no energy barrier existed in the FE profile of the S1 state from the
vertically excited FC geometries of the E-­ and Z-­forms (ϕ = 180 and 0°) to the twisted geometry
(ϕ = 90 °). The S0/S1 CIs could be reached efficiently and facilitate fluorescence quenching after
CN-­MBE photoexcitation; the molecule would show no emission when dispersed in dilute
solutions.
In the crystal state, due to packing mode being different for two isomers, the simulation crystal
structure of two forms of CN-­MBE showed that fractional free volumes of the E-­ and Z-­forms of
CN-­MBE were found to be 22.1 and 24.2%, respectively, which indicated that the E-­forms were
more densely packed than the Z-­forms in the aggregated phase.

NC

(E)-CN-MBE

CN Under the room light Under the UV light

(Z)-CN-MBE

Figure 3.34  Photos of (E)-­(above) and (Z)-­CN-­MBE (below) under room light and UV light. Source:
Reproduced with permission from Ref. [60]. Copyright 2013, American Chemical Society.
3.2  ­AIE Phenomena and Applications from RDBR Mechanis 87

(a) (b)
(Z)
45

30 (E)

15
Energy (kJ/mol)

0
100 S1

75 S0

50
25 (Z)
(E)
0

0 30 60 90 120 150 180


Torsion (ϕ) (°)

Figure 3.35  (a) Free-­energy profile of the changes in the torsional angle (ϕ) of the ethylenic C═C bond
site of CN-­MBE in THF solution. (b) CN-­MBE in THF obtained from MD simulations. Source: Reproduced with
permission from Ref. [61]. Copyright 2018, American Chemical Society.

Due to steric hindrance from the close stacking, it was energetically demanding for the rotation
around the ethylenic bond in order to reach the S0/S1 CIs when the (E)-­CN-­MBE crystal was
excited. Therefore, high emission is permitted in the solid phase. However, because of the loosely
packed structure that allowed for the rotation around the ethylenic bond to reach the S0/S1 CIs,
(Z)-­CNMBE did not exhibit fluorescence. Obviously, the restriction of the rotation of a double
bond of (E)-­CN-­MBE is crucial for its emission in aggregates.
Kimizuka et al. [62] demonstrated aggregation-­induced photon upconversion (iPUC) based on
control of the triplet energy landscape. Using AIEgen (2Z,2′Z)-­2,2′-­(1,4-­phenylene) bis(3-­
phenylacrylonitrile) (PPAN) (see Figure 3.36) as an acceptor and PtII octaethylporphyrin (PtOEP)
as a donor, when triplet states of acceptor were populated by a triplet sensitizer in solution, the
TTA-­UC emission was not observed. In contrast, crystalline powder samples displayed a clear UC
emission. For explaining such phenomena, the structure on the ground state (S0) and the excited
triplet state (T1) both in solution and in the crystal was simulated. It was revealed that in solution,
the double bond was dramatically twisted and stretched from S0min to T1min, resulting in the barrier-­
free T1-­to-­S0 intersystem crossing (ISC). And eventually, no fluorescence was observed. However,
in the rigid crystalline state, this conformational-­twisting-­driven transition was effectively
prohibited.
Gierschner et al. [63] prepared a series of dicyano-­distyrilbenzene (referred to DCS) derivatives
with two different CN substitution patterns (α and β in Figure 3.37). The α-­series compounds were
AIE-­active, but the β-­series were radiative in both solution and crystal states except for β-­
DCS. Evidently, this phenomenon contradicted to the principal of RIR; hence, computational stud-
ies were carried out to inspect the main reason for the difference between α-­and β-­series.
The TDDFT and CASSCF calculation results showed that there was a CI between S1 and S0 for
each compound in CHCl3 when the double bond twisted 90° (ϕDB = 90°; see Figure 3.38). The
energy of CI was identical for both α-­and β-­series (2.78 eV), but the barrier was different for them
to reach CI. In α-­series, due to the energies of initially excited FC (EFC), the structure range from
2.9 to 3.2 eV was higher than that of CI, making the CI available for nonradiative decay. In β-­series,
88 3  Aggregation-­induced Emission from the Restriction of Double Bond Rotation at the Excited State

3.0
S0
2.5 S0(T1 structure)
T1
Relative energy (eV)

2.0 1.88 eV
A
1.5 1.62 eV B B

1.0

0.5
Cʹ C
0.0
0 78°(T1 minute)
60 120 180
Dihedral angle θ(C1C2C3C4) (°)
1 cis form 1 trans form
CN CN
C1 θ = 180°
C2 C4 C2 C
C3 C1 C3 4
CN CN
θ = 0°

Figure 3.36  Potential energy curves of S0 at its optimized structures. Source: Reproduced with permission
from Ref. [62]. Copyright 2015, Wiley-­VCH Verlag GmbH &Co. KGaA.

QY (%) QY (%) QY (%)


crystal solution crystal
α-series β-series
NC
90 NC 0.2 0.01 69

CN
CN
α-DCS β-DCS
NC
70 NC OBu 0.2 54 OBu 84
BuO CN BuO
CN

α-DBDCS β-DBDCS

O O NC
66 NC 2 20 73

CN
O CN O
α-MODCS β-MODCS

O O NC
NC OBu OBu
42 0.3 31 46
BuO CN BuO
O CN O

α-MODBDCS β-MODBDCS

Figure 3.37  Molecular structure of α-­(left) and β-­series (right) and their fluorescence quantum yields in
CHCl3 and crystal [63].
3.2  ­AIE Phenomena and Applications from RDBR Mechanis 89

Figure 3.38  Left: TD-­DFT rigid torsional scans of one double bond ϕDB for α-­DBDCS and β-­DBDCS using
the optimized S0 state in CHCl3. Top: CASSCF calculated HOMO-­and LUMO-­like orbitals. Right: TD-­DFT-­
calculated FC energies and DFT-­calculated ground-­state energies. Source: Reproduced with permission from
Ref. [63]. Copyright 2017 American Chemical Society.

the energies of EFC are lower than CI, making it difficult to access CI and eventually showing high
emission in solution. This difference mainly comes from the distinct electronic nature of two
series; the negative charge of the cyano-­group dramatically stabilizes the FC structure of β-­series
due to their bigger resonance structures, but in α-­series, this effect is relatively weaker. In the crys-
tal state, every compound has a bright fluorescence because the significant intermolecular interac-
tion prevented the rotation of a double bond.
Diphenyl dibenzofulvene (DPDBF) is an AIEgen similar to TPE that was first reported by Tang
et al. in 2007 [64]. Probably, the rotation of a double bond in DPDBF is responsible for fluorescence
quenching of its solution similar to that of TPE. Shuai et al. [65] carried out a nonadiabatic dynam-
ics simulation for the excited-­state nonradiative decay processes in open-­and closed-­DPDBF and
showed that the former exhibits an AIE property in contrast to the normal ACQ effect of the latter.
The trajectory for open-­DPDBF showed that, after the initial excitation, the double bond length of
open-­DPDBF increased quickly from its initial value of 1.37 to 1.55 Å after 10 fs, and the double
bond rotation began correspondently (see Figure 3.39). There is a nonradiative transition point at
1206 fs; the energy for the S0 and S1 states approached each other with a gap of less than 0.5 eV. At
this point, the two phenyl rings are nearly coplanar and the DBF ring is approximately perpendicu-
lar to the two phenyl rings. In contrast to open-­DPDBF, the C═C bond of closed-­DPDBF was
restricted and the energy gap was relatively large at ~2 eV. Therefore, the energies of S1 could not
release to S0 through such a point and emission was observed in solution.
90 3  Aggregation-­induced Emission from the Restriction of Double Bond Rotation at the Excited State

(a)
41 37 13 39 35 13
14 14
35 12 34 12
39 37
15
34 15
11 11
36 35 33 16
16
33 21 32 20
5 5
23 22 4
22 4 21 6
6
25 24 3 24 23 3 1
1
2
29 27 2 28 26

(Open-DPDBF) (Closed-DPDBF)

(b)
3
1.55 C21=C33 Energy gap

1.50 1814 cm–1 2


Bond length (Å)

1.50
18.3 fs
1.45 Energy (eV)
1.45 1.388
23.3 fs
1.384 1429 cm–1 1
1.380
1.40 1.04 1.06 10.36 10.38

1.35 0
0 3 6 9 12 15
Time (ps)
(c)
210 3
C22–C33=C21–C4 Energy gap
180

150
150 2
Dihedral angle (°)

1817 f s 18 cm–1
Energy (eV)

120
100

90 50
1
60 0

0 1 2 3 4
30

0 0
0 3 6 9 12 15
Time (ps)

Figure 3.39  (a) The chemical structures for open-­and closed-­DPDBF. The temporal evolution of the
energy gap (S1–S0) (red) and the average values of coordinates (b) C21=C33 (green) and (c) C22–C33=C21–C4
(green). Source: Reproduced with permission from Ref. [65]. Copyright 2012, Royal Society of Chemistry.
3.2  ­AIE Phenomena and Applications from RDBR Mechanis 91

The studies reported above indicated that the single-­molecular nonradiative decay process of
DPDBF mainly resulted from the rotation of the C═C bond, but further supplemental theoretical
research of the AIE effect of DPDBF in the solid phase is needed. Blancafort et al. [66] combined
solution and crystal computational simulation of DPDBF. In solution, the rotation of the C═C
bond could reduce the energy of S1 and eventually decayed further to the ground state through a
(S1–S0) CI seam. But in crystal, the rotation was hindered by the surrounding molecules, which
caused the CI structure to show higher energy than S1. The CI seam is disfavored for solid DPDBF,
and fluorescent intensity is significantly enhanced (see Figure  3.40). In 2015, they further

(a)

S1, S2 40* 1.46 S2-Minac


3.52

59* 90*
1.49 1.51
Eexc
3.54

Eem S1-Minac (S1/S0)-Clac


Calc: 3.04 2.59 2.84
12* Exp: 2.81

1.36
Decay coordinate
S0-Minac +
0.00
Solution

(b)

1.43 1.71
S2-Mincry
4* 3.78 6*

S1
(S1/S0)-Clcry
1.50 4.83

6*

Eexc
S1-Mincry
3.54 3.09

Eem
1.36
Calc: 2.20
2*
Exp: 2.69

Decay coordinate
S0-Mincry
0.00
Crystal
Figure 3.40  Calculated mechanisms for the photophysics of DPDBF in acetonitrile (a) and in the solid
phase (b). Source: Reproduced with permission from Ref. [66]. Copyright 2013, Royal Society of Chemistry.
92 3  Aggregation-­induced Emission from the Restriction of Double Bond Rotation at the Excited State

investigated the MECI of DPDBF in the crystal state [67]. A cluster of 12 molecules (528 atoms)
surrounding each other was relaxed during the MECI optimization, with one molecule being
treated at the QM level. The results confirmed that the AIE effect of DPDBF was due to the packing
of the molecules. Even when the molecules surrounding the excited molecule were allowed to
relax, the rotation of the C═C bond was still hindered and the CI responsible for nonradiative
decay in solution is not accessible energetically.
In addition to these common AIE compounds discussed above, there are more examples to illus-
trate the importance of restricting the double bond rotation for certain AIEgens to render strong
fluorescence. Liu et al. [68] report a computational study on the fluorescence quenching in metha-
nol solution and fluorescence enhancement in crystal for 4-­diethylamino-­2 benzylidene malonic
acid dimethyl ester (BIM).
The push−pull substitution of BIM could lead to a charge-­transfer (CT) structure and result in
the fluorescence quenching of solution. The optimized results of the BIM molecule demonstrated
that the double bond of the S1 state was greatly stretched and its torsion was more serious than S0,
but the twisting of single bonds in the vicinity of a double bond was reduced. An S1 minimum
(referred to as S1-­EM; see Figure 3.41) rendered weak emission and was energetically more stable
than FC because of the torsion of a double bond. In addition, the rotation of both double bond and
adjacent single bond could lead to the S1 state geometry relaxing to the CT structure without bar-
riers. But the energetic decrease from the S1 state was much steeper for the former, suggesting that
the former was the dominant S1 decay channel. Due to the excited molecules decaying to a CT
state, the fluorescence of solution was quenched through the S1/S0 CI near the CT intermediate.
In the crystal state, the simulation works revealed that the energetic difference between FC and
S1-­EM state was much slighter than that of BIM in solution, suggesting that the surrounding mol-
ecules restricted the rotation of both double bond and single bond and blocked the energetic relax-
ation from the intramolecular motions. Moreover, the energy of the CT state was higher than that
of the FC state, and the energy barrier made it impossible for BIM nonradiative decay through
forming CT intermediate. Consequently, high emission channel was accessible for BIM molecules
in crystal states.
Tang et al. [69] prepared a series of benzylidene methyloxazolone (BMO) derivatives with AIE
activities. EZI process was observed in one BMO derivative BMO-­PH by 1H NMR spectra (see
Figure 3.42). When the Z-­isomer in CDCl3 was irradiated by a UV light at 365 nm, the fraction of
E-­isomer increased quickly in the first 35  minutes. To investigate the relationship between the
rotation of a double bond and solution fluorescence quenching, theoretical calculations were

O C OCH3
H3C OCH3
C C
N C
H3C H O
S1-EM S1-CT
S1/S0-Clb S1-EM
S1-CT
AIE
hv

Non-FL

Methanol S0-MIN Crystalline

Figure 3.41  Schematic representation of the conical intersection (left) and AIE mechanisms in BIM. Source:
Reproduced with permission from Ref. [68]. Copyright 2016, American Chemical Society.
3.3 ­Conclusion 93

Hγ Hβ O Hγ Hβ
N H
O 365 nm H
N O Hα
White light O

H H
Z isomer E isomer

100% Z isomer Hα
*

61% Z isomer + 39% E isomer


*

Hγ Hβ

8.5 8.0 7.5 7.0 2.5 2.0


Chemical shift (ppm)

Figure 3.42  EZI process of BMO-­PH that was monitored by 1H NMR spectra. No irradiation (upper spectra)
and irradiation (lower spectra) by a 365-­nm UV lamp for 35 minutes in CDCl3 (40 mM). Source: Reproduced
with permission from Ref. [69]. Copyright 2013, Royal Society of Chemistry.

carried out. The theoretical calculations of BMO-­PH via DFT/TD-­DFT showed that in the ground
state, the energy barrier of a double bond rotation was at least 1.0 eV higher than the single bond
rotation. But in the S1 state, the barrier for the former was dramatically reduced and even lower
than that for the latter. When torsion angles of the double bond were in the range of 70–120°, the
formation of CI of S1/S0 was mainly responsible for the nonradiative decay of BMO-­PH in the
­solution. In the crystal state, no EZI product was detected through 1H NMR, and high emission
was observed.

3.3  ­Conclusions

Most of the AIE molecules, especially the most extensively studied TPE and its derivatives, possess
a critical carbon–carbon double bond. Therefore, whether the RDBR process is involved and plays
a key role in the AIE mechanism is a very concerned question even from the moment when the
AIE phenomenon is discovered. From no effect, minor effect, to key effect and major effect on
the AIE phenomenon, a long time has been passed. Now, the importance of the RDBR process in
the AIE mechanism has been widely recognized and accepted. But the detailed relationship of the
94 3  Aggregation-­induced Emission from the Restriction of Double Bond Rotation at the Excited State

RDBR process and other nonradiative processes needs to be further unraveled and confirmed. The
status and degree of the RDBR process, in general, RIR AIE mechanism, also needs to be further
determined. More importantly, how the RDBR mechanism can be used to design better AIEgens is
what we will do in the future. It is believed that more AIEgens that are based on RDBR mechanism
and have exceptional properties will be developed in the near future.

­References

1 Birks, J. B. (1970). Photophysics of Aromatic Molecules. London: Wiley.


2 Mei, J., Hong, Y., Lam, J. W. et al. (2014). Aggregation-­induced emission: the whole is more
brilliant than the parts. Advanced Materials 26 (31): 5429–5479.
3 Lim, M. H. and Lippard, S. J. (2007). Metal-­based turn-­on fluorescent probes for sensing nitric
oxide. Accounts of Chemical Research 40 (1): 41–51.
4 Tang, C. W. and Vanslyke, S. A. (1987). Organic electroluminescent diodes. Applied Physics Letters
51 (12): 913–915.
5 Luo, J., Xie, Z., Lam, J. W. Y. et al. (2001). Aggregation-­induced emission of 1-­methyl-­1,2,3,4,5-­
pentaphenylsilole. Chemical Communications 381 (18): 1740–1741.
6 Hu, R., Lam, J. W. Y., Liu, Y. et al. (2013). Aggregation-­induced emission of tetraphenylethene-­
hexaphenylbenzene adducts: effects of twisting amplitude and steric hindrance on light emission
of nonplanar fluorogens. Chemistry A European Journal 19 (18): 5617–5624.
7 Tong, H., Hong, Y., Dong, Y. et al. (2006). Fluorescent “light-­up” bioprobes based on
tetraphenylethylene derivatives with aggregation-­induced emission characteristics. Chemical
Communications (35): 3705–3707.
8 An, B.-­K., Kwon, S.-­K., Jung, S.-­D. et al. (2002). Enhanced emission and its switching in
fluorescent organic nanoparticles. Journal of the American Chemical Society 124 (48): 14410–14415.
9 Kokado, K. and Chujo, Y. (2009). Polytriazoles with aggregation-­induced emission characteristics:
synthesis by click polymerization and application as explosive chemosensors. Macromolecules 42
(5): 1421–1424.
10 Wang, M., Zhang, G., Zhang, D. et al. (2010). Fluorescent bio/chemosensors based on silole and
tetraphenylethene luminogens with aggregation-­induced emission feature. Journal of Materials
Chemistry 20 (10): 1858–1867.
11 Chen, J., Law, C. C. W., Lam, J. W. Y. et al. (2003). Synthesis, light emission, nanoaggregation, and
restricted intramolecular rotation of 1,1-­substituted 2,3,4,5-­tetraphenylsiloles. Chemistry of
Materials 15 (7): 1535–1546.
12 Mei, J., Leung, N. L. C., Kwok, R. T. K. et al. (2015). Aggregation-­induced emission: together we
shine, united we soar! Chemical Reviews 115 (21):11718–11940.
13 Feng, H.-­T., Yuan, Y.-­X., Xiong, J.-­B. et al. (2018). Macrocycles and cages based on
tetraphenylethylene with aggregation-­induced emission effect. Chemical Society Reviews 47 (19):
7452–7476.
14 Hong, Y., Lam, J. W. Y., and Tang, B. Z. (2011). Aggregation-­induced emission. Chemical Society
Reviews 40 (11): 5361−5388.
15 Kwok, R. T. K., Leung, C. W. T., Lam, J. W. Y. et al. (2015). Biosensing by luminogens with
aggregation-­induced emission characteristics. Chemical Society Reviews 44 (33): 4228−4238.
16 Hu, R., Leung, N. L., and Tang, B. Z. (2014). AIE macromolecules: syntheses, structures and
functionalities. Chemical Society Reviews 43 (13): 4494−4562.
17 Hong, Y., Lam, J. W. Y., and Tang, B. Z. (2009). Aggregation-­induced emission: phenomenon,
mechanism and applications. Chemical Communications 45 (29):4332−4353.
  ­Reference 95

18 Hu, M., Yuan, Y., Wang, W. et al. (2020). Chiral recognition and enantiomer excess determination
based on emission wavelength change of AIEgen rotor. Nature Communications 11: 161.
19 Huang, J., Sun, N., Yang, J. et al. (2012). Benzene-­cored fluorophores with TPE peripheries: facile
synthesis, crystallization-­induced blue-­shifted emission, and efficient blue luminogens for
non-­doped OLEDS. Journal of Materials Chemistry 22 (24): 12001−12007.
20 Huang, J., Sun, N., Dong, Y. et al. (2013). Similar or totally different: the control of conjugation
degree through minor structural modifications, and deep-­blue aggregation-­induced emission
luminogens for non-­doped OLEDS. Advanced Functional Materials 23 (18): 2329−2337.
21 Yuan, Y.-­X., Xiong, J.-­B., Luo, J. et al. (2019). The self-­assembly and chiroptical properties of
tetraphenylethylene dicycle tetracholesterol with an AIE effect. Journal of Materials Chemistry C 7
(27): 8236–8243.
22 Geddes, C. D. and Lakopwicz, J. R. (2005). Advanced Concepts in Fluorescence Sensing. Norwell:
Springer.
23 Jares-­Erijman, E. A. and Jovin, T. M. (2003). Fret imaging. Nature Biotechnology 21 (11):
1387−1395.
24 Liu, Y., Tao, X., Wang, F. et al. (2007). Intermolecular hydrogen bonds induce highly emissive
excimers: enhancement of solid-­state luminescence. Journal of Physical Chemistry C 111 (17):
6544−6549.
25 An, B.-­K., Lee, D.-­S., Lee, J.-­S. et al. (2000). Microchannel networks for nanowire patterning.
Journal of the American Chemical Society 122 (41): 10232−10233.
26 Li, Y., Li, F., Zhang, H. et al. (2007). Tight intermolecular packing through supramolecular
interactions in crystals of cyano substituted oligo (para-­phenylene vinylene): a key factor for
aggregation-­induced emission. Chemical Communications 45 (3): 231−233.
27 Ren, Y., Kan, W. H., Henderson, M. A. et al. (2011). External-­stimuli responsive photophysics and
liquid crystal properties of self-­assembled “phosphole-­lipids”. Journal of the American Chemical
Society 133 (42): 17014−17026.
28 Xie, Z., Yang, B., Li, F. et al. (2005). Cross dipole stacking in the crystal of distyrylbenzene
derivative: the approach toward high solid-­state luminescence efficiency. Journal of the American
Chemical Society 127 (41): 14152−14153.
29 Zhang, J., Xu, B., Chen, J. et al. (2014). An organic luminescent molecule: what will happen when
the “butterflies” come together? Advanced Materials 26 (5): 739−745.
30 Yuan, Y.-­X., Wu, B.-­X., Xiong, J.-­B. et al. (2019). Exceptional aggregation-­induced emission from
one totally planar molecule. Dyes and Pigments 170: 107556.
31 Luo, J., Song, K., Gu, F. et al. (2011). Switching of non-­helical overcrowded tetrabenzoheptafulvalene
derivatives. Chemical Science 2 (10): 2029–2034.
32 Leung, N. L., Xie, N., Yuan, W. et al. (2014). Restriction of intramolecular motions: the general
mechanism behind aggregation-­induced emission. Chemistry–A European Journal, 20 (47):
15349–15353.
33 Zhao, Z., Zheng, X., Du, L. et al. (2019). Non-­aromatic annulene-­based aggregation-­induced
emission system via aromaticity reversal process. Nature Communications 10 (1): 1–10.
34 Yao, L., Zhang, S., Wang, R. et al. (2014). Highly efficient near-­infrared organic light-­emitting diode
based on a butterfly-­shaped donor–acceptor chromophore with strong solid-­state fluorescence and a
large proportion of radiative excitons. Angewandte Chemie International Edition 53 (8): 2119–2123.
35 Liu, J., Meng, Q., Zhang, X. et al. (2013). Aggregation-­induced emission enhancement based on 11,
11, 12, 12,-­tetracyano-­9, 10-­anthraquinodimethane. Chemical Communications 49 (12): 1199–1201.
36 Kamaldeep, K. S. n., Kaur, S., Bhalla, V. et al. (2014). Pentacenequinone derivatives for preparation
of gold nanoparticles: facile synthesis and catalytic application. Journal of Materials Chemistry A 2
(22): 8369–8375.
96 3  Aggregation-­induced Emission from the Restriction of Double Bond Rotation at the Excited State

37 Banal, J. L., White, J. M., Ghiggino, K. P. et al. (2014). Concentrating aggregation-­induced


fluorescence in planar waveguides: a proof-­of-­principle. Scientific Reports 4 (1): 1–5.
38 Irie, M., Fukaminato, T., Matsuda, K. et al. (2014). Photochromism of diarylethene molecules and
crystals: memories, switches, and actuators. Chemical Reviews 114 (24): 12174–12277.
39 Yuan, Y. X. and Zheng, Y. S. (2019). New acylhydrazone photoswitches with quantitative
conversion and high quantum yield but without hydrogen bond stabilizing (Z)-­isomer. ACS
Applied Materials & Interfaces 11 (7): 7303–7310.
40 Tseng, N. W., Liu, J., Ng, J. C. et al. (2012). Deciphering mechanism of aggregation-­induced
emission (AIE): Is E–Z isomerisation involved in an AIE process? Chemical Science 3 (2): 493–497.
41 Wang, J., Mei, J., Hu, R. et al. (2012). Click synthesis, aggregation-­induced emission, E/Z
isomerization, self-­organization, and multiple chromisms of pure stereoisomers of a
tetraphenylethene-­cored luminogen. Journal of the American Chemical Society 134 (24):
9956–9966.
42 Yang, Z., Qin, W., Leung, N. L. et al. (2016). A mechanistic study of AIE processes of TPE
luminogens: intramolecular rotation vs. configurational isomerization. Journal of Materials
Chemistry C 4 (1): 99–107.
43 Xiong, J.-­B., Feng, H.-­T., Sun, J.-­P. et al. (2016). The fixed propeller-­like conformation of
tetraphenylethylene that reveals aggregation-­induced emission effect, chiral recognition, and
enhanced chiroptical property. Journal of the American Chemical Society 138 (36): 11469–11472.
44 Xiong, J.-­B., Yuan, Y.-­X., Wang, L. et al. (2018). Evidence for aggregation-­induced emission from
free rotation restriction of double bond at excited state. Organic Letters 20 (2): 373–376.
45 Yuan, Y.-­X., Zhang, H.-­C., Hu, M. et al. (2020). Enhanced DNA sensing and chiroptical
performance by restriction of double-­bond rotation of AIE cis-­tetraphenylethylene macrocycle
diammoniums. Organic Letters 22: 1836–1840.
46 Debroy, P., Lindeman, S. V., and Rathore, R. (2009). A versatile synthesis of electroactive
stilbenoprismands for effective binding of metal cations. The Journal of Organic Chemistry 74 (5):
2080–2087.
47 Sinha, N., Stegemann, L., Tan, T. T. et al. (2017). Turn-­on fluorescence in tetra-­NHC ligands by
rigidification through metal complexation: an alternative to aggregation-­induced emission.
Angewandte Chemie International Edition 56 (10): 2785–2789.
48 Zeng, F., Zhao, S., Jiang, Y. et al. (2017). An emissive rigid tetraphenylethylene-­based molecule and
its thermal polymerization. Tetrahedron 73 (30): 4487–4492.
49 Qian, H., Cousins, M. E., Horak, E. H. et al. (2017). Suppression of Kasha’s rule as a mechanism for
fluorescent molecular rotors and aggregation-­induced emission. Nature Chemistry 9 (1): 83–87.
50 Kokado, K. and Sada, K. (2019). Consideration of molecular structure in the excited state to
design new luminogens with aggregation-­induced emission. Angewandte Chemie 131 (26):
8724–8731.
51 Peng, X.-­L., Ruiz-­Barragan, S., Li, Z.-­S. et al. (2016). Restricted access to a conical intersection to
explain aggregation induced emission in dimethyl tetraphenylsilole. Journal of Materials Chemistry
C 4 (14): 2802–2810.
52 Crespo-­Otero, R., Li, Q., and Blancafort, L. (2019). Exploring potential energy surfaces for
aggregation-­induced emission—­from solution to crystal. Chemistry–An Asian Journal 14 (6):
700–714.
53 Ding, W. L., Peng, X. L., Cui, G. L. et al. (2019). Potential-­energy surface and dynamics simulation
of THBDBA: an annulated tetraphenylethene derivative combining aggregation-­induced emission
and switch behavior. ChemPhotoChem 3 (9): 814–824.
  ­Reference 97

54 Zhao, G.-­J., Han, K.-­L., Lei, Y.-­B. et al. (2007). Ultrafast excited-­state dynamics of
tetraphenylethylene studied by semiclassical simulation. The Journal of chemical physics
127 (9): 094307.
55 Prlj, A., Došlić, N., and Corminboeuf, C. (2016). How does tetraphenylethylene relax from its
excited states? Physical Chemistry Chemical Physics 18 (17): 11606–11609.
56 Gao, Y.-­J., Chang, X.-­P., Liu, X.-­Y. et al. (2017). Excited-­state decay paths in tetraphenylethene
derivatives. The Journal of Physical Chemistry A 121 (13): 2572–2579.
57 Cai, Y., Du, L., Samedov, K. et al. (2018). Deciphering the working mechanism of aggregation-­
induced emission of tetraphenylethylene derivatives by ultrafast spectroscopy. Chemical Science
9 (20): 4662–4670.
58 Kokado, K., Machida, T., Iwasa, T. et al. (2018). Twist of C=C bond plays a crucial role in the
quenching of AIE-­active tetraphenylethene derivatives in solution. The Journal of Physical
Chemistry C 122 (1): 245–251.
59 Tasso, T. T., Furuyama, T., and Kobayashi, N. (2015). Dinitriles bearing AIE-­active moieties:
synthesis, E/Z isomerization, and fluorescence properties. Chemistry–A European Journal 21 (12):
4817–4824.
60 Chung, J. W., Yoon, S. J., An, B. K. et al. (2013). High-­contrast on/off fluorescence switching via
reversible E–Z isomerization of diphenylstilbene containing the α-­cyanostilbenic moiety. The
Journal of Physical Chemistry C 117 (21): 11285–11291.
61 Yamamoto, N. (2018). Mechanisms of aggregation-­induced emission and photo/thermal E/Z
isomerization of a cyanostilbene derivative: theoretical insights. The Journal of Physical Chemistry
C 122 (23): 12434–12440.
62 Duan, P., Yanai, N., Kurashige, Y. et al. (2015). Aggregation-­induced photon upconversion through
control of the triplet energy landscapes of the solution and solid states. Angewandte Chemie
International Edition 54 (26): 7544–7549.
63 Shi, J., Aguilar Suarez, L. E., Yoon, S. J. et al. (2017). Solid state luminescence enhancement in
π-­conjugated materials: unraveling the mechanism beyond the framework of AIE/AIEE. The
Journal of Physical Chemistry C 121 (41): 23166–23183.
64 Tong, H., Dong, Y., Hong, Y. et al. (2007). Aggregation-­induced emission: effects of molecular
structure, solid-­state conformation, and morphological packing arrangement on light-­emitting
behaviors of diphenyldibenzofulvene derivatives. The Journal of Physical Chemistry C 111 (5):
2287–2294.
65 Gao, X., Peng, Q., Niu, Y. et al. (2012). Theoretical insight into the aggregation induced emission
phenomena of diphenyldibenzofulvene: a nonadiabatic molecular dynamics study. Physical
Chemistry Chemical Physics 14 (41): 14207–14216.
66 Li, Q. and Blancafort, L. (2013). A conical intersection model to explain aggregation induced
emission in diphenyl dibenzofulvene. Chemical Communications 49 (53): 5966–5968.
67 Ruiz-­Barragan, S., Morokuma, K., and Blancafort, L. (2015). Conical intersection optimization
using composed steps inside the ONIOM (QM: MM) scheme: CASSCF: UFF implementation with
microiterations. Journal of Chemical Theory and Computation 11 (4): 1585–1594.
68 Wang, B., Wang, X., Wang, W. et al. (2016). Exploring the mechanism of fluorescence quenching
and aggregation-­induced emission of a phenylethylene derivative by QM (CASSCF and TDDFT)
and ONIOM (QM: MM) calculations. The Journal of Physical Chemistry C 120 (38): 21850–21857.
69 Jiang, M., He, Z., Zhang, Y. et al. (2017). Development of benzylidene-­methyloxazolone based
AIEgens and decipherment of their working mechanism. Journal of Materials Chemistry C 5 (29):
7191–7199.
99

The Expansion of AIE Thought: From Single Molecule


to Molecular Uniting
Qiuyan Liao1, Qianqian Li1, and Zhen Li1,2
1
 Department of Chemistry, Wuhan University, Wuhan, China
2
 Institute of Molecular Aggregation Science, Tianjin University, Tianjin, China

4.1 ­Aggregation-­Induced Emission

Organic luminescent materials have aroused a great deal of attention in the past few decades, due to
their versatile applications in display, chemical sensing, anticounterfeiting, biological imaging, and so
on [1]. Nevertheless, the practical applications of organic luminescent materials are still limited in
various areas where solid state is necessary. Commonly, it exhibits the phenomenon that materials
always show excellent luminescent performances in solution with high efficiency, while the lumines-
cent intensity sharply decreases or even disappears once aggregates. Taking perylene as an example, it
has bright blue emission in dilute solution while it shows nonemissive performance in aggregated
state with 90% water fraction. This is attributed to strong intermolecular π–π interactions, which can
quench the luminescence through nonradiative transition process as energy loss, and the phenome-
non is defined as aggregation-­caused quenching (ACQ). Gratifyingly, in 2001, Tang et al. proposed the
opposite concept of aggregation-­induced emission (AIE), which is defined from the fabulous emission
performance of hexaphenylsilole (HPS) in aggregated state different from nonemissive solution [2].
Since then, the ingenious AIE phenomenon has been implanted into plenty of organic optoelectronic
materials. Subsequently, with diversified materials floating up, the extensive applications of AIEgens
piques the interest in searching for the principle and inherent mechanism of luminescence [3]. The
AIE mechanism is generally attributed to the restriction of intramolecular motions (including rotation
and vibration), which can suppress the nonradiative transitions and promote the radiative process as
emission efficiently. It is legibly elucidated by HPS molecule, which combined six rotors as peripheral
substituents to form twisted structure. The molecular rotations occur in solution as energy loss easily,
which can be restricted in the solid state largely. Also, the twisted configuration is beneficial to
­suppressing the π–π interaction between adjacent molecules, resulting in the bright emission in the
aggregated states. Subsequently, another typical AIEgen is tetraphenylethene (TPE) with simple struc-
ture, which exhibits a highly twisted propeller-­shaped conformation by the peripheral four phenyl
rings [4, 5]. Following with the development of AIE family, plenty of arenes such as naphthalene,
anthracene, and pyrene derivatives with large π backbones are incorporated, which exhibit wonderful

Handbook of Aggregation-Induced Emission: Volume 1 Tutorial Lectures and Mechanism Studies, First Edition.
Edited by Youhong Tang and Ben Zhong Tang.
© 2022 John Wiley & Sons Ltd. Published 2022 by John Wiley & Sons Ltd.
100 4  The Expansion of AIE Thought: From Single Molecule to Molecular Uniting

Molecular set

UV on UV off
RTP ML TADF Interface

AIE

Aggregate

Non-emissive
PL

Scheme 4.1  The various applications of molecular aggregation: from single molecule to molecular set.

emission properties in aggregated state. The detail structure in solid state is further investigated by
crystal structure analysis, and the key role of different packing modes is highlighted (Scheme 4.1),
which can be proved by polymorphs with different molecular sets of the exactly same molecule [6].
For instance, the emission colors of photoluminescence (PL) materials can be adjustable through
­different molecular packing modes with appropriate molecular design. Also, they can display different
responsive emissions when excited by various stimulations, one is room temperature phosphores-
cence (RTP) property or thermally activated delayed fluorescence (TADF) property excited by
light, while the other one is mechanoluminescence (ML) excited by mechanical force. Furthermore,
the emission colors of ML can be tuned by the varied molecular structure and packing under
mechanical force, accompanying with dynamic color with mechanofluorescence and mechano-­
phosphorescence  [7]. Besides, another fascinating property under mechanical force defined to be
mechanochromism (MC) becomes prevailing, which elucidates the switchable emission colors in
crystal or amorphous state (ground state or as-­prepared state), deriving from the discrimination fea-
ture with or without special molecular packing, respectively [8]. Interestingly, RTP, ML, MC proper-
ties, and even TADF performance can be observed to exist simultaneously in distinguished molecular
packing modes with same molecule structure, manifesting the essential role of molecular packing in
solid state [9]. Apart from these, second-­order nonlinear optical materials are proved to be sensitive to
molecular arrangements with centrosymmetric or noncentrosymmetric modes. Therefore, the diverse
molecular packing modes like herringbone, sandwich herringbone, γ and β motifs bridge the molecu-
lar structure and optoelectronic performance, shedding light on the concept of “Molecular Uniting Set
Identified Characteristic (MUSIC)” [10]. Herein, some elaborate materials were selected to set forth
the relationship between molecular structure, molecular set, and related properties in solid state.
4.2  ­Photoluminescence Materials Based on Molecular Se 101

4.2 ­Photoluminescence Materials Based on Molecular Set

PL has been investigated for decades and it operates based on the Jablonski energy level diagram.
As shown in Figure 4.1, upon light excitation, the electrons jump to the excited singlet state and
then a majority of them can go back to ground state with fluorescence emission, while the others
will arrive at excited triplet state and then back to ground state with phosphorescence emission.
Both fluorescence and phosphorescence belong to radiative transition process, and the remaining
energy can be consumed through nonradiative transition process like thermal. Thus, the efficiency
of fluorescence and phosphorescence emission can be enhanced by increasing radiative rate with
high-­oscillator strength, as well as suppressing nonradiative transition with rigid surroundings.
Selecting large conjugated units seems to be a good method to acquire high oscillator strength;
however, the planar structures always bring strong π–π interactions, which are harmful to achieve
bright emission. Also, the solubility of compounds should be highly noticed. So, the balance
between advantages and disadvantages of conjugation should be taken into consideration for
molecular design, which can largely influence molecular packing in addition to molecular con-
figuration. After accomplishing this point, more exciting phenomena deriving from different
molecular arrangements would be dug out with simple but enjoyable molecule construction [11].
In order to investigate the influence of molecular set in fluorescence emission, different strate-
gies of molecule design were proposed to obtain excellent emission properties. Firstly, chiral seems
to be a good choice to obtain the tunable molecular uniting due to the flexible chiral flip process.
Therefore, tetrahydropyrimidines (THPs) derivatives 1–3 (Figure 4.2) possessing a nonaromatic
chiral central ring were synthesized. All of compounds THP 1–3 possessed different crystal shapes,
which almost showed two PL colors with blue and green emission [12]. Figure 4.2c and d showed
that the R-­enantiomer and S-­enantiomer were arranged in pairs for THP-­1b, while enantiomers in
THP-­1c appeared not. Moreover, every molecular set was reinforced through weak hydrogen
bonds, and all of them showed AIE properties in crystal. In addition, anthracene derivative APA

Sn
IC

Tn
ISC
IC

S1

T1
External conversion

External conversion
Phosphorescence
Fluorescence
Absorption

S0

Figure 4.1  Jablonski energy level diagram of photoluminescence.


102 4  The Expansion of AIE Thought: From Single Molecule to Molecular Uniting

(a) (c)
Br 1b
O
O O Aʹ B C
O N
O N O N
O RS
O
N
O N a
N Bʹ
O O
O Cʹ A c
b
Br
THP-1 THP-2 THP-3
(b) (d)
1b 2c 3b 3p 1c SS

b
C
1c 2cʹ 3c Aʹ
Bʹ c
A B
3 mm 1.5 mm 3 mm 1.5 mm

RR

Figure 4.2  (a) Molecular structure of THPs 1-­3; (b) the polymorphs under UV light (350 nm). Molecular
packing alignments of the R-­enantiomers (blue chiral carbon) and S-­enantiomers (yellow chiral carbon) of
(c) 1b and (d) 1c. 1b and 1c represent the polymorphs of THP-­1; 2c and 2c′ represent polymorphs of THP-­2;
3b, 3c, and 3p represent the polymorphs of THP-­3, respectively. Source: Reproduced with permission from
Ref. [12]. Copyright 2015 The Royal Society of Chemistry.

showed an augmentation with quantum yield (Φf ) of 39% in aggregated state compared to 0.6% in
dilute solution, which is constructed by a π conjugation core and a twisted substituent [13]. The
crystal structure clearly demonstrated a distance of 3.433  Å within two face-­to-­face stacking
anthracene moieties. The adjacent substituents could contact with each other through weak C–
H⋯π intermolecular interactions, providing a rigid surrounding to contain the molecular set with
sandwich herringbone packing mode in solid state (Figure 4.3a–c). The illustration could also be
confirmed by a (9-­anthryl) vinylstyrylbenzene derivative 1,4-­ASB (Figure 4.3d–f), which exhibited
much higher emission intensity in solid state than that in solution [14]. It can be ascribed to the
coexistence of C–H⋯π and π–π interactions in the molecular sets between adjacent molecules for
1,4-­ASB, decreasing the energy loss through nonradiative transition in rigid solid state.
Similarly, compound ditolyldibenzofulvene (DTDBF) was discovered to be AIE active with a bulky
conjugation core and peripheral tolyl groups [15]. Figure 4.4 showed tetra-­color emission depending
on its morphologies with blue crystal CB (461 nm), yellow crystal CY (545 nm), and orange crystal
CO (586 nm) and a faint yellow (557 nm) amorphous state (Am) emission, which might result from
the different torsion angles between bulky conjugation core and peripheral tolyl groups. The photo-
luminescence quantum yield (PLQY) of these four states behaved much difference with 28.1% (CB:
blue crystal), 23.3% (CY: yellow crystal), 16.2% (CO: orange crystal), and 2.9% (Am: amorphous state),
respectively, which can be explained by their crystal structures. The molecules in CB crystal were
fixed by multiple C–H⋯π intermolecular interactions of peripheral adjacent molecules, while others
exhibited a decrease in the sequence of CY, CB, and even no interaction in Am. Multiple C–H⋯π
intermolecular interactions created a rigid environment in CB crystal, impeding the nonradiative
transition process. Not only fluorescence emission, but also phosphorescence emission related to
molecular packing. Once a series of bipyridine/carbazole hybrid compounds with different linkage
positions were applied in phosphorescent organic light-­emitting diodes (PhOLEDs) as host materi-
als, only compound m-­BPySCz stood out with high external quantum efficiency for green (28.0%)
and sky-­blue (27.3%) PhOLEDs as well as low efficiency rolling off [16]. It was attributed to the netted
intermolecular interactions, creating an appropriate channel for electron or hole transportation
4.2  ­Photoluminescence Materials Based on Molecular Se 103

Figure 4.3  (a) Molecular structure and (b) the plot of αAIE (I/I0) versus fw (water fraction) at 461 nm.
(c) Crystal packing mode of APA. Source: Reproduced with permission from Ref. [13]. Copyright 2019 Wiley-­
VCH. (d) Molecular structure and (e) stacking image along the b axis and (f) stacking image in the ac plane of
1,4-­ASB. Source: Reproduced with permission from Ref. [14]. Copyright 2016 The Royal Society of Chemistry.

(a) (b)
CO Am

DTDBF

λem ΦF
CY CB
Material (nm) (%)

CB 461 28.1

CY 545 23.3

Am 557 2.9
CO 586 16.2

Figure 4.4  (a) Molecular structure and optical properties of DTDBF. (b) Photos of CO, Am, CY, and CB state
(taken under UV illumination). Source: Reproduced with permission from Ref. [15]. Copyright 2019 Wiley-­VCH.
104 4  The Expansion of AIE Thought: From Single Molecule to Molecular Uniting

(Figure 4.5). And the moderate distance within adjacent molecules could also suppress the exciton
quenching process, favoring to the stability in different conditions.
Besides C–H⋯π or π⋯π intermolecular interactions in the abovementioned materials, halogen–
hydrogen interactions or halogen–halogen interactions were also commonly acknowledged to be
strong enough in consolidating molecular uniting in aggregated state. Phenanthroimidazole deriva-
tive t-­PhIm-­Thi-­Br with multiple C–H⋯Br interactions emitted a large Stokes shift emission
(yellowish-­green emission) in crystal (Figure 4.6a–d), which showed a head-­to-­tail stacking mode
in dimer [17]. For comparison, the replacements of Br with H, phenyl ring, and Cl were found to be

N
48.47 ° 28.75 °

24.41 °

N
m-BPySCz

Figure 4.5  Molecular structure, single-­crystal structure, packing mode in dimer, and unit cell of m-­BPySCz.
Source: Reproduced with permission from Ref. [16]. Copyright 2018 Wiley-­VCH.

(a) (b)

0 20% 40% 60% 80% 90% 100%


N
S Br
N
0.007 0.019

DMF H2O

t-Phlm-Thi-Br
(c) (d)

Figure 4.6  (a) Molecular structure of t-­Phlm-­Thi-­Br. (b) The photographs in the mixtures of DMF and water
with different fraction taken under 365 nm irradiation. (c) Crystal packing and (d) illustration of head-­to-­tail
packing of t-­Phlm-­Thi-­Br. Source: Reproduced with permission from Ref. [17]. Copyright 2015 The Royal
Society of Chemistry.
4.2  ­Photoluminescence Materials Based on Molecular Se 105

bright blue emission in solution with nearly no AIE effect, which further confirmed the pivotal part
of molecular packing deriving from additional C–H⋯Br interactions, not the heavy atom effect.
However, another novel bodipy-­derived dye BDY-­IN was confirmed to show some difference in
emission colors with fluorine-­containing group or chlorine-­containing solvent [18]. Excitedly, two
different crystal forms (BDY-­O and BDY-­R) were obtained with orange (582 nm) and red (661 nm)
emission, respectively, deriving from distinguished molecular conformations and stacking modes.
For BDY-­R, it showed more planar configuration with small torsion angles between the indene
groups and bodipy moieties when compared to BDY-­O. The chlorine atoms supplied different C–
H⋯Cl interactions between solvent molecules and BDY-­IN molecules, directly affecting mole-
cules communications through C–H⋯π, π⋯π, or C–H⋯F intermolecular interactions
(Figure 4.7a–f). A series of 2,5-­distyrylfuran derivatives with fluorine or cyano substituents showed
enhanced fluorescence intensity in solid state, attributing to the rigid molecular set with multiple

(a) (b)
N F
B
F
N
N
F B
F N

BDY-IN

(c) (e)

c b

CHCI3 a
2
1

(d) (f)

b a

3 4 CH2CI2 c
5
6

Figure 4.7  (a) Molecular structure and (b) photographs of crystals of BDY-­IN. (c) The schematic interactions
between BDY-­IN molecules and solvents and (d) crystal packing diagram of BDY-­O. (e) The schematic
interactions between BDY-­IN molecules and solvent and (f) crystal packing diagram of BDY-­R. Source:
Reproduced with permission from Ref. [18]. Copyright 2011 Wiley-­VCH.
106 4  The Expansion of AIE Thought: From Single Molecule to Molecular Uniting

C–H⋯F, F⋯F and C–H⋯N intra-­ and inter-­molecular interactions [19]. Thus, apart from highly
twisted structures with enhanced luminescence efficiencies, a large amount of molecules with
large π conjugation units may surprise us with delightful performance benefiting from diversified
molecular sets in solid state.

4.3 ­Mechanoluminescence Materials Based on Molecular Set

Distinguished from PL materials, ML or triboluminescence (TL) materials are excited by external


mechanical force. They exhibit bright emission when stretched, rubbed, pressed, or under other
mechanical stimulus in solid state [20]. ML phenomenon has been investigated over hundreds of
years, which can be dated back to 1605 by Francis Bacon, a flash light was observed from a lump
of sugar when scraped. Pure organic materials with strong ML property were scarcely reported, the
major reason may be the ambiguous inherent mechanism of ML property and there was no defi-
nite theory to completely explain all kinds of ML materials. Moreover, the relationship between
ML intensity and mechanical force was also confused to scientists, which may lead us to ignore
some interesting phenomena when exerting force. On the other hand, ACQ phenomenon always
hampers the process of construction in ML molecules due to the π–π quenching in solid state,
which has been conquered after the arrival of the AIE era [21]. Hereafter, more efficient ML mate-
rials were obtained and then showed bright emission through combination with AIE moieties.
Gradually, with the expansion of ML family, various exciting results were discovered from ML
compounds, containing fluorescence-­phosphorescence dual ML emission, tricolor emission, and
the related quantitative research [22].

4.3.1  Mechanoluminescence Materials with Fluorescence Emission


Chi et  al. reported a compound AIE-­DF1 with an asymmetric structure, which emitted bright
green luminescence under scratching (Figure 4.8) [23]. Accordingly, the ML spectrum was in coin-
cidence with PL spectrum. The interactions between planar carbazole moieties and twisted pheno-
thiazine moieties endowed the molecules in an interlocked arrangement with close packing mode,
favoring to ML property. Similarly, tetraphenylethene derivatives TMPE and P4TA were designed
to explore ML properties. The relationship between molecular packing and ML effect was explored
from two polymorphs crystals of compound TMPE by our group (Figure 4.9) [24]. Excitedly, as
shown in Figure 4.9b, Cp-­form demonstrated wonderful ML properties but the Cc-­form was ML
inactive, and the possible reason can be explained by the further investigation of crystal structures.
As expected, the molecular conformations from Cp-­form and Cc-­form are different due to the flex-
ible molecular rotations through single bond, which induced the different torsion angles between
four benzene rings and ethylenic plane. Meanwhile, the intermolecular interactions between
methoxy groups and benzene rings of adjacent molecules were much different. The Cp-­form pos-
sessed much more kinds of C–H⋯π and C–H⋯O interactions than those in Cc-­form, leading to
more compact molecular packing. It was beneficial to suppress the possible energy loss under
external stimulus, resulting in the bright ML emission. Thus, the ML activity can be tunable by the
controllable molecular packing, which was confirmed by another tetraphenylethene derivative
P4TA with two crystal forms [25]. In Figure 4.10a–d, SCg-­form with interlaced molecular packing
showed ML active, in the meanwhile, SCb-­form with an antiparallel staking mode was lack of ML
properties. Unexpectedly, a bright green ML emission can be obtained from SCb-­form after being
exposed to DCM or acetone vapors for several minutes. And the XRD patterns explained the
4.3  ­Mechanoluminescence Materials Based on Molecular Se 107

(a) (c)
O
S N S N
O

AIE-DF1

(b) Mechanoluminescene

Packing
1.0 1.0

Normalized ML intensity
Normalized PL intensity
0.8 0.8

0.6 ML 0.6

0.4 0.4

0.2 PL 0.2

0.0 0.0
400 450 500 550 600
Wavelength (nm)

Figure 4.8  (a) Molecular structure and (b) molecular packing and interactions of AIE-­DF1. (c) Schematic of
ML, PL, and ML spectra of AIE-­DF1. Inset was the ML image. Source: Reproduced with permission from
Ref. [23]. Copyright 2015 Wiley-­VCH.

transformation from SCb-­form to SCg-­form before and after exposure to vapors, suggesting the
pivotal element of molecular set in efficient ML performance. The combination of thiophene
group as a self-­assembled unit seemed to be an efficient approach to capture crystalline poly-
morphs. Accordingly, a recyclable ML material tPE-­5-­MeTh was synthesized to broaden the ML
family, and the resultant crystal-­P1 and crystal-­P2 were cultured to display opposite ML property
(Figure 4.10e–j) [26]. ML-­active crystal-­P1 exhibited a tight and layer-­by-­layer noncentrosymmet-
ric packing mode, while crystal-­P2  without ML demonstrated a centrosymmetric mode with
antiparallel arrangement. The thiophene ring induced multiple C–H⋯π interactions, which not
only affect molecular conformation of tPE-­5-­MeThs in a more twisted structure, but also control
molecular packing in crystal-­P1 largely. More intriguingly, the ML emission would disappear after
continuous grinding, while it could be regained through simple thermal treatment. Powder X-­ray
diffraction (PXRD) pattern showed the ML turn on/off could be attributed to the phase transition
between crystal-­P1 and crystal-­P2, which originated from the flexibility of thiophene moiety and
the resultant assembly effect. Subsequently, a great quantity of materials bearing tetraphenyle-
thene exhibited obvious ML properties [27]. The chemical structures were shown in Figure 4.11,
p-­P4A, m-­P4A, and p-­P4A2 aldehydes bearing with TPE group were expected to be ML and AIE
active, while the ML effect completely disappeared of ketone-­type compound p-­P4Ac even with
the similar TPE core. p-­FP2A with the absence of TPE unit also expressed ML inactive property,
which urged us to think about the inherent mechanism of distinguished performance of these
analogs. The single crystal of these two compounds exhibited varied molecular packing with the
same nonpolar space groups, which manifested the importance of polar space group to ML
­property (detailed information is listed in Table  4.1). Another comparison between ML-­active
mm-TPE(PI)2 compound and ML-­inactive pp-­TPE(PI)2 analogs with different linkage positions
was carried out, which was mainly due to the tight molecular set of mm-­TPE(PI)2 with larger
dipole moment [28].
Apart from materials based on AIE building blocks, triphenylamine as a nontypical AIE moiety
was introduced into molecular construction. A terpyridine derivative BP-­TPY with branched con-
formation showed non-­AIE effect, but bright ML property as the near-­UV emission
(a) O O (c)
ML active Cp-form ML inactive

O O
TMPE
(b)
Cc-form
Normalized intensity (a.u.)

PL- Cp
PL- Am
ML- Cp

Under UV
(d)
400 450 500 550 600 650
Wavelength (nm)
(b) Without UV
Grinding Writing

Fuming Fuming

Under daylight In dark

Figure 4.9  (a) Molecular structure and (b) PL and ML spectra in different state of TMPE. PL-­Cp: PL spectrum of the Cp-­form crystal; PL-­Am: PL
spectrum of the amorphous; ML-­Cp: ML spectrum of the Cp-­form crystal. (c) Stacking mode, images of the ML-­active Cp-­form crystal and ML inactive
Cc-­form crystal. (d) Images of reversible mechanochromic fluorescence under grinding/fuming and writable mechanochromic fluorescence with letters
“ML” of TMPE. Source: Reproduced with permission from Ref. [24]. Copyright 2016 The Royal Society of Chemistry.
(e) c
b a
3.320Å
S C
3.549 Å A
H
(a) (c) b c
2.886 Å
3.257 Å
O a B
Crystal-P1
S C
H
S

tPE-5-MeTh (h)
a
Conformation SCb2
(f) c
b 3.398 Å
2.40Å C-H∙∙∙O
C-H∙∙∙π ML
C
3.639 Å
active A
ment

Conformation SCb1
al treat

C-H∙∙∙π
P4TA
Grindi

Therm

3.069 Å 3.203 Å
ng

(b) B
ML-Cb
ML Crystal-P2
inactive C
ML-Cg H
S
ML-Cbf
Normalized Intensity (a.u.)

First cycle Second cycle n-th cycle


(d) b
(i) (j)
a b
ML Turn on c b
a c
C-H∙∙∙S
C-H∙∙∙π
CH2Cl2 Vapors C-H∙∙∙O
2.66Å

450 500 550 600 Conformation SCg1 C C


Wavelength (nm)
Conformation SCg2 H H
S S

Figure 4.10  (a) Molecular structure of P4TA. (b) ML spectra of P4TA in different phases. Stacking modes and intermolecular interactions in (c) SCb and (d)
SCg. Source: Reproduced with permission from Ref. [25]. Copyright 2015 The Royal Society of Chemistry. (e) Molecular structure and (f) recyclable utilization
of tPE-­5-­MeTh. (g) The molecular geometry and intramolecular interactions and (i) stacking mode of crystal-­P1. (h) The molecular geometry and
intramolecular interactions and (j) stacking mode of crystal-­P2. Source: Reproduced with permission from Ref. [26]. Copyright 2019 The Royal Society of
Chemistry.
110 4  The Expansion of AIE Thought: From Single Molecule to Molecular Uniting

(Figure  4.12a)  [29]. Through the analysis of crystal structure (Figure  4.12b) in detailed, there
existed both intra-­and inter-­molecular interactions between coupled molecules, restricting nonra-
diative transitions during mechanical stimulus. Besides, triphenylamine derivative (TPA-­1BA)
with a benzaldehyde moiety were discovered to be ML active among several similar compounds,
the noncentrosymmetric molecular arrangement, and the high lattice stability should be the main
reason [30]. While TPA-­2BA and TPA-­3BA with two or three benzaldehyde moieties are ML inac-
tive (Figure 4.12c–g). When the benzaldehyde group was replaced by aldehyde unit to construct
compound TPA-­CHO, no ML property was obtained [31]. However, halogen-­substituted deriva-
tives TPA-­CHO-­2X with two fluorine, chlorine, and bromine atoms in another two phenyl rings at
para position were lucky to be ML-­active (Figure 4.13a, b). Taking TPA-­CHO and TPA-­CHO-­2F
as examples, the DFT calculation demonstrated that the halogen-­substituted phenyl rings exhib-
ited more positive charges, while more negative charges existed in halogen atoms or aldehyde
groups (Figure  4.13c). Hence, more efficient intermolecular interactions (C–H⋯π, C–H⋯O,
C–H⋯X, and X⋯X) could be induced to form intensive molecular packing, attributing to intense
ML emission (Figure 4.13d). The introduction of halogen atoms was certificated to be a salient way
to strengthen intermolecular interactions for efficient ML property. Consecutively, with the aim to
building ML active TPA-­X derivatives, the substituted position and types of halogen atoms were
taken into consideration, in which fluorine, chlorine, and bromine were introduced into TPA-­ph
core (Figure 4.13e) [32]. Their ML properties were listed in Table 4.2 and were shown in Figure 4.13f,
the intermolecular interactions in TPA-­mF and TPA-­pF crystal were stronger than those in TPA-
­oF crystal, and the HOMO and LUMO delocalized extremely on different molecules of TPA-­mF
and TPA-­pF, while TPA-­oF did not (Figure  4.13g), which may explain the difference of these
compounds in ML performance. Herein, the detailed investigation responded that three kinds of
halogens showed no obvious heavy atom effect in ML properties, but the position of halogen atoms
did account for the ML properties, resulting in the various intermolecular charge transfer and

H O O
O O O O
CH3 H
H H H

p-P4A m-P4A p-P4A2 p-P4Ac p-FP2A

Figure 4.11  Molecular structures of p-­P4A, m-­P4A, p-­P4A2, p-­P4Ac, and p-­FP2A. Source: Reproduced with
permission from Ref. [27]. Copyright 2016 The Royal Society of Chemistry.

Table 4.1  Data of single crystals.

Compound Crystal system Space group Symmetry Polarity ML activity

p-­P4A Monoclinic P2(1) Noncentrosymmetric Polar Active


p-­P4Ac Triclinic P-­1 Centrosymmetric Nonpolar Inactive
p-­FP2A Monoclinic P2(1)/c Centrosymmetric Nonpolar Inactive
m-­P4A Monoclinic P2(1) Noncentrosymmetric Polar Active
p-­P4A2 Orthorhombic Pna2(1) Noncentrosymmetric Polar Active

Source: Reproduced with permission from Ref. [27]. Copyright 2016 The Royal Society of Chemistry.
(a) (c) (d) Mechanoluminescence (ML)
R1
O O
B TPA-1BA:R1= CHO, R2=R3=H

TPA-2BA:R1=R2= CHO, R3=H


BP-TPY N TPA-1BA (ML)
TPA-3BA:R1=R2=R3= CHO
400 500 600 700 800
R2 R3 Wavelength (nm)
N
N N (e)
(b) a a
b
c a c

C–H∙∙∙ π C–H∙∙∙ π
2.98–3.44 Å 2.99–3.43 Å
C–H∙∙∙O C–H∙∙∙O
2.50 Å 2.68 Å
TPA-1BAa TPA-1BA (ML active) TPA-1BAb

(f) (g)
C–H∙∙∙ π C–H∙∙∙ π
2.92–3.43 Å 2.82–3.48 Å
C–H∙∙∙O C–H∙∙∙O
2.68–3.42 Å 2.59–3.00 Å

b a
c c

TPA-2BA (ML inactive) TPA-3BA (ML inactive)

Figure 4.12  (a) Molecular structure and (b) interactions in single crystal of BP-­TPY. Source: Reproduced with permission from Ref. [29]. Copyright 2018 The
Royal Society of Chemistry. (c) Molecular structures of TPA-­1BA, TPA-­2BA, and TPA-­3BA. (d) ML spectrum and images of TPA-­1BA. Crystal structures of (e)
TPA-­1BA, (f) TPA-­2BA, and (g) TPA-­3BA. Source: Reproduced with permission from Ref. [30]. Copyright 2017 The Royal Society of Chemistry.
112 4  The Expansion of AIE Thought: From Single Molecule to Molecular Uniting

(a) (b) (d)


c b
Halogen atoms TPA-CHOb
TPA-CHO-2F
a b
X c
c

TPA-CHO-2Cl
N TPA-CHOa
TPA-CHO (ML inactive)
O
X
TPA-CHO-2Br a
b
TPA-CHO: X = H c TPA-CHO-2Fb
TPA-CHO-2F:X = F c
TPA-CHO-2Cl:X = Cl “ML Emission” “ML Emission”
TPA-CHO-2Br:X = Br under daylight in dark
a
(c) c

–1.24 eV –1.16 eV TPA-CHO-2Fa


ΔE g = 4.32 eV
ΔEg = 4.07 eV
–5.49 eV TPA-CHO-2F (ML active)
–5.31 eV
(f)
TPA-CHO TPA-CHO-2F TPA-mF TPA-pF TPA-oF
c c c
b b b

(e) O

R1 R2

N R3
ML Weak ML No ML

O (g)
TPA-X (R1, R2, R3 = H or F or Cl or Br)
LUMO –0.26 eV –0.54 eV –0.48 eV
TPA-oF :R1= F, R2 = R3 = H TPA-pCl :R3 = Cl, R1 = R2 = H

TPA-oCl : R1= Cl, R2 = R3 = H TPA-pBr : R3 = Br, R1 = R2 = H


ΔEg = 4.33 eV ΔEg = 4.16 eV ΔEg = 4.15 eV
TPA-oBr : R1= Br, R2 = R3 = H TPA-mF : R2= F , R1 = R3 = H

TPA-ph : R1= R2 = R3 = H TPA-mCl : R2 = Cl , R1 = R3 = H HOMO –4.59 eV –4.70 eV –4.63 eV


TPA-pF : R3 = F, R2 = R3 = H TPA-mBr : R2 = Br, R1 = R3 = H

Figure 4.13  (a) Molecular structure and (b) ML images of TPA-­CHO-­2X. (c) Electrostatic potential diagram
and HOMO, LUMO distribution. (d) Crystal structure and intermolecular interactions of TPA-­CHO and
TPA-­CHO-­2F. Source: Reproduced with permission from Ref. [31]. Copyright 2019 The Royal Society of
Chemistry. (e) Molecular structures of TPA-­X. (f) Molecular packing and (g) HOMO and LUMO distribution of
TPA-­mF, TPA-­pF, and TPA-­oF. Source: Reproduced with permission from Ref. [32]. Copyright 2019 Wiley-­VCH.

Table 4.2  Optical properties and single-­crystal information of the compounds.

Compound Crystal system Space group Symmetry ML property

TPA-­oF Monoclinic P21/c1 Centrosymmetric Inactive


TPA-­oCl Monoclinic P21/n1 Centrosymmetric Inactive
TPA-­oBr Monoclinic P21/n1 Centrosymmetric Inactive
TPA-­ph Monoclinic P21/c1 Centrosymmetric Weak
TPA-­pF Monoclinic P21/c1 Centrosymmetric Weak
TPA-­pCl Monoclinic C1c1 Noncentrosymmetric Active
TPA-­pBr Monoclinic C1c1 Noncentrosymmetric Active
TPA-­mF Monoclinic C1c1 Noncentrosymmetric Active
TPA-­mCl Monoclinic C1c1 Noncentrosymmetric Active
TPA-­mBr Monoclinic C1c1 Noncentrosymmetric Active

Source: Reproduced with permission from Ref. [32]. Copyright 2019 Wiley-­VCH.


4.3  ­Mechanoluminescence Materials Based on Molecular Se 113

interactions in molecular set. With ortho substitution to triphenylamine core, the resultant
TPA-­o-­3COOMe exhibited different molecular conformations for the rotatable single bonds,
which are adjusted by the steric effect and intramolecular interactions between ester and phe-
nyl moieties, favoring to the formation of polymorphs  [33]. Luckily, an interlocked packing
mode and a layer stacking mode lied in two kinds of crystalline states A and B, respectively,
while crystal A exhibited ML-­inactive and crystal B exhibited bright blue ML emission
(Figure 4.14a). Through deeply analysis of structure of crystal A, ML performance is mainly
due to the six kinds of molecular conformations, and the irregular molecular packing with
interlocked mode, together with the moderate intermolecular interactions (Figure 4.14b–d).
Also, polymorphs of compound tPTI-­Bpin displayed that ML property is highly interrelated to
packing modes [34]. Crystalline tPTI-­B1 possessed triangle-­like molecular uniting with ML activ-
ity, while crystalline tPTI-­B2 with parallelogram-­like one was ML inactive. If the tert-­butyl moiety
was removed, the resultant molecule PTI-­Bpin was ML inactive with parallelogram-­like molecu-
lar packing (Figure 4.15a–d). Further detailed analysis of crystal structure revealed that tert-­butyl
group can induce multiple intermolecular interaction with optimized molecular arrangement, in
which tert-­butylbenzene group of tPTI-­B1 contacted with PTI core of adjoining molecule, and
phenyl rings in tPTI-­B2 and PTI-­Bpin kept close to the borate moiety. Thus, triangle-­like and

(a) (b)
TPA-o-3COOMe
Crystal B Crystal B
Normalized ML/PL intensity (a.u.)

H3
OC
CO

N
H3COOC COOCH3

ML
PL
Conformation
350 400 450 500 550 600 650 700
Wavelength (nm)
(c) (d)

Crystal A n m Crystal B
m m
n
m

Figure 4.14  (a) Molecular structure, PL and ML spectra with ML images of crystal B for TPA-­o-­3COOMe.
(b) Molecular conformations in crystal B. Molecular stacking modes of (c) crystal A and (d) crystal B. Source:
Reproduced with permission from Ref. [33]. Copyright 2019 Wiley-­VCH.
Figure 4.15  (a) Molecular structures and (b) PL and ML spectra of crystal tPTI-­B1, tPTI-­B2, and PTI-­Bpin (with images of crystal PTI-­B). (c) ML image of
tPTI-­B1. (d) Crystal images of tPTI-­B1 and tPTI-­B2. (e) Molecular stacking and intermolecular interactions in dimers of tPTI-­B1, tPTI-­B2, and PTI-­B. Source:
Reproduced with permission from Ref. [34]. Copyright 2019 The Royal Society of Chemistry.
4.3  ­Mechanoluminescence Materials Based on Molecular Se 115

(a) (b) (c)


1.0 1.0

Normalized intensity (a.u.)


Normalized intensity (a.u.)
PL-Py-Bpin PL-Py-Br
Monomer-ML ML-Py-Br
ML-Py-Bpin
O 0.8 0.8
Br Excimer-ML
B
O 0.6 0.6 Excimer-ML

0.4 0.4

0.2 0.2

0.0 0.0
400 450 500 550 600 650 400 500 600 700 800
Py-Bpin Py-Br Wavelength (nm) Wavelength (nm)

(d) (e)
End to face Edge to face
geometry geometry

3.511 Å

59°

(Side view) Crystal cell (Side view)

(Top view)
(Top view) Overlap ratio 76.9%
C-H•••π 3.332 Å (6) 3.349 Å (6) 3.424 Å (6) 3.436 Å (6) 3.601 Å (6)
C-H•••π 3.621 Å (6) 3.624 Å (7) 3.631 Å (6) 3.891 Å (5) 3.924 Å (5)
C-H•••π 2.770 Å (3) 3.289 Å (4) 3.391 Å (3) 3.542 Å (3)
3.140 Å (3) 3.173 Å (4) 3.785 Å (4) 3.560 Å (3) 3.577 Å (4) 3.688 Å (4) 3.778 Å (3)
3.786 Å (4) 3.859 Å (5) 3.951 Å (4) C-Br•••π 3.504 Å (8) 3.519 Å (8) 3.641 Å (8) 3.659 Å (8)
3.806 Å (4) 3.929 Å (4) 3.969 Å (4)
C-H•••O 3.025 Å (51) 3.832 Å (52) 3.863 Å (45) C-H•••O 3.765 Å (62) 3.673 Å (54) C-Br•••H 3.674 Å (9) 3.751 Å (9) 3.925 Å (9)

Figure 4.16  (a) Molecular structures of Py-­Bpin and Py-­Br. PL and ML spectra with ML images of
(b) Py-­Bpin and (c) Py-­Br. Crystal structures and intermolecular interactions in (d) Py-­Bpin and (e) Py-­Br.
Source: Reproduced with permission from Ref. [35]. Copyright 2018 Wiley-­VCH.

parallelogram-­like molecular uniting could be established in tPTI-­B1, tPTI-­B2, and PTI-­Bpin


­crystal, respectively, unveiling the relation between ML properties and molecular uniting with dif-
ferent forms (Figure 4.15e). More interesting, bright ML was observed from two pyrene derivatives
(Py-­Br and Py-­Bpin) with planar molecular configuration, which may be originated from its
unique excimer-­like emission (Figure 4.16a) [35]. As for Py-­Br, a bright blue ML phenomenon
appeared when being scratched, corresponding to excimer emission in solid state. The crystal
structure afforded the explanation that Py-­Br crystal existed face-­to-­face packing dimer with
3.604 Å interplay distance and large overlap as high as 76.9%, leading to excimer dominant emis-
sion under mechanical force in Figure 4.16d. However, ML spectra of Py-­Bpin exhibited two emis-
sion peaks, which can be ascribed to monomer and excimer emission, respectively (Figure 4.16b,
c). The crystal structure of Py-­Bpin presented weak overlap within two adjacent molecules, but
strong intermolecular interactions (C–H⋯π and C–H⋯O), allowing its contemporary monomer-­
like ML and excimer-­like ML (Figure  4.16e). Thus, the synergistic interplay of the overlap and
intermolecular interactions resulted in the largely unequal molecular uniting, accompanying with
differential ML properties.

4.3.2  Mechanoluminescence Materials with Mechanical Induced Dual-­or


Tri-­color Emission
Until now, a large number of ML materials have been explored, while they always exhibited the
fixed emission wavelength and forms, discouraging their applications in colorful areas. Is change-
able ML properties possible? With the puzzle in mind, the first switchable ML effect was achieved
116 4  The Expansion of AIE Thought: From Single Molecule to Molecular Uniting

by phenothiazine derivatives, which can realize dynamic molecular conformations with quasi-­
axial and quasi-­equatorial forms. As a result, compound FCO-­CzS presented dynamic ML from
blue (449  nm) to white and then to yellow (570  nm) upon continuous mechanical stimulus
(Figure 4.17a, b), which was related to two distinguished conformations of FCO-­CzS molecule [36].
The quasi-­axial one was ascribed to blue emission while the quasi-­equatorial one was responsible
for the yellow one. Continuous grinding can change the proportion of these two emissions, leading
to the formation of white emission in Figure  4.17c. The efficient intermolecular interactions,
including C–H⋯F and C–H⋯O interactions (Figure  4.17d) restricted the possible slippage as
energy loss under mechanical force, contributing to ML effect.
Although the tricolor ML emission was realized, the appropriate candidates for switchable ML by
changeable conformations are still limited. With the consideration of ML-­active phosphorescence
emission as tunable colors, the dual mechanical induced fluorescence and phosphorescence were
achieved, regardless of the transition forbidden principle from excited singlet state to triplet state.
The first example of fluorescence-­phosphorescence dual ML was observed from compound DPP-­BO
by our group [37]. The ML spectra of DPP-­BO showed two emission peaks with 350 and 450 nm,
which was coincident with PL spectra at 77 K (Figure 4.18a, b). The peak with shorter lifetime origi-
nated from fluorescence emission, while the other one was ascribed to phosphorescence emission
with the lifetime of several seconds. Similar to previous reports in our group, the brilliant ML prop-
erties can be explained by efficient molecular packing with strong intermolecular interactions. The
crystal structure of DPP-­BO (Figure 4.18c) illustrates two kinds of packing modes with different
strength of intermolecular interactions. Theoretical calculations supported that intersystem cross-
ing (ISC) channels were opened more possibly for molecular uniting with strong intermolecular
interactions, which explained the coexistence of fluorescence and phosphorescence emission under
mechanical stimulation (Figure 4.18d). Later, another ML system based on phenothiazine deriva-
tives CzS-­CH3 and CzS-­C2H5 was developed (Figure 4.19a) [38]. The heterocyclic phenothiazine
containing N and S atoms will provide enough n electrons to stimulate ISC process between excited
singlet state and excited triplet state, emitting efficient phosphorescence. Accordingly, fluorescence-­
phosphorescence dual emission upon grinding was achieved, closely relating to the edge-­to-­face
stacking mode with strong intermolecular C–H⋯π and C–H⋯N interactions in crystals
(Figure 4.19b). Zhang and coworkers reported CX crystal as an ultralong mechano-­phosphorescence
material, stemming from intermolecular electronic coupling (IEC) of molecular uniting, combined
with different molecular configurations in excited state (Figure 4.19c, d) [39]. There existed elec-
tronic coupling between n orbitals of carbonyl moiety and π orbitals of carbazole moiety in aggre-
gated state, which was previously proved to be effective in promoting spin-­orbital coupling (SOC)
(Figure  4.19e), and the proposal was further confirmed by theoretical calculations of transition
properties between S1 and Tn states in Figure  4.19f. Similarly, another boron-­containing pyrene
derivative showed unique mechanical force-­induced RTP and fluorescence emission, which was
related to the dimeric association through C–H⋯π interactions [40]. Moreover, the unique mechano-­
responsive phosphorescence can be obtained in amorphous state, which may result from the favora-
ble dimer formation upon mechanical force.
Apart from these, the formation of H-­aggregates can also be an efficient strategy to produce
mechao-­phosphorescence emission. The feasible function of H-­aggregates was confirmed by
­compound ImBr, which procured changes from blue fluorescence emission (460 nm) to dual ML
emission with blue fluorescence and yellow phosphorescence (527 nm) in one second (Figure 4.20a–
c) [41]. The dual ML spectrum is similar with PL spectrum, showing the same photophysical pro-
cess from excited state to ground state. The introduction of bromine atom and trifluoromethyl
group could supply enough intermolecular interactions to restrict intramolecular motions for the
4.3  ­Mechanoluminescence Materials Based on Molecular Se 117

(a) O (d)

N F
S

FCO-CzS
(b)
Normalized PL intensity (a.u.)

Crystal
Ground lightly
Ground heavily
Doped in PMMA Non centrosymmetric space group
Effective hydrogen bonds
(ML Active)
400 450 500 550 600 650 700
Wavelength (nm)
(c)

Crystal Ground lightly

Quasi-axai

Antiparallel arrangement

(Weak PL in crystal)
Doped in PMMA Ground heavily

Figure 4.17  (a) Molecular structure of FCO-­CzS. (b) Normalized PL spectra and (c) fluorescence pictures of
FCO-­CzS in different solid states. (d) Crystal structure and intermolecular interactions of FCO-­CzS. Source:
Reproduced with permission from Ref. [36]. Copyright 2018 Wiley-­VCH.

dimer formation, resulting in more efficient ISC channels and larger dipole moments compared to
single molecule (Figure 4.20e). Meanwhile, tricolor emission switching between blue, white, and
yellow colors can be easily achieved by UV excitation (Figure  4.20d), not mechanical force.
Alternatively, fluorine-­based derivative BrFlu-­CBr exhibited tricolor emission upon mechanical
stimulus (Figure  4.21a), which presented four Br⋯Br interactions shared by two molecules in
crystal [42]. Moreover, the heavy effect of bromine was verified to be an effective strategy to pro-
mote ISC process, providing possibilities to phosphorescence emission, while Br⋯Br interactions
(a) (b) (c)
Normalized PL/ML intensity (a.u.)

O O
B

Mechanoluminescent

DPP-BO Solid (298 K) Solid (77 K)


Weak coupled DPP-BO Crystal cell Strong coupled DPP-BO
(top view) (top view)
320 370 420 470 520
Wavelength (nm)
(d) 5.0
5.0
4.8 T18
T19
4.6 +0.3 eV T17
+0.3 eV S1 4.5 S1 T13
4.4 T11 T11
Energy level (eV)

T12
Energy level (eV)

–0.3 eV
4.2 –0.3 eV T10
T9,T10
T8
4.0 4.0 T5,T6 Weak coupled DPP-BO Strong coupled DPP-BO
3.8 (side view) (side view)
T4
T3,T4
3.6 T3
3.5 C-H···π 3.724 Å (2)
3.4 C-H···π 3.518 Å (2)
T1,T2 T1,T2 C-H···π 3.747 Å (2) C-H···π 3.761 Å (2)
3.2
C-H···O 3.540 Å (2) C-H···π 3.996 Å (2)
3.0 Weak coupled DPP-BO 3.0 Strong coupled DPP-BO
C-H···O 3.184 Å (2)

Figure 4.18  (a) Molecular structure and (b) PL spectra at different temperatures and ML spectra with ML picture of DPP-­BO. (c) Intermolecular interactions
in crystal cell and coupled molecules. (d) Energy level diagrams of coupled DPP-­BO. Source: Reproduced with permission from Ref. [37]. Copyright
2017 Wiley-­VCH.
4.3  ­Mechanoluminescence Materials Based on Molecular Se 119

(a) (c) O
S S

N
N N
O
CH3 C2H5
CzS-CH3 CzS-C2H5 CX

(b) (d)

n unit
o n

π
π unit

S*1
Hybrid ISC
C-H...π 3.00 Å, 3.20 Å, 3.32 T*1 nπ* ππ*
Å, 3.34 Å, 3.60 Å, 3.90 Å ππ* ππ*
C-H...N 3.34 Å, 3.36 Å
S*0

T5
(e) (f)
∠ θ = 55.7° T4
C +0.3 eV
B
A H L (27.9%)
H L + 1 (48.4%)
S1 T3
T2 H L (3.1%)
H L H L + 1 (14.2%)
H L+1 T1
H L (66.1%)
H L + 1 (23.8%)
–0.3 eV
A: 3.323 (Å) B: 3.370 (Å) C: 3.699 (Å) Coupled CX

Figure 4.19  (a) Molecular structure of CzS-­CH3 and CzS-­C2H5. (b) Intermolecular interactions in dimer of
CzS-­C2H5. Source: Reproduced with permission from Ref. [38]. Copyright 2017 Wiley-­VCH. (c) Molecular
structure and (d) schematic diagram of intermolecular electronic coupling of CX. (e) Intermolecular
interactions and (f) energy level diagram of coupled CX. Source: Reproduced with permission from Ref. [39].
Copyright 2018 The Royal Society of Chemistry.

could also be strong enough to maintain rigid surroundings to suppress nonradiative transition
under mechanical force (Figure 4.21b). The synergy between the promoted ISC and stabilization
of triplet states became the vital role to acquire persistent RTP, resulting in the dual emission under
external mechanical force. The emission color changed from cyan to blue with the different pro-
portion of mechanofluorescence and mechano-­phosphorescence emission during the continuous
grinding process. Excitedly, after stopping mechanical force, green-­white phosphorescence was
obtained, resulting in the tricolor emission by mechanical stimulus (Figure 4.21c).
Figure 4.20  (a) Molecular structure of ImBr. (b)–(d) PL and ML pictures and spectra of ImBr.
(e) Intermolecular interactions in ImBr crystal. Source: Reproduced with permission from Ref. [41].
Copyright 2018 Wiley-­VCH.

(a) (c) RTP UV on


Light produced Crystal
Br
At the very be
gin
nin
g
Br
Ground UV off
Stopping m

Br
Aft
er

Normalized intensity (a.u.)


e

ML-at the very beginning


a
ch

lo
ni ng
a

ca
TL
tim
ls
ti m u
e
lu s

Mechanical stimulus
BrFlu-CBr

(b)
a

c
3.528 Å
3.125 Å
ML-after stopping
mechanical stimulus
3.528 Å ML-after mechanical stimulus
3.125 Å
for a long time

300 350 400 450 500 550 600 650 700 750
Wavelength (nm)

Figure 4.21  (a) Molecular structure and multicolor switching of BrFlu-­CBr. (b) Molecular stacking and
Br⋯Br interactions in BrFlu-­CBr crystal. (c) PL and ML spectra upon grinding of BrFlu-­CBr (inset ML
pictures). Source: Reproduced with permission from Ref. [42]. Copyright 2018 Wiley-­VCH.
4.3  ­Mechanoluminescence Materials Based on Molecular Se 121

4.3.3  Quantitative Research of Mechanoluminescence Property


Although many ML materials have been reported with excellent performance, the practical appli-
cation is limited due to the elusive quantitative relationship between the external force and ML
intensity. Based on the recyclable ML property of compound tPE-­5-­MeTh as mentioned above,
our group further obtained a quantitative relationship between pressure and emissive intensity by
luminogen tPE-­2-­Th and tPE-­3-­Th, which was accomplished by ML devices in Figure 4.22a–d
(type-­A device constructed by crystals with big sizes and type-­B device constructed by crystals with
small sizes) [43]. The involved curves would be divided into two monotonically linear increasing
parts, and it would stay a stable intensity level, when pressure was above 14 N for type-­A device
and 6.5 N for type-­B device. Both of the unique quantitative analysis and the different performance
of devices were attributed to the varied molecular packing and switchable molecular structures.
The special self-­assemble property of thiophene groups can form many metastable aggregated
states to maintain a certain ML intensity under external stimulus (Figure 4.22c). The quantitative
ML property can be applied into encrypted communication, impact strength warning, information
storage, and heartbeat detection (Figure 4.23), which broadened the understanding and applica-
tion of ML materials.

Figure 4.22  (a) Molecular structure of tPE-­2-­Th and tPE-­3-­Th. (b) PL and ML spectra and (c) crystal images
and crystal structure of tPE-­2-­Th and tPE-­3-­Th. ML spectra under different pressure forces from (d) device A
and (e) device B. Source: Reproduced with permission from Ref. [43]. Copyright 2019 Elsevier.
122 4  The Expansion of AIE Thought: From Single Molecule to Molecular Uniting

Figure 4.23  Schematic diagram of practical application based on tPE-­2-­Th. Source: Reproduced with
permission from Ref. [43]. Copyright 2019 Elsevier.

4.4 ­Mechanochromism Materials

MC materials always engender intriguing emission color conversion ranging from deep blue to
red area under mechanical force, while they can be reversible through other ways such as heat,
fuming with organic solvent in most cases. They exhibited potential applications in sensitive
sensor, display, data storage, and anticounterfeiting. Normally, the transition of aggregated
state during mechanical stimulation should be responsible for MC materials, including one
crystalline to another crystalline, crystalline to amorphous states, from stable crystalline to
metastable liquid crystalline phase, and so on. All of these can be ascribed to the switchable
molecular conformations and molecular packing modes upon stimulus of an external
force [44].

4.4.1  Mechanochromism Materials Based on Polymorphs


Polymorphs with MC effect can afford crucial information to investigate the relation between
molecular packing and switchable luminescence properties, since they are from the same molec-
ular structures. Phenothiazine is a versatile unit for the changeable quasi-­axial and
4.4 ­Mechanochromism Material 123

quasi-­equatorial structure with folded and planar conformation, respectively, which can display
plentiful emission colors under different conditions. For molecule DOS illustrated in Figure 4.24,
10 kinds of crystals with distinguished emission colors from blue to orange can be obtained by the
various configurations of phenothiazine moiety [45]. However, all of them exhibited green emis-
sion in amorphous state after grinding, confirming the key role of molecular conformations, and
molecular packing to luminescence property. Polymorphism-­dependent fluorescence emission
can also occur in a diphenyldibenzofulvene derivative p-­FP2OC3 (Figure 4.25) [46]. The molecu-
lar structure adopted twisted configuration in crystal G with green emission and crystal Y with
yellow emission, which can be changed to more planar structure upon external mechanical force,
accompanying with the emission color changing to orange in amorphous state. Thus, three distin-
guished emission colors from the same molecule can be achieved, which can be switched by
grinding or heating. Similarly, a difluoroboron derivative BF2AVB was also reported to be
morphology-­dependent emission, accompanying with reversible MC property (Figure  4.26a–c)
[47, 48]. Large prism-­like crystal with green emission (505 nm) and needlelike crystal with cyan

(a) O O (c) DOS-O-EA


S

DOS-O-ACE DOS-O-nH
N N

O S DOS-O-TCM DOS-O-THF

G
DOS
H DOS-S-ACN DOS-O-TOL
H N
N
S
DOS-S-MeOH DOS-S-EtOH
S

Planar Folded DOS-S-DCM


(b)
DOS-O-TCM DOS-O-ACE DOS-O-EA DOS-O-nH DOS-O-THF

DOS-O-TOL DOS-S-EtOH DOS-S-DCM DOS-S-MeOH DOS-S-ACN

Figure 4.24  (a) Molecular structure and design strategy of DOS. (b) Luminescence images of 10 single
crystals of DOS. (c) Photographs of different crystal and amorphous forms of DOS under UV light.
Source: Reproduced with permission from Ref. [45]. Copyright 2019 Wiley-­VCH.
124 4  The Expansion of AIE Thought: From Single Molecule to Molecular Uniting

(a) (c) (d)

O O

p-FP2OC3

(b) (e) (f)


Normalized PL intensity

420 510 600 690


(g) C—H• • •π (h)

2.90Å

2.84Å 2.84Å

2.90Å
C—H• • • O
2.57Å
C—H• • •π

Figure 4.25  (a) Molecular structure of p-­FP2OC3. (b) Normalized PL spectra of p-­FP2OC3 before and after
grinding. (c)–(f) Photographs of p-­FP2OC3 before and after grinding. Intermolecular interactions in (g)
yellow-­emissive crystal and (h) green-­emissive crystal. Source: Reproduced with permission from Ref. [46].
Copyright 2011 Wiley-­VCH.

emission (470  nm) were cultivated (Figure  4.26b). In green crystal, the planar structure of
BF2AVB with strong conjugated effect should be the main reason for the red-­shifted emission.
However, the molecules in cyan crystal possessed more twisted conformation, which exhibited
parallel packing mode with the tert-­butyl groups on the same side (Figure 4.26f). Nevertheless,
both green and cyan crystals emit yellow emission upon grinding, and the MC process could be
easily reversed to initial state. The mechanism of MC behavior might depend on the changeable
molecular configuration and interlay interactions. All in all, the flexible rotation between
4.4 ­Mechanochromism Material 125

(a) (b)
F F
B
O O

O 1 mm 1 mm 1 mm
BF2AVB
(c) (f)

(001)

(d)
3 4
2
4

1
3 1
2

(e)
Slip plane (-OMe)

Slip plane

Slip plane (-t-Bu)


(π–π)

c
(–120)
a c*
b a
b

Figure 4.26  (a) Molecular structure of BF2AVB. (b) Photographs of green crystal, cyan crystal, and blue
solid under UV excitation. (c) Mechanochromic fluorescence of BF2AVB solid films. (d) Crystal packing and
Hirshfeld surfaces analysis for green and cyan BF2AVB crystals. Crystal packing of (e) green crystal and
(f) cyan crystal. Source: Reproduced with permission from Ref. [47]. Copyright 2013 Wiley-­VCH. Reproduced
with permission from Ref. [48]. Copyright 2010 American Chemical Society.

difluoroboron ring and phenyl ring, as well as tunable stacking mode induced by the strong C–
H⋯F interactions should be taken into consideration for MC-­active materials. Based on this,
another difluoroboron β-­diketonate derivative with four polymorphs and one amorphous state
was cultivated and confirmed to be MC-­active material [49]. From green-­emissive crystal (two
types) to yellow-­emissive crystal and then to orange-­emissive crystal, the molecular uniting var-
ied from dimer-­like structure with strong π–π interactions to monomeric stacking mode with no
π–π interaction. It exhibited red emission in the amorphous state after grinding, due to the change
of molecular conformations and molecular packing.

4.4.2  Mechanochromism Materials Based on Excimer Emission


The excimer emission is usually formed by the polycyclic moieties, such as anthracene, pyrene,
and perylene, with dimer structures in the aggregated state. Herein, three polymorphs of an
anthracene derivative BP2VA showed distinguished emissions with C1 (green, 527  nm), C2
126 4  The Expansion of AIE Thought: From Single Molecule to Molecular Uniting

Figure 4.27  (a) Molecular structure of BP2VA. (b) Photographs of the ground powder and the heated
powder under UV light. (c) Stacking modes in three single crystals. Source: Reproduced with permission
from Ref. [50]. Copyright 2012 Wiley-­VCH.

(orange, 579 nm), and C3 (red, 618 nm), respectively (Figure 4.27) [50]. The increased π–π inter-
actions from C1 to C3 were observed among anthracene cores. Interestingly, the green emission
would change to orange one after grinding (Figure 4.27b), corresponding to the crystal forms
from C1 to C2. Meanwhile, another two analogs of Py-­H and pyrene were also taken into con-
sideration. Py-­H and Py-­Bpin were MC active, while Py-­Br and pyrene were MC inactive.
Through intensive analysis of molecular packing and comprehensive characteristics of photo-
physical properties, the weak overlap in the dimer of Py-­H and Py-­Bpin caused them undergo
a transformation from monomer-­like packing mode to static excimer one under grinding, result-
ing in the red-­shift of emission with increased π–π interactions. Concerning excimer emission of
pyrene moiety, a lot of pyrene derivatives with tunable steric or electronic effect were developed,
exhibiting different photoluminescent properties for the controllable molecular packing
(Figure 4.28a) [51]. Two tert-­butyl units were introduced to 2,7-­site of pyrene as isolate group to
inhibit the π–π stacking. Multiple electron-­withdrawing moieties were also incorporated to tune
the electrostatic potential of the resultant molecules (Figure  4.28b, c). When stimulated by
Figure 4.28  (a) Molecular structure. (b) Electrostatic potential. (c) Molecular packing of PPOCH3, PPF, PPCN, PPCHO, and PPCF3. (d) The switch between
luminescence dark and bright state of PPCHO. (e) HOMO, LUMO distribution, and (f) energy diagram of coupled PPCHO. (g) Molecular structure of PPCN2,
PTCN2, and PTCHO. Source: Reproduced with permission from Ref. [51]. Copyright 2020 Wiley-­VCH.
128 4  The Expansion of AIE Thought: From Single Molecule to Molecular Uniting

external force, all of these crystals emitted red-­shifted fluorescence and marginally difference of
PLQY except for PPCHO. PPCHO showed an augmentation in PLQY from crystal (0.05%) to
ground state (20.07%) (Figure 4.28d, g, and related table), accompanying with monomer-­like to
excimer-­like packing mode. The weak emission of PPCHO as crystals mainly derived from the
improved ISC channels by aldehyde group and the obvious intermolecular charge transfer
(Figure  4.28e, f). Also, the emission lifetimes increased from 0.38 to 4.98  ns with off–on MC,
which made it possible for applications in encryption and pressure sensor. Apart from pyrene
derivatives, another tetrathiazolylthiophene fluorophore can realize mechaochromism property
from monomer-­like emission to excimer one under anisotropic stress of mechanical force, which
also confirmed the direct relation of MC property and molecular uniting with the aid of multiple
intermolecular interactions [52].

4.4.3  Other Kinds of Mechanochromism Materials


Thanks to the effort of scientists, fruitful achievements based on MC molecules were achieved.
Apart from those mentioned above, M-­4-­B exhibited sequential tricolor emission from deep-­
blue to bluish-­green, then to a reddish one during the ground process (Figure 4.29) [53]. The
initial deep-­blue emission is related to the antiparallel molecular packing with multiple N–B⋯π,
C–H⋯π, π–π, and C–H⋯O intermolecular interactions. The bluish-­green emission might result
from amorphous state of M-­4-­B, and subsequently a ring-­opening state of RhB should be respon-
sible for the reddish color. More interestingly, the switch of emission color for some MC materi-
als sometimes could be preliminary predicted based on the changeable dipole moments of polar
molecules (2,7-­diaryl-­TAPs) in Figure 4.30 [54]. The molecules with small dipole moments usu-
ally exhibited red-­shifted MC effect, while those with large dipole moment always displayed a
blue-­shifted emission under mechanical force. The inherent mechanism was investigated by
single-­crystal analysis and DFT calculations. Apparently, the introduction of electron-­donating
groups to phenyl A and electron-­withdrawing groups to phenyl B resulted in the smaller change
of dipole moments, while the larger changes were obtained through aryl-­exchange. Taking 5df
(electron-­donating group in phenyl ring A and electron withdrawing group in phenyl ring B) and
5fd (adverse to 5df) as examples, an obvious face-­to-­face antiparallel packing between two vici-
nal 2-­aryl-­TAP planes was induced by weak π–π interactions. This may be related to the twisted
molecular structure by the rotation of ring B. Actually, it would become planar upon grinding,
leading to red-­shifted emission from 524 to 588 nm. However, the interchanged 5fd displayed an
unusual cross-­parallel packing mode with the 3,5-­bis-­(trifluoromethyl)phenyl moiety and
4-­dimethylaminopheyl moiety of two different molecules in a cross-­oriented parallel arrange-
ment. Consequently, a red-­shifted fluorescence (624 nm) was acquired in crystal 5fd with strong
π–π interactions, which can be damaged after ground, resulting in the blue-­shifted emission
(563  nm) as amorphous state. In this system, the key point of MC properties is the packing
modes, and the deep mechanism of different dipole moments should be taken into considera-
tion, which was further related to molecular design with electron donating or withdrawing
groups in opposite position. Luckily, the distinctive performance can supply enough information
for molecular design and then make difference in particular red-­shifted or blue-­shifted MC
properties.
(a) (b) (c)
N O N
H H 1.0
H B
N Original powder
N

Normalized intensity
0.8 Slightly grinding
H O Heavily grinding
4.463 pm
0.6
4.209 pm
4.737 pm
0.4
4.715 pm
4.209 pm
0.2
4.463 pm
M-4-B
0.0
400 450 500 550 600 650 700
Wavelength (nm)

(d) Ring opened reaction of Rhodamine B


O O
Grinding N-R
Grinding Grinding N-R

Heating
N N N O N
O

Crystal Amorphous
highly twisted more planarized Ring closed Ring opened
emission at 441 nm emission at 468 nm non fluorescent emission at 576 nm

Figure 4.29  (a) Molecular structure and fluorescence images of the original deep-­blue powder, bluish-­green powder after slight grinding and reddish
powder. (b) The stacking mode of the unit cell of M-­4-­B crystal. (c) Normalized fluorescence intensity of three states. (d) The mechanisms of tricolored
switching of M-­4-­B. Source: Reproduced with permission from Ref. [53]. Copyright 2016 Wiley-­VCH.
Figure 4.30  (a) Molecular structure and (b) photographs of single crystals and ground samples of 5df and 5fd under UV light. (c) and (d) Molecular packing
and interactions of 5df and 5fd. (e) Calculated dipole moments and fluorescent images before and after grinding. Source: Reproduced with permission from
Ref. [54]. Copyright 2016 American Chemical Society.
4.5  ­Room Temperature Phosphorescence Materials Based on Molecular Unitin 131

4.5 ­Room Temperature Phosphorescence Materials Based


on Molecular Uniting

Organic RTP materials have aroused increasing interests over decades on account of their promi-
nent applications in organic light-­emitting diode, in vivo phosphorescence biological imaging, sen-
sitive chemical sensor, and anticounterfeiting fields  [55]. However, concerning the inherent
transition forbidden between excited singlet state and triplet state, and the quenching effect of
oxygen or water to the triplet excitons, the realization of organic RTP is still challenging. Thanks to
the persistent effort and enthusiasm of scientists, fruitful achievements have been fulfilled with
efficient strategies such as special electronic coupling and heavy atom effect, which can increase
SOC, resulting in the promoted ISC process. Besides, H-­aggregation, self-­assembling, and π–π
stacking can also stabilize excited triplet states, favoring to persistent RTP [56].

4.5.1  Room Temperature Phosphorescence Materials with Aromatics


Aromatics were the common building blocks for organic RTP materials with the combination of
carbonyls and heavy atoms. Basically, carbonyls provide n electrons, and the heavy atom effect
can enhance SOC, promoting ISC process. Through systemic investigation of luminescent materi-
als, it was found that a variety of substituents at different positions of aromatic ring could tune
the emission colors from blue to red with different molecular packing modes [57]. Nevertheless,
the limited phosphorescence emission efficiency and the lifetime still need more attention in
practical applications. Based on previous exploration, phosphorescence quantum yield is highly
related to ISC efficiency, and long phosphorescence lifetime mainly depends on the stability of
excited triplet states. Consequently, some moieties containing C, O, and P atoms were supposed
to support more n → π* transition so that ISC process was proposed [58]. Moreover, H-­aggregation
can be induced by different substituents forming stable excited triplet state (T*), and then long
RTP lifetimes in a magnitude of second can be obtained by DPhCzT (Figure 4.31) [59]. As dis-
played in single crystal of DPhCzT, the angle between transition dipoles and the interconnected
axis (θ) was larger than 54.7°, as defined to be H-­aggregation. Then different substituents resulted
in phosphorescence emission color varying from green (515  nm) to red (644  nm). Similarly, a
series of CN-­substituted phenylcarbazole isomers with one CN substituent in different position or
different numbers of CN groups were anticipated to show ultralong RTP lifetimes
(Figure  4.32a)  [60]. Taking compound mCNPhCz as a prototype, two kinds of single crystals
were picked out to show discriminate RTP performance. One (mCNPhCz) emitted bright phos-
phorescence with lifetime of 0.81 seconds, while 41 ms for the other one (mCNPhCz′). It is mainly
related to the H-­aggregation mode of mCNPhCz, and J-­aggregation dominant one for mCNPhCz′
(Figure 4.32b–d). As a result, the distinguished performance of the phenylcarbazole derivatives
manifested the superiority of H-­aggregation in ultralong RTP materials. Subsequently, a hybrid
strategy by the combination of intramolecular n → π* transition and intermolecular n → π* tran-
sition can be exploited by compound BCz-­BP (Figure  4.33a, b)  [61]. The anticipated largely
enhanced ISC process was observed in coupled dimers through time-­dependent density func-
tional theory calculation, which promoted bright long-­lived RTP emission. Inspired by this strat-
egy, Chi et al. recently reported a novel ultralong RTP compound o-­Cz with the carbazole moiety
as an electron donor (D)and carbonyl group as an electron acceptor (A), which inhibited intramo-
lecular charge transfer, not only through bond but also through space for the folded D–A struc-
ture (Figure 4.33d–f) [62]. Surprisingly, the synergy of intra-­or inter-­molecular interactions and
132 4  The Expansion of AIE Thought: From Single Molecule to Molecular Uniting

(a) Phosphoresence (b) 24FPB

τ = 0.92 s
2FPB F
Φ = 0.98 %
HO B OH 3.506 Å α = 60.5°

F
τ = 0.36 s θ θ
23FPB F
Φ = 0.09 % E**
HO B OH E
E*

F
J-aggregate H-aggregate
τ = 2.50 s G
24FPB F
0° 54.7° 90°
Φ = 2.42 %
HO B OH
S1ʺ
S1
S1ʹ
T1 T1
F
T*1
F
τ = 0.49 s
234FPB Exc Fluo Phos Ultralong
F Exc Fluo
Φ = 0.23 % phos
HO B OH S0 S0
Monomer H-aggregation

Figure 4.31  (a) The steady-­state photoluminescence and phosphorescence spectra of DEOPh, DECzT,
DPhCzT, CzDClT, and DCzPhP. (b) H-­aggregation in single crystal and schematic energy diagram of
H-­aggregation or J-­aggregation of DPhCzT. (c) Proposed energy transfer processes for fluorescence,
phosphorescence, and ultralong phosphorescence emission. Source: Reproduced with permission from
Ref. [58]. Copyright 2015 Nature publishing group.

multiple charge transfer channels sharply increased the ISC probability, compared to BCz-­BP, a
longer RTP lifetime of 0.84 seconds and a higher quantum efficiency of 16.6% were obtained.
As shown in Figure  4.34, three polymorphs of compound CzS-­CN achieved wide range RTP
lifetime (from 266 to 32 ms) with PLQY (from 22.6 to 6.9%), due to the distinguished molecular
packing modes and molecular configurations [63]. Molecules in crystal A and B exhibited more
twisted configuration with multiple intermolecular interactions (π–π, C–H⋯π, and C–H⋯N),
while they presented more planar configurations in crystal C. With respect to the influence of
molecular packing in RTP performance, three compounds were designed with similar photophysi-
cal properties in solution but different performance especially phosphorescence emission in solid
state. The only difference was the substituents to the carbazole moiety, which varied from phenyl
to p-­tolyl, then to 4-­methoxyphenyl in compound CPM, CMPM, and CMOPM [64]. Such mar-
ginal difference in molecular structure brought a decreased tendency in intermolecular interaction
of coupled molecules in crystal. As shown in Figure 4.35a, deeply investigation in crystal analysis
presented that the compact face-­to-­face packing mode favored the high PLQY and long lifetime
RTP of CPM (τ = 748 ms, ΦPL = 3.17%), better than those of CMPM (τ = 340 ms, ΦPL = 1.98%),
and CMOPM (τ = 114 ms, ΦPL = 0.51%). The excellent RTP performance can be ascribed to the
convenient orbital interactions in compact face-­to-­face packing (Figure  4.35b–e), which was
4.5  ­Room Temperature Phosphorescence Materials Based on Molecular Unitin 133

(a)
CN (b)
τ = 0.92 seconds
Φ = 1.8%
α = 38.99°
N

pCNPhCz
CN
τ = 0.81 seconds
PL intensity (a.u.)

Φ = 3.7%
N

mCNPhCz mCNPhCz
τ = 0.65 seconds
CN Φ = 2.9%
N (c)
α = 48.36°
oCNPhCz
NC CN τ = 0.48 seconds
Φ = 8.6%

DCNPhCz

350 450 550 650 750 mCNPhCz′


Wavelength (nm)

(d)

(e)

Figure 4.32  (a) Steady-­state PL and OURTP spectra with OURTP lifetime and PLQY of pCNPhCz, mCNPhCz,
oCNPhCz, and DCNPhCz. Molecular packing in (b) mCNPhCz crystal form and (c) mCNPhCz′ crystal form.
H-­aggregation structures in (d) mCNPhCz crystal form and (e) mCNPhCz′ crystal form. Source: Reproduced
with permission from Ref. [60]. Copyright 2019 The Royal Society of Chemistry.

further confirmed by theoretical calculation of ISC process from the lowest excited single state (S1)
to excited triplet states (Tn). Taking CPM for example, all kinds of dimers possessed more channels
for ISC process with smaller energy gaps between S1 and Tn states (Figure 4.35f). Also, the increased
overlap in face-­to-­face packing from CMOPM to CMPM and then to CPM can favor to
­persistent RTP.
134 4  The Expansion of AIE Thought: From Single Molecule to Molecular Uniting

(a) O (b) (d)


n O o-Cz UV
O

Br N
π Vis
N

IC
SCT NIR

T
k*ST
BCzBP S*1

T*n (e)
(c) T*1
k*r

S
S*0

2.983 (Å) Hybrid ISC:


3.524 (Å)
nπ* ππ*
ππ* ππ*

(f)

S1
T1

ISC
H-trap T1*
0.17 eV
0.35 eV
RISC
NIR

S0

Figure 4.33  (a) Molecular structure and (b) energy level diagram for electronic coupling of BCzBP.
(c) Intermolecular electronic coupling in single crystal of BCzBP. Source: Reproduced with permission from
Ref. [61]. Copyright 2016 Wiley-­VCH. (d) Chemical structure of o-­Cz. (e) Distribution of intramolecular
interaction regions in o-­Cz. (f) Proposed mechanism of two-­photon induced ultralong phosphorescence in
o-­Cz. Source: Reproduced with permission from Ref. [62]. Copyright 2019 The Royal Society of Chemistry.

Furthermore, for 10-­phenyl-­10H-­phenothiazine-­5,5-­dioxide-­based derivatives (CS-­CH3O,


CS-­CH3, CS-­H, CS-­Br, CS-­Cl, and CS-­F), substituents are adjusted from methoxy, methyl to
hydrogen, and then to bromine, chlorine, fluorine units with increased electron-­withdrawing abili-
ties gradually  [65]. Accordingly, the RTP lifetime increased from CS-­CH3O (88  ms), CS-­CH3
(96 ms) to CS-­H (188 ms), then to CS-­Br (268 ms), CS-­Cl (256 ms), and CS-­F (410 ms) (Figure 4.36a).
Apparently, as illustrated in Figure 4.36b, the combination of electron-­withdrawing groups induced
strong π–π interactions, which were beneficial to stabilize excited triplet state with ultralong
RTP. The electron-­donating groups of methoxy and methyl ones would increase π-­electronic
(a)
Crystal (A) Crystal (B) Crystal (C)

ФPL = 22.6% ФPL = 17.8% ФPL = 6.9%


S N CN
αAIE = 10.8 αAIE = 8.5 αAIE = 3.3
CzS-CN Persistent RTP Short RTP Short RTP
τP = 226 ms τP = 41 ms τP = 32 ms

UV on UV off

0.1 seconds 0.5 seconds 1.0 seconds 1.5 seconds

0.1 seconds 0.5 seconds 1.0 seconds 1.5 seconds

0.1 seconds 0.5 seconds 1.0 seconds 1.5 seconds

(b)

π–π 3.55 Å
C–H···π 2.57 Å
C–H···π 2.76–3.49 Å
C–H···N 3.01–3.90 Å
C–H···N 3.72 Å

π–π 5.18 Å
C–H···π 3.54 Å

π–π 4.59 Å C–H···π 2.95–3.11 Å


C–H···π 3.70 Å C–H···N 3.45–3.92 Å

π–π 3.84 Å
C–H···π 3.03–3.52 Å
C–H···N 3.54 Å
π–π 3.95 Å
C–H···N 2.84–3.95 Å

Figure 4.34  (a) Molecular structure and the related PL data, photographs of CzS-­CN. (b) Molecular packing
and intermolecular interactions in three crystal forms of CzS-­CN. Source: Reproduced with permission from
Ref. [63]. Copyright 2017 The Royal Society of Chemistry.
136 4  The Expansion of AIE Thought: From Single Molecule to Molecular Uniting

(a) O O O
(b) Dimer 1
5.53 Å
N N N 3.45 Å

O
CPM CMPM CMPM Dimer 2 Dimer 3 Dimer 4

7.52 Å 8.86 Å 9.06 Å

(c) Dimer 1 Dimer 2


τ = 748 ms, ΦPL = 3.17% τ = 340 ms, ΦPL = 1.98% τ = 114 ms, ΦPL = 0.51%
10.17 Å 9.09 Å
Intermolecular interaction Dimer 3 Dimer 4
Strong Weak
6.17 Å 7.04 Å

In crystal
Strong RTP Weak RTP (d) 6.02 Å 3.32 Å

(f) Dimer 1

Dimer 2 Dimer 3 Dimer 4


Energy level (eV)

4.0
8.87 Å 6.27 Å 7.54 Å

(e)
Dimer 1
5.57 Å
3.5 0.67 0.60 0.60 3.49 Å
0.49
0.42
Dimer 2 Dimer 3 Dimer 4

3.0
S1 Tn S1 Tn S1 Tn S1 Tn S1 Tn 7.56 Å 8.92 Å 9.10 Å

Monomer Dimer1 Dimer2 Dimer3 Dimer4

Figure 4.35  (a) Molecular structure and the carton modes for their packing style in crystal of CPM, CMPM,
and CMOPM. (b)–(e) Intermolecular interactions of dimers in CPM, CMPM, CMPM, and CPM-­A crystals. (f)
Energy level diagram of monomer and dimers in CPM crystal. Source: Reproduced with permission from
Ref. [64]. Copyright 2017 Wiley-­VCH.

density of phenothiazine, impeding π–π interactions for strong electrostatic repulsion between
adjacent π systems (Figure 4.36c). The speculation was well verified through ESP calculations in
Figure  4.36d, and the π–π distances in single-­crystal analysis. In the meanwhile, the dihedral
angles between involved phenyl rings were adjusted from 30.03° to 0°, which also intervened π–π
interactions. Thus, the synergy effect of short π–π distance and face-­to-­face packing modes contrib-
uted to the efficient π–π coupling, leading to persistent RTP with long lifetime and high PLQY of
CS-­F. The relationship between molecular structures (with different electronic structures), molec-
ular packing modes (different π–π interactions), and RTP performance can provide a guidance for
rational molecular design. Lately, luminogen CS-­2COOCH3 clearly revealed that a unique π–π
stacking mode prolonged the RTP lifetime [66]. With the aim to minimize the electronic effect of
substituents on the whole molecular backbone, a series of alkyl chains were integrated into pheno-
thiazine 5,5-­dioxide moiety (Figure 4.37a) [67]. They exhibited a particular odd-­even effect in RTP
performance, in agreement with the trend of melting points in alkanes with different lengths.
Luminogens with odd alkyl substituents (245 ms of CS-­CH3, 327 ms of CS-­C3H7, and 205 ms of
CS-­C5H11) presented longer RTP lifetimes than those with even ones (92 ms of CS-­C2H5, 181 ms
of CS-­C4H9, no exact lifetime of CS-­C6H13) (Figure 4.37b–d). Since odd alkyl substituents brought
about short distance or large overlap of π–π interactions, compared to even ones, the compact
molecular packing with strong π–π interactions should be responsible for the longer RTP lifetime
of CS derivatives (Figure 4.37e).
4.5  ­Room Temperature Phosphorescence Materials Based on Molecular Unitin 137

Figure 4.36  (a) Molecular structure and room temperature phosphorescence (RTP) behavior of CS-­CH3O,
CS-­CH3, CS-­H, CS-­Br, CS-­Cl, and CS-­F. (b) Proposed mechanism for persistent RTP. (c) Single-­crystal structures
and intermolecular interactions and (d) depiction of the electrostatic potential model of CS derivatives.
Source: Reproduced with permission from Ref. [65]. Copyright 2018 Nature publishing group.

Apart from the effect of strong π–π interactions to RTP property, other multiple intermolecular
interactions can also play the key role to excellent RTP properties. 9,9-­Dimethylxanthene was inte-
grated into organic RTP materials to promote intramolecular n → π* transitions [68]. Then the com-
monly used carbonyl group and chlorine atom were introduced to support ISC transitions and two
types of substituents were selected. One type is rigid phenyl ring (Ph and PhCl) in XCO-­Ph and
XCO-­PhCl, the other one is flexible alkanes (tBu and 1-­chloro-­2-­methylpropan-­2-­yl) in
XCO-­tBu and XCO-­PiCl (Figure  4.38a). These four compounds with the 9,9-­dimethylxanthene
core (XCO) showed RTP lifetimes ranging from 52 to 601 ms in crystal, regardless of their similar
photophysical properties in solution, which should be greatly related to molecular packing in crys-
tal. In Figure 4.38a, four compounds possessed similar interlaced molecular arrangement and an
antiparallel molecular set can be observed. However, the configuration of xanthene core varied
138 4  The Expansion of AIE Thought: From Single Molecule to Molecular Uniting

O O O O
(a) S
Removing “Ar”
(c)
S
N
Aromatic
core Ar
To simplify the structure N
CS–C2H5 CS–C3H7
n–1
CS-CnH2n+1 UV off
(b) 400 –200
C3H8
CH4 Odd-even effect 0.1 second
at RTP
C3H6
300 CS–C3H7
CS–CH3

Melting point (°C)


–160
RTP lifetime (ms)

CS–C4H9 CS–C5H10

Time
200
0.5 second

C4H10
100 CS–C2H5 C5H12 –120
1.0 second
C6H14
0
CS–C6H13
–80
1 2 3 4 5 6 1.5 second
Number of carbon atom (n)

(d) (e)
CS–CH3 τ = 245 ms
CS–C2H3 τ = 92 ms
3.70 Å
Crystal CS–C3H7 τ = 327 ms
@RT CS–C4H9 τ = 181 ms
Intensity (a.u.)

Distance
CS–C5H11 τ = 205 ms
CS–C6H13 /

3.93 Å
2.69 Å
43°

CS–C3H7 Overlap
0 500 1000 1500 2000 2500 3000 3500
Time (ms)

Figure 4.37  (a) Molecular design strategy of CS-­CnH2n+1. (b) Variation tendency of RTP lifetimes of
CS-­CnH2n+1. (c) The RTP behavior of CS-­C2H5 and CS-­C3H7 crystals. (d) Time-­resolved PL-­decay curves of
CS-­CnH2n+1 crystals. (e) Molecular packing in CS-­C3H7 crystal. Source: Reproduced with permission from
Ref. [67]. Copyright 2019 The Royal Society of Chemistry.

from XCO-­Ph to XCO-­PhCl, then to XCO-­tBu and XCO-­PiCl. When substituents varied from Ph
to tBu, the xanthene configuration behaved more planar, and the intermolecular distance between
two antiparallel molecules became shorter, resulting in the compact packing of XCO-­tBu. Similar
regulation can be obtained from PhCl to PiCl, deriving from that compounds with chlorine atoms
induced more C–H⋯Cl interactions. All in all, the alkyl group and chlorine atom induced more
intermolecular interaction sites to promote ISC process and the small size of alkyl group can main-
tain planar structure of core unit, promoting the efficient electron coupling (Figure 4.38b). Detailed
natural transition orbital analysis confirmed more n → π* transition in XCO-­tBu and XCO-­PiCl,
resulting in the longest RTP lifetime of XCO-­PiCl (Figure 4.38c). The detail investigation of sub-
stituent effect shed light on relationship between molecular packing and RTP lifetime. Apart from
9,9-­dimethylxanthene core unit showed promising development in constructing RTP materials, an
Figure 4.38  (a) Molecular structure, single-­molecular configuration, and antiparallel dimer packing of XCO-­derivatives. (b) Intermolecular interactions in
four crystals. (c) Natural transition orbital (NTO) analysis between singlet excited state (S1) and triplet excited state (T1) of XCO-­derivatives. Source:
Reproduced with permission from Ref. [68]. Copyright 2020 Wiley-­VCH.
140 4  The Expansion of AIE Thought: From Single Molecule to Molecular Uniting

Figure 4.39  (a) Molecular structure of TBBU. (b) and (c) Crystal structure of TBBU. (d) Digital images of the
TBBU doped film at various oxygen fractions. Source: Reproduced with permission from Ref. [69]. Copyright
2019 Wiley-­VCH.

unconventional triphenylamine derivative TBBU with a D-­π-­A structure accidently showed RTP
property (Figure 4.39a) [69]. The extraordinary intermolecular interactions between vicinal mole-
cules were indeed responsible for its RTP property, which could be confirmed by theoretical calcula-
tion and single-­crystal structure. In crystal, TBBU molecules with asymmetric noncoplanar
structure arranged in a head-­to-­tail packing mode, and the side-­by-­side stacking of adjacent mole-
cules cultivated abundant C–H⋯O, C–H⋯π interactions (Figure 4.39b, c). More importantly, the
doped polymer film with TBBU sample still existed RTP property, which can detect oxygen effi-
ciently (Figure  4.39d). Moreover, another compound TPA-­o-­3COOMe with antiparallel packing
mode of crystal A was confirmed to be RTP active, which was also benefited from multiple intermo-
lecular interactions. Similar strategy was applied in an ML-­active material BrFlu-­CBr, which
showed both mechano-­and photo-­stimulus phosphorescence with abundant intermolecular inter-
actions, unveiling the correlation of compact molecular packing and excellent RTP property.

4.5.2  Room Temperature Phosphorescence Materials with Simple or


Nonaromatic Structure
Apart from the persistent RTP property constructed by aromatic backbone with rigid and complex
structures, some materials with much simpler structures also exhibited fabulous RTP performance
by special aggregated structures. Recently, a series of boron-­containing phosphors were discovered
with long RTP lifetime (Figure 4.40a, b) [70]. Among these, (4-­methoxyphenyl)boronic acid (PBA-­
MeO) exhibited longest-­lived RTP with a lifetime of 2.24  seconds. Different from interactions
Time (second)
(a) PBA-MeO 0 2 4 6 8 10 (c)
Name Hydrogen bond Sketch of
(dn/g/cm3) length (Å) hydrogen bond

Ultralong phos. PBA-OH


(b) Fluorescence
(1.466)
OH

483 nm PBA-OH
τ = 0.71 second
B(OH)2
PBA-MeO
OMe (1.367)

483 nm PBA-MeO
τ = 2.24 second
PL intensity (a.u.)

B(OH)2
OEt
PBA-EtO
(1.307)
506 nm PBA-EtO
B(OH)2 τ = 1.11 second
OPr
PBA-BuO
493 nm PBA-PrO (1.249)

B(OH)2
τ = 0.13 second
OBu

PBA-PrO
492 nm PBA-BuO (1.201)
B(OH)2
τ = 1.28 second

300 350 400 450 500 550 600 650


Wavelength (nm)

Figure 4.40  (a) Photographs of PBA-­MeO crystal before and after irradiation. (b) Molecular structure and fluorescence, phosphorescence spectra of
PBA-­OH, PBA-­MeO, PBA-­EtO, PBA-­PrO, and PBA-­BuO crystals. (c) Hydrogen bonds and molecular packing in PBA-­derivatives crystals. Source: Reproduced with
permission from Ref. [70]. Copyright 2017 The Royal Society of Chemistry.
142 4  The Expansion of AIE Thought: From Single Molecule to Molecular Uniting

(a) (c)
HOOC CN CAA

UV off
b

UV on a c d = 3.0225 Å
(b)
do·o = 2.6337(16) Å 3.2674(25) Å 3.7368(17) Å

C
H 3.6394(16) Å
N 4.0589(28) Å
O Coupled A Coupled B Coupled C

Figure 4.41  (a) Molecular structure and photographs of CAA before and after UV irradiation. (b) Three
kinds of intermolecular coupling states in CAA crystal. (c) Molecular packing in layer-­by-­layer mode of CAA
crystal. Source: Reproduced with permission from Ref. [72]. Copyright 2018 The Royal Society of Chemistry.

between π planes, these arylboronic acids with plenty of oxygen atoms acting as weak electron
acceptors, which can supply numerous hydrogen bonds between adjacent molecules. The synergy
of suitable size of substituent groups and numerous hydrogen bonds in a regular form of molecular
arrangement, suppressed the nonradiative process largely, which was beneficial to RTP properties
(Figure 4.40c). Similarly, some arylboronic esters also behaved well in RTP properties, which could
improve SOC with molecular distortion in their T1 states [71]. Moreover, a commercialized com-
pound cyanoacetic acid (CAA) with no aromatic rings was exposed to be RTP active (Figure 4.41a),
which is the first example of a nonaromatic pure organic small molecule with persistent RTP prop-
erty  [72]. The gratifying phenomenon pushed us to pay more attention to its conformation and
molecular packing mode. Interestingly, there were three different kinds of coupled molecules with
strong electronic interactions between carboxyl and carboxyl, cyano and cyano, carboxyl and cyano
of adjacent molecules, leading to efficient ISC channels (Figure 4.41b). Moreover, the existence of
numerous classical O–H⋯O hydrogen bonds over the whole CAA crystal, and molecules arranged
in a layer-­by-­layer packing mode (Figure 4.41c). The layered structure was so similar to large π pla-
nar, benefiting to suppress nonradiative transition. Considering the above statements, the abundant
hydrogen bonds played an indispensable role to RTP property of CAA crystal [73].

4.5.3  Room Temperature Phosphorescence Materials with Multiple Emission


Normally, organic RTP materials exhibit only one emission color, restricting their development in
polychromatic applications. Recently, dual phosphorescence emission at room temperature can be
captured in a single component. Tang et al. designed a series of chlorobenzoyldibenzothiophene
derivatives decorated with halogen atoms, exhibiting salient dual phosphorescence emission
colors in crystalline state (Figure  4.42a)  [74]. The chlorine substituted compound ClBDBT
­displayed pure white phosphorescence with commission Internationale de l’Éclairage (CIE)
of (0.33, 0.35), which was close to standard white emission (0.33, 0.33). There existed two excited
4.5  ­Room Temperature Phosphorescence Materials Based on Molecular Unitin 143

(a) O (b)
UV on UV off
BDBT Sn
H S IC
ISC Tn
O S1
FBDBT
F S T1
O Phos2
Ex fast
CIBDBT
Phos1
Cl S slow
O S0
BrBDBT
Dual phosphorescent emission
Br S
(d)
(c) Transition f
θ1 = 33.0° S1
(%) (10–6)
θ2 = 33.2°
0.88 eV T2 H L 22.3 (π,π*)
θ1 = 39.0°
θ1 = 40.9° H-2 L-20.7 (n,π*)
θ2 = 21.8° 0.27 eV T1 T2 1.08
θ2 = 22.2° H-3 L-26.6 (n,π*)

3.36 eV
π-t π-t
o-π: o-π:
Å 3.3 3 H-4 L 12.8 (π,π*)
3.59 .33
S-to-:

2.48 eV

2.75 eV
2Å Å H L 35.7 (π,π*)
T1 H-1 L 10.8 (π,π*) 0.37
BDBT CIBDBT BrBDBT S0 H-2 L 43.0 (n,π*)

Figure 4.42  (a) Molecular structure and photographs of BDBT, FBDBT, ClBDBT, and BrBDBT before and after
UV irradiation. (b) Strategy to obtain dual phosphorescence emission. (c) Molecular configuration and
intermolecular interactions of BDBT, ClBDBT, and BrBDBT crystal. (d) Energy level diagram and electronic
transition characters of ClBDBT. Source: Reproduced with permission from Ref. [74]. Copyright 2017 Nature
publishing group.

triplet states (T1 and T2) below the lowest excited singlet state (S1), which implied the probability
of two transition channels from excited triplet states to ground state (S0). The characteristics of T1
and T2 states were manifested by hybrid quantum mechanics and molecular mechanics (QM/MM)
method. Apparently, T2 state possessed dominant (n, π*) transition with large transition coeffi-
cients, while T1 state contained more (π, π*) transition, leading to a low-­energy emission, and a
high-­energy emission with antikasha’s rule (Figure 4.42b, d). Detailed analysis conveyed that the
exceptional double transition channels might largely depend on the dibenzothiophene subunit. In
Figure 4.42c, the S⋯S or π–π interactions between dibenzothiophene moieties helped to enhance
the IEC of different excited states, contributing to the efficient dual phosphorescence emission.
Even more, several excitation wavelength-­dependent phosphors with color-­tunable (from violet to
green) emission were emerged. Compound TMOT based on triazine moiety (Figure 4.43a) can
exhibit the multiple and persistent phosphorescence emission with the color changing from green
(505 nm) to sky-­blue (452 nm), when they were excited by different wavelengths ranging from 365
to 320 nm (Figure 4.43b) [75]. The inherent mechanism could be explained by crystal analysis and
DFT calculation. The sky-­blue phosphorescence process (452 nm) derived from isolated molecule,
which was confirmed by phosphorescence spectra with one emission peak and long lifetime in
dilute solution or PMMA film. However, in single crystal, each TMOT molecule was constrained
by six adjacent molecules through abundant strong C–H⋯N intermolecular interactions between
triazine unit and methoxy groups, arranging these seven molecules in the same plane (Figure 4.43d).
The bright green phosphorescence emission (505 nm) mainly related to the H-­aggregation of tria-
zine moieties with strong π–π interactions. In addition, compounds of 2-­chloro-­4,6-­dimethoxy-­1,3,
144 4  The Expansion of AIE Thought: From Single Molecule to Molecular Uniting

Figure 4.43  (a) Molecular structure of TMOT, DMOT, and CYAD. (b) Trajectory of color modulation, recorded
by the change in the excitation from 300 to 360 nm, in the CIE coordinate diagram. (c) Excitation spectra
obtained by monitoring the emission at 430 and 470 nm for DMOT and at 380 and 450 nm for CYAD,
respectively. (d) Intermolecular interactions and molecular packing mode of TMOT crystal. (e) Intermolecular
interactions in DMOT and CYAD crystals. Source: Reproduced with permission from Ref. [75]. Copyright
2019 Nature publishing group.

5-­triazine (DMOT) and 1,3,5-­triazinane-­2,4,6-­trione (CYAD) showed the similar phenomena for
the special molecular packing mode in aggregated state (Figure 4.43c, e).

4.5.4  Photoinduced Room Temperature Phosphorescence Materials


With the rapid development of RTP materials, more interesting phenomenon was discovered. As for
the abovementioned 10-­phenyl-­10H-­phenothiazine-­5,5-­dioxide-­based derivatives, when the sub-
stituent was trifluoromethyl unit, no RTP was observed in single crystal of the resultant compound
CS-­CF3. However, after five minutes irradiation under a hand-­held UV lamp, it exhibited the persis-
tent RTP with lifetime of 299  ms, termed as photoinduced RTP. The inherent mechanism was
ascribed to the increased π–π interactions by light irradiation, rather than the change of molecular
structures, which could be further confirmed by varied π–π distances before and after irradiation
(Figure 4.44b). It can get back to initial state after two hours without UV irradiation, favoring to the
application in confidential information protection. Similarly, compound TPA-­B constructed by tri-
phenylamine and borate groups exhibited appealing photoinduced RTP phenomenon [76]. Initially,
almost no RTP emission could be observed of TPA-­B crystal at initial state, while a bright yellow
phosphorescence emission can be observed by naked eyes after 20 minutes UV irradiation (lasting
for one second) (Figure 4.45a). Conventional measurements showed an increase in RTP lifetime
from 5.32 to 211.13 ms of TPA-­B crystal (Figure 4.45b). The crystal analysis illustrated a tiny change
4.5  ­Room Temperature Phosphorescence Materials Based on Molecular Unitin 145

Figure 4.44  (a) Molecular structure and phosphorescence spectra before and after five minutes irradiation
of CS-­CF. (b) Crystal structures of coupled CS-­CF3 before and after five minutes irradiation. (c) Double
security protection applications by three kinds of components of CS-­CF3, CS-­F, and (4-­methoxyphenyl)
(phenyl)methanone. Source: Reproduced with permission from Ref. [64]. Copyright 2018 Nature
publishing group.

(a) (c)
TPA-B Crystal (i) Crystal (p)
RTP intensity (a.u.)

Molecule
B
O O

UV-irradiation

θA-B: 71.86° θA-B: 72.18°

Wavelength (nm)

Photo-induced RTP
Dimer

(b)

UV-irradiation time
d: 3.344 Å d: 3.319 Å
d: 3.848 Å d: 3.835 Å
20 minutes
RTP intensity

b b

0 minutes a a
Packing

450 500 550 600 650 700


Wavelength (nm)

Figure 4.45  (a) Molecular structure and phosphorescence spectra before and after UV irradiation of
TPA-­B. (b) RTP spectra of TPA-­B crystals after UV irradiation from 0 to 20 minutes. (c) The changes of
molecular configuration, intermolecular interactions and molecular packing in TPA-­B crystal before and
after UV irradiation. Source: Reproduced with permission from Ref. [76]. Copyright 2019 Wiley-­VCH.
146 4  The Expansion of AIE Thought: From Single Molecule to Molecular Uniting

of molecular conformation derived from the torsion angle between two phenyl rings (A and B),
which further affected the C–H⋯π and C–H⋯O interaction among adjacent molecules
(Figure 4.45c). A series of molecular rotors (PCzT, BCzT, and FCzT) containing a triazine core, a
carbazole unit, and various alkoxy chains can also realize photoactivated persistent RTP with simi-
lar mechanism (Figure 4.46a) [77]. Taking BCzT (with butoxyl chain) as example, the lifetime of
phosphorescence emission increased from 1.8 to 1330  ms after eight minutes irradiation by UV
lamp with high power (Figure 4.46b). As anticipated, the rotors can rotate around the single bond

(a) (c)
R N R
R
N N
–OC3H7 PCzT
N
–OC4H9 BCzT
–OC5H11 FCzT

(b) (d)
103 BCzT 105
Before photo-activation
104
Intensity

103
Intensity (counts)

102
102 101
100
0.0 0.5 1.0 1.5 2.0 2.5 3.0
Time (ms)

101

0 minutes
5 minutes
100
0 2 4 6 8 10
(e) Time (seconds)

1.5
tes ho
urs
nu
mi
10
Photo-activation region

Deactivation region
2.0 hours
5 minutes

UV on

UV off ho
3.0
tes urs
inu
1m

Figure 4.46  (a) Molecular structure of PCzT, BCzT, and FCzT. (b) Lifetime decay profiles of the emission
band around 530 nm of BCzT crystals before and after photoactivation under ambient conditions.
(c) Molecular packing in BCzT. (d) Intermolecular interactions in BCz-­T and BCzT-­a crystal. (e) Demonstration
of multilevel anticounterfeiting using the MCzT, PCzT, BCzT, and FCzT crystals. Source: Reproduced with
permission from Ref. [77]. Copyright 2018 Wiley-­VCH.
 ­Reference 147

between carbazole and triazine group, which formed different molecular conformations before and
after UV irradiation, leading to decreased distance between molecules and increased intermolecular
interactions (Figure 4.46c, d). Thus, more rigid surrounding can be formed, and the corresponding
nonradiative transitions would be restricted. Thus, more stable excited triplet states could be estab-
lished, which was beneficial to prolong RTP lifetime. Moreover, photoactivated prolonged RTP can
be deactivated to initial state after two days storage under ambient atmosphere (Figure  4.46e).
According to the above materials, the photoinduced RTP properties were closely related to change-
able molecular conformations and/or molecular packing before and after UV irradiation.

4.6 ­Conclusion and Perspectives

Hence, the rapidly development of pure organic luminescent materials urges us to consider the
crucial factor to fabulous performance of PL, ML, MC, and RTP materials. Many efficient strategies
in optimizing molecular structures are taken into account, and the key role of molecular packing
is highlighted, which is mainly detected by the electronic and steric effect of organic conjugated
system and substituent effect. The planar conjugated structures can easily induce π–π interactions,
prompting the formation of H-­aggregation, excimer and so on. While twisted structures usually
constructed the rigid surroundings by multiple C–H⋯π interactions, which can restrict the molec-
ular motions efficiently. The electronic property of conjugated core can be tuned by various sub-
stituents, resulting in the changeable electrostatic interactions and molecular packing. In many
cases, the substituent effect can not only influence the SOC, but also induce multiple intermolecu-
lar interactions, favoring to the bright emission as RTP. Moreover, the key role of molecular pack-
ing to ML, MC, and RTP properties can also be suitable to various luminescent materials, such as
electrochromic and electroluminescence materials, photoactivatable AIEgens, and so on.
Besides, some external force, including mechanical stimulus, light irradiation, and heating, can
vary the molecular packing, resulting in the dynamic luminescence property. The detailed infor-
mation of molecular arrangements can be detected through single crystals, and the inherent mech-
anism of luminescence could be suggested by theoretical calculation. Based on the systemic
investigation of relationship between molecular structures, packing modes and intermolecular
interactions, and luminescence properties, the concept of MUSIC was proposed by concerning the
similarity in the construction of optoelectronic materials and the symphonic performance as real
“music,” which can guide the molecular design of optoelectronic materials from the viewpoint of
molecular uniting set but not only the structure of single molecule.

­References

1 Mei, J., Leung, N. L., Kwok, R. T., et al. (2015). Aggregation-­induced emission: together we shine,
United We Soar! Chem. Rev. 115 (21): 11718–11940.
2 Luo, J., Xie, Z., Lam, J. W. Y., et al. (2001). Aggregation-­induced emission of 1-­methyl-­1,2,3,4,5-­
pentaphenylsilole. Chem. Commun. (18): 1740–1741.
3 Zhao, Z., Zhang, H., Lam, J. W., et al. (2020). Aggregation-­induced emission: new vistas at
aggregate level. Angew. Chem. Int. Ed. 59 (25): 9888–9907.
4 Huang, J., Sun, N., Dong, Y., et al. (2013). Similar or totally different: the control of conjugation
degree through minor structural modifications, and deep-­blue aggregation-­induced emission
luminogens for non-­doped OLEDs. Adv. Funct. Mater. 23 (18): 2329–2337.
148 4  The Expansion of AIE Thought: From Single Molecule to Molecular Uniting

5 Yang, J., Huang, J., Sun, N., et al. (2015). Twist versus linkage mode: which one is better for the
construction of blue luminogens with AIE properties? Chem. Eur. J. 21 (18): 6862–6868.
6 Li, Q., Li, Z. (2017). The strong light-­emission materials in the aggregated state: what happens from
a single molecule to the collective group. Adv. Sci. 4 (7): 1600484.
7 Xie, Y., Li, Z. (2020). The development of mechanoluminescence from organic compounds:
breakthrough and deep insight. Mater. Chem. Front. 4 (2): 317–331.
8 Wang, C., Li, Z. (2017). Molecular conformation and packing: their critical roles in the emission
performance of mechanochromic fluorescence materials. Mater. Chem. Front. 1 (11): 2174–2194.
9 Yang, J., Fang, M., Li, Z. (2020). Stimulus-­responsive room temperature phosphorescence in purely
organic luminogens. InfoMat. 2 (5): 791–806.
10 Li, Q., Li, Z. (2020). Molecular packing: another key point for the performance of organic and
polymeric optoelectronic materials. Acc. Chem. Res. 53 (4): 962–973.
11 Hong, Y., Lam, J. W., Tang, B. Z. (2011). Aggregation-­induced emission. Chem. Soc. Rev. 40 (11):
5361–5388.
12 Zhu, Q., Zhang, Y., Nie, H., et al. (2015). Insight into the strong aggregation-­induced emission of
low-­conjugated racemic C6-­unsubstituted tetrahydropyrimidines through crystal-­structure-­
property relationship of polymorphs. Chem. Sci. 6 (8): 4690–4697.
13 Tu, Y., Liu, J., Zhang, H., et al. (2019). Restriction of access to the dark state: a new mechanistic
model for heteroatom-­containing AIE systems. Angew. Chem. Int. Ed. 58 (42): 14911–14914.
14 Li, Y.-­X., Yang, X.-­F., Miao, J.-­L., et al. (2016). Effects of substitution position on crystal packing,
polymorphism and crystallization-­induced emission of (9-­anthryl)vinylstyrylbenzene isomers.
CrystEngComm 18 (12): 2098–2104.
15 Duan, Y., Ma, H., Tian, H., et al. (2019). Construction of a luminogen exhibiting high contrast and
multicolored emission switching through combination of a bulky conjugation core and tolyl
groups. Chem. Asian J. 14 (6): 864–870.
16 Wang, F., Liu, D., Li, J., et al. (2018). Molecular engineering of host materials for high-­performance
phosphorescent OLEDs: zig-­zag conformation with 3D gridding packing mode facilitating charge
balance and quench suppression. Adv. Funct. Mater. 28 (39): 1803193.
17 Zhang, Y., Wang, J.-­H., Zheng, J., et al. (2015). A Br-­substituted phenanthroimidazole derivative
with aggregation induced emission from intermolecular halogen–hydrogen interactions. Chem.
Commun. 51 (29): 6350–6353.
18 Zhang, Z., Xu, B., Su, J., et al. (2011). Color-­tunable solid-­state emission of 2, 2′-­biindenyl-­based
fluorophores. Angew. Chem. Int. Ed. 50 (49): 11654–11657.
19 Mallet, C., Moussallem, C., Faurie, A., et al. (2015). Rational topological design for fluorescence
enhancement upon aggregation of distyrylfuran derivatives. Chem. Eur. J. 21 (21): 7944–7953.
20 Xie, Y., Li, Z. (2018). Triboluminescence: recalling interest and new aspects. Chem 4 (5): 943–971.
21 Xie, Y., Tu, J., Zhang, T., et al. (2017). Mechanoluminescence from pure hydrocarbon AIEgen.
Chem. Commun. 53 (82): 11330–11333.
22 Wang, C., Li, Z. (2019). Mechanoluminescence materials with the characteristic of aggregation-­
induced emission (AIE). In: Tang, Y., Tang, B. (eds) Principles and Applications of Aggregation-­
Induced Emission. Springer, Cham.
23 Xu, S., Liu, T., Mu, Y., et al. (2015). An organic molecule with asymmetric structure exhibiting
aggregation-­induced emission, delayed fluorescence, and mechanoluminescence. Angew. Chem.
Int. Ed. 54 (3): 874–878.
24 Wang, C., Xu, B., Li, M., et al. (2016). A stable tetraphenylethene derivative: aggregation-­induced
emission, different crystalline polymorphs, and totally different mechanoluminescence properties.
Mater. Horiz. 3 (3): 220–225.
 ­Reference 149

25 Xu, B., He, J., Mu, Y., et al. (2015). Very bright mechanoluminescence and remarkable
mechanochromism using a tetraphenylethene derivative with aggregation-­induced emission.
Chem. Sci. 6 (5): 3236–3241.
26 Wang, C., Yu, Y., Chai, Z., et al. (2019). Recyclable mechanoluminescent luminogen: different
polymorphs, different self-­assembly effects of the thiophene moiety and recovered molecular
packing via simple thermal-­treatment. Mater. Chem. Front. 3 (1): 32–38.
27 Xu, B., Li, W., He, J., et al. (2016). Achieving very bright mechanoluminescence from purely organic
luminophores with aggregation-­induced emission by crystal design. Chem. Sci. 7 (8): 5307–5312.
28 Liu, F., Tu, J., Wang, X., et al. (2018). Opposite mechanoluminescence behavior of two isomers
with different linkage positions. Chem. Commun. 54 (44): 5598–5601.
29 Sun, Q., Tang, L., Zhang, Z., et al. (2017). Bright NUV mechanofluorescence from a terpyridine-­
based pure organic crystal. Chem. Commun. 54 (1): 94–97.
30 Fang, M., Yang, J., Liao, Q., et al. (2017). Triphenylamine derivatives: different molecular packing
and the corresponding mechanoluminescent or mechanochromism property. J. Mater. Chem. C
5 (38): 9879–9885.
31 Tu, J., Fan, Y., Wang, J., et al. (2019). Halogen-­substituted triphenylamine derivatives with intense
mechanoluminescence properties. J. Mater. Chem. C 7 (39): 12256–12262.
32 Yu, Y., Wang, C., Wei, Y., et al. (2019). Halogen-­containing TPA-­based luminogens: different
molecular packing and different mechanoluminescence. Adv. Opt. Mater. 7 (18): 1900505.
33 Wang, J., Chai, Z., Wang, J., et al. (2019). Mechanoluminescence or room-­temperature
phosphorescence: molecular packing-­dependent emission response. Angew. Chem. Int. Ed.
131 (48): 17457–17462.
34 Yu, Y., Fan, Y., Wang, C., et al. (2019). Phenanthroimidazole derivatives with minor structural
differences: crystalline polymorphisms, different molecular packing, and totally different
mechanoluminescence. J. Mater. Chem. C 7 (44): 13759–13763.
35 Gong, Y., Zhang, P., Gu, Y., et al. (2018). The influence of molecular packing on the emissive
behavior of pyrene derivatives: mechanoluminescence and mechanochromism. Adv. Opt. Mater.
6 (16): 1800198.
36 Yang, J., Qin, J., Geng, P., et al. (2018). Molecular conformation-­dependent mechanoluminescence:
same mechanical stimulus but different emissive color over time. Angew. Chem. Int. Ed. 57 (43):
14174–14178.
37 Yang, J., Ren, Z., Xie, Z., et al. (2017). AIEgen with fluorescence-­phosphorescence dual
mechanoluminescence at room temperature. Angew. Chem. Int. Ed. 56 (3): 880–884.
38 Yang, J., Gao, X., Xie, Z., et al. (2017). Elucidating the excited state of mechanoluminescence in
organic luminogens with room-­temperature phosphorescence. Angew. Chem. Int. Ed. 56 (48):
15299–15303.
39 Mu, Y., Yang, Z., Chen, J., et al. (2018). Mechano-­induced persistent room-­temperature
phosphorescence from purely organic molecules. Chem. Sci. 9 (15): 3782–3787.
40 Wakchaure, V. C., Ranjeesh, K. C., Goudappagouda, et al. (2018). Mechano-­responsive room
temperature luminescence variations of boron conjugated pyrene in air. Chem. Commun. 54 (47):
6028–6031.
41 Li, J. A., Zhou, J., Mao, Z., et al. (2018). Transient and persistent room-­temperature
mechanoluminescence from a white-­light-­emitting aiegen with tricolor emission switching
triggered by light. Angew. Chem. Int. Ed. 57 (22): 6449–6453.
42 Wang, J., Wang, C., Gong, Y., et al. (2018). Bromine-­substituted fluorene: molecular structure,
Br-­Br interactions, room-­temperature phosphorescence, and tricolor triboluminescence. Angew.
Chem. Int. Ed. 130 (51): 17063–17068.
150 4  The Expansion of AIE Thought: From Single Molecule to Molecular Uniting

43 Wang, C., Yu, Y., Yuan, Y., et al. (2020). Heartbeat-­sensing mechanoluminescent device based on a
quantitative relationship between pressure and emissive intensity. Matter 2 (1): 181–193.
44 Dong, Y. Q., Lam, J. W., Tang, B. Z. (2015). Mechanochromic luminescence of aggregation-­induced
emission luminogens. J. Phys. Chem. Lett. 6 (17): 3429–3436.
45 Li, W., Huang, Q., Mao, Z., et al. (2020). Selective expression of chromophores in a single molecule:
soft organic crystals exhibiting full-­colour tunability and dynamic triplet-­exciton behaviours.
Angew. Chem. Int. Ed. 132 (9): 3768–3774.
46 Luo, X., Li, J., Li, C., et al. (2011). Reversible switching of the emission of diphenyldibenzofulvenes
by thermal and mechanical stimuli. Adv. Mater. 23 (29): 3261–3265.
47 Krishna, G. R., Kiran, M. S. R. N., Fraser, C. L., et al. (2013). The relationship of solid-­state
plasticity to mechanochromic luminescence in difluoroboron avobenzone polymorphs. Adv. Funct.
Mater. 23 (11): 1422–1430.
48 Zhang, G., Lu, J., Sabat, M., et al. (2010). Polymorphism and reversible mechanochromic
luminescence for solid-­state difluoroboron avobenzone. J. Am. Chem. Soc. 132 (7): 2160–2162.
49 Zhu, J.-­Y., Li, C.-­X., Chen, P.-­Z., et al. (2020). A polymorphic fluorescent material with strong solid
state emission and multi-­stimuli-­responsive properties. Mater. Chem. Front. 4 (1): 176–181.
50 Dong, Y., Xu, B., Zhang, J., et al. (2012). Piezochromic luminescence based on the molecular
aggregation of 9,10-­bis((E)-­2-­(pyrid-­2-­yl)vinyl)anthracene. Angew. Chem. Int. Ed. 51 (43):
10782–10785.
51 Gong, Y., He, S., Li, Y., et al. (2020) partially controlling molecular packing to achieve off–on
mechanochromism through ingenious molecular design. Adv. Opt. Mater. 8 (8): 1902036.
52 Nagura, K., Saito, S., Yusa, H., et al. (2013). Distinct responses to mechanical grinding and
hydrostatic pressure in luminescent chromism of tetrathiazolylthiophene. J. Am. Chem. Soc.
135 (28): 10322–10325.
53 Ma, Z., Wang, Z., Meng, X., et al. (2016). A mechanochromic single crystal: turning two color
changes into a tricolored switch. Angew. Chem. Int. Ed. 55 (2): 519–522.
54 Wu, J., Cheng, Y., Lan, J., et al. (2016). Molecular engineering of mechanochromic materials by
programmed C–H arylation: making a counterpoint in the chromism trend. J. Am. Chem. Soc.
138 (39): 12803–12812.
55 Li, D., Lu, F., Wang, J., et al. (2018). Amorphous metal-­free room-­temperature phosphorescent
small molecules with multicolor photoluminescence via a host-­guest and dual-­emission strategy.
J. Am. Chem. Soc. 140 (5): 1916–1923.
56 Forni, A., Lucenti, E., Botta, C., et al. (2018). Metal free room temperature phosphorescence from
molecular self-­interactions in the solid state. J. Mater. Chem. C 6 (17): 4603–4626.
57 Bolton, O., Lee, K., Kim, H.-­J., et al. (2011). Activating efficient phosphorescence from purely
organic materials by crystal design. Nat. Chem. 3 (3): 205–210.
58 Su, Y., Phua, S. Z. F., Li, Y., et al. (2018). Ultralong room temperature phosphorescence from
amorphous organic materials toward confidential information encryption and decryption. Sci. Adv.
4 (5): eaas9732.
59 An, Z., Zheng, C., Tao, Y., et al. (2015). Stabilizing triplet excited states for ultralong organic
phosphorescence. Nat. Mater. 14 (7): 685–690.
60 Yuan, J., Wang, S., Ji, Y., et al. (2019). Invoking ultralong room temperature phosphorescence of
purely organic compounds through H-­aggregation engineering. Mater. Horiz. 6 (6): 1259–1264.
61 Yang, Z., Mao, Z., Zhang, X., et al. (2016). intermolecular electronic coupling of organic units for
efficient persistent room-­temperature phosphorescence. Angew. Chem. Int. Ed. 55 (6): 2181–2185.
62 Mao, Z., Yang, Z., Xu, C., et al. (2019). Two-­photon-­excited ultralong organic room temperature
phosphorescence by dual-­channel triplet harvesting. Chem. Sci. 10 (31): 7352–7357.
 ­Reference 151

63 Yang, J., Ren, Z., Chen, B., et al. (2017). Three polymorphs of one luminogen: how the molecular
packing affects the RTP and AIE properties? J. Mater. Chem. C 5 (36): 9242–9246.
64 Xie, Y., Ge, Y., Peng, Q., et al. (2017). How the molecular packing affects the room temperature
phosphorescence in pure organic compounds: ingenious molecular design, detailed crystal
analysis, and rational theoretical calculations. Adv. Mater. 29 (17) 1606829.
65 Yang, J., Zhen, X., Wang, B., et al. (2018). The influence of the molecular packing on the room
temperature phosphorescence of purely organic luminogens. Nat. Commun. 9 (1): 840.
66 Wang, Y., Yang, J., Tian, Y., et al. (2020). Persistent organic room temperature phosphorescence:
what is the role of molecular dimers? Chem. Sci. 11 (3): 833–838.
67 Yang, J., Gao, H., Wang, Y., et al. (2019). The odd-­even effect of alkyl chain in organic room
temperature phosphorescence luminogens and the corresponding in vivo imaging. Mater. Chem.
Front. 3 (7): 1391–1397.
68 Liao, Q., Gao, Q., Wang, J., et al. (2020). 9,9-­Dimethylxanthene derivatives with room-­temperature
phosphorescence: substituent effects and emissive properties. Angew. Chem. Int. Ed. 59 (25):
9946–9951.
69 Zhou, Y., Qin, W., Du, C., et al. (2019). Long-­lived room-­temperature phosphorescence for visual
and quantitative detection of oxygen. Angew. Chem. Int. Ed. 58 (35): 12102–12106.
70 Chai, Z., Wang, C., Wang, J., et al. (2017). Abnormal room temperature phosphorescence of purely
organic boron-­containing compounds: the relationship between the emissive behaviorand the
molecular packing, and the potential related applications. Chem. Sci. 8 (12): 8336–8344.
71 Shoji, Y., Ikabata, Y., Wang, Q., et al. (2017). Unveiling a new aspect of simple arylboronic esters:
long-­lived room-­temperature phosphorescence from heavy-­atom-­free molecules. J. Am. Chem. Soc.
139 (7): 2728–2733.
72 Fang, M., Yang, J., Xiang, X., et al. (2018). Unexpected room-­temperature phosphorescence from a
non-­aromatic, low molecular weight, pure organic molecule through the intermolecular hydrogen
bond. Mater. Chem. Front. 2 (11): 2124–2129.
73 Li, Q., Tang, Y., Hu, W., et al. (2018). Fluorescence of nonaromatic organic systems and room
temperature phosphorescence of organic luminogens: the intrinsic principle and recent progress.
Small 14 (38): 1801560.
74 He, Z., Zhao, W., Lam, J. W. Y., et al. (2017). White light emission from a single organic molecule
with dual phosphorescence at room temperature. Nat. Commun. 8 (1): 416.
75 Gu, L., Shi, H., Bian, L., et al. (2019). Colour-­tunable ultra-­long organic phosphorescence of a
single-­component molecular crystal. Nat. Photonics 13 (6): 406–411.
76 Dang, Q., Hu, L., Wang, J., et al. (2019). Multiple luminescence responses towards mechanical
stimulus and photo-­induction: the key role of the stuck packing mode and tunable intermolecular
interactions. Chem. Eur. J. 25 (28): 7031–7037.
77 Gu, L., Shi, H., Gu, M., et al. (2018). Dynamic ultralong organic phosphorescence by
photoactivation. Angew. Chem. Int. Ed. 57 (28): 8425–8431.
153

Clusterization-­Triggered Emission
Haoke Zhang1 and Ben Zhong Tang1,2,3
1
 Department of Chemistry, Hong Kong Branch of Chinese National Engineering Research Centre for Tissue Restoration and Reconstruction,
Institute for Advanced Study, and Department of Chemical and Biological Engineering, The Hong Kong University of Science & Technology,
Clear Water Bay, Kowloon, Hong Kong, China
2
 State Key Laboratory of Luminescent Materials and Devices, Guangdong Provincial Key Laboratory of Luminescence from Molecular
Aggregates, Center for Aggregation-Induced Emission, South China University of Technology, Guangzhou, China
3
 Shenzhen Institute of Aggregate Science and Technology, School of Science and Engineering, The Chinese University of Hong Kong,
Shenzhen, Guangdong, China

5.1 ­Introduction

Light plays a crucial role in the process of human civilization. Especially, in modern society, lumi-
nescence has greatly changed the way people live and work. In terms of the display device applica-
tion, developing new materials with desirable luminescent performance always be the top priority.
Traditionally, these ideal molecules are organic dyes with fused aromatic rings, which benefit from
their high emission efficiency [1–4]. Meanwhile, the emission color covers from ultraviolet to near-­
infrared and the emission wavelength could be subtly tuned by the molecular conjugation length
or dipole moment [5–8]. However, there are some drawbacks in these materials, that is, trouble-
some synthesis, high economic cost, and environmental pollution. Moreover, these fused-­ring-­
based structures always are toxic and difficult to be disposed of by nature and human bodies [9].
Apart from that, the planar conformation is one of the representative features of traditional dyes
that show the better conjugation than twist ones and these planar structures always possess high
luminescent quantum yield in their single-­molecule state  [10, 11]. However, the luminescence
becomes weaker or even quenched in the aggregate state, which is disadvantageous for the real
application as most of the material work in the solid state  [12]. This phenomenon is termed
aggregation-­caused quenching (ACQ) documented for more than a century since the discovery of
the concentration-­quenching effect by Perrin et al. in the 1920s [13–15].
In opposite to the ACQ, Tang et al. reported that pentaphenylsilole was nonluminescent in solu-
tion, but induced emission was observed in its aggregate state and this phenomenon was termed as
aggregation-­induced emission (AIE) [16, 17]. After 20 years of AIE research, varied luminogens with
AIE feature (AIEgens) are developed with tunable colors from ultraviolet to near-­infrared and emis-
sion efficiency is as high as 100% [18–22]. Furthermore, numerous advanced applications based on
these innovative materials are explored from optics and electronics to energy and bioscience [23–31].
For example, highly emissive AIEgens are the best candidates for the fabrication of organic
­light-­emitting devices and there is a rosy prospect in its biological application. The physiological

Handbook of Aggregation-Induced Emission: Volume 1 Tutorial Lectures and Mechanism Studies, First Edition.
Edited by Youhong Tang and Ben Zhong Tang.
© 2022 John Wiley & Sons Ltd. Published 2022 by John Wiley & Sons Ltd.
154 5  Clusterization-­Triggered Emission

environment is full of aqueous media in which most of the organic aromatic molecules form aggre-
gates or particles [32]. For traditional ACQ dyes, the bio-­imaging experiments need to perform in low
concentrations due to the quenching effect of high concentrations. Beyond that, the washing process
is required to remove the dissociative ACQ dyes. However, this problem is easily overcome by utiliz-
ing AIEgens [33, 34]. The imaging signal can be enhanced by increasing the concentration of AIE
probes and the AIE technique endows the bioexperiments with a wash-­free advantage. AIEgens also
show other advantages such as good photostability and large Stokes shift with whom this material
has been widely used in life science, optical electronics, chemoprobes, bioprobes, etc. [35–47].
However, the drawbacks of high toxicity and economic cost still exist in aromatic AIEgens
though the problem of quenching is solved [48, 49]. Generally speaking, nonaromatic systems are
the best candidates to overcome the problems of toxicity, because most of the natural products, if
not all, are nonaromatic structures such as starch, cellulose, and proteins [50, 51]. Some of them
may include phenyl rings but are isolated and negligible through-­bond electron delocalization [52,
53]. In contrast to synthetic aromatic systems, natural products have lots of advantages. For exam-
ple, these materials are always abundant in nature, resulting in a low economic cost. Meanwhile,
their biocompatibility is better than the artificial ones, which are environmentally friendly.
According to the accepted photophysical theories, these natural materials are nonluminescent due
to the absence of through-­bond conjugation and that is the reason why the fused-­ring-­based
chromophores dominate the area of photophysics. Whereas, dating back in history, the mechano-
luminescence of sugar was documented by Francis Bacon in his landmark book of “Advancement
of Learning” in 1605 [54]. Recently, several research groups also infrequently reported the lumi-
nescent behaviors of protein, starch, and sodium alginate, which are all nonconjugated poly-
mers  [55–59]. However, the emission mechanism behind this unusual phenomenon is rarely
mentioned and some of them even believe that the nonconventional luminescence is from impuri-
ties but not the nonconjugated structure. Meanwhile, the luminescence from the natural product
is always weak and in the ultraviolet region. Then, these factors hinder the further development
and application of this special luminescence.
In 2007, Tang et al. reported that the nonaromatic poly[(maleic anhydride)-­alt-­(vinyl acetate)]
emitted blue light in its colloid suspension, but its pure solution was nonluminescent under UV
irradiation, which is the same behavior with typical AIEgens [60, 61]. They gradually discovered
that some reported systems such as siloxane [62–67], poly(amidoamine) (PAMAM) [68–71] and
polypeptides [72–74] showed the same performance with PMV. According to the research experi-
ence on AIE, they independently defined this phenomenon as clusterization-­triggered emission
(CTE) and this special chromophore is termed as clusteroluminogen, then the emission is defined
as clusteroluminescence [61, 75–79]. Recently, Donald A. Tomalia et al. also used the term nontra-
ditional intrinsic luminescence (NTIL) to label this phenomenon  [80–83]. The photophysical
behaviors of NTIL, in essence, are similar to CTE, but NTIL focuses more on performance.
However, the aggregation process and mechanistic picture are also emphasized in the concept of
CTE. With the development of CTE, clusteroluminogens extend from natural/synthetic polymers
to small molecules and metal clusters (Figure 5.1). The emission color could be subtly turned from
blue to red and the clusteroluminescence quantum yield is as high as 90%. By summarizing all the
CTE-­related works, several common features are identified in clusteroluminogens.
(i) Clusteroluminogens consist of nonconjugated molecular motifs, where n or π electron-­based
groups are separated by nonconjugated linkers [84, 85]; (ii) In the isolated states, only the electron
transition belonging to n or π electron-­based groups can be detected in the PL spectra. However, an
intrinsic luminescence appears at longer wavelengths upon the formation of clusters [61]; (iii) The
excitation spectra of clusteroluminogens appear red-­shifted in comparison to the absorption spec-
tra. In other words, these molecules could be excited by light out of the range of absorption
5.1 ­Introductio 155

Clusteroluminogens
Natural/Synthetic polymers Small molecules Metal clusters
O R
CH2OH n
O H2N S
O OH
O OH O O R Au n
n H S
O O O N
O OH H
H
N N n Si O
n O n n O
H H H etc.
etc. etc.

Isolated Crosslinked Clustered


Eδ+ Eδ– : dipole or transient dipole

Small Large Eδ+ Eδ– Eδ+ Eδ–


Sn / Tn
Sn / T n Eδ– Eδ+ Eδ– Eδ+
Sn / Tn
Sn / Tn Eδ+ Eδ– Eδ+ Eδ–

E1 > E2 > E3 > E4 Eδ– Eδ+ Eδ– Eδ+

S0
Through-space conjugation (TSC)
S0
TSC: n–σ* interaction,
S0 n–π* interaction,
Non-conjugated linkers π–π* interaction,
S0
n electrons: N. O. P. S.../π electrons: aromatic rings, multiple bonds... H–bond interaction.....

Figure 5.1  The structures of different categories of clusteroluminogens and schematic illustration of the
working mechanism of CTE. Source: Reprinted with permission from Zhang et al. [79]. Copyright 2019
Elsevier Ltd.

peaks [86]; (iv) Clusteroluminogens show excitation wavelength-­dependent luminescence prop-


erty, where excitation at longer-­wavelength results in a red-­shifted emission  [87–89]; (v)
Clusteroluminogens show size-­dependent emission properties. The conception of “size” is related
to different generations of dendrimers, molecular weights of the polymers, diameter of the nano-
particles, etc. Generally, the increase of size would shift PL spectra to the redder range, and the
emission intensity will also be enhanced  [90, 91]; (vi) Clusteroluminogens, especially polymer
systems, are capable to phosphoresce and some of them even show room-­temperature phosphores-
cence [92, 93].
In this chapter, the CTE effect will be revisited and discussed through the view of mechanism.
As summarized in our recently published review article, through-­space conjugation (TSC) is dis-
closed as the working mechanism of clusteroluminescence (Figure 5.1) [79]. The TSC could exist
in all types of electrons, that is, σ, n, π, and the n–σ*, n–π* and π–π* interactions are all capable to
produce a narrow bandgap. It should be stressed here that these interactions could be inter/intra-
molecular and the biggest difference between clusteroluminogens and other conjugated AIEgens
is the channel of electron delocalization, that is, through-­space or through-­bond conjugation. In
156 5  Clusterization-­Triggered Emission

the following sections, the clusteroluminogens are introduced from four categories: pure n-­electron
system, pure π-­electron system, hybrid (n, π)-­electrons system and the other systems (i.e., nano-
clusters). Different kind of structures are presented to demonstrate the important roles of n–σ*,
n–π*, and π–π* interactions. Especially in the pure n-­electron system, nitrogen and oxygen-­based
clusteroluminogens are individually summarized. It is expected that this chapter can draw a clear
mechanistic picture for the CTE effect and also provide a guideline to design clusteroluminogens
with high emission efficiency.

5.2 ­Pure n-­Electron Systems

In 1970, Halpern et al. reported the unexpected fluorescence in tertiary amine derivative 1 in which
a broad peak with maximum emission wavelength (λem) of 310 nm was observed at 250 nm excita-
tion [94]. They attributed this intrinsic fluorescence to a weak V ← N state, where N for the ground
state and V state is one electron in the n orbital and one in the σ* [95–97]. In 2009, Imae et al. dis-
covered the blue fluorescence in triethylamine (2) in the process of fluorescent PAMAM study [98].
In this work, they proposed that the blue emission of tertiary amine came from complex formation
between oxygen and nitrogen even if there was no oxygen in 2. Recently, we observed blue fluores-
cence in aliphatic primary amine 3, which proved that the tertiary amine was not an indispensable
structure for clusteroluminescence [99]. Hexamethylenediamine is completely nonemissive at low
concentration (i.e., 0.1 mM, 1 mM). However, its concentrated solution and powder fluoresce a
bright blue color. Further control experiments exclude the interferences of impurity and oxidation.
In these simple amine systems, the n–σ* interaction and hydrogen-­bond interaction are proposed
to play an important role in their clusteroluminescence (Figure 5.2a).
In 2007, Stiriba et al. investigated that the photophysical properties of linear polyethyleneimine
(LPEI) and their hyperbranched counterpart (PEI) [100]. Meanwhile, their methylated derivatives
LPEIMe and PEIMe were also synthesized and studied (Figure 5.2b). All these four polymers show
inherent blue fluorescence at the clustering state. Based on the experimental data, they concluded
that the creation of amine-­rich nanocluster and electron-­hole recombination processes generated
clusteroluminescence. In other words, strong N–N interaction is produced in clustered polymers
and a transient dipole is formed at excited state, which narrows the bandgap between valence and
conduction bands. Meanwhile, from polyethylenimine to methylated polyethylenimine, a
bathochromic shift of λem is observed which is attributed to the difference of polymer polarities
and the polymer-­solvent interaction, In 2008, Imae et  al. reported the luminescent behavior of
NH2-­terminated third-­generation poly(propylene imine) dendrimers (PPI) (Figure 5.2c) [101, 102].
Similar to linear and hyperbranched polyethyleneimine, PPI dendrimer also shows intrinsic blue
fluorescence. Meanwhile, excitation wavelength-­dependent emission (EWDM) is observed that
λem at 400 and 450 nm are recorded with 340 and 370 nm excitation, respectively. They also noticed
that the emission intensity of PPI dendrimer was gradually increased when the media pH decreased
from 10.6 to 7.5 and 3.8. The authors stated that at acidic conditions the protonation of tertiary
amine groups filled the whole dendritic interior with cations, and the strong charge–charge repul-
sion made the structure of PPI dendrimer more rigid. This phenomenon further demonstrates that
structural rigidification is one of the key factors for CTE.
Since the clusteroluminescence is observed in the nitrogen-­based nonaromatic polymers and
small molecules, it is easy to come up with the question that whether the oxygen-­based nonaro-
matic systems are with the same property or not because of the rich n electron in oxygen. In 2018,
Yuan et  al. systematically investigated the photophysical performances of polyethylene glycol
5.2 Pure n-Electron Systems 157

(a)
NH2
N H2N
N

Quinuclidine

1 2 3

(b)
H H
N N N N
H2N N NH2 N N N
H n n

LPEI (λem = 456 nm) LPEIMe (λem = 466 nm)

H2N H2N NH2 N N N

NH N N N

N N NH2 N N N
H2N N N N N N
H
H
H2N N N N N N
N N NH2 N N N

NH N N N

H2N H2N NH2 N N N

PEI (λem = 422 nm) PEIMe (λem = 462 nm)

(c) H2N NH2 H2N


H2N
H2N
NH2
NH2
H2N N N H2N
N
H2N N
NH2
N N
H2N N
NH2
N
H2N
N N N
N
N
NH2
N
H2N N NH2
N N
H2N N
N
N N NH2
N
PPI N
H2N
N NH2
N
H2N N
N NH2
N
NH2 N NH2
H2N N
H2N NH2
NH2
NH2 H2N NH2

Figure 5.2  (a) Chemical structures of quinuclidine (1), trimethylamine (2) and hexamethylenediamine (3).
Refs. [94, 98, 99]. (b) Chemical structures of linear (LPEI) and hyperbranched polyethyleneimine (PEI) and
their methylated derivatives LPEIMe and PEIMe. Ref. [100]. (c) Chemical structures of polypropyleneimine
(PPI). Ref. [102].
158 5  Clusterization-­Triggered Emission

(PEG), poly(ethylene oxide) (PEO)–poly(propylene oxide) (PPO)–PEO triblock copolymer (F127),


and xylitol (Figure  5.3a)  [103]. As expected, all of these compounds show blue emission at the
concentrated solution and solid state whose λem locates around 400 nm. PEG exhibits the highest
solid-­state luminescence quantum yield (ΦL) of 8.1% among them. Meanwhile, cryogenic-­
persistent phosphorescence is observed in their solid state and even room temperature phospho-
rescence is detected in xylitol’s crystals. The above results suggest that oxygen-­based nonaromatic
polymers and monomers are also capable to produce clusteroluminescence. Seeing from the struc-
tures of xylitol, the oxygen and hydroxy group are believed to play an important role in the CTE
effect. If that is the case, the ice should also be emissive due to the existence of amounts of hydroxy
groups. According to the previous research experiences, the water must be nonluminescent as it is
the commonly used solvent for UV and PL measurements. However, in 1999, Lobyshev et  al.
reported the intrinsic luminescence of water and its λem is recorded at 350 and 410 nm under the
260 and 310  nm excitation, respectively, but the accuracy of this data is still in debate  [104].
Recently, Tang and Yuan et al. observed the blue fluorescence and green phosphorescence from
frozen water [79, 105]. As shown in Figure 5.3b, when a deionized water-­filled NMR was put into
liquid nitrogen, the ice was formed on the outer wall of the tube above the liquid nitrogen level,
and then a weak blue emission was observed under 365 nm irradiation. Meanwhile, green phos-
phorescence was also captured after stopping the UV irradiation and the phosphorescence was
disappeared at 2.0 seconds (Figure 5.3c). A reliable PL measurement is needed to record the emis-
sion spectra in the future.
To clarify the working mechanism of clusteroluminescence in oxygen clusters, Yuan et al. have
analyzed the packing structures of xylitol single crystals and numerous inter/intramolecular O⋯O
interactions were observed with short distance (2.7–2.8 Å). Oxygen atoms have a van der Waals
radius of 1.52 Å, so a relatively strong O⋯O interaction is generated once the O⋯O distance is
shorter than 3.04 Å. All these data suggested that the inter/intramolecular O⋯O interactions in
oxygen clusters produced partial electronic channels and narrowed the bandgap. Apart from the
O⋯O interactions, recently Wang’s group has summarized that the involved hydrogen bonds in
these systems not only perform an electrostatic interaction but they also induce electron density
delocalization [106]. As shown in Figure 5.3d, first-­principles computer modeling of ring-­shaped
water clusters suggests that increasing cluster size leads to increased electron delocalization and a
characteristic orbital shift in occupied orbitals. It seems that the hydrogen-­bond interaction is
another factor that caused the CTE effect.
In the periodic table of the elements, fluorine locates after oxygen and shows three lone pair
electrons. Theoretically, F–F interaction could also induce the inter/intramolecular electron delo-
calization and narrow the bandgap. However, no CTE effect is observed in fluorinated ethylene-­
propylene resins which is worthy of further exploration  [57]. Since the N-­ and O-­based pure
n-­electron systems have been revealed with clusteroluminescence, Yan et al. recently investigated
the “hybrid” structure of oligomeric siloxanes with both oxygen and nitrogen elements (Figure 5.4).
As expected, strong blue luminescence is observed by 365 nm UV irradiation and oxygen–nitrogen
clusters are proposed as the chromophore of clusteroluminescence. Meantime, the EWDM effect
is also detected in oligomeric siloxanes that the λem is bathochromic shift for 50 nm when the exci-
tation wavelength increases from 318 to 418 nm [107].
In summary, pure n-­electron structures are capable of emitting visible light due to the formation
of electron channels by through-­space n-­electron conjugation. Based on the above discussions,
there are two possible mechanisms for inter/intramolecular n-­electron delocalization. (i) Strong
O–O and N–N interactions are formed in the oxygen and nitrogen clusters, respectively, then tran-
sient dipoles will be generated at the excited state and result in the n electrons overlap. (ii) Hydrogen
(a) O Quantum yield (d)
H n OH –8
(HOMO) (HOMO) (HOMO) (HOMO)
PEG 8.1%
MOs energy (eV)

–10

O O –12
H O OH
x y z
–14
F127 5.8% (HOMO-16)
(HOMO-11) (HOMO-14)
(HOMO-8)
–16
OH

HO OH
OH OH
Xylitol 1.5% (H2O)3 (H2O)4 (H2O)5 (H2O)6

(b) (c)

0.5 second 1.0 second 2.0 second

UV on UV off

Figure 5.3  (a) Chemical structures of polyethylene glycol (PEG), poly(ethylene oxide) (PEO)–poly(propylene oxide) (PPO)–PEO triblock copolymer (F127)
and xylitol. Source: Ref. [103]. Copyright 2018 WILEY-­VCH Verlag GmbH & Co. KGaA, Weinheim. (b and c) Photographs of frozen water taken under (b)
365 nm irradiation and (c) immediately after stopping UV irradiation with 0.5, 1.0, and 2.0 seconds time delay. Source: Reprinted with permission from
Zhang et al. [79]. Copyright 2019 Elsevier Ltd. (d) Energy level diagrams of various small water clusters. The red lines correspond to the energy levels of the
characteristic occupied. Ref. [106].
160 5  Clusterization-­Triggered Emission

O O
(CH2)3NH2 (CH2)3NH2
Si Si
O O
O O O O
Si Si
(CH2)3NH2 (CH2)3NH2

Oligomeric siloxane
Figure 5.4  Chemical structures of oligomeric siloxane. Ref. [107].

bonds not only shorten the distance of O–O, N–N, or O–N but also induce the electron density
delocalization in the H-­bond region. Theoretical studies have proved that in the N⋯H–O struc-
ture, the electron of N will transfer from N orbital to O–H antibonding orbitals and the delocalized
electronic structures narrow the bandgap of clusteroluminogen. These two factors work synergisti-
cally to produce the unusual luminescence in pure n-­electron systems.

5.3 ­Pure π-­Electron Systems


In 1963, Lumry et  al. published a research article in Nature with the title of “Fluorescence of
Styrene Homopolymers and Copolymers” [108]. In this work, they discovered the unconventional
luminescence in polystyrene [PS]. They synthesized two PS polymers with isotactic and atactic
conformations, respectively. Atactic PS exhibited almost the same absorption and emission spectra
with benzene (Figure 5.5a) [85, 109]. However, a red-­shifted emission was observed in isotactic PS
compared to the atactic one. Meanwhile, the poly(styrene-­co-­methyl methacrylate) showed a simi-
lar λem with atactic PS and the λem cut down with the decrease of styrene fraction. All these results
suggest that intra/intermolecular interaction among neighboring isolated benzene rings plays an
important role in the red-­shifted emission. However, this work has not been widely noticed as the
unclear mechanistic picture of the abnormal fluorescence in isotactic PS.

(a) (b)

CMP
Atactic PS

n
Isotactic PS

Figure 5.5  Chemical structures of (a) atactic and isotactic polystyrene, Ref. [108] and (b) CMP consisting of
tetrasubstituted [2.2]paracyclophane. Ref. [110].
5.3 Pure π-­Electron Systems 161

In 2013, Morisaki et al. synthesized conjugated microporous polymers (CMPs) from the tetrasu-
bstituted [2,2]paracyclophane skeleton as a tetra-­functional building block (Figure 5.5b) [110]. The
CMPs consists of relatively uniform particles in organic solvents and shows a strong yellow-­green
fluorescence. [2,2]Paracyclophane is one of the archetypical systems with through-­space conjuga-
tion. However, they do not clarify whether the strong yellow-­green shows a direct relationship with
the through-­space conjugation of [2,2]paracyclophane or not. Recently, Tang et al reported that
the [2,2]paracyclophane-­based structure could induce a 38 nm bathochromic shift of λem in com-
parison to the corresponding phenyl rings derivative, which further proved that the through-­space
π-­electron conjugation is another pathway to extend the conjugation in addition to the traditional
through-­bond conjugation [111].
To clarify the important role of through-­space π-­electron conjugation in the CTE effect, Tang
et al. synthesized the small molecule of 1,1,2,2-­Tetraphenylethane (s-­TPE) which had no through-­
bond conjugation among four benzene rings (Figure 5.6a) [85, 112–116]. s-­TPE showed a 290 nm
emission peak in its THF solution, which was assigned to the benzene’s 0–0 electron transition.
The addition of water firstly increased the benzene emission which is attributed to the increase of

(a) (b)
3 400
290 nm
460 nm R
300
2 R
H
200
Ill0

Ill0

H H
H
1 R
s-TPE 100 R

Through-space conjugation
0 0
0 30 60 90
Water fraction (vol%)
(c)

Ph3C3H Ph5C5H Ph5C5H Ph7C7H

Figure 5.6  (a) Plots of the PL intensity change at 290 and 460 nm, respectively, versus the water fraction
of 1,1,2,2-­Tetraphenylethane (s-­TPE) in THF/water mixtures, inset: photograph of its powder taken under
365 nm UV light. Source: Ref. [85]. Copyright 2017, American Chemical Society (b) Schematic illustration
of the through-­space conjugation in s-­TPE. (c) Chemical structures of sym-­Triphenylcyclopropene
(Ph3C3H), sym-­pentaphenylcyclopentadiene (Ph5C5H), hexaphenylbenzene (Ph6C6), and sym-­
heptaphenylcycloheptatriene (Ph7C7H). Ref. [117].
162 5  Clusterization-­Triggered Emission

polarity. At fw = 70%, a new peak appeared around 460 nm along with a decrease of 290 nm emis-
sion intensity. Further increasing the fw made the longer-­wavelength emission stronger and
stronger. It has been proved by the dynamic light scattering measurement that aggregates of s-­TPE
were formed at fw ≥ 70%. Inset in Figure 5.6a shows a bright cyan emission of s-­TPE’s powder. The
460 nm emission in s-­TPE suggests a typical CTE effect. The crystal structure analysis on s-­TPE and
the relevant theoretical calculation reveals the existence of through-­space conjugation among four
benzene rings (Figure 5.6b).
At about the same time McGonigal et al. also reported the excited-­state aromatic interactions in
phenyl-­ring molecular rotors (Figure 5.6c) [117]. In this work, they used the steady-­state PL spec-
troscopy and theoretical calculation method to investigate the through-­space π-­electron conjuga-
tion in sym-­Triphenylcyclopropene (Ph3C3H), sym-­pentaphenylcyclopentadiene (Ph5C5H),
hexaphenylbenzene (Ph6C6), and sym-­heptaphenylcycloheptatriene (Ph7C7H). As a result,
through-­space conjugation is only discovered in Ph5C5H and Ph7C7H. They explained that the
existing SP3 carbon in the central carbocycles of Ph5C5H and Ph7C7H provided the excited-­state
molecular mobility to form the excited-­state through-­space aromatic dimers (ESTSAD). Meanwhile,
the crowded phenyl rings in Ph7C7H facilitate the formation of ESTSAD, then Ph7C7H shows
stronger through-­space conjugation than Ph5C5H. Although SP3 carbon is also involved in Ph3C3H,
the longer distance between two neighboring phenyl rings limits the formation of ESTSAD. Moreover,
the rigid core in Ph6C6 makes the excited-­state molecular motion difficult and shows no through-­
space conjugation in the excited state. At the same time, the temperature-­dependent PL experi-
ments for these four molecules demonstrates that the peak energy of Ph5C5H and Ph7C7H is
decreased with the increasing temperature, however, Ph3C3H and Ph6C6 show a constant bandgap
at different temperature. Theoretically, high thermal energy could promote the possibility of
ESTSAD formation and reduce the peak energy.
Above phenyl-­ring molecular rotors focus more on the intramolecular through-­space π-­electron
conjugation, so is it possible that the intermolecular through-­space π-­electron conjugation could
also generate the clusteroluminescence? Recently, Tang et al. synthesized two nonconjugated mol-
ecules 1,2-­diphenylethane (s-­DPE) and 1,2-­bis(2,4,5-­trimethylphenyl)ethane (s-­DPE-­TM) which
are the half structures of s-­TPE and its methylated derivative [118–120]. Theoretical calculation
and absorption spectra data suggest that the intramolecular through-­space π-­electron conjugation
is difficult to form in s-­DPE and s-­DPE-­TM (Figure 5.7a). However, similar to the tetraphenyle-
thane system, powder of s-­DPE and s-­DPE-­TM show dual-­emission signal, the short-­wavelength
one is from the isolated phenyl rings and the other one locates at the longer-­wavelength range
(Figure 5.7b). Experimental and calculation results reveal that s-­DPE and s-­DPE-­TM are capable of
performing light-­driven solid-­state molecular motion to form excited-­state through-­space com-
plexes (ESTSC). The radiative decay of ESTSC generates the unexpected longer-­wavelength emis-
sion classified as clusteroluminescence, meanwhile, the original packing structure can be recovered
from ESTSC after the removal of light irradiation (Figure 5.7c). However, why do the s-­DPE and
s-­DPE-­TM perform directional intermolecular motions and produce ESTSC upon light irradiation,
and what is the driving force? Partial atomic charges (δ) of s-­DPE were mapped based on Mulliken
population analysis. As shown in Figure  5.7d, atoms C1 and C1′ exhibited positive δ (δ+)
(δC1 = 0.026, δC1′ = 0.012) and the other carbons exhibited negative δ (δ−). It is proposed that the
opposite δ among these carbons will generate a transient dipole, similar to the London dispersion
force, and attract the neighboring monomers getting close to each other, then the δ of C1 and C1′
become more positive in ESTSC than the monomer. So, the intermolecular transient dipole
between two neighboring phenyl rings serves as the driving force for the generation of ESTSC.
(a) (b) (d)
1.0 355 nm 390 nm
1.0 δ– δ δ+
s-DPE s-DPE
H
H s-DPE-TM s-DPE-TM –0.557 0.100
l390 : l297 = 3
0.8
Normalized absorbance

H 0.8
Normalized PL intensity

H l355 : l285 = 26

S-DPE
0.6 0.6
H
H
268 nm 297 nm
0.4 H 0.4
H
280 nm
S-DPE-TM
0.2 0.2
285 nm
0.0 0.0
200 250 300 350 400 270 310 350 390 430 470 Ground-state monomer
(c) Wavelength (nm) Wavelength (nm)

Clusteroluminescence

UV light

Remove UV light

No/Weak intermolecular Excited-state through-space


interaction ESTSC
complexes (ESTSC)

Figure 5.7  (a) Absorption and (b) photoluminescence spectra of 1,2-­diphenylethane (s-­DPE) and 1,2-­bis(2,4,5-­trimethylphenyl)ethane (s-­DPE-­TM) in the
solid state. (c) Schematic illustration of the formation of excited-­state through-­space complexes (ESTSC). (d) Mulliken population analysis of s-­DPE monomer
and ESTSC, partial charge (δ) range: −0.557 to 0.100. Source: Ref. [119]. Copyright 2019, American Chemical Society.
164 5  Clusterization-­Triggered Emission

In short, different from the traditional theories on π-­bond conjugation, through-­space π-­electron
conjugation is another efficient way to construct luminescent materials with many advantages,
that is, good processability and low economic cost. Investigation on the model structures suggests
that both the inter and intramolecular through-­space π-­electron conjugation could be utilized to
improve the emission efficiency of clusteroluminescence. However, it needs to be emphasized
that [2,2]paracyclophane, serving as the typical through-­space system, is not the best conformation
for the through-­space π-­electron delocalization though it has a completely cofacial geometry. In
contrast, through-­space aromatic dimers dislocated with a certain angle will be beneficial to the
through-­space π-­electron conjugation, which is similar to the H-­and J-­aggregates [121]. The differ-
ence is that the H-­and J-­aggregates focus more on the ground-­state properties but through-­space
π-­electron conjugation, in most cases, exists in the excited state.

5.4 ­(n, π)-­Electrons Systems

Poly[(maleic anhydride)-­alt-­(vinyl acetate)] (PMV) is a famous copolymer that shows perfect


microspheres with uniform sizes, reactive surface, and desirable solubility  [122–124]. PMV is a
nonconjugated polymer but involves both n and π electrons. In 2007, Tang et  al. observed blue
intrinsic fluorescence from PMV under 365  nm UV irradiation and its emission intensity was
enhanced with the increase of concentration [60]. Then, a series of poly(maleic anhydride) deriva-
tives were disclosed with clusteroluminescence but the underlying mechanism was unclear. In
2017, Tang et  al. synthesized two oligomers poly[(maleic anhydride)-­alt-­(2,4,4-­trimethyl-­1-­
pentene)] (PMP) and oligo(maleic anhydride) (OMAh4) (Figure 5.8a) [125]. Experimentally, PMP
almost shows no fluorescence both in solution and solid state. In contrast, OMAh4 exhibits blue
and yellow colors in solution and powder under 365 nm UV irradiation, even if OMAh4 is just an
oligomer whose Mw is around 1000. PL spectrum of OMAh4 shows two emission peaks at 410 and
500 nm with a 365 nm excitation which suggests that OMAh4 has multiple chromophores in its
clustering state but only one emission peak is observed at 458 nm excitation. Theoretical calcula-
tion results suggest that the short distance between two neighboring succinic anhydride groups
rigidifies the oligomeric chain and increases the inter/intra-­chain interactions. Meanwhile, the
electrophilicity of carbon in C=O groups and nucleophilicity of oxygen in C=O and –O– groups
generate a strong electrostatic interaction inside the oligomer. As a result, the distance of
O=C⋯O=C is as short as 2.84 Å. It has been proved that this kind of electrostatic interaction could
induce through-­space electron delocalization. It is noteworthy that the n–π* interaction can also
induce a narrow band gap, where n electron comes from oxygen and π electron from the C=O
double bond [126]. In PMP, the 2,4,4-­trimethyl-­1-­pentene branches separate the neighboring suc-
cinic anhydride groups and prevent them from forming through-­space conjugation. The optimized
conformation of PMP shows a longer distance of O=C⋯O=C (>5 Å). That is the reason why the
PMP is nonluminescent both in solution and clustering state.
Since ester groups could be used for the construction of clusteroluminogens, amide units should
also work. In 2017, Tang et al. have synthesized two polymers poly(N-­isopropylacrylamide) (R1;
PNIPAM) and poly(N-­tert-­butylacrylamide) (R2; PNtBA) as shown in Figure 5.8b that side chains
of PNIPAM and PNtBA both are amide groups [84]. As a result, monomers of PNIPAM and PNtBA
are almost nonluminescent in the solid-­state but their polymeric powders fluoresce bright blue
emission under 365  nm UV irradiation, which performs the polymerization-­induced emission
(PIE) effect [127]. Recently, Yuan et al. have systematically investigated the CTE effect of various
polypeptides and reported that most of them are capable of emitting clusteroluminescence.
Moreover, they discovered the clusteroluminescence and RTP in some amino acids [128].
5.4 (n, π)-Electrons Systems 165

(a)
O O O O OMAh4 (365 nm)
O O O O O OMAh4 (458 nm)
PMP

O O O O O O O O
O O O O

PL intensity (a.u.)
PMP OMAh4

PMP OMAh4 PMP OMAh4

Daylight

UV light
400 500 600 700
Wavelength (nm)

(b)
H AIBN, dry THF R1 = –CH(CH3)2
N n
R
70 °C, 4 hours NH R2 = –C(CH3)3
O
O R

Monomer Polymer Polymer (zoom in)

R1

R2

Polymerization-induced emission (PIE)

Figure 5.8  (a) Chemical structures of poly[(maleic anhydride)-­alt-­(2,4,4-­trimethyl-­1-­pentene)] (PMP) and


oligo(maleic anhydride) (OMAh4), photographs of PMP and OMAh4 in THF and their powder taken under
daylight and 365 nm UV light irradiation, PL spectra of PMP (λex = 365 nm) and OMAh4 (λex = 365 and
458 nm) in THF with a concentration of 0.01 M. Ref. [125]. (b) Reaction scheme for the synthesis of
poly(N-­isopropylacrylamide) (R1; PNIPAM) and poly(N-­tert-­butylacrylamide) (R2; PNtBA), physical
appearance of the monomers (NIPAM and NtBA) and polymers (PNIPAM and PNtBA) under a 365 nm UV
lamp. Ref. [84].

Above-­mentioned (n, π)-­electrons systems are generally based on ester or amide units where the
n and π electrons are from heteroatoms and carbonyl, respectively. To broaden the (n, π)-­electrons
systems, Tang et al. synthesized compound 1 with multiple n (N, O, S) and π (C=N, S=O, benzene)
electrons [129, 130]. Packing structures in its crystal indicate that the two phenyl rings are stacked
in a nearly face-­to-­face manner with a small tilted angle (Figure 5.9a). The distance between two
benzenes is 3.330  Å, indicating a strong intramolecular through-­space π-­electron conjugation.
Meanwhile, oxygen and nitrogen atoms are arranged nearby and electron overlap is increased,
then the whole structure becomes conjugated and serves as the chromophore for the
166 5  Clusterization-­Triggered Emission

(a) (b)
O Clustering and
S rigidification of
N
O nonconventional
N R chromophores
N
1 Electron overlap
R
Through-space conjugation

(c) Br
O 2c O 3b 3p

O N O N
O O
N N
O 2cʹ O 3c

1.5 mm 3 mm 1.5 mm
2 Br 3

Figure 5.9  Schematic illustration of the unconventional luminescence of (a) model compound 1 and
(b) polymers. Source: Ref. [129]. Copyright 2018, American Chemical Society. (c) Molecular structures of
compounds 2 and 3 and their polymorphs under UV light (350 nm); 2c and 2c′ represent different cyan-­
fluorescence polymorphs of 2; 3b, 3c, and 3p represent the blue-­, cyan-­and purple-­fluorescence polymorphs
of 3, respectively. Ref. [131].

clusteroluminescence at the clustering state (Figure 5.9b). Similar to compound 1, Zhu et al. syn-
thesized the C-­6 unsubstituted tetrahydropyrimidines 2 and 3 which also possessed various n and
π electrons [131]. As expected, clusteroluminescence is observed in these two structures and theo-
retical calculations verify the existence of through-­space n/π-­electron conjugation. At the same
time, polymorphs-­dependent clusteroluminescence is obtained in their different crystals. As
shown in Figure 5.9c, two polymorphs (2c and 2c′) for 2 and three polymorphs (3b, 3c, and 3p) for
3 are measured with different emission colors. Crystal-­packing analysis and the relevant calcula-
tion suggest that the polymorphs-­dependent clusteroluminescence is ingeniously tuned by the
synergistic contribution of through-­space and through-­bond conjugation.

5.5 ­Other Systems

The above three parts have introduced organic clusteroluminogens with pure n/π and hybrid (n,
π)-­electrons. Apart from that, inorganic and organometallic systems are also disclosed with the
CTE effect. For example, in 2012, Xie et al. reported that the AIE effect of oligomeric Au(I)−thi-
olate complex, which was prepared from HAuCl4 and reduced l-­Glutathione (GSH) (Figure 5.10),
GSH is consists of carboxylic acids, amine, amide, and S–S groups [25, 132]. The Au(I)−thiolate
complex shows no emission in aqueous solution, but a bright yellow emission appears with the
addition of ethanol. There is no π-­conjugated structure in this complex, which is a new system of
clusteroluminogens. Au(0)@Au(I)−thiolate nanoclusters (NCs) are obtained from the same reac-
tants with Au(I)−thiolate complex but different reaction conditions. The resultant Au−thiolate
5.6 ­Summar 167

Aggregation-induced emission (AIE)


Au(0)@Au(l)-thiolate NCs with AIE
of oligomeric Au(I)-thiolate complexes

R
Solvent-induced HAuCl4 ∆
S +
R Au
S n aggregation GSH 70 °C

fe 0 30 60 65 70 75 80 85 90 95% 5

1
(+)

Figure 5.10  CTE of Au(I)–thiolate complexes to ultrabright Au(0)@Au(I)–thiolate core–shell nanoclusters.


Source: Ref. [132]. Copyright 2012, American Chemical Society.

NCs have strong orange luminescence both in solution and solid states, which suggests that the
clusters have already formed in the solution. The performances of Au(I)−thiolate complex and
Au−thiolate NCs are in analogy with linear and dendritic polymers, respectively. The authors have
tried to separate the Au−thiolate NCs with native polyacrylamide gel electrophoresis (PAGE). Five
bands 1–5 are observed. From bands 1 to 5, their molecular weight is increased. However, contrary
to the prediction from the Jellium model where larger-­sized NCs should emit longer wavelengths,
they found a blue-­shifted λem from 1 to 5.
Since GSH has multiple heteroatoms (N, O, S) and carbonyl groups, the NCs should possess strong
inter/intramolecular through-­space (n, π)-­electron conjugation according to the studies on com-
pound 1. However, currently, it is still unclear whether ligand-­to-­metal (LMCT), metal-­to-­ligand
charge transfer (MLCT), or surface plasmon resonance contribute to the clusteroluminescence of
NCs or not. In other words, what’s the role of central metals in their luminescent behavior, that is,
a quantum effect or electronic effect? In practical use, Xie et  al. come up with an experienced
method to prepare emissive Au NCs through regulating the thiolate-­to-­Au ratio. Conventional Au–
thiolate NCs have short Au(I)–thiolate motifs on the surface because of a low thiolate-­to-­Au ratio,
resulting in a large metallic core and thin Au(I)–thiolate surface. By increasing the thiolate-­to-­Au
ratio, the oligomeric Au(I)–thiolate surface to the metal core ratio is also improved. And the size of
the metallic core is smaller than the conventional Au–thiolate NCs. Finally, a luminescent Au(0)@
Au(I)−thiolate NCs are achieved. Based on the above results, it suggests that the surface of the Au–
thiolate NCs plays a more important role than the metallic core in the clusteroluminescence.

5.6 ­Summary

In this chapter, we have summarized the CTE effect from four systems, pure n-­electron, pure π-­
electron, hybrid (n, π)-­electrons, and metal clusters based on different working mechanisms.
Different from conventional conjugated chromophores, studies on the clusteroluminescence start
from the late twentieth century, and only the phenomenon are reported without a clear photo-
physical theory. Because of that, this direction did not form a systematic research area and gradu-
ally fade away. Thanks to the flourishing development of AIE, aggregate-­state studies have become
as important as molecular science [133]. Meanwhile, the proposal of the CTE concept of Tang’s
168 5  Clusterization-­Triggered Emission

group promotes unconventional luminescence to a well-­known property in chromophores.


Currently, not just the typical CTE systems introduced in this chapter, the contribution through-­
space conjugation are also considered in many conjugated emitters. For example, tetraphenyleth-
ylene (TPE), acting as the star AIEgen, its aggregate-­state emission at 470 nm is ascribed to the
through-­bond conjugation among four benzene rings and the middle double bond. However,
recent studies indicate that the 470 nm emission in TPE is caused by intramolecular through-­space
π-­electron conjugation  [134]. So, we believe that the summarized CTE effect and the through-­
space conjugation not only apply to nonconjugated clusteroluminogens but also change people’s
perceptions of the traditional π-­conjugated chromophores.
To summarize, we put forward the following six characteristic features of CTE which were pro-
posed in our recently published review article. It is possible that they may not be applicable to all
the clusteroluminogens simultaneously. (i) Clusteroluminogens are based upon nonconjugated
molecular motifs, where n or π electron-­based groups are separated by nonconjugated linkers;
(ii) in the isolated states, only the electron transition belonging to n or π electron-­based groups can
be detected in the PL spectra of clusteroluminogens. However, an intrinsic luminescence appears
at longer wavelengths upon clustering of these molecules, which is termed as the clusterolumines-
cence; (iii) the excitation spectra of clusteroluminogens appear red-­shifted in comparison with
their absorption spectra. In other words, these molecules are excited by light whose wavelength is
out of the range of strong absorption peaks; (iv) clusteroluminogens show excitation-­dependent
luminescence properties, where excitation at longer-­wavelength results in a red-­shifted emission;
(v) clusteroluminogens show size-­dependent emission properties. The concept of “size” is related
to different generations of dendrimers, molecular weights of the polymers, the diameter of the
nanoparticles, concentration, etc. Generally, the increase in size would shift PL spectra to the red,
and the emission intensity will be enhanced; (vi) clusteroluminogens, especially polymer systems,
are capable to phosphoresce, and some of them show room-­temperature phosphorescence.
Currently, the concept of carbon dots, nitrogen dots, oxygen dots, and sulfur dots, etc. are rising
with great attention because of their high emission efficiency and low economic cost. In these
systems, the reported properties are almost similar to the above-­mentioned six characteristics in
clusteroluminogens but the working mechanism is seldom discussed due to their indistinct com-
ponents and chemical structure. Based on the theories illustrated in this chapter, these different
elemental dots should belong to the concept of CTE, but further works are required to clarify the
emitting species inside these structures. Serving as the emission mechanism for the CTE effect,
through-­space conjugations are sensitive to the clustering conformation and that is the reason why
the clusteroluminogens always show mechanochromic luminescence effect. It is foreseeable that
CTE will be applicable in several areas such as process monitoring of cellular uptake and reactions,
structure visualization, and sensing. Meanwhile, in comparison to existing types of probe mole-
cules, clusteroluminogens show incomparable advantages (i.e., low toxicities, good biocompatibil-
ities) which are conducive to biological applications.

­References

1 Zhao, Z., Lam, J.W.Y., Tang, B.Z. (2013). Self-­assembly of organic luminophores with gelation-­
enhanced emission characteristics. Soft. Matter. 9 (18): 4564.
2 Song, F., Xu, Z., Zhang, Q.S., Zhao, Z., Zhang, H., Zhao, W. et al. (2018). Highly efficient circularly
polarized electroluminescence from aggregation-­induced emission luminogens with amplified
chirality and delayed fluorescence. Adv. Funct. Mater. 28 (17): 1800051.
 ­Reference 169

3 Li, Y.C., Liu, J.Y., Zhao, Y.D., Cao, Y.C. (2017). Recent advancements of high efficient donor-­
acceptor type blue small molecule applied for OLEDs. Mater. Today. 20 (5): 258–266.
4 Zhao, F., Ma, D. (2017). Approaches to high performance white organic light-­emitting diodes for
general lighting. Mater. Chem. Front. 1 (10): 1933–1950.
5 Xu, W.H., Lee, M.M.S., Zhang, Z.H., Sung, H.H.Y., Williams, I.D., Kwok, R.T.K. et al. (2019). Facile
synthesis of AIEgens with wide color tunability for cellular imaging and therapy. Chem. Sci. 10
(12): 3494–3501.
6 Shen, X.Y., Wang, Y.J., Zhang, H.K., Qin, A.J., Sun, J.Z., Tang, B.Z. (2014). Conjugates of
tetraphenylethene and diketopyrrolopyrrole: tuning the emission properties with phenyl bridges.
Chem. Comm. 50 (63): 8747–8750.
7 Li, L.Z., Chen, M., Zhang, H.K., Nie, H., Sun, J.Z., Qin, A.J. et al. (2015). Influence of the number
and substitution position of phenyl groups on the aggregation-­enhanced emission of benzene-­
cored luminogens. Chem. Comm. 51 (23): 4830–4833.
8 Liu, S.J., Zhou, X., Zhang, H.K., Ou, H.L., Lam, J.W.Y., Liu, Y. et al. (2019). Molecular motion in
aggregates: manipulating TICT for boosting photothermal theranostics. J. Am. Chem. Soc. 141 (13):
5359–5368.
9 Zhu, C., Kwok, R.T.K., Lam, J.W.Y., Tang, B.Z. (2018). Aggregation-­induced emission: a trailblazing
journey to the field of biomedicine. ACS Appl. Bio. Mater. 1 (6): 1768–1786.
10 Chen, M., Li, L.Z., Wu, H.Q., Pan, L.X., Li, S.W., He, B.R. et al. (2018). Unveiling the different
emission behavior of polytriazoles constructed from pyrazine-­based AIE monomers by click
polymerization. ACS Appl. Mater. Interfaces. 10 (15): 12181–12188.
11 Zhang, H., Liu, J., Du, L., Ma, C., Leung, N.L.C., Niu, Y. et al. (2019). Drawing a clear mechanistic
picture for the aggregation-­induced emission process. Mater. Chem. Front. 3 (6): 1143–1150.
12 Birks, J.B., Braga, C.L., Lumb, M.D. (1965). “Excimer” fluorescence. VI. Benzene, toluene, p-­xylene
and mesitylene. Proc. R. Soc. A: Math. Phys. Eng. Sci. 283 (1392): 83–99.
13 Förster, T., Kasper, K. (1954). Ein Konzentrationsumschlag der Fluoreszenz. Z. Phys. Chem.
(Muenchen, Ger). 1: 275−277.
14 Perrin, J.B. (1923). Chimie Physique-­Observations sur la fluorescence. Académie des sciences. 177.
15 Birks, J.B. (1970). Photophysics of Aromatic Molecules. London: Wiley-­Interscience.
16 Luo, J., Xie, Z., Lam, J.W.Y., Cheng, L., Chen, H, Qiu, C. et al. (2001). Aggregation-­induced
emission of 1-­methyl-­1,2,3,4,5-­pentaphenylsilole. Chem. Comm. 0 (18): 1740–1741.
17 Hong, Y.N., Lam, J.W.Y., Tang, B.Z. (2011). Aggregation-­induced emission. Chem. Soc. Rev. 40 (11):
5361–5388.
18 Hu, Z., Zhang, H., Chen, Y., Wang, Q., Elsegood, M.R.J., Teat, S.J. et al. (2020).
Tetraphenylethylene-­based color-­tunable AIE-­ESIPT chromophores. Dyes Pigm. 175: 108175.
19 Qi, J., Li, J., Liu, R., Li, Q., Zhang, H., Lam, J.W.Y. et al. (2019). Boosting fluorescence-­photoacoustic-­
raman properties in one fluorophore for precise cancer surgery. Chem. 5 (10): 2657–2677.
20 Li, Y., Cai, Z., Liu, S., Zhang, H., Wong, S.T.H., Lam, J.W.Y. et al. (2020). Design of AIEgens for
near-­infrared IIb imaging through structural modulation at molecular and morphological levels.
Nat. Commun. 11 (1): 1255.
21 Liu, S., Chen, C., Li, Y., Zhang, H., Liu, J., Wang, R. et al. (2020). Constitutional isomerization
enables bright NIR-­II AIEgen for brain-­inflammation imaging. Adv. Funct. Mater. 30 (7): 1908125.
22 Zheng, Z., Liu, H., Zhai, S., Zhang, H., Shan, G, Kwok, R.T.K. et al. (2020). Highly efficient singlet
oxygen generation, two-­photon photodynamic therapy and melanoma ablation by rationally
designed mitochondria-­specific near-­infrared AIEgens. Chem. Sci. 11 (9): 2494–2503.
23 Mei, J., Leung, N.L., Kwok, R.T., Lam, J.W.Y., Tang, B.Z. (2015). Aggregation-­induced emission:
together we shine, United We Soar! Chem. Rev. 115 (21): 11718–11940.
170 5  Clusterization-­Triggered Emission

24 Wang, Z.Y., Cheng, X., Qin, A.J., Zhang, H.K., Sun, J.Z., Tang, B.Z. (2018). Multiple stimuli
responses of stereo-­isomers of AIE-­active ethynylene-­bridged and pyridyl-­modified
tetraphenylethene. J. Phys. Chem. B. 122 (7): 2165–2176.
25 Zhang, J., Liu, Q.M., Wu, W.J., Peng, J.H., Zhang, H.K., Song, F.Y. et al. (2019). Real-­time
monitoring of hierarchical self-­assembly and induction of circularly polarized luminescence from
achiral luminogens. ACS Nano. 13 (3): 3618–3628.
26 Zhang, H.K., Li, H.K., Wang, J., Sun, J.Z., Qin, A.J., Tang, B.Z. (2015). Axial chiral aggregation-­
induced emission luminogens with aggregation-­annihilated circular dichroism effect. J. Mater.
Chem. C. 3 (20): 5162–5166.
27 Yuan, W.Z., Lu, P., Chen, S.M., Lam, J.W.Y., Wang, Z.M., Liu, Y. et al. (2010). Changing the
behavior of chromophores from aggregation-­caused quenching to aggregation-­induced emission:
development of highly efficient light emitters in the solid state. Adv. Mater. 22 (19): 2159–2163.
28 Yan, X.Z., Cook, T.R., Wang, P., Huang, F.H., Stang, P.J. (2015). Highly emissive platinum(II)
metallacages. Nat. Chem. 7 (4): 342–348.
29 Yan, X.Z., Wang, H.Z., Hauke, C.E., Cook, T.R., Wang, M., Saha, M.L. et al. (2015). A suite of
tetraphenylethylene-­based discrete organoplatinum(II) metallacycles: controllable structure and
stoichiometry, aggregation-­induced emission, and nitroaromatics sensing. J. Am. Chem. Soc. 137
(48): 15276–15286.
30 Liu, H.J., Zeng, J.J., Guo, J.J., Nie, H., Zhao, Z.J., Tang, B.Z. (2018). High-­performance non-­doped
OLEDs with nearly 100% exciton use and negligible efficiency roll-­off. Angew. Chem. Int. Ed. 57
(30): 9290–9294.
31 Kwok, R.T.K., Leung, C.W.T., Lam, J.W.Y., Tang, B.Z. (2015). Biosensing by luminogens with
aggregation-­induced emission characteristics. Chem. Soc. Rev. 44 (13): 4228–4238.
32 Li, K., Liu, B. (2014). Polymer-­encapsulated organic nanoparticles for fluorescence and
photoacoustic imaging. Chem. Soc. Rev. 43 (18): 6570–6597.
33 Zhao, E.G., Chen, Y.L., Chen, S.J., Deng, H.Q., Gui, C., Leung, C.W.T. et al. (2015). A luminogen
with aggregation-­induced emission characteristics for wash-­free bacterial imaging, high-­throughput
antibiotics screening and bacterial susceptibility evaluation. Adv. Mater. 27 (33): 4931–4937.
34 Chen, Y.C., Zhang, W.J., Zhao, Z., Cai, Y.J., Gong, J.Y., Kwok, R.T.K. et al. (2018). An easily
accessible ionic aggregation-­induced emission luminogen with hydrogen-­bonding-­switchable
emission and wash-­free imaging ability. Angew. Chem. Int. Ed. 57 (18): 5011–5015.
35 Mei, J., Hong, Y.N., Lam, J.W.Y., Qin, A,J., Tang, Y.H., Tang, B.Z. (2014). Aggregation-­induced
emission: the whole is more brilliant than the parts. Adv. Mater. 26 (31): 5429–5479.
36 Wang, L., Zhang, H.K., Qin, A.J., Jin, Q., Tang, B.Z., Ji, J. (2016). Theranostic hyaluronic acid
prodrug micelles with aggregation-­induced emission characteristics for targeted drug delivery. Sci.
China Chem. 59 (12): 1609–1615.
37 Liu, S.J., Cheng, Y.H., Zhang, H.K., Qiu, Z.J., Kwok, R.T.K., Lam, J.W.Y. et al. (2018). In situ
monitoring of RAFT polymerization by tetraphenylethylene-­containing agents with aggregation-­
induced emission characteristics. Angew. Chem. Int. Ed. 57 (21): 6274–6278.
38 Cheng, Y.H., Liu, S.J., Song, F.Y., Khorloo, M., Zhang, H.K., Kwok, R.T.K. et al. (2019). Facile
emission color tuning and circularly polarized light generation of single luminogen in engineering
robust forms. Mater. Horiz. 6 (2): 405–411.
39 Zhang, H., Kwok, R.T.K., Lam, J.W.Y., Tang, B.Z. (2018). Aggregation and chirality. Liquid Crystals
XXII. 107350H.
40 Alam, P., Leung, N.L.C., Cheng, Y.H., Zhang, H.K, Liu, J.K., Wu, W.J. et al. (2019). Spontaneous
and fast molecular motion at room temperature in the solid state. Angew. Chem. Int. Ed. 58 (14):
4536–4540.
 ­Reference 171

41 Zhao, Z., Chen, C., Wu, W.T., Wang, F.F., Du, L.L., Zhang, X.Y. et al. (2019). Highly efficient
photothermal nanoagent achieved by harvesting energy via excited-­state intramolecular motion
within nanoparticles. Nat. Commun. 10 (1): 768.
42 Wang, Z.Y., Zhang, P.F., Liu, H.X., Zhao, Z., Xiong, L.H., He, W. et al. (2019). Robust serum
albumin-­responsive AIEgen enables latent bloodstain visualization in high resolution and
reliability for crime scene investigation. ACS Appl. Mater. Interfaces. 11 (19): 17306–17312.
43 Liu, S., Zhang, H., Li, Y., Liu, J., Du, L., Chen, M. et al. (2018). Strategies to enhance the
photosensitization: polymerization and the donor–acceptor even–odd effect. Angew. Chem. Int. Ed.
57 (46): 15189–15193.
44 Niu, G., Zheng, X., Zhao, Z., Zhang, H., Wang, J., He, X. et al. (2019). Functionalized acrylonitriles
with aggregation-­induced emission: structure tuning by simple reaction-­condition variation,
efficient red emission, and two-­photon bioimaging. J. Am. Chem. Soc. 141 (38): 15111–15120.
45 Peng, H.Q., Liu, B., Wei, P., Zhang, P., Zhang, H., Zhang, J., et al. (2019). Visualizing the initial step
of self-­assembly and the phase transition by stereogenic amphiphiles with aggregation-­induced
emission. ACS Nano. 13 (1): 839–846.
46 Zhang, H., Sun, J.Z., Liu, J., Kwok, R.T.K., Lam, J.W.Y., Tang, B.Z. (2019) Visualizing and
monitoring interface structures and dynamics by luminogens with aggregation-­induced emission.
J. Appl. Phys. 126 (5): 050901.
47 Zhang, H., Zheng, X., Kwok, R.T.K., Wang, J., Leung, N.L.C., Shi, L. et al. (2018). In situ
monitoring of molecular aggregation using circular dichroism. Nat. Commun. 9 (1): 4961.
48 Qian, J., Tang, B.Z. (2017). AIE luminogens for bioimaging and theranostics: from organelles to
animals. Chem. 3 (1): 56–91.
49 Qi, J., Chen, C., Ding, D., Tang, B.Z. (2018). Aggregation-­induced emission luminogens: union is
strength, gathering illuminates healthcare. Adv. Healthc. Mater. 7 (20): 1800477.
50 Gong, Y.Y., Tan, Y.Q., Mei, J., Zhang, Y.R., Yuan, W.Z., Zhang, Y.M. et al. (2013). Room temperature
phosphorescence from natural products: crystallization matters. Sci. China Chem. 56
(9):1178–1182.
51 Zhu, S., Zhang, J., Wang, L., Song, Y., Zhang, G., Wang, H. et al. (2012). A general route to make
non-­conjugated linear polymers luminescent. Chem. Commun. 48 (88): 10889–10891.
52 Yan, J.J., Wang, Z.K., Lin, X.S., Hong, C.Y., Liang, H.J., Pan, C.Y. et al. (2012). Polymerizing
nonfluorescent monomers without incorporating any fluorescent agent produces strong
fluorescent polymers. Adv. Mater. 24 (41): 5617–5624.
53 Chen, X., He, Z., Kausar, F., Chen, G., Zhang, Y., Yuan, W.Z. (2018). Aggregation-­induced dual
emission and unusual luminescence beyond excimer emission of poly(ethylene terephthalate).
Macromolecules. 51 (21): 9035–9042.
54 Bacon, F. (1605). George Fabyan Collection (Library of Congress). The proficience and aduancement
of learning, diuine and humane. At London: Printed for Henrie Tomes. 1 (45): 118–121.
55 Wang, P., Liu, C., Tang, W.Q., Ren, S.X., Chen, Z.J., Guo, Y.R. et al. (2019). Molecular glue strategy:
large-­scale conversion of clustering-­induced emission luminogen to carbon dots. ACS Appl. Mater.
Interfaces. 11 (21): 19301–19307.
56 Du, L.L., Jiang, B.L., Chen, X.H., Wang, Y.Z., Zou, L.M., Liu, Y.L. et al. (2019). Clustering-­triggered
emission of cellulose and its derivatives. Chinese J. Polym. Sci. 37 (4): 409–415.
57 Yuan, W.Z., Zhang, Y. (2017). Nonconventional macromolecular luminogens with aggregation-­
induced emission characteristics. J. Polym. Sci. Polym. Chem. 55 (4): 560–574.
58 Zink, J.I., Hardy, G.E., Sutton, J.E. (1976). Triboluminescence of sugars. J. Phys. Chem. 80 (3): 248–249.
59 Li, M., Li, X., An, X., Chen, Z., Xiao, H. (2019). Clustering-­triggered emission of carboxymethylated
nanocellulose. Front. Chem. 7: 447.
172 5  Clusterization-­Triggered Emission

60 Xing, C.M., Lam, J.W.Y., Qin, A., Tang, B.Z. (2007). Unique photoluminescence from
nonconjugated alternating copolymer poly[(maleic anhydride)-­alt-­(vinyl acetate)]. Polym. Mater.
Sci. Eng. 96: 418–419.
61 Yuan, W.Z., Zhang, Y.M. (2017). Nonconventional macromolecular luminogens with aggregation-­
induced emission characteristics. J. Polym. Sci. Pol. Chem. 55 (4): 560–574.
62 Bekiari, V., Lianos, P. (1998). Characterization of photoluminescence from a material made by
interaction of (3-­aminopropyl)triethoxysilane with acetic acid. Langmuir. 14: 3459–3461.
63 Bekiari, V., Lianos, P., Stangar, U.L., Orel, B., Judeinstein, P. (2000). Optimization of the intensity
of luminescence emission from silica/poly(ethylene oxide) and silica/poly(propylene oxide)
nanocomposite gels. Chem. Mater. 12 (10): 3095–3099.
64 Carlos, L.D., Bermudez, V.D., Amaral, V.S., Nunes, S.C., Silva, N.J.O., Ferreira, R.A.S. et al. (2007).
Nanoscopic photoluminescence memory as a fingerprint of complexity in self-­assembled alkyl/
siloxane hybrids. Adv. Mater. 19 (3): 341–348.
65 Lu, H., Zhang, J., Feng, S.Y. (2015) Controllable photophysical properties and self-­assembly of
siloxane-­poly(amidoamine) dendrimers. Phys. Chem. Chem. Phys. 17 (40): 26783–26789.
66 Niu, S., Yan, H.X., Li, S., Xu, P.L., Zhi, X.L., Li, T.T. (2016). Bright blue photoluminescence emitted
from the novel hyperbranched polysiloxane-­containing unconventional chromogens. Macromol.
Chem. Phys. 217 (10): 1185–1190.
67 Lu, H., Feng, L., Li, S., Zhang, J., Lu, H., Feng, S. (2015). Unexpected strong blue
photoluminescence produced from the aggregation of unconventional chromophores in novel
siloxane–poly(amidoamine) dendrimers. Macromolecules. 48 (3): 476–482.
68 Larson, C.L., Tucker, S.A. (2001). Intrinsic fluorescence of carboxylate-­terminated polyamido
amine dendrimers. Appl. Spectrosc. 55: 679–683.
69 Varnavski, O., Ispasoiu, R.G., Balogh, L., Tomalia, D., Goodson, T. (2001). Ultrafast time-­resolved
photoluminescence from novel metal–dendrimer nanocomposites. J. Chem. Phys. 114 (5):
1962–1965.
70 Lee, W.I., Bae, Y.J., Bard, A.J. (2004). Strong blue photoluminescence and ECL from OH-­
terminated PAMAM dendrimers in the absence of gold nanoparticles. J. Am. Chem. Soc. 126 (27):
8358–8359.
71 Liu, X.Y., Zeng, Y., Liu, J., Li, P., Zhang, D.S., Zhang, X.H. et al. (2015). Highly emissive
nanoparticles based on AIE-­active molecule and PAMAM dendritic “Molecular Glue”. Langmuir.
31 (15): 4386–4393.
72 Al-­Jamal, K.T., Ruenraroengsak, P., Hartell, N., Florence, A.T. (2006). An intrinsically fluorescent
dendrimer as a nanoprobe of cell transport. J. Drug Target. 14 (6): 405–412.
73 Marek, P., Gupta, R., Raleigh, D.P. (2008). The fluorescent amino acid p-­cyanophenylalanine
provides an intrinsic probe of amyloid formation. Chembiochem. 9 (9): 1372–1374.
74 Turoverov, K.K., Kuznetsova, I.M. (2003). Intrinsic fluorescence of actin. J. Fluoresc. 13 (1): 41–57.
75 Schilter, D. (2017). Fluorescence: isolated rings do big things. Nat. Rev. Chem. 1: 0097.
76 He, Z.K., Ke, C.Q., Tang, B.Z. (2018). Journey of aggregation-­induced emission research. ACS
Omega. 3 (3): 3267–3277.
77 Wang, H., Zhao, E., Lam, J.W.Y., Tang, B.Z. (2015). AIE luminogens: emission brightened by
aggregation. Mater. Today. 18 (7): 365–377.
78 Zhang, Y., Shen, P., He, B., Luo, W., Zhao, Z., Tang, B.Z. (2018). New fluorescent through-­space
conjugated polymers: synthesis, optical properties and explosive detection. Poly. Chem. 9 (5):
558–564.
79 Zhang, H., Zhao, Z., McGonigal, P.R., Ye, R., Liu, S., Lam, J.W.Y. et al. (2020). Clusterization-­
triggered emission: uncommon luminescence from common materials. Mater. Today. 32: 275–292.
 ­Reference 173

80 Tomalia, D.A., Klajnert-­Maculewicz, B., Johnson, K.A.M., Brinkman, H.F., Janaszewska, A.,
Hedstrand, D.M. (2019). Non-­traditional intrinsic luminescence: inexplicable blue fluorescence
observed for dendrimers, macromolecules and small molecular structures lacking traditional/
conventional luminophores. Prog. Polym. Sci. 90: 35–117.
81 Studzian, M., Pułaski, Ł., Tomalia, D.A., Klajnert-­Maculewicz, B. (2019). Non-­Traditional Intrinsic
Luminescence (NTIL): dynamic quenching demonstrates the presence of two distinct fluorophore
types associated with NTIL behavior in pyrrolidone-­terminated PAMAM dendrimers. J. Phys.
Chem. C. 123 (29): 18007–18016.
82 Konopka, M., Janaszorska, A., Johnson, K.A.M., Hedstrand, D., Tomalia, D.A., Klajnert-­
Maculewicz, B. (2018). Determination of non-­traditional intrinsic fluorescence (NTIF) emission
sites in 1-­(4-­carbomethoxypyrrolidone)-­PAMAM dendrimers using CNDP-­based quenching
studies. J. Nanopart. Res. 20: 220.
83 Konopka, M., Janaszewska, A., Klajnert-­Maculewicz, B. (2018). Intrinsic fluorescence of PAMAM
dendrimers-­quenching studies. Polymers. 10 (5): 540.
84 Ye, R., Liu, Y., Zhang, H., Su, H., Zhang, Y., Xu, L. et al. (2017). Non-­conventional fluorescent
biogenic and synthetic polymers without aromatic rings. Poly. Chem. 8 (10): 1722–1727.
85 Zhang, H.K., Zheng, X.Y., Xie, N., He, Z.K., Tiu, J.K, Leung, N.L.C. et al. (2017). Why do simple
molecules with “Isolated” phenyl rings emit visible light? J. Am. Chem. Soc. 139 (45): 16264–16272.
86 Shang, C., Wei, N., Zhuo, H., Shao, Y., Zhang, Q., Zhang, Z. et al. (2017). Highly emissive
poly(maleic anhydride-­alt-­vinyl pyrrolidone) with molecular weight-­dependent and excitation-­
dependent fluorescence. J. Mater. Chem. C. 5 (32): 8082–8090.
87 van Dam, B., Nie, H., Ju, B., Marino, E., Paulusse, J.M.J., Schall, P. et al. (2017). Excitation-­
dependent photoluminescence from single-­carbon dots. Small. 13:1702098.
88 Li, C., Liu, W.J., Sun, X.B., Pan, W., Yu, G.F., Wang, J.P. (2018). Excitation dependent emission
combined with different quenching manners supports carbon dots to achieve multi-­mode sensing.
Sens. Actuat. B Chem. 263: 1–9.
89 Zhao, Z., Chen, X., Wang, Q., Yang, T., Zhang, Y., Yuan, W.Z. (2019). Sulphur-­containing
nonaromatic polymers: clustering-­triggered emission and luminescence regulation by oxidation.
Poly. Chem. 10: 3639–3646
90 Li, Y., Zhao, Y., Song, Y., Chang, Y., Liu, L. (2011). The study on the UV and fluorescence properties
of PAMAM dendrimers. Spectrosc. Spect. Anal. 31 (02): 422–426.
91 Deng, L., Wang, X.L., Kuang, Y., Wang, C., Luo, L., Wang, F. et al. (2015). Development of hydrophilicity
gradient ultracentrifugation method for photoluminescence investigation of separated non-­sedimental
carbon dots. Nano Res. 8 (9): 2810–2821.
92 Zhou, Q., Wang, Z., Dou, X., Wang, Y., Liu, S., Zhang, Y. et al. (2019). Emission mechanism
understanding and tunable persistent room temperature phosphorescence of amorphous
nonaromatic polymers. Mater. Chem. Front. 3 (2): 257–264.
93 Zhang, T., Zhao, Z., Ma, H., Zhang, Y., Yuan, W.Z. (2019). Polymorphic pure organic luminogens
with through-­space conjugation and persistent room-­temperature phosphorescence. Chem. Asian
J. 14 (6): 884–889.
94 Halpern, A.M. (1970). The vapor state emission from a saturated amine. Chem. Phys. Lett. 6 (4): 296–298.
95 Halpern, A.M., Roebber, J.L., Weiss, K. (1968). Electronic structure of cage amines: absorption
spectra of triethylenediamine and quinuclidine. J. Chem. Phys. 49 (3): 1348–1357.
96 Basch, H., Robin, M.B., Kuebler, N.A. (1967). Electronic states of the amide group. J. Chem. Phys.
47 (4): 1201–1210.
97 Colle, R., Moscardo, F., Riani, P., Salvetti, O. (1977). V-­N Vertical transition of planar ethylene.
Theor. Chim. Acta. 44 (1): 1–7.
174 5  Clusterization-­Triggered Emission

98 Chu, C.C., Imae, T. (2009). Fluorescence investigations of oxygen-­doped simple amine compared
with fluorescent PAMAM dendrimer. Macromol. Rapid. Commun. 30 (2): 89–93.
99 Zhang, H. (2018). Aggregation-­Induced Emission: Mechanistic Study, Clusteroluminescence and
Kinetic Resolution. PhD Thesis of HKUST.
100 Pastor-­Pérez, L., Chen, Y., Shen, Z., Lahoz, A., Stiriba, S.E. (2007). Unprecedented blue intrinsic
photoluminescence from hyperbranched and linear polyethylenimines: polymer architectures
and pH-­Effects. Macromol. Rapid. Comm. 28 (13): 1404–1409.
101 Wang, D., Imae, T. (2004). Fluorescence emission from dendrimers and Its pH dependence. J. Am.
Chem. Soc. 126 (41): 13204–13205.
102 Tamano, K., Imae, T. (2008). Investigation of luminescent poly(propylene imine) dendrimer.
J. Nanosci. Nanotechnol. 8 (9): 4329–4334.
103 Wang, Y., Bin, X., Chen, X., Zheng, S., Zhang, Y., Yuan, W.Z. (2018). Emission and emissive
mechanism of nonaromatic oxygen clusters. Macromol. Rapid. Comm. 39 (21): 1800528.
104 Lobyshev, V.I., Shikhlinskaya, R.E., Ryzhikov, B.D. (1999). Experimental evidence for intrinsic
luminescence of water. J. Mol. Liq. 82 (1–2): 73–81.
105 Wang, Q., Dou, X., Chen, X., Zhao, Z., Wang, S., Wang, Y. et al. (2019). Reevaluating protein
photoluminescence: remarkable visible luminescence upon concentration and insight into the
emission mechanism. Angew Chem. Int. Ed. 58 (36): 12667–12673.
106 Zhang, Z., Li, D., Jiang, W., Wang, Z. (2018). The electron density delocalization of hydrogen bond
systems. Adv. Phys-­X. 3 (1): 1428915.
107 Du, Y., Bai, T., Ding, F., Yan, H., Zhao, Y., Feng, W. (2019). The inherent blue luminescence from
oligomeric siloxanes. Polym. J. 51 (9): 869–882.
108 Yanari, S.S., Bovey, F.A., Lumry, R. (1963). Fluorescence of styrene homopolymers and copolymers.
Nature. 200 (4903): 242–244.
109 Hirayama, F. (1965). Intramolecular excimer formation. I. Diphenyl and triphenyl alkanes.
J. Chem. Phys. 42 (9): 3163–3171.
110 Morisaki, Y., Gon, M., Chujo, Y. (2013). Conjugated microporous polymers consisting of
tetrasubstituted [2,2]paracyclophane junctions. J. Polym. Sci. Pol. Chem. 51 (10): 2311–2316.
111 Dang, D., Zhang, H., Xu, Y., Xu, R., Wang, Z., Kwok, R.T.K. et al. (2019). Super-­resolution
visualization of self-­assembling helical fibers using aggregation-­induced emission luminogens in
stimulated emission depletion nanoscopy. ACS Nano. 13 (10): 11863–11873.
112 Zhang, Y.Y., He, B.R., Luo, W.W., Peng, H.R., Chen, S.M., Hu, R.R. et al. (2016). Aggregation-­
enhanced emission and through-­space conjugation of tetraarylethanes and folded
tetraarylethenes. J. Mater. Chem. C. 4 (39): 9316–9324.
113 Martinez, A.G., Barcina, J.O., Cerezo, A.D., Rivas, R.G. (1998). Hindered rotation in
diphenylmethane derivatives. Electrostatic vs charge-­transfer and homoconjugative aryl-­aryl
interactions. J. Am. Chem. Soc. 120 (4): 673–679.
114 Chen, L., Wang, Y.H., He, B.R., Nie, H., Hu. R.R., Huang, F. et al. (2015). Multichannel
conductance of folded single-­molecule wires aided by through-­space conjugation. Angew. Chem.
Int. Ed. 54 (14): 4231–4235.
115 Shen, P.C., Zhuang, Z.Y., Jiang, X.F., Li, J.S., Yao, S.N., Zhao, Z.J. et al. (2019). Through-­space
conjugation: an effective strategy for stabilizing intramolecular charge-­transfer states. J. Phys.
Chem. Lett. 10 (11): 2648–2656.
116 Shen, P.C., Zhuang, Z.Y., Zhao, Z.J., Tang, B.Z. (2016). Recent advances of folded tetraphenylethene
derivatives featuring through-­space conjugation. Chinese Chem. Lett. 27 (8): 1115–1123.
117 Sturala, J., Etherington, M.K., Bismillah, A.N., Higginbotham, H.F., Trewby, W., Aguilar,
J.A. et al. (2017). Excited-­state aromatic interactions in the aggregation-­induced emission of
molecular rotors. J. Am. Chem. Soc. 139 (49): 17882–17889.
 ­Reference 175

118 Guo, J., Fan. J., Liu, X., Zhao, Z., Tang, B.Z. (2019). Photomechanical luminescence from
through-­space conjugated AIEgens. Angew. Chem. Int. Ed. DOI:10.1002/anie.201913383.
119 Zhang, H., Du, L., Wang, L., Liu, J., Wan, Q., Kwok, R.T.K. et al. (2019). Visualization and manipulation
of molecular motion in the solid state through photoinduced clusteroluminescence. J. Phys. Chem. Lett.
10 (22): 7077–7085.
120 Wang, H., Xing, H., Gong, J., Zhang, H., Zhang, J., Wei, P. et al. (2020). “Living” luminogens: light
driven ACQ-­to-­AIE transformation accompanied with solid-­state actuation. Mater. Horiz. DOI:
10.1039/D0MH00447B.
121 Heyne, B. (2016). Self-­assembly of organic dyes in supramolecular aggregates. Photochem.
Photobiol. Sci. 15 (9): 1103–1114.
122 Xing, C.M., Yang, W.T. (2004). A novel, facile method for the preparation of uniform, reactive
maleic anhydride vinyl acetate copolymer micro-­and nanospheres. Macromol. Rapid. Comm. 25
(17): 1568–1574.
123 Bosma, M., Vorenkamp, E.J., Tenbrinke, G., Challa, G. (1988). Influence of the degree of
hydrolysis of poly(styrene-­alt-­maleic anhydride) on miscibility with polyvinyl acetate). Polymer.
29 (9): 1694–1698.
124 Sunel, V., Popa, M., Stoican, A.D., Popa, A.A., Uglea, C.V. (2008). Poly (maleic anhydride-­alt-­vinyl
acetate) conjugate with alkylating agents. Mater. Plast. 4 5(2): 149–153.
125 Zhou, X., Luo, W., Nie, H., Xu, L., Hu, R., Zhao, Z. et al. (2017). Oligo(maleic anhydride)s: a
platform for unveiling the mechanism of clusteroluminescence of non-­aromatic polymers.
J. Mater. Chem. C. 5 (19): 4775–4779.
126 Wang, R.B., Yuan, W.Z., Zhu, X.Y. (2015). Aggregation-­induced emission of non-­conjugated
poly(amido amine)s: discovering, luminescent mechanism understanding and bioapplication.
Chinese J. Polym. Sci. 33 (5): 680–687.
127 Liu, B., Zhang, H., Liu, S., Sun, J., Zhang, X., Tang, B.Z. (2020). Polymerization-­induced emission.
Mater. Horiz. 7 (4):987–998.
128 Chen, X., Luo, W., Ma, H., Peng, Q., Yuan, W.Z., Zhang, Y. (2018). Prevalent intrinsic
emission from nonaromatic amino acids and poly(amino acids). Sci. China Chem. 61 (3):
351–359.
129 Han, T., Deng, H.Q., Qiu, Z.J., Zhao, Z., Zhang, H.K., Zou, H. et al. (2018). Facile multicomponent
polymerizations toward unconventional luminescent polymers with readily openable small
heterocycles. J. Am. Chem. Soc. 140 (16): 5588–5598.
130 Tu, Y., Liu, J., Zhang, H., Peng, Q., Lam, J.W.Y., Tang, B.Z. (2019). Restriction of access to the dark
state: a new mechanistic model for heteroatom-­containing AIE systems. Angew. Chem. Int. Ed. 58
(42): 14911–14914.
131 Zhu, Q., Zhang, Y., Nie, H., Zhao, Z., Liu, S., Wong, K.S. et al. (2015). Insight into the strong
aggregation-­induced emission of low-­conjugated racemic C6-­unsubstituted
tetrahydropyrimidines through crystal-­structure–property relationship of polymorphs. Chem. Sci.
6 (8): 4690–4697.
132 Luo, Z., Yuan, X., Yu, Y., Zhang, Q., Leong, D.T., Lee, J.Y. et al. (2012). From aggregation-­induced
emission of Au(I)-­thiolate complexes to ultrabright Au(0)@Au(I)-­thiolate core-­shell nanoclusters.
J. Am. Chem. Soc. 134 (40): 16662–16670.
133 Zhao, Z., Zhang, H., Lam, J.W.Y., Tang, B.Z. (2020). Aggregation-­induced emission: new vistas at
aggregate level. Angew. Chem. Int. Ed. doi: 1002/anie.201916729.
134 Cai, Y., Du, L., Samedov, K., Gu, X., Qi, F., Sung, H.H.Y. et al. (2018). Deciphering the working
mechanism of aggregation-­induced emission of tetraphenylethylene derivatives by ultrafast
spectroscopy. Chem. Sci. 9 (20): 4662–4670.
177

Crystallization-­induced Emission Enhancement


Yong Qiang Dong, Yingying Liu, Mengyang Liu, Qian Wang, and Kang Wang
Beijing Key Laboratory of Energy Conversion and Storage Materials, College of Chemistry, Beijing Normal University, Beijing, China

6.1 ­Introduction

Organic luminogens have attracted great attention due to their applications in organic light-­
emitting diodes (OLEDs), data storage, photoswitches, bio-­imaging, and sensors. Of particular
interest are those luminogens exhibiting high performance in solid state, as luminogens are mostly
used in solid states in real application.
However, many traditional luminogens are highly emissive in solution, but weakly or nearly
nonemissive in the solid state because of the notorious aggregation-­caused quenching (ACQ) [1–4].
ACQ effect has severely restrained the practical applications of organic luminogens as they are
often used in solid state rather than discrete states. In contrast, Tang and coworkers found that
some organic luminogens exhibit much enhanced emission from solution to aggregation, which is
exactly opposite to ACQ phenomenon. Tang coined it as aggregation-­induced emission (AIE), and
then many AIE active luminogens (AIEgen) have been developed and the mechanism of AIE has
been attributed to restriction of intramolecular motion (RIM) [5–7].
Crystallization of luminogens normally weakens and red-­shifts their luminescence which fur-
ther deteriorates the performance of luminogens as molecules of luminogens pack more tightly in
crystalline state. Scientists have tried to fabricate stable amorphous films through various chemical
modification and physical doping strategies, although crystalline films are known to exhibit higher
charge mobilities and are beneficial for organic light-­emitting transistors and organic laser [8–10].
However, amorphous film is in a metastable state and may transform to crystalline one slowly due
to the Joule heat generated in an electrical device, hence breakdown of the device [11, 12].
Instead of fighting against the crystallization, Tang’s group found that the PL intensity of amor-
phous film of hexaphenylsilole (HPS) increased by ~2.5-­fold with the PL spectrum blue-­shifted to
462 nm upon fuming with solvent vapor, which is caused by the transformation of the HPS film
from amorphous to crystalline [13, 14]. With the development of AIEgens, they found that many
of them exhibit similar effect, and they coined this phenomenon as crystallization-­induced emis-
sion enhancement (CIEE) [15]. The CIEE phenomenon is a wonderful effect: (i) it is different from
traditional luminogens, and demonstratration of CIEE process may widen and deepen under-
standing of photophysical processes and improve the optical theory; (ii) knowledge gained from

Handbook of Aggregation-Induced Emission: Volume 1 Tutorial Lectures and Mechanism Studies, First Edition.
Edited by Youhong Tang and Ben Zhong Tang.
© 2022 John Wiley & Sons Ltd. Published 2022 by John Wiley & Sons Ltd.
178 6  Crystallization-­induced Emission Enhancement

study of CIEE effect may generate many high-­performance luminophores, such as crystals with
high quantum yield, luminogens exhibit high contrast emission switching through reversible tran-
sition between amorphous and crystalline state which may find application in sensor and optical
data storage and so on. As many CIEE active luminogens have been reported, it is necessary to give
a summary and generate some idea of high-­performance luminogens and disclose the photophysi-
cal process of the CIEE process, which will be carried out in this chapter.

6.2 ­Tetraphenylethylene Derivatives

Tetraphenylethylene (TPE) is a star AIEgen as many excellent AIEgens have been developed from
TPE core as functional TPE derivatives can be readily synthesized. We select some TPE derivatives
list in Chart 6.1.
Emissions of amorphous solids of TPE1 and TPE2 peaked at 495 and 505 nm with quantum yield
(ΦF) of 12.7 and 10.2%, respectively (Figure  6.1e, j). Crystals of TPE1 (Figure  6.1a) and TPE2
(Figure 6.1f), obtained through slow evaporation of their solutions, emit blue lights peaked at 455
and 444 nm, respectively, with ΦF of 34.45 and 28.58%. It is clear that both luminogens are CIEE
active [16].
TPE3 and TPE4 also are also CIEE active. Both of them can form deep-­blue emissive crystals
(448 nm with quantum yield of 56% for TPE3, 446 nm with ΦF of 54% for TPE4) and sky-­blue emis-
sive crystals (462 nm with ΦF of 51% for TPE3, 460 nm with ΦF of 48% for TPE4). Amorphous

O O O O

O O O O
TPE1 TPE2 TPE3 TPE4

O O N N N N N N

F F N N
TPE5 TPE6 TPE7 TPE8
O
N HN N

HO OH

N
HN
NH
N

HO OH
N NH
TPE9 TPE10

C
C

X N

C C
C C

TPE11 TPE12 TPE13

Chart 6.1  Structure of TPE derivatives.


6.2 ­Tetraphenylethylene Derivative 179

(a) (e) 455 nm a


II b
I
c
I d
(c)
(b)
I: quenching melt
II: 95 °C, seven minutes
III: fuming with III
acetone, five minutes
I 495 nm

PL intensity (a.u.)
(d)
390 460 530 600
(f) (h) (j) 444 nm f
IV g
I
h
I i

(g)
(i)
IV: 150 °C, seven minutes
V: fuming with
V
CHCI3, seven minutes

I
505 nm

390 460 530 600


Wavelength (nm)

Figure 6.1  Images of TPE1 (a)–(d) and TPE2 (f)–(i): (a), (f) pristine crystal; (b), (g) amorphous solids; (c), (h)
thermal-­treated and (d), (i) solvent-­fumed amorphous solids; PL spectra of samples of TPE1 (e); and TPE2 (j)
in the images. Excitation wavelength: 350 nm. Photographs were taken under UV illumination. Source:
Adapted with permission from Ref. [16] under the terms of the Creative Commons Attribution License.

solids of TPE3 and TPE4 obtained through quenching of their melt emit green light with ΦF of 47
and 44%, which is slightly lower than their crystalline cousins, indicating that TPE3 and TPE4 are
CIEE active [17]. Although many TPE-­cored CIEE luminogens are reported, the emission contrast
between crystalline and amorphous states is still low. The high CIEE contrast can be achieved
through increasing the ΦF of crystalline state or decreasing that of amorphous state. The ΦF of
amorphous state can be decreased through construction of a looser molecular packing, for exam-
ple, two luminogens may form amorphous state with different molecular packing patterns and
different disorder degree that is affected by the molecular structure.
Dong and coworkers constructed a TPE5  with less symmetry  [18]. The luminogen can form
deep-­blue (420 nm ΦF = 50.4%, TPE5CA), sky-­blue (TPE5CB, 460 nm ΦF = 8.3%) emissive crystals,
and green emissive amorphous solid (TPE5Am, 496 nm ΦF = 7.6%) (Figure 6.2). The ΦF of TPE5CA
is about six times higher than that of TPE5CB and TPE5Am. TPE5 exhibit CIEE effect with high
contrast in both emission color and efficiency.
Normally, it is impossible to compare the molecular conformations and packing patterns of a
luminogen in crystalline and amorphous state, as the molecules in amorphous solid pack ­randomly.
Two crystalline states with sharp contrast in both emission color and ΦF facilitate the disclosure of
the effect of molecular packing pattern and conformation on their optical properties, which may
be extended to the disclosure of CIEE mechanism. Bond length alternation (BLA) was used to
180 6  Crystallization-­induced Emission Enhancement

(a) (d) a

Normalized PL intensity (a.u.)


b
c

(b)

(c)

370 440 510 580 650


Wavelength (nm)

Figure 6.2  Photos of (a) TPE5CA, (b) TPE5CB, (c) TPE5Am under 365 nm UV illumination; (d) normalized PL
spectra. Excitation wavelength: 350 nm; exposure time: 1/25 seconds. Source: Reproduced by permission of
Ref. [18]. The Royal Society of Chemistry.

estimate the conjugation difference in TPECA and TPECB based on the exact molecular conforma-
tion in the crystals. BLA was calculated to be 0.145 47 for TPECA and 0.139 24 for TPECB. The
smaller BLA value for TPECB suggests the better molecular coplanarity and conjugation, which
coincide with its emission, thus, the red-­shifted PL spectrum of amorphous solid may also be
induced by the more planar conformation.
In addition to the remarkable contrast in emission color among different morphologies of TPE5,
the three morphologies of TPE5 also exhibit high contrast in ΦF. The calculated density of TPE5CA
(1.185 g/cm3) is higher than that of TPE5CB (1.127 g/cm3), suggesting tighter molecules packing
in TPE5CA. The tighter-­packing pattern may further hinder the rotation and vibration of the phe-
nyl rings, which will block the nonradiative transition and induce higher ΦF.
Similar to other TPE derivatives, there are no strong intermolecular interactions, such as π–π
interaction or H/J-­aggregation in both TPE5CA and TPE5CB due to the twisted molecular confor-
mation. As shown in Figure 6.3, in TPE5CA, there exist four C−H⋯O interactions with identical
bond length of 2.56 Å, 14 C−H⋯π interactions with the phenyl rings of nearby molecules measur-
ing 2.978, 3.435, 3.447, and 3.474 Å (Figure 6.3). Meanwhile, there are two intramolecular C−H⋯π
interactions with a bond length of 3.442 Å. Hence, so many intermolecular and intramolecular
weak interactions highly hinder the rotation and vibration of the phenyl rings in TPE5CA, which
will block the nonradiative pathway, resulting in high ΦF.
Differently, there were no C−H⋯π interactions found in TPE5CB. Only four C−H⋯O interac-
tions exist in butoxy groups of adjacent molecules with bond lengths of 2.595, 2.711, 2.814, and
2.939 Å (Figure 6.4). That is, in the TPE5CB, every phenyl ring of the molecule can rotate in some
degree, which vastly consumes the energy of excited state and leads to low ΦF. After destroying the
crystal structure, TPE5Am exhibits lower ΦF of 7.6%, which may be ascribed to loose-­packing pat-
tern and less interactions in the amorphous solid.
Through comparison of TPE4 and TPE5, it is clearer that luminogen with lower symmetry
exhibits high contrast CIEE effect due to the lower ΦF caused by the looser molecular packing of
amorphous solid. In addition to TPE4 and TPE5, other examples have also been reported. TPE6
and TPE7 are TPE derivatives with low symmetry and both of them are nearly nonemissive in
amorphous state (on TLC plates, freshly prepared suspension, or ground powder), but both of
them emit intense light of cyan color with high ΦF of 44.8 and 100% in crystalline state,
6.2 ­Tetraphenylethylene Derivative 181

(a)

(b) (c)

C
H
O

Figure 6.3  (a) View of C−H⋯O (green dashed line) and (b) C−H⋯π (red dashed line) between molecules and
(c) intramolecular C−H⋯π (dark-­red dashed line) interactions in single crystal of TPE5CA. The blue dots refer
to the center of benzene rings. Source: Reproduced by permission of Ref. [18]. The Royal Society of Chemistry.

C
H
O

Figure 6.4  View of C−H⋯O (green dashed line) interactions between molecules in single crystal of
TPE5CB. Source: Reproduced by permission of Ref. [18]. The Royal Society of Chemistry.
182 6  Crystallization-­induced Emission Enhancement

respectively [19]. However, the similar derivative, TPE8, with high symmetry, is AIE active and
exhibits low efficiency contrast between amorphous and crystalline state [20]. Thus, it is clear that
TPE derivatives with low symmetry may afford high-­contrast CIEE effect.
Chi and coworkers developed a TPE9 with AIE and CIEE effect. Crystals of TPE9 emit intense
blue light with ΦF of 85%, while upon grinding, emission of TPE9 red-­shifted to 513 nm with ΦF
decreased to 66%. Intermolecular interactions of C–H⋯π, C–H⋯N, and C–H⋯O in the crystals of
TPE9 help to solidify the conformation of molecules and restrain intramolecular motion, hence
block the nonradiative process. Thus, TPE9 is CIEE active  [21]. Wang synthesized tetra(3-­1H-­
benzo[d]imidazol-­2-­yl)-­4-­hydroxyphenyl)ethene (TPE10) which is also CIEE active. The powder
of TPE10 is nearly nonemissive with ΦF of 0.25%, while its crystals emit bright-­green light with ΦF
of 5.6% [22].
Lain and coworkers synthesized two o-­carborane–tetraphenylethenes which isare CIEE active.
Crystals of TPE11 exhibited a strong green emission with emission that peaked at 522 nm with ΦF
of 95%, owing to the intrinsic TICT transition from the TPE donor to the C1–C2 bond of the o-­
carborane acceptor. The emission of the crystals was red-­shifted to by 25 nm with a slight decrease
of ΦF from 95 to 88% upon grinding due to the transformation from crystalline to amorphous solid.
Crystalline and amorphous solids for TPE12 emit light peaked at 534 and 528 nm with ΦF of 90 and
85%, respectively [23].
Liu and coworkers functionalized spiropyran with TPE unit [24] and obtained luminogen TPE13
exhibiting high contrast CIEE effect. Microcrystals of TPE13 obtained by recrystallization exhibit
orange emission (601 nm, ΦF = 10.17%), while the amorphous solid of TPE13 is nearly nonemis-
sive (678 nm, ΦF < 0.01%) (Figure 6.5). The HOMO of TPE13 distributes on TPE unit, whereas the
LUMO locates on the indolium moiety, thus, TPE moiety as an electron donor and the orange
emission of crystals of TPE13 is from the ICT state in the protonated merocyanine form. Although,
the single crystal of TPE13 was not obtained, the authors found that the optimized structure of
TPE13 is nearly a planar conjugated system through the theoretical calculation. They argue that
molecules are interlocked by weak π–π interactions in the crystal structure of TPE13. The nonra-
diation transition is blocked and the π–π stacking is not strong enough to quench the fluorescence
of ICT. If the crystal is crushed, the molecules will be more planar in conformation, and the inter-
molecular distances will be smaller, which results in the enhancement of π–π interactions and

CROF
AROF
450
Intensity

300

Ground
150
EtOAc
0
400 500 600 700 800
Wavelength (nm)

Figure 6.5  Fluorescent images and PL spectra of TPE13 converting reversibly between crystal and
amorphous in the solid states. Source: Adapted with permission Ref. [24]. Copyright 2016, John Wiley
and Sons.
6.3 ­CIEE Active Luminogens with Bulky Conjugation Cor 183

dipole–dipole interactions, so a bathochromic effect and quenching of the fluorescence appear.


The red-­shift of the absorption spectra for the crystals (>100 nm) after grinding is consistent with
the deduction that a more planar conformation and enhancement of the π–π interactions are
induced. The authors attribute the fluorescence quenching in the amorphous states of TPE13 to
the change of π–π interaction and dipolar interaction in TPE13 system.

6.3  ­CIEE Active Luminogens with Bulky Conjugation Core

A series of luminogens constructed with bulky conjugation core and peripheral aryl groups exhibit
high-­contrast CIEE effect. Here we select some of them and try to disclose the mechanism behind
the phenomenon.

6.3.1  Dibenzofulvene (DBF) Derivatives (Chart 6.2)


A series of DBFs have been found to exhibit distinct CIEE effect. Tang and coworkers found that
when large amount of water is poured into acetonitrile solution of DBF1, the luminogen aggre-
gated promptly in the mixture as water is nonsolvent. The freshly prepared suspension is nearly
nonemissive, though the aggregates are already large enough to be seen even with naked eyes.
However, after about one hour, the emission of the aggregates is turned on. The aggregates in
freshly prepared suspension are amorphous, while they transform to crystalline aggregates with
elapse of time. Though pure amorphous solid of DBF1 cannot be obtained through quenching of
its melt, as it crystallizes very quickly, it is clear that DBF1 is CIEE active.
As amorphous solid of DBF1 cannot be obtained directly, it is difficult to study its optical proper-
ties and its application may be restricted. Therefore, Tang and coworkers attached a phenyl ring to
DBF1 and obtained DBF2 with lower symmetry. DBF2 emits faintly in amorphous state (550 nm,
ФF = 0.5%) with but strongly in the crystalline phase (450 nm, ФF = 16%), with a difference in fluo-
rescence ΦF as high as 32 times [15].

DBF1 DBF2 DBF3 DBF4

O O O O O O

DBF5 DBF6 DBF7

Chart 6.2  Structure of DBF derivatives.


184 6  Crystallization-­induced Emission Enhancement

(e) a
(a) (b)
b
c
d

Intensity (a.u.)
(c) (d)

400 460 520 580 640


Wavelength (nm)

Figure 6.6  Photos of (a) DBF4CO, (b) DBF4Am, (c) DBF4CY, (d) DBF4CB, (e) PL spectra (excitation
wavelength: 360 nm). Photos were taken under UV illumination. Source: Adapted with permission Ref. [26].
Copyright 2019, John Wiley and Sons.

In addition to DBF2, many DBFs have been synthesized and their CIEE effects were investi-
gated. Through introducing one or two methyl group to DBF1, DBF3, and DBF4 are obtained. Both
of them are CIEE active. DBF3 can form crystals with sky-­blue emission (DBF3CA, 465  nm,
ΦF = 13.9%) or blue–green emission (DBF3CB, 465 nm, ΦF = 14.6%), while the amorphous solid
of DBF3 emit faintly yellow-­greenish light (DBF3Am, 522 nm, ΦF = 2.1%) [25].
Crystals of DBF4  with blue (DBF4C1, 461  nm, ΦF  =  28.1%), yellow (DBF4C2, 545  nm,
ΦF = 23.3%), and orange (DBF4C3, 586 nm, ΦF = 16.2%) emission can be obtained through slow
evaporation of solution. The amorphous solid of DBF4 exhibit orange emission with low ΦF of
2.9% (Figure 6.6) [26].
The high-­contrast emission color of the three crystals of DBF4 is attributed to the molecular
conformation difference. Amorphous solid of many CIEE active luminogens exhibit redder emis-
sion in comparison with their crystalline phases, while emission of DBF4Am falls between two
crystalline phases of DBF4C2 and DBF4C3. Thus, the average torsion angle of molecules in
DBF4Am may be between those of DBF4C2 and DBF4C3.
The available of three single crystals with high-­contrast emission of one luminogen will facili-
tate the disclosure of the CIEE mechanism. The asymmetric units of DBF4CB, DBF4CY, and
DBF4CO contain one, three, and six molecules, respectively, of DBF4, which is marked in blue
color in the crystal structure image (Figure 6.6). In the asymmetric unit of DBF4CB, the conforma-
tion of one molecule was solidified by eight C−H⋯π interactions. However, in the asymmetric
units of DBF4CY and DBF4CO, the numbers of C−H⋯π interaction is 12 for 3 molecules and 15
for 6 molecules, respectively (Figure 6.7). Then, it is clear that the emission efficiencies of different
crystalline phases depend on the average number of C−H⋯π interactions per molecule. More
interactions help to solidify the molecular conformation more tightly and block the nonradioactive
pathways, and then increase the ΦF. Compared with the crystals of DBF4, the ΦF of DBF4Am is
lowered to 2.9%, which may be attributed to the loose packing pattern and less rotation restriction
6.3 ­CIEE Active Luminogens with Bulky Conjugation Cor 185

(a) (c)

(b)

Figure 6.7  Analysis of the C–H⋯π interactions (green line) in single crystal structures of (a) DBF4CB,
(b) DBF4CY, and (c) DBF4CO. Source: Reproduced with permission Ref. [26]. Copyright 2016, John Wiley
and Sons.

in amorphous solid. Following this line, high ΦF luminogens may be generated through introduc-
tion of groups exhibiting strong interactions to the twisted conjugation core.
Similar to other DBFs, DBF5-­7 also exhibit high-­contrast CIEE [27, 28]. For example, two crys-
tals of DBF5 are blue (466 nm) and green (518 nm) emissive with ΦF of 95 and 62%, respectively,
and the higher ΦF of the blue crystal is attributed to the stronger intermolecular interactions. The
amorphous solid of DBF5 emit orange light faintly with low ΦF of 5%.

6.3.2  9-­([1,1′-­Biphenyl]-­4-­ylphenylmethylene)-­9H-­xanthene
To develop more CIEE luminogens with high contrast, Dong and coworkers compared the CIEE
efficiency of TPE1 and DBF2. They found that the cores of TPE1 and DBF2 are a double bond and
dibenzofulvene, while the CIEE contrast (ΦC/ΦAm) of TPE1 and DBF2 are 2.7 and 32, respectively.
That is, tiny differences in the chemical structure greatly influence the CIEE effect of some lumi-
nogens, and combination of the bulky conjugation core and peripheral phenyl rings may afford
luminogens with high-­contrast CIEE effect. Therefore, they replaced the dibenzofulvene core with
9-­methylenexanthene, both of which are large conjugation core, and get 9-­([1,1′-­biphenyl]-­4-­ylph
enylmethylene)-­9H-­xanthene (BPPX). BPPX can form deep-­blue (432  nm) and green (492  nm)
emissive crystals with ΦF of 42.2 and 59.3%, respectively (Figure  6.8). Similar to DBF2, BPPX
exhibits high-­contrast CIEE effect with value of ΦC/ΦAm up to about 148. As aforementioned, dif-
ferent luminogen may form different amorphous solid with varied molecular packing density. For
TPE2, when forming amorphous solid, the phenyl rings may adjust the torsion angle in response
the microenvironment, and forming denser molecular packing. However, for DBF2 and BPPX, the
bulky core cannot rotate as the phenyl rings in TPE1, and may form looser amorphous solid, in
which the phenyl rings can still rotate in some degree and quench the emission. Thus, the combi-
nation of bulky core and peripheral phenyl rings may be a possible design strategy for high contrast
CIEE luminogens [29].
186 6  Crystallization-­induced Emission Enhancement

(d)
(c) 432 492 584 a
b

Norimalized PL intensity
c

I I
II III
(a) (b)
O IV
BPPX
380 460 540 620 700
Wavelength (nm)
Figure 6.8  Photos of luminogen BPPX in different aggregate states: (a) BPPXBC, (b) BPPXGC, and
(c) BPPXAm. (d) Normalized PL spectra of samples show in panel (excitation wavelength: 360 nm). Photos
were taken under UV illumination. Conditions: (i) heating to melt and quickly cooling; (ii) fuming with
methanol vapor, six hours; and (iii) fuming with acetone vapor, five hours. Source: Adapted by permission of
Ref. [29]. The Royal Society of Chemistry.

N
N N N

N N

N N
N N N N N N

DCNAC1 DCNAC2 DCNAC3 DCNAC4

Chart 6.3  Molecular structures of the DCNAC derivatives.

6.3.3  Dicyanomethylenated Acridones


A series dicyanomethylenated acridones (DCNAC) exhibiting high-­contrast CIEE effect were
reported by Zhang and coworkers (Chart 6.3) [30]. All the DCNAC derivatives are nonemissive in
solution or in amorphous state obtained by grinding the crystals, however, their crystals exhibit
intense luminescence (Figure 6.9) with different emission colors. The redder emission of DCNAC1C
(647 nm, ΦF = 5%) and DCNAC4C (707 nm, ΦF = 16%) are caused by excimer and ICT process,
respectively. The π-­π stacking and excimeric coupling are significantly weakened for luminogens
with longer alkyl chains, resulting in the blue-­shifted emission and enhanced ΦF of DCNAC2C
(572 nm, ΦF = 30%) and DCNAC3C (562 nm, ΦF = 41%) crystals.
To disclose the origin of the high-­contrast CIEE, the author established a three-­dimensional
molecular arrangement and interaction image of amorphous solid of DCNAC1 by using molecular
mechanics force field by fitting quantum mechanical calculations. The theoretically simulated
structure demonstrated that DCNAC1  molecules adopt a random-­orientation packing feature
without hydrogen bonding and π–π stacking in the amorphous phase. These intermolecular inter-
actions play a key role in restricting the conformational transformation of the molecules. Therefore,
the nonemissive properties of the amorphous solid are attributed to the excited-­state quenching
induced by torsional vibrations. While in crystal lattices, these torsional vibrations can be restricted
by multiple intermolecular interactions and thus the emission is turned on.
6.3 ­CIEE Active Luminogens with Bulky Conjugation Cor 187

DCNAC1 DCNAC2 DCNAC3 DCNAC4

(a)

(b)

(c)

Grinding

Heating or
vapor fuming

Figure 6.9  Fluorescent photographs of the DCNAC derivatives (a) in solution (b) and in the crystalline
state under 365 nm light irradiation. (c) Fluorescent images of the pristine crystalline, ground, and ground/
heated solids under 365 nm light irradiation. Source: Adapted with permission from Ref. [30]. Copyright
2016, American Chemical Society.

6.3.4  Bis(diarylmethylene)dihydroanthracene [31]
Dong and coworkers, using the design three derivatives of bis(diphenyl-­methylene)-­
dihydroanthracene (R-­DHA), 9,10-­bis-­(diphenylmethylene)-­9,10-­dihydroanthracene (PDHA),
9,10-­bis(di-­4′-­methoxylphenylmethylene)-­9,10-­dihydro-­anthracene (ADHA), and 9,10-­bis(di-­4′-­m
ethylphenylmethylene)-­9,10-­dihydro-­anthracene (TDHA) (Chart 6.4).
Crystals of PDHA emits blue light at 446 nm with ΦF of 15.1%, and its amorphous solid cannot
be obtained due to its fast crystallization ability. However, ADHA, exhibits a low ΦF of ∼2−3% in
188 6  Crystallization-­induced Emission Enhancement

O O

O O
PDHA ADHA TDHA

Chart 6.4  Molecular structures of the R-­DHA derivatives.

(a) (b)
TDHA-am TDHA-b TDHA-g TDHA-am

Normalized intensity
I I
II III
TDHA-b TDHA-g

III

II

380 455 530 605 680


Wavelength (nm)
Figure 6.10  (a) Photos and their transformation diagram scheme of TDHA-­b, TDHA-­g, and TDHA-­am. (b) PL
spectra of TDHA-­b, TDHA-­g, and TDHA-­am states. (Photos were taken under UV illumination. Conditions:
(i) heating to melt and cooling by liquid nitrogen, (ii) heating at 170 °C for more than one hour or fuming
with CH2Cl2 or CHCl3 vapor for three hours; (iii) fuming with acetone vapor for three hours.) Source: Adapted
with permission from Ref. [31]. Copyright 2015, American Chemical Society.

both crystalline and amorphous states. Indeed, after removal of the solvent molecules by heating
at 100 °C under vacuum for 12 hours, crystals of ADHA becomes more emissive, with a higher ΦF
of 26.2%.
Two types of TDHA crystals (TDHA-­b and TDHA-­g) were obtained from its n-­hexane/CHCl3
mixture with blue (425 nm, ΦF = 27.7%) and sky-­blue (464 nm and 76.8%) emission, respectively.
Compared to their crystalline counterparts, the amorphous powder of TDHA is weakly green
emissive with PL spectrum peaked at 520 nm and a lower ΦF of 3.3%. Thus, it is clear that both
ADHA and TDHA exhibit high-­contrast CIEE effect (Figure 6.10).
It is believed that luminogenic molecules are poorly packed in the amorphous state. This may
generate many free spaces for the phenyl rings and dihydroanthracene blade to rotate and vibrate,
6.3 ­CIEE Active Luminogens with Bulky Conjugation Cor 189

thus leading to partial emission quenching. On the other hand, because of the free of constraint in
the crystal lattice, the molecules in the amorphous powders may assume a more planar conforma-
tion and, hence, show a redder emission. It is noteworthy that the ΦF of ADHA-­am and TDHA-­am
are much lower than those of amorphous powders of TPE derivatives (∼40%), possibly due to their
comparative poorer molecular packing, which is caused by their butterfly-­like shapes.
To exclude the effect of molecular structure and size, the author examined crystal structure of
TDHA-­b and TDHA-­g, and found that the average volume occupied by a single luminogen mole-
cule in the crystal lattice of TDHA-­b and TDHA-­g is much larger that its molecular size (805.6 Å3
for TDHA-­b and 778.3 Å3 for TDHA-­g) (Figure 6.11). This suggests a loosely molecular packing,
which should enable the molecules to have some sort of motion. However, multiple CH⋯HC,

(a) a
b

CH···HC
C···C
CH···C

(b)
b

Figure 6.11  (a) Molecular packing of crystal TDHA-­g. (b) Molecular packing of crystal TDHA-­b as viewed
along the x-­axis of the unit cell. Carbon and hydrogen atoms are shown in gray and white, respectively.
Source: Adapted with permission from Ref. [31]. Copyright 2015, American Chemical Society.
190 6  Crystallization-­induced Emission Enhancement

CH⋯C, and C⋯C interactions are found throughout the crystal structures. On the other hand, the
smaller occupied space and, hence, higher packing density in TDHA-­g than in TDHA-­b should
help to further rigidify the molecular conformation and block the nonradiative pathways, thus
resulting in a much higher emission efficiency. And the low ΦF of amorphous solid may also be
attributed to the loose packing pattern, in which the molecules may still have some motions.
As aforementioned, in addition to the DBF derivatives, DHA, DCNAC, and BPPX derivatives
with large conjugation core do exhibit high-­contrast CIEE effect. The large conjugation core may
facilitate looser packing patterns in which some groups such as phenyl rings may still have enough
free space to rotate in some degree, hence the quenching of emission. Thus, the large conjugation
core with peripheral phenyl rings may be a possible design strategy for high contrast CIEE
luminogens.

6.4  ­Other High-­contrast CIEE Luminogens

6.4.1  4-­Dimethylamino-­2-­Benzylidene Malonic Acid Dimethyl Ester


Cariati and coworkers synthesized 4-­dimethylamino-­2-­benzylidene malonic acid dimethyl ester
(DMAME) with CIEE behavior (Figure 6.12) [32]. Amorphous films of DMAME exhibit very faint
greenish-­blue luminescence with ΦF of 1% at room temperature, while its crystalline powder is
highly emissive and peaked at 468 nm with a ΦF of 38.0%. The weak emission of DMAME in the
amorphous phase is attributed to the nonradiative relaxation processes favored by molecular tor-
sional mobility as molecules pack loosely in amorphous solid. When rotation around the aryl main
axis is frozen by lowering the temperature, the intensity of the emission strongly increases and is
quite broad since different molecular conformations and local environments contribute to the
overall emission. The enormous intra-­and intermolecular short distance contacts (2.484−3.106 Å)
in the crystals of DMAME stiffen the molecular conformations and meanwhile activate the RIM
process to turn on the emission.

Crystalline phase
Amorphous
1 Crystalline
O
O O
PL intensity (a.u.)

N O

DMAME O
O O

N O
0
440 480 520 560 600
Amorphous phase Wavelength (nm)

Figure 6.12  Photo and PL spectra of DMAME film in the crystalline and amorphous phases excited at
350 nm, the spectra are normalized to the crystalline film emission. Source: Adapted from Ref. [32] with
permission from the PCCP Owner Societies.
6.4 ­Other High-­contrast CIEE Luminogen 191

6.4.2  Diphenyl Maleimide Derivatives [33]


A group of BADPMA derivatives exhibit high-­contrast CIEE effect (Figure 6.13). Their amorphous
films are nearly nonemissive with ΦF lower than 0.1%, while their crystals are highly green emis-
sive with ΦF of 28–80%. The single crystal of BADPMA4 suggests that there are no specific strong
intermolecular interactions (such as π-­π stacking or H/J-­aggregates) due to their twisted
conformations.
The twisted conformation molecules are arranged into molecular columns through weak
N–H⋯O (2.08  Å) interactions and the molecular columns are held together by weak C–H⋯π
(3.44  Å) and C–H⋯O (2.53 and 2.59  Å) interactions, form a 2D and 3D structure. These weak
intermolecular interactions fix the molecular conformation of BADPMA4 in the crystals, and the
rigidification of the twisted conformation of BADPMA4 not only prevents formation of intramo-
lecular between N–H and C=O group but also inhibits the internal rotations and blocks the nonra-
dioactive relaxation, which contributes to the high ΦF. For amorphous solid, through theoretical
calculation, the authors suggest that the benzamide group of the molecules in the excited states
has rotated drastically, accompanied by the formation of an intramolecular hydrogen bond between
N–H and C=O, which quenches the emission.

(a) ΦF of film ΦF of crystals


O
1:R=H 0.1% 28%
R
O N O 2:R=OMe 0.1% 37%

3:R=NO2 0.1% 39%

BADPMA 4:R=Br 0.1% 53%, 80% for single crystal

1 2 3 4
(b)

(c)

Figure 6.13  (a) Molecular structures and ΦF of the benzamide-­based diphenyl maleimide (BADPMA)
derivatives, photos of (b) amorphous film and (c) crystalline powder of the BADPMA derivatives under UV
light. Source: Adapted from Ref. [33] with permission from The Royal Society of Chemistry.
192 6  Crystallization-­induced Emission Enhancement

6.4.3  3,4-­Bisthienylmaleic Anhydride [34]

ΦF
O O O Solution 59%

S S Amorphous 8%

Crystal 85%
BTMA

BTMA is strongly emissive in both solution and crystalline state, peaking at 549 and 522 nm with
high ΦF of 59 and 85%, respectively. However, the amorphous solid of BTMA obtained through
grinding crystals is weakly emissive, peaking at 624 nm with ΦF of 8%. That is, BTMA is CIEE
active. BTMA is different from other CIEE luminogen, as it is highly emissive in solution. The
optimized conformation of BTMA shows small twist angle between thiophene and the central ring
that reduces the distance between the S atom and the adjacent O atom to 2.91 Å, which causes the
formation of weak S⋯O interaction. The interaction can enhance molecular rigidity and suppress
the nonradiative relaxation, resulting in the high ΦF of BTMA in solution. Due to the twisted
molecular conformation of BTMA, there is no strong intermolecular packing affecting emission of
its crystals, such as π-­π stacking or H/J-­aggregation. The high ΦF of BTMA crystals is attributed to
the weak intermolecular interactions, S⋯S (3.51 Å), and partly overlapping π-­π (3.39 Å) short con-
tact, which fix the conformation and suppress the nonradiative relaxation. The information of
molecular conformation or packing pattern of BTMA in amorphous solid is difficult to be obtained
experimentally or through theoretical calculation. The greatly red-­shifted and weakened emission
of amorphous solid of BTMA may be due to the ICT process.
Wang and coworkers designed and synthesized five cyanostilbene derivatives (CSCzs) with
D–π–A structure containing carbazole as donor (Chart 6.5) [35]. All the five luminogens are CIEE
active, while each of them shows different ΦF contrast. Luminogens CSCz1 and CSCz5 exhibit
high contrast. Weak interactions such as C–H⋯C, C–H⋯N, C–H⋯O, and C–H⋯Cl are found
between molecules in the crystals of the five luminogens. The conformations of molecules are
locked by the weak interaction networks, and the nonradiative relaxation channels are blocked.
While in amorphous state, some weak interactions disappear, the emission of luminogens are
partly quenched.

ΦF of Am ΦF of crystals
R 1: R = H 3.5% 77%

2: R = OMe 14.1% 57.9%

N 3: R = NO2 9.5% 57.8%


N
4: R = Me 6.1% 25.7%

CSCz 5: R = Cl 3% 73.6%

Chart 6.5  Molecular structures and ΦF of the CSCz derivatives.


6.4 ­Other High-­contrast CIEE Luminogen 193

6.4.4  Boron-­containing CIEE Luminogens


Organoboron complexes are one of the important classes of luminogen. Most of them are emissive
in dilute solution, but exhibit quenched or reduced fluorescence intensity in the solid state.
However, some systems exhibiting AIE and CIEE activity have been developed.
Chujo and coworkers synthesized a series of boron diiminate derivatives (Chart  6.6) both of
which are AIE and CIEE active [36, 37]. Comparing to boron diketonates and boron ketoiminates,
boron diiminates can be readily functionalized through bonding various substituents to two nitro-
gen atoms, which facilitate the regulation of intermolecular interactions in solid state and develop-
ment of CIEE active luminogens. Crystals of luminogens Bo1-­5 emit bluer light more intensely
than their amorphous cousins, indicating all of them are CIEE active, and Bo3 exhibit the highest
contrast between its crystalline and amorphous solid with approximately 15-­fold increase of ΦF
(Figure 6.14). The authors attributed the weakened emission of amorphous solid of Bo1-­5 to the

R2 R4 Bo1: R1 = H, R2 = H, R3 = Me, R4 = H
F F Bo2: R1 = H, R2 = H, R3 = Ph, R4 = H
B
N N Bo3: R1 = H, R2 = OMe, R3 = Ph, R4 = H
R3 Bo4: R1 = H, R2 = H, R3 = Ph, R4 = NO2
Bo5: R1 = H, R2 = NMe, R3 = Ph, R4 = H
R1

R1 R3 Bo6: R1 = I, R2 = H, R3 = I, R4 = H
F F
Bo7: R1 = I, R2 = I, R3 = H, R4 = H
B
N N Bo8: R1 = H, R2 = I, R3 = I, R4 = H
Bo9: R1 = H, R2 = I, R3 = H, R4 = I

R2 R4

Chart 6.6  Molecular structures of the Bo1-­9.

Figure 6.14  Photographs of boron


diiminates in crystalline and Crystalline sample
amorphous states under UV
irradiation. Source: Adapted with
permission from Ref. [36]. Copyright
2014, American Chemical Society.

Amorphous sample

Bo1 Bo2 Bo3 Bo4 Bo5


194 6  Crystallization-­induced Emission Enhancement

formation of excimers via π−π stacking of phenyl units, which often causes red-­shifts and quench-
ing of emission. In crystals of these boron diiminates, little degrees of π−π interactions were
observed between their phenyl rings. Intermolecular interactions of CH⋯F with distance of
2.31−2.65 Å were found in crystal of these boron diiminates, which are shorter than that of the
sum of van der Waals radii (2.67 Å). Thus, the molecular conformation in the crystal could be sta-
bilized and locked by the multiple CH⋯F hydrogen bonds, hence the blue-­shifted emissions and
the increased ΦF of the crystals of Bo1-­5. Through introduction of iodine to phenyl rings of BO3,
four boron diiminate derivatives Bo6-­9 are obtained [38]. All the four luminogen are CIEE active
with low contrast as iodine groups greatly weaken the emission of crystalline states due to heavy
atom effect and the steric hindrance effect of iodine groups.

H3C C
C C C

C C
C
CBoA C CH3
CBMe

o-­Carborane (C2B10H12) is a polyhedral boron cluster compound which includes two adjacent
carbon atoms in the cluster cage. The applications of carboranes for use in boron neutron capture
therapy and for heat-­resistant materials have been extensively studied because of their high boron
content and thermal and chemical stability. Chujo and coworkers introduced o-­carborane to 9-­and
10-­positions of anthracene and obtained CBoA, changing the anthracene from highly emissive in
solution to a luminogen with AIE and CIEE properties [39]. The emission of CBoA in the dilute
THF solution is weak peaking at 650 nm with ΦF of less than 1%, which is derived from charge
transfer (CT) emission of the di(o-­carboranyl)-­anthracene moiety From the aggregates in mixture
of THF/H2O (v/v 1 : 99), only CT emission (643 nm) was observed at with a slightly higher ΦF of
7%. The most efficient emission was observed from the crystal of CBoA, with a ΦF of 77%, that is,
CBoA is both AIE and CIEE active and may arise from the suppression of C–C bond vibration in
the o-­carborane in the aggregates and crystals. In addition, the emission or crystals of CBoA is
quenched upon grinding due to the transformation from crystalline to amorphous.
In addition, Tanaka, Chujo and coworkers synthesized bis-­o-­carborane-­substituted
1,4-­bis(phenylethynyl)benzene  [40]. Interestingly, CBMe showed high ΦF in the aggregate state
(542 nm, ΦF = 81%), the crystalline state (493 nm, 78%) and even in solution (540 nm, ΦF = 84%).
It is assumed that excitation deactivation induced by the intramolecular motion at the o-­carborane
unit could be restricted by the substituent effect. Emission of crystals of CBMe were weakened and
red-­shifted upon grinding due to the transformation from crystalline to amorphous. And the ΦF of
amorphous solid of CBMe decreased to 0.61, which is ascribed to the π stacking of molecules in the
amorphous solid. That is, CBMe exhibit CIEE effect.

F O
F B
N

N
C8H17 TCB
6.4 ­Other High-­contrast CIEE Luminogen 195

Zhan and coworkers designed and synthesized 6,8-­Di-­tert-­butyl-­2,2-­difluoro-­3-­(9-­octyl-­9H-­carb


azol-­3-­yl)-­2H-­2l4,3l4-­benzo[e] [1,3,2]oxazaborinine (TCB). Amorphous solid of TCB emitted yel-
low–green light with ΦF of 6.4%. In contrast, the crystals of TCB showed intense green light emis-
sion with ΦF of 69.0%. It is clear that TCB exhibited obvious CIEE effect. The tert-­butyl group may
block the close π–π stacking of oxazaborinine groups in crystals which may quench the emission
of luminogen, while affording a looser packing pattern in amorphous solid, which facilitates the
motion of groups and nonradiative transition [41].
Jiang and coworkers synthesized a series of boron-­dipyrromethenes [42] and one of them exhibit
CIEE effect. Amorphous powder of Bodipye-­TPA-­BH-­COOH prepared by quenching the melt in
the freeze (−78 °C) emits quite a faint light with fluorescent ΦF of 1.38%, which is probably due to
the active intramolecular twisting motion in the relatively loosing amorphous phase for this com-
pound (Figure 6.15). The crystals of Bodipy-­TPA-­BH-­COOH exhibit bright red emission with ΦF of
12.3% due to the two aspects: (i) there is few effective intermolecular π-­π stacking interaction
between Bodipy chromophores which may quench the emission of luminogens; (ii) there are mul-
tiple intermolecular C–H⋯π/O/F interactions which restrain the intramolecular twisting motion
between functionalized-­triphenylamine moiety around Bodipy framework in Bodipy-­TPA-­BH-­
COOH and in turn leads to the CIEE effect.

N
B O
N O
F

BoSAF

Tanaka, Chujo, and coworkers reported a multi-­functional boron complex with the loosely-­fused
azomethine structure (BoSAF)  [43]. The amorphous solid, yellow emissive crystal and orange
emissive crystal of BoSAF have different ΦF of 0.3, 39 and 34%, respectively. That is, BoSAF exhibit
high contrast CIEE effect. From the mechanistic and theoretical investigations, it was proposed
that large structural relaxation in the excited state of the boron complex could play a critical role in
emission annihilation in the absence of structural restriction. In addition, the UV absorption of the
amorphous solid indicates the available of π–π interaction in amorphous state, which induces the
energy loss in the excited state through nonradiative decay to the ground state. The protruded fluo-
rine atoms and curved structures of the complex prevent the formation of π–π stacking in the
crystalline states. Moreover, the CH⋯F and CH⋯π intermolecular interactions in the crystals
inhibit the active molecular motions and contribute to increase in intermolecular interaction

N
COOH
N

N N
B
F F Single crystal Amorphous
Bodipy-TPA-BH-COOH form form

Figure 6.15  The emission colors of Bodipy-­TPA-­BH-­COOH in single crystal and amorphous forms under UV
light, respectively. Source: Reprinted Ref. [42]. Copyright 2016, with permission from Elsevier.
196 6  Crystallization-­induced Emission Enhancement

involving π-­electronic systems. As a result, radiative decay should be facilitated, leading to rela-
tively high fluorescence quantum efficiencies.

6.5  ­Potential Applications

6.5.1  Volatile Organic Compounds (VOCs) Sensor


Some VOCs are dangerous to human health or cause harm to the environment. As aforemen-
tioned, organic solvent vapor can activate the dynamic transformation of luminogens from amor-
phous to crystalline. Thus, CIEE compounds may be used as VOCs sensor. The amorphous films of
CIEE active luminogens can be obtained through prompt solvent evaporation. Normally, when the
amorphous films of CIEE luminogens are exposed to VOCs, double signal will be given: increased
PL intensity and blue-­shifted emission color due to the crystallization of film upon fuming. If the
CIEE luminogen exhibit high contrast in ΦF, it may be a turn-­on mode sensor for VOCs.
For example, when a thin film of DBF2 coated on the inner wall of a quartz cell, it is nearly non-
emissive due to the amorphous essence of the film. The emission of the film is turned on upon
exposure to solvent vapor due to transformation from amorphous to crystalline state (Figure 6.16).
This demonstrates that the vapochromism of CIEE compound can be used for VOCs
sensing [15].

6.5.2  OLED
Normally, crystallization of organic molecules in OLED causes the degradation of device, partly
due to the weakening of ΦF and red-­shifted EL spectra induced by crystallization, so organic mate-
rials with higher glass-­transition temperatures have therefore been synthesized to overcome the
problem. However, CIEE compounds behave totally opposite to conventional luminogens and are
expected to exhibit different performance from normal dyes in OLEDs.

60
Crystalline bright state
Relative emission intensity

40

20

0 Amorphous dark state

0 1 2 3

Figure 6.16  Repeated switching between dark and bright states of the emission of the thin solid films of
DBF2 coated on quartz plates by fuming-­heating. Source: Adapted by permission of Ref. [15]. The Royal
Society of Chemistry.
6.5 ­Potential Application 197

Si

HPS2,4

(a) (b)
800 2500 6

2000
600

Current efficiency (cd/A)


Current density (A/m2)

4
Luminance (cd/m2)
Current density
1500 EL
Luminance
PL(a)
400
2 PL(k)
1000

200
500 0

400 450 500 550 600 650


0 0
4 8 12 16 0 25 50 75 100
Voltage (V) Current density (A/m2)

Figure 6.17  (a) I–V–L characteristic and (b) current ΦF of an LED device of HPS2,4. Shown in the inset of
panel b are the EL and PL spectra of its amorphous (a) and crystalline (k) films. Source: Adapted with
permission from Ref. [44]. Copyright 2015, American Chemical Society.

HPS2,4 is green emissive in amorphous state but blue in crystalline state, which makes it promis-
ing for OLED applications, as the carrier mobility is higher in crystal than in amorphous solid.
An OLED of Al/LiF/AlQ3/HPS2,4/NPB/ITO using HPS2,4 as the active layer was fabricated. The
I–V–L characteristic of the device is shown in Figure 6.17a. The EL spectrum of HPS2,4 is similar to
that of its crystal but different from that of its amorphous film (Figure 6.17b, inset), indicating that
in the EL device the molecules of HPS2,4 are mainly in the crystalline state. Although the configura-
tion of the OLED device is not optimized, it already exhibits an efficiency as high as 5.86 cd/A [44].
Whereas researchers in the area are trying hard to avoid crystal formation in the device fabrica-
tion processes, the crystalline layers made of the CIEE luminogens may offer excellent device
performances due to the high charge mobility in the crystals. This feature distinguishes CIEE
luminogens from the conventional fluorophores and promises the development of OLEDs with
high efficiencies.

6.5.3  High-­density Data Storage


The change in reflectivity has been commonly used as output in the rewritable optical media. A
luminescence-­based process is of advantage because of its higher sensitivity, lower background
198 6  Crystallization-­induced Emission Enhancement

noises, and potential for two-­dimensional imaging. The amorphous and crystalline films of CIEE
compound emit weak and strong lights with different color, respectively. If CIEE luminogens
could be reversibly switched between amorphous and crystalline states, their emission could be
reversibly modulated between dark and bright states, and they may be a group of promising candi-
dates for innovative applications in optical information storage systems.

6.5.4  Mechanochromic (MC) Luminescent Sensor


MC luminogens exhibit emission color or intensity change in response to mechanical stimuli such
as pressing, grinding, crushing, or rubbing and have promising applications in mechanosensors,
security papers, and optical storage. Few MC luminogens had been reported before 2008 probably
due to the lack of clear design strategy and the ACQ effect of traditional luminogen. The twisted
conformations of CIEE luminogens afford loosely packing patterns, which facilitates the phase
transformation in the solid state upon exposure to mechanical stimuli, thus many CIEE lumino-
gens exhibit MC luminescence. And some reviews on MC luminogen have been published for
reader’s reference [45–47].

6.6  ­Summary and Perspective

CIEE luminogens exhibit higher ΦF in stable crystalline state than the metastable amorphous solid
which may facilitate its real application. CIEE phenomenon is different from traditional lumino-
gens; disclosure of the mechanism of CIEE process may improve the optical theory, afford design
strategy of high-­performance luminogens, and so on. More and more researchers have been
attracted by CIEE and many CIEE luminogens have been reported, and some possible design strat-
egies for high-­performance CIEE luminogens have been proposed, such as larger conjugation core
with peripheral phenyl rings, luminogens with low symmetry. And application of CIEE luminogen
have also been explored in OLED, VOC sensors, mechanical stimuli sensors, data storage, and
security papers.
Although much work has been done in the area, many intriguing possibilities remain to be
explored. (i) Detailed information of molecular packing patterns of amorphous solid is difficult to
obtained directly, Raman spectroscopy has been used to detect the motion modes of phenyl
rings  [48, 49], and more experimental method should be developed to obtained information
of molecular microenvironmet, and to disclose the CIEE mechanism. (ii) OLED based on crys-
tals of CIEE luminogen had rarely been reported, and detailed technology to control the morphol-
ogy of the crystals and special structure of OLED may be necessary to pave the real application of
CIEE luminogen in OLED. (iii) Effective and detailed design strategy for high-­performance CIEE
luminogens should be developed based on the disclosure of the CIEE mechanism.

­References

1 Thomas, S. W., Joly, G. D., Swager, T. M. Chemical sensors based on amplifying fluorescent
conjugated polymers. Chem. Rev. 2007; 107 (4): 1339–86.
2 Grimsdale, A. C., Chan, K. L., Martin, R. E., Jokisz, P. G., Holmes, A. B. Synthesis of light-­emitting
conjugated polymers for applications in electroluminescent devices. Chem. Rev. 2009; 109 (3):
897–1091.
  ­Reference 199

3 Liu, J., Lam, J. W. Y., Tang, B. Z. Acetylenic polymers: syntheses, structures, and functions. Chem.
Rev. 2009; 109 (11): 5799–867.
4 Shimizu, M., Hiyama, T. Organic fluorophores exhibiting highly efficient photoluminescence in
the solid state. Chem. Asian J. 5 (7): 1516–31.
5 Luo, J. D., Xie, Z. L., Lam, J. W. Y., Cheng, L., Chen, H. Y., Qiu, C. F., et al. Aggregation-­induced
emission of 1-­methyl-­1,2,3,4,5-­pentaphenylsilole. Chem. Commun. 2001; (18): 1740–1.
6 Mei, J., Hong, Y. N., Lam, J. W. Y., Qin, A. J., Tang, Y. H., Tang, B. Z. Aggregation-­induced emission:
the whole is more brilliant than the parts. Adv. Mater. 2014; 26 (31): 5429–79.
7 Mei, J., Leung, N. L. C., Kwok, R. T. K., Lam, J. W. Y., Tang, B. Z. Aggregation-­induced emission:
together we shine, united we soar! Chem. Rev. 2015; 115 (21): 11718–940.
8 Zhao, Z., Gao, S., Zheng, X., Zhang, P., Wu, W., Kwok, R. T. K., et al. Rational design of
perylenediimide-­substituted triphenylethylene to electron transporting aggregation-­induced
emission luminogens (AIEgens) with high mobility and near-­infrared emission. Adv. Funct. Mater.
2018; 28 (11): 1705609.
9 Zhao, Z., Nie, H., Ge, C., Cai, Y., Xiong, Y., Qi, J., et al. Furan is superior to thiophene: a furan-­
cored AIEgen with remarkable chromism and OLED performance. Adv. Sci. 2017; 4 (8): 1700005.
10 He, J. T., Xu, B., Chen, F. P., Xia H. J., Li, K. P., Ye, L., et al. Aggregation-­induced emission in the
crystals of 9,10-­distyrylanthracene derivatives: the essential role of restricted intramolecular
torsion. J. Phys. Chem. C. 2009; 113 (22): 9892–9.
11 Yin, S., Shuai, Z., Wang, Y. A quantitative structure−property relationship study of the glass
transition temperature of OLED materials. J. Chem. Inform. Comput. Sci. 2003; 43 (3): 970–7.
12 Liu, S., He, F., Wang, H., Xu, H., Wang, C., Li, F., et al. Cruciform DPVBi: synthesis, morphology,
optical and electroluminescent properties. J. Mater. Chem. 2008; 18 (40): 4802.
13 Dong, Y. Q., Lam, J. W. Y., Li, Z., Tong, H., Law, C. W., Feng, X. D., et al. Vapochromism of
hexaphenylsilole and its blends with poly(methyl methacrylate). Polym. Mater. Sci. Eng. 2004;
91: 707–8.
14 Dong, Y. Q., Lam, J. W. Y., Li, Z., Qin, A. J., Tong, H., Dong, Y. P., et al. Vapochromism of
hexaphenylsilole. J. Inorg. Organomet. Polym. Mater. 2005; 15 (2): 287–91.
15 Dong, Y., Lam, J. W. Y., Qin, A., Li, Z., Sun, J., Sung, H. H. Y., et al. Switching the light emission of
(4-­biphenylyl)phenyldibenzofulvene by morphological modulation: crystallization-­induced
emission enhancement. Chem. Commun. 2007; (1): 40–2.
16 Shi, J., Zhao, W., Li, C., Liu, Z., Bo, Z., Dong, Y., et al. Switching emissions of two
tetraphenylethene derivatives with solvent vapor, mechanical, and thermal stimuli. Chin. Sci. Bull.
2013; 58 (22): 2723–7.
17 Luo, X., Zhao, W., Shi, J., Li, C., Liu, Z., Bo, Z., et al. Reversible switching emissions of
tetraphenylethene derivatives among multiple colors with solvent vapor, mechanical, and thermal
stimuli. J. Phys. Chem. C. 2012; 116 (41): 21967–72.
18 Tian, H., Wang, P., Liu, J., Duan, Y., Dong, Y. Q. Construction of a tetraphenylethene derivative
exhibiting high contrast and multicolored emission switching. J. Mater. Chem. C. 2017; 5 (48):
12785–91.
19 Lin, Y., Chen, G., Zhao, L., Yuan, W. Z., Zhang, Y., Tang, B. Z. Diethylamino functionalized
tetraphenylethenes: structural and electronic modulation of photophysical properties,
implication for the CIE mechanism and application to cell imaging. J. Mater. Chem. C. 2015; 3
(1): 112–20.
20 Wang, Z., Nie, H., Yu, Z., Qin, A., Zhao, Z., Tang, B. Z. Multiple stimuli-­responsive and reversible
fluorescence switches based on a diethylamino-­functionalized tetraphenylethene. J. Mater. Chem.
C. 2015; 3 (35): 9103–11.
200 6  Crystallization-­induced Emission Enhancement

21 Xu, B., Xie, M., He, J., Xu, B., Chi, Z., Tian, W., et al. An aggregation-­induced emission
luminophore with multi-­stimuli single-­and two-­photon fluorescence switching and large two-­
photon absorption cross section. Chem. Commun. 2013; 49 (3): 273–5.
22 Zeng, X., Yu, Z., Wang, F., Wang, X., Wang, C., Lu, H., et al. THF-­induced emission enhancement
and reversible stimulus response of a tetraphenylethylene luminogen dependent ESIPT
mechanism. Dyes Pigm. 2019; 171: 107699.
23 Li, J., Yang, C., Peng, X., Chen, Y., Qi, Q., Luo, X., et al. Stimuli-­responsive solid-­state emission
from o-­carborane-­tetraphenylethene dyads induced by twisted intramolecular charge transfer in
the crystalline state. J. Mater. Chem. C. 2018; 6 (1): 19–28.
24 Su, X., Wang, Y., Fang, X., Zhang, Y.-­M., Zhang, T., Li, M., et al. A high contrast tri-­state fluorescent
switch: properties and applications. Chem. Asian J. 2016; 11 (22): 3205–12.
25 Duan, Y., Xiang, X., Dong, Y. Diphenyldibenzofulvene derivatives exhibiting reversible
multicolored mechanochromic luminescence with high contrast. Acta Chimica Sinica 2016; 74
(11): 923–8.
26 Duan, Y., Ma, H., Tian, H., Liu, J., Deng, X., Peng, Q., et al. Construction of a luminogen exhibiting
high contrast and multicolored emission switching through combination of a bulky conjugation
core and tolyl groups. Chem. Asian J. 2019; 14 (6): 864–70.
27 Luo, X. L., Li, J. N., Li, C. H., Heng, L. P., Dong, Y. Q., Liu, Z. P., et al. Reversible switching of the
emission of diphenyldibenzofulvenes by thermal and mechanical stimuli. Adv. Mater. 2011; 23
(29): 3261–5.
28 Gu, X., Yao, J., Zhang, G., Yan, Y., Zhang, C., Peng, Q., et al. Polymorphism-­dependent emission for
di(p-­methoxylphenyl)dibenzofulvene and analogues: optical waveguide/amplified spontaneous
emission behaviors. Adv. Funct. Mater. 2012; 22 (23): 4862–72.
29 Zhao, Z., Chen, T., Jiang, S., Liu, Z., Fang, D., Dong, Y. Q. The construction of a multicolored
mechanochromic luminogen with high contrast through the combination of a large conjugation
core and peripheral phenyl rings. J. Mater. Chem. C. 2016; 4 (21): 4800–4.
30 Chen, W., Wang, S., Yang, G., Chen, S., Ye, K., Hu, Z., et al. Dicyanomethylenated acridone based
crystals: torsional vibration confinement induced emission with supramolecular structure
dependent and stimuli responsive characteristics. J. Phys. Chem. C. 2016; 120 (1): 587–97.
31 He, Z., Zhang, L., Mei, J., Zhang, T., Lam, J. W. Y., Shuai, Z., et al. Polymorphism-­dependent and
switchable emission of butterfly-­like bis(diarylmethylene)dihydroanthracenes. Chem. Mater. 2015;
27 (19): 6601–7.
32 Cariati, E., Lanzeni, V., Tordin, E., Ugo, R., Botta, C., Schieroni, A. G., et al. Efficient crystallization
induced emissive materials based on a simple push-­pull molecular structure. Phys. Chem. Chem.
Phys. 2011; 13 (40): 18005–14.
33 Zheng, R., Mei, X., Lin, Z., Zhao, Y., Yao, H., Lv, W., et al. Strong CIE activity, multi-­stimuli-­
responsive fluorescence and data storage application of new diphenyl maleimide derivatives.
J. Mater. Chem. C. 2015; 3 (39): 10242–8.
34 Mei, X., Wang, J., Zhou, Z., Wu, S., Huang, L., Lin, Z., et al. Diarylmaleic anhydrides: unusual
organic luminescence, multi-­stimuli response and photochromism. J. Mater. Chem. C. 2017; 5 (8):
2135–41.
35 Zhao, H., Wang, Y., Harrington, S., Ma, L., Hu, S., Wu, X., et al. Remarkable substitution influence
on the mechanochromism of cyanostilbene derivatives. RSC Adv. 2016; 6 (71): 66477–83.
36 Yoshii, R., Hirose, A., Tanaka, K., Chujo, Y. Functionalization of boron diiminates with unique
optical properties: multicolor tuning of crystallization-­induced emission and introduction into the
main chain of conjugated polymers. J. Am. Chem. Soc. 2014; 136 (52): 18131–9.
  ­Reference 201

37 Yoshii, R., Hirose, A., Tanaka, K., Chujo, Y. Boron diiminate with aggregation-­induced emission
and crystallization-­induced emission-­enhancement characteristics. Chem. Eur. J. 2014; 20
(27): 8320–4.
38 Yamaguchi, M., Ito, S., Hirose, A., Tanaka, K., Chujo, Y. Modulation of sensitivity to mechanical
stimulus in mechanofluorochromic properties by altering substituent positions in solid-­state
emissive diiodo boron diiminates. J. Mater. Chem. C. 2016; 4 (23): 5314–9.
39 Naito, H., Morisaki, Y., Chujo, Y. O-­carborane-­based anthracene: a variety of emission behaviors.
Angew. Chem. Int. Ed. Engl. 2015; 54 (17): 5084–7.
40 Mori, H., Nishino, K., Wada, K., Morisaki, Y., Tanaka, K., Chujo, Y. Modulation of luminescence
chromic behaviors and environment-­responsive intensity changes by substituents in bis-­o-­
carborane-­substituted conjugated molecules. Mater. Chem. Front. 2018; 2 (3): 573–9.
41 Zhan, Y., Xu, Y., Yang, P., Zhang, H., Li, Y., Liu, J. Carbazole-­based salicylaldimine difluoroboron
complex with crystallization-­induced emission enhancement and reversible piezofluorochromism
characteristics. Tetrahedron Lett. 2016; 57 (48): 5385–9.
42 Zhang, L., Chen, Y., Jiang, J. Solid state fluorescent functionalized-­triphenylamine BODIPY
detector for HCL vapor with high stability and absolute fluorescent quantum yield. Dyes Pigm.
2016; 124: 110–9.
43 Ohtani, S., Gon, M., Tanaka, K., Chujo, Y. A flexible, fused, azomethine–boron complex:
thermochromic luminescence and therosalient behavior in structural transitions between
crystalline polymorphs. Chem. Eur. J. 2017; 23 (49): 11827–33.
44 Li, Z., Dong, Y., Mi, B. X., Tang, Y. H., Haussler, M., Tong, H., et al. Structural control of the
photoluminescence of silole regioisomers and their utility as sensitive regiodiscriminating
chemosensors and efficient electroluminescent materials. J. Phys. Chem. B. 2005; 109 (20): 10061–6.
45 Sagara, Y., Kato, T. Mechanically induced luminescence changes in molecular assemblies. Nat.
Chem. 2009; 1 (8): 605–10.
46 Chi, Z., Zhang, X., Xu, B., Zhou, X., Ma, C., Zhang, Y., et al. Recent advances in organic
mechanofluorochromic materials. Chem. Soc. Rev. 2012; 41 (10): 3878–96.
47 Dong, Y. Q., Lam, J. W. Y., Tang, B. Z. Mechanochromic luminescence of aggregation-­induced
emission luminogens. J. Phys. Chem. Lett. 2015; 6 (17): 3429–36.
48 Zheng, C., Zang, Q., Nie, H., Huang, W., Zhao, Z., Qin, A., et al. Fluorescence visualization of
crystal formation and transformation processes of organic luminogens with crystallization-­induced
emission characteristics. Mater. Chem. Front. 2018; 2 (1): 180–8.
49 Zhang, T., Ma, H., Niu, Y., Li, W., Wang, D., Peng, Q., et al. Spectroscopic signature of the
aggregation-­induced emission phenomena caused by restricted nonradiative decay: a theoretical
proposal. J. Phys. Chem. C. 2015; 119 (9): 5040–7.
203

Surface-­fixation Induced Emission


Yohei Ishida1 and Shinsuke Takagi2
1
Division of Materials Science and Engineering, Faculty of Engineering, Hokkaido University, Sapporo, Hokkaido, Japan
2
Department of Applied Chemistry, Graduate Course of Urban Environmental Sciences, Tokyo Metropolitan University, Minami-­Oshawa,
Hachiohji, Tokyo, Japan

7.1 ­Introduction

The photochemical behavior of dyes can be affected by complex formation with two-­dimensional
(2D) nanosheet materials such as clay mineral nanosheets [1–25]. The absorption and emission
behavior of molecules on the clay surface have been reported by many researchers  [26–39]. In
some cases, a non-­ or weakly emissive molecule becomes fluorescent on the clay surface and/or
the fluorescence becomes much stronger as shown in Figure 7.1 [35]. In this chapter, the fluores-
cence enhancement phenomena of dyes on the clay mineral nanosheet surfaces are described. We
named this effect as surface-­fixation induced emission (S-­FIE) because of its mechanism.
Particularly, the composition of synthetic clay minerals is clear and its purity is high. Sumecton
SA (Kunimine Ind. Co., Japan) and Laponite (BYK Chemie GmbH, Moosburg, Germany) are com-
mercially available clays. Clay minerals are composed of many elements including Si, Mg, Al, and
O. The composition of Sumecton SA is [(Si4+7.2Al3+0.8)(Mg6)O20(OH)4] 0.8-­0.8Na+, and its structure
is shown in Figure 7.2. Isomorphous substitution of Si4+ with Al3+ in the tetrahedral layer pro-
duces anionic charges in the structure. The unit structure can be extended widely in two dimen-
sions. Typically, the average inter-­charge distance on the clay mineral surfaces is 1.2 nm in the
hexagonal array (calculation from its cation-­exchange capacity (CEC) of 0.997  meq/g), and its
theoretical surface area is 7.5 × 102 m2/g. It is possible to change the replacement ratio of Al3+, and
thus, control CEC through hydrothermal synthesis. Synthetic saponite-­type clay minerals have a
typical particle diameter of around 20–60 nm. Although the nanosheets stack in the solid state,
they can be swelled and exfoliated in an aqueous solution. When the synthetic saponite clay-­type
minerals are dispersed as single nanosheets, the solution can be substantially transparent. Since
the solution and membrane are transparent, the dye–clay complex can be optically observed and
used as colored and photochemically useful materials.
The features of clay mineral nanosheets as host materials for guest molecules are: (i) the flat
surface at atomic level, (ii) positive or negative charges in the structure, and (iii) optical transpar-
ency in solution state when its particle size is small (<50 nm), and the concentration is not high
(<1.0 × 10−1 g/l). Due to these features, the molecules on the surface could change their structure,

Handbook of Aggregation-Induced Emission: Volume 1 Tutorial Lectures and Mechanism Studies, First Edition.
Edited by Youhong Tang and Ben Zhong Tang.
© 2022 John Wiley & Sons Ltd. Published 2022 by John Wiley & Sons Ltd.
204 7  Surface-­fixation Induced Emission

TPAB aqueous solution (365nm irr.)

Without
With clay
clay

Figure 7.1  Photograph of triphenylbenzene derivative (TPAB) without clay (left) and with clay (right) in
aqueous solution. Source: Tokieda et al. [40]. Reproduced with permission of Elsevier.

O
Tetrahedral
Al3+or Si4+ sheet
Mg2+ Octahedral
sheet
Tetrahedral
sheet

Figure 7.2  Left: Unit structure of synthetic saponite-­type clay minerals. This unit is 2-­dimensionally
connected and forms disc-­shaped particles. Replacement of Si4+ with Al3+ affords an anionic charge to the
structure. Right: What happened to the molecular characteristics of molecules on the atomically flat
inorganic surfaces? Source: Tokieda et al. [40]. Reproduced with permission of Elsevier.

intramolecular electronic distribution, vibrational motion, and so on, then they could change their
fluorescence properties.
One of the early examples of this effect is the fluorescence enhancement of methylviologen on
clay mineral surfaces reported by Villemure, Detellier, and Szabo [41, 42]. While methylviologen is
not emissive in aqueous solutions, it is when adsorbed on the clay surface. Although the fluores-
cence suffers concentration quenching as adsorption density increases, it is enhanced by more
than one hundred times. Although this phenomenon is interesting, little research on it has been
conducted. Combining dye and clay without any designs or considerations frequently causes an
aggregation of the dyes, and the fluorescence becomes weak. Thus, regardless of its uniqueness,
this function of clay has not received attention.
A unique fluorescence behavior, aggregation-­induced emission (AIE), has been reported by Tang
et al. [43, 44]. In AIE, the suppression of vibrational motion leading to radiationless dissipation of
the exciton energy in an aggregation state could play an important role in emission enhancement.
Although S-­FIE has some similarities to AIE in terms of the phenomenon and mechanism, each
effect possesses different advantages. The advantages of S-­FIE for the differentiations between
them are mentioned in this chapter.
7.2  ­What Happened to the Characteristics of Molecules on the Clay Mineral Nanosheets 205

7.2  ­What Happened to the Characteristics of Molecules


on the Clay Mineral Nanosheets

The properties of a molecule are affected by its surrounding field. The photochemical (excited
state) properties are particularly affected by the surrounding field, including the molecular struc-
ture, solvation, and intermolecular interactions. A conceptual depiction of typical potential curves
of a molecule under the S0 and S1 state are shown in Figure 7.3. For diatomic molecules, the hori-
zontal axis indicates the inter-­atomic distance. In typical molecules, the curve indicates a multi-­
dimensional molecular conformation. In Figure 7.3, req. and r'eq. indicate the most stable structures
for S0 and S1, respectively. Because the electronic state in S0 and S1 is different, their most stable
structure is not the same (req.   r'eq.). At the S1 state, there are two relaxation pathways: radiative
and nonradiative deactivations. The probability is governed by the overlap of wave functions
between (S1, ν0) and (S0, ν0, 1, 2,) for radiative deactivation and by the overlap of wave functions
between (S1, ν0) and (S0, νhigh) for nonradiative deactivation. Depending upon the surrounding
chemical field, the value of req., r'eq., E(S0), E(S1), and the shapes of Ψ(S0) and Ψ(S1) change. These
effects of the chemical field around the molecule change the photochemical properties of the mol-
ecule such as absorption wavelength, fluorescence wavelength, absorption coefficient, and fluores-
cence quantum yield. For example, when ΔE (=  |E(S0)  −  E(S1)|) is decreased, the fluorescence
quantum yield tends to decrease because of the energy gap law  [45–50]. In contrast, when
Δr (=  |req.  −  r'eq.|) is decreased, the fluorescence quantum yield tends to be large because the
­overlap of wave functions between (S1, ν0) and (S0, ν0) increases while that between (S1, ν0) and
(S0, νhigh) decreases.

Figure 7.3  Conceptual potential energy


curve of molecule under the S0 and S1 state. Ψ(S1)
Energy

E(S1) V0

r′eq.

Vhigh

Ψ(S0)

E(S0) V0

req.

Nuclear coordinates
206 7  Surface-­fixation Induced Emission

How do clay mineral nanosheets act as a chemical field on the photochemical (excited state)
property of guest molecules? The major effects of clay mineral nanosheets are categorized as such:
(i) The molecule itself suffers changes in the molecular structure and molecular motion. The
molecular structure becomes more planar, and its motion suppressed on the layered silicate’s sur-
face because of the flat surface of the clay mineral nanosheets at the atomic level. (ii) The molecule
is in a high-­local concentration and tends to form aggregates. Molecular aggregation affects photo-
chemical properties because of the interaction between transition dipole moments. The details of
these effects are described in the next section.

7.3  ­Clay–Molecular Complexes

The dye molecule and clay mineral nanosheets form complexes through attractive effects such as
electrostatic and hydrophobic interactions. Cationic molecules adsorb on the surface of anioni-
cally charged clay mineral nanosheets mainly by electrostatic interactions. For example, it is well
known that methylene blue (MB, Figure 7.4 left) easily adsorbs on clay mineral nanosheets in an
aqueous dispersion  [53–56]. Complexation can be monitored by UV–Vis absorption measure-
ments. The absorption spectrum of MB changes with time after mixing each solution (Figure 7.4
left) [51, 56]. There is an interaction not only between the clay minerals nanosheets, and MB but
also between the MB molecules that cause the formation of aggregates. Generally, aggregation is
easily formed on clay mineral nanosheet surfaces or other inorganic surfaces in aqueous suspen-
sion mainly because of hydrophobic interactions between MB molecules. Typical aggregates are
H-­type and J-­type aggregates that are non-­emissive and emissive, respectively [53]. Although an
assembly including H-­type and J-­type aggregates on clay mineral nanosheets is unique, the excited-­
state lifetime of molecules on layered silicates tends to be short even for J aggregates; therefore,
molecules on clay mineral nanosheets are photochemically inactive or complicated in general.
Monomers

N 0.5
+
+N
N
N S N+
H -Aggreggates

H -Dimers

0.4 N
Cl– H
N N
H
Absorbance

J -Aggreggates

N
Absorbance

0.3
+
N N
+

0.2
C. [p-TMPyP]=
150
b. 0.1
10% vs CEC
a.
0
500 600 700 800 300 400 500 600 700
Wavelength(nm) Wavelength (nm)

Figure 7.4  Left: Absorption spectra of methylene blue in Kunipia montmorillonite dispersion in aqueous
solution. The spectra of the dispersion when measured at: (a) one minute and, (b) 24 hours after mixing the
dye solution with the clay dispersion; (c) a comparison with the MB solution. Source: Bujdák [51].
Reproduced with permission of Elsevier. Right: Absorption spectra of clay–p-­TMPyP complex at various dye
concentrations up to 150% vs. CEC. Source: Takagi et al. [52]. Adapted from the American Chemical Society.
7.4  ­Absorption Spectra of Clay–Molecular Complexe 207

Many researchers have found it difficult to use clay mineral nanosheets as a platform to construct
photochemical reaction systems, and photofunctional materials.
Recently, techniques to control the assembly of structures on clay mineral nanosheets have been
developed [3, 57–63]. Using a combination of anionic saponite and di-­, tri-­, and tetra-­cationic por-
phyrins, aggregation was completely suppressed. In the combined systems, the porphyrins were
adsorbed on the surface of the saponite at up to 100% vs. CEC of the clay minerals without causing
aggregation.
The absorption spectra of 5,10,15,20-­tetrakis(1-­methylpyridinium-­4-­yl) porphyrin (p-­TMPyP)
with a loading range of 10–150% vs. CEC are shown in Figure 7.4 (right). At up to 100% vs. CEC,
the spectral shape remained the same, indicating nonaggregate formation. Above 100% vs. CEC,
a new band appeared in the absorption spectra at a shorter wavelength ascribed to nonadsorbed
dye species. When the adsorption rate is 100% vs. CEC, the average intermolecular distance of p-­
TMPyP is 2.4 nm for a typical saponite-­type clay mineral nanosheet. The suggested mechanism
for suppressing aggregate formation on saponite is as follows: the inter-­anionic charge distance
on the saponite surface is 1.2 nm based on the hexagonal array, and the inter-­cationic distance in
p-­TMPyP is around 1.1 nm. Because of the good distance matching of charged positions between
the clay mineral nanosheet and guest molecule, the interaction is quite strong, resulting in aggre-
gation suppression between molecules. Because the relative inter-­charge charge-­matching plays
an important role, this mechanism is called the size-­matching effect [3, 57–61]. In this complex,
p-­TMPyP is photochemically active; it has a similar excited state lifetime compared with a mono-
mer in solution, and was thus subjected to physicochemical observation. It is possible to control
intermolecular distance by changing the charge density of clay  [60]. This is because such
­high-­density adsorption structures  – and enough long excited-­state lifetimes  – are favored for
intermolecular photochemical reactions, such as energy transfer and/or electron transfer
reactions [64–76].

7.4  ­Absorption Spectra of Clay–Molecular Complexes

Before describing emission enhancement behavior, we will briefly summarize UV–Vis absorption
properties of dye molecules on a clay mineral nanosheet. The absorption spectrum of guest mole-
cules on a clay nanosheet surface is much different from that in a bulk solution as a monomer. This
is shown in Figure  7.5  [57, 58, 77–79]. The reasons for spectral change are categorized into:
(i) aggregation [54, 80–84], (ii) protonation [33], (iii) isomerization [31], and (iv) molecular struc-
ture change [79]. Here, we focus on a system where (i) ~ (iii) can be excluded. In the case of spec-
tral change induced by (iv), the spectral shift to a longer wavelength (red shift) was often observed.
A typical example showing the absorption spectral change is summarized in Table 7.1 (see also the
list of molecular structures in Figure 7.6). Many researchers, including us, suggest that molecular
structure change on the clay surface plays an important role for the spectral shift. As the clay sur-
face is extremely flat at the atomic level, an adsorbed molecule could be flattened compared to a
free condition in solution. If the guest molecule possesses more than two cationic parts, a two-­
point interaction between the clay and guest molecule could induce the flattening of the molecule.
Even in the case of a mono-­cationic guest molecule, molecular flattening is expected. In an aque-
ous condition, there is a hydrophobic interaction between the clay surface and organic parts in the
adsorbed molecule. Thus, both the electrostatic and hydrophobic interaction could occur, and
the molecular flattening could be induced. As seen in Table 7.1, molecules with rotatable parts in
the molecule tend to exhibit a large spectral shift to a longer wavelength. If the molecule has
208 7  Surface-­fixation Induced Emission

30 nm Red shift by flattening of the molecule


on the inorganic surface
Absorbance

In water On the clay surface

300 400 500 600


Wavelength(nm)

Figure 7.5  Absorption spectra of p-­TMPyP4+ with (red) and without (blue) clay in water. Source: Ishida
et al. [79]. Reproduced with permission of the American Chemical Society.

aromatic rotatable substituents, the extension of the conjugate system leading to a spectral shift to
a longer wavelength could be enlarged by the molecular flattening.

7.5  ­Emission Enhancement Phenomenon in Clay–Molecular


Complexes: S-­FIE

In addition to UV–Vis absorption, emission is also affected by complexation with clay mineral
nanosheets. The most stationary stable structure as well as the motion of the molecule can be sup-
pressed by adsorption on the clay mineral nanosheet, and in the interlayer space of stacked clay
nanosheets. In fact, suppression of the non-­radiative deactivation process was observed in many
types of molecules complexed with clay mineral nanosheets [27, 32, 34, 37–38, 40, 85–87]. In con-
trast, the radiative deactivation process could be enhanced by an increase in the Franck–Condon
factor. These effects directly change the emission behavior of the molecule. Fluorescence quantum
yield (ϕf ) is expressed in Equation 7.1,

φ f = kf / ( knr + kf ) , (7.1)

where kf and knr denote the radiative deactivation rate constant, and the sum of the nonradiative
deactivation rate constant (equal to the sum of the thermal deactivation rate constant and intersys-
tem crossing rate constant), respectively. To enhance the ϕf, an increase in kf or a decrease in knr is
required. kf and knr can be calculated using the values of ϕf and fluorescence lifetimes (τ) according
to Equations 7.2 and 7.3.
k f f/
(7.2)

k nr 1 f / (7.3)

The effect of adsorption on clay mineral nanosheets is discussed using potential energy curves –
a simplistic model that is useful for a qualitative discussion. The plausible conceptual potential
energy curves for a molecule in the electronically ground and excited state on clay mineral
nanosheets – (a) and in solution – (b) are shown in Figure 7.7. On (a), it is expected that the station-
ary stable structure of the excited state will be similar to that of the ground state because the
7.5  ­Emission Enhancement Phenomenon in Clay–Molecular Complexes: S-­FI 209

Table 7.1  Absorption maxima (λmax) of molecules with and without clay in watera.

Compound λWAbs (nm) λCAbs (nm) λCAbs − λWAbs (nm)

Dimidium 478 527 +49


MPPb 371 384 +13
c
BDMAD 530 577 +47
MPd 362 372 +10
e
DASM 449 480 +31
MPSMf 396 419 +23
g
DPSM 397 419 +22
DASDMh 446 480 +34
i
DASMQ 509 543 +34
TPABj 254 264 +10
k
TMAB 296 316 +20
Pzl 687 698 +11
v 1+ m
[Sb (TPP)(OH)2]   417 431 +14
[Sbv(DMPyP)(OH)2]3+ n 418 431 +13
v 5+ o
[Sb (TMPyP)(OH)2]   414 424 +10
[Sbv(MPyP)(OH)2]1+ p 416 431 +15
v 1+ q
[Sb (DPyP)(OH)2]   415 429 +14
[Sbv(TPyP)(OH)2]1+ r 411 423 +12
a W C
λ Abs and Abs are λAbs of molecules without clay and with clay, respectively.
b
MPP is 5-­methyl-­6-­phenylphenanthridinium chloride.
c
BDMAD is 3,8-­bisdimethylamino-­5-­methyl-­6-­phenylphenanthridinium chloride.
d
MP is 5-­methylphenanthridinium chloride.
e
DASM is trans-­4-­[(dimethylamino)styryl]-­1-­methylpyridinium chloride.
f
MPSM is trans-­4-­(4-­methylpiperazin)styryl-­1-­methylpyridinium chloride.
g
DPSM is trans-­4-­(4-­dimethylpiperazin)styryl-­1-­methylpyridinium chloride.
h
DASDM is trans-­4-­[(dimethylamino)styryl]-­1,3-­dimethylpyridinium chloride.
i
DASMQ is trans-­4-­[(dimethylamino)stylyl]-­1-­methyl-­quinolinium chloride.
j
TMAB is 1,3,5-­tris(N,N,N-­trimethylanilinium-­4-­yl)benzene trichloride.
k
TPAB is 1,3,5-­tris[(N-­pyridinium)aniline-­4-­yl]benzene trichloride.
l
Pz is tetrapyridino[3,4-­b: 3′,4′-­g: 3″,4″-­l: 3′″,4′″-­q]porphyrazine tetrachloride.
m
[SbV(TPP)(OH)2]+ is dihydroxo(5,10,15,20-­tetraphenylporphyrinato)antimony(V) trichloride.
n
[SbV(DMPyP)(OH)2]3+ is dihydroxo[5,10-­diphenyl-­15,20-­di(N-­methyl-­pyridinium-­4-­yl)porphyrinato]antimony(V)
pentachloride.
o
[SbV(TMPyP)(OH)2]5+ is dihydroxo[5,10,15,20-­tetrakis(N-­methyl-­pyridinium-­4-­yl)porphyrinato]-­antimony(V)
pentachloride.
p
[SbV(MPyP)(OH)2]+ is dihydroxo[5,10,15-­triphenyl-­20-­mono(4-­pyridyl)-­porphyrinato]antimony(V) chloride.
q
[SbV(DPyP)(OH)2]+ is dihydroxo[5,15-­diphenyl-­10,20-­di(4-­pyridyl)porphyrinato]antimony(V) chloride.
r
[SbV(TPyP)(OH)2]+ is dihydroxo[5,10,15,20-­tetra(4-­pyridyl)porphyrinato]antimony(V) chloride.
Source: Tokieda et al. [40]. Reproduced with permission of Elsevier.

molecule is held on the nanosheet surfaces by strong electrostatic interactions. In such cases, kf
and knr tend to be large and small, respectively, compared to (b). Although the curve shape is
expressed in the same manner, it is expected to be sharper on nanosheet surfaces. These effects can
induce various changes in the photochemical properties of the molecule.
210 7  Surface-­fixation Induced Emission

H2N NH2 N N

N+ N+ N+ N+ Cl–

Cl– Cl– Cl–

Dimidium MPP BDMAD MP

+
N N N N N
+ +
+N N N
Cl– 2Cl– Cl–
DASM DPSM MPSM

N N
+ +
N N
Cl– Cl–

DASDM DASMQ

+ + N+
N N
N+
N

NH N

3Cl– N N 4Cl–
3Cl–
N HN

N
+
N
+
N N N
+N +
+ + N

TPAB TMAB Pz

N N+

OH N OH
N N OH
+ +
+
N Sb N N Sb N
N Sb N
Cl– 5Cl–
N 3Cl–
HO N HO N
HO

+N N +N N
+
+

[Sbv(TPP)(OH)2]1+ [Sbv(DMPyP)(OH)2]3+ [Sbv(TMPyP)(OH)2]5+


N N N

N OH N OH N OH
+ + +
N Sb N N Sb N N Sb N
Cl– Cl– Cl–
HO N HO N HO N

N N N N

[Sbv(MPyP)(OH)2]1+ [Sbv(DPyP)(OH)2]1+ [Sbv(TPyP)(OH)2]1+

Figure 7.6  Structures of the dyes examined in this chapter.


7.6  ­Mechanism of Surface-­Fixation Induced Emissio 211

(a) (b)
Ψ′(S1) Ψ′(S1)

Ψ(S0) Ψ(S0)
Energy

vhigh vhigh

v0 v0

Figure 7.7  Plausible conceptual potential energy curves of the ground and excited states of a molecule:
(a) On a clay mineral nanosheet, and (b) in solution. Source: Tsukamoto et al. [35]. Adapted from American
Chemical Society.

We have observed the emission enhancement behavior of many types of molecules on the clay
mineral nanosheets. A typical example, the emission enhancement of cationic phthalocyanine
derivative (Pz) by complexation is shown in Figure 7.8 left [37]. The fluorescence of Pz was more
strongly enhanced on the clay mineral nanosheet (sample b) and intercalated in the interlayer
space of nanosheets (sample c), compared to in a bulk solution as a monomer (without clay, sam-
ple a). In this case, the major factor for the fluorescence enhancement was the decrease of knr of Pz
on the clay surface (knr of samples a, b, and c: 99 × 108 s−1, 9.0 × 108 s−1, 2.4 × 108 s−1, respectively),
while kf was almost constant (kf of samples a, b, and c: 1.0 × 108 s−1, 1.0 × 108 s−1, 0.56 × 108 s−1,
respectively). In other cases, we observed an increase of kf for the fluorescence enhancement ϕf of
1,3,5-­tris[(N-­pyridinium)aniline-­4-­yl]-­benzene (TPAB) increased from 0.077 to 0.42 by complexa-
tion with clay mineral nanosheets (Figure 7.8 right) [35], where kf increased from 0.081 × 109 s−1
to 0.13 × 109 s−1 and knr decreased from 0.97 × 109 s−1 to 0.19 × 109 s−1, respectively. The effects of
clay mineral nanosheets on both kf and knr caused fluorescence enhancement. Based on a system-
atic demonstration of the many types of molecular structures listed in Figure 7.6 (also see Table 7.2;
ϕf of molecules listed in Figure 7.6 in a solution, and on the exfoliated clay mineral nanosheets),
this phenomenon has been named as a S-­FIE). Like the absorption spectral shift (Figure 7.5), the
emission spectral shift was also observed (Figure  7.9)  [27]. A complexation of clay mineral
nanosheets with m-­BIIITMPySp and p-­BIIITMPySp showed vivid color changes – from green to yel-
low and from orange to red, respectively.

7.6  ­Mechanism of Surface-­Fixation Induced Emission

Molecules on the clay mineral surfaces suffer molecular structural change and immobilization due
to interactions such as electrostatic and/or hydrophobic interactions. Because both the ground and
excited state is affected by atomically flat clay mineral surfaces, the structural change between the
ground and excited state becomes small, i.e. the structure under the excited state could be
212 7  Surface-­fixation Induced Emission

600 472 (0.42) TPAB


1 c ϕf = 0.19
Fluorescence intensity (a.u.)

Fluorescene intensity (a.u.)


Without clay
500
With clay

b ϕf = 0.10 400

300

200
482 (0.077)

100
a ϕf = 0.01
0 0
650 750 850 350 400 450 500 550 600
Wavelength (nm) Wavelength (nm)

Figure 7.8  Left: Fluorescence spectra and fluorescence quantum yields (ϕf) of Pz in: water (a), on the
exfoliated clay mineral nanosheet (b), and in the interlayer space of stacked clays, (c). The CEC value for b
and c is 0.5%. Source: Ishida et al. [37]. Adapted from the American Chemical Society. Right: The
fluorescence spectra of 1,3,5-­tris[(N-­pyridinium)aniline-­4-­yl]-­benzene (TPAB) after complexing with clay.
Source: Tsukamoto et al. [35]. Adapted from the American Chemical Society.

Table 7.2  The fluorescence quantum yields of various molecules in a bulk solution ( f ), on


w
exfoliated clay
mineral nanosheet ( fc), and the fraction of the two yields.

w c c w
Compound f f f /  f

Dimidium 0.041 0.14 3.4


MPP 0.16 0.40 2.5
BDMAD 0.017 0.12 7.1
MP 0.26 0.22 0.85
DASM 0.0031 0.091 30
MPSM 0.040 0.61 15
DPSM 0.055 0.64 12
DASDM 0.0048 0.20 42
DASMQ —­ 0.0060 —­
TPAB 0.077 0.42 5.5
TMAB 0.51 0.56 1.1
Pz 0.010 0.10 10
[Sbv(TPP)(OH)2]1+ 0.052 0.079 1.5
[Sbv(DMPyP)(OH)2]3+ 0.049 0.061 1.2
v 5+
[Sb (TMPyP)(OH)2] 0.032 0.032 1
[Sbv(MPyP)(OH)2]1+ 0.037 0.043 1.2
v 1+
[Sb (DPyP)(OH)2] 0.035 0.038 1.1
[Sbv(TPyP)(OH)2]1+ 0.027 0.030 1.1

Source: Tokieda et al. [40]. Reproduced with permission of Elsevier.


7.6  ­Mechanism of Surface-­Fixation Induced Emissio 213

90 50
[m-BIII(TMPySp)(OH)]Cl3 [p-BIII(TMPySp)(OH)]Cl3
80

Fluorescence intensity (a.u.)


Fluorescence intensity (a.u.)

Without clay Without clay


70 40 600
630
584 With clay With clay
60 540 (2% vs. CEC)
(2%vs. CEC)
30
50 515
40 (sh)
20
30
20 550
(sh) 10
10
0 0
450 550 650 750 500 600 700 800
Wavelength (nm) Wavelength (nm)

UV light irr: + UV light irr: N


+
N

HO HO

N
BN N
B
N

N N
N
+ +
N + N N
+

Without With clay Without With clay


clay clay

In water In water

Figure 7.9  Top: Fluorescence spectra normalized by absorbance at the excitation wavelength. The
excitation wavelengths are 370, and 400 nm for m-­BIIITMPySp; and 398 and 426 nm for p-­BIIITMPySp
without, and with the clay mineral nanosheets in water, respectively. Fluorescence maxima are shown in the
spectra. CEC of molecules was 2.0%. Bottom: Changes in fluorescence color of m-­BIIITMPySp and p-­
BIIITMPySp without and with the saponite excited at 365 nm in water. Source: Tsukamoto et al. [27]. Adapted
from ACS Applied Materials & Interfaces.

relatively similar to that under the ground state. Therefore, the overlap of wave functions for 0–0
transition between ground and excited state, i.e. the Franck–Condon factor, could become large.
An increase of the Franck–Condon factor could contribute to fluorescence enhancement.
As described in Equation 7.1, an enhancement of ϕf requires an increase in kf or a decrease in
knr, and we observed both. The plausible conceptual energy curves in Figure 7.10 left represent the
increase in kf due to complexation with clay mineral nanosheets. The potential curves for the
ground and excited states are shown for the dye molecule with and without clay mineral nanosheets,
respectively. An increase of kf means that the molecular structural difference between the ground
and excited state becomes small when the molecule adsorbs on the clay surface. We named this
effect as structure resembling effect (SRE). SRE increased kf because of the effective overlap of
wave functions (the Franck–Condon factor) between the ground and excited state.
The decrease of knr could be attributed to the suppression of intramolecular mobility, such as a
vibration or rotation, on the clay mineral nanosheet surfaces. Such effects can be conceptualized
in terms of potential curves in Figure 7.10 right. The slope of the potential is sharper with a clay
system than without it. The sharper curve means that the smaller intramolecular structure change
increases the energy. Such a change in the potential energy curve may decrease the overlap of wave
214 7  Surface-­fixation Induced Emission

Without clay With clay


S0 S1 S0 S1
Energy

v0′ v0′

v0 v0

Nuclear coordinate

Without clay With clay


S0 S1 S0 S1
Energy

Nuclear coordinate

Figure 7.10  Plausible conceptual potential energy curves of the ground and excited states of guest
molecules without and with clay, showing a structure resembling effect (left) and a structure fixing effect
(right). Source: Tsukamoto et al. [38]. Adapted from The Royal Society of Chemistry.

functions between the lowest vibrational state in the electrically excited, and high vibrational state
in the ground state although this result depends on the relation between these most stable struc-
tures. In other cases, the curve sharpening could make the activation energy for other states in the
excited state higher. These phenomena should lead to a decrease of knr. We named this effect a
structure-­fixing effect.

7.7  ­Summary and Outlook

In addition to the molecules listed in Figure 7.6, S-­FIE holds for a very wide variety of molecules
such as viologen, RuIIbpy3, or triphenylbenzene derivatives [24, 41–42, 63, 77]. This general versa-
tility is one of the unique characteristics of S-­FIE as compared to AIE [43, 44]. In the case of AIE,
  ­Reference 215

the suppression of vibrational motion leading to radiationless dissipation of the exciton energy in
an aggregation state could play an important role in emission enhancement. Although S-­FIE has
some similarities to AIE in terms of the actual phenomenon and mechanism, each process pos-
sesses different advantages. In S-­FIE, the photophysical observation and analysis is easy because
the sample solution is transparent, and the photochemical behavior is simple; for example, a fluo-
rescence decay curve of monomerically adsorbed molecules on clay mineral nanosheets can be
analyzed by a single exponential fitting. Moreover, it is easier to predict S-­FIE than AIE due to its
simpler mechanism that depends on the flat inorganic surfaces, and we can thus simply design the
photophysically enhanced system. By using these unique characteristics of S-­FIE, an artificial pho-
tosynthesis system based on arranging an artificial light-­harvesting system, and molecular photo-
catalysts has been reported [72, 87–93]. Moreover, the flexible clay mineral-­based materials can be
potentially used for applications in reversible environmental responsive emissive materials [94].
In this chapter, only clay mineral nanosheets are considered as 2D host materials, but use of
other 2D materials such as semiconducting nanosheets, 2D chalcogenide, and graphene are also
interesting [95–100]. The authors believe that S-­FIE and the unique functionality of clay mineral
nanosheets or related 2D materials can aid in the development of both the basic science, and
application-­based aspects.

­Acknowledgment

Yohei Ishida thanks financial supports from JSPS KAKENHI grant number 21K04805.

­References

1 Ogawa M, Kuroda K. Photofunctions of intercalation compounds. Chem. Rev. 1995;95:399–438.


Shichi T, Takagi K. Clay minerals as photochemical reaction fields. J. Photochem. Photobiol. A Chem.
2
2000;1:113–30.
3 Takagi S, Eguchi M, Tryk DA, Inoue H. Porphyrin photochemistry in inorganic/organic hybrid
materials: clays, layered semiconductors, nanotubes, and mesoporous materials. J. Photochem.
Photobiol. A Chem. 2006;7:104–26.
4 Bujdák J, Komadel P. Interaction of methylene blue with reduced charge montmorillonite. J. Phys.
Chem. B 1997;101:9065–8.
5 López Arbeloa F, Martínez Martínez V, Arbeloa T, López Arbeloa I. Photoresponse and anisotropy of
rhodamine dye intercalated in ordered clay layered films. J. Photochem. Photobiol. A Chem.
2007;8:85–108.
6 Čeklovský A, Czímerová A, Pentrák M, Bujdák J. Spectral properties of TMPyP intercalated in thin
films of layered silicates. J. Coll. Inter. Sci. 2008;324:240–5.
7 Suzuki Y, Tenma Y, Nishioka Y, Kawamata J. Efficient nonlinear optical properties of dyes confined
in interlayer nanospaces of clay minerals. Chem. Asian J. 2012;7:1170–9.
8 Zhou C-­H, Shen Z-­F, Liu L-­H, Liu S-­M. Preparation and functionality of clay-­containing films.
J. Mater. Chem. 2011;21:15132–53.
9 Bergaya F, Van Damme H. Stability of metalloporphyrins adsorbed on clays: a comparative study.
Geochim. Cosmochim. Acta 1982;46:349–60.
10 Podsiadlo P, Shim BS, Kotov NA. Polymer/clay and polymer/carbon nanotube hybrid organic–
inorganic multilayered composites made by sequential layering of nanometer scale films. Coord.
Chem. Rev. 2009;253:2835–51.
216 7  Surface-­fixation Induced Emission

11 Carrado KA. Synthetic organo-­and polymer–clays: preparation, characterization, and materials


applications. Appl. Clay Sci. 2000;17:1–23.
12 Teepakakorn AP, Bureekaew S, Ogawa M. Adsorption-­induced dye stability of cationic dyes on clay
nanosheets. Langmuir 2018;34:14069–75.
13 Liu P. Polymer modified clay minerals: A review. Appl. Clay Sci. 2007;38:64–76.
14 Choy J, Choi S, Oh J, Park T. Clay minerals and layered double hydroxides for novel biological
applications. Appl. Clay Sci. 2007;36:122–32.
15 Ruiz-­Hitzky E, Aranda P, Darder M, Rytwo G. Hybrid materials based on clays for environmental
and biomedical applications. J. Mater. Chem. 2010;20:9306–16.
16 Drozd D, Szczubiałka K, Skiba M, Kepczynski M, Nowakowska M. Porphyrin–nanoclay
photosensitizers for visible light induced oxidation of phenol in aqueous media. J. Phys. Chem. C
2014;118:9196–202.
17 Okada T, Ide Y, Ogawa M. Organic-­inorganic hybrids based on ultrathin oxide layers: designed
nanostructures for molecular recognition. Chem. Asian J. 2012;7:1980–92.
18 Ruiz-­Hitzky E, Aranda P, Darder M, Ogawa M. Hybrid and biohybrid silicate based materials:
molecular vs. block-­assembling bottom–up processes. Chem. Soc. Rev. 2011;40:801–28.
19 Chakraborty C, Dana K, Malik S. Intercalation of perylenediimide dye into LDH clays:
enhancement of photostability. J. Phys. Chem. C 2011;115:1996–2004.
20 Bizaia N, de Faria EH, Ricci GP, Calefi PS, Nassar EJ, Castro KADF, et al. Porphyrin-­kaolinite as
efficient catalyst for oxidation reactions. ACS Appl. Mater. Interfaces 2009;1:2667–78.
21 Čeklovský A, Czímerová A, Lang K, Bujdák J. Layered silicate films with photochemically active
porphyrin cations. Pure Appl. Chem. 2009;81:1385–96.
22 Sasai R, Itoh T, Ohmori W, Itoh H, Kusunoki M. Preparation and characterization of rhodamine
6G/alkyltrimethylammonium/Laponite hybrid solid materials with higher emission quantum
yield. J. Phys. Chem. C 2009;113:415–21.
23 Madhavan D, Pitchumani K. Photoreactions in clay media: singlet oxygen oxidation of electron-­
rich substrates mediated by clay-­bound dyes. J. Photochem. Photobiol. C Photochem. Rev.
2002;153:205–10.
24 Czímerová A, Jankovič L, Madejová J, Čeklovský A. Unique photoactive nanocomposites based on
rhodamine 6G/polymer/montmorillonite hybrid systems. J. Polym. Sci. Part B Polym. Phys.
2013;51:1672–9.
25 Takagi S, Shimada T, Ishida Y, Fujimura T, Masui D, Tachibana H, et al. Size-­matching effect on
inorganic nanosheets: control of distance, alignment, and orientation of molecular adsorption as a
bottom-­up methodology for nanomaterials. Langmuir 2013;29:2108–19.
26 Eguchi M, Shimada T, Inoue H, Takagi S. Kinetic analysis by laser flash photolysis of porphyrin
molecules’ orientation change at the surface of silicate nanosheet. J. Phys. Chem. C 2016;120:7428–34.
27 Tsukamoto T, Shimada T, Takagi S. Photophysical properties and adsorption behaviors of novel
tri-­cationic boron(III) subporphyrin on anionic clay surface. ACS Appl. Mater. Interfaces
2016;8:7522–8.
28 Sato N, Fujimura T, Shimada T, Tani T, Takagi S. J-­Aggregate formation behavior of a cationic
cyanine dye on inorganic layered material. Tetrahedron Lett. 2015;56:2902–5.
29 Konno S, Fujimura T, Otani Y, Shimada T, Inoue H, Takagi S. Microstructures of the porphyrin/
viologen monolayer on the clay surface: segregation or integration? J. Phys. Chem. C
2014;118:20504–10.
30 Ohtani Y, Ishida Y, Ando Y, Tachibana H, Shimada T, Takagi S. Adsorption and photochemical
behaviors of the novel cationic xanthene derivative on the clay surface. Tetrahedron Lett.
2014;55:1024–7.
  ­Reference 217

31 Umemoto T, Ohtani Y, Tsukamoto T, Shimada T, Takagi S. Pinning effect for photoisomerization of


a dicationic azobenzene derivative by anionic sites of the clay surface. Chem. Commun.
2014;50:314–6.
32 Tsukamoto T, Shimada T, Takagi S. Photochemical properties of mono-­, tri-­, and penta-­cationic
antimony(V) metalloporphyrin derivatives on a clay layer surface. J. Phys. Chem. A
2013;117:7823–32.
33 Czı́merová A, Jankovič L, Bujdák J. Effect of the exchangeable cations on the spectral properties of
methylene blue in clay dispersions. J. Coll. Inter. Sci. 2004;274:126–32.
34 Hagiwara S, Ishida Y, Masui D, Shimada T, Takagi S. Photochemical properties of cationic pyrene
derivative and energy transfer reaction between pyrene and porphyrin on the clay surface. Clay Sci.
2013;17:7–10.
35 Tsukamoto T, Shimada T, Takagi S. Unique photochemical properties of p-­substituted cationic
triphenylbenzene derivatives on a clay layer surface. J. Phys. Chem. C 2013;117:2774–9.
36 Hagiwara S, Ishida Y, Masui D, Shimada T, Takagi S. Unique photochemical behavior of novel
tetracationic pyrene derivative on the clay surface. Tetrahedron Lett. 2012;53:5800–2.
37 Ishida Y, Shimada T, Takagi S. “Surface-­Fixation Induced Emission” of porphyrazine dye by a
complexation with inorganic nanosheets. J. Phys. Chem. C 2014;118:20466–71.
38 Tsukamoto T, Shimada T, Takagi S. Structure resembling effect of clay surface on photochemical
properties of meso-­phenyl or pyridyl-­substituted monocationic antimony(v) porphyrin derivatives.
RSC Adv. 2015;5:8479–85.
39 Ishida Y, Shimada T, Ramasamy E, Ramamurthy V, Takagi S. Room temperature phosphorescence
from a guest molecule confined in the restrictive space of an organic–inorganic supramolecular
assembly. Photochem. Photobiol. Sci. 2016;15:959–63.
40 Tokieda D, Tsukamoto T, Ishida Y, Ichihara H, Shimada T, Takagi S. Unique fluorescence behavior
of dyes on the clay minerals surface: surface fixation induced emission (S-­FIE). J. Photochem.
Photobiol. C Photochem. Rev. 2017;339:67–79.
41 Villemure G, Detellier C, Szabo AG. Fluorescence of clay-­intercalated methylviologen. J. Am.
Chem. Soc. 1986;108:4658–9.
42 Villemure G, Detellier C, Szabo AS, 1991. Fluorescence of methylviologen intercalated into
montmorillonite and hectorite aqueous suspensions. Langmuir 1991;7:1215–21.
43 Luo J, Xie Z, Lam JWY, Cheng L, Tang BZ, Chen H, et al. Aggregation-­induced emission of
1-­methyl-­1,2,3,4,5-­pentaphenylsilole. Chem. Commun. 2001;1740–1.
44 Mei J, Leung NLC, Kwok RTK, Lam JWY, Tang BZ. Aggregation-­induced emission: together we
shine, united we soar! Chem. Rev. 2015;115:11718–940.
45 Lin SH. Energy gap law and Franck–Condon factor in radiationless transitions. J. Chem. Phys.
1970;53:3766–7.
46 Tripathy U, Kowalska D, Liu X, Velate S, Steer RP. Photophysics of Soret-­excited tetrapyrroles
in solution. I. Metalloporphyrins: MgTPP, ZnTPP, and CdTPP. J. Phys. Chem. A
2008;112:5824–33.
47 Wilson JS, Chawdhury N, Muna A-­MRA, Younus M, Muhammad SK, Paul RR, et al. The energy
gap law for triplet states in Pt-­containing conjugated polymers and monomers. J. Am. Chem. Soc.
2001;123:9412–7.
48 Christensen RL, Goyette M, Gallagher L, Joanna D, Beverly D, Johan L, et al. S1 and S2 States of
apo-­and diapocarotenes. J. Phys. Chem. A 1999;103:2399–407.
49 Grosshenny V, Harriman A, Francisco RM, Raymond Z. Electron delocalization in ruthenium (II)
and osmium (II) 2, 2′-­bipyridyl complexes formed from ethynyl-­bridged ditopic ligands. J. Phys.
Chem. 1996;100:17472–84.
218 7  Surface-­fixation Induced Emission

50 Hochstrasser RM, Marzzacco C. Perturbations between electronic states in aromatic and


heteroaromatic molecules. J. Chem. Phys. 1968;49:971–84.
51 Bujdák J. Effect of the layer charge of clay minerals on optical properties of organic dyes. A review.
Appl. Clay Sci. 2006;34:58–73.
52 Takagi S, Shimada T, Eguchi M, Yui T, Yoshida H, Tryk DA, et al. High-­density adsorption of
cationic porphyrins on clay layer surfaces without aggregation: the size-­matching effect. Langmuir
2002;18:2265–72.
53 Bujdák J. The effects of layered nanoparticles and their properties on the molecular aggregation of
organic dyes. J. Photochem. Photobiol. C Photochem. Rev. 2018;35:108–33.
54 Czı́merová A, Bujdák J, Gáplovský A. The aggregation of thionine and methylene blue dye in smectite
dispersion. 2004;243:89–96.
55 Schoonheydt RA, Heughebaert L. Clay adsorbed dyes: methylene blue on Laponite. Clay Miner.
2018;27:91–100.
56 Neumann MG, Schmitt CC, Gessner F. Time-­dependent spectrophotometric study of the
interaction of basic dyes with clays II: Thionine on natural and synthetic montmorillonites and
hectorites. J. Coll. Inter. Sci. 1996;177:495–501.
57 Takagi S, Shimada T, Yui T, Inoue H. High density adsorption of porphyrins onto clay layer without
aggregation: characterization of smectite-­cationic porphyrin complex. Chem. Lett. 2001;30:128–9.
58 Eguchi M, Takagi S, Tachibana H, Inoue H. The “size matching rule” in di-­, tri-­, and tetra-­cationic
charged porphyrin/synthetic clay complexes: effect of the inter-­charge distance and the number of
charged sites. J. Phys. Chem. Solids 2004;65:403–7.
59 Takagi S, Aratake Y, Konno S, Masui D, Shimada T, Tachibana H, et al. Effects of porphyrin
structure on the complex formation behavior with clay. Micropor. Mesopor. Mat. 2011;141:38–42.
60 Egawa T, Watanabe H, Fujimura T, Ishida Y, Yamato M, Masui D, et al. Novel methodology to
control the adsorption structure of cationic porphyrins on the clay surface using the “size-­
matching rule”. Langmuir 2011;27:10722–9.
61 Fujimura T, Shimada T, Hamatani S, Onodera S, Sasai R, Inoue H, et al. High density intercalation
of porphyrin into transparent clay membrane without aggregation. Langmuir 2013;29:5060–5.
62 Auwärter W, Écija D, Klappenberger F, Barth JV. Porphyrins at interfaces. Nature Chem.
2015;7:105–20.
63 Yui T, Takagi K, Inoue H. Microscopic environment and molecular orientation of guest molecules
within polyfluorinated surfactant and clay hybrids: photochemical studies of stilbazolium
derivatives. J. Photochem. Photobiol. C Photochem. Rev. 2018;363:61–7.
64 Czímerová A, Iyi N, Bujdák J. Energy transfer between rhodamine 3B and oxazine 4 in synthetic-­
saponite dispersions and films. J. Coll. Inter. Sci. 2007;306:316–22.
65 Czímerová A, Bujdák J, Iyi N. Fluorescence resonance energy transfer between laser dyes in
saponite dispersions. J. Photochem. Photobiol. C Photochem. Rev. 2007;187:160–6.
66 Bujdak J, Czímerová A, Arbeloa FL. Two-­step resonance energy transfer between dyes in layered
silicate films. J. Coll. Inter. Sci. 2011;364:497–504.
67 Madhavan D, Pitchumani K. Efficient triplet–triplet energy transfer using clay-­bound ionic
sensitizers. Tetrahedron 2002;58:9041–4.
68 Fujii K, Kuroda T, Sakoda K, Iyi N. Fluorescence resonance energy transfer and arrangements of
fluorophores in integrated coumarin/cyanine systems within solid-­state two-­dimensional
nanospace. J. Photochem. Photobiol. C Photochem. Rev. 2011;225:125–34.
69 Lezhnina MM, Bentlage M, Kynast UH. Nanoclays: two-­dimensional shuttles for rare earth
complexes in aqueous solution. Optical Materials 2011;33:1471–5.
  ­Reference 219

70 Tsukamoto T, Ramasamy E, Shimada T, Takagi S, Ramamurthy V. Supramolecular surface


photochemistry: cascade energy transfer between encapsulated dyes aligned on a clay nanosheet
surface. Langmuir 2016;32:2920–7.
71 Ohtani Y, Shimada T, Takagi S. Artificial light-­harvesting system with energy migration
functionality in a cationic dye/inorganic nanosheet complex. J. Phys. Chem. C
2015;119:18896–902.
72 Ohtani Y, Nishinaka H, Hoshino S, Shimada T, Takagi S. Anisotropic photochemical energy
transfer in clay/porphyrin system prepared by size-­matching effect and Langmuir. J. Photochem.
Photobiol. A Chem. 2015;313:15–8.
73 Tsukamoto T, Shimada T, Shiragami T, Takagi S. Photochemical chlorination and oxygenation
reaction of cyclohexene sensitized by Ga(III) porphyrin–clay minerals system with high durability
and usability. Bull. Chem. Soc. Jpn. 2015;88:578–83.
74 Ishida Y, Kulasekharan R, Shimada T, Ramamurthy V, Takagi S. Supramolecular-­surface
photochemistry: supramolecular assembly organized on a clay surface facilitates energy transfer
between an encapsulated donor and a free acceptor. J. Phys. Chem. C 2014;118:10198–203.
75 Eguchi M, Watanabe Y, Ohtani Y, Shimada T, Takagi S. Switching of energy transfer reaction by the
control of orientation factor between porphyrin derivatives on the clay surface. Tetrahedron Lett.
2014;55:2662–6.
76 Ishida Y, Kulasekharan R, Shimada T, Takagi S, Ramamurthy V. Efficient singlet-­singlet energy
transfer in a novel host-­guest assembly composed of an organic cavitand, aromatic molecules, and
a clay nanosheet. Langmuir 2013;29:1748–53.
77 Chernia Z, Langmuir DG. Flattening of TMPyP adsorbed on laponite. Evidence in observed and
calculated UV−Vis spectra. Langmuir 1999;15:1625–33.
78 Kuykendall VG, Langmuir JT. Photophysical investigation of the degree of dispersion of aqueous
colloidal clay. Langmuir 1990;6:1350–6.
79 Ishida Y, Masui D, Shimada T, Tachibana H, Inoue H, Takagi S. The mechanism of the porphyrin
spectral shift on inorganic nanosheets: the molecular flattening induced by the strong host–guest
interaction due to the “size-­matching rule.” J. Phys. Chem. C 2012;116:7879–85.
80 Bujdák J, Iyi N, Fujita T. The aggregation of methylene blue in montmorillonite dispersions. Clay
Miner. 2018;37:121–33.
81 Miyamoto N, Kawai R, Kuroda K, Ogawa M. Adsorption and aggregation of a cationic cyanine dye
on layered clay minerals. Appl. Clay Sci. 2000;16:161–70.
82 Lu L, Jones RM, McBranch D, Langmuir DW. Surface-­enhanced superquenching of cyanine dyes
as J-­aggregates on Laponite clay nanoparticles. Langmuir 2002;18:7706–13.
83 Bhattacharjee D, Hussain SA, Chakraborty S, Schoonheydt RA. Effect of nano-­clay platelets on the
J-­aggregation of thiacyanine dye organized in Langmuir-­Blodgett films: a spectroscopic
investigation. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2010;77:232–7.
84 Ogawa M, Kawai R, Kuroda K. Adsorption and aggregation of a cationic cyanine dye on smectites.
J. Phys. Chem. B 1996;100:16218–21.
85 Kudo N, Tsukamoto T, Tokieda D, Shimada T, Takagi S. Fluorescence enhancement behavior of
hemicyanine derivatives on the clay nanosheets: aggregation induced emission (AIE) vs. surface-­
fixation induced emission (S-­FIE). Chem. Lett. 2018;47:636–9.
86 Nakazato R, Shimada T, Ohtani Y, Ishida T, Takagi S. Adsorption and emission enhancement
behavior of 4,4′-­bipyridine on dispersed montmorillonite nano-­sheets under aqueous conditions.
Tetrahedron Lett. 2018;59:2459–62.
87 Takagi S, Eguchi M, Tryk DA, Inoue H. Light-­harvesting energy transfer and subsequent electron
transfer of cationic porphyrin complexes on clay surfaces. Langmuir 2006;22:1406–8.
220 7  Surface-­fixation Induced Emission

88 Tsukamoto T, Shimada T, Takagi S. Artificial photosynthesis model: photochemical reaction


system with efficient light-­harvesting function on inorganic nanosheets. ACS Omega
2018;3:18563–71.
89 Takagi S, Tryk DA, Inoue H. Photochemical energy transfer of cationic porphyrin complexes on
clay surface. J. Phys. Chem. B 2002;106:5455–60.
90 Ishida Y, Shimada T, Masui D, Tachibana H, Inoue H, Takagi S. Efficient excited energy transfer
reaction in clay/porphyrin complex toward an artificial light-­harvesting system. J. Am. Chem. Soc.
2011;133:14280–6.
91 Ishida Y, Shimada T, Takagi S. Artificial light-­harvesting model in a self-­assembly composed of
cationic dyes and inorganic nanosheet. J. Phys. Chem. C 2013;117:9154–63.
92 Suzuki S, Tatsumi D, Tsukamoto T, Honna R, Shimada T, Inoue H, et al. Active species transfer-­
type artificial light harvesting system in the nanosheet – dye complexes: utilization of longer
wavelength region of sunlight. Tetrahedron Lett. 2018;59:528–31.
93 Tatsumi D, Tsukamoto T, Honna R, Hoshino S, Shimada T, Takagi S. Highly selective
photochemical epoxidation of cyclohexene sensitized by Ru(II) porphyrin–clay hybrid catalyst.
Chem. Lett. 2017;46:1311–4.
94 Takagi S, Shimada T, Masui D, Tachibana H, Ishida Y, Tryk DA, Inoue H. Unique
solvatochromism of a membrane composed of a cationic porphyrin-­clay complex. Langmuir
2010;26:4639–41.
95 Sonotani A, Shimada T, Takagi S. “Size-­matching effect” in a cationic porphyrin–titania
nanosheet complex. Chem. Lett. 2017;46:499–501.
96 Sano K, Sonotani A, Tatsumi D, Ohtani Y, Shimada T, Takagi S. Characterization of dispersed
titania nanosheet under aqueous conditions and its complex formation behavior with cationic
porphyrin. J. Photochem. Photobiol. A Chem. 2018;353:597–601.
97 Sonotani A, Sano K, Wakayama S, Shimada T, Takagi S. Efficient electron injection from excited
porphyrin to titania nanosheet. Chem. Lett. 2018;47:803–5.
98 Dong R, Zhang T, Feng X. Interface-­assisted synthesis of 2D materials: trend and challenges.
Chem. Rev. 2018;118:6189–235.
99 Mas-­Ballesté R, Gómez-­Navarro C, Gómez-­Herrero J, Zamora F. 2D materials: to graphene and
beyond. Nanoscale 2011;3:20.
100 Schulman DS, Arnold AJ, Das S. Contact engineering for 2D materials and devices. Chem. Soc.
Rev. 2018;47:3037–58.
221

Aggregation-­induced Delayed Fluorescence


Yan Fu1, Hao Chen1, Zujin Zhao1, and Ben Zhong Tang1,2,3
1 
State Key Laboratory of Luminescent Materials and Devices, Guangdong Provincial Key Laboratory of Luminescence from Molecular
Aggregates, South China University of Technology, Guangzhou, China
2
 Department of Chemistry, Hong Kong Branch of Chinese National Engineering Research Centre for Tissue Restoration and
Reconstruction, Institute for Advanced Study, and Department of Chemical and Biological Engineering, The Hong Kong University of
Science & Technology, Clear Water Bay, Kowloon, Hong Kong, China
3
 Shenzhen Institute of Aggregate Science and Technology, School of Science and Engineering, The Chinese University of Hong Kong,
Shenzhen, Guangdong, China

8.1 ­Introduction

Photoelectronic materials are important information carriers, which are closely related to the pro-
gress of human civilization. Organic light-­emitting materials have the advantages of structural
diversity, various colors, good controllability, and so on. So far, they have rapidly developed in the
fields of organic light-­emitting diodes (OLEDs) [1–4], bioimaging [5, 6], and chemical sensors [7,
8]. Among them, OLEDs have many advantages such as high flexibility, self-­luminescence, short
response time, etc., which are considered to be the carriers of next-­generation displays and lighting
technologies. However, in the fabrication of OLED devices, common fluorescent materials often
have problems of emission quenching and low-­exciton utilization. The doping technique is widely
used to solve these problems [9, 10]. Although the doping technique can alleviate the concentra-
tion quenching problem, but at high voltages, the high-­concentration excitons will undergo bimo-
lecular quenching in the doped host materials [11], resulting in a drawback of efficiency roll-­off.
Meanwhile, the doping process is more complicated, which is not conducive to large-­scale produc-
tion in the industry. Therefore, the development of robust materials with high efficiency and good
stability for nondoped OLEDs needs to be addressed urgently, and is also an important issue to
promote the large-­scale commercial application of OLEDs.
Conventional organic luminescent materials mostly possess large π-­conjugation planes. In solu-
tion, the luminescence intensity is high, but in the neat film, the luminescence quenching caused
by the strong intermolecular interactions makes the luminescence intensity significantly reduced
that is known as aggregation-­caused quenching (ACQ) [12–14], and has limited the application of
organic light-­emitting materials in devices significantly. Since the first report of aggregation-­
induced emission (AIE) concept in 2001 [15, 16], many researchers have devoted themselves to
studying the mechanism and design principles of luminogens with AIE characteristics (AIEgens),
as AIE paves a new avenue towards ACQ-­free materials. The mechanism of the restriction of intra-
molecular motion (RIM) is generally recognized for the AIE phenomenon [15–19]. The emergence

Handbook of Aggregation-Induced Emission: Volume 1 Tutorial Lectures and Mechanism Studies, First Edition.
Edited by Youhong Tang and Ben Zhong Tang.
© 2022 John Wiley & Sons Ltd. Published 2022 by John Wiley & Sons Ltd.
222 8  Aggregation-­induced Delayed Fluorescence

of AIE luminescent materials provides a good idea for solving concentration quenching and exci-
ton annihilation, and exhibits superior properties when applied in nondoped OLEDs.
In OLEDs, the singlet and triplet excitons generated by electro-­excitation are in a ratio of 25 %:
75 %. Conventional fluorescent materials can only utilize singlet excitons, and thus, the theoretical
external quantum efficiency (EQE) is limited to 5–7.5%. The thermally activated delayed fluores-
cence (TADF) materials can solve this problem to a large degree. The materials with TADF prop-
erty generally have a twisted donor-­acceptor (D-­A) structure that allows the HOMO and LUMO to
be separated sufficiently to produce a small singlet–triplet energy splitting (ΔEst). Consequently,
triplet excitons can be converted into singlet excitons by the process of reverse intersystem crossing
(RISC) under thermal activation, leading to a maximum internal quantum efficiency (IQE) of
100% [20]. Therefore, TADF materials have attracted considerable attention, and developed into
the third generation of luminescent materials. However, there are also some potential problems
with TADF materials. Due to the high concentration of excitons at high voltages, the OLEDs based
on TADF materials still suffer from serious efficiency roll-­off.
Recently, a molecular design strategy of integrating the advantages of AIE and TADF emitters to
create robust luminescent materials was proposed  [21]. The generated materials possess
aggregation-­induced delayed fluorescence (AIDF); namely, they show weak emission and negligi-
ble delayed fluorescence in solution but strong emission with prominent delayed component upon
aggregate formation or in a neat film. The newly emerged AIDF luminogens have exhibited
impressive electroluminescence (EL) properties, in terms of excellent EL efficiency and greatly
suppressed efficiency roll-­off that may bring about a breakthrough for OLEDs. In this chapter, we
systematically summarize the recent advancements of AIDF materials, including the molecular
design, photophysical properties, and EL performances, and provide valuable insights into the
development of efficient light-­emitting materials for high-­performance OLEDs.

8.2 ­Novel Aggregation-­induced Delayed


Fluorescence Luminogens

The traditional TADF emitters can harvest both singlet and triplet excitons to achieve 100% exciton
utilization, but suffer from concentration-­caused exciton annihilation; and thus, serious efficiency
roll-­off occurred at high voltages due to the intermolecular interaction, as described by Dexter
energy transfer (DET) [22]. AIEgens have twisted configurations and can congenitally avoid ACQ
effect by suppressing strong intermolecular interactions [23]. Yasuda and coworkers [24] exploited
interesting Janus carborane triads with excellent AIE and TADF properties to fabricate nondoped
OLEDs in 2016, and suggested the possibility of endowing AIEgens with delayed fluorescence
(DF) characteristic.
To study the mechanism of AIDF a step further, a tailor-­made luminogen was synthesized,
namely, DBT-­BZ-­DMAC (Figure  8.1), with unsymmetrical Donor–Acceptor–Donor′ (D-­A-­D′)
structure by introducing 9,9-­dimethyl9,10-­dihydroacridine (DMAC) and dibenzothiophene (DBT)
into benzoyl (BZ) core [25]. The BZ serves as an electron acceptor, and DMAC and DBT function
as electron donors, respectively. According to crystallography analysis and molecular orbital
amplitude plots (Figure 8.2), the DBT-­BZ-­DMAC was in a highly twisted conformation, and the
HOMO and LUMO were well separated. DBT-­BZ-­DMAC showed strong emission in the solid state
without concentration or ACQ and the photoluminescence (PL) quantum yield (ΦPL) of the neat
film was 80.2%, being about tenfold higher than that in THF solution. The doped films, prepared
by codeposition of DBT-­BZ-­DMAC and 4,4′-­di(9H-­carbazol-­9-­yl)-­1,1′-­biphenyl (CBP) matrix, also
showed strong PL, but the moderately lowered the ΦPL values to some extent, and the ΔEST became
8.2 ­Novel Aggregation-­induced Delayed Fluorescence Luminogen 223

O O
O
O N S
N O
N N
S
O
S
S
DBT-BZ-PTZ DMF-BP-PXZ
DBT-BZ-DMAC DBT-BZ-PXZ

O O
O
O N
N

N N
N N
O
O
O S
DPF-BP-PXZ SBF-BP-PXZ CP-BP-PXZ CP-BP-PTZ

O
O O
N
N
N
N

CP-BP-DMAC DMF-BP-DMAC DPF-BP-DMAC

Figure 8.1  Molecule structures of aggregation-­induced delayed fluorescence materials with carbonyl


group as electron acceptor.

(a) (c)
EL efficiencies: 43.3 cd/A, 35.7 lm/W and 14.2%

O1 87°
101
External quantum efficiency (%)

N1
O

S
S1
DBT-BZ-DMAC
(b) LUMO
100
TAPC DBT-
(25 BZ-
ITO DMAC LiF/AI
nm) TmPyPB
(35 nm)
(55 nm)
HOMO

10–1
102 103 104
Luminance (cd/m2)

Figure 8.2  (a) Crystal structures of DBT-­BZ-­DMAC (CCDC 1483487); (b) molecular orbital amplitude plot of
DBT-­BZ-­DMAC, calculated by B3LYP/6-­31G (d, p); (c) EQE−luminance curves of nondoped OLED based on
DBT-­BZ-­DMAC.

smaller as the doping concentration increased. The DBT-­BZ-­DMAC possessed both the AIE and
TADF properties with great solid state PL efficiency and high-­exciton utilization. The nondoped
OLEDs based on DBT-­BZ-­DMAC were prepared with the configuration of ITO/TAPC (25  nm)/
DBT-­BZ-­DMAC (35 nm)/TmPyPB (55 nm)/LiF (1 nm)/Al, and the doped devices were also fabri-
cated by codepositing CBP and DBT-­BZ-­DMAC with different doping concentrations. The best
maxima current efficiency (ηC), power efficiency (ηP), and external quantum efficiency (ηext) of
224 8  Aggregation-­induced Delayed Fluorescence

nondoped OLEDs were up to 43.3 cd/A, 35.7 lm/W, and 14.2%, respectively. Notably, the EL effi-
ciencies of the nondoped OLED were still maintained at 43.1 cd/A, 33.1 lm/W, and 14.2% as lumi-
nance increased to 1000 cd/m, realizing almost negligible efficiency roll-­off. Compared to the
doped OLEDs, the nondoped OLED shows relatively lower EL efficiencies, but lower turn-­on volt-
age of 2.7 V, higher luminance of 27 270 cd/m2 and negligible efficiency roll-­off (Table  8.2).
Furthermore, another two similar molecules DBT-­BZ-­PXZ and DBT-­BZ-­PTZ with an asymmetric
D-­A-­D′ configuration were studied, which were composed of a central benzoyl as acceptor, and
phenoxazine (PXZ) and phenothiazine (PTZ) as donor units [26]. Both molecules also exhibited an
excellent AIDF property. The nondoped devices with DBT-­BZ-­PXZ as the emitter showed high-­EL
efficiencies of 22.6 cd/A, 27.9 lm/W, and 9.2%, with small efficiency roll-­off. For the device based
on DBT-­BZ-­PTZ, the EL efficiencies were 26.5 cd/A, 29.1 lm/W, and 9.7%. Since DMAC group is
more branched than PTZ and PXZ groups, and thus can efficiently suppress strong intermolecular
interactions, and thus lead to a higher ΦPL than that of DBT-­BZ-­PTZ and DBT-­BZ-­PXZ, accounting
for the superior EL efficiencies of DBT-­BZ-­DMAC. In short, these results indicated that the mole-
cules with an asymmetric D-­A-­D′ structure were excellent emitters for nondoped OLEDs, and
endowing AIEgens with DF property could be a promising molecular design principle to explore
robust light emitters for efficient and stable OLEDs with the simplified structures.
Endowing AIEgens with DF characteristics was considered as a promising strategy for designing
novel robust luminescent materials for efficient and stable nondoped OLEDs. In order to explore
the internal mechanism of AIDF, three new AIDF materials (DMF-­BP-­PXZ, DPF-­BP-­PXZ, and
SBF-­BP-­PXZ) containing an electron-­withdrawing benzoyl (BP) core were designed and investi-
gated [27]. By comparing the photophysical transition rates of DMF-­BP-­PXZ, DPF-­BP-­PXZ, and
SBF-­BP-­PXZ in solutions and in neat films, including radiative decay rate (kF), internal conversion
rate (kIC), intersystem crossing rate (kISC), and reverse intersystem crossing (RISC) rate (kRISC)
(Table  8.1). It was found that the intramolecular rotation was highly active in solution, which
played an overwhelming role in a primary internal conversion (IC) channel to inactivate the
excited state quickly and nonradiatively. Conversely, the intramolecular motions were restricted
greatly in the aggregated state, and thus, the nonradiative IC channels were blocked, which pro-
moted the intersystem crossing (ISC) and RISC processes under the condition of a small ΔEST,
resulting in the occurrence of AIDF (Figure 8.3). Moreover, according to theoretical calculations,
in the aggregated state, the ΔEST decreased and the spin-­orbit coupling (SOC) value increased rela-
tive to those in solution, which were also conducive to the occurrence of AIDF.
As we know, serious bimolecular quenching processes, such as triplet–triplet annihilation (TTA)
and singlet–triplet annihilation (STA), can make the EL efficiency decrease drastically, and the
DET with short-­range nature can lead to a severe exciton annihilation in high concentration.

Table 8.1  Photophysical parameters of DMF-­BP-­PXZ, DPF-­BP-­PXZ, and SBF-­BP-­PXZ.

DMF-­BP-­PXZ DPF-­BP-­PXZ SBF-­BP-­PXZ

THF Film THF Film THF Film

kF[×106 s−1] 11.1 10.6 8.8 13.5 10.1 13


7 −1
kIC[×10 s ] 41.7 1.3 30.5 1.4 55.2 1.6
kISC[×106 s−1] —­ 9.7 —­ 8.2 —­ 7.7
6 −1
kRISC[×10 s ] —­ 1.0 —­ 1.2 —­ 1.1
8.2 ­Novel Aggregation-­induced Delayed Fluorescence Luminogen 225

Solution Intersystem
crossing
S*
T*

Fluorescence

Non-radiative
conversion
Energy

relaxation
Internal
S0

Solid Intersystem
crossing
S*
T*
Reverse intersystem
Fluorescence

Fluorescence

Non-radiative
Energy

conversion

crossing

relaxation
Delayed

Internal

S0

Aggregation-induced delayed fluorescence


Figure 8.3  The proposed photophysical mechanisms for AIDF molecules in solution and solid state.

Therefore, endowing AIEgens with DF property should be a feasible strategy, owing to the rela-
tively long distance between excitons that are located at the center of branched AIEgens (Figure 8.4).
The carbonyl is a simple, but efficient electron withdrawing core for the AIDF materials because it
can increase the SOC value by means of n-­π* transition. In addition, in order to increase carrier
transport ability, electron-­donating group carbazole was also adopted to create AIDF materials.
The generated materials CP-­BP-­PXZ, CP-­BP-­PTZ, and CP-­BP-­DMAC had highly twisted conforma-
tions [21] that were favored to impede close packing and weaken intermolecular interactions. All
of these materials showed AIE characteristics with strong emissions as the aggregates formed
upon the addition of a large amount of the poor solvent (water) into their THF solutions. The DF
also became prominent with the formation of aggregates. Their neat films exhibited prompt fluo-
rescence with short lifetimes of 19.1–23.5  ns, and distinct DF with long lifetimes of 2.1–5.7 μs.
These luminogens showed very small ΔEST values of 0.016–0.033 eV in neat films, allowing RISC
processes to occur efficiently at ambient temperature. Their nondoped devices with a three-­layer
configuration of ITO/TAPC (25 nm)/emitting layer (EML) (35 nm)/TmPyPB (55 nm)/LiF (1 nm)/
Al were fabricated by vacuum deposition. As for the nondoped device of CP-­BP-­PXZ, the turn-­on
voltage was as low as 2.5 V, and the maximum ηC and ηext were as high as 59.1 cd/A and 18.4%,
respectively. Furthermore, when the luminance was increased to 1000 cd/m2, the corresponding
226 8  Aggregation-­induced Delayed Fluorescence

(a) + –

25%
75%
STA
RISC
TTA
S1
T1

IC PF DF
Heat
Serious bimolecular quenching processes
decreased exciton utillization
S0

(b)

25%
75%
RISC

S1

IC PF DF
Suppressed bimolecular quenching processes
Increased exciton utillization T1
S0

Figure 8.4  The main exciton dynamic processes for (a) a conventional luminogen with DF and (b) an AIDF
emitter in nondoped OLEDs, with prompt fluorescence (PF), delayed fluorescence (DF), internal conversion
(IC), singlet–triplet annihilation (STA), triplet–triplet annihilation (TTA), and reverse intersystem
crossing (RISC).

current efficiency roll-­off was as small as 1.2%, demonstrating the impressive efficiency stability.
For comparison, doped devices were also fabricated with the same configuration except for the
EML that used doped films with 6 and 10 wt% emitters doped in CBP host, and the EL performance
data was shown in Table 8.2. It was found that the nondoped OLEDs of these AIDF luminogens
offered lowered turn-­on voltages and enhanced luminance with respect to the doped OLEDs,
while the high-­EL efficiencies attained in these nondoped devices were comparable to those of
doped devices.
AIDF luminogens had shown great potentiality for nondoped OLEDs due to their ACQ-­free
property and 100% theoretical utilization of the electrically generated excitons. Recently, two new
Table 8.2  The key electroluminescence data of the AIDF materials.

Electroluminescence data

ηext,max (%)/ηC,max
Emitters Device configurationa Vonb (cd/A)/ηP,max (lm/W)c λELd (nm) Lmaxe (cd/m2) ROf (%) CIEg

DBZ-­BP-­DMAC ITO/TAPC/EML/TmPyPB/LiF/Al 2.7 14.2/43.3/35.7 516 27 270 0.46 (0.260, 0.550)


DBT-­BZ-­PXZ ITO/TAPC/EML/TmPyPB/LiF/Al 2.9 9.2/26.6/27.9 —­ —­ 26.1 (0.434, 0.542)
DBT-­BZ-­PTZ ITO/TAPC/EML/TmPyPB/LiF/Al 2.7 9.7/26.5/29.1 —­ —­ 12.4 (0.448, 0.531)
DMF-­BP-­PXZ ITO/TAPC/EML/TmPyPB/LiF/Al 2.7 13.3/39.9/38 —­ 27 331 6.0 (0.440, 0.543)
DPF-­BP-­PXZ ITO/TAPC/EML/TmPyPB/LiF/Al 2.6 14.3/41.6/45.0 —­ 31 422 1.4 (0.458, 0.530)
SBF-­BP-­PXZ ITO/TAPC/EML/TmPyPB/LiF/Al 2.5 12.3/36.8/37.9 —­ 33 990 0.8 (0.456, 0.528)
CP-­BP-­PXZ ITO/TAPC/EML/TmPyPB/LiF/Al 2.5 18.4/59.1/65.7 548 100 290 1.2 (0.40, 0.57)
CP-­BP-­PTZ ITO/TAPC/EML/TmPyPB/LiF/Al 2.5 15.3/46.1/55.7 554 46 820 16.7 (0.42, 0.55)
CP-­BP-­DMAC ITO/TAPC/EML/TmPyPB/LiF/Al 2.7 15.0/41.6/37.9 502 376 80 0.2 (0.23, 0.49)
CC6-­DBP-­PXZ ITO/PEDOT:PSS/PVK/EML/TmPyPB/LiF/Al 4.2 9.02/25.08/11.25 505 14 366 9.6 (0.265, 0.501)
CC6-­DBP-­DMAC ITO/PEDOT:PSS/PVK/EML/TmPyPB/LiF/Al 2.9 7.73/22.23/16.11 568 30 644 4.8 (0.454, 0.523)
DCB-­BP-­PXZ ITO/TAPC/EML/TmPyPB/LiF/Al 2.5 22.6/79.2/81.8 548 95 577 11.4 (0.39, 0.57)
CBP-­BP-­PXZ ITO/TAPC/EML/TmPyPB/LiF/Al 2.5 21.4/69.0/75.0 546 98 089 10.1 (0.39, 0.57)
mCP-­BP-­PXZ ITO/TAPC/EML/TmPyPB/LiF/Al 2.5 22.1/72.3/79.0 542 100 126 10.8 (0.39, 0.57)
mCBP-­BP-­PXZ ITO/TAPC/EML/TmPyPB/LiF/Al 2.5 21.8/70.4/76.5 542 96 815 9.9 (0.38, 0.57)
DMF-­BP-­DMAC ITO/HATCN//TAPC/TcTa/EML/TmPyPB/LiF/Al 3.2 5.7/19.9/16.4 528 23 270 —­ (0.39, 0.57)
DMF-­BP-­DMAC ITO/HATCN/mCP/EML/TPBi/LiF/Al 3.9 6.4/21.6/14.6 526 32 460 —­ (0.39, 0.57)
DPF-­BP-­DMAC ITO/HATCN/TAPC/TcTa/EML/TmPyPB/LiF/Al 2.8 13.2/43.8/40.2 534 50 170 —­ (0.39, 0.57)
DPF-­BP-­DMAC ITO/HATCN/mCP/EML/TPBi/LiF/Al 3.1 14.4/42.3/30.2 524 52 560 —­ (0.38, 0.57)
TATC-­BP ITO/PEDOT:PSS/EML/TmPyPB/LiF/Al 2.6 5.9/17.8/20.0 549 —­ 18.6 (0.41, 0.54)
TATP-­BP ITO/PEDOT:PSS/EML/TmPyPB/LiF/Al 2.7 6.0/46.4/47.2 541 —­ 3.3 (0.38, 0.55)

(Continued)
Table 8.2  (Continued)

Electroluminescence data

ηext,max (%)/ηC,max
Emitters Device configurationa Vonb (cd/A)/ηP,max (lm/W)c λELd (nm) Lmaxe (cd/m2) ROf (%) CIEg

4,4-­CzSPz ITO/PEDOT:PSS/mCP/EML/TPBi/LiF/Al 3.6 20.7/61.2/38.4 526 —­ —­ —­


PTZ2PTO ITO/MoO3/TAPC/mCP/EML/DPEPO/TPBi/LiF/Al 4.3 16.4/44.9/32.0 504 1 568 7.9 (0.27, 0.50)
34AcCz-­PM ITO/MoO3/TAPC/TCTA/EML/TmPyPB/LiF/Al 3.1 14.1/45.2/40.0 534 —­ —­ —­
PTZ-­DMNI ITO/PEDOT:PSS/mCP/EML/TPBi/LiF/Al 6.0 7.13/16.8/9.3 635 4 634 —­ (0.59, 0.40)
DMAC-­DMNI ITO/PEDOT:PSS/mCP/EML/TPBi/LiF/Al 5.0 5.38/12.4/5.2 570 2 312 —­ (0.37, 0.42)
Cz-­AQ ITO/PEDOT:PSS/EML/TPBi/Cs2CO3/Al 3.5 5.8/10.8/—­ 572 3 200 —­ (0.50, 0.49)
TPA-­AQ ITO/PEDOT:PSS/EML/TPBI/Cs2CO3/Al 3.8 7.5/10.6/—­ 612 2 200 —­ (0.60, 0.40)
NCFMCz1 ITO/MoO3/TCTA/mCP/EML/TSPO1/TPBi/Ca/Al 3.9 7.2/20.3/—­ 503 —­ 7.0 —­
NCFTCz ITO/MoO3/TCTA/mCP/EML/TSPO1/TPBi/Ca/Al 4.4 3.2/5.4/—­ 466 —­ —­ —­
TRZ-­HPB-­PXZ ITO/HATCN//TAPC/TcTa/EML/TmPyPB/LiF/Al 2.5 12.7/41.2/44.9 544 40 382 2.7 (0.39, 0.57)
TRZ-­HPB-­DMAC ITO/HATCN//TAPC/TcTa/EML/TmPyPB/LiF/Al 3.1 6.5/21.4/17.6 521 15 460 7.0 (0.28, 0.58)
CzTAZPO ITO/PEDOT:PSS/EML/TmPyPB/Ca/Al 4.5 12.8/29.1/—­ 537 9 776 1.8 (0.37, 0.56)
sCzTAZPO ITO/PEDOT:PSS/EML/TmPyPB/Ca/Al 4.1 9.6/20.6/—­ 531 8 283 0.97 (0.36, 0.56)
Ac3TRZ3 ITO/PEDOT:PSS/EML/TSPO1/TmPyPB/LiF/Al 3.4 3.5/11.4/—­ 520 —­ 14.2 (0.30, 0.54)
TAc3TRZ3 ITO/PEDOT:PSS/EML/TSPO1/TmPyPB/LiF/Al 2.9 3.1/10.2/—­ 538 —­ 22.5 (0.26, 0.48)
P1-­05 ITO/PEDOT:PSS/EML/TSPO1/TmPyPB/LiF/Al 3.2 7.1/10.6/—­ 455 1 902 —­ (0.17, 0.17)
P2-­05 ITO/PEDOT:PSS/EML/TSPO1/TmPyPB/LiF/Al 3.2 12.1/24.8/—­ 472 6 150 —­ (0.18, 0.27)
P3-­05 ITO/PEDOT:PSS/EML/SPPO13/LiF/Al 3.0 16.2/50.3/—­ 525 10 273 —­ (0.34, 0.55)
P4-­05 ITO/PEDOT:PSS/EML/SPPO13/LiF/Al 3.2 7.8/21.1/—­ 568 5 339 —­ (0.46, 0.50)
P5-­05 ITO/PEDOT:PSS/EML/SPPO13/LiF/Al 4.6 1.0/1.7/—­ 616 1 283 —­ (0.57, 0.41)
Electroluminescence data

ηext,max (%)/ηC,max
Emitters Device configurationa Vonb (cd/A)/ηP,max (lm/W)c λELd (nm) Lmaxe (cd/m2) ROf (%) CIEg

NCFMCz1 ITO/MoO3/TCTA/mCP/EML/TSPO1/TPBi/Ca/Al 3.9 7.2/20.3/—­ 503 —­ 7.0 —­


NCFTCz ITO/MoO3/TCTA/mCP/EML/TSPO1/TPBi/Ca/Al 4.4 3.2/5.4/—­ 466 —­ —­ —­
TRZ-­HPB-­PXZ ITO/HATCN//TAPC/TcTa/EML/TmPyPB/LiF/Al 2.5 12.7/41.2/44.9 544 40 382 2.7 (0.39, 0.57)
TRZ-­HPB-­DMAC ITO/HATCN//TAPC/TcTa/EML/TmPyPB/LiF/Al 3.1 6.5/21.4/17.6 521 15 460 7.0 (0.28, 0.58)
CzTAZPO ITO/PEDOT:PSS/EML/TmPyPB/Ca/Al 4.5 12.8/29.1/—­ 537 9 776 1.8 (0.37, 0.56)
sCzTAZPO ITO/PEDOT:PSS/EML/TmPyPB/Ca/Al 4.1 9.6/20.6/—­ 531 8 283 0.97 (0.36, 0.56)
Ac3TRZ3 ITO/PEDOT:PSS/EML/TSPO1/TmPyPB/LiF/Al 3.4 3.5/11.4/—­ 520 —­ 14.2 (0.30, 0.54)
TAc3TRZ3 ITO/PEDOT:PSS/EML/TSPO1/TmPyPB/LiF/Al 2.9 3.1/10.2/—­ 538 —­ 22.5 (0.26, 0.48)
P1-­05 ITO/PEDOT:PSS/EML/TSPO1/TmPyPB/LiF/Al 3.2 7.1/10.6/—­ 455 1 902 —­ (0.17, 0.17)
P2-­05 ITO/PEDOT:PSS/EML/TSPO1/TmPyPB/LiF/Al 3.2 12.1/24.8/—­ 472 6 150 —­ (0.18, 0.27)
P3-­05 ITO/PEDOT:PSS/EML/SPPO13/LiF/Al 3.0 16.2/50.3/—­ 525 10 273 —­ (0.34, 0.55)
P4-­05 ITO/PEDOT:PSS/EML/SPPO13/LiF/Al 3.2 7.8/21.1/—­ 568 5 339 —­ (0.46, 0.50)
P5-­05 ITO/PEDOT:PSS/EML/SPPO13/LiF/Al 4.6 1.0/1.7/—­ 616 1 283 —­ (0.57, 0.41)
a
 Device configuration: ITO, indium tin oxide; PEDOT:PSS, poly(3,4-­ethylenedioxythiophene)-­poly(styrenesulfonate); MoO3, molybdenum trioxide; TAPC, 1,1′-­bis(di4-­
tolylaminophenyl)cyclohexane; TCTA, tris(4-­carbazoyl-­9-­ylphenyl)amine; mCP, 1,3-­bis(N-­carbazolyl)benzene; TmPyPB, 1,3,5-­tri(m-­pyrid-­3-­yl-­phenyl)benzene; TPBi,
2,2′,2″-­(1,3,5-­benzinetriyl)-­tris(1-­phenyl-­1hbenzimidazole); DPEPO, bis[2-­(diphenylphosphino)phenyl] ether oxide; Cs2CO3, caesium carbonate; TSPO1, diphenyl-­4-­
triphenylsilylphenyl-­phosphine oxide; LiF, lithium fluoride.
b
 Von, turn-­on voltage at 1 cd/m2
c
 ηext,max, maximum external quantum efficiency; ηC,max, maximum current efficiency, ηP,max, maximum power efficiency
d
 λEL, electroluminescence emission peak
e
 Lmax, maximum luminance
f
 RO, efficiency roll-­off from the maximum value to that at 1000 cd/m2
g
 CIE, Commission Internationale de I’Eclairage coordinates.
230 8  Aggregation-­induced Delayed Fluorescence

luminogens (DMF-­BP-­DMAC and DPF-­BP-­DMAC) with an asymmetric structure were further


reported [28], in which the twisted electron-­donating group DMAC was used as a donor to form a
major D-­A structure with the electron-­accepting carbonyl group, and 9,9-­dimethylfluorene (DMF)
and 9,9-­diphenylfluorene (DPF) served as functional groups to depress the intermolecular π–π
stacking, and increase carrier transport ability. Both the luminogens showed AIDF property, and
their ΔEST values of the neat films were 0.14 and 0.09 eV, respectively, which were small enough
for an effective RISC process. In view of the excellent solid state PL efficiencies, prominent delayed
fluorescence as well as good thermal stability, DMF-­BP-­DMAC and DPF-­BP-­DMAC were used to
fabricate nondoped OLEDs to investigate their EL properties. The devices of ITO/HATCN (5 nm)/
NPB (30 nm)/mCP (10 nm)/EML (30 nm)/TPBi (50 nm)/LiF (1 nm)/Al were fabricated. It was
worth noting that the devices based on DPF-­BP-­DMAC turned on at low voltages of 2.8 V and emit-
ted strong green light at 534 nm with the maximum L, ηC, ηP, and ηext of 50 170 cd/m2, 43.8 cd/A,
40.2  lm/W, and 13.2%, respectively. Moreover, they retained high ηext of 12.6% at 1000 cd/m2,
revealing a very small efficiency roll-­off.
Doping technique suffers from bimolecular quenching process (DET) at high voltage [29], before
the high-­concentration excitons is formed in host diffuse to guest molecules for radiative decay and
results in sharp efficiency loss. To address this issue, a new concept of integrating delayed fluores-
cence moiety within host materials was proposed  [30]. A series of novel luminogens, namely,
DCB-­BP-­PXZ, CBP-­BP-­PXZ, mCP-­BP-­PXZ, and mCBP-­BP-­PXZ, were synthesized by grafting
4-­(phenoxazin-­10-­yl) benzoyl, a moiety with AIDF feature, to common host materials, including
1,4-­di(carbazol-­9-­yl) benzene (DCB), 4,4′-­bis(carbazol-­9-­yl)biphenyl (CBP), 1,3-­bis(carbazol-­9-­yl)
benzene (mCP), and 3,3′-­di(car-­bazol-­9-­yl)biphenyl (mCBP). All of these luminogens possessed
highly twisted geometry, which facilitated the separation of HOMO and LUMO. Their neat films
showed strong PL peaks at 530–532 nm with high ФPL values of 66.0–71.6%, and had small ΔEST
values (0.016–0.024 eV) to realize fast RISC process with high kRISC values (5.0–6.0 × 105 s−1) at
room temperature. The nondoped OLEDs with a three-­layer configuration of ITO/TAPC (25 nm)/
EML (35  nm)/TmPyPB (55  m)/LiF (1  nm)/Al were prepared by vacuum deposition. All of the
devices turned on at a low voltage of 2.5 V with yellow emission. Remarkably, the maximum ηC, ηP,
and ηext of DCB–BP–PXZ were up to 72.9 cd/A, 81.8 lm/W, and 22.6%, respectively, and the effi-
ciency roll-­off was only 11.4% at 5000 cd/m2, which were the most efficient among nondoped
OLEDs reported so far (Figure 8.5). These results eloquently demonstrated the feasibility and ver-
satility of the presented molecular design strategy of integrating DF moiety within the host materi-
als, to create robust luminescent materials for nondoped OLED with advanced efficiency and
stability.
Solution-­processed film preparation techniques possess many advantages compared to vacuum
deposition, such as feasible to manufacture large area OLEDs, flexible substrates, and better con-
trol of the doping process [31]. Recently, two new organic small molecules, CC6-­DBP-­PXZ, and
CC6-­DBP-­DMAC (Figure 8.6), consisting of electron-­withdrawing benzoyl and electron-­donating
9-­hexylcarbazole and PXZ (or DMAC) moieties  [32], were developed for efficient solution-­
processed OLEDs. They were weak emitters in solution, but showed strong emission in spin-­coated
neat films with greatly enhanced delayed fluorescence. Both the compounds showed excellent
thermal and morphological stabilities, and good film-­forming abilities that were favorable to
solution-­proceeded OLEDs. The introduction of flexible alkyl chains (CC6) can improve the solu-
bility and film-­formation ability of the molecules by solution-­process method, but has little impact
on the configurations of the two compounds. CC6-­DBP-­PXZ and CC6-­DBP-­DMAC exhibited
strong emissions at 547 and 510 nm and the ФPL values in neat films were 38.3 and 59.5%, respec-
tively. The nondoped OLEDs with a configuration of ITO/PEDOT:PSS (50  nm)/PVK (30  nm)/
8.2 ­Novel Aggregation-­induced Delayed Fluorescence Luminogen 231

(a) (b)
500
105 DCB-BP-PXZ
CPB-BP-PXZ 102 103
mPC-BP-PXZ
mCBP-BP-PXZ 400

Current density (mA/cm2)


104

Current efficiency (cd/A)

Power efficiency (Im/W)


Luminance (cd/m2)

102
300 101
103
DCB-BP-PXZ
101
200 CPB-BP-PXZ
102 mCP-BP-PXZ
100 mCBP-BP-PXZ

100 100
101
380 480 580 680 780
Wavelength (nm)
100 0 10–1 10–1
2 4 6 8 100 101 102 103 104 105
Voltage (V) Luminance (cd/m2)

(c)
1.0

500 cd/m2
0.9
Relative luminance (l/l0)

1000 cd/m2
2000 cd/m2
0.8

0.7

0.6

0.5
0 20 40 60 80 100

Time (hours)

Figure 8.5  (a) Luminance–current density–voltage and (b) current efficiency–power efficiency–luminance


curves of nondoped OLEDs. Inset: EL spectra at luminance of 5000 cd/m2. (c) Relative luminance–time
curves of CBP-­BP-­PXZ-­based nondoped device at different initial luminance.

EML/TmPyPB (40 nm)/LiF (1 nm)/Al were fabricated by solution-­processed technique, in which


the neat films of CC6-­DBP-­PXZ and CC6-­DBP-­DMAC prepared by spin-­coating served as EML. The
device based on CC6-­DBP-­DMAC radiated green light at 505 nm, with the turn-­on voltages of 4.2
V and the maximum L, ηC, ηP, and ηext of 14 366 cd/m2, 25.08 cd/A, 11.25 lm/W, and 9.02%, respec-
tively. The device based on CC6-­DBP-­PXZ showed yellow light with the emission peak at 568 nm,
with a smaller turn-­on voltage of 2.9 and the maximum L, ηC, ηP, and ηext of 30 644 cd/m2, 22.23
cd/A, 16.11 lm/W, and 7.73%, respectively. In addition, these nondoped OLEDs showed small effi-
ciency roll-­off of 2.8–9.6% at 5000 cd/m2, which was much better than most solution-­processed
OLEDs based on TADF emitters (Figure 8.7). Moreover, both the compounds could also be used to
fabricate doped OLEDs. They also behaved efficiently as dopants in a CBP host, endowing the
solution-­processed doped OLEDs with a high maximum ηext of up to 12.1%. These solution-­
processed doped OLEDs present extremely small efficiency roll-­off of down to 0.08%, demonstrat-
ing their outstanding efficiency stability. These results indicate the great potential of small
molecules with AIDF for the fabrication of high-­performance solution-­processed OLEDs.
232 8  Aggregation-­induced Delayed Fluorescence

O O O

N N N
O N
O N
N N
N N
O

DCB-BP-PXZ CBP-BP-PXZ mCBP-BP-PXZ

O N O O
N
O O
O
N

N N N N N
N C6H13
O C6H13
O
mCP-BP-PXZ CC6-DBP-PXZ CC6-DBP-DMAC
O
O O
N
C6H13 N N
N
N S
N
N G6H13 N
N N
N N
N
C6H13
C6H13
TATC-BP TATP-BP OPPh

O O
O

N N
N
S
S
S
OPNa OPPy
OPAn

Figure 8.6  Molecule structures of aggregation-­induced delayed fluorescence materials with carbonyl


group as an electron acceptor.

It is obvious that the molecules selecting benzoyl as electron-­withdrawing groups exhibit excel-
lent performance when applied to OLEDs. Not only high maximum ηext but also negligible effi-
ciency roll-­off was also obtained. Many research groups have invested their time and effort in
expanding this system, and gained decent achievements. Wang and coworkers [33] designed two
symmetric molecules with AIDF property, TATC-­BP and TATP-­BP, which consisted of carbonyl as
electron-­accepting group and triazatruxene as the electron-­donating groups. The difference
between the compounds was that the substituents on the triazatruxene group for TATC-­BP was
hexyl and for TATP-­BP was phenyl, respectively. Compared to TATC-­BP, TATP-­BP with four phe-
nyl substituents had a longer distance between triplet excitons, which contributed to relieve the
triplet–triplet annihilation. The PL peaks in neat films for TATC-­BP and TATP-­BP were at 524 and
520 nm, with ΦPL values of 22.0 and 24.2%, respectively. Interestingly, mechanochromic lumines-
cence (MCL) phenomenon was found in these compounds. By grinding the crystal, the emission
color was changed from blue-­green (483 nm) to yellow (542 nm) for TATC-­BP, and from green
(498 nm) to yellow (534 nm) for TATP-­BP, while the disparate emission colors stayed the same over
seven days. The emission color could change back to original after fuming in dichloromethane
vapor for 12 hours, demonstrating that the MCL property of these compounds was stable.
Whereafter, the nondoped devices were fabricated by solution-­processing with the configuration
of ITO/PEDOT:PSS (25 nm)/EML (25 nm)/TmPyPB (55 nm)/LiF (1 nm)/Al (150 nm), in which
the EML were the neat films of these compounds. As for nondoped device I (TATC-­BP) and device
II (TATP-­BP), all of them exhibited yellow-­green emission with the EL peaks at 549 and 541 nm.
Meanwhile, the maximum ηext and ηC of device I and device II were 5.9%, 17.8 cd/A, and 6.0%, 18.9
cd/A, respectively. It was worth noting that the efficiency roll-­off of device II was only 3.3% at 1000
8.2 ­Novel Aggregation-­induced Delayed Fluorescence Luminogen 233

(a) (b)
300
Device CIE (x,y)
1A (0.272,0.503)
1B (0.265,0.501)
Device 1A
104
1C (0.454,0.523)

Current density (mA/cm2)


1D (0.459,0.518) Device 1B
Device 1C
EL intensity (a.u.)

Device 1B 200

Luminance (cd/m2)
Device 1D 103

102
Device 1C
100

101

0 100
400 500 600 700 0 4 8 12

Wavelength (nm) Voltage (V)

(c)
102 102
Device 1A

External quantum efficiency (%)


Device 1B
Device 1C
Current efficiency (cd/A)

Device 1D

101 101

100 100
0 10000 20000 30000
Luminance (cd/m2)

Figure 8.7  (a) EL spectra, (b) current density–luminance–voltage, and (c) current efficiency–external
quantum efficiency–luminance of the nondoped devices. EML = CC6-­DBP-­DMAC (Device 1A); CC6-­DBP-­
DMAC (Device 1B); CC6-­DBP-­PXZ (Device 1C); CC6-­DBP-­PXZ (Device 1D).

cd/m2 that could be ascribed to the shorter delayed lifetime (0.91 μs), weak intermolecular interac-
tion, and long distance between triplet excitons of TATP-­BP (Figure 8.8).
Four asymmetric molecules were synthesized with attaching diverse bulky groups to the
(4-­(10H-­phenothiazin-­10-­yl)phenyl)enzoyl moiety. The bulky groups were phenyl, naphthyl,
anthryl, and pyrenyl, corresponding to these four compounds, OPPh, OPNa, OPAn, and OPPy.
Both of them had a dual-­emission property (Figure 8.9), and the emission peaks were derived from
intramolecular charge transfer (426–459  nm) in a single compound and intermolecular charge
transfer (465–475  nm) in a dimer that was formed through C–H⋯O intermolecular hydrogen
bonding. Meanwhile, the ratio of emission peaks was regulated by attaching bulky groups with
different steric hindrances. As the steric hindrance of these substituents increased (from phenyl to
anthracenyl group), the emission intensity for intermolecular charge transfer displayed a noticea-
ble decrease, which was attributed to the fact that the strength of dimers’ packing tightness was
234 8  Aggregation-­induced Delayed Fluorescence

54°
(a)

65° 66°

TATC-BP TATP-BP

(b)

LUMO

HOMO

(c)

Figure 8.8  (a) Optimized geometries, (b) HOMO and LUMO distributions, and (c) spin density distributions
of TATC-­BP (left) and TATP-­BP (right). Source: Adapted with permission from Ref. [33] copyright 2018 Royal
Society of Chemistry.

decreased as the number of aromatic rings increased (Figure 8.9). As a result, the emission colors
of crystalline powders for OPPh, OPNa, OPAn, and OPPy were blue, bluish-­white, white, and yel-
low, with CIEx,y coordinates of (0.22, 0.23), (0.25, 0.24), (0.30, 0.25), and (0.49, 0.51), respectively. It
was worth noting that OPNa not only achieved pure white with CIEx,y coordinates of (0.32, 0.33)
by grinding, but also had an excellent AIDF property. These results demonstrated that adjusting
the ratios of intramolecular charge transfer and intermolecular charge transfer was a feasible
method to obtain pure white emission in a single compound.
In addition to these luminogens with carbonyl core, some other AIDF molecules with highly
twisted connection between donors and acceptors (DPS-­PXZ, DBTO-­PXZ, DPS-­PTZ, and DBTO-­
PTZ) were also reported, in which PXZ and PTZ served as the donors and diphenylsulphone (DPS)
and dibenzothiophene-­S,S-­dioxide (DBTO) served as the acceptors (Figure 8.10) [35]. The values
of ФPL for DPS-­PXZ, DBTO-­PXZ, DPS-­PTZ, and DBTO-­PTZ were 3.1, 1.0, 1.6, and 1.2% in dilute
THF solutions, and increased to 16.3, 6.7, 20.3, and 10.7% in solid state, respectively. At the same
(a) (b)

OPPh Grinding

OPNa

OPAn

OPPy
400 450 500 550 600 650 700 750 800

(c)
C–H∙∙∙O : 2.37Å
π∙∙∙π : 3.14Å
C–H∙∙∙O : 2.58Å

C–H∙∙∙π : 2.46Å:2.98Å

Figure 8.9  (a) Emission spectra of microcrystals of OPPh, OPNa, OPAn, and OPPy with molecular alignments of interlocked, opposite, and linear; (b)
fluorescent images of these molecular under grinding; (c) molecular packing of OPPh, OPNa, and OPPy in single-­crystal structures (C–H⋯π or C–H⋯O
weak hydrogen bonds and π–π interactions are shown). Source: Adapted with permission from Ref. [34] copyright 2019 Royal Society of Chemistry.
236 8  Aggregation-­induced Delayed Fluorescence

O O O O O O O O
S S S S

N N N N N N N
N
O O O S
S S S
O

DPS-PXZ DBTO-PXZ DPS-PTZ DBTO-PTZ

O O
O O
S
S O O
S N
N N
N N N
N N N
O
S S O N

4, 4-CzSPz PTSOPO PXZ2PTO 34AcCz-PM

N N
N
N N N N CN O
O
N N N
S N N N N
N
O O

34AcCz-Trz Pz-DMNI DMAC-DMNI


ND-AC CND-AC

O O

O O
O N N
O N N N
N CF3
CF3 CF3 CF3

O CN
O CN CN CN

Cz-AQ TPA-AQ NCFCz NCFMCz2 NCFMCz1 NCFTCz

Figure 8.10  Molecule structures of aggregation-­induced delayed fluorescence materials with heterocyclic


group as electron acceptor.

time, they exhibited apparent delayed components with lifetimes of 0.8, 0.7, 52.3, and 2.3 μs,
respectively. These results demonstrated that they had AIE property and delayed fluorescence.
Their ΔEST values were in the range of 0.06–0.26 eV, which enabled the RISC process to occur.
Their twisted conformations were also helpful to inhibit strong π–π intermolecular interactions,
and thus reduced nonradiative decay in the aggregated state. All data suggested that they might be
promising candidates as light emitters for nondoped OLEDs.
At the same time, some other research groups also developed AIDF materials with the diphenyl-­
sulfone group as core, and achieved high EL efficiency. Zhao and coworker [36] designed an asym-
metric D-­A-­D′-­type compound, 4,4-­CZSPz, with the diphenyl-­sulfone group as electron acceptor,
phenothiazine, and carbazole groups as electron donors. The ΔEST value of this compound was
0.25 eV, which was small enough to go through the RISC process, and the transient decay spectra
of the neat film exhibited a distinct delayed fluorescence at room temperature (Figure  8.11).
Thanks to the AIDF property of 4,4-­CZSPz, in a vacuum, the ΦPL in neat film reached up to 97.3%
with the emission peak at 530 nm, indicating it was an excellent potential emitter for nondoped
OLEDs. This value was higher than this of the symmetric compounds with two phenothiazine
groups or two carbazole groups as donors ( 52.8%), demonstrating the design strategy of exploiting
asymmetric efficient AIDF emitters was viable. In solution, as the solvent polarity increased, the
emission peak exhibited a remarkable red shift. The relation between the stokes shift (νa−νf ) and
8.2 ­Novel Aggregation-­induced Delayed Fluorescence Luminogen 237

(a) (b)
1
17K
External quantum efficiency (%)

77K

Normalized intensity (a.u.)


101 127K
O O
S 177K
N N 227K
100 S 4,4-CzSPz 0.1 277K
298K

10–1
Doped OLED (EQEm = 26.2%)
Non-doped OLED (EQEm = 20.7%)
10–2 0.01

101 102 103 104 4000 5000 6000 7000


Luminance (cd/m2) Time (µs)

Figure 8.11  (a) Transient PL decay curves of 4,4-­CZSPz film carried out from 17 to 298 K; (b) EQE−
luminance curves of nondoped and doped devices for 4,4-­CZSPz. Source: Adapted with permission from
Ref. [36] copyright 2019 American Chemical Society.

the solvent polarity (f) was linear with a large slope of 10 544 according to the Lippert−Mataga
model, suggesting that the emission of the molecules originated from a strong ICT process. Both
the doped and nondoped OLEDs were fabricated with the configuration of ITO/PEDOT: PSS
(30 nm)/mCP (20 nm)/4,4CZSPz (or 10 wt% 4,4-­CZSPz:CBP, for the doped device) (30 nm)/TPBI
(40 nm)/LiF (1 nm)/Al (100 nm). All of them displayed green emission peaks at 518 nm for doped
device and 526  nm for nondoped device. Their efficiencies were outstanding among the green
emission devices, with maximum ηext, ηC, ηP of 26.2%, 85.1 cd/A, and 53.5 lm/W for doped device
and 20.7%, 61.2 cd/A, and 38.4 lm/W for nondoped device. Meanwhile, the molecule had mecha-
noluminescence property that was advantageous to broaden its application.
Lee et al. [37] also synthesized an asymmetric AIDF emitter, PTSOPO, which used diphenyl-­
sulfone as an electron acceptor, and phenothiazine and phenoxazine as electron donors. The value
of ΔEST was only 0.09 eV, which benefited from the efficient separation of HOMO and LUMO and
contributed to facilitating the RISC process. In the neat film, the ΦPL of PTSOPO was relatively
high among AIDF-­based materials, up to 80%. For comparison, they also synthesized a symmetric
compound PTSOPT. However, this compound did not exhibit a significant AIDF property, as the
ΔEST and ΦPL were only 0.41 eV and 14%, respectively. As for PTSOPO, the maximum ηext values of
nondoped device (17.0%), and doped device (17.7%) were almost equivalent that could be attrib-
uted to AIE property. At the same time, the maximum ηext of the devices based on the PTSOPO was
higher than that of PTSOPT device. As mentioned, in terms of photophysical property and electri-
cal performance, the asymmetric compound performed much better than the symmetric com-
pound, which verified the feasibility of the design method to develop asymmetric compounds
once again.
In consideration of the outstanding performance of AIDF materials in OLEDs, more novel
molecular systems are being developed. Recently, a new electron-­withdrawing group  [38] was
designed, namely, 10-­phenyl-­10H-­phenothiazine-­5,5-­dioxide (2PTO), by oxidizing the S atom of
phenothiazine and attaching a phenyl to the N atom, and this group had a distorted construction
that contributed to develop efficient AIDF materials. Then, they connected the 2PTO with pheno-
thiazine, constructing a D-­A type compound PXZ2PTO. As expected, PXZ2PTO had a highly
twisted configuration (Figure 8.12), for which the dihedral angles were 70° between phenothiazine-­
oxidic moiety and phenothiazine, and 85.5° between phenothiazine-­oxidic moiety and benzene,
238 8  Aggregation-­induced Delayed Fluorescence

18
16.4% Nondoped devices

Highly twisted
12
EQE (%)

85.5°
AIDF
70.0°

Efficient OLED
PXZ2PTO
0
1 10 100 1000
Luminance (cd/m2)

Figure 8.12  Molecular structure of PXZ2PTO and EQE–luminance characteristics of the nondoped device
based on PXZ2PTO. Source: Adapted with permission from Ref. [38] copyright 2018 Royal Society of Chemistry.

respectively. Therefore, the overlap between HOMO and LUMO was small, which was conducive
to obtaining a low ∆EST value (0.03 eV). Meanwhile, this compound had relatively high ΦPL values
of 61.54 and 68.75%, and short delayed lifetimes of 2.49 and 2.61 μs for neat film and doped film,
respectively. With the good photophysical properties, both the nondoped and doped devices dis-
played brilliant EL performance, and the efficiency of nondoped device was comparable with that
of doped device. The nondoped and doped devices displayed green light with the emission peaks
at 504 and 500 nm, respectively. The maximum ηext, ηC, and ηP were found to be 16.4%, 44.9 cd/A,
and 32.0  lm/W for nondoped device, and 16.3%, 43.8 cd/A, and 35.2  lm/W for doped device.
Moreover, the efficiency roll-­off of both the nondoped and doped devices were small that verified
the potential of AIDF materials to receive stable and efficient OLEDs once again.
Not only novel electron acceptors were designed to develop AIDF emitters, but also some inno-
vative electron donors were also reported. Zhang and coworker [39] synthesized an ­acridine–carbazole
hybrid donor, namely, 34AcCZ, which was taken shape with two complex donors, DMAc and
CZ. Afterward, 34AcCZ-­PM and 34AcCZ-­Trz were prepared by using 34AcCZ as an electron-­
donating group, pyrimidine and triazine as electron-­accepting groups, and both of them exhibited
prominent AIDF property. As shown in their crystalline-­packing diagrams (Figure  8.13), there
were no π–π interactions between two adjacent compounds, but many C–H⋯π interactions were
found, which led to a more rigid molecular structure, and thus, suppressed nonradiative attenua-
tion. Owing to the weaker electron-­accepting ability of pyrimidine, the LUMO energy level of
34AcCZ-­PM (−2.53 eV) was shallower as compared to 34AcCZ-­Trz (−2.68 eV). In neat films, the
emission peaks were 538 nm for 34AcCZ-­PM and 556 nm for 34AcCZ-­Trz. Meanwhile, on account
of the rigid structure and the more prominent AIE property of 34AcCZ-­PM (Figure 8.13), the neat
film of 34AcCZ-­PM (67%) obtained a much higher ΦPL value as compared to 34AcCZ-­Trz (42%).
The nondoped and doped devices were fabricated with the structure of ITO/MoO3/TAPC/TCTA/
EML/TmPyPB/LiF/Al. The nondoped device based on 34AcCZ-­PM (3.10 V) had a lower turn-­on
voltage as compared to 34AcCZ-­Trz (3.35 V), which can be attributed to the smaller LUMO energy
barrier between 34AcCZ-­PM and electron transport layer. Compared to the devices based on
34AcCZ-­Trz, the devices based on 34AcCZ-­PM had superior EL performance. For the devices
8.2 ­Novel Aggregation-­induced Delayed Fluorescence Luminogen 239

(a)

d1 = 3.331 Å, d2 = 3.335 Å,
d1 = 3.277 Å, d2 = 3.593 Å, d3 = 2.697 Å, d4 = 3.595 Å, d5 = 3.662 Å; d3 = 3061 Å, d4 = 3.506 Å, d5 = 2.517 Å

(b) (c)
12 34AcCz-PM 34AcCz-Trz 12 A B
fw (vol%) C D
20
10 90% 10
80%
Counts (x104)

8 70% 8 15

EQE (%)
60%
6 50% 6
40% 10
4 30% 4
20%
10% 5
2 2

0 0 0
500 600 700 500 600 700 100 101 102 103 104
Wavelength (nm) Wavelength (nm) 2
Luminance (cd/m )

Figure 8.13  (a) (Left) Packing diagrams of 34AcCZ-­PM (CCDC 1877512) and (right) 34AcCZ-­Trz (1877513);
(b) PL spectra for 7.35 × 10−3 M 34AcCZ-­PM and 34AcCZ-­Trz in THF/water mixtures of different fw; (c) EQE as
a function of the luminance. Source: Adapted with permission from Ref. [39] copyright 2019 Royal Society of
Chemistry.

based on 34AcCZ-­PM, the maximum ηext, ηC, and ηP were 14.1%, 45.2 cd/A and 40.0 lm/W for non-
doped device, and 22.6%, 73.3 cd/A, and 72.4 lm/W for doped device. It was worth noting that the
efficiency roll-­off of the doped device was negligible with the high EL efficiencies of 65.5 cd/A,
27.9 lm/W, and 20.1% at an enormously higher luminance of 5000 cd/m2. This work verified that
the AIDF materials exhibited an excellent EL performance especially in terms of reducing the
efficiency roll-­off.
Zhou and co-­workers  [40] synthesized two D–A type AIDF emitters, namely, ND-­AC and
CND-­AC, selecting naphthyridine (ND) and cyano-­naphthyridine (CND) as the electron acceptor
and acridine as the electron donor. Due to the presence of ortho-­substituted cyano, the dihedral
angle between the donor and the acceptor units of CND-­AC was 85.8°, and it was larger than that
of ND-­AC (80.7°). However, compared with ND-­AC, the more distorted configuration of CND-­AC
led to a much lower oscillator strength (f: 0.0001 of CND-­AC; 0.0011 of ND-­AC), which further
led to a lower ΦPL value. The ΦPL of ND-­AC was 55%, but that of CND-­AC was only 5%. This result
demonstrated the significance of the trade-­off between spatial configuration and oscillator
strength. Then, the authors explained deeply about why the ΦPL was so different by using the rate
constants of radiative/nonradiative decay (kr/knr). As for ND-­AC, the kr/knr was 1.22, but for
CND-­AC, it was only 0.05, which demonstrated that the nonradiative transition was the major
decay process for CND-­AC. In the neat films, the emission peaks of ND-­AC and CND-­AC were
located at 551 and 632 nm, respectively. The large red shift of emission peaks could be attributed
to the introduction of cyano, which could enhance the electron-­withdrawing ability of the
240 8  Aggregation-­induced Delayed Fluorescence

acceptor group, and lead to a deeper LUMO energy level. Subsequently, doped OLEDs were fab-
ricated with the structure of ITO/TAPC (30 nm)/TCTA (5 nm)/ND-­AC (9 wt%):CBP (15 nm) or
CND-­AC (1.5 wt%):CBP (15 nm) (EML)/TmPyPB (65 nm)/LiF(1 nm)/Al(100 nm). The devices
based on ND-­AC and CND-­AC provided maximum ηext values of 16.8 and 8.4% with EL emission
peaks at 542 and 588 nm, respectively. Owing to the relatively larger ΦPL of ND-­AC than CND-­AC,
the nondoped device based on ND-­AC was also fabricated with the structure of ITO/TAPC
(30 nm)/TCTA (5 nm)/ND-­AC (15 nm)/TmPyPB (65 nm)/LiF (1 nm)/Al (100 nm), and possessed
a satisfactory EL performance with the EL peak at 558 nm, and the maximum ηext, ηC and ηP of
12.0%, 38.5 cd/A and 30.2 lm/W, respectively (Figure 8.14).
As for red fluorescent material, the energy gap between HOMO and LUMO is smaller, resulting
in a larger kIC as compared to blue, green, and yellow ones. Therefore, it is difficult to achieve high-­
external quantum efficiency for red emitters. However, in the recent years, the field of red emitters
had made significant progress, and some red AIDF materials with outstanding performance had
been developed. Chi and coworkers [41] prepared two novel D-­π-­A-­type compounds, PZ-­DMNI
and DMAC-­DMNI, in which the PTZ and DMAC served as electron donors, and 1,8-­naphthalimide
(NI), which had a strong electron-­withdrawing ability, acted as electron acceptor, and the methyl-
phenyl (DM) group was selected as a π bridge (Figure 8.15). Through introducing the DM group,
the steric hindrance and the rigidity of these compounds were enhanced. Therefore, the intramo-
lecular vibration was restrained, leading to suppressed nonradiative transition, and making more
excitons attenuate through the radiative transition. The configurations of PZ-­DMNI and DMAC-­
DMNI were twisted, and there was almost no overlap between HOMO and LUMO orbitals, result-
ing in small ΔEST values of 0.05 eV for PZ-­DMNI and 0.13 eV for DMAC-­DMNI, which was
favorable to the RISC process. In the neat film, these molecules had excellent ΦPL values of 55% for
Pz-­DMNI and 39% for DMAC-­DMNI, with emission peaks at 624 and 595  nm, respectively.
Moreover, the doped devices exhibited good performance with the structure of ITO/PEDOT:PSS

Non-doped device Max EQE = 12.0%


External quantum efficiency (%)

101

TADF AIE

0 95%
100 80.7°

10–1

10–1 100 101 102 103


Luminance (cd/m2)

Figure 8.14  EQE–luminance characteristics of the nondoped device based on ND-­AC. Inset: molecule
structure and the photos in THF/water mixtures with different water fractions (fw = 0% and 90%) taken
under 365 nm excitation. Source: Adapted with permission from Ref. [40] copyright 2019 Royal Society of
Chemistry.
8.2 ­Novel Aggregation-­induced Delayed Fluorescence Luminogen 241

105
λEL = 635 nm

O
104
N N S
PL intensity (a.u.)

103

400 500 600 700 800


Wavelength (nm)

102

AIE + TADF
101
0 100 200 300 400 500
Time (μs)

Figure 8.15  PL decay curve of the neat films of Pz-­DMNI. Inset in plot: molecule structure, EL spectra
at 10 V and photos in THF/water mixtures with different water fractions (fw = 0 and 99%) taken under
365 nm excitation. Source: Adapted with permission from Ref. [41] copyright 2019 Wiley-­VCH.

(40 nm)/mCP (20 nm)/CBP:red TADF emitter (15 nm, 12% doping)/TPBi (40 nm)/LiF (1 nm)/Al


(120 nm), and the maximum ηext was 7.13 and 5.38% for PTZ-­DMNI and DMAC-­DMNI, respec-
tively. At the same time, the doping concentration in emitting layer was higher than the common
doping concentrations of doped OLEDs. The change of EL spectra of these doped devices was
extremely small at different voltages, demonstrating that they can provide high-­color stability.
Sun and coworkers [42] developed two D–A type compounds with AIDF property, namely, CZ–
AQ and TPA–AQ, which consisted of anthraquinone as electron acceptor, and carbazole and triph-
enylamine as electron donors. Compared with CZ–AQ, TPA–AQ had a longer π-­conjugation system
that favored for the generation of red emission. The values of delayed fluorescence quantum yields
(Φd) were 16% for CZ–AQ and 25% for TPA–AQ. The much larger Φd of TPA–AQ could be ascribed
to the more efficient RISC process, which benefited from the smaller ΔEST (0.05 eV for TPA–AQ;
0.11 eV for CZ–AQ). Meanwhile, the kRISC was 2.62 × 105 S−1 for CZ–AQ and 7.34 × 105 S−1 for
TPA–AQ, and it was consistent with the Φd values. Then, the doped devices were fabricated by
solution-­processing method, with the structure of ITO/PEDOT:PSS (40 nm)/CZ–AQ (10 wt%):CBP
(40 nm) or TPA–AQ (10 wt%):CBP (40 nm)/TPBI (30 nm)/Cs2CO3(2 nm)/Al (120 nm). The device
of CZ–AQ showed orange light at 572  nm, while the device of TPA–AQ displayed red light at
612 nm. Meanwhile, the devices had low turn-­on voltages of 3.5 V for CZ–AQ and 3.8 V for TPA–
AQ. The device based on TPA–AQ had better EL performance, with high maximum ηext, ηC, and L
of 7.8%, 10.6 cd/A, and 2200 cd/m, respectively.
Three primary colors (red, green, and blue) are significant to flat-­panel displays. Nevertheless,
the exploitation of blue light-­emitting materials with high efficiency and pure emission color is
still challenging, due to the large energy gap between HOMO and LUMO, which leads to a difficult
charge injection from adjacent charge-­transporting layers to light-­emitting layers. Therefore, the
performance of blue device is not satisfactory. Recently, some blue-­emitting materials with
242 8  Aggregation-­induced Delayed Fluorescence

excellent performance had been developed. Gintare and coworker [43] synthesized four D–A type
compounds, NCFCZ, NCFMCZ1, NCFMCZ2, and NCFTCZ, in which trifluoromethyl-­substituted
benzonitrile was selected as electron acceptor, and the carbazole derivatives acted as electron
donors. Among them, the compounds with substituents on the carbazole (NCFMCZ1, NCFMCZ2,
and NCFTCZ) had AIDF property. In neat films, the emission of NCFTCz was blue (459 nm) with
a ΦPL of 52.53%, while the emission peaks of 9NCFMCz1 and NCFMCz2  were red-­shifted to
513 nm with the ΦPL values of 52.17 and 16.95%, respectively. Time-­of-­flight technique was used to
detect the charge mobility of these compounds, and the results indicated that NCFMCz1 and
NCFTCz had the bipolar charge transfer capability. Owing to the outstanding performance of
NCFMCz1 and NCFTCz, the non-­doped devices using NCFMCz1 and NCFTCz as light-­emitting
layers were fabricated with a configuration of ITO/MoO3 (2 nm)/TCTA (40 nm)/mCP (8 nm)/EML
(32 nm)/TSPO1 (8 nm)/TPBi (40nm)/Ca/Al. The turn-­on voltages were 4.4 V for NCFTCz and 3.9
V for NCFMCz1, and the maximum ηext and EL peak were found to be 3.2% and 466 nm, and 7.2%
and 503 nm for NCFTCz and NCFMCz1, respectively.
Recently, two new luminogens, TRZ-­HPB-­PXZ, TRZ-­HPB-­DMAC, which comprised of electron
donors (acridine and phenoxazine) and acceptor (triazine) bridged by the through space conju-
gated hexaphenylbenzene (HPB) were prepared [44]. Both the molecules showed apparent AIDF
property. They exhibited weak emission with negligible-­delayed component in solutions, but
strong fluorescence with greatly enhanced delayed component upon aggregate formation.
According to crystal and electronic structures, it was found that the peripheral phenyls in HPB
were closely aligned in a propeller-­like fashion, rendering efficient through-­space charge transfer
between donor and acceptor moieties to balance the ΔEST and ΦPL. The screwy architecture was of
great benefit to the separation of frontier orbitals, resulting in a small ΔEST of 0.0017 and 0.0013 eV
for TRZ-­HPB-­PXZ and TRZ-­HPB-­DMAC, respectively, which were small enough for RISC process.
Both the luminogens showed low ΦPL values of 5.5 and 9.1% in solutions, but high ΦPL values of
61.5 and 51.8% in neat films, respectively. Owing to their excellent solid state PL efficiencies, mul-
tilayer nondoped devices with the configuration of ITO/HATCN (5  nm)/TAPC (20  nm)/TCTA
(5 nm)/EML/TmPyPB (55 nm)/LiF (1 nm)/Al were fabricated, in which the neat films of TRZ-­
HPB-­PXZ and TRZ-­HPB-­DMAC served as the emitter layer. Device I (TRZ-­HPB-­PXZ) and device
II (TRZ-­HPB-­DMAC) turned on at low voltages of 2.5 and 3.1 V, respectively. Device I emitted
bright yellow light at 544 nm and the maximum L, ηc, ηp, and ηext were 40382 cd/m2, 41.2 cd/A,
44.9  lm/W, and 12.7%, respectively. The ηc, ηp, and ηext remained as 40.1 cd/A, 31.5  lm/W, and
12.3%, respectively, when the luminance increased to 1000 cd/m2 (Figure 8.16). Device II emitted
bright at 521 nm, and the maximum ηext was 6.5%, which was lower than those of TRZ-­HPB-­PXZ
due to the lower ΦPL and the larger ΔEST of TRZ-­HPB-­DMAC. Both the molecules showed eminent
AIDF property, which was conductive to suppressing emission quenching, exciton annihilation in
neat films as well as small efficiency roll-­off for nondoped devices.
In addition to small molecules with AIDF property, some AIDF dendritic molecules and poly-
mers had also been developed and exhibited excellent performance in solution-­processed OLEDs.
Wang and coworkers [45] synthesized two tortile D-­π-­A-­type compounds, namely, CzTAZPO and
sCzTAZPO. Both the compounds used triazine (TRZ) as electron-­withdrawing group, carbazole
dendrite as electron-­donating group, and phenyl as the π bridge. To improve their electron trans-
porting ability, diphenylphosphine oxide groups were attached to TRZ. The difference between the
two compounds was that CzTAZPO had two electron-­donating groups attached to TRZ, while sCz-
TAZPO had only one. With the multiplication of the number of carbazole dendrites, the increase
of fluorescence oscillator strength (f) and the decrease of ΔEST were realized simultaneously. The f
and ΔEST were 0.236 and 0.084 eV for CzTAZPO, and 0.159 and 0.104 eV for sCzTAZPO,
8.2 ­Novel Aggregation-­induced Delayed Fluorescence Luminogen 243

(a)

–1.83 eV
–1.86 eV
LUMO

∆EST = 0.0017 eV ∆EST = 0.0013 eV

Eg = 2.59 eV Eg = 2.87 eV

HOMO

–4.45 eV –4.70 eV

TRZ-HPB-PXZ TRZ-HPB-DMAC

(b) 102
External quantum efficiencies (%)

101

100
TRZ-HPB-PXZ
TRZ-HPB-DMAC

10–1
102 103 104
Luminance (cd/m2)

Figure 8.16  (a) Spatial distributions of HOMO and LUMO of the HPB-­based molecules, calculated by DFT/
TDDFT method; (b) external quantum efficiency-­luminance of the nondoped OLEDs. Inset: EL spectra at
luminance of 1000 cd/m2.

respectively. With the larger f and smaller ΔEST, CzTAZPO exhibited better performance in both PL
and EL properties than sCzTAZPO. CzTAZPO had an excellent ΦPL of up to 71%. Meanwhile, with
a smaller value of ΔEST, CzTAZPO gained a larger kRISC, which was advantageous for the genera-
tion of delayed fluorescence. Then, three-­layered nondoped devices were fabricated by solution-­
processing method with the configuration of ITO/PEDOT:PSS (40nm)/CzTAZPO and sCzTAZPO
(50 nm)/TmPyPB (20 nm)/Ca (10 nm)/Al (100 nm), in which emitting layers were the neat films
of the compounds. The EL peak and CIE coordinates were 537 nm and (0.37, 0.36), and 531 nm
and (0.36, 0.36) for CzTAZPO and sCzTAZPO, respectively. Notably, the maximum ηext of CzTAZPO
reached 12.8%, and at 1000 cd/m2, the ηext still remained as 12.6%, corresponding to a small effi-
ciency roll-­off of only 1.8% (Figure 8.17), which could be attributed to the fast RISC process and
efficient radiative decay.
Wang et al. [46] designed two AIDF dendritic compounds, Ac3TRZ3 and TAc3TRZ3, for which
alternately arrayed electron donors (acridan or dendritic triacridan) and electron acceptors (tria-
zine) in the circumjacent of the hexaphenylbenzene core (Figure 8.18). As shown in Figure 8.19,
the HOMOs and LUMOs achieved physical separation, but the spatial distance between the fron-
tier molecular orbitals was short. Therefore, small ΔEST values were achieved (Ac3TRZ3: 0.08 eV;
244 8  Aggregation-­induced Delayed Fluorescence

CzTAZPO
AIDF emitters
Solution process

Ca/Al

TmPyPB CzTAZPO sCzTAZPO


102
CzTAZPO or sCzTAZPO
101 sCzTAZPO

EQE (%)
PEDOT:PSS
100 CzTAZPO
ITO sCzTAZPO
10–1
Glass
10–2
102 103 104
Negligible efficiency roll-off Luminance (cd/A)

Figure 8.17  Device configuration, molecular structures and EQE–luminance characteristics of the


nondoped solution-­processed devices based on CzTAZPO and sCzTAZPO. Source: Adapted with permission
from Ref. [45] copyright 2019 Wiley-­VCH.

N
N N N N N

N N
N N N
O N

N N N N
N N
N N N O
N P
N N N
TRZ-HPB-PXZ TRZ-HPB-DMAC
sCzTAZPO
N

Ac3TRZ3
N N
N N

1-x x n

N
N N
N N
N N N N N
N N
N N
P1
N
N O
N P N N
1-x x n
N N

N
N
N N
N N N
N

N N P2
CzTAZPO

TAc3TRZ3

x n x n 1-x x n
1-x 1-x
CF3 CN N CN

N N N

N N N N N N N N
N

N
P3
CF3 P4 CN P5 CN

Figure 8.18  Molecule structures of aggregation-­induced delayed fluorescence materials with heterocyclic


group as electron acceptor.
8.2 ­Novel Aggregation-­induced Delayed Fluorescence Luminogen 245

20
10−0 TADF AIE
15
PL intensity (a.u.)

10−1
Td = 3.16 μs
10

I/I0
10–2

10–3 5

10–4 0
0 10 20 30 0.0 0.2 0.4 0.6 0.8 1.0
Time (μs) Through-space charge transfer
Water fraction (fw)
hexaarylbenzenes (TSCT-HABs)

Figure 8.19  PL decay curve (left), molecule structure (center) and (C) the relative emission intensity of the
TSCT-­HABs (right), inset: the photos in THF/water mixtures with different water fractions (fw = 0 and 90%)
taken under 365 nm excitation.

TAc3TRZ3: 0.04 eV), while the efficient radiative transition was also obtained through spatial π–π
interaction between the adjacent donor and acceptor. As a result, in the doped films, the ratio of
delayed components reached 77% for Ac3TRZ3 and 87% for TAc3TRZ3, with the ΦPL values of 54
and 63%, respectively. Furthermore, Ac3TRZ3 displayed a green light with a peak at 486 nm, and
due to the stronger electron-­donating ability of dendritic teracridan, the emission peak of
TAc3TRZ3  was red shifted to 508  nm. Then, the doped devices were fabricated by solution-­
processing method with the structure of ITO/PEDOT:PSS (40 nm)/EML (40 nm)/TSPO1 (8 nm)/
TmPyPB (42 nm)/LiF (1 nm)/Al (100 nm), in which the EML was the film with 10 wt% Ac3TRZ3
or 10 wt% TAc3TRZ3 doped in Ac6 (six acridan arrayed around the hexaphenylbenzene) host. The
emission peaks were located at 492  nm for Ac3TRZ3 and 503  nm for TAc3TRZ3, respectively.
Meanwhile, the maximum ηext values were 11.0% for Ac3TRZ3 and 14.2% for TAc3TRZ3, which
were in the front rank among solution-­processed OLEDs and indicated again that TSCT was a
promising strategy to design light-­emitting materials for OLEDs.
On account of the potential advantages of the compounds with TSCT property, a series of poly-
mers were further designed [47], namely, P1, P2, P3, P4, and P5, whose emission color could be
regulated by attaching different electron acceptors into nonconjugated polystyrene backbone
(Figure 8.20). The emission colors of these polymers were ranged from deep-­blue to red, and the
white emission emitting from a single polymer was also obtained. These polymers adopted acridan
(Ac) as electron-­donating group, and the triazine derivatives with different substituents as electron-­
accepting groups, which were 2,4-­diphenyl-­6-­cyclohexyl-­1,3,5-­triazine (TRZ-­Cy),
2,4,6-­triphenyl-­1,3,5-­triazine (TRZ-­H), 2-­phenyl-­4,6-­bis(4-­[trifluoromethyl]phenyl)-­1,3,5-­triazine
(TRZ-­CF3), 4,4’-­(6-­phenyl-­1,3,5-­triazine-­2,4-­diyl)dibenzonitrile (TRZ-­CN), and 5,5’-­(6-­phenyl-­1,3,
5-­triazine-­2,4-­diyl)dipicolinonitrile (TRZ-­PyCN), corresponding to P1, P2, P3, P4, and P5, respec-
tively. As the electron-­accepting ability of triazine derivatives was increased in order of TRZ-­Cy,
TRZ-­H, TRZ-­CF3, TRZ-­CN, and TRZ-­PyCN, the charge transfer effect of the polymers was
enhanced and the Eg between HOMO and LUMO energy level was decreased from 2.95 to 1.90 eV,
and thus, the emission colors were tuned from deep-­blue (453 nm) to red (616 nm) in solid state.
In neat films, the values of ΦPL were 29, 51, 74, 34, and 6, for P1-­05, P2-­05, P3-­05, P4-­05, and P5-­05,
respectively (the mean of PX-­05 was the feed ratio of donor/acceptor was 95 mol%/5 mol%). All of
the polymers were demonstrated to be AIDF materials, especially for P3-­05, whose emission
246 8  Aggregation-­induced Delayed Fluorescence

(a) Through-space charge transfer polymers with full-color emission

1-x x n 1-x x n 1-x x n 1-x x n CF3

N
N
D A N N
N N
N N
N N N N

Through-space
charge transfer (TSCT) P1-05, x = 0 .05 P2-05, x = 0.05 P3-05, x = 0.05 CF3
(full-color emission) P1-50, x = 0.50 P2-50, x = 0.50 P3-50, x = 0.50

1-x x n 1-x x n
Electron transfer CN N CN

LUMO N N
N N
N N N N

hv B G R N

P4-05, x = 0.05 CN P5-05, x = 0.05 CN

HOMO
D
A P1-05 P2-05 P3-05 P4-05 P5-05

(b) Through-space charge transfer polymers with white emission

1-x-y x y n CN

D A1 D A2 N N
N N N N

Blue TSCT Yellow TSCT


emission emission
CN
WP-0.4, x = 0.05, y = 0.004
White emission WP-0.6, x = 0.05, y = 0.006
WP-0.8, x = 0.05, y = 0.008

Figure 8.20  Molecular design and chemical structures of through-­space charge-­transfer polymers with
(a) full-­color and (b) white emission. D: donor, A: acceptor, B: blue, G: green, R: red. Source: Adapted with
permission from Ref. [47] copyright 2019 Wiley-­VCH.

intensity was enhanced by 117 times when the water fraction was increased from 0 to 99%, and the
ratio of delayed component was increased to 91% with a kRISC of 8.5 × 106 S−1. Then, nondoped
devices were fabricated with two different configurations. The devices based on P1-­05 and
P2-­05  were fabricated with the configuration of ITO/PEDOT:PSS (40  nm)/polymer (40  nm)/
TSPO1 (8  nm)/TmPyPB(42  nm)/LiF (1  nm)/Al (100  nm), and the configuration of the devices
based on P3-­05, P4-­05, and P5-­05 was ITO/PEDOT:PSS (40 nm)/polymer (40 nm)/SPPO13 (50 nm)/
LiF (1 nm)/Al (100 nm). The emission peaks and CIEx,y coordinates for P1-­05 to P5-­05 were 455 nm
and (0.17, 0.17), 472 nm and (0.18, 0.27), 525 nm and (0.34, 0.55), 568 nm and (0.46, 0.50), and
 ­Reference 247

616 nm and (0.57, 0.41), respectively. It was noticeable that the device based on P1-­05 with a CIEy
< 0.20 exhibited outstanding EL performance, whose maximum ηext and ηC were 7.1% and 10.6
cd/A, respectively. Meanwhile, the maximum ηext of the device based on P3-­05 was 16.2% and the
ηext maintained in 13.0% at 1000 cd/m2, exhibiting a small efficiency roll-­off. White emissive mate-
rials from the dual-­emission of blue and yellow were developed by introducing two kinds of
electron-­accepting groups (5 mol% TRZ-­H and 0.4 mol% TRZ-­CN) on polystyrene backbone. It was
remarkable that this white polymer exhibited excellent EL performance with a maximum ηext of
14.1%, providing the most efficient OLEDs based on white emission polymers ever reported. It
strongly demonstrated the feasibility of applying TSCT property to design outstanding AIDF lumi-
nogens and provided a strategy to develop dual-­emission materials for OLEDs.

8.3  ­Conclusion and Outlook

Recently, as one of the most favorable alternatives of display techniques, OLEDs have developed
with an explosive growth. An increasing number of researchers devote their effort to explore novel
materials systems, and improve the performance of OLEDs to achieve commercial production and
large-­scale application. Nonetheless, the devices based on TADF emitters, for which the develop-
ment momentum is more prosperous in recent year, need to be doped in host materials with pre-
cise concentration control, and the efficiency roll-­off is still severe at high voltages. Thanks to the
appearance of AIDF luminogens, not only can they realize almost 100% IQE, but also restrain
concentration quenching and exciton annihilation. According to the reported results, the devices
using AIDF materials as light-­emitting layers exhibited excellent EL performance with very small
efficiency roll-­off at high voltages, which are superior to those of TADF emitters to a large degree.
AIDF materials have achieved the coverage of full-­color region, and white emission devices are
also realized based on the AIDF materials. However, there is still much room for the improvement
in terms of efficiency and stability for non-­doped devices, and the deeper mechanism deciphering
of AIDF phenomenon is also needed to put effort into. In addition, the emission colors of currently
reported AIDF materials are mainly in the green and yellow regions, and high-­performance blue
and red AIDF materials are insufficient. The multiple resonance effect of the boron and nitrogen
atoms and the through-­space charge transfer effect can be adopted to improve the performance of
blue AIDF materials. Enhancing the electron-­withdrawing ability and increasing the twisted intra-
molecular charge transfer effect by designing new electron acceptors are helpful to acquire effi-
cient red AIDF materials. For increasing the stability of the devices based on AIDF materials, there
may be two feasible methods. One is increasing the distance between the excitons, which is condu-
cive to weakening the bimolecular-­quenching processes. Another includes grafting groups with
good exciton transfer ability when designing molecules, which can be used to improve the perfor-
mance of doped systems. Finally, through this chapter, we hope more research groups will pay
attention to the field of AIDF luminogens and bring about further exciting achievement for OLEDs.

­References

1 Tang, C. W., Vanslyke, S. A. (1987). Organic electroluminescent diodes. Applied Physics Letters
51(12): 913–915.
2 Pope, M., Kallmann, H. P., Magnante, P. (1963). Electroluminescence in organic crystals. Journal of
Chemical Physics 38(8): 2042–2043.
248 8  Aggregation-­induced Delayed Fluorescence

3 D’andrade, B. W., Forrest, S. R. (2004). White organic light-­emitting devices for solid-­state lighting.
Advanced Materials 16(18): 1585–1595.
4 Grimsdale, A. C., Chan, K. L., Martin, R. E., et al. (2009). Synthesis of light-­emitting conjugated
polymers for applications in electroluminescent devices. Chemical Reviews 109(3): 897–1091.
5 Mei, J., Hong, Y., Lam, J. W., et al. (2014). Aggregation-­induced emission: the whole is more
brilliant than the parts. Advanced Materials 26(31): 5429–5479.
6 Kwok, R. T., Leung, C. W., Lam, J. W., et al. (2015). Biosensing by luminogens with aggregation-­
induced emission characteristics. Chemical Society Reviews 44(13): 4228–4238.
7 Scott, J. C., Bozano, L. D. (2007). Nonvolatile memory elements based on organic materials.
Advanced Materials 19(11): 1452–1463.
8 McQuade, D. T., Pullen, L. D., Swager., T. M. (2000). Conjugated polymer-­based chemical sensors.
Chemical Reviews 100(7): 2537–2574.
9 Tao, Y., Yuan, K., Chen, T., et al. (2014). Thermally activated delayed fluorescence materials
towards the breakthrough of organoelectronics. Advanced Materials 26(47): 7931–7958.
10 Dos Santos, P. L., Ward, J. S., Bryce, M. R., et al. (2016). Using guest-­host interactions to optimize
the efficiency of TADF OLEDs. The Journal of Physical Chemistry Letters 7(17): 3341–3346.
11 Zhang, D., Wei, P., Zhang, D., et al. (2017). Sterically shielded electron transporting material with
nearly 100% internal quantum efficiency and long lifetime for thermally activated delayed
fluorescent and phosphorescent OLEDs. ACS Applied Materials & Interfaces 9(22): 19040–19047.
12 Zhelev, Z., Ohba, H., Bakalova, R. (2006). Single quantum dot-­micelles coated with silica shell as
potentially non-­cytotoxic fluorescent cell tracers. Journal of the American Chemical Society 128(19):
6324–6325.
13 Jenekhe, S. A., Osaheni, J. A. (1994). Excimers and exciplexes of conjugated polymers. Science
265(5173): 765–768.
14 Bradley, D. D. C., Friend, R. H., Gymer, R. W., et al. (1999). Electroluminescence in conjugated
polymers. Nature 397(6715): 121–128.
15 Luo, J., Xie, Z., Lam, J. W. Y., et al. (2001). Aggregation-­induced emission of 1-­methyl-­1,2,3,4,5-­
pentaphenylsilole. Chemical Communications (18): 1740–1741.
16 Tang, B. Z., Zhan, X., Yu, G., et al. (2001). Efficient blue emission from siloles. Journal of Materials
Chemistry 11(12): 2974–2978.
17 Qin, A., Lam, J. W. Y., Mahtab, F., et al. (2009). Pyrazine luminogens with “free” and “locked”
phenyl rings: Understanding of restriction of intramolecular rotation as a cause for aggregation-­
induced emission. Applied Physics Letters 94(25): 253308.
18 Yu, G., Yin, S., Liu, Y., et al. (2005). Structures, electronic states, photoluminescence, and carrier
transport properties of 1,1-­disubstituted 2,3,4,5-­tetraphenylsiloles. Journal of the American
Chemical Society 127(17): 6335–6346.
19 Chen, J., Law, C. C. W., Lam, J. W. Y., et al. (2003). Synthesis, light emission, nanoaggregation, and
restricted intramolecular rotation of 1,1-­substituted 2,3,4,5-­tetraphenylsiloles. Chemistry of
Materials 15(7): 1535–1546.
20 Godumala, M., Choi, S., Cho, M. J., et al. (2019). Recent breakthroughs in thermally activated
delayed fluorescence organic light emitting diodes containing non-­doped emitting layers. Journal
of Materials Chemistry C 7(8): 2172–2198.
21 Huang, J., Nie, H., Zeng, J., et al. (2017). Highly efficient nondoped OLEDs with negligible
efficiency roll-­off fabricated from aggregation-­induced delayed fluorescence luminogens.
Angewandte Chemie International Edition 56(42): 12971–12976.
22 Kondakov, D. Y. (2009). Role of triplet-­triplet annihilation in highly efficient fluorescent devices.
Journal of the Society for Information Display 17(2): 137–144.
 ­Reference 249

23 Guo, J., Zhao, Z., Tang, B. Z. (2018). Purely organic materials with aggregation-­induced delayed
fluorescence for efficient nondoped OLEDs. Advanced Optical Materials 6(15): 1800264.
24 Furue, R., Nishimoto, T., Park, I. S., et al. (2016). Aggregation-­induced delayed fluorescence based
on donor/acceptor-­Tethered Janus carborane triads: Unique photophysical properties of nondoped
OLEDs. Angewandte Chemie International Edition 55(25): 7171–7175.
25 Guo, J., Li, X.-­L., Nie, H., et al. (2017). Achieving high-­performance nondoped OLEDs with
extremely small efficiency roll-­off by combining aggregation-­induced emission and thermally
activated delayed fluorescence. Advanced Functional Materials 27(13): 1606458.
26 Guo, J., Li, X.-­L., Nie, H., et al. (2017). Robust luminescent materials with prominent aggregation-­
induced emission and thermally activated delayed fluorescence for high-­performance organic
light-­emitting diodes. Chemistry of Materials 29(8): 3623–3631.
27 Guo, J., Fan, J., Lin, L., et al. (2019). Mechanical insights into aggregation-­induced delayed
fluorescence materials with anti-­kasha behavior. Advanced Science 6(3): 1801629.
28 Zeng, J., Guo, J., Liu, H., et al. (2019). Aggregation-­induced delayed fluorescence luminogens for
efficient organic light-­emitting diodes. Chemical-­An Asian Journal 14(6): 828–835.
29 Lee, J., Aizawa, N., Numata, M., et al. (2017). Versatile molecular functionalization for inhibiting
concentration quenching of thermally activated delayed fluorescence. Advanced Materials 29(4):
1604856.
30 Liu, H., Zeng, J., Guo, J., et al. (2018). High-­performance non-­doped OLEDs with nearly 100 %
exciton use and negligible efficiency roll-­off. Angewandte Chemie International Edition 57(30):
9290–9294.
31 Wu, H. B., Ying, L., Yang, W., et al. (2009). Progress and perspective of polymer white light-­
emitting devices and materials. Chemical Society Reviews 38(12): 3391–3400.
32 Huang, J., Xu, Z., Cai, Z., et al. (2019). Robust luminescent small molecules with aggregation-­
induced delayed fluorescence for efficient solution-­processed OLEDs. Journal of Materials
Chemistry C 7(2): 330–339.
33 Chen, Y. H., Wang, S. M., Wu, X. F., et al. (2018). Triazatruxene-­based small molecules with
thermally activated delayed fluorescence, aggregation-­induced emission and mechanochromic
luminescence properties for solution-­processable nondoped OLEDs. Journal of Materials
Chemistry C 6(46): 12503–12508.
34 Xie, Z., Su, T., Ubba, E., et al. (2019). Achieving tunable dual-­emissive and high-­contrast
mechanochromic materials by manipulating steric hindrance effects. Journal of Materials
Chemistry C 7(11): 3300–3305.
35 Gan, S., Luo, W., He, B., et al. (2016). Integration of aggregation-­induced emission and delayed
fluorescence into electronic donor-­acceptor conjugates. Journal of Materials Chemistry C 4(17):
3705–3708.
36 Zhao, J., Chen, X. J., Yang, Z., et al. (2019). Highly-­efficient doped and nondoped organic light-­
emitting diodes with external quantum efficiencies over 20% from a multifunctional green
thermally activated delayed fluorescence emitter. Journal of Physical Chemistry C 123(2):
1015–1020.
37 Lee, I. H., Song, W., Lee, J. Y. (2016). Aggregation-­induced emission type thermally activated
delayed fluorescent materials for high efficiency in non-­doped organic light-­emitting diodes.
Organic Electronics 29: 22–26.
38 Xiang, S. P., Huang, Z., Sun, S. Q., et al. (2018). Highly efficient non-­doped OLEDs using
aggregation-­induced delayed fluorescence materials based on 10-­phenyl-­10H-­phenothiazine
5,5-­dioxide derivatives. Journal of Materials Chemistry C 6(42): 11436–11443.
250 8  Aggregation-­induced Delayed Fluorescence

39 Zhang, Q., Sun, S., Liu, W., et al. (2019). Integrating TADF luminogens with AIE characteristics
using a novel acridine–carbazole hybrid as donor for high-­performance and low efficiency roll-­off
OLEDs. Journal of Materials Chemistry C 7(31): 9487–9495.
40 Zhou, X., Yang, H., Chen, Z., et al. (2019). Naphthyridine-­based emitters simultaneously exhibiting
thermally activated delayed fluorescence and aggregation-­induced emission for highly efficient
non-­doped fluorescent OLEDs. Journal of Materials Chemistry C 7(22): 6607–6615.
41 Wu, Y., Chen, X., Mu, Y., et al. (2019). Two thermally stable and AIE active 1,8-­naphthalimide
derivatives with red efficient thermally activated delayed fluorescence. Dyes and Pigments
169: 81–88.
42 Huang, B., Ji, Y. G., Li, Z. J., et al. (2017). Simple aggregation-­induced delayed fluorescence
materials based on anthraquinone derivatives for highly efficient solution-­processed red OLEDs.
Journal of Luminescence 187: 414–420.
43 Grybauskaite-­Kaminskiene, G., Volyniuk, D., Mimaite, V., et al. (2018). Aggregation-­enhanced
emission and thermally activated delayed fluorescence of derivatives of 9-­phenyl-­9h-­carbazole:
effects of methoxy and tert-­butyl substituents. Chemistry-­A European Journal 24(38): 9581–9591.
44 Zhang, P., Zeng, J., Guo, J., et al. (2019). New aggregation-­induced delayed fluorescence
luminogens with through-­space charge transfer for efficient non-­doped OLEDs. Frontiers in
Chemistry 7: 199.
45 Wang, J., Liu, C., Jiang, C., et al. (2019). Solution-­processed aggregation-­induced delayed
fluorescence (AIDF) emitters based on strong π-­accepting triazine cores for highly efficient
nondoped OLEDs with low efficiency roll-­off. Organic Electronics 65: 170–178.
46 Wang, X. D., Wang, S. M., Lv, J. H., et al. (2019). Through-­space charge transfer hexaarylbenzene
dendrimers with thermally activated delayed fluorescence and aggregation-­induced emission for
efficient solution-­processed OLEDs. Chemical Science 10(10): 2915–2923.
47 Hu, J., Li, Q., Wang, X., et al. (2019). Developing through-­space charge transfer polymers as a
general approach to realize full-­color and white emission with thermally activated delayed
fluorescence. Angewandte Chemie International Edition 58(25): 8405–8409.
251

Homogeneous Systems to Induce Emission of AIEgens


Kenta Kokado1,2,3 and Kazuki Sada4,5
1
 Research Institute for Electronic Science, Hokkaido University, Sapporo, Hokkaido, Japan
2
 Graduate School of Environmental Science, Hokkaido University, Sapporo, Hokkaido, Japan
3
 PRESTO, JST, Chiyodaku, Tokyo, Japan
4
 Faculty of Science, Hokkaido University, Sapporo, Hokkaido, Japan
5
 Graduate School of Chemical Sciences and Engineering, Hokkaido University, Sapporo, Hokkaido, Japan

9.1 ­Introduction

In 2001, Tang et al. reported the aggregation-­induced emission (AIE) of 1-­methyl-­1,2,3,4,5-­pentap


henylsilole (1), which was nonemissive in solution but strongly emissive in the aggregated state [1].
Since this demonstration, various luminogens showing AIE (AIEgens) have been reported  [2]
including tetraphenylethylene (TPE, 2)  [3], 1-­cyano-­1,2-­bis(4′-­methylbiphenyl)ethylene (CN-­
MBE, 3) [4], hexaphenylbenzene (4) [5], and aryl-­o-­carborane (5) (Figure 9.1) [6]. The accumula-
tion of aryl substituents on the π plane or π bond appeared to be the requirement for AIE. In 2003,
Tang et al. proposed the restriction of intramolecular rotation (RIR) as the working mechanism of
the AIE of siloles [7]; this idea was derived from the relationship between the peak broadening of
the NMR spectrum and the emission enhancement in fluorescence spectroscopy at low tempera-
ture. It appears unreasonable to connect the results of nuclear magnetic resonance (NMR) spec-
troscopy and fluorescence spectroscopy, as fluorescence is a result of a photophysical process in the
excited state, whereas NMR spectra illustrate the condition in the ground state. Nevertheless, RIR
has been applied for other AIEgens such as TPE to explain the origin of AIE, and this explanation
is widely accepted in the newly formed AIE community, leading to confusion and disagreement
over the interpretation of the AIE mechanism among traditional photochemists [8, 9]. For photoi-
somerizable AIEgens such as TPE, there is a wealth of photochemical information collected
through other techniques, such as ultrafast spectroscopy  [10–14], optical calorimetry  [15], and
time-­resolved microwave conductometry analyses [16], indicating that π twist in the excited state
(including photoisomerization), rather than RIR, is the more probable AIE mechanism. (After
accumulation of researches of AIEgens, Tang’s group extended the concept of RIR to restriction of
intramolecular motion [17], including restriction of intramolecular vibration [18].)
Recent progress in theoretical computation has provided information on the molecular structure
in the excited state, disclosing the relaxation pathway including the minimum energy conical inter-
section (MECI). We elucidated the quenching mechanism of TPE in the solution state using pho-
toirradiation and theoretical computation, in which π twist of the central C═C bond in the excited

Handbook of Aggregation-Induced Emission: Volume 1 Tutorial Lectures and Mechanism Studies, First Edition.
Edited by Youhong Tang and Ben Zhong Tang.
© 2022 John Wiley & Sons Ltd. Published 2022 by John Wiley & Sons Ltd.
252 9  Homogeneous Systems to Induce Emission of AIEgens

Me
Si CN
Me
Me

Pentaphenylsilole (1) CN-MBE (3)

C C

Tetraphenylethylene Hexaphenylbenzene (4) Aryl-o-carborane (5)


(TPE, 2)

Figure 9.1  Representative AIEgens, including pentaphenylsilole [1], tetraphenylethylene (TPE) [3],


1-­cyano-­1,2-­bis(4′-­methylbiphenyl)ethylene (CN-­MBE) [4], hexaphenylbenzene [5], and aryl-­o-­carborane [6].

state was shown to be responsible for the quenching [19]. The established theoretical ­computations
also revealed the quenching mechanism of AIEgens without photoisomerizable groups, in which π
twist and distortion in the excited state is responsible for the quenching  [20]. Therefore, a large
structural change, such as π twist in the excited state, is essential for quenching of AIEgens in the
solution state, and prohibiting this large structural change, such as through aggregation, leads to the
strong emission in AIEgens [21]. From this principle, it is apparent that aggregation is not a require-
ment for the strong emission of AIEgens but only a sufficient condition. Thus, even in a homogene-
ous condition such as a solution, the strong emission of AIEgens can be observed by applying
appropriate molecular design to prohibit the large structural change in the excited state.
In this chapter, we introduce examples of AIEgens that show strong emission in homogeneous
systems such as in solutions, liquids, gels, or crystals (Figure 9.2). To focus on homogeneous sys-
tems inducing the emission of AIEgens, examples in which the emission is started after the forma-
tion of aggregates or fibers are mostly omitted in this chapter.

9.2 ­Homogeneous Solution

Although most research on AIEgens adopt their aggregates to induce emission, there is a moderate
amount of research on inducing the emission of AIEgens in homogeneous solutions. In the work-
ing mechanism of AIE, the fixation of the molecular structure of AIEgens in the excited state is
essential to induce the emission. From this viewpoint, one can easily imagine strategies to fix the
molecular structure of AIEgens in homogeneous solution using the following three approaches:
(i) supramolecular interaction, (ii) bulky steric hindrance, and (iii) covalent linkage. For (i) supra-
molecular interaction, ionic interactions, inclusion complexes, and adhesion on macromolecules
have been applied for the fixation. In the following sections, we introduce these ­examples in order.
9.2 ­Homogeneous Solutio 253

(a) Anions (b)


Cations

Homogeneous solution
Complexation Complexation with anions
Complexation with cations
Strong emission Inclusion complexes
Hosts
Adhesion on macromolecules
Steric hindrance
Covalent linkage

Inclusion
Liquid
Strong emission
AIEgen
Gels and network polymers
Adhesion
Chemically crosslinked gels
Physically crosslinked gels
Strong emission

Crystalline materials

Gelation
Strong emission

Figure 9.2  Homogeneous systems reviewed in this chapter. (a) Schematic illustration of selected systems,
and (b) summary of the systems.

9.2.1  Complexation with Anions


Anions have a great effect on both life health and environmental protection; thus, many efforts
have been made to develop consistent methods to monitor anions, flourishing the chemistry of
many anion receptors [22]. This concept is practically applicable to induce the emission of AIEgens
in homogeneous solution (Figure 9.3). Shinkai et al. first reported TPE with four zinc-­dipicolylamine
(6) exhibiting cyclization-­induced emission (CIE) when complexed with a dicarboxylic acid such
as tartaric acid in HEPES-­buffered H2O [23]. The emission intensity was enhanced by 170 times
compared with that in the absence of tartarate, and monocarboxylic acid such as acetate could not
induce the emission. The first association constant (KA1) for tartarate was smaller (819 M−1) than
second association constant (KA2, 3935 M−1), indicating a “positive allosteric effect.” Miljanić et al.
also reported a trigonal fluorinated trispyrazole specifically detecting phthalic acid from the iso-
mers and other dicarboxylic or monocarboxylic acids in DMF[24]. The association of two mole-
cules of 7 and phthalic acid resulted in the strong emission in homogeneous solution. Maleic acid
could also induce the strong emission; however, the trans isomer, fumaric acid could not because
of the lack of proximity of the two molecules of 7.
In addition to carboxylates, oxyanions are known to be bound by various anion receptors with
multiple hydrogen bonding sites, and this feature has also been applied to induce the emission of
AIEgens in homogeneous solution. For example, Wu et  al. synthesized bisurea-­type TPE (8),
which specifically binds phosphate (PO43−), hydrogen phosphate (HPO42−), and sulfate (SO42−) in
DMSO, resulting in strong emission[25]. Notably, dihydrogenphosphate (H2PO4−), carboxylate
(CO32−), or other halogenates cannot induce the strong emission. In the crystal structure of
254 9  Homogeneous Systems to Induce Emission of AIEgens

N N
N NH
Zn2+ Zn2+
N N N N

F F

O O
F F

F N N F
8NO3–
F F
N
N
O O F F
HN
N F F NH

N N N N
Zn2+ Zn2+
N N

6 7

H3C CH3 H9C4 C4H9


O O NH HN

N N N N O NH HN O
H H H H
HN O O NH

HN NH O O

HN NH O O

HN O O NH
H H H H
N N N N O NH HN O

O O NH HN
H3C CH3 H9C4 C4H9
8 9

Figure 9.3  AIEgens showing strong emission in the presence of anions in homogeneous solution.

[K([18] crown-­6)]8[(HPO4)4(8)2], molecule 8 has an anion/ligand ratio of 4/2; thus, two molecules
of 3 bind the anion HPO42− via two bisurea arms derived from each ligand. Pigge et al. also syn-
thesized TPE with urea groups (9), which exhibited emission enhancement with the addition of
various anions in THF[26]. The specificity was moderate; however, fluoride exhibited the greatest
emission owing to the high basicity or hydrogen bonding ability of fluoride.
Although it has not been reported, the application of this phenomenon to various other anions
such as metal oxyanions will expand the usability of emission from AIEgens. The chemistry of
anion receptors should be further investigated to develop the anion-­responsive emission of
AIEgens in homogeneous solution.

9.2.2  Complexation with Cations


Thanks to the development of coordination chemistry, various organic ligands that can interact
with metal cations dependent on their valence and orbital shape have been explored. The metal
coordination bond can also be utilized to restrict the twist and distortion of AIEgens in the excited
state (Figure 9.4).
9.2 ­Homogeneous Solutio 255

N N
N N
MeO N N OMe

N N MeO N N OMe
N N

10 11

N N
N N

OH
N
HO
N N OH
N N 14
12 13

Figure 9.4  AIEgens showing strong emission in the presence of cations in homogeneous solution.

Hahn et al. reported the synthesis of TPE with N-­heterocyclic carbene (NHC) groups (10), which
can bind precious metal cations such as Ag+ and Au+ [27]. After complexation of 10 with these
metal cations, they started to enhance the emission at ~500 nm in acetonitrile. The crystal struc-
ture of [Ag2·10]·(PF6)2 shows that the cation is chelated by two NHC groups derived from one
molecule of 10. Xu et  al. synthesized TPE with 2-­pyridyl groups (11, which showed emission
enhancement by complexation with Ag+ in THF). Compound 11 specifically bound Ag+ via two
vicinal 2-­pyridyl groups, as revealed by its crystal structure [28]. Other metal ions such as Al3+,
Ba2+, Ca2+, Cu2+, Fe2+, K+, Li+, Mg2+, Na+, and Ni+ did not result in enhancement. From Hill’s
equation, the Hill coefficient was determined to be 2.47, indicating strong cooperativity on Ag+
binding. Sun et  al. reported the formation of a metal complex of 12 and Pd2+, which exhibited
enhanced emission after complexation in acetone/H2O (1/1, v/v) [29]. Notably, the ligand with two
3-­pyridyl groups on the geminal position resulted in no emission enhancement. Although the
authors depicted complexation with Pd2+ via the geminal two 3-­pyridyl groups, the complexation
via the vicinal two 3-­pyridyl groups appears more plausible from the quenching mechanism of
TPE in solution. Liu et al. synthesized a metal complex of 13 and Pt2+, showing yellow emission in
methanol or acetonitrile) [30]. The aggregation drastically changed the emission color to blue. In
other words, the formation of the supramolecular assembly is sufficient to induce emission; how-
ever, the latitude of twist or distortion in the excited state remains large, resulting in the emission
at relatively longer wavelength. Chattopadhyay et al. reported the synthesis of an iminomethylben-
zene derivative (14) exhibiting emission enhancement through complexation with Al3+ in DMF/
H2O (9/1) [31]. It showed specific binding of Al3+, and no enhancement was observed for Na+, K+,
Ca2+, Mg2+, Mn2+, Fe2+, Co2+, Cu2+, Zn2+, Hg2+, Sr2+, Ni2+, Ba2+, Pb2+, Al3+, Cr3+, and Fe3+.
Compared with an anion, the molecular structures of ligands for metal cations are simple and
small because of the small size of target metal cations and the omnidirectional orientation of their
orbitals. Because of the variety of metal cations, a variety of responses has been reported; however,
there are no examples for cations other than metal cations, such as organic cations. In addition, the
turn-­on emission response to early transition metal cations would be desirable using AIEgens,
256 9  Homogeneous Systems to Induce Emission of AIEgens

which is difficult to achieve using conventional organic luminogens whose frontier orbitals change
their energy after complexation with early-­transition metal cations in principle, resulting in
quenching.

9.2.3  Inclusion Complexes


In addition to anion receptors and metal-­complex cages, many inclusion complexes have been
explored because of the growing supramolecular chemistry. Using such an inclusion complex,
there have been several reports of the twist or distortion of AIEgens being restricted in the excited
state (Figure 9.5).
Zheng et al. reported that TPE having oligo(ethylene glycol) chains (15) forms an inclusion com-
plex with γ-­cyclodextrin (CD), resulting in the induction of monomer emission at 388 and 405 nm
in H2O [32]. Compound 15 is soluble in H2O; however, the emission was first broadened, presum-
ably owing to the formation of a small aggregate by the hydrophobic TPE core. The addition of γ-­
CD caused the hypsochromic shift of the emission, whereas the smaller CDs, α-­CD, and β-­CD,
caused no shift. The interaction between compound 15 and γ-­CD was confirmed by nuclear
Overhauser effect spectroscopy (NOESY) spectrum between the aromatic protons of 15 and the
methine protons of γ-­CD. The monomer emission was readily returned to aggregate emission by
the addition of a competitive guest, potassium ursodeoxycholate, indicating that the inclusion of
15 by γ-­CD is essential to induce the monomer emission. Li et al. synthesized TPE with trimethyl-
ammonium groups (16), showing inclusion in cucurbit[7] uril (CB7) and subsequent emission
enhancement in H2O [33]. The addition of a competitive guest, 1-­adamantanamine hydrochloride,
resulted in the hypsochromic shift of the emission maxima from 464 to 396 nm, indicating the dis-
sociation of the complex of 16 and CB7. A similar molecule, TPE with trimethylammonium groups
(17), was synthesized by Guo et  al., which showed emission enhancement in the presence of
p-­sulfonatocalix[4] arene in H2O [34]. The inclusion of other molecules, such as α-­CD, β-­CD, CB7,
and CB8, could not enhance the emission. A similar phenomenon was observed with the addition

O O + O O +
O O N N
4 4
4Br–

+ +
O O N N
O O O O
4 4
15 16

+ +
N O O N

4Br–

+ +
N O O N
17

Figure 9.5  AIEgens showing strong emission by the formation of inclusion complex with host compound
in homogeneous solution.
9.2 ­Homogeneous Solutio 257

of heparin, which is a highly sulfonated polysaccharide. Lei et al. reported the in situ imaging of a
membrane-­associated glycoprotein mucin 1 (MUC1) using an AIE-­active Cu nanocluster (CuNC)
covered with mono-­(6-­mercapto-­6-­deoxy)-­β-­CD (SH-­βCD)  [35]. After treatment of HepG2 cells
with a MUC1-­binding aptamer containing an adamantanyl group (Ad-­Apt) and the addition of a
SH-­βCD-­covered CuNC and an adamantane dimer, di(1-­adamantan-­yl)phosphine (Ad–Ad),
MUC1 was selectively imaged without any nonspecific adsorption of CuNC on the cell membrane,
which is beneficial for long-­term image-­guided diagnosis and therapy.

9.2.4  Adhesion on Macromolecules


Häner et al. reported a TPE derivative incorporated in DNA [36]. The TPE derivative initially had
a CEP (2-­cyanoethyl-­N,N-­diisopropyl-­phosphoramidite)-­tethering 4-­oxybutynyl chain (18,
Figure  9.6), and then, it was incorporated in DNA using the phosphoramidite method. After
hybridization with the complementary chain, the emission was enhanced 3–9 times compared
with that of the single strand in THF/H2O (1/1,v/v). The melting temperature (Tm) of the duplex
was improved by +2–5 °C, most likely due to the favorable stacking of the TPE moieties. Lei et al.
prepared an oligonucleotide with a TPE moiety on the 3′-­terminus (19), and emission enhance-
ment was observed when 19 was hybridized with a tetrapod DNA quadraplex consisting of a

O O
NC P
N

N
P CN
O O N3
18

N
O– O N
5′ AGTTGGAGACGTAAG O P O N N
N
O H
O
19
OH OH
N N
COOH
HO HO

20

+
N N N
N O O

BF4–

21 22

Figure 9.6  AIEgens showing strong emission after adhesion to macromolecules in homogeneous solution.
258 9  Homogeneous Systems to Induce Emission of AIEgens

complementary chain of 19 and six G bases in TE reaction buffer [37]. The emission intensity was
finely tuned by changing the spacer length between the TPE moiety and the hybridizing oligonu-
cleotide chain as well as the number of G-­quartets (up to 6). The presence of G-­quartets was essen-
tial for the emission enhancement, and other bases (A, C, T) could not enhance the emission. This
method is promising for signal transaction of biosensors. Hong et al. reported emission enhance-
ment of 1,2-­bis(2,4-­dihydroxybenzylidene)hydrazine (20) when it adsorbed on a polyrotaxane con-
sisting of Jeffamine and β-­CD in DMSO  [38]. Compound 20 adsorbed on the polyrotaxane via
multiple hydrogen bonding, and the use of a larger amount of polyrotaxane significantly increased
the emission from compound 20. Wang et al. synthesized TPE with a triazolium moiety (21), which
exhibited emission enhancement when mixed with heparin in acetonitrile/Tris-­HCl buffer (5/95,
v/v) [39]. The emission was also enhanced in the presence of other sulfonic polysaccharides such
as chondroitin (Chs) and sulfonate cyclodextrin. The addition of a positively charged peptide, pro-
tamine, effectively quenched the emission owing to the dissociation of 21 and heparin. Liu et al.
reported the emission enhancement of 6-­(1,3-­dioxo-­4,5,6,7-­tetraphenylisoindolin-­2-­yl) hexanoic
acid (22) when mixed with protamine in Tris-­HCl buffer  [40]. The addition of heparin readily
quenched the emission. The detection limit (4.78 ng/ml) and linear range (0–6 μg/ml) for prota-
mine were excellent compared with those of other reported systems.

9.2.5  Steric Hindrance


The introduction of steric hindrance is an effective way to restrict the twist and distortion of a
molecular structure in the excited state (Figure 9.7). As the earliest example, Tang et al. reported
tetra(aryl)ethylenes having several naphthalene rings as the aryl groups (23) [41]. As the number

HO

O
O O OH
OH OH OH
HO O
OH
OH
O O
OH
OH O
O
O HO
HO OH
O O
O OH OH
23 O HO
HO O
OH O
O OH
HO
24

25 26

Figure 9.7  AIEgens showing strong emission due to steric hindrance restricting π twist in the excited
state in homogeneous solution.
9.2 ­Homogeneous Solutio 259

of naphthyl groups increased, the fluorescence quantum yield (ΦF) increased from 0.24 to
1.66% in THF.
The direct connection of TPE and CD was also employed to introduce steric hindrance to
TPE [42]. Although mixing of TPE with CDs resulted in no emission enhancement, their cova-
lently connected systems resulted in emission enhancement for α-­CD and β-­CD, not γ-­CD, in
DMSO. The addition of adamantane to TPE and β-­CD adduct (24) resulted in further emission
enhancement, indicating the steric hindrance function of the CD moiety, not only including
the host.
The substitution of the ortho position of phenyl rings on TPE is another effective way to intro-
duce steric hindrance to TPE [43]. Zhao et al. reported stereoselective synthesis of o-­aryl-­substituted
TPE (25) by McMurry coupling of 2-­arylbenzophenone. The reaction specifically produced (Z)-­
isomer, most likely due to the repulsion of the unsubstituted phenyl rings on TPE and π–π stacking
of the substituent phenyl rings. The resulting (Z)-­isomer exhibited strong emission in THF; how-
ever, it also exhibited emission enhancement upon aggregation. These features, absorption max-
ima and AIEE, were also well predicted and described by theoretical computations by the same
research group [44], although the calculation is mainly related to the ground state.
Zhu et  al. reported the synthesis of 1,1-­bis(2,6-­dimethylphenyl)-­2,2-­diphenyl-­ethylene (26),
which exhibited emission enhancement in THF. Interestingly, no emission was observed for the
1,1-­bis(2-­methylphenyl) derivative [45]. The decay pathway from the excited state was proposed by
Thiel et al. [46]. As a result, the barrier along the pathway from the Franck–Condon point to the
conical intersection, where cyclization or pyramidal inversion occurs, was relatively high for com-
pound 26 (6.2 and 8.4 kcal/mol, respectively), whereas those for typical AIEgens were lower (0 and
1.8 kcal/mol, respectively).

9.2.6  Covalent Linkage


The covalent linkage of the twistable moiety in AIEgen is the third and intuitive way to achieve
emission enhancement in homogeneous solution (Figure 9.8).
Zheng et  al. reported the synthesis of 9-­diphenyldibenzofulvene and its dimerization via two
imidazolium to form a macrocycle compound (27) [47]. Compound 26 exhibited monomer emis-
sion in DMSO at 450 nm. Diphenyldibenzofulvene itself is AIE-­active [48]; thus, the covalent link-
age would restrict twist or distortion of the molecular structure in the excited state, although
geminal linkage appeared less effective to restrict them. Recently, Zhen et al. reported the impor-
tance of vicinal linkage of TPE derivatives to enhance the emission [49]. In this case, the vicinal
phenyl rings were linked by resorcinol (28), xylene, or benzene, resulting in emission enhance-
ment in THF. Geminal linkage resulted in no emission enhancement. Femtosecond transient

OMe MeO
N N MeO OMe O
+
O O
2Cl–
O O
+
N N MeO OMe O
OMe MeO
27 28 29

Figure 9.8  AIEgens showing strong emission by covalent linkage restricting π twist in the excited state in
homogeneous solution.
260 9  Homogeneous Systems to Induce Emission of AIEgens

absorption spectroscopy revealed the initial twist of the central C═C bond in the excited state, and
larger twist for the compound with geminal linkage was proposed, indicating the importance of
vicinal linkage in restricting the twist or distortion of the molecular structure in the excited state.
In earlier reports, the importance of geminal linkage for the RIR process has been suggested. For
example, compound 29 exhibited strong emission in THF; thus, the authors reported that RIR was
achieved even in the solution state [50]. However, the O-­linkage considerably affected the molecu-
lar structure in the ground state, and the central C═C bond was already highly twisted because of
the rigid and wide π plane [51]. As a result, the molecular structure in the ground and excited state
were similar, leading to strong emission in homogeneous solution despite the central C═C bond.

9.3 ­Liquid

Bulk materials realize the fixation of a molecular structure compared to that in homogeneous solu-
tion. A pure liquid is one of the softest and fluent materials; thus, one can imagine that it is too soft
to restrict the twist of the molecular structure in the excited state. However, the high viscosity of a
liquid such as honey is enough to restrict the twist of the molecular structure in the excited state
(Figure 9.9).
In 2017, we first reported an AIE liquid comprising a TPE moiety and long and branched alkyl
chains [52]. The linkage of the alkyl chain and TPE moiety is critical for the resulting physical
properties. Ether linkage (30) resulted in a stable liquid even at low temperature; however, ester
linkage resulted in a supercooled liquid, which was liquid in the as-­prepared state or after heating
but became a crystalline solid after cooling. The subtle hydrogen bonding derived from the ester
group is responsible for this difference, as confirmed by the crystal structures of model compounds.
Moreover, the AIE liquid derived from ether linkage exhibited quite high ΦF, whereas that derived
from ester linkage produced lower ΦF. This difference in ΦF resulted from the aforementioned
intermolecular interaction; thus, the relatively stronger hydrogen bonding of the ester linkage
resulted in face-­to-­face packing of the TPE moiety. In other words, concentration quenching effect
is inevitable even for AIEgens, unless appropriate molecular design is applied.
Takeda et al. reported the thermoresponsive emission behavior of TPE with tetra(ethylene gly-
col) (TEG) chains (31)  [53]. By heating from liquid N2 temperature, the emission maxima
bathochromically shifted by 60  nm from 466 to 517  nm. The glass-­transition temperature (Tg,
−60 °C) was significantly lowered by tethering the TEG chains compared with that for straight
alkyl chains (amide linkage: 252 °C, ester linkage: 62 °C).
Examples of liquid crystals are omitted here because in a liquid crystal, an AIEgen would behave
as a hard segment, meaning the formation of aggregates of the AIEgen. This situation is not far
from that of typical aggregates of AIEgens in a poor solvent, as can be conventionally observed.

O O
O O O O
O O
4 4

O O
O O O O
4 O O 4
30 31

Figure 9.9  Liquid AIEgens.


9.4  ­Gels and Network Polymer 261

9.4 ­Gels and Network Polymers

Another important approach for the fixation of the molecular structure is the use of macromolecu-
lar systems. Although macromolecular or supramolecular networks exist in gels, solvents, or small
molecules, other chemical species such as ions can enter and leave a gel, analogously to solution.
For AIEgens, which are conventionally incorporated in a swollen gel network, emission is usually
difficult unless the twist or distortion in the excited state is restricted. This difficulty arises because
the environment surrounding the AIEgen is similar to a solution state. In other words, a change
of the swelling of an AIEgen-­incorporated gel often involves a change of emission from the gel.
From the viewpoint of the cross-­linking type, gels can be classified into two categories, chemically
crosslinked and physically crosslinked gels, and are thus characterized by covalent linkage or
supramolecular interaction, respectively.

9.4.1  Chemically Crosslinked Gels


In 2010, we presented the first chemically crosslinked gel (32, Figure 9.10) containing an AIEgen
as the crosslinker [54]. The AIEgen was derived from o-­carborane, an icosahedral cluster with the
chemical formula C2B10H12 [6], in which the cluster is highly distorted in the excited state in solu-
tion. The gel mainly consisted of poly(γ-­glutamic acid) (PGA, MW = 200 000–500 000) crosslinked
by diglycidyl compounds. An increase of the crosslinker resulted in emission enhancement in the
swollen state, owing to the restriction of distortion of the AIEgen. The gel with less crosslinker was
basically not emissive in the swollen state in H2O; however, it started to emit after being dried.
Additionally, increasing the ionic strength enhanced the emission, owing to the decrease of the
swelling degree. The PGA-­diglycidyl system was also employed for another AIEgen, boron keto-
iminate (33), by Tanaka et  al.  [55]. The emission maxima were observed at longer wavelength,
500  nm, in H2O or low polar solvents, whereas highly polar organic solvents hyprochromically

HO O
O H HO O
O H O
N H
N N N
m H n N
O l H m n
O O O
O O O O
OH
OH OH

O
O O

O
F
OH O
B
O O O HO F
N
O O
N
O H m N
C O H l
C
32 33

m n
O O O O
O Si O O
m

O O O O O O
Si O
mn
34 n
35 36

Figure 9.10  Chemically crosslinked gels and crosslinkers derived from AIEgens.


262 9  Homogeneous Systems to Induce Emission of AIEgens

shifted the emission maxima owing to the local aggregation in the gel. Additionally, mixing with a
protein, bovine serum albumin, resulted in hypsochromic shift even in H2O.
We were the first to explore a conventional vinyl crosslinker, in which TPE with two or four acry-
loyl groups was used to synthesize network polymers derived from polystyrene (34), polyacrylate, or
polymethacrylate [56]. The ΦF of the obtained network polymers increased upon increasing Young’s
moduli, and this dependence was emphasized for samples derived from the crosslinker with two
acryloyl groups owing to relatively sparse environment around the crosslinker. Fukushima et al.
reported that a poly(acrylic acid) (PAA) gel containing TPE monoacrylate exhibited emission
enhancement in response to Ca2+ concentration [57]. The linear polymer without a crosslinker also
exhibited an increase of hydrodynamic diameter in the presence of Ca2+ [58]. Recently, Tang et al.
reported that the swelling degree of PAA gel and poly(vinyl alcohol) (PVA) gel corresponded well to
the emission intensity of the contained TPE [59]. Tang et al. also reported shrinkage of a polymer
gel consisting of a random copolymer of acrylamide and sodium 4-­styrene sulfonate, which was
mediated by protonation of the included tetra(4-­pyridyl)TPE (13) [60]. The protonated 13 acted as a
potential crosslinker of the gel; thus, the swelling degree gradually decreased with bathochromic
emission color change. Additionally, they also prepared a double-­layered gel, with one layer con-
taining 13. The gel exhibited gradual bending upon the protonation of 13.
Poly(dimethyl siloxane) (PDMS) is an important elastomer. We synthesized TPE with
3-­butenyloxy groups as a crosslinker for H-­terminated PDMS. The resulting PDMS elastomer (35)
exhibited strong emission when collapsed by drying or a poor solvent, and the emission was
quenched after swelling in a good solvent, which was repeated without any fatigue. Additionally,
the emission was drastically enhanced at low temperature [61]. Ma et al. reported the preparation
of a PDMS elastomer containing bare TPE. The composite exhibited temperature-­dependent emis-
sion intensity, where the emission almost disappeared at 100 °C. Additionally, the emission inten-
sity decreased upon stretching the composite; after applying 70% strain, the emission intensity
decreased by half.
A crystal-­crosslinked gel (CCgel) is one of the most structured gels templated by a metal–organic
framework (MOF)  [62–64]. By synthesizing TPE with propargyl groups, we prepared a CCgel
crosslinked by an AIEgen (36)  [65]. Unlike soft conventional gels, the emission was enhanced
when the CCgel was swollen, owing to the increase of scattering when it was collapsed. The CCgel
exhibited solvent-­dependent emission intensity; thus, good solvents resulted in high emission
intensity. Moreover, the abundant carboxylates in the CCgel, which were originally used for coor-
dination to metal ions, could be reused for coordination to other metal ions. In addition, coordina-
tion to luminescent rare-­earth ions such as Eu3+ and Tb3+ resulted in a change of the emission
color of the CCgel to white owing to the mixture of the two colors from the CCgel and the rare-­
earth ions.
In these examples, the AIEgens were basically homogeneously dispersed in chemically
crosslinked gels, and the change of the environment drastically affected their emission behavior.

9.4.2  Physically Crosslinked Gels


Many examples of physically crosslinked AIEgen gels have been reported; however, herein, we
would like to introduce examples of physically crosslinked AIEgen gels in which the supramolecu-
larly interacting groups apparently existed apart from the AIEgen moiety (Figure 9.11).
Liang et al. synthesized TPE with an oligopeptide chain including the QQEFQFQFKQQ sequence
(37), which is known to form a β-­sheet fibril structure [66]. The addition of an aqueous salt readily
resulted in gelation with emission enhancement, and the emission intensity increased with
9.4  ­Gels and Network Polymer 263

O
O
N KRKR SGSG QQEFQFQFKQQ
H

2Br–
+ +
N N
O O
37 38

(HO)2B
O NH

H
N O
B(OH)2
N O HN
39
40 N

N+
O 2Br– O
O O O O
O O O
O
O
O
O O O O
O O
O O
O
O O

+N
41

Figure 9.11  AIEgens forming physically cross-­linked gels.

increasing salt concentration. The use of trypsin resulted in transformation of the gel into a sol after
12 hours; however, the strong emission remained owing to the partial digestion of the fibril structure.
Huang et  al. reported the formation of a gel consisting of poly(sodium 4-­styrene sulfonate)
and TPE with trimethylammonium groups (38) via electrostatic interaction [67]. Upon increas-
ing the concentration of the ionic polymer, the emission intensity was enhanced, and the hydro-
gel was formed. The addition of ATP to the hydrogel decreased the emission intensity, and the
hydrogel was destroyed, owing to the multivalent ATP interacting more strongly with 38 than
the ionic polymer. Further addition of ATPase led to recovery of the hydrogel through the diges-
tion of ATP to phosphates and adenine. Tang et  al. presented a novel method to observe the
gelation process of a PVA hydrogel physically crosslinked by TPE diboronic acid (39) [68]. The
repeating freeze–thaw cycles increased the emission intensity and mechanical strength and
decreased the swelling degree.
Lin et  al. reported the formation of a supramolecular hydrogel consisting of N1,N3,N5-­
tri(pyridin-­4-­yl)benzene-­1,3,5-­tricarboxamide (40) treated with 3-­fluorobenzaldehyde  [69]. The
hydrogel was melted by heating owing to the cleavage of hydrogen bonding. The same research
group also reported the formation of a supramolecular organogel from the pillar[5] arene-­based
compound 41 in 1-­butanol [70]. The organogel exhibited strong emission, and the emission drasti-
cally decreased with the addition of Fe3+, Cr3+, or Hg2+ compared with Zn2+, Pd2+, Cd2+, Ni2+,
Co2+, Ag+, Ca2+, Cu2+, Mg2+, Ba2+, Tb3+, Eu3+, La3+, or Al3+. Moreover, recovery of the emission of
264 9  Homogeneous Systems to Induce Emission of AIEgens

the Fe3+ or Cr3+-­absorbed organogel was observed after the addition of cyanide (CN−) or phos-
phate (PO43−), respectively. Therefore, such an organogel could be applied as a metal ion or anion
sensor with ultrasensitive limit of detection (LOD).

9.5 ­Crystalline Materials

In a crystal, molecules are located adjacent to each other; however, the adjacent molecule may be
a different molecule or chemical species. Using such an environment for AIEgen crystals has ena-
bled the emission from AIEgens to be tuned using a conjugated microporous polymer (CMP),
covalent-­organic framework, MOF, periodic mesoporous organosilica, or cocrystal (Figure 9.12).

O O
Br Br
HO OH

Br Br HO OH
42 43 44
O O

O OH HO O

O O

OH OH HO OH

O 46 O

HO OH
OH OH

O O

45
O OH HO O HO OH
47

EtO OEt
OEt EtO
Si Si
EtO OEt
NH HN
O O
O O

N N
O O
EtO N N OEt
Si N O O N Si
EtO H H EtO OEt
OEt 48

Figure 9.12  Molecules forming crystalline materials derived from AIEgens.


9.5 ­Crystalline Material 265

Jiang et al. reported the preparation of CMP synthesized by Yamamoto coupling of tetra(4-­
bromophenyl)ethylene (42)  [71]. The obtained CMP exhibited a very high surface area
(~1665  m2/g) after a 72-­hours reaction. Although only weak emission was observed for the
monomer (440  nm), a linear polymer (506  nm) derived from di(4-­bromophneyl)ethylene in
THF, the CMP exhibited strong emission at 550 nm in THF, indicating the restriction of twist
and distortion in the excited state. Zhao et al. reported the synthesis of CMP from TPE ­diboronic
acid (39) or diethynyl TPE (43) reacting with tetra(4-­bromophenyl)methane or tetra(4-­
bromophenyl)adamantane via Suzuki–Miyaura or Sonogashira–Hagihara coupling  [72]. The
obtained CMP exhibited a high surface area up to 900 m2/g and strong emission. The emission
was further increased by the inclusion of benzene, toluene, or mesitylene but quenched by the
inclusion of nitrobenzene.
Dinča et al. presented a series of work on MOF using an organic ligand derived from TPE. In
2012, they reported the preparation of MOF from tetra(4-­carboxyphenyl)ethylene (44) and Zn2+ or
Cd2+ exhibiting high surface area (~300 m2/g) [73]. The obtained MOF exhibited strong emission
(ΦF = 35–39%) in the presence or absence of N,N-­diethyl formamide (DEF) as the crystallization
solvent. The emission maxima slightly changed with the inclusion of other solvents such as ethyl-
enediamine, cyclohexanone, and acetaldehyde. A molecular dynamics simulation was performed
in which the 16 protons on 44 were substituted with deuterium  [74]. Variable-­temperature 2H
NMR spectroscopy and simulation indicated that the activation energy for phenyl ring flipping was
43(6)  kJ/mol with 2.2  ×  1011  Hz, which can easily proceed at room temperature. The enforced
activation energy originated from the coordination in a rigid MOF lattice. The researchers then
synthesized another organic ligand, tetrakis[4-­(3,5-­dicarboxyohenylethynyl)-­phenyl]ethylene (45),
and prepared a MOF from 45 and Zn2+ [75]. Despite the increased number of connecting points to
the metal nodes, ΦF decreased to 9%, most likely due to the high intrinsic strain energy. The authors
also presented the usage of the MOF as a sensor for ammonia [76]. This MOF exhibited a selective
bathochromic shift of ~25 nm after exposure to ammonia at 100 °C, whereas triethylamine, ethyl-
enediamine, and nitrogen gas did not result in this bathochromic shift.
Zhao et al. reported the preparation of a MOF from 4,4′-­(2,2′-­diphynylethen-­1,1-­diyl)dibenzoic
acid (46) and Zn2+ [77]. The obtained MOF possessed a Kagome-­type 2D sheet structure and exhib-
ited a high surface area (~1389 m2/g). The MOF exhibited strong emission with ΦF of 15%, which
was much less than that of 46 (79%). The inclusion of solvents other than the crystallization sol-
vent, DEF, resulted in a slight shift of the emission maxima and enhancement of the emission
intensity.
We investigated emission control by the component of cocrystals including the AIEgen tetra(4-­
hydroxyphenyl)ethylene (47) and various cyclic compounds including nitrogen atom as the hydro-
gen bonding acceptor (HBA)  [78]. In accordance with the orbital energy of HBA, the emission
intensity was drastically controlled; thus, HBA with higher LUMO energy than that of 45 resulted
in strong emission owing to prohibition of photo-­induced energy transfer-­derived quenching.
Han et al. reported the incorporation of the tetranitrilomethylidyne–hexaphenyl derivative (48)
to PMO [79]. The obtained PMO had a high surface area (~300 m2/g). The incorporated AIEgen
exhibited enhanced emission, analogously to the original AIEgen, and the ΦF doubled (46%→84%).
PMO exhibited selective quenching for Cu2+ even in the presence of other metal ions such as Ag+,
Al3+, Ba2+, Ca2+, Cd2+, Co2+, Cr3+, Fe2+, Fe3+, Hg2+, K+, Mg2+, Mn2+, Na+, Ni2+, Pb2+, and Zn2+.
The coordination of the AIEgen to Cu2+ was responsible for the quenching with a low LOD
of 40 nM.
Although the AIEgens were fixed in a rigid crystal lattice or matrix in these examples, they could
contact other molecules or chemical species. Thus, the emission of AIEgens can be feasibly tuned
by various factors such as a change in torsional energy or orbital energy.
266 9  Homogeneous Systems to Induce Emission of AIEgens

9.6 ­Outlook and Future Perspectives

As reviewed in this chapter, there have been many reports of homogeneous systems that exhibit
the strong emission of AIEgens without forming an aggregate of the AIEgens. Therefore, “aggrega-
tion” is only one means to induce strong emission of AIEgens, and the name “AIE” often causes
misunderstanding of this emission phenomenon. The most important factor to induce the emis-
sion is restriction of the π twist and distortion of AIEgens in the excited state, which leads to the
relaxation pathway to MECI.
From the revealed working mechanism of the strong emission of AIEgens, many new AIEgens
can be designed and synthesized based on theoretical computations and using state-­of-­art syn-
thetic techniques. However, the research examples are unevenly distributed on TPE; thus, new
AIEgens are expected to be discovered in the near future.
In addition, there are few examples of true AIE, where the materials exhibit strong emission
only in aggregate such as polystyrene [80], polymaleimide [81], tetraphenylethane [82], or polysac-
charide [83]. The formation of supramolecular orbitals would be responsible for this phenomenon,
and this “true AIE” should be explored next to identify the origin of this strong emission.

­References

1 Luo, J., Xie, Z., Lam, J.W.Y., Cheng, L., Tang, B.Z., Chen, H., et al. (2001). Aggregation-­induced
emission of 1-­methyl-­1,2,3,4,5-­pentaphenylsilole. Chemical Communications 381(18): 1740–1741.
2 Mei, J., Leung, N.L.C., Kwok, R.T.K., Lam, J.W.Y., Tang, B.Z. (2015). Aggregation-­induced
emission: Together we shine, United We Soar! Chemical Reviews 115(21): 11718–11940.
3 Tong, H., Hong, Y., Dong, Y., Häußler, M., Lam, J.W.Y., Li, Z., et al. (2006). Fluorescent “light-­up”
bioprobes based on tetraphenylethylene derivatives with aggregation-­induced emission
characteristics. Chemical Communications (35): 3705–3707.
4 An, B.-­K., Kwon, S.-­K., Jung, S.-­D., Park, S.Y. (2002). Enhanced emission and its switching in
fluorescent organic nanoparticles. Journal of the American Chemical Society 124(48): 14410–14415.
5 Hu, R., Lam, J.W.Y., Liu, Y., Zhang, X., Tang, B.Z. (2013). Aggregation-­induced emission of
tetraphenylethene-­hexaphenylbenzene adducts: Effects of twisting amplitude and steric hindrance
on light emission of nonplanar fluorogens. Chemistry -­A European Journal 19(18): 5617–5624.
6 Kokado, K., Chujo, Y. (2009). Emission via aggregation of alternating polymers with o-­carborane
and p-­phenylene-­ethynylene sequences. Society 42(5): 1418–1420.
7 Chen, J., Law, C.C.W., Lam, J.W.Y., Dong, Y., Lo, S.M.F., Williams, I.D., et al. (2003). Synthesis,
light emission, nanoaggregation, and restricted intramolecular rotation of 1,1-­substituted
2,3,4,5-­tetraphenylsiloles. Chemistry of Materials 15(7): 1535–1546.
8 Yang, Z., Qin, W., Leung, N.L.C., Arseneault, M., Lam, J.W.Y., Liang, G., et al. (2015). A
mechanistic study of AIE processes of TPE luminogens: Intramolecular rotation vs.
configurational isomerization. Journal of Materials Chemistry C 4: 99–107.
9 Shi, J., Suarez, L.E.A., Yoon, S., Varghese, S., Serpa, C., Park, S.Y., et al. (2017). Solid state
luminescence enhancement in π-­conjugated materials: Unraveling the mechanism beyond the
framework of AIE/AIEE. The Journal of Physical Chemistry C 121: 23166–23183.
10 Sharafy, S., Muszkat, K.A. (1971). Viscosity dependence of fluorescence quantum yields. Journal of
the American Chemical Society 93(17): 4119–4125.
11 Barbara, P.F., Rand, S.D., Rentzepis, P.M. (1981). direct measurements of tetraphenylethylene
torsional motion by picosecond spectroscopy. Journal of the American Chemical Society 103(9):
2156–2162.
 ­Reference 267

12 Greene, B.I. (1981). Observation of a long-­lived tristed intermediate following pocosecond UV


excitation of tetraphenylethylene. Chemical Physics Letters 79(1): 51–53.
13 Schilling, C.L., Hilinski, E.F. (1988). Dependence of the lifetime of the twisted excited singlet state
of tetraphenylethylene on solvent polarity. Journal of the American Chemical Society 110(7):
2296–2298.
14 Morais, J., Ma, J., Zimmt, M.B. (1991). Solvent dependence of the twisted excited state energy of
tetraphenylethylene: Evidence for a zwitterionic state from picosecond optical calorimetry. Journal
of Physical Chemistry 95(10): 3885–3888.
15 Ma, J., Zimmt, M.B., Dutt, G.B., Waldeck, D.H. (1994). The excited state potential energy surface
for the photoisomerization of tetraphenylethylene: A fluorescence and picosecond optical
calorimetry investigation. Journal of the American Chemical Society 116(23): 10619–10629.
16 Schuddeboom, W., Jonker, S.A., Warman, J.M., de Haas, M.P., Vermeulen, M.J.W., Jager, W.F., et al.
(1993). Sudden polarization in the twisted, phantom state of tetraphenylethylene detected by
time-­resolved microwave conductivity. Journal of the American Chemical Society 115(8):
3286–3290.
17 Luo, J., Song, K., Gu, F. L., Miao, Q. (2011). Switching of non-­helical overcrowded
tetrabenzoheptafulvalene derivatives. Chemical Science 2(10): 2029–2034
18 Leung, N.L.C., Xie, N., Yuan, W., Liu, Y., Wu, Q., Peng, Q. (2014). Restriction of intramolecular
motions: The general mechanism behind aggregation-­induced emission. Chemistry -­A European
Journal 20: 15349–15353.
19 Peroche, S. (2003). Novel fluorinated amphiphilic cyclodextrin derivatives: Synthesis of mono-­,
di-­and heptakis-­(6-­deoxy-­6-­perfluoroalkylthio) -­b-­cyclodextrins. Tetrahedron Letters 44: 241–245.
20 Peng, X.L., Ruiz-­Barragan, S., Li, Z.S., Li, Q.S., Blancafort, L. (2016). Restricted access to a conical
intersection to explain aggregation induced emission in dimethyl tetraphenylsilole. Journal of
Materials Chemistry C 4(14): 2802–2810.
21 Kokado, K., Sada, K. (2019). Consideration of molecular structure in the excited state to design
new luminogens with aggregation-­induced emission. Angewandte Chemie -­International Edition
58(26): 8632–8639.
22 Gale, P.A. (2011). Anion receptor chemistry. Chemical Communications 47(1): 82–86.
23 Noguchi, T., Roy, B., Yoshihara, D., Tsuchiya, Y., Yamamoto, T., Shinkai, S. (2014). Cyclization-­
induced turn-­on fluorescence system applicable to dicarboxylate sensing. Chemistry -­A European
Journal 20(2): 381–384.
24 Zhang, Z., Hashim, M.I., Wu, C.H., Wu, J.I., Miljanić, O. (2018). Discrimination of dicarboxylic
acids: Via assembly-­induced emission. Chemical Communications 54(82): 11578–11581.
25 Zhao, J., Yang, D., Zhao, Y., Yang, X.J., Wang, Y.Y., Wu, B. (2014). Anion-­coordination-­induced
turn-­on fluorescence of an oligourea-­functionalized tetraphenylethene in a wide concentration
range. Angewandte Chemie -­International Edition 53(26): 6632–6636.
26 Kassl, C.J., Christopher Pigge, F. (2014). Anion detection by aggregation-­induced enhanced
emission (AIEE) of urea-­functionalized tetraphenylethylenes. Tetrahedron Letters 55(34):
4810–4813.
27 Sinha, N., Stegemann, L., Tan, T.T.Y., Doltsinis, N.L., Strassert, C.A., Hahn, F.E. (2017). Turn-­on
fluorescence in tetra-­NHC ligands by rigidification through metal complexation: An alternative to
aggregation-­induced emission. Angewandte Chemie -­International Edition 56(10): 2785–2789.
28 Wang, N., Zhang, J., Xu, X.D., Feng, S. (2020). Turn-­on fluorescence in a pyridine-­decorated
tetraphenylethylene: The cooperative effect of coordination-­driven rigidification and silver ion
induced aggregation. Dalton Transactions 49(6): 1883–1890.
29 Yan, Q.Q., Hu, S.J., Zhang, G.L., Zhang, T., Zhou, L.P., Sun, Q.F. (2018). Coordination-­enhanced
luminescence on tetra-­phenylethylene-­based supramolecular assemblies. Molecules 23(2): 363.
268 9  Homogeneous Systems to Induce Emission of AIEgens

30 Li, H., Xie, T.Z., Liang, Z., Shen, Y., Sun, X., Yang, Y., et al. (2019). Adjusting emission wavelength
by tuning the intermolecular distance in charge-­regulated supramolecular assemblies. Journal of
Physical Chemistry C 123(37): 23280–23286.
31 Mondal, A., Ahmmed, E., Chakraborty, S., Sarkar, A., Lohar, S., Chattopadhyay, P. (2020).
Aggregation induced emission enhancement (AIEE) of naphthalene-­appended organic moiety:
An Al3+ ion selective turn-­on fluorescent probe. ChemistrySelect 5(1): 147–155.
32 Song, S., Zheng, H.F., Li, D.M., Wang, J.H., Feng, H.T., Zhu, Z.H., et al. (2014). Monomer emission
and aggregate emission of TPE derivatives in the presence of γ-­cyclodextrin. Organic Letters 16(8):
2170–2173.
33 Jiang, R., Wang, S., Li, J. (2016). Cucurbit[7]uril-­tetraphenylethene host-­guest system induced
emission activity. RSC Advances 6(6): 4478–4482.
34 Liu, Y.C., Wang, Y.Y., Tian, H.W., Liu, Y., Guo, D.S. (2016). Fluorescent nanoassemblies between
tetraphenylethenes and sulfonatocalixarenes: A systematic study of calixarene-­induced
aggregation. Organic Chemistry Frontiers 3(1): 53–61.
35 Huang, Y., Ji, J., Zhang, J., Wang, F., Lei, J. (2019). Host-­guest recognition-­regulated aggregation-­
induced emission for in situ imaging of MUC1 protein. Chemical Communications 56(2): 313–316.
36 Li, S., Langenegger, S.M., Häner, R. (2013). Control of aggregation-­induced emission by DNA
hybridization. Chemical Communications 49(52): 5835–5837.
37 Zhu, L., Zhou, J., Xu, G., Li, C., Ling, P., Liu, B., et al. (2018). DNA quadruplexes as molecular
scaffolds for controlled assembly of fluorogens with aggregation-­induced emission. Chemical
Science 9(9): 2559–2566.
38 Huang, P.C., Lin, L.Y., Yang, D.J., Hong, J.L. (2015). Rigid Jeffamine-­included polyrotaxane as
hydrogen-­bond template for salicylideneazine with aggregation-­enhanced emission. RSC Advances
5(48): 37979–37987.
39 Kang, Q., Xiao, Y., Hu, W., Wang, Y. (2018). Smartly designed AIE triazoliums as unique targeting
fluorescence tags for sulfonic biomacromolecule recognition: Via “electrostatic locking.” Journal of
Materials Chemistry C 6(46): 12529–12536.
40 Wang, X., Jiang, Q., Man, Y., Feng, S., Lee, Y.I., Liu, H.G. (2018). A novel amphiphilic pH-­
responsive AIEgen for highly sensitive detection of protamine and heparin. Sensors and Actuators,
B: Chemical 261: 233–240.
41 Zhou, J., Chang, Z., Jiang, Y., He, B., Du, M., Lu, P., et al. (2013). From tetraphenylethene to
tetranaphthylethene: Structural evolution in AIE luminogen continues. Chemical Communications
49(25): 2491–2493.
42 Liang, G., Lam, J.W.Y., Qin, W., Li, J., Xie, N., Tang, B.Z. (2014). Molecular luminogens based on
restriction of intramolecular motions through host-­guest inclusion for cell imaging. Chemical
Communications 50(14): 1725–1727.
43 Zhao, Z., He, B., Nie, H., Chen, B., Lu, P., Qin, A., et al. (2014). Stereoselective synthesis of folded
luminogens with arene-­arene stacking interactions and aggregation-­enhanced emission. Chemical
Communications 50(9): 1131–1133.
44 Luo, W., Nie, H., He, B., Zhao, Z., Peng, Q., Tang, B.Z. (2017). Spectroscopic and theoretical
characterization of through-­space conjugation of foldamers with a tetraphenylethene hinge.
Chemistry -­A European Journal 23(71): 18041–18048.
45 Zhang, G.F., Chen, Z.Q., Aldred, M.P., Hu, Z., Chen, T., Huang, Z., et al. (2014). Direct validation of
the restriction of intramolecular rotation hypothesis via the synthesis of novel ortho-­methyl
substituted tetraphenylethenes and their application in cell imaging. Chemical Communications
50(81): 12058–12060.
 ­Reference 269

46 Gao, Y.J., Chang, X.P., Liu, X.Y., Li, Q.S., Cui, G., Thiel, W. (2017). Excited-­state decay paths in
tetraphenylethene derivatives. Journal of Physical Chemistry A 121(13): 2572–2579.
47 Wang, J.H., Feng, H.T., Luo, J., Zheng, Y.S. (2014). Monomer emission and aggregate emission of
an imidazolium macrocycle based on bridged tetraphenylethylene and their quenching by C60.
Journal of Organic Chemistry 79(12): 5746–5751.
48 Tong, H., Dong, Y., Hong, Y., Haussier, M., Lam, J.W.Y., Sung, H.H.Y., et al. (2007). Aggregation-­
induced emission: Effects of molecular structure, solid-­state conformation, and morphological
packing arrangement on light-­emitting behaviors of diphenyldibenzofulvene derivatives. Journal
of Physical Chemistry C 111(5): 2287–2294.
49 Xiong, J.B., Yuan, Y.X., Wang, L., Sun, J.P., Qiao, W.G., Zhang, H.C., et al. (2018). Evidence for
aggregation-­induced emission from free rotation restriction of double bond at excited state.
Organic Letters 20(2): 373–376.
50 Shi, J., Chang, N., Li, C., Mei, J., Deng, C., Luo, X., et al. (2012). Locking the phenyl rings of
tetraphenylethene step by step: Understanding the mechanism of aggregation-­induced emission.
Chemical Communications 48(86): 10675–10677.
51 Kokado, K., Machida, T., Iwasa, T., Taketsugu, T., Sada, K. (2018). Twist of C═C bond plays a
crucial role in the quenching of AIE-­active tetraphenylethene derivatives in solution. Journal of
Physical Chemistry C 122(1): 245–251.
52 Machida, T., Taniguchi, R., Oura, T., Sada, K., Kokado, K. (2017). Liquefaction-­induced emission
enhancement of tetraphenylethene derivatives. Chemical Communications 53(15): 2378–2381.
53 Takeda, T., Yamamoto, S., Mitsuishi, M., Akutagawa, T. (2018). Thermoresponsive amphipathic
fluorescent organic liquid. Journal of Physical Chemistry C 122(17): 9593–9598.
54 Kokado, K., Nagai, A., Chujo, Y. (2010). Poly (γ-­glutamic acid) hydrogels with water-­sensitive
luminescence derived from aggregation-­induced emission of o-­carborane. Society 43(15):
6463–6468.
55 Suenaga, K., Yoshii, R., Tanaka, K., Chujo, Y. (2016). Sponge-­type emissive chemosensors for the
protein detection based on boron ketoiminate-­modifying hydrogels with aggregation-­induced
blueshift emission property. Macromolecular Chemistry and Physics 217(3): 414–421.
56 Kokado, K., Taniguchi, R., Sada, K. (2015). Rigidity-­induced emission enhancement of network
polymers crosslinked by tetraphenylethene derivatives. Journal of Materials Chemistry C 3(33):
8504–8509.
57 Ishiwari, F., Hasebe, H., Matsumura, S., Hajjaj, F., Horii-­Hayashi, N., Nishi, M., et al. (2016).
Bioinspired design of a polymer gel sensor for the realization of extracellular Ca2+ imaging.
Scientific Reports 6: 1–12.
58 Morishima, K., Ishiwari, F., Matsumura, S., Fukushima, T., Shibayama, M. (2017). Mesoscopic
structural aspects of Ca2+-­triggered polymer chain folding of a tetraphenylethene-­appended
poly(acrylic acid) in relation to its aggregation-­induced emission behavior. Macromolecules 50(15):
5940–5945.
59 Tavakoli, J., Zhang, H.P., Tang, B.Z., Tang, Y. (2019). Aggregation-­induced emission lights up the
swelling process: A new technique for swelling characterisation of hydrogels. Materials Chemistry
Frontiers 3(4): 664–667.
60 Li, Z., Liu, P., Ji, X., Gong, J., Hu, Y., Wu, W., et al. (2020). Bioinspired simultaneous changes in
fluorescence color, brightness, and shape of hydrogels enabled by AIEgens. Advanced Materials
1906493: 1–10.
61 Taniguchi, R., Yamada, T., Sada, K., Kokado, K. (2014). Stimuli-­responsive fluorescence of AIE
elastomer based on PDMS and tetraphenylethene. Macromolecules 47(18): 6382–6388.
270 9  Homogeneous Systems to Induce Emission of AIEgens

62 Ishiwata, T., Furukawa, Y., Sugikawa, K., Kokado, K., Sada, K. (2013). Transformation of metal-­
organic framework to polymer gel by cross-­linking the organic ligands preorganized in metal-­
organic framework. Journal of the American Chemical Society 135(14): 5427–5432.
63 Furukawa, Y., Ishiwata, T., Sugikawa, K., Kokado, K., Sada, K. (2012). Nano-­and microsized cubic
gel particles from cyclodextrin metal-­organic frameworks. Angewandte Chemie -­International
Edition 51(42): 10566–10569.
64 Ishiwata, T., Kokado, K., Sada, K. (2017). Anisotropically swelling gels attained through axis-­
dependent crosslinking of MOF crystals. Angewandte Chemie -­International Edition 56(10):
2608–2612.
65 Oura, T., Taniguchi, R., Kokado, K., Sada, K. (2017). Crystal crosslinked gels with aggregation-­
induced emissive crosslinker exhibiting swelling degree-­dependent photoluminescence. Polymers
9(1): 19.
66 Zhang, C., Liu, C., Xue, X., Zhang, X., Huo, S., Jiang, Y., et al. (2014). Salt-­responsive self-­assembly
of luminescent hydrogel with intrinsic gelation-­enhanced emission. ACS Applied Materials and
Interfaces 6(2): 757–762.
67 Wang, H., Ji, X., Li, Y., Li, Z., Tang, G., Huang, F. (2018). An ATP/ATPase responsive
supramolecular fluorescent hydrogel constructed: Via electrostatic interactions between
poly(sodium p-­styrenesulfonate) and a tetraphenylethene derivative. Journal of Materials
Chemistry B 6(18): 2728–2733.
68 Tavakoli, J., Gascooke, J., Xie, N., Tang, B.Z., Tang, Y. (2019). Enlightening freeze–thaw process of
physically cross-­linked poly(vinyl alcohol) hydrogels by aggregation-­induced emission fluorogens.
ACS Applied Polymer Materials 1(6): 1390–1398.
69 Fan, Y.Q., Huang, Q., Zhang, Y.M., Wang, J., Guan, X.W., Chen, Y.Y., et al. (2019). Forming a
water-­soluble supramolecular polymer and an AIEE hydrogel: Two novel approaches for highly
sensitive detection and efficient adsorption of aldehydes. Polymer Chemistry 10(47): 6489–6494.
70 Wei, T.B., Ma, X.Q., Fan, Y.Q., Jiang, X.M., Dong, H.Q., Yang, Q.Y., et al. (2019). Aggregation-­
induced emission supramolecular organic framework (AIE SOF) gels constructed from tri-­
pillar[5]arene-­based foldamer for ultrasensitive detection and separation of multi-­analytes. Soft
Matter 15(33): 6753–6758.
71 Xu, Y., Chen, L., Guo, Z., Nagai, A., Jiang, D. (2011). Light-­emitting conjugated polymers with
microporous network architecture: Interweaving scaffold promotes electronic conjugation,
facilitates exciton migration, and improves luminescence. Journal of the American Chemical
Society 133(44): 17622–17625.
72 Dong, J., Tummanapelli, A.K., Li, X., Ying, S., Hirao, H., Zhao, D. (2016). Fluorescent porous
organic frameworks containing molecular rotors for size-­selective recognition. Chemistry of
Materials 28(21): 7889–7897.
73 Shustova, N.B., McCarthy, B.D., Dinc , M. (2011). Turn-­on fluorescence in tetraphenylethylene-­
based metal-­organic frameworks: An alternative to aggregation-­induced emission. Journal of the
American Chemical Society 133(50): 20126–20129.
74 Shustova, N.B., Ong, T.C., Cozzolino, A.F., Michaelis, V.K., Griffin, R.G., Dinc , M. (2012). Phenyl
ring dynamics in a tetraphenylethylene-­bridged metal-­organic framework: Implications for the
mechanism of aggregation-­induced emission. Journal of the American Chemical Society 134(36):
15061–15070.
75 Shustova, N.B., Cozzolino, A.F., Dinc , M. (2012). Conformational locking by design: Relating
strain energy with luminescence and stability in rigid metal-­organic frameworks. Journal of the
American Chemical Society 134(48): 19596–19599.
 ­Reference 271

76 Shustova, N.B., Cozzolino, A.F., Reineke, S., Baldo, M., Dinc , M. (2013). Selective turn-­on
ammonia sensing enabled by high-­temperature fluorescence in metal-­organic frameworks with
open metal sites. Journal of the American Chemical Society 135(36): 13326–13329.
77 Zhang, M., Feng, G., Song, Z., Zhou, Y.P., Chao, H.Y., Yuan, D., et al. (2014). Two-­dimensional
metal-­organic framework with wide channels and responsive turn-­on fluorescence for the
chemical sensing of volatile organic compounds. Journal of the American Chemical Society 136(20):
7241–7244.
78 Jimbo, T., Tsuji, M., Taniguchi, R., Sada, K., Kokado, K. (2018). Control of aggregation-­induced
emission from a tetraphenylethene derivative through the components in the co-­crystal. Crystal
Growth and Design 18(7): 3863–3869.
79 Gao, M., Han, S., Hu, Y., Zhang, L. (2016). Enhanced fluorescence in tetraylnitrilomethylidyne-­
hexaphenyl derivative-­functionalized periodic mesoporous organosilicas for sensitive detection of
copper(II). Journal of Physical Chemistry C 120(17): 9299–9307.
80 Yanari, S.S., Bovey, F.A., Lumry, R. (1963). Fluorescence of styrene homopolymers and
copolymers. Nature 200(1): 242–244.
81 Zhao, E., Lam, J.W.Y., Meng, L., Hong, Y., Deng, H., Bai, G., et al. (2015). Poly[(maleic anhydride)-­
alt-­(vinyl acetate)]: A pure oxygenic nonconjugated macromolecule with strong light emission and
solvatochromic effect. Macromolecules 48(1): 64–71.
82 Zhang, H., Zheng, X., Xie, N., He, Z., Liu, J., Leung, N.L.C., et al. (2017). why do simple molecules
with “isolated” phenyl rings emit visible light? Journal of the American Chemical Society 139(45):
16264–16272.
83 Ruff, Y., Lehn, J.M. (2008). Glycodynamers: Fluorescent dynamic analogues of polysaccharides.
Angewandte Chemie -­International Edition 47(19): 3556–3559.
273

10

Hetero-­aggregation-­induced Tunable Emission (HAITE) Through


Cocrystal Strategy
Yinjuan Huang1 and Qichun Zhang2
1
 School of Materials Science and Engineering, Nanyang Technological University, Singapore, Singapore
2
 Department of Chemistry, Department of Materials Science and Engineering, City University of Hong Kong, Hong Kong, China

10.1 ­Introduction

Organic luminescent materials with several charming advantages including easy processability,
enough brightness, high flexibility, tunable photophysical properties [1–3], and appreciable bio-
compatibility [4, 5] are crucial for human beings to illuminate information [6, 7] via optoelectronic
devices, photonics, and biological probes. However, the well-­known aggregation-­caused quench-
ing (ACQ) effect widely exists in the conventional organic chromophores [4, 8–10]. The origins of
ACQ effect mainly come from π–π stacking, energy transfer, excited states reactions, and inter-­/
intramolecular charge transfer [11]. This effect greatly impedes the optoelectronic and biological
applications of these traditional chromophores [6, 7]. Therefore, developing some new strategies
to address this issue is very important and highly desirable.
In fact, a few approaches toward the reduction of ACQ effect have been reported [4, 10], ­including
controlling the formation of J-­aggregates, introducing amorphous domains to decrease crystalline
degree, managing the triplet formation via intersystem crossing process (ISC) [12–14], and grafting
bulky groups onto the framework of luminescent cores to prevent π–π interaction between conju-
gated planes [15]. On the other hand, different from the molecules with ACQ effect, a large num-
ber of fluorescent materials with the famous aggregation-­induced emission (AIE) property have
been discovered and widely applied in various fields recently [8, 9, 15–17].
Nevertheless, there are still some serious problems remaining in the above-­mentioned methods
toward solid/aggregated fluorescent materials. For instance, the molecular stacking orientation
and conformation were difficult to be controlled [18, 19], or avoiding slow ISC process may lead to
retain significant loss of the residual triplet state [20]. As to the AIE molecules, some of them are
very difficult to be synthesized, which causes much higher cost [4, 10, 20, 21]. Therefore, develop-
ing facile strategies to overcome above-­mentioned issues is very important.
Recently, a new phenomenon of hetero-­aggregation-­induced tunable emission (HAITE) through
a facile cocrystal strategy emerged, where nonfluorescent/poor fluorescent molecules are used as
molecular barriers to form cocrystals with common chromophores via noncovalent intermolecular
interactions. The cocrystal strategy can not only enhance the solid-­state luminescence of these

Handbook of Aggregation-Induced Emission: Volume 1 Tutorial Lectures and Mechanism Studies, First Edition.
Edited by Youhong Tang and Ben Zhong Tang.
© 2022 John Wiley & Sons Ltd. Published 2022 by John Wiley & Sons Ltd.
274 10  Hetero-­aggregation-­induced Tunable Emission (HAITE) Through Cocrystal Strategy

chromophores greatly but also result in easily tunable emission color [5, 20, 22–27]. We believe
that this HAITE via cocrystal strategy may open more avenues for the applications of originally
nonfluorescent or poor fluorescent chromophores in bioimaging or optical devices and provide
constructive guidelines to fabricate more highly emissive solid materials through using traditional
ACQ molecules as raw materials. This approach would greatly expand the family of solid emissive
materials for optoelectronic devices, photonics, and biological probes.
In this chapter, we will focus on the introduction of organic cocrystals with HAITE including the
preparation and characterization methods as well as their luminescence properties. Moreover, we will
also emphasize on the description and summary of the HAITE mechanism through cocrystal strategy.

10.2 ­Interactions Within Organic Cocrystals

Organic cocrystals can form from two or more different molecules via noncovalent intermolecular
interactions [28–31], which are accompanied by an ordered stacking of donors (D) and acceptors
(A) and the designable multifunctional synergistic effects  [32]. Organic cocrystals, which are
regarded as an effective and convenient way to fabricate excellent multifunctional optoelectronic
materials, were discovered in 1844 [33]. But, they began to attract much more attention till 1973,
when the charge-­transfer cocrystals with excellent conductivity were obtained [34, 35]. After that,
organic cocrystals have become active fields [36] and have attracted increasing attention in both
optoelectronic devices [2, 37–40] and biological fields [41].
Generally, there are one or more noncovalent interactions within a cocrystal, such as hydrogen
and halogen bonds, π–π interactions, and charge-­transfer (CT) interactions (Figure 10.1). Actually,
CT interactions may act as the main driving force to form cocrystals when a strong charge acceptor
crystallizes with a proper charge donor (Figure 10.1d) [34, 35, 42], where the CTs occur from the
highest occupied molecular orbital (HOMO) of the donors (electron-­rich) to the lowest unoccu-
pied molecular orbital (LUMO) of the acceptors (electron-­deficient) [43]. The strength of CT inter-
actions can be quantified through infrared (IR) and Raman measurements  [44, 45], which are
determined by the electron affinity energy of acceptors, the ionization potential of donors, and the
electrostatic Coulomb force between the donor–acceptor pairs  [43]. Although numerous CT
cocrystals have been developed, the limited CT complexes make it difficult to achieve expected
functions. To overcome this situation, arene–perfluoroarene (AP) complexes are proposed as alter-
natives, in which perfluoroarene may cocrystallize with arene or its derivatives to form 1 : 1 com-
plex with nearly parallel molecular stacking alternately in the cocrystal [3].

(a) (b) (c) (d)

X H Y R X Y
R = C, halogen, N,...
X = F, O, N,...
X = I, Br, Cl, F,... Electron-rich donor
Y = F, O, N,... Electron-deficient
Y = N, O, S, Se,..
I–, Br–, Cl–, F–,... acceptor
π–π Interactions Hydrogen bonds
Halogen bonds CT interactions

Figure 10.1  (a) Schematics of π–π stacking, (b) hydrogen bonds, (c) halogen bonds, and (d) CT interactions.
Source: Reprinted with permission from Ref. [46]. Copyright 2019 Wiley-­VCH.
10.3  ­Preparation of Organic Cocrystal 275

Whereas, the stable hydrogen bonds may exist in most of the cocrystals (Figure 10.1b), which
can direct good orientation, and thus help us to predict the structural consequences easily [46, 47].
Another common interaction in organic cocrystals, namely π–π interactions (Figure 10.1a), can be
formed between π-­donors (e.g. porphyrins and acene-­based derivatives) and π-­acceptors (fuller-
enes) [22, 48–50]. As a thriving force for cocrystal constructions, halogen bonds may form between
halogen atom (donor) and nucleophilic region (acceptor) (Figure 10.1c) [43, 51, 52], and is more
characteristic due to the excellent directional, hydrophobic, and strength-­tunable properties as
well as the higher effectiveness than a hydrogen bond in cocrystal formation  [43, 52]. Strictly
speaking, the cocrystals containing halogen bonds are one kind of CT complexes (involve two
kinds of CT interactions: π and σ) in which the lone electron pairs in the electron-­rich atoms (e.g.
F−, I−, Br−, Cl−, O, S, N, and Se) donate to the halogen acceptor (Cl, F, I, Br) (n → σ* donation) [53].
Generally, the driving forces for cocrystallization are the simultaneous combination of two or more
kinds of the above-­mentioned interactions [43, 52]. In addition, the morphologies of the cocrystals
are correlated to the intermolecular interactions within the crystals. Generally, the orientation of
the largest dimension is determined by the direction along the strongest intermolecular interac-
tions, in which the maximum intermolecular overlap of the π orbits exists.

10.3  ­Preparation of Organic Cocrystals

The main strategies toward organic cocrystals include vapor phase (Figure 10.2a) [54], liquid phase
(Figure 10.2b) [2, 33], as well as solid-­phase (mechanochemical method through grinding the mix-
ture of donors and acceptors with/out adding several drops of solvents) [41] methods. The most

(a) Low temperature


High temperature
sublimation co-crystallization

Carrier gas
Pump

Donor Acceptor Cocrystal

(b) Liquid–liquid Nano-


Solvent interfacial Hot saturated
precipitation
evaporation precipitation solution: cooling

Poor
D and A
solvent
Poor in good
solvent solvent

Figure 10.2  Preparation methods. (a) vapor-­phase method, (b) liquid-­phase methods. Source: Reprinted
with permission from Ref. [46]. Copyright 2019 Wiley-­VCH.
276 10  Hetero-­aggregation-­induced Tunable Emission (HAITE) Through Cocrystal Strategy

widely used method is a solution-­based technique, which is the simplest way (convenient and low
cost) to prepare regular organic cocrystals with high quality. Liquid-­phase technique is a method
to precipitate crystals when supersaturating the solution via varying external conditions, which
include liquid–liquid interfacial precipitation [55], solvent evaporation [56], nanoprecipitation [2],
and cooling of hot saturated solution (Figure 10.2b). Such liquid methods usually require planar
molecular structures and strong intermolecular interactions between donor and acceptor blocks,
most importantly, similar solubilities of donors and acceptors [50]. In contrast, the vapor-­phase
methods are seldom reported [54], although they are essential for cocrystallizing of the poor solu-
bilized organics. Both cosublimation (high-­temperature region) and codeposition (low-­temperature
region) should be performed; therefore, the similar sublimation points of the insoluble molecules
are necessary (Figure  10.2b)  [43, 52]. Solid-­phase technique is an ancient one, which mainly
includes grinding and solvent-­assisted grinding [41], and has been rediscovered recently to meet
the requirement for environmentally friendly and solvent-­free processes [41, 46]. However, unfor-
tunately, the resulting materials from solid-­phase method are labeled with low-­quality phases and
may present different properties from that of the single crystalline phases.

10.4 ­Molecular Stacking Modes Within Organic Cocrystals

The interaction between donors and acceptors as well as their arrangement in the cocrystals have
enormous impact on the properties of the resulted materials. Therefore, the structural regulation
of organic cocrystals is particularly important.
For instance, different D–A arrangement may result in different properties of cocrystals.
Currently, the most common arrangement is the stacking between donor and acceptor (1 , 1, mole
ratio) with two packing modes: mixed stacking mode (. . .DADA. . .) (Figure 10.3a) [57] and segre-
gated stacking mode (DDDD. . .AAAA) (Figure 10.3b) [35] Cocrystals with mixed stacking modes
can be semiconductors or even insulator, and, by contrast, segregated stacking complexes usually
display high electrical conductivity [58]. In addition, packing ratios other than 1 : 1 have also been
found in organic cocrystals if appropriate donors/acceptors as well as the right preparation condi-
tions are applied during cocrystal formation [43]. Within these cocrystals, the extra acceptor/donor
molecules may either exist within the costacks or form a new segregated stack [59].

(a) (b)
Mixed stacking (1 : 1) Segregated stacking (1 : 1)

Figure 10.3  Common packing modes of cocrystals formed via 1 : 1 stoichiometric ratio. Source: Reprinted
with permission from Ref. [46]. Copyright 2019 Wiley-­VCH.
10.6 ­HAITE Through Cocrystal Strateg 277

The structure and morphology of cocrystals can be determined by molecular structures of both
donors and acceptors (intrinsic factors) as well as crystal growth conditions (extrinsic factors).
Different molecular structures may lead to different intermolecular interactions, which ultimately
leads to different molecular arrangements. Therefore, the molecular packing mode can be tuned
by regulating the molecular structure and configuration of either donors or acceptors. For exam-
ple, the crystallization characteristics of the on-­demand cocrystals can be realized by selecting
appropriate molecular structures, suitable molecular configuration, substituents type, etc.  [60].
Although molecular structures play a decisive function in the interactions between donors and
acceptors, external factors can still change the structure and morphology of the resulted cocrystals
to some extent. When using solution strategy, different external conditions (such as solvents, sub-
strate species, and growth temperature) may lead to different single cocrystals with different struc-
tures or morphologies [61].

10.5  ­Characterization of Organic Cocrystals

Single crystals with high quality can be screened through structural and morphological characteri-
zations, such as X-­ray diffraction (XRD), scanning tunneling microscope (STM) and transmission
electron microscopy (TEM), which will be briefly introduced. XRD is a common method for char-
acterizing the inner structure of materials. With the help of XRD, both the molecular packing
mode and the detailed structural information such as the interaction kinds and strength between
D and A molecules can be obtained. A rocking curve can be used to characterize the quality of
cocrystals. Under such mode, the defects in cocrystals can be easily reflected as an increased peak
width. As reported in the literature, most of the crystals produced by vapor-­phase strategy usually
present high crystal quality, and their peak widths are much narrower (less than 0.05) [62]. STM
has high atomic-­level and high-­resolution, which makes it possible to visualize the molecular
alignment on the surface of a single crystal [63]. Except for XRD and STM, the characterizations
by Fourier transform infrared (FTIR), Raman spectroscopy, and differential scanning calorimetry
(DSC) are usually demonstrated as auxiliary characterizations to confirm the interaction between
donors and acceptors within a cocrystal. In addition, scanning electron microscope (SEM) and
TEM are usually used to observe the morphology of crystals. The selected area diffraction (SAED)
under TEM may also give the structural information of the crystal, and it is usually used for deter-
mining the growth direction of a single crystal.

10.6 ­HAITE Through Cocrystal Strategy

Recently, cocrystallization strategy has been proven to be an effective strategy toward both tunable
luminescence colors and regulated intensity. Due to their effectiveness, cocrystals can provide new
properties rather than a simple combination of the properties of donor and acceptor [64, 65].
In order to achieve desirable fluorescent complexes, it is essential to clearly understand which
factors may affect the properties of cocrystals. The direct factors include molecular species, molec-
ular packing methods, degree of charge transfer, and band structure. Specifically, the kinds of
donors and acceptors are decisive factors [66]. Although crystalline complexes usually exhibit dif-
ferent properties from both donors and acceptors, they do maintain the intrinsic characteristic
properties of each coformer to some extent [46]. In this situation, the characteristic properties of
both donors and acceptors should be considered during fabricating cocrystals with new, desired
278 10  Hetero-­aggregation-­induced Tunable Emission (HAITE) Through Cocrystal Strategy

properties. More importantly, in addition to introduce intrinsic characteristics of coformers, the


molecular types of each component can also lead to different CT degree, different crystal struc-
tures, and different band structures, which eventually influences the optoelectronic properties of
cocrystals  [43]. In addition, variable stacking modes of donors and acceptors can also result in
different properties, which has been described in above-­mentioned in Section  10.4  [24].
Furthermore, the degree of charge transfer (q, 0   q  1), a character of CT cocrystal, can be related
to luminescence and conductivity properties [67]. Specifically, the cocrystals with larger q value
(dominated by CT interactions) normally exhibit high conductivity [67]. In addition, q value can
also be related to the molecular packing; e.g. the cocrystals dominated by CT interactions usually
possess strong ionic character, leading to a segregated stacking, conversely resulting in a mixed
stacking [50, 68]. For the band structures factor, it has been proven that the HOMO of donors pro-
vides the largest contribution to that of the ultimate cocrystal, while the LUMO of acceptors mainly
determine that of cocrystals [69]. Moreover, the band structures can also affect the CT behaviors as
well as the stacking structures of the cocrystals. Therefore, in order to implement suitable tailoring
of band structures and obtain a cocrystal with particularly desired properties, it is crucial to have a
deeper understanding of band structures. Except for the above-­mentioned direct factors, there still
is an indirect factor, namely, preparation conditions, which is indispensable. As detailed above,
different preparation conditions may lead to diverse crystal structures, which further results in dif-
ferent CT degree and band structures and ultimately affects the properties of the cocrystals.
Based on the above-­mentioned relationship between molecular structures of coformers and the
properties of resulted cocrystals, a variety of solid materials with tunable fluorescence based on
cocrystal strategy have been reported very recently. Such cocrystals can be named as HAITE mate-
rials due to their characters of crystallizing two or more different molecules as well as the different
luminescence (color, intensity, and lifetime) from that of single-­component chromophore in solid
state after forming solid cocrystals (Figure  10.4 and Table  10.1)  [2, 22, 23, 70–72], which may
greatly expand the family of new solid materials with high-­efficiency luminescence. The details
will be described and discussed below.

10.6.1  HAITE with Tunable Color and Enhanced Emission


Cocrystals, which are formed through crystallizing two or more compounds, have been developed
to modulate and control the physicochemical properties by means of varying the intermolecular
noncovalent interactions (Figure 10.5) [22, 50, 65, 70, 71, 73–75].

10.6.1.1  Insignificant Changed Intensity but Tuned Color


Yan and the coworkers chose suitable coformers to prepare co-­crystals, in which both tunable pho-
toemission and high luminescent efficiency were achieved [22]. 4-­(1-­Naphthylvinyl)pyridine (NP)
and three coformers (1,4-­diiodotetrafluorobenzene [DITFB], 4-­bromo-­2,3,5,6-­tetrafluorobenzoic
acid [BTFBA], and 4-­benzoylbenzoic acid [BBA]) were chosen as donor and acceptors, respectively
(Figure 10.6a). After cocrystallization driven by hydrogen or halogen bonding interactions as well
as π–π interactions between NP and the acceptors, the optical properties of these cocrystals can be
highly tuned, accompanied with HAITE (Figure 10.6a). Specifically, all the absorption and emission
of NP/DITFB, NP/BTFBA, and NP/BBA red-­shifted compared with NP (purple luminescence), and
the corresponding emission colors were tuned to blue, green, and cyan, respectively. The 23 nm red-­
shift of NP/DITFB can be attributed to the enhanced exciton coupling between the pairs of neigh-
boring chromophores, while the evident shifts for NP/BTFBA and NP/BBA result from the O−H⋯N
intermolecular interaction between carboxylate acid groups and pyridine units (Figure 10.6b).
10.6 ­HAITE Through Cocrystal Strateg 279

Donor Acceptor Cocrystal with HAITE


chromophore

+ Cocrystallization
(a)

Tuned color
Insignificant changed PLQY

Cocrystallization
(b) +

Tuned color
Increased PLQY

Cocrystallization
+
(c)

Intrinsic color
Increased PLQY

Figure 10.4  Schematic diagram of HAITE based on cocrystal, represented by the mixed-­stacking (1 : 1)
mode. Three main cases are demonstrated, (a) insignificant changed PLQY but tuned color, (b) increased
PLQY together with tuned color, and (c) increased PLQY but intrinsic color, respectively. The case of
prolonged lifetime (thermally activated delayed fluorescence and phosphorescence) is not presented here.

Table 10.1  Comparison between current HAITE and AIE.

Molecular Emission in Emission in


Names speciesa Preparation method solutionb solid statec Tunable emissiond

AIE One Organic synthesis Weak/Non Strong Enhanced intensity


HAITE Two or Cocrystallization Strong/ Strong Tunable color or/and enhanced
more Weak/Non intensity or/and prolonged lifetime
a
 Number of molecular species required for AIE/HAITE occurrence.
b
 Emission of molecular solution for AIE, and Emission of cocrystal solution for HAITE.
c
 Emission of AIE molecules in solid state for AIE, and Emission of cocrystals for HAITE.
d
 For AIE, compared with the emission of molecular solution; For HAITE, compared with the emission of the
single-­component chromophores in solid state.

In addition, the photophysical properties can also be modified via CT interactions. Yan and the
coworkers put forward vapor deposition method during a continuous heating–cooling process
to fabricate micro-­sized cocrystals based on dibenzothiophene (DBT, donor) and
1,2,4,5-­tetracyanobenzene (TCNB, acceptor)  [76]. The prepared DBT/TCNB cocrystals demon-
strated an obvious red-­shifted emission (λem  =  510  nm) and changed the new emission color
(bright green emission) compared with the pure DBT (λem  =  358  nm, white luminescence)
(Figure  10.7a, b). The band gap of DBT/TCNB cocrystal was calculated to be 2.13  eV and was
Figure 10.5  (a) Structures of the reported donors and acceptors used in HAITE with tunable color and enhanced emission. Each molecule is denoted by a
capital letter. (b) Fluorescence microscopy images of the cocrystals base on the donors and acceptors in (a). Source: Reprinted with permission from Ref. [46].
Copyright 2019 Wiley-­VCH.
(a)

NP NP + DITFB NP + BTFBA NP + BBA

NP/DITFB NP/BTFBA NP/BBA

0% 90%
Powder Powder 0% 90% Powder 0% 90% Powder 0% 90%

(b)
1.
69
2.6

2.79 Å Å 2.58 Å

2.
72
Å

1.63 Å
Å
2.50


2.4

NP/DITFB NP/BTFBA NP/BBA

Figure 10.6  (a) The images of the powders as well as the solutions (in CH3OH/CHCl3–H2O mixture) of NP, NP/DITFB, NP/BTFBA, and NP/BBA under UV (365 nm),
with water fractions of 0 and 90%, respectively. (b) The intermolecular interactions of cocrystals NP/DITFB, NP/BTFBA, and NP/BBA. Source: Reprinted with
permission from Ref. [22]. Copyright 2018 Wiley-­VCH.
(a) (c)
Ex
DBT Em
S S NC
Normalized intensity

CN

NC CN

DBT
DBT/TCNB

π*
250 300 350 400
Wavelength
(b)
Ex
DBT/TCNB Em
π*
Normalized intensity

3.48 eV

2.13 eV

π π

300 400 500 600


Wavelength

Figure 10.7  Fluorescence spectra of (a) DBT (excitation wavelength: 300 nm) and (b) DBT/TCNB microtubes (excitation wavelength: 360 nm) at room
temperature. The black line represents the excitation spectra, and the red line represents the emission spectra. (c) Calculated molecular orbital diagrams for the
pristine DBT (left) and DBT/TCNB cocrystal (right). Source: Reprinted with permission from Ref. [76]. Copyright 2017 Royal Society of Chemistry.
10.6 ­HAITE Through Cocrystal Strateg 283

greatly reduced compared with that of the single-­component DBT (3.48 eV), which is consistent
with the experimental results (2.43 and 3.46 eV). These results were attributed to the formation of
CT states introduced by TCNB, which can be further confirmed by molecular orbital analysis. The
charge transfer from DBT to TCNB was demonstrated by the theoretical calculations (Figure 10.7c).
Furthermore, effective energy transfer (ET) can also change the photophysical properties of
organic luminescent materials [25, 27, 77]. Fu and coworkers [77] developed convenient solution-­
phase method toward coassembly of two structurally similar semiconducting molecules
(5,12-­bis(phenylethynyl)naphthacene (BPN) and 9,10-­bis(phenylethynyl)-­anthracene (BPA)),
where two-­component cocrystals (BPA)x(BPN)1−x with different luminescence from that of BPN
and BPA were achieved. Specifically, BPA tubes emit strong yellow–green light when excited by
blue light (Figure 10.8b), while BPN tubes present dark-­red luminescence when excited by green
light (Figure  10.8c). By comparison, (BPA)x(BPN)1−x cocrystals prepared from different ratios
between BPA and BPN demonstrate tunable emission colors varying from orange–red to near
infrared (NIR). Figure 10.8d shows the (BPA)0.8(BPN)0.2 alloy with NIR emission. Given the effi-
cient intermixing of BPA and BPN in (BPA)x(BPN)1−x coassemblies as well as the good overlap
between the absorption spectrum of BPN and the photoluminescence (PL) spectrum of BPA
(Figure 10.8e), there should be efficient ET process from BPA to BPN in the (BPA)x(BPN)1−x coas-
semblies, which would make the emission colors of the cocrystals more close to that of single-­
component BPN upon decreasing m/m (BPA/BPN). The PL spectra of (BPA)x(BPN)1−x cocrystals,
formed at m/m = 100 : 1, 20 : 1 and 4 : 1 (Figure 10.8f), were collected to confirm this process. The
PL peaks, designated to BPA in the region of 500–600 nm, are almost quenched, whereas the new
peaks from BPN in the region of 600–800 nm gradually red-­shift when the m/m ratio decreases.
Surprisingly, the ET efficiency can be tailored to 100% through the control of m/m below 100 : 1.
Although various CT and ET cocrystals have been prepared, they are nonetheless limited to the
specified D/A pairs, which makes it difficult to achieve more expected functional materials. Recently,
researchers found an exciting result in cocrystals [2, 3, 40], where CT and ET processes can be com-
bined together and multiple luminescent colors including white light emission (WLE) can be
achieved through using three different kinds of components. For example, Liao et al. [2] fabricated
pyrene-­doped naphthalene/TCNB micro cocrystals, which contain both CT and ET processes, and
result in highly efficient blue, WLE and orange. Inspired by this masterpiece, Hu and coworkers [3]
fabricated another tricomponent cocrystal with distinct optical properties by rationally modulating
the two competitive interactions, CT and AP, respectively. Specifically, various emission from blue
through WLE to yellow can be realized via introducing different dozes of TCNB into the pyrene–
octafluoronaphthalene (OFN) host. The AP interactions exist in pyrene/OFN, where the CT interac-
tions lead to the formation of pyrene/TCNB confirmed by spectrum measurements and
density-­functional theory (DFT) calculations. Interestingly, the PL spectrum of pyrene/OFN blue-­
shifts by 59 nm compared with that of pure pyrene, while the PL spectrum of pyrene/TCNB red-­
shifts by 102 nm (Figure 10.9a, b). The small Mulliken atom charge (+0.026), centered on the pyrene
component in pyrene/OFN cocrystals, proves that almost no CT interactions exist between pyrene
and OFN (Figure 10.9c). This is further verified by the calculated LUMO and HOMO (Figure 10.9d),
both of which mainly distribute on pyrene. By comparison, the larger Mulliken atom charge
(+0.046) from pyrene moiety as well as the calculated HOMO and LUMO of pyrene/TCNB confirm
the CT interactions in pyrene–TCNB, in which the LUMO orbital ­distributes on the TCNB moiety,
whereas the HOMO orbital almost locates on the pyrene moiety (Figure 10.9c, d). The molecular
arrangements of pyrene/TCNB and pyrene/OFN display similar stacking modes because of the
strong AP or CT interactions between pyrene and the acceptors (Figure 10.9e, f).
284 10  Hetero-­aggregation-­induced Tunable Emission (HAITE) Through Cocrystal Strategy

(a)

BPA

BPN (BPA)0.72(BPN)0.28
(b) (c) (d)

BPA BPN (BPA)0.8(BPN)0.2

(e) (f)
BPA m/m = 100:1 m/m = 20:1
solution BA tubes
Relative intensity (a.u.)

Relative intensity (a.u.)

m/m = 4:1

BN tubes

BPN
solution

400 500 600 500 600 700 800 900


Wavelength (nm) Wavelength (nm)

Figure 10.8  (a) Schematic illustration of the molecular stacking of (BPA)0.72(BPN)0.28 alloy. Fluorescence
microscopy images of (b) BPA tubes (blue light irradiation), (c) BPN tubes (green light irradiation), and (d)
(BPA)0.8(BPN)0.2 alloy (blue light irradiation), respectively. Scale bars, 25 μm in (b), 50 μm in (c), 50 μm in (d). (e)
Normalized absorption and PL spectra of (red curves) BPN and (blue curves) BPA solutions in tetrahydrofuran
(THF). Inset demonstrates the photographs of the solutions under irradiation of 365 nm light. (f) PL spectra
of BPN tubes, BPA tubes, and (BPA)x(BPN)1−x coassemblies formed at m/m = 100 : 1, 20 : 1, and 4 : 1,
λex = 480 nm. Source: Reproduced with permission from Ref. [77]. Copyright 2018 Springer Nature.

In the TCNB-­doped pyrene/OFN cocrystals, an efficient ET process from pyrene/OFN to pyrene/


TCNB occurred due to their well-­matched spectra and similar crystal structures. According to the
energy-­level diagrams (Figure 10.9d), pyrene–OFN with blue emission was served as energy donor,
where pyrene–TCNB with orange-­emission acts as energy acceptor (Figure 10.10a). Surprisingly,
the TCNB-­doped pyrene–OFN with various doping ratios from 0.5 to 3% demonstrated tunable
emission colors (Figure  10.10a, b). Emission bands at 403  nm derived from pyrene–OFN
10.6 ­HAITE Through Cocrystal Strateg 285

(a) (b)

Pyrene/OFN Pyrene/TCNB
Relative intensity (a.u.)

Relative intensity (a.u.)


CT transistion
Pyrene Pyrene

OFN TCNB

300 400 500 600 700 300 400 500 600 700
(c) Wavelength (nm) (d) Wavelength (nm)
LUMO –1.67 eV LUMO –3.34 eV

+0.026 –0.026 –0.046 +0.046

1.59 D
Pyrene/OFN Pyrene/TCNB

0.96 D

Pyrene/OFN Pyrene/TCNB
HOMO –5.59 eV HOMO –5.84 eV

(e) b c (f) a b

3.35 A

a 0

F 3.33 A

–CN
Pyrene/OFN AP CT Pyrene/TCNB

Figure 10.9  (a) The absorption spectra of pyrene/OFN, (b) pyrene/TCNB cocrystals and (c) their respective
coformers. (d) The calculated static dipole moments and Mulliken charge distributions of pyrene/TCNB and
pyrene/OFN. (e) The calculated molecular orbital diagrams of pyrene/TCNB and pyrene/OFN. Molecular
packing structures of pyrene/OFN (f) and pyrene/TCNB. Source: Reprinted with permission from Ref. [3].
Copyright 2017 Wiley-­VCH.

component reduces gradually, whereas the two splitting peaks at 535 and 575 nm originating from
pyrene–TCNB system enhance upon increasing the doping amount of TCNB (Figure  10.10b).
Furthermore, the fluorescence lifetime of pyrene–OFN can be reduced from 66.2 to 32.3 ns with a
3% doping ratio, further confirming the energy-­transfer process from pyrene/OFN to pyrene/TCNB.
A possible energy-­transfer mechanism was proposed (Figure 10.10c). First, pyrene/OFN donor
and pyrene/TCNB acceptor fulfill the well-­matched crystal structures (Figure 10.10c). Second, the
emission spectrum of pyrene/OFN was well overlapped with the absorption spectrum of pyrene/
TCNB (Figure 10.10d). Therefore, an efficient Förster resonance energy transfer (FRET) process
(a) (b)
Pyrene
Pyrene/OFN 0.5% 1% 3% /TCNB
Relative intensity (a.u.)

Pyrene/OFN
0.5%
1%
3%
Pyrene/TCNB

400 450 500 550 600 650


Wavelength (nm)
(c) (d)
hv 365 nm Absorption
Pyrene/TCNB
Emission
Relative intensity (a.u.)

Pyrene/OFN

403 nm WLE

TCNB OFN Pyrene 250 300 350 400 450 500 550 600
Wavelength (nm)

Figure 10.10  (a) Fluorescence microscopy images and (b) PL spectra of pyrene/OFN micro crystals with various TCNB doping ratios (insets: the corresponding
photographs excited by 365 nm light). (c) Schematic illustration of the proposed energy-­transfer mechanism. (d) The absorption spectrum (black line) of pyrene/
TCNB and emission spectrum (red line) of pyrene/OFN. Source: Reprinted with permission from Ref. [3]. Copyright 2017 Wiley-­VCH.
10.6 ­HAITE Through Cocrystal Strateg 287

could occur from pyrene/OFN moiety to pyrene/TCNB phase. Specifically, when a pyrene/OFN
donor pair is excited at 365 nm light, the resultant excitation energy will be transferred to a nearby
pyrene/TCNB (acceptor pair) within 2–6 nm [2, 3]. Importantly, the two competitive intermolecu-
lar interactions, CT and AP interactions, enable the doping process because of the larger associa-
tive ability of CT interaction. The competitive interaction strategy could be utilized to fabricate
many organic optoelectronic materials.

10.6.1.2  Enhanced Emission and Tuned Color


Except for tunable emission colors, many cocrystal formations could also be accompanied by a
significant increase in the photoluminescence quantum yield (PLQY) compared with each single
component [22–27, 70, 74–76, 78]. Several mechanisms are proposed for the increased PLQY as
well as the corresponding varied fluorescence lifetime [22–27], which will be detailed below.
The NP/BTFBA cocrystal, prepared from 4-­(1-­naphthylvinyl)pyridine and
4-­bromotetrafluorobenzoic acid, presents the PLQY value of 50.38%, which was greatly enhanced
compared with that of pristine NP (11.95%) as well as the solution of mixed NP/BTFBA
(Figure 10.6a) [22]. Such high HAITE performance can be attributed to the strong hydrogen-­bond
interactions between two adjacent molecules (Figure 10.6b), which may effectively inhibit nonra-
diative decay pathways, configuration relaxation, and thermal vibration  [22, 23]. The lifetime of
pure NP is calculated to be 1.57 ns, while that of NP/DITFB (0.26 ns) can be declined for ~6 times
relative to pure NP, which should be attributed to the effect of heavy atoms in DITFB that slightly
reduce the radiative transition process. For NP/BTFBA and NP/BBA, the lifetimes are enhanced to
be 8.85 and 4.35 ns, respectively, indicating a delayed deactivation process in the excitation state [79].
Another proposed reason is the isolation effect of the acceptors, which can effectively restrict the
formation of aggregations (such as excimers) between the donor chromophores [24, 27, 70]. For
instance, Wang et al. [24] reported three CT complexes based on 1-­acetyl-­3-­(4-­methoxyphenyl)-­5-­(
9-­anthryl)-­2-­pyrazoline (AMAP)/1-­acetyl-­3-­naphthyl-­(9-­anthryl)-­2-­pyrazoline(ANNP)/1-­acetyl-­3-­
thiophene-­5-­(9-­anthryl)-­2-­pyrazoline (ATAP) as donors (D) and TCNB as an acceptor (A)
(Figure 10.11a). Remarkably, the solid-­state PLQYs of these three CT cocrystals have been improved
to 41.38% for ANNP/TCNB, 38.85% for AMAP/TCNB, and 34.06% for ATAP/TCNB compared with
the pristine donors (3.5, 4.0, and 2.8% for AMAP, ANNP, and ATAP, respectively) (Figure 10.11a).
As revealed by the crystal structural analysis, each acceptor molecule is sandwiched face-­to-­face
between two anthracene moieties via hydrogen bond, CT process, and π⋯π interactions, leading to
CT cocrystals with typical mixed stacking modes of D⋯A⋯D⋯A⋯D (AMAP/TCNB and ATAP/
TCNB) and DAD⋯DAD (ANNP/TCNB) arrangement (Figure 10.11b). The mixed stacking feature
between π-­stacked donors and acceptor should be responsible for the greatly enhanced PLQY due
to the facts that it can effectively restrict the formation of aggregations (such as excimers) between
donor chromophores. Furthermore, the fluorescence lifetime of CT complexes has been improved
by ~100 times compared with the pristine donors, which should be attributed to the more stable CT
state generated by TCNB in comparison to the S0 − S1 [70].
There is another example for the suppression of concentration quenching [27], in which micro-
crystals were prepared from mixed carbon-­bridged oligo-­para-­phenylenevinylenes (COPVs)
(Figure 10.12a). Upon excitation with 380 nm light, the pure microcrystals of COPV2 and COPV3
barely demonstrate the characteristic emission band from 0 to 0 vibronic-­level because of the self-­
absorption (Figure 10.12b, c). By comparison, the COPV3-­doped microcrystals with fCOPV3 (mole
fraction) from 0.005 to 0.13 almost show identical PL spectra as that of a COPV3 solution, which
contains 0−0 (476  nm), 0−1 (506  nm), and 0−2 (545  nm) bands (Figure  10.12d). The above-­
mentioned results indicate that COPV3 should molecularly disperse in the solid “solvent” of
Figure 10.11  (a) Chemical structures of acceptor and donors, and the as-­formed three cocrystal structures as well as the corresponding fluorescence
microscopy images (λex = 365 nm), the PLQY are indicated. (b) The crystal stacking modes of three cocrystals. In ANNP/TCNB, the molecules in cyan and
lavender are designated to the two kinds of chiral molecules of donor. In AMAP/TCNB,the molecules in gray and pink color present two kinds of chiral AMAP. In
ATAP/TCNB, the structures in pink and orange colors present two kinds of ATAP with different chirality. Source: Reprinted (adapted) with permission from
Ref. [24]. Copyright 2017 American Chemical Society.
10.6 ­HAITE Through Cocrystal Strateg 289

Figure 10.12  (a) Molecular structures of COPV2 and COPV3, and the fluorescent micrographs of the
microcrystals of COPV2 and COPV3 as well as their cocrystals with different fCOPV3 (0.019, 0.063, 0.105,
0.130, as indicated in each image). λex = 350−390 nm, scale bars: 10 μm. (b, c) Absorption (dashed) and PL
(solid) spectra of COPV2 (a) and COPV3 (b) in CHCl3. Concentration: 20 mM. (d) PL spectra of thin films of
COPV2 (blue) and COPV3 (green) as well as cocrystals with fCOPV3 = 0.005−0.13 (black), and that of COPV3
solution in CHCl3. λex = 380 nm. Source: Reprinted (adapted) with permission from Ref. [27]. Copyright 2018
American Chemical Society.

COPV2, the PLQYs of the cocrystals (0.65−0.76) are much higher than those of the pure COPV2
(0.45) and COPV3 (0.40) due to the suppression of concentration quenching.
Actually, ET from a host molecule (donor) to a guest molecule (acceptor) is also an effective
strategy to increase solid-­state fluorescent PLQY of the guest, which could enlarge the Stokes shift
to eliminate self-­absorption [25, 26]. You et al. [25] reported such kinds of work based on NCSQ-­g
with solid-­state fluorescence and CCSQ-­1 that only exhibits fluorescence in solution but no fluo-
rescence in solid state (Figure 10.13a, b). The large spectral overlap between absorption of CCSQ-­1
and emission of NCSQ-­g as well as their structural similarity make them excellent ET pairs
(Figure 10.13c, inset). Therefore, a series of CCSQ-­1-­doped NCSQ-­g ET complexes were success-
fully prepared. When increasing the doping dose of CCSQ-­1, the peak centered at 548 nm charac-
teristic for NCSQ-­g decreases gradually and blue-­shifts slightly, the featured emission of CCSQ-­1
emerges and then increases gradually (Figure 10.13c), which means that the energy (365 nm light)
is absorbed by NCSQ-­g transfers to CCSQ-­1 and then emits fluorescence. The maximum emission
of CCSQ-­1 appeared when CCSQ-­1/NCSQ-­g ratio is 0.01 mol%. Surprisingly, a quite small doping
amount (0.05 mol%) of CCSQ-­1 can lead to a ultra-­high ET efficiency of 96% (see calculation of
Equation 10.1), in which the emission intensity of the ET complex excited by 365 nm light can be
increased by nearly threefold compared with that of CCSQ-­1 excited by 560 nm light (Figure 10.13e,
inset, fluorescence microscope image under 365 nm).
The ET efficiency (EET) was calculated via the following Equation 10.1 [80].

I DA
EET 1 (10.1)
ID

where IDA and ID are the emission intensities of donor in the presence and absence of acceptor,
respectively.
290 10  Hetero-­aggregation-­induced Tunable Emission (HAITE) Through Cocrystal Strategy

Figure 10.13  (a) Molecular structures of CCSQ-­1 and NCSQ-­g. (b) Photos of CCSQ-­1 and NCSQ-­g in THF
(1.0 × 10−5 M) as well as crystals of NCSQ-­g under 365 nm light. (c) Fluorescence spectra of CCSQ-­1/NCSQ-­g
cocrystals with different ratios (0−1.0 mol%), Inset, spectral overlap of absorption of CCSQ-­1 in THF (blue)
and emission of NCSQ-­g in solid (red). (d) Maximum emission value of CCSQ-­1 (blue) and NCSQ-­g (red)
versus the CCSQ-­1/NCSQ-­g ratios. Inset, plot of the ET efficiency against the doping ratio. (e) Fluorescence
spectra of CCSQ-­1/NCSQ-­g crystals with 0.05 mol%, excited by 365 (black) and 560 nm (red). Inset, the
corresponding photo of CCSQ-­1/NCSQ-­g powder under 365 nm light. (f) Lifetime decay profiles of CCSQ-­1/
NCSQ-­g complexes under different doping ratios. (g) A linear Stern–Volmer fit of (I0/I − 1) against the mole
ratio of CCSQ-­1/NCSQ-­g at 0.0001 – 0.1 mol%. Source: Reprinted (adapted) with permission from Ref. [25].
Copyright 2012 American Chemical Society.

Energy-­transfer rate constant (kET) can be calculated through Equation 10.2 [81].

1 1
kET (10.2)
DA D

where τDA and τD are the lifetimes of the donor doped and undoped by the acceptor, respectively.
The fluorescence decay profile of NCSQ-­g presents two-­exponential decays (τ1, τ2). Therefore, the
average lifetime (< τ >) is needed here, which can be determined by below Equation 10.3 [80].
2 2
1 1 2 2
(10.3)
1 1 2 2

where α1 and α2 are the amplitudes of components τ1 and τ2, respectively.


As showed in the fluorescence lifetime decay profiles measured at the emission of NCSQ-­g, the
lifetime decreased to 0.66 ns for the CCSQ-­1-­doped NCSQ-­g with 0.05 mol% doping ratio com-
pared with that of pure NCSQ-­g (3.72 ns) (Figure 10.13f), which further confirms the occurrence
of efficient ET process. The decreased lifetime of CCSQ-­1-­doped NCSQ-­g also can be related to the
absence of excimers in the cocrystals [73, 82]. In the pure NCSQ-­g crystal, the NCSQ-­g molecules
will present face-­to-­face packing and easily form an excimer, which is usually formed by the com-
bination of a molecule in the ground state with the same molecule in the excited state. The radia-
tive transition process from an excimer to a bimolecule in ground state is usually a symmetry
resistance process, leading to a lower radiative decay rate and a larger lifetime. Conversely, in the
cocrystals, the doped CCSQ-­1 molecules will certainly prevent the close packing of NCSQ-­g mol-
ecules to a certain extent, which may hinder the formation of NCSQ-­g excimers and result in a
10.6 ­HAITE Through Cocrystal Strateg 291

smaller lifetime. The calculated Stern−Volmer quenching constant (KSV) is 65 800, indicating a
highly efficient ET pair formed by CCSQ-­1/NCSQ-­g (see calculation of Equation 10.4 and Equation
10.5) (Figure 10.13g) [83, 84]. Specifically, a doping ratio of 0.01 mol% can result in ET efficiency
of 85%, which means ~8500 NCSQ-­g molecules should be quenched by one CCSQ-­1 molecule [85].
In this case, the distance between two adjacent CCSQ-­1 molecules can be calculated to be 17 nm
(see calculation of Equation 10.4). Such long distance indicates that both exciton migration and
FRET should be involved in this ET process [86].
The distance between two CCSQ-­1 can be calculated according to Equation 10.4 [86].

d nVD ,
3
(10.4)

where VD is the volume of one donor (NCSQ-­g) molecule calculated from single-­crystal data and n
is designated to the molar ratio of donor/acceptor.
The Stern–Volmer quenching constant (KSV) can be calculated via the Stern–Volmer
equation [84].

I0
1 KSV A , (10.5)
I

where I and I0 are fluorescence intensities of CCSQ-­1-­doped NCSQ-­g and undoped NCSQ-­g,
respectively. [A] represents the concentration of acceptor. KSV can be determined by the slope of
the linear fit from the Stern–Volmer plot.
The same method was also reported to develop luminescent solid-­state perylene successfully,
which may be promising because pure perylene has a significantly high PLQY (~0.94) in dilute
solution, but it can self-­quench in high concentration or solid  [26]. A similar molecule,
1,3,6,8-­tetramethylpyrene (TMPY), was utilized as a host material due to its much larger energy
gap and higher LUMO level compared with that of perylene, as well as the well-­matched spectra
and molecular structures. After doping 5 mol% of TMPY, a high PLQY near 80% was observed in
the ET complex.

10.6.2  HAITE with Increased PLQY but Intrinsic Color


Although various achievements of organic chromophores with tunable emission colors via cocrys-
tal strategy is of utmost importance for the development of organic optics and photonics, these
kinds of materials actually present different optical properties from those of each coformer, which
undoubtedly cannot fulfill the applications that require the intrinsic properties of donors.
Therefore, keeping the intrinsic emission color of donors when improving their emission effi-
ciency through cocrystal strategies is highly desirable for optoelectrics. Such strategy remains a
great challenge and hasn’t appeared until the reports by Zhang et al. [5, 20] recently.
As we all know, most of the π-­conjugated organic chromophores suffer from the serious ACQ
effect due to the well-­known inevitable loss induced by the long-­lived and nonemissive triplet
states [20]. There are mainly two pathways to generate triplets: one is the ISC process from singlet
and another is spin-­allowed, singlet-­fission process requiring a timescale of ~100  fs, which is
related to the chromophore interactions [13]. Several efforts have been made to impede the ISC
process and the formation of triplets, e.g. using the intermittent excitation strategies [12, 18] or
utilizing scavengers that may partially quench the triplets to blend with chromophores [14, 87].
Nevertheless, it is still difficult to fully eliminate the emission loss induced by triplets.
292 10  Hetero-­aggregation-­induced Tunable Emission (HAITE) Through Cocrystal Strategy

To solve such problem, Zhang et  al.  [5, 20] report a new and facile cocrystal method where
molecular barriers with high energy gap are intercalated in the periodic crystal matrices of π-­
conjugated chromophores (Figure 10.14a) to effectively stop the interaction of two neighboring
chromophores (low energy gap) stacked via face-­to-­face and then block the singlet-­to-­triplet pro-
cess, leading to the absence of triplet states and near-­unity PLQY of the organic chromophores.
Specifically, two active optical materials coronene (Cor) and anthracene (Ant) with optical energy
edges of c. 2.76 and 3.10 eV, respectively [20], were used as the chromophores. OFN with higher
energy gap of c. 3.78 eV is chosen as a molecular barrier to prevent intermolecular interaction and
electron exchange between two adjacent chromophores, which will contribute to the ISC [88–90].
The molecular structures of the Ant/OFN and Cor/OFN cocrystals all demonstrate a mixed-­stack
mode (1  :  1), in which the OFN molecular barriers are periodically intercalated in the packing
arrays (Figure 10.14b, c). The distance between two neighboring chromophore planes are enlarged
to 6.86 and 6.88 Å for Ant/OFN and Cor/OFN, respectively, which are almost double of the com-
mon π–π distance.
In addition, the OFN barriers lead to a blue-­shifted absorption of the chromophores compared
with the pristine crystals (Figure 10.14d, e). The PL blue-­shifts (Figure 10.14f, g) come from the
screening of the π–π interactions between two chromophores by OFN barriers, which may decrease
the PL red-­shift induced by exciton delocalization [91]. Such screening effect was assessed through
one-­electron coupling, which indicates weaker interactions between the chromophores [92]. The
calculation suggests that photoexcitation could induce low-­frequency molecular distortion and
then trigger the exciton interactions in Cor/OFN and Ant/OFN (Figure 10.15a) [20], which was
confirmed by the calculated natural transition orbital analysis and transition energies
(Figure  10.15b). The interaction thus results in a bright trap (exciplex) state (SS1) with a lower
energy than that of S1 in isolated chromophores (Figure 10.15c). More importantly, thus-­prepared
Ant–OFN and Cor–OFN show much higher PLQYs than those of the pristine crystals (Ant,81 ± 5%;
Cor, 5 ± 5%) or even those of chromophores in solution (Ant in CHCl3,27 ± 5%; Cor in CHCl3,
3 ± 5%), which are 100 ± 5% and 79 ± 5%, respectively.
Transient absorption spectroscopy was measured to prove the absence of triplet formation in the
cocrystals (Figure 10.15d–f). Evidently, the triplet formation of Ant/OFN cocrystal is suppressed,
while the triplet-­fission-­induced, pump-­induced absorption (PIA) of the Ant crystal centered at c.
440 nm is observed (Figure 10.15d–f), which can be rationalized that the OFN barriers impede the
singlet fission (SF) when the chromophore pairs are screened. Similarly, the triplet character of
Cor also disappears, but the singlet PIA remains in Cor/OFN cocrystals. In addition, the ISC in two
cocrystal systems is also eliminated, which still exists in both pristine chromophore solutions and
crystals, and gradually populates the triplets (Figure 10.15g–i). Based on above analysis, it is rea-
sonable to believe that the introduction of OFN molecular barriers can effectively prevent the
singlet–triplet quenching, leading to almost 100% PLQY in cocrystals.
Afterwards, the emission of more π-­conjugated chromophores (pyrene [Pyr], perylene [Per], and
anthracene [Ant]) with ACQ effect have also been enhanced through above-­mentioned cocrystal
method using OFN as a molecular barrier, which further verifies the universality of such method [5].
Differently, nano-­/microscale cocrystals have been prepared by using P123 as a surfactant to control
the dimensions of the cocrystals. As expected, after stacking with electron-­deficient OFN barrier,
which has a higher band gap of ~3.78 eV than each chromophore, the photophysical properties of
the resulted micrococrystals have been changed, e.g. blue-­shifted absorption and emission of Per/
OFN and Cor/OFN (Figure 10.16) as well as greatly enhanced emission.
Notably, the solid PLQYs of the thus-­obtained four micro-­/nano-­cocrystals (Per/OFN MS, Cor/
OFN MW, Per/OFN NP and Cor/OFN NR) are enhanced by 254, 235, 474, and 582%, respectively,
(a) (b)
EM EM

b
MB MB MB
a
S1 S1 Mean distance between Ant/OFN 6.86 Å
=3.43 Å
3.78 eV
T T (c)
<3.1 eV

MB EM MB EM MB a

c
MB: OFN; EM: Ant or Cor Mean distance between Cor/OFN
6.88 Å
=3.44 Å

(d) Ant/OFN
(e) (f) Ant/OFN (g)
Cor/OFN Cor/OFN
Ant Ant
Cor
PL Intensity (a.u.)

PL Intensity (a.u.)

Cor
Absorption (a.u.)

Absorption (a.u.)

∼100% ∼81%
PLQE PLQE ∼100% ∼5%
PLQE PLQE

300 400 500 600 300 400 500 600 400 500 600 700 400 500 600 700
Wavelength (nm) Wavelength (nm) Wavelength (nm) Wavelength (nm)

Figure 10.14  (a) Schematic diagram of the electronic energies in the periodic packing structure of the cocrystals. The blue arrows refer to the transition of
singlet emission. The orange arrows refer to singlet trapping via singlet–singlet annihilation (SSA). (a, b) Crystal structures of Ant/OFN (b) and Cor/OFN (c)
cocrystals. They all demonstrate mixed stack (1 : 1) mode. (d, e) Absorptions of Ant, Ant/OFN, Cor, and Cor/OFN in the crystal state as well as their (f, g) PL spectra
upon 325 nm continuous-­wave photoexcitation. Source: Reprinted with permission from Ref. [20]. Copyright 2018 Wiley-­VCH.
(a) (d) ΔT/T (10–3) (f) Ant/OFN (GA) ~59 ps Ant/OFN RISC induced DFL
Ant/OFN 45
Ant-OFN-Ant trimer 28 Ant (GA) ~790 ps Ant for Ant
1000 SSA induced DFL Ant in CHCI3(GA) Inf s Ant in CHCI3
8
Time (ps)

100 –11 0 TA at 7 ns
Normalized ΔT/T

Singlet PIA –30


Singlet PIA for
10 –50 Ant-OFN
–69 Singlet PIA
1 –89
Displ. = 0 Å Displ. = 4 Å (e) 420 500 600 700 780
Ant 2
–4
(b) SS1“state” in Ant-OFN-Ant trimer 1000 DFL RISC induced DFL
–10 0
Time (ps)

–16
100 Singlet PIA –23
–29 Singlet PIA for
Triplet PIA for Ant-OFN and Ant in CHCl3
10 Triplet PIA –35
Ant and Ant in CHCl3
–41
1 –48
420 500 600 700 780 420 430 440 450 650 700 750 780
Wavelength (nm) Wavelength (nm) Wavelength (nm)

(g) 40 Kinctic at 425 nm (h) Kinctic at 427 nm


Ant (i) 0
Ant in CHCL3
Kinctic at 580 nm Kinctic at 439 nm
HOMO LUMO 0 Kinctic at 580 nm
20
(c) SSA induced DFL
–40
S1 0
ΔT/T (10–3)

ΔT/T (10–3)
ΔT/T (10–3)

2 –20
(local) –20 –80 Singlet PIA
~0.4 eV
SS1 Singlet PIA
–40
(exciplex) Triplet PIA –120 Triplet PIA
–60 Singlet PIA –40 via ISC
1 Kinctic at 603 nm
3 –80 Ant/OFN Triplet PIA
–60 Kinctic at 740 nm
Kinctic at 431 nm

–100 –60
S0 0.4 1 10 100 1000 1000000 1 10 100 1000 7700 1 10 100 1000 7700

Time (ps) Time (ps) Time (ps)

Figure 10.15  (a) Molecular distortion model for the planar Ant–OFN–Ant trimer to study the potential role of low-­frequency intermolecular distortions upon
photoexcitation. (b) Molecular orbital diagrams for the SS1 “state” in Ant–OFN–Ant trimer under a displacement of 1 Å, indicating strong interaction between
OFN and Ant LUMO orbitals. (c) Energy diagram for the SS1 (exciplex) formation including photo absorption (1), intermolecular distortion (2) and fluorescent
emission (3). (d, e) 3D transient absorption (TA) spectra of Ant and Ant/OFN. (f) TA spectra (right) and global analysis-­extracted spectra signature (left) of Ant, Ant/
OFN and Ant in CHCl3. (g–i) The kinetics of Ant, Ant/OFN crystal, and Ant in CHCl3 at specific wavelengths. Source: Reprinted with permission from Ref. [20].
Copyright 2018 Wiley-­VCH.
10.6 ­HAITE Through Cocrystal Strateg 295

(a) 1.2 (c)


Per MS Per MS
1.0 OFN crystal OFN crystal
Absorbance (a.u.)

Per/OFN MW

PL intensity (a.u.)
Per/OFN MW
0.8

0.6

0.4
29 nm
0.2
15 nm
0.0
300 400 500 600 300 400 500 600 700 800 900
Wavelength (nm) Wavelength (nm)

(b) 1.2 (d)

1.0
Cor MW
Cor MW
Absorbance (a.u.)

PL intensity (a.u.)
0.8 Cor/OFN MW
Cor/OFN MW

0.6

0.4
16 nm
0.2 9 nm

0.0
300 400 500 600 400 500 600 700
Wavelength (nm) Wavelength (nm)

Figure 10.16  (a) Absorption and (c) PL spectra of Per MS, OFN and Per/OFN MW at excitation of 460 nm.
(b) Absorption and (d) PL spectra of Cor/OFN MW and Cor MW at excitation of 360 nm. MS and MW refer to
microsheet and microwire, respectively. Source: Reprinted with permission from Ref. [5]. Copyright 2019
Springer Nature.

in comparison with those of the corresponding raw chromophores as received, where MS, microw-
ire (MW), nanoparticle (NP), and nanorod (NR) refer to microsheet, microwire, nanoparticle and
nanorod, respectively. As detailed above, these results are undoubtedly attributed to the OFN
molecular barriers that periodically intercalate between two chromophore planes and then hinder
their intermolecular interaction and electron exchange, which eventually leads to the absence of
triplet states. In addition, the larger PLQY enhancements of nanoscale cocrystals compared with
that of microscale ones should be due to the surfactant of P123, which could effectively block the
π–π interaction or aggregation between polycyclic aromatic hydrocarbons (PAH) molecules as well
as the small crystals, resulting in the decreased dimensions accompanied with weakening the ACQ
effect to a certain extent. It is worth noting that the introduction of OFN barriers cannot cause the
color change of the chromophores, instead it just results in a color-­diluted but brighter lumines-
cence compared with that of the corresponding pristine powders (Figure  10.17c). More impor-
tantly, the highly improved PLQY as well as the superb water dispersibility endow the nanococrystals
with good biocompatibility, cellular permeability, and considerable cell-­imaging performance
(Figure 10.17d–g).
(a) F F (b) F F
F F F F

F F F F
F F F F

(c) Per MS Per/OFN MW OFN crystal Cor MW Cor/OFN MW

365 nm on

(d) Per NS (e) Per/OFN NP (f) Cor NR (g) Cor/OFN NR

Figure 10.17  Cocrystallization schemes of (a) Per/OFN and (b) Cor/OFN. (c) The photographs of the five powders (Per/OFN MW, Per MS, OFN crystals, Cor MW,
Cor/OFN MW) as well as the corresponding patterned “Nanyang Technological University (NTU)” and aqueous dispersions under 365 nm. (d–f) Cell-­imaging
results. (d) Confocal images of MCF-­7 cells labeled by Per NS, (e) Per/OFN NP, (f) Cor NR, and (g) Cor/OFN NR. λex = 405 nm. Scale bars in (d–g) are 75 μm. MS, MW,
NS, NP, and NR refer to microsheet, microwire, nanosheet, nanoparticle, and nanorod, respectively. Source: Reprinted with permission from Ref. [5]. Copyright 2019
Springer Nature.
10.6 ­HAITE Through Cocrystal Strateg 297

These researches will provide constructive guidelines for fabricating highly emissive solid-­state
materials by using just traditional ACQ molecules as well as OFN as raw materials, which would
greatly expand the families of the solid fluorescent materials. More importantly, they may create
more opportunities for the applications of originally nonfluorescent or poor fluorescent chromo-
phores in optical devices or bioimaging.

10.6.3  HAITE: Thermally Activated Delayed Fluorescence


Harvesting nonemissive excitons of organic semiconducting materials is extremely essential for
developing effective devices, which can be achieved based on reversed intersystem crossing (RISC)
between different excited states [93]. As reported, the CT nature of excited states can lead to spatial
separation of orbitals, and then result in a small singlet–triplet energy gap (DEST, ~0.1 eV), which
undoubtedly enhances the singlet–triplet spin coupling [94–97]. Given that CT can easily occur in
cocrystals [98] formed from two or more different components via noncovalent bonds, cocrystal
strategy would be a potential way to harvest nonemissive excitons.
Hu et al. [93] first reported the thermally activated delayed fluorescence (TADF) in an organic
cocrystal system, where a 1  :  2 CT cocrystal (TSB/TCNB, yellow color) was based on trans-­1,2-­
diphenylethylene (TSB, a triplet-­generating material) and TCNB (Figure  10.18b, c). Various

TSB/TCNB
(a) (b)
IC TSB
Sn ISC/RISC
VR LE (T2)
3

CT (S1)
1 3
LE (T1)
ISC/RISC
Absorption

Prompt fluo.

Delayed fluo. 2
Delayed fluo. 1

NC CN
Phos.

TCNB

NC CN
S0
(d) 1.0 TSB/TCNB UV
(c) TSB/TCNB
Prompt PL
Delayed PL
0.8
Normalized intensity

5907 cm–1
0.6

116 nm
0.4

0.2
100 μm
0.0
300 400 500 600 700 800
Wavelength (nm)

Figure 10.18  (a) Jablonski diagram of TSB/TCNB cocrystal, all possible photophysical processes are
indicated. At RT, RISC from intermediate triplets T1 and T2 with similar energies allows delayed fluorescence
to be the dominant decay channel. (b) Molecular structures of TSB and TCNB. (c) Optical image of TSB/TCNB
cocrystal. (d) Absorption and PL spectra of TSB/TCNB at RT, including prompt and delayed (delay for 50 ms)
components. Source: Reprinted with permission from Ref. [93]. Copyright 2019 Wiley-­VCH.
298 10  Hetero-­aggregation-­induced Tunable Emission (HAITE) Through Cocrystal Strategy

characterizations (e.g. single-­crystal XRD, absorption spectroscopy, variable-­temperature, time-­


resolved photoluminescence, steady-­state spectroscopy, and calculations based on first-­principles)
confirm that strong intermolecular CT interactions in the rigid crystalline TSB/TCNB narrow the
singlet–triplet energy gap and then facilitate RISC process to generate TADF. Specifically, the red-­
shifted broad absorption at 432 nm of TSB/TCNB compared with that of the single-­component
TSB crystal indicates a typical CT peak (Figure 10.18d). The CT nature of the excited states can be
further revealed by the large Stokes shift (5907  cm−1) determined by above absorption and the
broad emission peak at 580  nm (Figure  10.18d)  [99]. Moreover, TADF should be the dominant
decay channel at room temperature (RT), because the delayed component of TSB/TCNB is almost
the same as that of the prompt one (Figure 10.18d).
PL spectra at varying time delays under RT were measured to understand the excited states in the
delayed and prompt processes. The monoexponential decay of prompt component shows a nano-
second lifetime (T) (31.5  ns), indicating the prompt fluorescence from only one excited state
(Figure 10.19a). The delayed component with a millisecond lifetime (T1 = 0.69 ms, T2 = 2.94 ms)
presents double-­exponential decays, revealing a fluorescence from two excited states (Figure 10.19b).
Furthermore, the intensity of the delayed PL spectrum at 580 nm decreases monotonically with

(a) (b)
297 K 297 K
Fitted Fitted
104

104
Intensity (a.u)

Intensity (a.u)

102
Milli-second-scale

101 T1 = 0.69 ms
T = 31.5 ns
10 3
Nano-second-scale T2 = 2.94 ms
100
0 40 80 120 160 200 0.0 3.0 6.0 9.0 12 .0
(c) Decay time (ns) (d) Decay time (ns)

4×105 297 K 297 K


Steady 7×104 Delayed
277 K 277 K
state 257 K 6×104 50 μs 257 K
3×105 237 K 237 K
217 K 5×104 217 K
Intensity (a.u)

Intensity (a.u)

197 K
197 K
4×104 177 K
177 K
2×105 157 K
157 K 3×104 137 K
137 K 117 K
117 K 2×104
1×105
1×104

0
0
450 500 550 600 650 700 750 800 450 500 550 600 650 700 750 800
Wavelength (nm) Wavelength (nm)

Figure 10.19  Measured photoluminescence spectra of TSB/TCNB cocrystal. (a) Nano-­and (b) millisecond-­
scale decay curves of the intensity of time-­resolved emission spectra at 580 nm under RT. Fitting results
give single-­and double-­exponential forms for (a) and (b), respectively. The lifetimes are indicated. Inset:
snapshot of real-­time decay observation. (c) Steady-­state and (d) delayed (delay for 50 ms) PL spectra from
117 to 297 K. Source: Reprinted with permission from Ref. [93]. Copyright 2019 Wiley-­VCH.
10.6 ­HAITE Through Cocrystal Strateg 299

decreasing temperature (Figure 10.19d), which is a typical character of TADF emitters [97]. The


same trend was found in steady-­state PL results (Figure  10.19c), further confirming that the
delayed component is the dominant radiative decay process at room temperature. All the photo-
physical processes of TSB/TCNB were confirmed by theoretical calculations.
The possible mechanism of TADF for TSB/TCNB was proposed. The strong intermolecular
interactions in the rigid crystalline TSB/TCNB shows compact stacking orientation different from
those of single-­component crystals (Figure 10.20a, b), which can greatly reduce possible energy
loss from nonradiative relaxations, avoid invasion of atmospheric O2, and therefore stabilize the
triplet excitons [100]. In addition, the π–π interactions were enhanced and strong hydrogen bonds
were formed after introducing TCNB acceptor in 1 : 2 ratio, which facilitates the intermolecular CT
interactions and thus results in a small DEST and then enhanced RISC. Contribution of the strong
intermolecular interactions in TSB/TCNB to the small DEST was further confirmed by below
experiments (Figure 10.20c). The emission spectra of TSB/TCNB solution show several peaks in
the region of 300–420 nm that are almost a simple addition of emission of solid TCNB and TSB,
which is distinctively different from those of the TSB/TCNB cocrystal (c. 580 nm). Moreover, the

(a) (b)

c
b
o

TSB/TCNB, side view TSB/TCNB, top view

(c) 348 297 K TSB/TCNB solution


267 K
237 K
332 207 K
177 K
147 K
Intensity (a.u.)

365 117 K
87 K
77 K

321 380 480 560 640 720


497 540
460

320 400 480 560 640


Wavelength (nm)

Figure 10.20  (a, b) Crystal structure of TSB/TCNB (TSB in blue color, TCNB in green color), the main
average distances are indicated, units, Å. (c) PL spectra of TSB/TCNB solution from 297 to 77 K. Inset:
Enlarged view for the region of 400–800 nm. Source: Reprinted with permission from Ref. [93]. Copyright
2019 Wiley-­VCH.
300 10  Hetero-­aggregation-­induced Tunable Emission (HAITE) Through Cocrystal Strategy

blue emission from the TSB/TCNB solution demonstrated much faster decay than that of the
cocrystal at 77and 297 K, referring to the typical feature of fluorescence. In sharp contrast, a series
of new peaks in the region of 500–650 nm appeared when cooling the solution to low temperatures
(below 207 K [Figure 10.20c]).
This research opens a new strategy (cocrystal) toward the exciton-­utilization efficiency as well as
the development of novel TADF materials. Because such kind of research is the only one case of
organic TADF cocrystals, there should be plenty of rooms left to further develop for organic TADF
cocrystals.

10.6.4  HAITE-­phosphorescence
Generally, most of the π-­conjugated chromophores can emit strong fluorescence in solution, which
would be greatly quenched in solid or aggregated state due to the exciplexes or excimers forma-
tion [6, 101]. On the other hand, it is usually difficult for the common π-­conjugated chromophores
to present phosphorescence because of the forbidden spin of singlet–triplet transition. Therefore,
developing a strategy toward singlet–triplet conversion and then solid organic phosphorescent
materials is an extremely challenging task [102]. The initial method is to encapsulate the chromo-
phores through constrained systems (e.g. cyclodextrins, cavitands, polymer matrices) that could
block diffusion-­controlled quenching process  [103], which often need heavy-­atom coguests.
Crystal engineering has been reported to achieve organic phosphorescent materials because the
densely rigid environment in the crystalline structure is beneficial to the intermolecular electronic
coupling, which could stabilize the triplet states to promote persistent room temperature phospho-
rescence (RTP). This phenomenon is known as the crystallization-­induced RTP effect  [104]. In
addition, carbonyl groups and halogen atoms have been proven to be essential substituents in
organic phosphorescent materials, which can enhance spin−orbit coupling [105]. Based on above
considerations, utilizing organic molecules containing I or Br atoms as halogen bond donors to
form cocrystals with appropriate chromophores would be a potential strategy to fabricate metal-­
free phosphorescent materials [23, 102, 105–107]. Recently, several organic phosphorescent cocrys-
tals based on halogen bonds have been reported [23, 102, 105–107], where the external heavy-­atom
effect from halogen bonds facilitates the spin–orbital coupling and intersystem crossing, therefore
enabling efficient RT phosphorescence in cocrystals.
For instance, Jin et  al.  [102, 107] reported a series of phosphorescent cocrystals based on
1,4-­diiodotetrafluorobenzene (DITFB, halogen bonding donor) and PAH, such as
3-­ring-­N-­heterocyclic hydrocarbons (3-­R-­NHHs), naphthalene (Nap) and phenanthrene (Phe). As
expected, five cocrystals demonstrate orange–yellow, green, orange, green, and orange phospho-
rescence. The phosphorescence lifetimes of these cocrystals are in millisecond scale. Taking Nap/
DITFB and Phe/DITFB as representatives, their molecular structures, phosphorescence spectra as
well as the photos under 365  nm are illustrated in Figure  10.21a–d. By comparison, each pure
PAHs cannot demonstrate solid-­state phosphorescence, further confirming the effective external
heavy-­atom effect from iodine atoms in DITFB. Ono et al. [23] obtained a tricomponent cocrystal
based on EBPDI, TPFB and iodobenzene (ETIC) assembled from 5,5′-­(ethyne-­1,2-­diyl)-­bis(2-­pyrid
in-­3-­yl-­isoindoline-­1,3-­dione) (EBPDI), tris(pentafluorophenyl)borane (TPFB), and iodobenzene
(Figure 10.21e), which exhibited the dual emission of RT fluorescence and phosphorescence. The
phosphorescence peak locates at a larger wavelength (640 nm) and present a much longer lifetime
(tav, 499  ms) compared with that of fluorescence. As expected, the phosphorescence emissions
(500–700 nm) disappeared when iodobenzene was removed (EBPDI/TPFB) (Figure 10.21f, g). In
this case, the fluorescence was regenerated due to the suppression of ISC, leading to a high PLQY
O
(a) F F (b) F F (e) N F
F
FF
F
F
N
O F F
I I I I O F
B
F
N F F
I
N
F F F O F F
Nap/DITFB Phe/DITFB TPFB TPFB F
iodobenzene

(c) 323 481 514


ETIC
FL PH (g) FL PH
(f) 7000

Normalized Intensity (a.u.)


Nap/DITFB 1.0
6000 TPFB/TPFB TPFB/TPFB
Nap 0.8
552 5000 ETIC ETIC
Intensity

4000 0.6
3000 0.4
2000
585 0.2
1000
0.0
Intensity

0
400 450 500 550 600 650 700 400 450 500 550 600 650 700
(d) Wavelength (nm) Wavelength (nm)
360 568
(h) TPFB/TPFB (i) ETIC Heavy atom
Sn Sn
Phe/DITFB effect
S1 ISC (S1→Tn) S1 ISC (S1→Tn)
615 675 Tn Tn
Phe T1
Ex. Ex. T1
NR FL NR FL
737 NR PH NR PH

S0 S0
200 300 400 500 600 700 800
Wavelength (nm)

Figure 10.21  (a, b) Molecular structures of Nap/DITFB (a) and Phe/DITFB (b). (c, d) Phosphorescent excitation (blue) and emission (green, Nap/DITFB; orange,
Phe/DITFB) spectra of cocrystals. The vibration bands were separated via assuming a Gaussian distribution of each band. Insets: the corresponding photos of the
patterned “Beijing Normal University (BNU)” under 365 nm light. (e) Molecular structures of ETIC. (f) Emission spectra of ETIC (red) and that of removal
iodobenzene (TPFB/TPFB, blue), and (g) the corresponding normalized emission spectra. Excitation at 340 nm. Jablonski diagrams in (h) iodobenzene-­free crystal
(TPFB/TPFB) and (i) ETIC. FL and phosphorescence (PH) refer to fluorescence and phosphorescence emissions, respectively, observed at room temperature in air.
Source: Reprinted (adapted) with permission from Refs. [23, 102, 107]. Copyright 2018 Wiley-­VCH, 2014 and 2012, the Royal Society of Chemistry.
302 10  Hetero-­aggregation-­induced Tunable Emission (HAITE) Through Cocrystal Strategy

of 27% (Figure 10.21f). These phosphorescent characters indicate the effective heavy-­atom effect of
iodobenzene in the supramolecular cocrystal. In the presence of iodobenzene, triplet states can
generate from singlet excited states via ISC, resulting in both fluorescence and RTP in air
(Figure 10.21i). Removal of iodobenzene results in regeneration of strong fluorescence from sin-
glet excited states (Figure 10.21h).
Moreover, 2,7-­di-­(N,N-­diphenylamino)-­9,9-­dimethyl-­9H-­fluorene (DDF) doped
2,8-­bis(diphenylph­osphoryl)dibenzo[b,d] thiophene (PPT) with good single crystalline structure
was prepared very recently, which presented visible, long, persistent green luminescence (LPL,
containing fluorescence and RTP) with a duration of more than six seconds at RT (Figure 10.22a,
b) [108]. A different mechanism was proposed, in which long-­lived charge-­separated (CS) states
should be generated due to the strong electronic diffusibility and relatively stable anion radical of
PPT. Specifically, when a trace amount of DDF was doped into the crystalline structure of PPT, the
CS state DDF•+–PPT•− could generate by photo-­induced electron/hole transfer within the D/A
system. Transient fluorescence spectra were measured to confirm this CS process. The emission at
405  nm from DDF blocks gives a biexponential profile. The faster component (0.15  ns, 86.9%)
refers to photo-­induced electron transfer from excited DDF to PPT blocks to generate CS states,
and the slower component (3.9 ns, 13.1%) is designated to the singlet–triplet conversion. At RT, the
PPT : DDF cocrystals present effective LPL emission after removing the excitation of 365 nm and
the emission peak was almost overlapped with the RTP emission from DDF blocks (Figure 10.22c),
indicating that the RTP of the DDF triplet state mainly contributes to this LPL emission, contain-
ing a small amount of delayed fluorescence from the CS state. The possible mechanism was pro-
posed. The conversion from the CS state to the DDF triplet state was allowed because of the higher
energy of the CS state (2.5  eV) than that of DDF triplet state (2.4  eV). Compared with the ISC
process from the singlet to triplet of DDF, the long-­lived CS state could greatly delay the conversion
process, leading to the LPL emission composed of the delayed fluorescence and persistent RTP
(Figure 10.22d).
When the RTP-­ and LPL-­based luminescent materials are still in the embryonic stage and the
luminescence mechanisms are still unclear, more efforts are required in this field.

10.7  ­Summary and Outlook

Cocrystal strategies really provide an ideal way to design and achieve various organic solid-­state
fluorescent materials with desirable emission. The optical properties of a large number of π-­
conjugated molecules have been tuned effectively through such convenient method by choosing
suitable coformers or preparing methods. These properties include emission color, fluorescent effi-
ciency and lifetime, as well as the luminescent types. Notably, more than four emission colors
(even white) can be easily achieved from only one two-­component or three-­component system just
via slightly varying the doping amount of the coformers in cocrystals. Moreover, the PLQY of sev-
eral poor luminescent chromophores have been greatly enhanced, even near-­unity PLQY of
organic chromophores was achieved. In addition, new luminescent species, possessing the delayed
fluorescence and phosphorescence that are absent for the single-­component chromophores, can be
introduced in the pristine molecules. Therefore, cocrystal strategies with low cost and convenient
operation have been considered as an emerging functional material-­design strategy, which may
provide many more opportunities to achieve more novel photophysical phenomena as well as
more solid luminescent materials. Moreover, the HAITE properties make the cocrystalized lumi-
nescent materials potential candidates for aggregated-­/solid-­state optoelectronic applications,
(a) (b)

PPT:DDF UV On UV Off
DDF

1 mol% N N

S After one second Six seconds

O O 50 μm
(d) 4.0
P P
1PPT*:DDF
PPT:DDF
PPT (3.7 eV)
3.5

(c) PPT:1DDF*
PPT:DDF 1.0
0.5 (3.2 eV)
3.0
Energieas (eV)

ET
3PPT*:DDF
ISC
Emission intensity (a.u.)

0.4 0.8
Absorbance (a.u.)

(2.8 eV)
+
DDF* -(PPT)n-PPT
0.3 0.6 2.5 PPT:3DDF*
(2.5 eV)
hv

FL

(2.4 eV)
0.2 0.4
LPL

2.0
LPL
PL

0.1 0.2
0.5
0.0 0.0
300 400 500 600 0.0
Wavelenth (nm) PPT:DDF

Figure 10.22  (a) Chemical structures of PPT and DDF as well as optical microscope image of the PPT : DDF cocrystal under day light irradiation, (b) Photographs
of DDF : PPT film at RT under no irradiation, 365 nm excitation and after turning off 365 nm light for one and six seconds. (c) Absorption (black), phosphorescence
(blue), fluorescence (red), photoluminescence (dark-­yellow), and LPL (mauve) emission spectra of PPT : DDF crystalline film (excitation, 365 nm). All the spectra
were measured at RT. (d) Energy diagram and possible photophysical processes of PPT : DDF cocrystal when excited by 365 nm. The (PPT)n (n ≥ 0) represents
several PPT molecules. Source: Reprinted with permission from Ref. [108]. Copyright 2019 Wiley-­VCH.
304 10  Hetero-­aggregation-­induced Tunable Emission (HAITE) Through Cocrystal Strategy

including the advanced organic optics and photonics, solid luminescent sensors, informational
displays, organic lasers, glow-­in-­the-­dark paintings, fluorescent antiforgery devices, and bioimag-
ing. However, longstanding problems still exist in this field: (i) Effective tuning of optical proper-
ties via controlling molecular orientation, stacking mode, and definite stoichiometric ratio remains
at an early stage; (ii) the research on the mechanism of tunable luminescence is still in the primary
stage, and further investigation of the relationships between intermolecular interactions and lumi-
nescent properties are highly desirable in order to achieve cocrystal emitters with desirable lumi-
nescent properties; (iii) it is essential to find more novel cocrystals beyond the old D–A systems,
which will undoubtedly introduce new luminescent properties to cocrystal field; (iv) the investiga-
tions in enhancing PLQY as well as developing delayed fluorescence and phosphorescence of
cocrystals are still in their infancy, which offer opportunity for potential research and need to be
greatly explored; (v) as a newly emerging field, great effort to realize practical applications of
organic cocrystals should be made. It is undeniable that the field of HAITE through cocrystal strat-
egy is booming and will demonstrate an excellent prospect.

­References

1 Zhang, G. Y., Zhao, J. B., Chow, P. C. Y. et al. (2018). Nonfullerene acceptor molecules for bulk
heterojunction organic solar cells. Chemical Reviews 118 (7): 3447–3507.
2 Lei, Y. L., Jin, Y., Zhou, D. Y. et al. (2012). White-­light emitting microtubes of mixed organic
charge-­transfer complexes. Advanced Materials 24 (39): 5345–5351.
3 Sun, Y. Q., Lei, Y. L., Liao, L. S. et al. (2017). Competition between arene-­perfluoroarene and
charge-­transfer interactions in organic light-­harvesting systems. Angewandte Chemie International
Edition 56 (35): 10352–10356.
4 Ding, D., Li, K., Liu, B. et al. (2013). Bioprobes based on AIE fluorogens. Accounts of Chemical
Research 46 (11): 2441–2453.
5 Huang, Y. J., Xing, J., Gong, Q. Y. et al. (2019). Reducing aggregation caused quenching effect
through co-­assembly of PAH chromophores and molecular barriers. Nature Communcations
10: 1–9.
6 Ostroverkhova, O. (2016). Organic optoelectronic materials: mechanisms and applications.
Chemical Reviews 116 (22): 13279–13412.
7 Grimsdale, A. C., Chan, K. L., Martin, R. E. et al. (2009). Synthesis of light-­emitting conjugated
polymers for applications in electroluminescent devices. Chemical Reviews 109 (3): 897–1091.
8 Mei, J., Leung, N. L. C., Kwok, R. T. K. et al. (2015). Aggregation-­induced emission: together we
shine, united we soar! Chemical Reviews 115 (21): 11718–11940.
9 Nielsen, C. B., Holliday, S., Chen, H. Y. et al. (2015). Non-­fullerene electron acceptors for use in
organic solar cells. Accounts of Chemical Research 48 (11): 2803–2812.
10 Borisov, S. M., Wolfbeis, O. S. (2008). Optical biosensors. Chemical Reviews 108 (2): 423–461.
11 Sosorev, A. Y., Parashchuk, O. D., Zapunidi, S. A. et al. (2013). Intrachain aggregation of charge-­
transfer complexes in conjugated polymer: acceptor blends from photoluminescence quenching.
The Journal of Physical Chemistry C 117 (14): 6972–6978.
12 Sandanayaka, A. S. D., Matsushima, T., Bencheikh, F. et al. (2017). Toward continuous-­wave
operation of organic semiconductor lasers. Science Advances 3 (4): e1602570.
13 Bakulin, A. A., Morgan, S. E., Kehoe, T. B. et al. (2016). Real-­time observation of multiexcitonic
states in ultrafast singlet fission using coherent 2D electronic spectroscopy. Nature Chemistry
8: 16–23.
  ­Reference 305

14 Kéna-­Cohen, S., Sivan, Y., Bradley, D. D. C. et al. (2011). Plasmonic sinks for the selective removal
of long-­lived states. ACS Nano 5 (12): 9958–9965.
15 Meher, N., Iyer, P. K. (2018). Spontaneously self-­assembled naphthalimide nanosheets:
aggregation-­induced emission and unveiling a-­PET for sensitive detection of organic volatile
contaminants in water. Angewandte Chemie International Edition 57 (28): 8488–8492.
16 An, B. K., Kwon, S. K., Jung, S. D. et al. (2002). Enhanced emission and its switching in fluorescent
organic nanoparticles. Journal of the American Chemical Society 124 (48): 14410–14415.
17 Zhang, W., Liu, W., Li, P. et al. (2015). Rapid-­response fluorescent probe for hydrogen peroxide in
living cells based on increased polarity of C-­B bonds. Analytical Chemistry 87 (19): 9825–9828.
18 Ju, H. J., Wang, K., Zhang, J. et al. (2017). 1,6-­and 2,7-­trans-­β-­styryl substituted pyrenes exhibiting
both emissive and semiconducting properties in the solid state. Chemistry of Materials 29 (8):
3580–3588.
19 Zhang, C. C., Chen, P. L., Hu, W. P. (2016). Organic light-­emitting transistors: materials, device
configurations, and operations. Small 12 (10): 1352–1294.
20 Ye, H. Q., Liu, G. F., Liu, S. et al. (2018). Molecular-­barrier-­enhanced aromatic fluorophores in
cocrystals with unity quantum efficiency. Angewandte Chemie International Edition 57 (7):
1928–1932.
21 Bu, F., Duan, R. H., Xie, Y. J. et al. (2015). Unusual aggregation-­induced emission of a coumarin
derivative as a result of the restriction of an intramolecular twisting motion. Angewandte Chemie
International Edition 54 (48): 14492–14497.
22 Li, S. Z., Yan, D. P. (2018). Two-­component aggregation-­induced emission: two-­component
aggregation-­induced emission materials: tunable one/two-­photon luminescence and stimuli-­
responsive switches by co-­crystal formation. Advanced Optical Materials 6 (19): 1870076.
23 Ono, T., Taema, A., Goto, A. et al. (2018). Switching of monomer fluorescence, charge-­transfer
fluorescence, and room-­temperature phosphorescence induced by aromatic guest inclusion in a
supramolecular host. Chemistry A European Journal 24 (66): 17487–17496.
24 Khan, A., Wang, M. L., Usman, R. et al. (2017). Molecular marriage via charge transfer interaction
in organic charge transfer co-­crystals toward solid-­state fluorescence modulation. Crystal Growth &
Design 17 (3): 1251–1257.
25 Yang, S. J., You, J. S., Lan, J. B. et al. (2012). Facile access to extremely efficient energy-­transfer
pairs via an unexpected reaction of squaraines with ketones. Journal of the American Chemical
Society 134 (29): 11868–11871.
26 Li, J. P., Takaishi, S., Fujinuma, N. et al. (2011). Enhancement of luminescence intensity in TMPY/
perylene co-­single crystals. Journal of Materials Chemistry 21: 17662–17666.
27 Okada, D., Azzini, S., Nishioka, H. et al. (2018). π-­Electronic co-­crystal microcavities with selective
vibronic-­mode light amplification: toward förster resonance energy transfer lasing. Nano Letters 18
(7): 4396–4402.
28 Tseng, K. P., Fang, F. C., Shyue, J. J. et al. (2011). Spontaneous generation of highly emissive RGB
organic nanospheres. Angewandte Chemie International Edition 50 (31): 7032–7036.
29 Morimoto, M., Irie, M. (2010). A diarylethene cocrystal that converts light into mechanical work.
Journal of the American Chemical Society 132 (40): 14172–14178.
30 Liao, Q., Fu, H. B., Wang, C. et al. (2011). Cooperative assembly of binary molecular components
into tubular structures for multiple photonic applications. Angewandte Chemie International
Edition 50 (21): 4942–4946.
31 Giansante, C., Raffy, G., Schäfer, C. et al. (2011). White-­light-­emitting self-­assembled nanofibers
and their evidence by microspectroscopy of individual objects. Journal of the American Chemical
Society 133 (2): 316–325.
306 10  Hetero-­aggregation-­induced Tunable Emission (HAITE) Through Cocrystal Strategy

32 Horiuchi, S., Ishii, F., Kumai, R. et al. (2005). Ferroelectricity near room temperature in co-­crystals
of nonpolar organic molecules. Nature Materials 4: 163–166.
33 Wöhler, F. (1844). Untersuchungen über das Chinon. Annalen 51 (2): 145–163.
34 Odom, S. A., Caruso, M. M., Finke, A. D. et al. (2010). Restoration of conductivity with TTF-­TCNQ
charge-­transfer salts. Advanced Functional Materials 20 (11): 1721–1727.
35 Ferraris, J., Walatka, V., Perlstei, J. H. et al. (1973). Electron transfer in a new highly conducting
donor–acceptor complex. Journal of the American Chemical Society 95 (3): 948–949.
36 Hasegawa, T., Mattenberger, K., Takeya, J. et al. (2004). Ambipolar field-­effect carrier injections in
organic Mott insulators. Physical Review B 69: 245115.
37 Xiao, J. H., Yin, Z. Y., Li, H. et al. (2010). Postchemistry of organic particles: when TTF
microparticles meet TCNQ microstructures in aqueous solution. Journal of the American Chemical
Society 132 (20): 6926–6928.
38 Liu, G. F., Liu, J., Ye, X. et al. (2017). Self-­healing behavior in a thermo-­mechanically responsive
cocrystal during a reversible phase transition. Angewandte Chemie International Edition 56 (1):
198–202.
39 Qin, Y. K., Zhang, J., Zheng, X. Y. et al. (2014). Charge-­transfer complex crystal based on extended-
­π-­conjugated acceptor and sulfur-­bridged annulene: charge-­transfer interaction and remarkable
high ambipolar transport characteristics. Advanced Materials 26 (24): 4093–4099.
40 Zhang, J., Tan, J. H., Ma, Z. Y. et al. (2013). Fullerene/sulfur-­bridged annulene cocrystals: two-­
dimensional segregated heterojunctions with ambipolar transport properties and
photoresponsivity. Journal of the American Chemical Society 135 (2): 558–561.
41 Braga, D., Maini, L., Grepioni, F. (2013). Mechanochemical preparation of co-­crystals. Chemical
Society Reviews 42: 7638–7648.
42 Wang, Y., Zhu, W. G., Du, W. N. et al. (2018). Cocrystals strategy towards materials for near-­
infrared photothermal conversion and imaging. Angewandte Chemie International Edition 57 (15):
3963–3967.
43 Wang, Y., Zhu, W.G., Dong, H.L., Zhang, X.T., Li, R.J., Hu, W. (2016). Organic cocrystals: new
strategy for molecular collaborative innovation. Topics in Current Chemistry (Z), 374 (83): 1–34.
44 Chi, X., Besnard, C., Thorsmølle, V. K. et al. (2004). Structure and transport properties of the
charge-­transfer salt coronene-­TCNQ. Chemistry of Materials 16 (26): 5751–5755.
45 Kistenmacher, T. J., Emge, T. J., Bloch, A. N. et al. (1982). Structure of the red, semiconducting
form of 4,4′,5,5′-­tetramethyl-­Δ2,2′-­bi-­1,3-­diselenole-­7,7,8,8-­tetracyano-­p-­quinodimethane,
TMTSF-­TCNQ. Acta Crystallographica Section B: Structural Science, Crystal Engineering and
Materials 38: 1193–1199.
46 Huang, Y.J., Wang, Z. R., Zhang, Q. C. et al. (2019). Organic cocrystals: beyond electrical
conductivities and field-­effect transistors (FETs). Angewandte Chemie International Edition 58 (29):
9696–9711.
47 Fourmigu, M. (2008). Halogen bonding in conducting or magnetic molecular materials. Structural
Bonding (Berlin) 126: 181–207.
48 Lin, H. Y., Chang, X. P., Yan, D. P. et al. (2017). Tuning excited-­state-­intramolecular-­proton-­transfer
(ESIPT) process and emission by cocrystal formation: a combined experimental and theoretical
study. Chemical Science 8: 2086–2090.
49 Zhang, J., Xu, W., Sheng, P. et al. (2017). Organic donor–acceptor complexes as novel organic
semiconductors. Accounts of Chemical Research 50 (7): 1654–1662.
50 Zhu, W. G., Zheng, R. H., Zhen, Y. G. et al. (2015). Rational design of charge-­transfer interactions
in halogen-­bonded co-­crystals toward versatile solid-­state optoelectronics. Journal of the American
Chemical Society 137 (34): 11038–11046.
  ­Reference 307

51 Cavallo, G., Metrangolo, P., Milani, R. et al. (2016). The halogen bond. Chemical Reviews 116 (4):
2478–2601.
52 Zhang, J., Gu, P. Y., Long, G. K. et al. (2016). Switching charge-­transfer characteristics from p-­type
to n-­type through molecular “doping” (co-­crystallization). Chemical Science 7: 3851–3856.
53 Lucassen, A. C. B., Karton, A., Leitus, G. et al. (2007). Co-­crystallization of sym-­triiodo-­
trifluorobenzene with bipyridyl donors: consistent formation of two instead of anticipated three
N⋯I halogen bonds. Crystal Growth & Design 7 (2): 386–392.
54 Black, H. T., Perepichka, D. F. (2014). Crystal engineering of dual channel p/n organic
semiconductors by complementary hydrogen bonding. Angewandte Chemie International Edition
53 (8): 2138–2142.
55 Wakahara, T., D’Angelo, P., Miyazawa, K. I. et al. (2012). Fullerene/cobalt porphyrin hybrid
nanosheets with ambipolar charge transporting characteristics. Journal of the American Chemical
Society 134 (17): 7204–7206.
56 Liu, C., Minari, T., Lu, X. B. et al. (2011). Solution-­processable organic single crystals with bandlike
transport in field-­effect transistors. Advanced Materials 23 (4): 523–526.
57 Zhu, W. G., Yi, Y. P., Hu, W. P. (2015). Precisely tailoring the stoichiometric stacking of perylene-­
TCNQ co-­crystals towards different nano and microstructures with varied optoelectronic
performances. Small 11 (18): 2150–2156.
58 Zhu, L. Y., Yi, Y. P., Li, Y. et al. (2012). Prediction of remarkable ambipolar charge-­transport
characteristics in organic mixed-­stack charge-­transfer crystals. Journal of the American Chemical
Society 134 (4): 2340–2347.
59 Yoshida, Y., Kumagai, Y., Mizuno, M. et al. (2015). Improved dynamic properties of charge-­
transfer-­type supramolecular rotor composed of coronene and F4TCNQ. Crystal Growth & Design
15 (11): 5513–5518.
60 Shirota, Y., Kageyama, H. (2007). Charge carrier transporting molecular materials and their
applications in devices. Chemical Reviews 107 (4): 953–1010.
61 Li, R. J., Hu, W. P., Zhu, D. B. et al. (2010). Micro-­and nanocrystals of organic semiconductors.
Accounts of Chemical Research 43 (4): 529–540.
62 Zeis, R., Siegrist, T., Kloc, Ch. (2005). Single-­crystal field-­effect transistors based on copper
phthalocyanine. Applied Physical Letters 86: 022103.
63 Menard, E., Marchenko, A., Podzorov, V. et al. (2006). Nanoscale surface morphology and
rectifying behavior of a bulk single-­crystal organic semiconductor. Advanced Materials 18 (12):
1552–1556.
64 Wuest, J. D. (2012). Co-­crystals give light a tune-­up. Nature Chemistry 4: 74–75.
65 Yan, D. P., Delori, A., Lloyd, G. O. et al. (2011). A cocrystal strategy to tune the luminescent
properties of stilbene-­type organic solid-­state Materials. Angewandte Chemie International Edition
50 (52): 12483–12486.
66 Fan, G. L., Yan, D. P. (2014). Positional isomers of cyanostilbene: two-­component molecular
assembly and multiple-­stimuli responsive luminescence. Scientific Reports 4: 4933.
67 Torrance, J. B. (1979). The difference between metallic and insulating salts of
tetracyanoquinodimethone (TCNQ): how to design an organic metal. Accounts of Chemical
Research 12 (3): 79–86.
68 Torrance, J. B., Vazquez, J. E., Mayerle, J. J. et al. (1981). Discovery of a neutral-­to-­ionic phase
transition in organic materials. Physical Review Letters 46: 253–257.
69 Shokaryev, I., Buurma, A. J. C., Jurchescu, O. D. et al. (2008). Electronic band structure of
tetracene-­TCNQ and perylene-­TCNQ compounds. The Journal of Physical Chemistry A 112 (11):
2497–2502.
308 10  Hetero-­aggregation-­induced Tunable Emission (HAITE) Through Cocrystal Strategy

70 Li, S. Z., Lin, Y. J., Yan, D. P. (2016). Two-­component molecular cocrystals of 9-­acetylanthracene
with highly tunable one-­/two-­photon fluorescence and aggregation induced emission. Journal of
Materials Chemistry C 4: 2527–2534.
71 Sun, H., Peng, J., Zhao, K. et al. (2017). Efficient luminescent microtubes of charge-­transfer
organic cocrystals involving 1,2,4,5-­tetracyanobenzene, carbazole derivatives, and pyrene
derivatives. Crystal Growth & Design 17 (12): 6684–6691.
72 Gujrati, M. D., Kumar, N. S. S., Brown, A. S. et al. (2011). Luminescent charge-­transfer complexes:
tuning emission in binary fluorophore mixtures. Langmuir 27 (11): 6554–6558.
73 Wu, J. J., Li, Z. Z., Zhuo, M. P. et al. (2018). Tunable emission color and morphology of organic
microcrystals by a “cocrystal” approach. Advanced Optical Materials 6 (9): 1701300.
74 Yan, D. P., Yang, H. J., Meng, Q. Y. (2014). Two-­component molecular materials of
2,5-­diphenyloxazole exhibiting tunable ultraviolet/blue polarized emission, pump-­enhanced
luminescence, and mechanochromic response. Advanced Functional Materials 24 (5): 587–594.
75 Usman, R., Khan, A., Wang, M. L. et al. (2018). Investigation of charge-­transfer interaction in
mixed stack donor-­acceptor cocrystals toward tunable solid-­state emission characteristics. Crystal
Growth & Design 18 (10): 6001–6008.
76 Fang, X. Y., Yang, X. G., Yan, D. P. (2017). Vapor-­phase π–π molecular recognition: a fast and
solvent-­free strategy towards the formation of co-­crystalline hollow microtube with 1D optical
waveguide and up-­conversion emission. Journal of Materials Chemistry C 5: 1632–1637.
77 Lei, Y. L., Sun, Y. Q., Zhang, Y. et al. (2018). Complex assembly from planar and twisted π-­
conjugated molecules towards alloy helices and core-­shell structures. Nature Communications
9: 4358.
78 Ye, X., Liu, Y., Guo, Q. et al. (2019). 1D versus 2D cocrystals growth via microspacing in-­air
sublimation. Nature Communications 10: 761.
79 Zhu, W. G., Zheng, R. H., Hu, W. P. et al. (2015). Revealing the charge-­transfer interactions in
self-­assembled organic cocrystals: two-­dimensional photonic applications. Angewandte Chemie
International Edition 54 (23): 6785–6789.
80 Lakowicz, J. R. Principles of Fluorescence Spectroscopy. 3rd ed.; Springer: New York, 2006.
81 Li, F., Lin, J. L., Feng, J. et al. (2003). Electrical and optical characteristics of red organic light-­
emitting diodes doped with two guest dyes. Synthetic Metals 139 (2): 341–346.
82 Son, M. J., Park, K. H., Shao, C. Z. et al. (2014). Spectroscopic demonstration of exciton dynamics
and excimer formation in a sterically controlled perylene bisimide dimer aggregate. The Journal
Physical Chemistry Letters 5 (20): 3601–3607.
83 Bhattacharyya, S., Paramanik, B., Patra, A. (2011). Energy transfer and confined motion of dyes
trapped in semiconducting conjugated polymer nanoparticles. The Journal Physical Chemistry C
115 (43): 20832–20839.
84 Wu, C. F., Peng, H. S., Jiang, Y. F. et al. (2006). Energy transfer mediated fluorescence from blended
conjugated polymer nanoparticles. The Journal Physical Chemistry B 110 (29): 14148–14154.
85 Guerzo, A. D., Olive, A. G. L., Reichwagen, J. et al. (2005). Energy transfer in self-­assembled
[n]-­acene fibers involving 100 donors per acceptor. Journal of the American Chemical Society 127
(51): 17984–17985.
86 Van Der Meer, B. W., Coker, G., Chen, S. Y. Resonance Energy Transfer; VCH: New York, 1994.
87 Schols, S., Kadashchuk, A., Heremans, P. et al. (2009). Triplet excitation scavenging in films of
conjugated polymers. Chemphyschem: A European Journal of Chemical Physics and Physical
Chemistry 10 (2): 1071–1076.
88 Freed, K. F. (1976). Theory of collision induced intersystem crossing. The Journal of Chemical
Physics 64: 1604–1611.
  ­Reference 309

89 Giacobbe, E. M., Mi, Q. X., Colvin, M. T. et al. (2009). Ultrafast intersystem crossing and spin
dynamics of photoexcited perylene-­3,4:9,10-­bis(dicarboximide) covalently linked to a nitroxide
radical at fixed distances. Journal of the American Chemical Society 131 (10): 3700–3712.
90 Green, S. A., Simpson, D. J., Zhou, G. et al. (1990). Intramolecular quenching of excited singlet
states by stable nitroxyl radicals. Journal of the American Chemical Society 112 (20): 7337–7346.
91 Ahn, T. S., Mgller, A. M., Al-­Kaysi, R. O. et al. (2008). Experimental and theoretical study of
temperature dependent exciton delocalization and relaxation in anthracene thin films. The
Journal of Chemical Physics 128: 054505.
92 Zhao, Y. Y., Luo, X., Li, H. et al. (2013). Interlayer breathing and shear modes in few-­trilayer MoS2
and WSe2. Nano Letters 13 (3): 1007–1015.
93 Sun, L. J., Hua, W. J., Liu, Y. et al. (2019). Thermally activated delayed fluorescence in an organic
cocrystal: narrowing the singlet–triplet energy gap via charge transfer. Angewandte Chemie
International Edition 58 (33): 11311–11316.
94 Zhang, S. T., Yao, L., Peng, Q. M. et al. (2015). Achieving a significantly increased efficiency in
nondoped pure blue fluorescent OLED: a quasi-­equivalent hybridized excited state. Advanced
Functional Materials 25 (11): 1755–1762.
95 Li, W. J., Liu, D. D., Shen, F. Z. et al. (2012). A twisting donor-­acceptor molecule with an
intercrossed excited state for highly efficient, deep-­blue electroluminescence. Advanced
Functional Materials 22 (13): 2797–2803.
96 Geng, Y., D’Aleo, A., Inada, K. et al. (2017). Donor-­σ-­acceptor motifs: thermally activated delayed
fluorescence emitters with dual upconversion. Angewandte Chemie International Edition 56 (52):
16536–16540.
97 Endo, A., Ogasawara, M., Takahashi, A. et al. (2009). Thermally activated delayed fluorescence
from Sn4+-­porphyrin complexes and their application to organic light emitting diodes-­a novel
mechanism for electroluminescence. Advanced Materials 21 (47): 4802–4806.
98 Sun, L. J., Zhu, W. G., Hu, W. P. (2017). Intermolecular charge-­transfer interactions facilitate
two-­photon absorption in styrylpyridine–tetracyanobenzene cocrystals. Angewandte Chemie
International Edition 56 (27): 7831–7835.
99 Park, S. K., Cho, I., Gierschner, J. et al. (2016). Stimuli-­responsive reversible fluorescence
switching in a crystalline donor–acceptor mixture film: mixed stack charge-­transfer emission
versus segregated stack monomer emission. Angewandte Chemie International Edition 55 (1):
203–207.
100 Bian, L. F., Shi, H. F., An, Z. F. et al. (2018). Simultaneously enhancing efficiency and lifetime of
ultralong organic phosphorescence materials by molecular self-­assembly. Journal of the American
Chemical Society 140 (34): 10734–10739.
101 Hofbeck, T., Monkowius, U., Yersin, H. (2015). Highly efficient luminescence of Cu(I)
compounds: thermally activated delayed fluorescence combined with short-­lived
phosphorescence. Journal of the American Chemical Society 137 (1): 399–404.
102 Wang, H., Hu, R. X., Pang, X. et al. (2014). The phosphorescent co-­crystals of
1,4-­diiodotetrafluorobenzene and bent 3-­ring-­N-­heterocyclic hydrocarbons by C-­I⋯N and C-­I⋯π
halogen bonds. CrystEngComm 16: 7942–7948.
103 Lee, D., Bolton, O., Kim, B. C. et al. (2013). Room temperature phosphorescence of metal-­free
organic materials in amorphous polymer matrices. Journal of the American Chemical Society 135
(16): 6325–6329.
104 Chen, G. L., Feng, H., Feng, F. F. et al. (2018). Photophysical tuning of organic ionic crystals from
ultralong afterglow to highly efficient phosphorescence by variation of halides. The Journal of
Physical Chemistry Letters 9 (21): 6305–6311.
310 10  Hetero-­aggregation-­induced Tunable Emission (HAITE) Through Cocrystal Strategy

105 Ventura, B., Bertocco, A., Braga, D. et al. (2014). Luminescence properties of 1,8-­naphthalimide
derivatives in solution, in their crystals, and in co-­crystals: toward room-­temperature
phosphorescence from organic materials. The Journal of Physical Chemistry C 118 (32):
18646–18658.
106 Li, L. L., Wang, H., Wang, W. Z. et al. (2017). Interactions between haloperfluorobenzenes and
fluoranthene in luminescent cocrystals from π-­hole⋯π to σ-­hole⋯π bonds. CrystEngComm 19:
5058–5067.
107 Shen, Q. J., Pang, X., Zhao, X. R. et al. (2012). Phosphorescent cocrystals constructed by
1,4-­diiodotetrafluorobenzene and polyaromatic hydrocarbons based on C-­I⋯π halogen bonding
and other assisting weak interactions. CrystEngComm 14: 5027–5034.
108 Han, J. L., Feng, W. H., Muleta, D. Y. et al. (2019). Small-­molecule-­doped organic crystals with
long-­persistent luminescence. Advanced Functional Materials 29 (30): 1902503.
311

11

Anti-­Kasha Emission from Organic Aggregates


Wenbin Huang and Zikai He
School of Science, Harbin Institute of Technology, Shenzhen, HIT Campus of University Town, Shenzhen, China

11.1 ­Introduction

Light plays an irreplaceable role in daily life and industrial production. The study of luminescence
has been in progress with the development of the society for tens of thousands of years from sun-
light to chemoluminescence, electroluminescence, bioluminescence, etc.  [1–3]. Many emerging
luminescent systems stimulate the study of working principles and thus further promote the
development and application of the area. Based on vast experimental evidences, some classical
theories have been summarized and extended to the explanation of novel luminescent systems. As
one of the famous theories, the Kasha’s rule has become an indispensable, analytical model to
explain luminescent phenomena of organic molecules. This rule, summarized and proposed by
Kasha in 1950s, describes that the photophysical and photochemical processes in condensed phase
have the significant yields only in the lowest excited state of a given multiplicity [4, 5]. In other
words, neither the photoemissions nor the photoreactions in the photoexcited process involves the
initial higher-­order excited states but the lowest excited states. It has been proved that most pho-
toemissions and photoreactions do follow this rule, which greatly facilitates people to understand
and develop corresponding applications [6, 7]. However, some exceptions are considered as anti-­
Kasha rule. It indicates that emissions, reactions, or other processes occur from upper rather than
the lowest singlet and triplet excited states, and thus are comparatively rare and extremely difficult
to observe, owing to significantly fast vibrational or collisional relaxation of upper excited states.
The anti-­Kasha research is mainly focused on the emissive properties of traditional systems such
as azulenes, polyenes, quinones, aromatic carbonyl compounds, and silyl ketones during the past
decades [8]. Due to the concentration-­quenching effect in the traditional systems, the emission
was weakened or it completely disappeared as the molecules form aggregates, which is summa-
rized as aggregation-­caused quenching (ACQ) effect [9] and which greatly hinders the practical
applications. Because ACQ considerably prevails in luminescence systems  [10–13], many anti-­
Kasha emissions only exist when the molecules are dispersed in solvents [14, 15], which require
additional processing before the practical applications. In 2001, Tang et al. reported another phe-
nomenon in contrast to the ACQ and first coined it as aggregation-­induced emission (AIE) [16].
Upon aggregation, molecular vibrations and rotations are intensively restricted resulting in the

Handbook of Aggregation-Induced Emission: Volume 1 Tutorial Lectures and Mechanism Studies, First Edition.
Edited by Youhong Tang and Ben Zhong Tang.
© 2022 John Wiley & Sons Ltd. Published 2022 by John Wiley & Sons Ltd.
312 11  Anti-­Kasha Emission from Organic Aggregates

weakening of the nonradiative transitions and the enhancement of the emissions [17–19]. AIE has
great value in biological imaging and medical treatment [13].
AIE would be an ideal strategy to develop the luminescent system with anti-­Kasha emission in
aggregate. The system simultaneously endows the AIE characteristics and anti-­Kasha emission.
The following basic criteria is required to achieve the anti-­Kasha emissions where the rate of inter-
nal conversion and vibrational relaxation from upper excited states to lowest excited state (Sn → S1
or Tn → T1) is quite small and the rate of radiative and other nonradiative depopulation of these
higher states, such as intersystem crossing, should be fast enough [20]. Although several molecules
with AIE and anti-­Kasha emission characteristics have been reported, the research of the anti-­
Kasha phenomenon in the aggregate state is still in the primary stage due to limited reported
examples and incomprehensive theoretical guidance.
In this chapter, we will review some selected examples of anti-­Kasha emission in aggregate.
According to their structure, we categorize them into three aspects and carefully scrutinize the
factors on the performance of anti-­Kasha emission. At the end of this chapter, we hope to establish
theoretical guidance and offer structure models to realize anti-­Kasha emission for future research-
ers. In addition, through the modulation of photoexcitation energy, anti-­Kasha emission has dem-
onstrated the advantages in frontier applications, which will also be described in the following
examples.

11.2  ­Anti-­Kasha Emission from Aromatic Carbonyl


Compounds in Aggregates

Aromatic carbonyl compounds have been widely concerned in recent years for developing anti-­
Kasha emission due to their n electrons and varied π-­conjugated system. In general, partially or
completely, charge-­transfer (CT) character can be obtained by constructing donor–acceptor (D–A)
structure with carbonyl groups as electron acceptors, which is of great role in adjusting the arrange-
ment of energy levels. In addition, the nonbond orbital and n electrons participate in the photoex-
citation process, which is constructive to the exciton spin-­flip during intersystem crossing (ISC)
process. In this section, we will focus on the anti-­Kasha properties of these aromatic carbonyl
compounds in aggregate state and discuss the influence of external and internal factors on their
luminous behaviors.
Arindam Banerjee’s group reported a naphthalene diamide polymer NDI 1 (Figures  11.1a
and 11.2) [21] and studied its luminescence performance. The results indicated that the mole-
cule exhibited weak fluorescence when dissolved in CHCl3 solution. With the addition of non-­
polar methylcyclohexane (MCH) solvent, the molecules gradually formed H-­aggregates [22–24],
and the obvious blue–white fluorescence was greatly enhanced, indicating the AIE characteris-
tics (Figure  11.2a). With the enhancing degree of the aggregation, the molecules in the
H-­aggregates reduce the intermolecular distance and form lower energy excimer-­like states
after photo excitation, which can result in long-­lived fluorescence with a distinct broad emis-
sion peak [25].
The time-­dependent density functional theory (TD-­DFT) calculations based on the H-­aggregate
dimer manifested that the total energy is the lowest when the intermolecular center distance is
4.06  Å  [26, 27]. Surprisingly, when calculating the oscillator strength of the transitions among
frontier orbitals, it is found that the oscillator strength of S1 → S0 transition is almost zero, indicat-
ing that the radiative decay of S1 is very weak. The fluorescence mainly comes from the bright S2
state, which is obviously the opposite of the Kasha’s rule (Figure 11.2b). The fact that anti-­Kasha
11.2  ­Anti-­Kasha Emission from Aromatic Carbonyl Compounds in Aggregates 313

(a) (b)
O O DMF-BP-PXZ: R =
H
N
N O
n
O O
O O
O
O N
nN
H R DPF-BP-PXZ: R =
O O
n = 10
N
NDI 1 O

(c) SBF-BP-PXZ: R =
O

(d)
Cl S
SDPE-OO: R1 = O, R2 = O
O R1 SDPE-OS: R1 = S, R2 = O
ClBDBT
R2 SDPE-SS: R1 = S, R2 = S
O

(e)
(f)

3
N
O N O
R
N CZNI-PLA: R =

N
O R O
O O
BCZ1: R = O
nH
O
O CZBP-PLA: R =
BCZ2: R =

O
O

BCZ3: R =

Br O
CZAQ-PLA: R =
O O
S O
BCZ4: R =
H

Figure 11.1  Structures of aromatic carbonyl compounds with anti-­Kasha emission in aggregates. Source:
(a) origined from Ref. [21], (b) origined from Ref. [28], (c) origined from Ref. [34], (d) origined from Ref. [47],
(e) origined from Ref. [54], and (f) origined from Ref. [57].

emission comes from the high lying excited state suggests that competing internal conversion of
S2 → S1 is such a slow process that the radiative transition of S2 → S0 dominates. The emission from
bright S2 rather than dark S1 is rare in aggregates. In fact, for a molecule, the calculated zero oscil-
lator strength of S1 cannot mean that the emission certainly coming from the S2 with large oscilla-
tor strength. There are possibly other complicated processes. Abundant direct evidences are needed
to fully validate the source of the anti-­Kasha emission. Nevertheless, this chapter provided us with
a feasible method. The formation of H-­aggregate to affect the luminescence process is a new tool
to obtain anti-­Kasha emission.
314 11  Anti-­Kasha Emission from Organic Aggregates

(a) (b)
CHCI3 : MCH Bright state
20 S2
10 : 90
IC
Dark state
15 S1
Intensity (a.u.)

100 : 0

10

Absorption

Emission
5

400 450 500 550 600 650 S0


Wavelength (nm)

Figure 11.2  (a) The fluorescence spectra of NDI 1 at different compositions of CHCl3 and
methylcyclohexane (MCH) at 2 mM concentration. (b) Relaxation mechanism for excited states of the
NDI 1 dimer (IC = Internal Conversion). Source: Adapted with permission from Ref. [21], Royal Society of
Chemistry, 2015.

Tang et al. designed and synthesized three D–A luminescent molecules with phenoxazine (PXZ)
as donor and aryl benzophenones as acceptors (Figures 11.1b and 11.3) [28]. Photoluminescence
spectroscopy suggested that these molecules displayed weak fluorescence in solution. The fluores-
cence intensity was enhanced and the lifetime was prolonged when water was added as a poor
solution to form aggregates indicating the AIE characteristics (Figure 11.3a). In particular, these
molecules exhibited stronger luminescence after doping into thin films, which were attributed to
aggregation-­induced delayed fluorescence (AIDF). Transient absorption spectra showed that the
triplet-­state excitons of the molecules in the solution did not participate in the process of reverse
ISC (rISC). Because the aggregation state inhibits the internal conversion (IC) and nonradiative
transition, the triplet-­state excitons can participate in the rISC, and finally the delayed fluorescence
in the aggregation state is enhancement.
According to DFT, TD-­DFT and the combined quantum mechanics and molecular mechanics
(QM/MM) method, S1 of DMF-­BP-­PXZ molecule in aggregate state is a dark state and therefore
fluorescence of DMF-­BP-­PXZ possibly comes from S2, which can be proved by calculating oscilla-
tor strength of the S2, which is in better agreement with the experimental data. This phenomenon
violates the Kasha’s rule. Calculations based on Fermi’s Golden Rule indicates that upper triplet
state T3 is involved in the process of ISC and rISC to upper singlet state S2, leading to the generation
of delayed fluorescence of S2 (Figure 11.3c). It is also found that spin-­orbit coupling (SOC) is larger
and recombination energy is smaller in solid state than that in solution state of DMF-­BP-­PXZ,
which benefits to improve the efficiency of delayed fluorescence. Further research shows that S2
exhibits typical CT characteristics, T3 is locally excited (LE) characteristic, and the energy differ-
ence of S2 − T3 is quite small [29]. Together with the results of other research groups [30–33], the
combined influence of these factors can effectively regulate the fluorescence performance, espe-
cially the improvement of quantum efficiency.
The crystal structure analysis displayed that these molecules present a twisted conformation in
the solid state to avoid the formation of π–π stacking, which could suppress concentration-­caused
emission quenching and triplet excitons annihilation (Figure 11.3b). In addition, a large number
11.2  ­Anti-­Kasha Emission from Aromatic Carbonyl Compounds in Aggregates 315

(a) (b)
fw (vol%)
99
90 d1
80 d2
70
PL intensity (a.u.)

60 d3
d5
50
40 d4
d3 d2
30
20
10 d3 d2
d5
0 d6

d1 = 2.834 Å, d2 = 2.733 Å, d3 = 2.621 Å,


450 500 550 600 650 d4 = 3.123 Å, d5 = 3.156 Å, d6 = 3.177 Å,
Wavelength (nm)
(c)
3.5 3.5 kISC
kISC
S2 S2
3.0 3.0
T4 T3
T3 T2
T2 2.5 kRISC
2.5
Excitation energy (eV)

S1 T1
S1 T1
2.0 2.0
Excitation energy (eV)

kF kIC kF kIC
kISC = 6.6 × 106 s–1
1.5 kISC(S2→T3) = 4.1 × 106 s–1 1.5
kRISC = 7.2 × 104 s–1
kF(S2→S0) = 1.2 × 106 s–1 kF (S2→S0) = 1.3 × 107 s–1
1.0 kIC(S2→S1→S0) 1.0
kIC(S2→S1→S0)
≈ kIC(S1→S0) = 3.1 × 1010 s–1 ≈ kIC(S1→S0) = 8.8 × 106 s–1
0.5 0.5

0.0 S0 0.0 S0

Solution Solid

Figure 11.3  (a) PL spectra of DMF-­BP-­PXZ in THF/water mixtures with different water fractions (fw),
measured under nitrogen. (b) Packing pattern of DMF-­BP-­PXZ in crystals. (c) Adiabatic excitation energies
for DMF-­BP-­PXZ in THF solution (left) and in solid (right). Source: Adapted with permission from Ref. [28],
Wiley-­VCH, 2019.

of weak hydrogen bonds have been observed, and resulting synergistic effect can simultaneously
rigidify the conformation of the molecules in favor of restricting intramolecular motions. In this
case, it is known that aggregation-­induced delayed anti-­Kasha fluorescence can be ascribed to the
suppression of IC channels and the promotion of ISC and rISC process in solid, owing to benefits
from restriction of intramolecular motion and rigidification of molecular structure. This D–A
molecular design provides us with a new strategy to construct structure model with high quantum
efficiency anti-­Kasha emission.
Tang et al. also designed and synthesized four dibenzothiophene derivatives with heavy atoms
and carbonyl groups, in which chlorine-­containing molecule crystals exhibit white phosphores-
cence in ambient conditions (Figures 11.1c and 11.4) [34]. Photoluminescence spectroscopy sug-
gested that these molecules displayed no or weak luminescence in solution. The luminescence
intensity was enhanced when single crystals form indicating the AIE characteristics, which was
316 11  Anti-­Kasha Emission from Organic Aggregates

(a) (b)
300 K Prompt Delayed
Temperature (K)
BDBT
FBDBT 50
CIBDBT 100
Normalized PL intensity

BrBDBT 150
200
250

PL intensity
275
300

400 500 600 700 400 500 600 700


Wavelength (nm) Wavelength (nm)
(c) (d)
Sn
At 467 nm temperature (K) S5
50 Tn
100 IC
100
150
200 S1
275 T2
300 ISC
10–1
Excitation
Energy

T1

10–2

Phosphorescence
10–3 Violation of Kasha’s rule
0 5 10 15 20
Time (ms) S0

Figure 11.4  (a) The prompt (solid line) and delayed (dash line, 10 ms) PL spectra of the crystalline
powders of BDBT, FBDBT, ClBDBT, and BrBDBT at 300 K. (b) PL spectra of ClBDBT crystals measured at
different temperatures from 50 to 300 K. (c) Time-­resolved PL decay curves of ClBDBT measured at 467 nm
from 50 to 300 K. (d) Relaxation mechanism for excited states of the crystalline powders of ClBDBT at
300 K. Source: Adapted with permission from Ref. [34], Springer Nature, 2017; and Ref. [41], American
Chemical Society, 2019.

ascribed to dual room-­temperature phosphorescence from T2 and T1 (Figure 11.4a) [35–37]. The


delayed spectrum indicated that the two emission peaks have different lifetimes, one in millisec-
onds and the other in seconds, without the generation of fluorescence. Further research suggested
that the lifetime of short-­wavelength emission T2 is affected by temperature, while the lifetime of
bathochromic emission T1 was not, and the proportion of short-­wavelength emission in total emis-
sion decreased as the temperature decreased, which manifested that T2 emission may be derived
from the thermal activation equilibrium of T1 (Figure 11.4b, c) [37–40]. According to analysis of
crystal structures, strong intermolecular π–π effect and intramolecular conjugate action in ClBDBT
11.2  ­Anti-­Kasha Emission from Aromatic Carbonyl Compounds in Aggregates 317

crystal resulted in hybridization of (n, π*) transition on the carbonyl group with (π, π*) transition
on the conjugate aromatic fraction to generate different energy levels and orbital configurations in
favor of thermal activation equilibrium of T1.
Although these experimental results indicate that T2 is generated by the thermal activation equi-
librium of T1, Swapan Chakrabarti et al. used TD-­DFT-­based response theory calculations to dis-
cover the existence of an alternative source of T2, which is the faster S1  →  T2 ISC process
(Figure 11.4d) [41]. The calculated fluorescence intensity of S1 → S0 was very weak, which was
consistent with the experimental results.
Since the large ET2 T1 may prevent T1 excitons from reaching T2 by Boltzmann population, the
generation of most T2 excitons should be realized by S1 → T2 ISC process. The ISC rate was calcu-
lated by using the cumulative expansion method by incorporating the Gaussian damping func-
tion [42, 43], and the ISC rate of S1 → T1 and S1 → T2 were calculated to be 7.71 × 108 and 2.13 × 1011,
respectively. Combined with an energy difference of S1 − T2 of 0.09 eV, these results further indi-
rectly confirmed that most T2 excitons are derived from ISC process. Absolute values of the
Duschinsky [44] rotation matrix for S1 → T1 and S1 → T2 transition indicated that the extent of state
mixing is quite strong for the former and even stronger for the latter because of apparent structural
changes of S1 and T2 resulting in larger vector displacements of the general modes of these
states [42]. By utilizing the high-­temperature limit approximation [45] to calculate radiative rate
and Marcus−Levich−Jortner theory [46] to calculate nonradiative rate, it was found that the non-
radiative intensity of T2 is much stronger than that of T1, and therefore the lifetime of T1 is longer
than T2, corroborating nicely with the experimental data. These evidences support that the ISC of
S1 compared with the thermal activation of T1 contributed more to the whole process. These results
indicate that the research of anti-­Kasha mechanism is of considerable complexity, and the overall
logic and details of the luminescent process can be fully explained by the high consistency between
experiment and theoretical results.
Wu’s group reported that three molecules with dual fluorescence from S2 and S1, dual phospho-
rescence from T2 and T1, and thermally activated delayed fluorescence (TADF) from S2 in solid
state (Figures 11.1d and 11.5), and finally observed single-­molecule white light at 77 K by decreas-
ing the temperature [47]. Under UV excitation, the single crystals of SDPE-­OO exhibited emission
of 391 and 502  nm, which depends on the excitation wavelength in ambient conditions
(Figure 11.5a). In addition, its emission spectra showed a mirror-­image relationship with the cor-
responding excitation spectra, indicating that the two emissions come from different electronic
energy levels, that is, S2 and S1. It was interesting that the energy gap between S2 and S1 have a good
logarithmic relationship with quantum yield of their high-­energy emission from S2 for this series
of compounds, which has been found for azulene derivatives in the gas phase [8]. This trend also
well conformed to the rule of energy gap [48]. The nonradiative efficiency from Sn to S1 decreases
with the increase of the separation energy of the two electronic states, and vice versa  [49, 50].
Delayed spectroscopy under low temperature suggested that the phosphorescence of 450 and
550  nm in SDPE-­OO crystal have been found, and it attribute to T2 and T1, respectively
(Figure 11.5b). In addition, there is a delayed fluorescence at 391 nm, and its proportion of total
luminescence decreases with the decrease of temperature, indicating that this emission serves as
TADF. Similar phenomena has been observed in crystals of other molecules.
Similarly, analysis of these crystals revealed the presence of abundant intramolecular and inter-
molecular hydrogen bonding in the crystals, which greatly inhibited the vibrations and rotations
of molecules in high-­energy states and thus effectively suppressed the nonradiative processes and
IC (Figure 11.5c). Further density functional calculation and quantum mechanical calculation [34,
51, 52] showed that the energy difference of S2 − S1 and T2 − T1 of these three crystals is large, and
(a) (b)
S1-S0 Excitation wavelength 1.0 Prompt
SDPE -OO 250 nm
OO Delay 0.5 ms
7000 270 nm
S2-S0 290 nm 0.5
310 nm Delay 2.0 ms
350 nm Normalized intensities
0 0.0
FL intensities

S2-S0 250 nm 1.0 Prompt


4000 SDPE -OS 270 nm OS Delay 0.5 ms
290 nm
310 nm 0.5 Delay 2.0 ms
S1-S0 350 nm
0 0.0
S2-S0 250 nm 1.0 Prompt
4000 SDPE -SS 255 nm SS
260 nm Delay 0.5 ms
S1-S0 0.5
265 nm Delay 2.0 ms
0 0.0
300 350 400 450 500 550 600 650 700 300 400 500 600 700
Wavelength (nm) Wavelength (nm)
(c) (d) ISC
Sm

Tn

rISC

S2
T2

S1
Anti-Kasha F

T1
Absorption

Anti-Kasha P
F

Intramolecular H-bonding: C-S...H: 2.45 A C-O...H: 2.61 A


Intermolecular H-bonding: C-S...H: 2.57 A 2.61 A
S0

Figure 11.5  (a) Excitation-­wavelength-­dependent fluorescence spectra of SDPE-­OO, SDPE-­OS, and SDPE-­SS crystalline powders at room
temperature under air. (b) Prompt and delayed emission spectra of SDPE-­OO, SDPE-­OS, and SDPE-­SS crystalline powders under air. (c) Top
view of the crystal structure with multiple intermolecular hydrogen bonding interactions in the unit cells of the SDPE-­SS single crystal
structure. (d) Suggested mechanism for the upper excited state fluorescence from neat solid of SDPE-­OO, SDPE-­OS, and SDPE-­SS based on
DFT calculations. Source: Adapted with permission from Ref. [47], Wiley-­VCH, 2020.
11.2  ­Anti-­Kasha Emission from Aromatic Carbonyl Compounds in Aggregates 319

thus reduces the rate of IC. In addition, it is observed that their different singlet and triplet elec-
tronic configurations, such as (π, π*) of S2 and (n, π*) and (π, π*) of Tn in SDPE-­OO crystal, are
beneficial to the ISC process owing to El-­Sayed rule [53]. Due to the small ES2 Tn contributing to
rISC from Tn to S2, synergistic effects resulted in excellent delayed fluorescence of SDPE-­OO crys-
tals (Figure 11.5d).
Hence, a simple single-­molecule system that shows multiple emissions by suppressing of Kasha
rule has been proven by the matched energy levels between S2 with its adjacent triplet excited
states. The intra-­and intermolecular interactions can suppress the IC, which plays the crucial role
in design of multiple-­emissive solid-­state organic materials with anti-­Kasha characteristics.
Zhang’s group designed and synthesized four dicarbazole derivatives (Figures 11.1e and 11.6) in
2019  [54]. Some molecules in solid state exhibited strong, dual room-­temperature phosphores-
cence properties, with the highest phosphorescence efficiency up to 64%. Photoluminescence spec-
troscopy suggested that no or weak luminescence of these molecules were observed in solution.
The PL intensity was greatly enhanced with the addition of water to form aggregation, indicating
the AIE characteristics. Interestingly, in the aggregative process of BCZ1, fluorescence can be con-
verted into phosphorescence due to the coupled vibration or exciton splitting. The initial band was
(n, π*) characteristic rather than CT state (Figure 11.6b) [55]. The steady-­state and delayed spec-
trum at low temperature indicated that the short-­wavelength and long-­wavelength emission of
BCZ1, BCZ2, BCZ3 were derived from anti-­Kasha phosphorescence of the upper state with (π, π*)
character and Kasha phosphorescence of the lowest triplet excited state T1 with (π, π*) character,
respectively (Figure  11.6a). However, the wavelength of their emission was shifted due to the

(b)
(a)
1.0 × 106
BCZ1 95
90
8.0 × 105 80
O BCZ1 70
50
Intensity

6.0 × 105 40
30
Normalized emission intensity

20
4.0 × 105 10
0
Normalized absorbance

O
BCZ2
2.0 × 105
O
0.0
400 450 500 550 600 650 700
O (c) Wavelength (nm)
BCZ3
S2
Br T4
S1
T2,T3
O O
S BCZ4 T1
3.94 eV

3.23 eV

3.34 eV

3.12 eV

2.11 eV

300 400 500 600 700 800


Wavelength (nm)
S0

Figure 11.6  (a) Absorption spectra (gray lines), steady-­state PL emission spectra in air at 298 K (red lines)
and at 77 K (blue lines), and delayed emission spectra at 77 K (green lines) of BCZ1, BCZ2, BCZ3, and BCZ4.
(b) Steady-­state photoluminescence (PL) spectra of BCZ1 in the course of AIE or AIEE process. (c) Vertical
excitation energy levels of BCZ1. Source: Adapted with permission from Ref. [54], Wiley-­VCH, 2019.
320 11  Anti-­Kasha Emission from Organic Aggregates

special molecular aggregation state or electronic energy level change for the BCZ derivatives with
different substituents. It was worth noting that different from BCZ1, BCZ2 and BCZ3, BCZ4 exhib-
ited distinct fluorescence and less-­obvious phosphorescence.
The calculations performed using a range-­separated hybrid function [56] were implemented to
elaborate the potential mechanism of totally different performance of BCZ1, BCZ2, BCZ3, and
BCZ4. The calculation results illustrated that there are three triplet excited states: T1 and the degen-
erate T2 and T3 which all below S1. The SOC process occurs favorably between S1 and T3 and S1 and
T1, respectively, because of (n, π*) character for S1 and (π, π*) character for both T1 and T3 according
to well-­known El-­Sayed rule [53]. Following the degeneration and fast IC between T3 and T2, the
relatively large energy gap between T1 and T2 gives rise to the competitive radiative process of T2
with IC from T2 to T1. Dual phosphorescence was thus observed in BCZ1. It is similar to that of
BCZ2 and BCZ3 (Figure 11.6c). As for BCZ4, (π, π*) character for S1, T1, and T2 cannot give rise to
large SOC between S1 and T2 and S1 and T1, respectively, which elucidates that BCZ4 mainly shows
prompt fluorescence. Based on these results, we know that the phenyl carbonyl structure plays a
decisive role in the construction of molecules with anti-­Kasha emission. The energy level and
quantum efficiency can obviously be potentially improved by the modification of the substituent.
Zhang’s group reported the design and synthesis of three kinds of organic polymers with D–A
structure by the carbazole substitution with PLA as donor and different types of carbonyl deriva-
tives as acceptor (Figures 11.1f and 11.7). Their luminescence properties have been systematically
studied and applied to cell imaging [57]. Photoluminescence and delayed spectra at room tempera-
ture or low temperature indicated that the thin film of CZNI-­PLA exhibited single fluorescence
with CT character at 512 nm and single phosphorescence with (π, π*) character at 525 nm at room
temperature (Figure 11.7c). Significantly distinct to CZNI-­PLA, only dual phosphorescence and no
fluorescence were observed both in thin film of CZBP-­PLA and CZAQ-­PLA. The emissions of
CZBP-­PLA locate at 445 nm with (n, π*) character from T2 and at 484 nm with (π, π*) character
from T1, respectively. For CZAQ-­PLA thin film, the photoluminescence spectra show peaks at
575 nm with (π, π*) character from T2 and at 586 nm from T1 (Figure 11.7d, e).
Calculation based on range-­separated hybrid function [56] found that ISC in CZNI-­PLA from S1
to T1 and S1 to T2 were forbidden processes because S1, T1, and T2 all are (π, π*) character and can-
not improve SOC according to El-­Sayed rule [53]. In addition to these, the large ΔEST also further
blocked this process, and finally all comprehensive effects would render CZNI-­PLA to exhibit
strong fluorescence and weak phosphorescence (Figure  11.7a). Compared with CZNI-­PLA, the
ISC between S1 and T1, S1, and T2 in CZBP-­PLA are favorably allowed by quantum mechanics
owing to S1 with (n, π*) character, T1 and T3 with (π, π*) character, and T2 with nearly equal propor-
tions of (n, π*) and (π, π*) character. Due to relatively large energy difference between T1 and T2,
anti-­Kasha phosphorescence of T2 can be observed because radiative process of T2 possibly com-
petes with the IC to T1 (Figure 11.7b). A similar understanding applies to CZAQ-­PLA, but the dif-
ference is that the lifetime of T2 is longer than that of T1 because T2 is dominated by (π, π*)
transitions and T1 is mainly mixed (π, π*) and (n, π*) configurations [58].
The analysis of orbital characteristics and theoretical calculations on these molecules conveyed
the critical information that changing substituent group on donor with different aromatic carbonyl
accepter can greatly modulate the luminescence property and adjust the transition characters of
the frontier orbitals to enhance ISC capacity, which is of significant value in molecular design to
achieve anti-­Kasha emission.
Here, we review several typical examples of aromatic carbonyl derivatives exhibiting efficiently
anti-­Kasha fluorescence or phosphorescence in aggregate state, which were explained universally
as large energy differences between upper excited state and lowest excited state to ensure that
Sn Sn
(a) (b)
CZBP-PLA
CZNI-PLA CZAQ-PLA

H.E
S1 L. T1 S1 . T2 T1
E.

EI-Sayed rules
Ex Ex
FL Weak Weak
RTP Phos. 2
Phos. 1

(c) (d) (e)


512 526 576 Air 298 K 445 484 Air 298 K Air 298 K
1.0 Vacuum 298 K 1.0 422 Vacuum 298 K 1.0 Vacuum 298 K
Normalized intensity

Normalized intensity

Normalized intensity

Steady at 77 K Steady at 77 K Steady at 77 K


Delayed at 77 K Delayed at 77 K Delayed at 77 K
0.8 0.8 0.8

0.6 0.6 0.6

0.4 0.4 0.4

0.2 0.2 0.2

0.0 0.0 0.0


400 450 500 550 600 650 700 750 350 400 450 500 550 600 650 450 500 550 600 650 700
Wavelength (nm) Wavelength (nm) Wavelength (nm)

Figure 11.7  The relaxation mechanism for excited states of thin films of (a) CZNI-­PLA and (b) CZBP-­PLA and CZAQ-­PLA. Steady-­
state emission and delayed emission spectra of (c) CZNI-­PLA, (d) CZBP-­PLA, and (e) CZAQ-­PLA thin films. Source: Adapted with
permission from Ref. [57], Royal Society of Chemistry, 2019.
322 11  Anti-­Kasha Emission from Organic Aggregates

competitive radiative rate is comparable with IC rate. However, the anti-­Kasha pathway becomes
more complex due to the involvement of SOC and restricted molecular motions. It is this complexity
that makes it difficult to identify anti-­Kasha features, especially in aggregate state. Therefore, the
combined explanations of experimental phenomena and theoretical calculations are necessary,
which will greatly help us to solve this dilemma. Through conjugation effect, coordination effect,
substitution effect, spatial effect, and so on, aromatic carbonyl derivatives will have great potential
for development of anti-­Kasha emissive systems.

11.3  ­Anti-­Kasha Emission from Azulene Compounds


in Aggregate

Azulene, a classical structure with permanent dipole moment consisting of an electron deficient,
five-­membered ring and an electron-­rich, seven-­membered ring exhibits anti-­Kasha emission from
upper excited state S2 or even higher excited states because of large energy difference between
upper excited states and lowest excited state [14, 15, 49, 59–62]. Most studies on the luminescence
properties of azulene were carried out in solution, but few ones in the aggregate state due to the
ACQ effect.
Zhu’s group demonstrated that azulene derivatives substituted by aldehyde groups can influence
anti-­Kasha emission by the effect of intermolecular hydrogen bonds (Figure 11.8a) [63]. TA with
three aldehyde groups dissolved in dimethylformamide (DMF) exhibited intense fluorescence at
416 nm from S3 and medium fluorescence at 465 nm from S2. As water was gradually added, the
fluorescence at 416 nm was quenched, while the fluorescence at 465 nm was red-­shifted to 495 nm.
The luminescent color of whole system changed from blue to green as intermolecular hydrogen
bonds were formed between TA and H2O (Figure 11.9a) and was confirmed by NMR, IR, DLS, and
TEM. For MA with one aldehyde group and DA with two aldehyde groups, no emission was
observed under the same conditions because these molecules form much weaker hydrogen bonds
with water than TA. It was found that the energy difference between S3 and S2 in TA is 0.35 eV,
which is so large that the rate of IC was reduced in favor of radiative process of S3 in solution state.

(a) (b)
TA: R1 = CHO, R2 = CHO, R3 = CHO O R
R1 R
DA: R1 = H, R2 = CHO, R3 = CHO O
R2
MA: R1 = H, R2 = H, R3 = CHO S S O
R3 R
S S
R
(c) (d) 1: R1 = H, R2 = H O S S
CN 2: R1 = Me, R2 = H
N 3: R1 = H, R2 = CN O
N F
B 4: R1 = H, R2 = NO2 R
F N 5: R1 = H, R2 = COOMe R O
N R1 R1 6: R1 = H, R2 = OMe
7: R1 = H, R2 = NH(CO)Me
8: R1 = H, R2 = NMe2 NC
R= O
R2
6
DPPZ BODIHY

Figure 11.8  Structures of azulene analogues and other unconventional aromatic compounds. Source: (a)
origined from Ref. [63], (b) origined from Ref. [64], (c) origined from Ref. [65], and (d) origined from Ref. [66].
11.3  ­Anti-­Kasha Emission from Azulene Compounds in Aggregat 323

(a)

OHC

CHO

S3-to-S0 OHC
H2O S2-to-S0

(b)
Non-radiative internal coversion
S4, S5 n–π* Fluorescence

S3 π–π* S5 π–π*

S3, S4 n–π*

S2 n–π*
S2 n–π*

S1 π–π* S1 π–π*

S0 S0

Figure 11.9  (a) A proposed process for the formation of intermolecular H-­bonding of compound TA
upon addition of a small amount of H2O, accompanied by an anti-­Kasha’s rule luminescent conversion.
(b) Proposed mechanism for the tuning of the TA luminescence in DMF/H2O mixtures. Source: Adapted with
permission from Ref. [63], American Chemical Society, 2018.

The theoretical calculations suggested that S4, S5, and S2 in TA are (n, π*) character, while S3 is
(π, π*) character. Therefore, the energy of S4 and S5 can be reduced until they are lower than S3, and
finally quenched S3 emission through increasing polarity via the addition of water. Meanwhile, the
increased polarity can also lead to the decrease of energy level of S2, which eventually gave rise to
the red-­shift emission of S2 (Figure 11.9b). In contrast, the H-­bonding is not sufficient for the shift-
ing of the dark (n, π*) states to be lower than the emissve S3 (π, π*) state in DA and MA.
Zhu’s group designed and reported a novel molecule, consisting of hexathiobenzene core as
energy donor and six cyanostyryl-­modified azulenes as energy accepter, with the characteristics of
fluorescence resonance energy transfer (FRET), anti-­Kasha emission, AIE and photoisomerization
(Figures 11.8b and 11.10) [64]. Under the UV excitation, the target molecule in the solution exhib-
ited the fluorescence at 360 nm from S3 and at 480 nm from S2 as same as that of the precursor
cyanostyryl-­modified azulene. The emission character of the precursor hexathiobenzene were not
observed, indicating that there was an energy-­transfer process (Figure 11.10a, d). The fact that the
mixed solutions of cyanostyryl-­modified azulene and hexathiobenzene core exhibited their respec-
tive emission and emission of target molecule was further enhanced confirmed the occurrence of
the energy-­transfer process (Figure  11.10b). The energy-­transfer process of FRET is facilitated
because there is a large overlap between the emission of hexathiobenzene and the absorption of
target molecule, which was verified by observing the enhanced emission of target molecule at
480  nm. Surprisingly, by adding water into the solution of the target molecule or the target
324 11  Anti-­Kasha Emission from Organic Aggregates

(a) (b)
Target molecule Target molecule

Cyanostyry-modified azulene Cyanostyryl-modified azulene


+Hexathiobenzene
Intensity

Intensity
400 500 600 700 400 450 500 550 600 650
Wavelength (nm) Wavelength (nm)

(c) (d)
0% Ex 310 nm
10%
20% S3
30%
40%
50% Ex 365 nm
60%
S2
Intensity

70%
80%
90%
95%
S1

400 450 500 550 600 S0


Wavelength (nm) 360 nm 480 nm 480 nm

Figure 11.10  Under 365 nm excitation (a) Fluorescence spectra of the target molecule and cyanostyryl-­
modified azulene and corresponding photographs. (b) Emission spectra of the target molecule and mixture
of cyanostyryl-­modified azulene and hexathiobenzene. (c) Emission spectra of the target molecule in THF/
H2O with different water fractions. (d) Proposed mechanism for the dual-­wavelength emission of the target
molecule. Source: Adapted with permission from Ref. [64], American Chemical Society, 2019.

molecule were doped in PMMA to prepare the solid film, the anti-­Kasha fluorescence intensity was
significantly enhanced, indicating the characteristics of aggregation-­induced anti-­Kasha emission
(Figure 11.10c). In this example, by integrating subunits with different functions, new properties
are obtained in addition to the inherited properties of the subunits, which is of great significance
to the design of functional molecules.
Normally, intrinsic anti-­Kasha emission of azulene in the aggregate state were difficult to be
observed due to ACQ effect. To overcome this defect, structure modifications were employed to
enhance the luminescence in aggregates, which is not only beneficial to improve quantum yield
but also endow new functionalities such as the response to environment modulations.

11.4  ­Anti-­Kasha Emission from Other Unconventional


Aromatic Compounds in Aggregates

In addition to the conventional structures mentioned above, some other molecules with special
structures can also exhibit anti-­Kasha emission. However, there is little progress in this field,
appealing for researchers to promote further development. A simple aza-­aromatic compound
11.4  ­Anti-­Kasha Emission from Other Unconventional Aromatic Compounds in Aggregates 325

dibenzo[a,c]phenazine (DPPZ) was synthesized by Yang’s group, and it was found that it exhibited
single-­molecule white-­light emission consisting of ternary bands of single, simultaneous fluores-
cence and dual RTP (Figures 11.8c and 11.11) [65]. Photoluminescence spectra showed four dis-
tinct peaks in DPPZ crystals (Figure  11.11a). After grinding the crystals, the oxygen-­quenching
effect resulted in the decrease of the intensity of two peaks in long-­wavelength range, indicating
that the two emission bands were triplet excited states related phosphorescence. Also, the emission
was dependent on the excitation wavelength, which was consistent with the characteristics of anti-­
Kasha emission. The phosphorescence was attributed to emission from Tn and emission from T1,
respectively.
Photoluminescence spectra at different temperatures indicated that phosphorescence from T1
was temperature-­dependent and was enhanced with the rise of temperature, while phosphores-
cence from T2 was not influenced indicating that independent of temperature, which ruled out the
possibility of that T2 excitons were formed through thermal activation from T1 (Figure 11.11b) [8,
34]. The calculations based on the NTOs of relevant excited states suggested that the ISC between S1
and T1 and S2 and T2 in DPPZ were favorably allowed by El-­Sayed rule [53] owing to S1 and T2 with

(a) (b)
Peak fit of DPPZ PL spectrum Original PL of DPPZ 80 K
R-square = 0.997 Peak #1 (S1 Emission)
100 K
Normalized PL intensity (a.u.)

Peak #2 (Tn Emission)


120 K
Peak #3 (Tn Emission)
Peak #4 (T2 Emission)
S1 140 K
T1 160 K
PL intensity

Fitted PL spectrum
180 K
Tn 200 K
220 K
240 K

400 450 500 550 600 650 700 750 400 450 500 550 600 650 700 750
Wavelength (nm) Wavelength (nm)
(c) (d)
S2
Hole Particle Transition
ISC
99% n→π*
S1 state

ISC
S1
S2 state 98%
π→π* T2

ISC X
Absorption

T1 state 91%
π→π*
F

T1
Anti-Kasha P

99%
P

T2 state n→π*
S0

Figure 11.11  (a) PL spectrum and peak fitting of DPPZ neat powder under ambient conditions.
(b) Temperature-­dependent PL spectra from 80 to 240 K. (c) Natural transition orbitals (NTO) images of the
S1, S2, T1, and T2 excited states of DPPZ. (d) Schematic Jablonski diagram with SOC matrix elements of
RTP-­related ISC processes in DPPZ. Source: Adapted with permission from Ref. [65], Wiley-­VCH, 2018.
326 11  Anti-­Kasha Emission from Organic Aggregates

(n, π*) character, S2 and T1 with (π, π*) character, respectively (Figure 11.11c). Due to the different
transition patterns of T2 and T1, the internal conversion between T2 and T1 should be probably
slowed down, and thus resulted in that the radiative transition of T2 can compete with the limited
IC process between T2 and T1, and finally lead to the dual phosphorescent emission from T2 with a
short lifetime and RTP from T1 with a little long lifetime (Figure 11.11d). This finding provided a
prototype strategy to achieve high-­efficiency, inexpensive, pure, white-­light emission in organic sin-
gle molecule with three standard primary colors through the precise modulation of excited states.
A series of fluorescent BF2-­hydrazine-­based rotatory dyes (BODIHY) with anti-­Kasha emission
and AIE were designed and synthesized by Ivan Aprahamian group (Figures 11.8d and 11.12) [66].
The electron-­withdrawing and electron-­donating groups were introduced at different positions on
the benzene ring of the molecule, and BODIHY molecule could rotate along the conjugated bonds
between the donor moiety and the acceptor moiety.
It was found that the luminescence intensity of these molecules was significantly related to the
viscosity of the solvent. In other words, the emission intensity was enhanced with the increasing
of viscosity (Figure 11.12a). The double logarithmic plot of maxima emission intensities versus
solvent viscosities yielded a good, linear correlation for these compounds. In addition, different
substituents also played a great role on the emission wavelength. It was found that the blue-­shift
emission and absorption occurred when the para position was substituted by the electron-­
withdrawing group. On the contrary, if the para position was substituted by electron-­donating
groups, molecules exhibited the bathochromic-­shifted emission and absorption [67].
DFT calculations and time-­dependent DFT models suggested that S1 of these molecules are dark
states and all emissions came from S2, because the radiative relaxation from S1 was very slow with
minimal oscillator strength and small energy gap between S1 and S0, which resulted in the ultrafast
nonradiative relaxation [48]. Thus, different energy gaps with various substituent groups resulted the
distinct IC from S2 to S1, exhibiting different brightness and quantum yields. Investigation of the
ground and excited state potential-­energy surface of these compounds suggested that the lowest
energy conformations were similar in the S0 and S1, ruling out twisted intramolecular CT mechanism.
There was a distinct correlation between the viscosity sensitivities and barriers to rotation in their
emissive states, which means that excited states would decrease the energy gap between emissive state

(a) Wavenumber (cm–1) (b)


35 000 30 000 25 000 20 000 15 000
30 000
1.8 2 500
25 000
2 000
Viscosity 20 000
Energy (cm–1)

1.2
1 500 15 000
A

Compound 1
1000 10 000
0.6 Ground state
Excited state 0.0 1.0
5 000 fOSC
500

0
0.0 0
300 350 400 450 500 550 600 650 0 50 100 150 200 250 300 350
λ (nm) Rotor angle (°)

Figure 11.12  (a) The absorption (dashed line) and fluorescence (solid lines) spectra of compound
BODIHY-­4 in different viscosity. (b) Ground (blue diamonds) and excited (shaded red diamonds) PBE TD-­DFT
state energies of compound BODIHY-­1 as a function of rotor angle. Source: Adapted with permission from
Ref. [66], Springer Nature, 2017.
  ­Reference 327

in S2 and dark state in S1 by rotation and give rapid IC from S2 to S1 in low-­viscosity solvents at room
temperature. At high viscosities, rotations were hindered and IC was decreased to improve fluores-
cence (Figure 11.12b). On account of these experimental results, it was proposed that the viscosity
sensitivity of BODIHY fluorescence is attributed to barrier to rotation in the emissive excited state,
which suggested the suppression of Kasha’s rule. This mechanism enriched the research area of AIE
and anti-­Kasha rule, and provided a model for the design of novel, luminescent, rotatable molecules.

11.5  ­Conclusions

In summary, we reviewed the examples of anti-­Kasha emission in aggregates. Several key factors
in realizing anti-­Kasha emission were carefully analyzed through full text. We classified them as
fast radiative process of upper excited states, limited IC process caused by energy gaps between
upper excited state and the lowest excited state, decreasing nonradiative depopulation caused by
rigid environment, the favorably allowed ISC and rISC process, and FRET and barrier to rotation
in the emissive excited state. Some novel prototype structures and strategies are provided as the
guidance to realize anti-­Kasha emission in aggregate state. However, there are still many unsolved
issues, such as low quantum yield, difficulty in practical applications.
An overview of these anti-­Kasha examples indicates that specific molecular structure is a prerequi-
site for anti-­Kasha emission. Limited specific molecular structure makes the development of materials
with anti-­Kasha emission proceeding laboriously. Another consideration is the proper methodology to
realize anti-­Kasha emission including aggregation, crystallization, forming supramolecular systems,
etc. The development of anti-­Kasha emission initiates an unexpectedly broad range of possibilities
from photochemical to material science. Nevertheless, it is expected that the anti-­Kasha luminescent
materials will play a valuable role in academic research and practical applications.

­References

1 Tsien, R.Y. (1998). The green fluorescent protein. Annual Review of Biochemistry 67: 509–44. https://
doi.org/10.1146/annurev.biochem.67.1.509.
2 Mitschke, U. and Bäuerle, P. (2000). The electroluminescence of organic materials. Journal of
Materials Chemistry 10: 1471–1507. https://doi.org/10.1039/A908713C.
3 Miao, W. (2008). Electrogenerated chemiluminescence and its biorelated applications. Chemical
Reviews 108: 2506–2553. https://doi.org/10.1021/cr068083a.
4 Kasha, M. (1950). Characterization of electronic transitions in complex molecules. Discussions of the
Faraday Society 9: 14–19. https://doi.org/10.1039/DF9500900014.
5 Kasha, M. and McGlynn, S.P. (1956). Molecular electronic spectroscopy. Annual Review of Physical
Chemistry 7: 403–424. https://doi.org/annurev.pc.07.100156.002155.
6 Klán, P. and Wirz, J. (2009). Photochemistry of organic compounds: From concepts to practice.
Hoboken, NJ: Wiley-­Blackwell.
7 Turro, N.J., Ramamurthy, V., Ramamurthy, V., et al. (2009). Principles of molecular photochemistry:
An introduction. Sausalito, CA: University Science Books.
8 Itoh, T. (2012). Fluorescence and phosphorescence from higher excited states of organic molecules.
Chemical Reviews 112: 4541–4568. https://doi.org/10.1021/cr200166m.
9 Förster, T.T. and Kasper, K. (1954). Ein Konzentrationsumschlag der Fluoreszenz. Zeitschrift fur
Physikalische Chemie 1: 275–277. https://doi.org/10.1524/zpch.1954.1.5_6.275.
328 11  Anti-­Kasha Emission from Organic Aggregates

10 Birks, B.J. (1970). Photophysics of aromatic molecules. Berichte Der Bunsengesellschaft Für
Physikalische Chemie 74: 1294–1295. https://doi.org/10.1002/bbpc.19700741223.
11 Bakalova, R., Zhelev, Z., Aoki, I., et al. (2006). Silica-­shelled single quantum dot micelles as
imaging probes with dual or multimodality. Analytical Chemistry 78: 5925–5932. https://doi.
org/10.1021/ac060412b.
12 Zhelev, Z., Ohba, H. and Bakalova, R. (2006). Single quantum dot-­micelles coated with silica shell
as potentially non-­cytotoxic fluorescent cell tracers. Journal of the American Chemical Society 128:
6324–6325. https://doi.org/10.1021/ja061137d.
13 Mei, J., Leung, N.L.C., Kwok, R.T.K., et al. (2015). Aggregation-­induced emission: Together we
shine, united we soar! Chemical Reviews 115: 11718–11940. https://doi.org/10.1021/acs.
chemrev.5b00263.
14 Zhou, Y., Zhuang, Y., Li, X., et al. (2017). Selective dual-­channel imaging on cyanostyryl-­modified
azulene systems with unimolecularly tunable visible-­near infrared luminescence. Chemistry-­A
European Journal 23: 7642–7647. https://doi.org/10.1002/chem.201700947.
15 Zhou, Y., Zou, Q.Z., Qiu, J., et al. (2018). Rational design of a green-­light-­mediated unimolecular
platform for fast switchable acidic sensing. The Journal of Physical Chemistry Letters 9: 550–556.
https://doi.org/10.1021/acs.jpclett.7b03233.
16 Luo, J., Xie, Z., Lam, J.W.Y., et al. (2001). Aggregation-­induced emission of 1-­methyl-­1,2,3,4,5-­
pentaphenylsilole. Chemical Communications 381: 1740–1741. https://doi.org/10.1039/b105159h.
17 Chen, J. and Tang, B.Z. (2013). Aggregation-­induced emission: Fundamentals and applications.
Wiley Online Library.
18 Leung, N.L.C., Xie, N., Yuan, W., et al. (2014). Restriction of intramolecular motions: The general
mechanism behind aggregation-­induced emission. Chemistry-­A European Journal 20: 15349–
15353. https://doi.org/10.1002/chem.201403811.
19 Tu, Y., Liu, J., Zhang, H., et al. (2019). Restriction of access to the dark state: A new mechanistic
model for heteroatom-­containing AIE systems. Angewandte Chemie International Edition 58:
14911–14914. https://doi.org/10.1002/anie.201907522.
20 Demchenko, A.P., Tomin, V.I. and Chou, P.T. (2017). Breaking the Kasha rule for more efficient
photochemistry. Chemical Reviews 117: 13353–13381. https://doi.org/10.1021/acs.
chemrev.7b00110.
21 Basak, S., Nandi, N., Bhattacharyya, K., et al. (2015). Fluorescence from an H-­aggregated
naphthalenediimide based peptide: Photophysical and computational investigation of this rare
phenomenon. Physical Chemistry Chemical Physics 17: 30398–30403. https://doi.org/10.1039/
c5cp05236j.
22 Rösch, U., Yao, S., Wortmann, R., et al. (2006). Fluorescent H-­aggregates of merocyanine dyes.
Angewandte Chemie International Edition 45: 7026–7030. https://doi.org/10.1002/anie.200602286.
23 Kumar, N.S.S., Varghese, S., Suresh, C.H., et al. (2009). Correlation between solid-­state
photophysical properties and molecular packing in a series of indane-­1,3-­dione containing
butadiene derivatives. The Journal of Physical Chemistry C 113: 11927–11935. https://doi.
org/10.1021/jp902482r.
24 Kumar, M. and George, S.J. (2011). Spectroscopic probing of the dynamic self-­assembly of an
amphiphilic naphthalene diimide exhibiting reversible vapochromism. Chemistry-­A European
Journal 17: 11102–11106. https://doi.org/10.1002/chem.201101642.
25 Kumar, M. and George, S.J. (2011). Green fluorescent organic nanoparticles by self-­assembly
induced enhanced emission of a naphthalene diimide bolaamphiphile. Nanoscale 3: 2130–2133.
https://doi.org/10.1039/c1nr10151j.
  ­Reference 329

26 Jissy, A.K. and Datta, A. (2013). Can arsenates replace phosphates in natural biochemical
processes? A computational study. The Journal of Physical Chemistry B 117: 8340–8346. https://doi.
org/10.1021/jp402917q.
27 Jissy, A.K. and Datta, A. (2013). What stabilizes the LinPn inorganic double helices? The Journal of
Physical Chemistry Letters 4: 1018–1022. https://doi.org/10.1021/jz400263y.
28 Guo, J., Fan, J., Lin, L., et al. (2019). Mechanical insights into aggregation-­induced delayed
fluorescence materials with anti-­Kasha behavior. Advanced Science 6: 1801629. https://doi.
org/10.1002/advs.201801629.
29 Leitl, M.J., Krylova, V.A., Djurovich, P.I., et al. (2014). Phosphorescence versus thermally activated
delayed fluorescence. Controlling singlet-­triplet splitting in brightly emitting and sublimable Cu(I)
compounds. Journal of the American Chemical Society 136: 16032–16038. https://doi.org/10.1021/
ja508155x.
30 Chen, X.K., Zhang, S.F., Fan, J.X., et al. (2015). Nature of highly efficient thermally activated
delayed fluorescence in organic light-­emitting diode emitters: Nonadiabatic effect between excited
states. The Journal of Physical Chemistry C 119: 9728–9733. https://doi.org/10.1021/acs.
jpcc.5b00276.
31 Etherington, M.K., Gibson, J., Higginbotham, H.F., et al. (2016). Revealing the spin-­vibronic
coupling mechanism of thermally activated delayed fluorescence. Nature Communications 7:
13680. https://doi.org/10.1038/ncomms13680.
32 Marian, C.M. (2016). Mechanism of the triplet-­to-­singlet upconversion in the assistant dopant
ACRXTN. The Journal of Physical Chemistry C 120: 3715–3721. https://doi.org/10.1021/acs.
jpcc.6b00060.
33 Park, I.S., Matsuo, K., Aizawa, N., et al. (2018). High-­performance dibenzoheteraborin-­based
thermally activated delayed fluorescence emitters: Molecular architectonics for concurrently
achieving narrowband emission and efficient triplet-­singlet spin conversion. Advanced Functional
Materials 28: 1802031. https://doi.org/10.1002/adfm.201802031.
34 He, Z., Zhao, W., Lam, J.W.Y., et al. (2017). White light emission from a single organic molecule
with dual phosphorescence at room temperature. Nature Communications 8: 416–423. https://doi.
org/10.1038/s41467-­017-­00362-­5.
35 Chu, S.Y. and Goodman, L. (1975). A simple theoretical model for dual phosphorescence. Chemical
Physics Letters 32: 24–27. https://doi.org/10.1016/0009-­2614(75)85160-­8.
36 Prieto, M.F.R., Nickel, B., Grellmann, K. H., et al. (1988). Dual phosphorescence from 2-­(2′-­
hydroxyphenyl)benzoxazole due to keto-­enol tautomerism in the metastable triplet state. Chemical
Physics Letters 146: 387–392. https://doi.org/10.1016/0009-­2614(88)87464-­5.
37 Chaudhuri, D., Sigmund, E., Meyer, A., et al. (2013). Metal-­free OLED triplet emitters by side-­
stepping Kasha’s rule. Angewandte Chemie International Edition 52: 13449–13452. https://doi.
org/10.1002/anie.201307601.
38 Wagner, P.J., May, M.J., Haug, A., et al. (1970). Phosphorescence of phenyl alkyl ketones. Journal of
the American Chemical Society 92: 5269–5270. https://doi.org/10.1021/ja00720a073.
39 Wagner, P.J., Kemppainen, A.E. and Schott, H.N. (1973). Effects of ring substituents on the type II
photoreactions of phenyl ketones. How interactions between nearby excited triplets affect chemical
reactivity. Journal of the American Chemical Society 95: 5604–5614. https://doi.org/10.1021/
ja00798a027.
40 Itoh, T. (2004). Successive occurrence of the T1(π, π*) and T2(n, π*) phosphorescence and the
S1(n, π*) fluorescence observed for p-­cyanobenzaldehyde in a solid matrix. Journal of Luminescence
109: 221–225. https://doi.org/10.1016/j.jlumin.2004.03.002.
330 11  Anti-­Kasha Emission from Organic Aggregates

41 Paul, L., Moitra, T., Ruud, K., et al. (2019). Strong Duschinsky mixing induced breakdown of
Kasha’s rule in an organic phosphor. The Journal of Physical Chemistry Letters 10: 369–374. https://
doi.org/10.1021/acs.jpclett.8b03624.
42 Etinski, M., Tatchen, J. and Marian, C.M. (2011). Time-­dependent approaches for the calculation
of intersystem crossing rates. The Journal of Chemical Physics 134: 154105. https://doi.
org/10.1063/1.3575582.
43 Föller, J., Kleinschmidt, M. and Marian, C.M. (2016). Phosphorescence or thermally activated
delayed fluorescence? Intersystem crossing and radiative rate constants of a three-­coordinate
copper (I) complex determined by quantum-­chemical methods. Inorganic Chemistry 55: 7508–
7016. https://doi.org/10.1021/acs.inorgchem.6b00818.
44 Moitra, T., Alam, M.M. and Chakrabarti, S. (2018). Intersystem crossing rate dependent dual
emission and phosphorescence from cyclometalated platinum complexes: A second order
cumulant expansion based approach. Physical Chemistry Chemical Physics 20: 23244–23251.
https://doi.org/10.1039/c8cp03111h.
45 Younker, J.M. and Dobbs, K.D. (2013). Correlating experimental photophysical properties of
iridium (III) complexes to spin–orbit coupled TDDFT predictions. The Journal of Physical
Chemistry C 117: 25714–25723. https://doi.org/10.1021/jp410576a.
46 Brédas, J.L., Beljonne, D., Coropceanu, V., et al. (2004). Charge-­transfer and energy-­transfer
processes in π-­conjugated oligomers and polymers: A molecular picture. Chemical Reviews 104:
4971–5003. https://doi.org/10.1021/cr040084k.
47 Wu, Y., Xiao, H., Chen, B., et al. (2020). Multiple-­state emissions from neat, single-­component
molecular solids: Suppression of Kasha’s rule. Angewandte Chemie International Edition 59:
10173–10178. https://doi.org/10.1002/anie.202000608.
48 Englman, R. and Jortner, J. (1970). Energy gap law for radiationless transitions in large molecules.
Molecular Physics 18: 145–164. https://doi.org/10.1080/00268977000100171.
49 Murata, S., Iwanaga, C., Toda, T., et al. (1972). Fluorescence yields of azulene derivatives. Chemical
Physics Letters 13: 101–104. https://doi.org/10.1016/0009-­2614(72)80054-­X.
50 Griesser, H.J. and Wild, U.P. (1980). The energy gap dependence of the radiationless transition
rates in azulene and its derivatives. Chemical Physics 52: 117–131. https://doi.
org/10.1016/0301-­0104(80)85190-­1.
51 Ma, H., Peng, Q., An, Z., et al. (2019). Efficient and long-­lived room-­temperature organic
phosphorescence: Theoretical descriptors for molecular designs. Journal of the American Chemical
Society 141: 1010–1015. https://doi.org/10.1021/jacs.8b11224.
52 Zhang, Y., Yang, H., Ma, H., et al. (2019). Excitation wavelength dependent fluorescence of an
ESIPT triazole derivative for amine sensing and anti-­counterfeiting applications. Angewandte
Chemie International Edition 58: 8773–8778. https://doi.org/10.1002/anie.201902890.
53 El-­Sayed, M.A. (1968). The triplet state: Its radiative and nonradiative properties. Accounts of
Chemical Research 1: 8–16. https://doi.org/10.1021/ar50001a002.
54 Wang, T., Su, X., Zhang, X., et al. (2019). Aggregation-­induced dual-­phosphorescence from organic
molecules for nondoped light-­emitting diodes. Advanced Materials 31: 1904273. https://doi.
org/10.1002/adma.201904273.
55 Sun, X.X., Wang, X.J., Li, X.Y., et al. (2015). Polymerization-­enhanced intersystem crossing: New
strategy to achieve long-­lived excitons. Macromolecular Rapid Communications 36: 298–303.
https://doi.org/10.1002/marc.201400529.
56 Chai, J. and Head Gordon, M. (2008). Long-­range corrected hybrid density functionals with
damped atom–atom dispersion corrections. Physical Chemistry Chemical Physics 10: 6615–6620.
https://doi.org/10.1039/b810189b.
  ­Reference 331

57 Wang, T., Su, X., Zhang, X., et al. (2019). A combinatory approach towards the design of organic
polymer luminescent materials. Journal of Materials Chemistry C 7: 9917–9925. https://doi.
org/10.1039/c9tc02266j.
58 Zhao, W., He, Z., Lam, Jacky W.Y., et al. (2016). Rational molecular design for achieving persistent
and efficient pure organic room-­temperature phosphorescence. Chem 1: 592–602. https://doi.
org/10.1016/j.chempr.2016.08.010.
59 Griesser, H.J. and Wild, U.P. (1980). The fluorescence lifetimes of carbonyl derivatives of azulene
showing dual emission. Journal of Photochemistry 13: 309–318. https://doi.org/10.1016/
0047-­2670(80)80022-­0.
60 Makinoshima, T., Fujitsuka, M., Sasaki, M., et al. (2004). Competition between intramolecular
electron-­transfer and energy-­transfer processes in photoexcited azulene-­C60 dyad. The Journal of
Physical Chemistry A 108: 368–375. https://doi.org/10.1021/jp0342428.
61 Murai, M., Ku, S.-­Y., Treat, N.D., et al. (2014). Modulating structure and properties in organic
chromophores: Influence of azulene as a building block. Chemical Science 5: 3753–3760. https://
doi.org/10.1039/c4sc01623h.
62 Xin, H. and Gao, X. (2017). Application of azulene in constructing organic optoelectronic
materials: New tricks for an old dog. ChemPlusChem 82: 945–956. https://doi.org/10.1002/
cplu.v82.7.
63 Zhou, Y., Baryshnikov, G., Li, X., et al. (2018). Anti-­Kasha’s rule emissive switching induced by
intermolecular H-­bonding. Chemistry of Materials 30: 8008–8016. https://doi.org/10.1021/acs.
chemmater.8b03699.
64 Gong, Y., Zhou, Y., Yue, B., et al. (2019). Multiwavelength anti-­Kasha’s rule emission on self-­
assembly of azulene-­functionalized persulfurated arene. The Journal of Physical Chemistry C 123:
22511–22518. https://doi.org/10.1021/acs.jpcc.9b06731.
65 Zhou, C., Zhang, S., Gao, Y., et al. (2018). Ternary emission of fluorescence and dual
phosphorescence at room temperature: A single-­molecule white light emitter based on pure
organic aza-­aromatic material. Advanced Functional Materials 28: 1802407. https://doi.
org/10.1002/adfm.201802407.
66 Qian, H., Cousins, M.E., Horak, E.H., et al. (2017). Suppression of Kasha’s rule as a mechanism for
fluorescent molecular rotors and aggregation-­induced emission. Nature Chemistry 9: 83–87.
https://doi.org/10.1038/nchem.2612.
67 Mei, J., Hong, Y., Lam, J.W.Y., et al. (2014). Aggregation-­induced emission: The whole is more
brilliant than the parts. Advanced Materials 26: 5429–5479. https://doi.org/10.1002/
adma.201401356.
333

12

Aggregation-enhanced Emission: From Flexible to Rigid Cores


Harnimarta Deol, Gurpreet Singh, Vandana Bhalla, and Manoj Kumar
Department of Chemistry, UGC Sponsored Centre for Advanced Studies-­II, Guru Nanak Dev University, Amritsar, PB, India

12.1 ­Introduction

It is a well-­known fact that the fluorescence of traditional organic dyes is weakened at higher con-
centrations due to the formation of aggregates. The reason for this weakening of fluorescence is the
intermolecular π–π stacking between the molecules that leads to aggregation and hence causes
quenching [1]. However, the vast applications of the solid-­state emitters in various fields such as
organic electronics, optical sensing, and bioimaging demand the development of those assemblies
which have strong emission on aggregation [2].
In 2001, Tang and coworker proposed the novel concept of “aggregation-­induced emission”
(AIE) in silole derivatives, and the saying “together we shine” was materialized [3]. Another
way round to aggregation-­caused quenching (ACQ), the silole derivatives were nonemissive in
dilute solution, but emitted strongly in the aggregated state. In simple words, through AIE
phenomenon one can overcome the limitations of self-­aggregation and actually take the ben-
efit of as prepared solid-­state emitters. Recently, the term aggregation-­enhanced emission
(AEE) has been introduced, which is also known as AIEE (aggregation-­induced emission
enhancement), where molecules having twisted geometry overcome their weak emission via
coplanarization leading to the generation of emissive J-­aggregates [4]. In these materials, the
relaxed planar excited state in aggregates undergoes the suppressed nonradiative decay via
molecular stacking that results in enhanced emission. The key requirements for the develop-
ment of AIE/AEE active materials are twisted geometry and the presence of free rotors.
The restricted intramolecular rotation of free rotors and reduced intermolecular π–π stacking
due to twisted geometry results in emission enhancement upon aggregation. The restricted
intramolecular rotation (RIR) effect can be induced by simple approaches such as: (i) intro-
duction of freely moving rotors (ii) complexation with analytes such as metal ions/anions/
other molecules. Other mechanisms that can introduce AEE effect are restriction of ICT/TICT
(see Figure 12.1).

Handbook of Aggregation-Induced Emission: Volume 1 Tutorial Lectures and Mechanism Studies, First Edition.
Edited by Youhong Tang and Ben Zhong Tang.
© 2022 John Wiley & Sons Ltd. Published 2022 by John Wiley & Sons Ltd.
334 12  Aggregation-enhanced Emission: From Flexible to Rigid Cores

Interaction
with different
Introduction of analytes such
freely moving as
rotors cations/anions/
biomolecules

Mechanism of
AEE

LE state Restriction of
intramolecular
ICT/TICT charge transfer
state (ICT)/twisted
Emission from intramolecular
Excitation LE state charge transfer H2O
Acceptor Donor Upon
(TICT) ICT
aggregation
ICT/TICT
Acceptor Donor
inhibited
TICT and
emission
no due to AEE
r
Acceptor Do

Figure 12.1  Different mechanism to induce AEE phenomena.

12.2 ­Freely Moving Rotors-­induced Emission Enhancement

The free rotors present in traditional hexaphenylsilole  [5], tetraphenylethylene (TPE)  [6], and
distyrlanthracene [7] induce AIE characteristics in the molecules (see Figure 12.2). These AIEgens
exhibited various fascinating properties, and were utilized as OLEDs, chemo/biosensors, and
photosensitizers [8].
To widen the scope of AEE active fluorophores, from our laboratory in 2012, we reported het-
erooligomer derivative 1 having peripheral heteroaromatic rotors covalently linked to the central

Si

Hexaphenylsilole Tetraphenylethylene Distyrylanthracene

Figure 12.2  Traditional AIEgens.


12.2  ­Freely Moving Rotors-­induced Emission Enhancemen 335

N N

O O
CH3 H3C
1

(a) (b)
1000 % of H2O
900
800 0
700 10
20
Intensity

600
30
500 40
400 50
300 60
200 70
80
100
90 1μm
0
250 300 350 400 450 500 550 600
Wavelength (nm)

Figure 12.3  (a) Fluorescence emission spectra of 1 in different ratios of H2O : THF. (b) SEM image of
aggregates of derivative 1 in H2O : THF (80 : 20). Source: Reproduced with permission from Ref. [9].
Copyright 2012 Wiley-­VCH.

benzene stator [9]. The derivative 1 emitted weakly in tetrahydrofuran (THF); however, upon addi-
tion of water enhancement in the emission intensity (ϕAEE = 1.15) was observed (see Figure 12.3a)
along with the formation of spherical aggregates that was confirmed by scanning electron micro-
scope (SEM) studies (see Figure 12.3b). In the presence of water as a cosolvent, the intramolecular
rotations are restricted that leads to release of energy through radiative channel (see Scheme 12.1).
This is the first report in which fully aromatic AEE active material was prepared by covalently con-
necting aromatic rotors to the aromatic stator. Further, the fluorescent aggregates of derivative 1
were utilized as a sensor for the detection of biomolecules such as bovine serum albumin (BSA)
and cytochrome c. The aggregates of derivative 1 showed emission enhancement in the presence
of BSA, while fluorescence was quenched in the presence of cytochrome c. The aggregates of deriv-
ative 1 were also utilized for the detection of DNA over RNA in aqueous media. The emission of
the aggregates of derivative 1 was quenched in the presence of DNA while no significant change in
the emission intensity was observed in the presence of RNA. Building on same concept, we synthe-
sized heterooligomer derivatives 2 and 3 [10] by substituting the pyridyl moieties with sterically
hindered carbazole unit. Both the derivatives exhibited AEE characteristics (see Figures  12.4a
and  12.5a) as supported by viscosity-­/concentration-­dependent fluorescence studies and
concentration-­dependent 1H NMR studies. In the isolated state, derivative 3 exhibited higher
quantum yield (ϕ  =  0.006) than derivative 2 (ϕ  =  0.0018). However, a higher AEE effect was
observed upon aggregation in the case of derivative 2 (ϕ  =  0.59) in comparison to derivative 3
(ϕ = 0.43). Actually, in the case of derivative 3, rotation of rotors is already restricted in an isolated
state (THF) because of the presence of higher number of sterically hindered rotors, and therefore,
lesser emission enhancement is observed upon aggregation (see Figures  12.4b and  12.5b). The
Rotors

H2O

Monomeric form Fluorescent aggregates

Scheme 12.1  Formation of fluorescent aggregates of derivative 1 upon addition of water.

C6H13
N

H3C
O

O
CH3
2

(a) (b)
% H2O in THF
450
80
400
70
350 Rotors Sterically
90
300 restricted
60
rotor
Intensity

250 50
200 H2O
40
150 30
100 20
50 10
0 Monomeric form Fluorescent aggregates
0
340 360 380 400 420 (ϕ = 0.0018) (ϕAEE = 0.59)
Wavelength (nm)

(c)

2 μm

Figure 12.4  (a) Fluorescence emission spectra of 2 in different ratio of H2O : THF at λex = 290 nm.


(b) Schematic representation of formation of fluorescent aggregate of derivative 2. (c) SEM image of
aggregates of derivative 2 in H2O : THF (80 : 20). Source: Reprinted with permission from Ref. [10]. Copyright
2012 American Chemical Society.
12.2  ­Freely Moving Rotors-­induced Emission Enhancemen 337

C6H13
N

H3C
O

O
CH3
3

(a) % H2O in THF


80
450
70
400 90
350 60
50 (b)
300
Intensity

40 Rotors Sterically
250 A B 30 restricted
200 20 rotor
150 10 H2O
0
100
50
Monomeric form
0
(ϕ = 0.006) (ϕAEE = 0.43)
345 395 445
Wavelength (nm)

(c)

2 μm

Figure 12.5  (a) Fluorescence emission spectra of 3 in different ratio of H2O : THF. λex = 290 nm. The inset
shows the difference in fluorescence of 3 in THF (A) and 80% H2O in THF (b) Schematic representation of
formation of fluorescent aggregate of derivative 3. (c) SEM image of aggregates of derivative 3 in H2O : THF
(80 : 20). Source: Reprinted with permission from Ref. [10]. Copyright 2012 American Chemical Society.

SEM images in both the cases showed the formation of spherical aggregates (see Figures  12.4c
and 12.5c). These studies highlighted the important contribution of number of rotors as well as
their steric bulk towards endowing AEE characteristics in the molecules. We utilized same strategy
to induce AEE characteristics in derivative 4 based on the rigid terphenyl core  [11]. Terphenyl
derivative 4 showed emission enhancements upon addition of water due to restriction to all the
motions (see Figure 12.6). To explore the potential of polyaromatic scaffolds for the preparation of
emissive materials, we aimed at developing AEE active materials using highly rigid and flat hexa-­
peri-­hexabenzocoronene (HBC) core [12]. HBC derivatives are known for their spontaneous aggre-
gation through intermolecular π–π stacking of aromatic groups which eventually results in ACQ. To
make HBC core AEE active, we covalently attached rotors to the core and synthesized HBC
338 12  Aggregation-enhanced Emission: From Flexible to Rigid Cores

O O O O

O O
O O
C C
NH HN

O O

O C HN NH C O
O O
O O

HN NH
C C
O O
O O

O O O O

600

500 % H2O
0
400
Intensity (a.u.)

60
300 70

200 80
90
100

0
375 425 475 525 575
Wavelength (nm)

Figure 12.6  Fluorescence spectrum of derivative 4 in different fractions of water in THF. Source:


Reproduced from Ref. [11]. Copyright 2016 with permission from Royal Society of Chemistry.

derivatives 5 and 6. Interestingly, the presence of rotors induced the AEE characteristics (see
Figures 12.7a and 12.8a) in coronene-­based molecules. The THF solution of derivative 5 was found
to be emissive (ϕ = 0.41) and upon introducing water fractions upto 40, 67% enhancement in the
emission intensity (ϕ  =  0.68) was observed. On the contrary, in the case of derivative 6, 86%
12.2  ­Freely Moving Rotors-­induced Emission Enhancemen 339

(a)
1000
900 H2O in THF
800 0
700 20
600

Intensity
40
500 50
400
70
300
200 90
100
0
450 500 550 600 650
Wavelength (nm)

(b) (c)
R R
5, R = NH2
O
6, R = C12H25
N N
H H
5 µm 1 µm

Figure 12.7  (a) Fluorescence spectrum of compound 5 in different H2O : THF ratios. SEM images of
derivatives 5 (b) in H2O : THF, 4 : 6 and (c) in H2O : THF, 9 : 1 solution. Source: Reprinted with permission
from Ref. [12]. Copyright 2013 American Chemical Society.

(a)
1000 % of H2O
0
800 20
40
600 50
Intensity

70
400 90

200

0
450 500 550 600
Wavelength (nm)
(b) (c)

5 µm 1 µm

Figure 12.8  (a) Fluorescence spectrum of compound 6 in different H2O : THF ratios. SEM images of
derivatives 6 (b) in H2O : THF, 4 : 6 and (c) in H2O : THF, 9 : 1 solution. Source: Reprinted with permission
from Ref. [12]. Copyright 2013 American Chemical Society.
340 12  Aggregation-enhanced Emission: From Flexible to Rigid Cores

R
R
R
R
H2O
R

R
Monomeric in R
THF R

(5, ϕ = 0.41) R
(6, ϕ = 0.47) Spherical aggregates in R
THF: H2O (6 : 4)
(5, ϕ = 0.68)
(6, ϕ = 0.73)

Scheme 12.2  Schematic representation of formation spherical aggregates of derivatives 5 and 6 on


increasing water content.

enhancement in the emission intensity (ϕ = 0.73) was observed upon increasing the water content.
The higher AEE characteristics observed in the case of derivative 6 is assigned to the presence of
higher number rotatable bonds in the molecule. However, a further increase in water content,
resulted in decrease in emission intensity in both the derivatives that may be attributed to π–π
interactions between the coronene units which affects the intramolecular conjugation [13–17] (see
Scheme 12.2). The SEM images of derivatives 5 and 6 supported the aforementioned explanation.
The formation of spherical aggregates (see Figures 12.7b and 12.8b) in H2O : THF (4 : 6) solvent
mixture is observed in derivatives 5 and 6, however, their SEM images in THF solution having
water content upto 90% showed the formation of wire-­shaped assemblies (see Figures  12.7c
and 12.8c), possibly because of the π−π intermolecular interactions between discotic molecules.
Moving in the same direction, we aimed at developing AEE active materials using rigid and rela-
tively electron-­deficient pentacenequinone scaffold. To overcome the ACQ induced due to efficient
intermolecular π–π stacking of aromatic groups, we focused on developing J-­aggregates. The strat-
egy worked very well in the case of donor–acceptor–donor (D–A–D) system 7 based on pentacene-
quinone scaffold. The electron-­rich thiophene moieties were introduced as donors/rotors at the
periphery of the rigid pentacenequinone core  [18]. The derivative 7 exhibited very interesting
emission behavior due to combined effect of TICT and AEE phenomenon. Derivative 7 exhibited a
weak emission (ϕ = 0.2) band at 460 nm in THF and this band was redshifted to 492 nm upon addi-
tion of 10% water with drastic emission enhancement. Further addition of the water fraction up to
50% led to decrease in emission intensity (see Figure 12.9a). Very interestingly, upon addition of
water fractions upto 60%, an enhancement in the emission intensity and the redshifting of the
emission band to 560 nm was observed. Nearly twofold enhancement in emission intensity was
observed on increasing the water content to 90% (see Figure 12.9b). In H2O : THF (1 : 1) mixture,
derivative 7 adopts a twisted conformation in polar protic media and thus exhibits TICT state in
excited state which due to its susceptibility to nonradiative decay led to decrease in emission inten-
sity before the onset of expected aggregation of molecules in mixed aqueous media. In H2O : THF
(9  :  1) mixture, formation of J-­nanoaggregates of the derivative 7 is the reason behind all the
changes in photophysical properties (see Figure  12.9c). To get more insight into mechanism of
AEE phenomena, we synthesized derivatives 8–10, which differ from each other in terms of a
number of rotors as well as the electronic nature of rotors. In the synthesized mono, di, and tetra
substituted pentacenequinone derivatives 8–10, we evaluated the relation between position/­
number of molecular rotors, and emission enhancement [19]. Interestingly, derivative 10 having
12.2  ­Freely Moving Rotors-­induced Emission Enhancemen 341

O
S

S
O
7

(a) (b)
160 350 Water fraction (%)
Water fraction (%)
140 50 300 90

120 40 80
250 70
30
Intensity

100
20 200 60
80 10
150
60 0
100
40

20 50

0 0
400 450 500 550 600 650 450 500 550 600
Wavelength (nm) Wavelength (nm)

(c) TICT

TICT
Donor
H2O : THF
(9 : 1)
Acceptor

J-aggregates
(ϕ = 0.2) (ϕ = 0.4)

Figure 12.9  Fluorescence spectra of compound 7 showing the variation of fluorescence intensity in


H2O : THF mixture: (a) from 0 to 50%; (b) from 60 to 90% volume fractions of water in THF; λex = 328 nm.
(c) Schematic representation of inhibition of TICT phenomenon and formation of J aggregates upon addition
of water. Source: Reprinted with permission from Ref. [18]. Copyright 2012 American Chemical Society.

four pyridyl rotors showed a 16-­fold emission enhancement that was greater than the other two
derivatives (see Figure 12.10a). The viscosity-­dependent studies, concentration-­dependent fluores-
cent studies, and time-­resolved fluorescent studies supported the RIR mechanism in these deriva-
tives. The SEM studies proved the formation of aggregates in mixed aqueous media (see
Figure 1.10b). The dynamic light scattering (DLS) studies suggested the formation of smaller-­sized
aggregates in the case of derivative 8, while bigger-­sized aggregates were formed in the case of
derivatives 9 and 10 (see Table 12.1). It is proposed that due to stronger π–π intermolecular interac-
tions in case of derivative 8, small-­sized aggregates are observed; whereas in case of derivatives 9
and 10, presence of higher number of pyridyl groups at the periphery impaired the face-­to-­face
intermolecular π–π stacking; and thus, aggregation-­driven growth of molecules is observed
342 12  Aggregation-enhanced Emission: From Flexible to Rigid Cores

O O
N N

N
O O
8 9

O
N N

N
N O
10

(a)
900 Water fraction (%) (b)
50
750 40
30
600 20
Intensity

10
450 0
300

150 2 μm
0
400 450 500 550 600
Wavelength (nm)

(c)

THF : H2O
(1 :1)

Monomeric form
ϕ = 0.027
Fluorescent aggregates
Aggregation driven growth of Size = 768 nm
the aggregates ϕ = 0.43

Figure 12.10  (a) Fluorescence spectra of derivative 10 showing the variation of fluorescence intensity in
H2O : DMSO mixtures. λex = 322 nm. (b) SEM image 10 showing the formation of aggregates. (c) Schematic
representation of change in morphology of pentacenequinone derivative 10 on increasing the number of
rotors. Source: Reproduced from Ref. [19]. Copyright 2014 with permission from Royal Society of Chemistry.

(see Figure 12.10c). The formation of bigger-­sized aggregates extended the delocalization of the


electron cloud, and hence, greater enhancement of the emission intensity is observed in the case
of derivatives 9 and 10. Based on all the studies, a hindrance to face-­to-­face intermolecular π–π
stacking and RIR is the proposed AEE mechanism in these derivatives. To understand the AEE
mechanism further, we prepared D–A–D systems 11 and 12. Being D–A–D systems, both the deriv-
atives formed luminescent J-­aggregates (see Scheme 12.3) that were not observed in the case of
derivatives 8–10 due to the presence of electron-­deficient pyridyl groups at the periphery. Thus, the
electronic nature of substituents greatly influences the type of packing of the molecules. Both the
12.2  ­Freely Moving Rotors-­induced Emission Enhancemen 343

Table 12.1  Comparative photophysical properties of derivatives 8-­10.

Quantum yield Quantum yield Emission DLS Kf (s−1) Knr


Derivative in solution in aggregates enhancement (size) (nm) τF1 (ns) τF2 (ns) (109 s−1) (109 s−1)

8 0.030 0.22 7.3 220 0.30 2.02 0.73 2.6


9 0.029 0.41 14.13 650 0.078 1.63 0.25 0.36
10 0.027 0.43 15.92 768 0.13 1.84 0.233 0.30

τF1 and τF1 biexponential life times, Kf radiative rate constant, Knr nonradiative rate constant.

ICT
ICT
OC6H13
O Donor H2O : THF
(9 : 1)

Acceptor

O
OC6H13
ϕ = 0.24
11 ϕ = 0.0039
J-aggregates

ICT
ICT
O Donor H2O : THF
(9 : 1)
Acceptor

12 ϕ = 0.0064 ϕ = 0.15
J-aggregates

Scheme 12.3  Schematic representation of inhibition of ICT and formation of J-­aggregates of derivatives
11 and 12 upon addition of water content.

(a) (b)
400 Water fraction 140
90 Water fraction (%)
350 70 120
90
300 60 100 70
250 30
Intensity

80 60
Intensity

200 10 30
0 60
150 10
40 0
100
50 20

0 0
450 470 490 510 530 550 570 590 430 450 470 490 510 530 550 570 590
Wavelength (nm) Wavelength (nm)

Figure 12.11  Fluorescence spectra of (a) derivative 11 and (b) 12 showing the variation of fluorescence
intensity in H2O : THF mixtures with different water fractions. Source: Reproduced from Ref. [19]. Copyright
2014 with permission from Royal Society of Chemistry.

derivatives 11 and 12 are weakly emissive in THF, and a dual emission band was observed on
increasing the water fraction more than 50%. An increase in fluorescence intensity of the emission
band at longer wavelength was observed upon increasing the water content to 90% in THF solution
of both the derivatives 11 (see Figure 12.11a) and 12 (see Figure 12.11b). The emission band at a
shorter wavelength corresponds to locally excited state, and the band at a longer wavelength
344 12  Aggregation-enhanced Emission: From Flexible to Rigid Cores

corresponds to an intermolecular charge transfer state. In the time-­resolved fluorescence studies,


observed a higher lifetime in case of derivative 12 upon aggregation as compared to the aggregates
of derivative 11. This is attributed to effective π-­stacking of phenyl rings of derivative 12 as the
presence of alkyl side chains in derivative 11 prevents the inter chain π-­stacking (see Scheme 12.3).
Thus, the conformational fixation upon aggregation controls the extent of emission enhancement.

12.3 ­Guest-­induced Emission Enhancement

In continuation of our efforts in the same direction, we developed a variety of weakly emitting
materials in which restriction to rotations is induced strategically by the presence of different ana-
lytes such as cations, anions, and molecules. The general approach involved developing small
building blocks with rotors as well as specific binding sites for target analyte. Due to the presence
of rotors, these building blocks self-­assemble in mixed aqueous media. However, to achieve the
desired restriction, an ordered-­planarized architecture is attained only in the presence of the target
analyte. The AEE behavior of these building blocks in the presence of different analytes such as
metal ions, anions, and small molecules has been evaluated.
To begin with, we synthesized hexaphenylbenzene (HPB) derivative 13 having imine linkages as
binding sites for soft metal ions [20]. The HPB derivative 13 is nonemissive (ϕ = 0.03) in mono-
meric form because of the rapid isomerization of C–N bond in excited state, however, upon addi-
tion of water, enhancement in emission (ϕ = 0.48) was observed due to the formation of aggregates
(see Scheme 12.4). Further, upon the addition of Hg2+ ions to the solution of 13, an increase in
quantum yield (ϕ = 0.72) was observed (see Figure 12.12a). The detection limit of aggregates of
derivative 13 for Hg2+ ions was found to be 35.9 ppb. The 1H NMR studies confirmed the interac-
tions of Hg2+ ions with the imine linkages of the assemblies. In the aggregated state, derivative 13
is emissive due to RIR. The Hg2+ ions–imine linkages coordination further rigidified these aggre-
gates 13 (see Scheme 12.4) and restricted the rotational relaxation of the excited state via nonradia-
tive modes. The SEM images of assemblies of derivative 13 in the presence of mercury ions showed
the formation of bigger aggregates, but morphological changes were not very significant (see
Figure 12.12b). To understand the structure–activity relationship with regards to metal-­assemblies
interaction in emission enhancement phenomena, we synthesized another HPB derivative 14
appended with pyrene units through imine linkages  [21]. The presence of pyrene units in the
building block 14 induced interesting morphological features in the assembled architecture. The
derivative 14 showed weak monomeric emission at 376 nm and strong excimer emission at 456 nm
(ϕ = 0.076) due to pyrene moieties in H2O : EtOH (1 : 1) solvent mixture. Upon addition of Hg2+
ions, nearly tenfold emission enhancement (ϕ = 0.74) was observed (see Figure 12.13a), and the
detection limit was found to be 4.5 nm. Moreover, morphology of derivative 14 changed from the
irregular aggregates to cross-­linked nanofibers (see Figure 12.13b). The morphological transforma-
tion is attributed to the combined effect of restriction to intramolecular motion of derivative 14 due
to interaction of Hg2+ ions with imine linkages and Hg2+ ions induced pronounced π–π intermo-
lecular stacking of pyrene units (see Scheme 12.5). Very interestingly, upon switching pyrene units
with quinolone units in HPB derivative 15, selective affinity of fluorescent assemblies towards
Zn2+ ions was observed (see Scheme  12.6)  [22]. The Zn2+ ions induced emission changes (see
Figure 12.14a) were also accompanied by modulation of the morphology change of fluorescent
aggregates changed from spherical to rods (see Figure 12.14b), and the detection limit was found
to be 14.7 nm. The time-­resolved studies showed increase in lifetime decay from 1.1 to 4.5 ns in the
presence of Zn2+ ions that indicated zinc-­induced rigidification of 15-­Zn2+ complex. To
12.3 ­Guest-­induced Emission Enhancemen 345

N N

N N
13

H2O

Scheme 12.4  Schematic representation for mercury ions induced emission enhancement in HPB
derivative 13.

Hg2+ ions

(a) (b)
600

500 Hg2+
150 μm
400
Intensity

300
1 μm 5 μm

200 0 μm
100

0
420 450 480 510 540
10 μm
Wavelength (nm)

Figure 12.12  (a) Fluorescence spectra of 13 in the presence of Hg2+ ions in H2O : THF (4 : 6) at
λex = 380 nm. (b) Change in morphology of aggregates of derivative 13 upon addition of mercury.
Source: Reprinted with permission from Ref. [20]. Copyright 2013 American Chemical Society.
346 12  Aggregation-enhanced Emission: From Flexible to Rigid Cores

N N

14

(a) (b)
800
Hg2+
700
600
20
Intensity (a.u.)

500 Hg2+
400 (equiv.)
0
300
200
100
0
350 400 450 500 550 600
Wavelength (nm)

Figure 12.13  (a) Change in fluorescence spectra of compound 14 with the addition of Hg2+ ions in
H2O : EtOH (1 : 1, v/v), λex = 342 nm. (b) SEM images of aggregates derivative 14 showing change in
morphology upon addition of mercury in H2O : THF (1 : 1). Source: Reprinted from Ref. [21]. Copyright 2013,
with permission from Elsevier.

Hg2+ ions

Scheme 12.5  Schematic presentation of the formation of fibrous aggregate of 14-­Hg2+ ensemble in
H2O : EtOH (1 : 1, v/v). Source: Reprinted from Ref. [21]. Copyright 2013, with permission from Elsevier.
12.3 ­Guest-­induced Emission Enhancemen 347

MeO OMe

N N

N N

15

THF:
water Zn2+

HPB derivative
Zinc induced emission
enhancement
Fluorescent aggregates

Scheme 12.6  Schematic representation of Zn2+ ions induced emission enhancement in HPB derivative 15.
Source: Reproduced from Ref. [22]. Copyright 2017 with permission from Royal Society of Chemistry.

(a)
1000 (b)
Zn2+

800 Zn2+ (equiv.)


12
Intensity (a.u.)

600

400 0

200
1 μm 1 μm

0
350 400 450 500 550
Wavelength (nm)

Figure 12.14  (a) Fluorescence spectra showing the change in the emission intensity of derivative 15 upon
addition of Zn2+ ions in the H2O : EtOH (6 : 4, v/v) mixture buffered with HEPES, λex = 350 nm. (b) SEM
images showing the modulation of morphology of aggregates of derivative 15 upon addition of Zn2+ ions in
H2O : EtOH (6 : 4, v/v). Source: Reproduced from Ref. [22]. Copyright 2017 with permission from Royal Society
of Chemistry.
348 12  Aggregation-enhanced Emission: From Flexible to Rigid Cores

understand the role of structural features of an initial building block in controlling the metal-­
induced “ordered” aggregation, we synthesized donor–acceptor system 16, which showed Zn2+
ions induced emission enhancement (see Figure  12.15)  [23]. The observed AEE phenomena is
outcome of two processes, first coordination of Zn2+ ions with pyridyl groups reduced the electron
density on the nitrogen atoms, and hence, ICT effect is exhibited. Second, metal-­ligand complexa-
tion enhanced the planar rigidity of the system, and resulted in emission enhancement (see
Scheme  12.7). The time-­resolved studies revealed that presence of Zn2+ ions accelerated the

N
C6H13
N
N

16
0 equiv. 45 equiv.
600

500 Zn2+ Zn2+

400 0 equiv.
Intensity

45 equiv.
300

200

100

0
347 397 447 497 547
Wavelength (nm)

Figure 12.15  Change in fluorescence spectra of 16 with the addition of Zn2+ ions in H2O : THF (8 : 2, v/v)
λex = 300 nm. Source: Reproduced from Ref. [23], Copyright 2014 with permission from Royal Society of Chemistry.

ESICT

OH O

Zn2+ OH
NH2

Zinc increses rigidity in system


Threonine is weakly interacting with Zn2+ ions but not displacing it from ensemble

Scheme 12.7  Schematic representation of Zn2+ ions and threonine induced rigidity in derivative 16.
Source: Reproduced from Ref. [23]. Copyright 2014 with permission from Royal Society of Chemistry.
12.3 ­Guest-­induced Emission Enhancemen 349

(a) (b)

600 75 equiv. 0 equiv.


Thr
480 Thr
Thr
Intensity

360
50 equiv. 50 equiv.
240
120
200 nm
0
340 380 420 460 500 540 580
Wavelength

Figure 12.16  (a) Fluorescence emission spectra of the 16-­Zn2+ ensemble with the addition of threonine in
H2O : THF (8 : 2, v/v). (b) The SEM images showing change in morphology of 16-­Zn2+ ensemble after
addition of threonine. Source: Reproduced from Ref. [23]. Copyright 2014 with permission from Royal Society
of Chemistry.

decrease in nonemissive rate constant, thus, indicating that rigidity of the system is major cause of
emission enhancement. The detection limit of derivative 16 for Zn2+ ions was found to be 110 nm.
Further, the supramolecular 16-­Zn2+ ensemble exhibited further aggregation in the presence of
threonine with high sensitivity as the detection limit was found to be 120 nm. In the presence of
threonine, emission intensity of the band centered at 490 nm was quenched, and a new emission
band appeared at 440 nm (see Figure 12.16a). The emission changes were accompanied by mor-
phology transformation of the spherical aggregates of the supramolecular 16-­Zn2+ ensemble to
cross-­linked flakes (see Figure  12.16b). The time-­resolved studies showed a threonine-­induced
increase in the value of nonradiative rate constant and a decrease in the value of radiative rate
constant of the 16-­Zn2+ ensemble. The 1H NMR studies of 16-­Zn2+ ensemble in the presence of
threonine showed an enhancement in π–π intermolecular stacking of aromatic groups. Thus, in
the presence of threonine, the interaction between 16 and Zn2+ ions is weakened and π–π intermo-
lecular stacking is enhanced. These changes resulted in the transformation of packing of mole-
cules in the ensemble, and subsequently, emission changes were observed. To further understand
the relation between packing of molecules and emission behavior, we synthesized cyclopentadien-
one derivative 17 having pyridyl groups that act as rotors as well as binding sites [24]. Interestingly,
derivative 17 formed H-­type fluorescent aggregates of spherical shape in mixed aqueous media.
The viscosity-­dependent and 1H NMR studies suggested that RIR and weakening of intermolecu-
lar π–π stacking is the reason behind emission enhancement (ϕ = 0.57). The time-­resolved studies
showed an increase in decay time from 0.43 to 1.68 ns on aggregation. Upon addition of Fe2+ ions
to the solution of aggregates of derivative 17, the intensity of band at 465 nm decreased and new
band appeared at 525 nm with low intensity (ϕ = 0.049) (see Figure 12.17a). Actually, in the pres-
ence of Fe2+ ions, a donor–acceptor system is generated, and hence, emission band is redshifted
and its intensity is decreased. In the time-­resolved studies, an increase in lifetime is observed from
0.5 to 0.93 ns. Further, a significant change in the magnitude of a nonradiative constant (knr) was
observed (from 1.95 × 109 s−1 to 1.018 × 109 s−1) in the presence of Fe2+ ions. Further, the morphol-
ogy of aggregates changed from spherical to rod shaped in the presence of Fe2+ ions (see
Figure 12.17b). All these results indicate iron-­induced modulation of aggregates from H-­type to J-­
type aggregates (see Scheme 12.8). Further, the supramolecular ensemble of iron later showed the
emission enhancement in the presence of creatinine (see Figure 12.18). In the fluorescence stud-
ies, in the presence of creatinine the band at 525 nm disappeared and a new band appeared at
475  nm. The detection limit of 17-­Fe2+ ensemble for creatinine was found to be 10  pM. In the
350 12  Aggregation-enhanced Emission: From Flexible to Rigid Cores

N 17 N

(a)
500 000 450 nm
450 000
400 000
Fe2+ ions
350 000
(150 equiv.)
Intensity

300 000 (b)


250 000 60000 Fe2+ ions
50000
40000
200 000 30000

150 000 20000


10000
0
100 000 370 420 470 520 570 620

50 000
500 nm 1 um
0
370 470 570
Wavelength (nm)

Figure 12.17  (a) Change in fluorescence spectra of derivative 17 on the addition of Fe2+ ions; the inset
showing the appearance of a new band at 525 nm. (b) TEM images of aggregates of derivative 17 in
presence of Fe2+ ions in mixed aqueous media. Source: Reprinted with permission from Ref. [24]. Copyright
2019 American Chemical Society.

Fe2+

Fe2+

Fe2+

Fe2+

Fe2+ = Fe2+ Fe2+

Fe2+

Fe2+

Fe2+

Scheme 12.8  Schematic representation of modulation of aggregates of derivative 17 from H type to J type
in presence of Fe2+ ions. Source: Reprinted with permission from Ref. [24]. Copyright 2019 American
Chemical Society.
12.3 ­Guest-­induced Emission Enhancemen 351

475 nm

500 000
Creatinine
400 000 (30 equiv.)
Intensity

300 000

200 000
525 nm

100 000

0
400 500 600
Wavelength (nm)

Figure 12.18  Fluorescence spectra of 17-­Fe2+ ensemble showing the variation of fluorescence intensity
upon addition of creatinine solution. Source: Reprinted with permission from Ref. [24]. Copyright 2019
American Chemical Society.

time-­resolved studies, an increase in the average lifetime of 17-­Fe2+ ensemble was observed in the
presence of creatinine from 0.43 to 1.042 ns, and a significant change in the magnitude knr was
observed (from 2.25 × 109 s−1in 17-­Fe2+ to 0.45 × 109 s−1). All these studies confirmed the forma-
tion of “more ordered” assemblies in the presence of creatinine.
To further understand the metal ions induced emission and energy transfer processes in donor–
acceptor systems, derivative 18 was synthesized having HPB unit as the donor and rhodamine
moieties as an acceptor [25]. Derivative 18 showed AEE characteristics (ϕ = 0.20), and the TEM
studies of derivative 18 in CH3CN  :  H2O (1  :  1, v/v) solvent mixture showed the formation of
spherical aggregates (see Figure 12.19a). Among different metal ions examined, in the presence of
Hg2+ ions, interesting coupling of AEE phenomena and energy transfer was observed. The ratio-
metric fluorescence response was observed upon addition of Hg2+ ions to the solution of aggre-
gates of derivative 18 because of AEE-­förster resonance energy transfer (FRET) phenomenon with
a detection limit of 100  nm. The spirolactam ring of rhodamine moiety was opened (see
Scheme 12.9) in the presence of Hg2+ ion, and resulted in the appearance of a new emission band
at 582 nm (see Figure 12.19b). The intensity of the emission band at 475 was gradually decreased
and intensity of the band centered at 582 nm was linearly increased (ϕ = 0.62) in the presence of
Hg2+ ions. The fluorescence lifetime studies of aggregates of derivative 18 in the absence of
Hg2+ions showed triexponential decay with lifetimes of 1.89  ns (42.97%), 7.28  ns (37.20%), and
0.34 ns (19.82%) that get shortened to 1.19 ns (51.39%), 5.94 ns (12.65%), and 0.06 ns (19.08%) in
the presence of Hg2+ ions. The decrease in lifetime on addition of Hg2+ ions indicated the fast
energy transfer from donor HPB to acceptor rhodamine moiety. Interestingly, the derivative 18 exhib-
ited only 25% mercury-­induced energy transfer in the molecular form, however, 97.14% energy
transfer was observed in the assembled state, which clearly showed the importance of AEE.
Next, we explored the potential of AEE phenomena for the development of fluorescent materi-
als, which besides showing metal-­induced morphological/emission changes also possessed suffi-
cient reduction potential to act as nanoreactors to reduce metal ions to metallic species.
352 12  Aggregation-enhanced Emission: From Flexible to Rigid Cores

N N

N N
N O O N
O O

N N
18

(a) (b)
800
700
600
Hg2+
500
Intensity

400
600 nm 300
200
100
0
400 450 500 550 600
Wavelength (nm)

Figure 12.19  (a) TEM image of derivative 18 in H2O : CH3CN (1 : 1, v/v). (b) Fluorescence spectra of
derivative 18 in the presence of Hg2+ ions in H2O : CH3CN (1 : 1, v/v); λex = 327 nm. Source: Reprinted from
Ref. [25]. Copyright 2017, with permission from Elsevier.

Ex = 327 nm Em = 475 nm
Ex = 327 nm

AIEE active
Part and FRET
Donor

“FRET” “FRET”
OFF Hg2+ ON

KI

N N N Hg2+ Hg2+ N
N N HO N
O O N OH
N N N N
FRET
O O O O
Acceptor

N N N N

Em = 582 nm

2+
Scheme 12.9  AEE active Hg ions induced FRET phenomenon in derivative 18. Source: Reprinted from
Ref. [25]. Copyright 2017, with permission from Elsevier.
12.3 ­Guest-­induced Emission Enhancemen 353

NC CN

N N

OH 19 HO

(b)
700 350 510 nm

600 Zn2+
(equiv.)
(a) 500
Intensity (a.u.)

0
400 447 nm

300

200

100
200 nm

0
400 450 500 550 600 650
Wavelength (nm)

Figure 12.20  (a) SEM image of pot shaped aggregates of derivative 19 in H2O : THF (8 : 2, v/v) mixture.
(b) Time dependent fluorescence spectra of compound 19 in the presence of Zn2+ ions (0–350 equiv.) in
H2O : THF (8 : 2, v/v) mixture; λex = 390 nm. Source: Reproduced from Ref. [26]. Copyright 2015 with
permission from Royal Society of Chemistry.

Dicyanopentafulvene (DCF) derivative 19 appended with β naphthyl units and having rotatable
C–C and C–N bonds was synthesized [26]. This derivative 19 exhibited emission enhancement and
formed pot-­shaped fluorescent aggregates in the mixed aqueous media (see Figure 12.20a). The
fluorescent aggregates of DCF derivative 19 selectively interacted with Zn2+ ions because of the
presence of imine linkages and reduced Zn2+ ions to ZnO nanoparticles (NPs) during sensing
event. The UV–Vis absorption, powder X-­ray diffraction (XRD), and transmission electron micros-
copy (TEM) studies confirmed the reduction of Zn2+ ions to ZnO NPs during the molecular recog-
nition event. The time-­dependent fluorescence studies of derivative 19 showed 6.4-­fold
enhancement in the emission intensity of the band centered at 510 nm upon addition of Zn2+ ions
within 180 minutes (see Figure 12.20b). The excitation spectra of derivative 19 showed two peaks
at 300 and 378 nm for the emission at 510 nm in the presence of Zn2+ ions. These studies indicated
that origin of the emission band at 510 nm was because of the localized surface plasmon state of
ZnO NPs. The detection limit of aggregates of derivative 19 for Zn2+ ions was found to be
95 × 10−8 M. This is the first report where the weakly fluorescent aggregates of DCF derivative 19
exhibited ZnO NPs-­induced AEE characteristics. In the time-­resolved studies, in the absence of
Zn2+ ions major fraction (56%) of molecules decayed through fast pathway (0.74 ns), and in the
presence of Zn2+ ions major fraction (95%) decayed through slower pathway (9.6 ns). Furthermore,
a significant decrease in nonradiative constant (from 3.6 × 108 to 0.9 × 108 s−1) in the presence of
Zn2+ ions was observed. All these studies suggest that the presence of Zn2+ ions further restricted
354 12  Aggregation-enhanced Emission: From Flexible to Rigid Cores

ZnO
ZnO
ZnO

Zn2+
ZnO

H2O : THF

Scheme 12.10  Formation of luminescent ZnO NPs using aggregates of derivative 19.

C12H25
O N O

HO

N
OH

O N O
C12H25
20

Fluorophores

Water

Rotors
with binding sites
Closely packed assemblies

Scheme 12.11  Schematic representation formation of self-­assemblies of PBI derivative 20 upon addition
of water. Source: Reproduced from Ref. [28]. Copyright 2015 with permission from Royal Society of
Chemistry.

the intramolecular rotation and deactivated the nonradiative decay that resulted in emission
enhancement (see Scheme 12.10).
To understand the role of rigid scaffold in designing of materials showing metal NPs-­induced
emission enhancement, perylene bisimide (PBI) derivative 20 was synthesized having β naphthyl
groups at bay position through a phenyl spacer and rotatable C–C/C–N bonds. The rigid and pla-
nar fluorescent dye PBI derivatives are known for their weak emission in aggregated states due to
ACQ. Tang et al. reported solid-­state–emitting PBI derivative by appending it with AIE active TPE
units  [27]. Unlike this, PBI derivative 20 showed AEE characteristics without having any AEE
active core (see Scheme  12.11)  [28]. The viscosity-­dependent emission studies, time-­resolved
12.3 ­Guest-­induced Emission Enhancemen 355

Figure 12.21  Fluorescence spectra of compound 500


20 showing the response to the Hg2+ ion in the Equiv.
H2O : THF (1 : 1, v/v) mixture. Source: Reproduced 400
from Ref. [28]. Copyright 2015 with permission 18
from Royal Society of Chemistry.

Intensity
300
Hg2+
200
0
100

0
500 530 560 590 620 650
Wavelength (nm)

fluorescence studies and concentration-­dependent 1H NMR studies suggested that PBI molecules
formed π–π stacked units interconnected through H-­bonding to form closely packed fluorescent
supramolecular assemblies. The close packing of PBI molecules disfavored the structural relaxa-
tion pathways in the excited state, which are known for the formation of nonfluorescent aggre-
gates. The aggregation-­driven growth and restriction to rotation is the mechanism behind the AEE
phenomenon. The fluorescent aggregates of derivative 20 have imine linkages as binding sites
exhibit “on–on” response towards the Hg2+ ions (see Figure 12.21) and detection limit was found
to be 7.15 nm. The assemblies served as nanoreactor and stabilizers for reduction of Hg2+ ions to
Hg NPs. The time-­resolved studies showed longer decay time of derivative 20 in the presence of
mercury ions, thus suggesting formation of “more ordered” aggregates.
In the search of new AEE active scaffolds showing metal NPs-­induced emission enhancement,
pyrazine-­based donor–acceptor system 21 was synthesized [29]. The pyrazine derivative 21 exhibited
emission band at 555 nm in THF corresponding to TICT state. The viscosity-­dependent fluorescence
studies of derivative 21 confirmed the formation of TICT state in the molecule. Upon introducing
water as cosolvent, the intensity of band at 555 nm decreased, and a band at 455 nm appeared due
to aggregation corresponding to LE state with low-­emission intensity (see Figure 12.22a). The addition
of water upto 99.9% did not completely restrict TICT state, and hence weakly emissive spherical
aggregates were formed (see Figure 12.22b). On addition of Cu2+ ions to solution of derivative 21
in water, the transition from LE state to TICT state is completely restricted, and emission intensity
of band at 455 nm is increased gradually (see Figure 12.23). In the presence of copper ions, struc-
tural rigidification of the system is achieved that was responsible for the emission enhancement of
supramolecular ensemble (see Scheme 12.12). During interaction process, the aggregates of deriv-
ative 21 act as reactors and stabilizers for the generation of CuO NPs, and themselves get oxidized
to form polyamine 22. Interestingly, the oxidized species 22 in combination with copper oxide NPs
served as light-­harvesting antenna and exhibited excellent photocatalytic efficiency in Sonogashira
coupling under mild and ecofriendly conditions (room temperature, aqueous media, aerial condi-
tion, and visible light irradiation). Building on same strategy [30], a variety of AEE active materials
were developed based on HPB, oligophenylene, pyrazine scaffold to prepare metal NPs [31–37],
metal nanoclusters [38, 39], core shell NPs [40, 41], and nano hybrid composite [42–44]. Further,
similar strategy was adapted to develop AEE active photoredox assemblies [45].
The polyamine 22 also exhibited AEE characteristics upon addition of water fraction (see
Figure 12.24a) [46]. The polyamine 22 undergoes solvent-­dependent living self-­assembly behavior
accompanied with morphology transformation. The TEM image of derivative 22 in THF shows the
presence of a fibrous network (see Figure 12.24b). Upon increasing the volume of water fractions
in the THF solution, the morphology and the size of the aggregates changed. The TEM image of
356 12  Aggregation-enhanced Emission: From Flexible to Rigid Cores

N N

H2N NH2
21

(a) (b)
1000
0% water
400
60% water 60
800 350 99.5% water
70
300
Intensity (a.u.)

80
Intensity (a.u.)

600 250 90
200 99.5
400
150
50% water
100
200
50
0 0
400 450 500 550 600 650 400 450 500 550 600 650
Wavelength (nm) Wavelength (nm)

Figure 12.22  Fluorescence spectra of compound 21 (5.0 μM) in different water/THF fraction (a) 0–50%
and (b) 60–99.5% (slit width 3 : 5); λex = 360 nm. Source: Reprinted with permission from Ref. [29]. Copyright
2016 American Chemical Society.

1000 445 nm

50
800
Cu2+
(equiv.)
Intensity (a.u.)

600 0
50
400 Cu2+
(equiv.)
555 nm
0
200

0
400 450 500 550 600 650
Wavelength (nm)

Figure 12.23  Florescence spectra of derivative 21 (5.0 M) show ratiometric response upon addition of
Cu2+ ions (0–50 equiv.) in HEPES buffer with (Ph = 7.05, λex = 358 nm). Source: Reprinted with permission
from Ref. [29]. Copyright 2016 American Chemical Society.
12.3 ­Guest-­induced Emission Enhancemen 357

NH2
NH2
Cu2+

NH2 N NH2 NH2


N NH2 N
N NH2
N Cu2+
Water N N Cu2+
N N 10 min N

N NH2
N NH2
N NH2
NH2 N NH2
Cu2+

NH2

NH2
TICT active 21 Random Aggregates Complexation of
of 21(TICT-AEE) aggregates of 21

NH2

H
N N CuO
H n
N 60 min
N N NH2
N CuO
N CuO NH2 NH2
N H
NH2 N N
H N CuO
N N n
N NH2
N CuO
CuO
N NH2
NH2 H NH2
H N
N N
N n
N N CuO
N
N CuO CuO
NH2
NH2
NH2

Supramolecular ensemble 22:CuO Nps

Scheme 12.12  Probable schematic representation of CuO NPs formation by using aggregates of derivative
21 in aqueous media and resulting in the formation of supramolecular ensemble 22:CuO NPs.
Source: Reprinted with permission from Ref. [29]. Copyright 2016 American Chemical Society.

the H2O : THF (1 : 1) solution of derivative 22 shows the existence of worm-­like nanorods. Upon
increasing the water fraction upto 90%, the degree of association further increased, and the TEM
image clearly showed the coalescence of nanorods to form closely packed globular aggregates (see
Figure 12.24c). The aggregated polyamine 22 has a number of acidic NH groups and shows “turn
on” response in the presence of fluoride ions via hydrogen bonding (see Figure  12.25a) with a
detection limit of 40 ppb. The time-­resolved fluorescence studies indicate a decrease in nonradia-
tive rate constant (from knr = 3.95 × 109 s−1 to knr = 0.58 × 109 s−1) in the presence of fluoride. All
these studies suggest that in the presence of fluoride ions, supramolecular aggregates of derivative
22 are more rigidified and fluoride-­induced restriction to rotation is the reason behind the emis-
sion enhancement of polyamine derivative 22 in aqueous media. Further, the TEM image of solu-
tion of derivative 22 in the presence of fluoride ions showed the presence of aggregates having
crystal-­like morphology with well-­distributed fluoride ions deposited on the surface of aggregates
(see Figure 12.25b) which further support the formation of “more ordered” aggregates in the pres-
ence of fluoride ions (see Scheme 12.13).
358 12  Aggregation-enhanced Emission: From Flexible to Rigid Cores

NH2

H
N N
H
N N
N

N
N
NH2
N
NH2
n
NH2 22

(a)
80%
600 90
Water fraction 80
500 70
0% 60
Intensity (a.u.)

400 50
40
300 30
20
200
10
100 0

0
400 450 500 550 600 650
Wavelength (nm)
(b) (c)

Figure 12.24  (a) Fluorescence spectra of derivative 22 in different fractions of H2O : THF (λex = 360 nm).


TEM images of derivative 22, (b) in the presence of THF show fibrous network, and (c) in the presence of
9 : 1 H2O : THF fraction show globular type structure. Source: Reprinted from Ref. [46]. Copyright 2018, with
permission from Elsevier.

To understand the influence of anions on aggregation behavior of HPB derivatives, derivative 23


appended with β naphthyl units were synthesized. Derivative 23 exhibited AEE characteristics in
the mixed aqueous media and formed irregular-­shaped aggregates as confirmed by TEM studies
(see Scheme  12.14)  [47]. Among various anions examined, the fluorescent aggregates showed
12.3 ­Guest-­induced Emission Enhancemen 359

(a) (b)
50 ppm
800
700 F–

600 0 ppm
Intensity (a.u.)

500
400
300
200
100
0
370 470 570
Wavelength (nm)

Figure 12.25  (a) Fluorescence spectra of derivative 22 upon addition of F− ions in 9 : 1 (H2O : THF)
buffered with HEPES (pH = 7.05 and λex = 360 nm). (b) TEM image shows deposition of fluoride ions
deposited on the surface of supramolecular aggregates of derivative 22. Source: Reprinted from Ref. [46].
Copyright 2018, with permission from Elsevier.

F–

ϕ = 0.02 ϕ = 0.4

Scheme 12.13  Formation of ordered and fluorescent aggregates of derivative 22 in the presence of
fluoride. Source: Reprinted from Ref. [46]. Copyright 2018, with permission from Elsevier.

CN− ions-­induced emission enhancement (see Figure 12.26) and detection limit of these aggre-
gates for CN− ions was found to be 55 nm. The cyanide ions interacted with hydroxyl groups of the
aggregates of derivative 23 and resulted in the formation of more rigid fluorescent assemblies.
Interestingly, cyanide-­induced deprotonation of the hydroxyl groups was not observed. These
results indicate that because of the cyanide-­controlled supramolecular architecture, the rotational
relaxation of the excited state via nonradiative modes is restricted, hence, emission enhancement
was observed (see Scheme 12.14). The cyanide-­induced emission enhancement was also accompa-
nied by modulation of irregular aggregates of derivative 23 to generate rod-­shaped assemblies as
suggested by TEM studies (see Scheme 12.14). Using the same strategy, cyclopentadienone-­based
assemblies having β naphthyl units which exhibited cyanide-­induced emission enhancement and
generated porous supramolecular anionic assemblies were developed [48].
360 12  Aggregation-enhanced Emission: From Flexible to Rigid Cores

N N

OH 23
HO

ϕ = 0.13 ϕ = 0.57

(a) (b)
= H-bond
CN–
Supramolecular ensemble

Spherical aggregates Modulated nanorods

Scheme 12.14  The schematic presentation of 23-­CN− adduct showing the formation rods shaped
aggregates along with SEM images (a) aggregates of derivative 23 and (b) 23-­CN− adduct.
Source: Reproduced from Ref. [47]. Copyright 2013 with permission from Royal Society of Chemistry.

CN– (equiv.) Figure 12.26  Change in fluorescence spectra of 23 with the


1000 CN– addition of CN− ions in H2O : EtOH (6 : 4, v/v) buffered with
200 HEPES, pH = 7.0; λex = 322 nm. Inset showing the fluorescence
800
intensity; (i) before and (ii) after the addition of CN− ions.
600 Source: Reproduced from Ref. [47]. Copyright 2013 with
Intensity

i ii permission from Royal Society of Chemistry.


400 0

200

0
350 400 450 500 550 600
Wavelength (nm)

In an effort to study AEE phenomena in assemblies that show analyte-­induced structural


changes, HPB derivative 24 having amide linkages were synthesized [49]. The HPB derivative 24
showed AEE characteristics as evidenced from viscosity-­dependent fluorescent studies and
concentration-­dependent 1H NMR studies. The concentration-­dependent 1H NMR showed upfield
12.3 ­Guest-­induced Emission Enhancemen 361

500 380 nm 0

CN–
400 (equiv.)

Intensity (a.u.)
300 250 250

CN–
200 485 nm (equiv.)
0
100
HN NH
O O 0
F3C CF3 350 400 450 500 550
24 Wavelength (nm)

Figure 12.27  Fluorescence spectra of compound 24 showing the response to the CN− ion (0−250 equiv.) in
H2O : EtOH (8 : 2, v/v) mixture, λex = 282 nm. Source: Reprinted with permission from Ref. [49]. Copyright
2014 American Chemical Society.

380 nm
(b)
282 nm 485 nm
O

N CF3 282 nm
H NC
O N CF3
O
H O–
N N CF3 1 μm
F3C H O–
H H
O N CF3 NC
O CN N CF3
N –
N CF3 H O
F3C
H H
O
CN– NC
H O–
O N CF3 N CF3
N CF3 O– CN NC
N H N
F3C CF3
H H H O–
O H O–
O N CF3
N NC H O–
N CF3 NC N CF
F3C N
NC 3
(a) H H
O
H
CF3
O–
O
H
N N CF3 O–
F3C H N CF3
H NC
O
N
F3C H
1 μm

Scheme 12.15  The probable schematic presentation of 24-­CN− adduct suggesting the formation spherical
nanoaggregates along with SEM images (a) aggregates of derivative 24 and (b) 24-­CN− adduct. Source:
Reprinted with permission from Ref. [49]. Copyright 2014 American Chemical Society.

shift for aromatic protons and downfield shift for amide protons. These studies suggested that the
π–π stacking between the molecules of derivative 24 and amide linkages were involved in
H-­bonding which resulted in formation of large-­sized weakly fluorescent aggregates. Interestingly,
on addition of CN− ions to solution of derivative 24, the intensity of band centered at 385 nm was
decreased and new band appeared at 485  nm (see Figure  12.27). Various experimental studies
indicate that nucleophile addition of cyanide ions on amide linkages of aggregates of derivative 24
leads to generation of new species 24-­CN−, which showed an emission band at 485 nm. The detec-
tion limit of the aggregates of derivative 24 for CN− ions was found to be 11 nm. The SEM image
of the 24-­CN− adduct in H2O : EtOH (8 : 2) showed the disruption of closely packed aggregates to
form dispersed nanoaggregates (see Scheme 12.15). Upon addition of CN− ions to the aggregates
362 12  Aggregation-enhanced Emission: From Flexible to Rigid Cores

N3 25 N3
SNa
– N
N+ N N H
Hb
N NH N Ha NH2
SNa
Hb
Na2S –N2 NaHS Ha

H2O Ha
Hb
SNa
N NH N Ha NH2
N+
N– N H Hb
N
SNa

Scheme 12.16  Probable mechanistic pathway of reduction of probe 25 in presence of


Na2S. Source: Reproduced from Ref. [50]. Copyright 2015 with permission from Royal Society of Chemistry.

of derivative 24 the -­NH hydrogens are no longer free for intermolecular H-­bonding, and hence
closely packed architecture is collapsed to form smaller aggregates (see Scheme  12.15). The
cyanide-­induced restriction to long ordered π–π intramolecular stacking resulted in the formation
of highly fluorescent nanoaggregates.
Building on the strategy of examining aggregation behavior in the molecule having a reaction
center, HPB derivative 25 having azide units as a reaction center at the periphery were synthe-
sized [50]. HPB derivative 25 showed weak emission in mixed aqueous media because of the pres-
ence of azido groups at the periphery of the aggregates that increases the solubility of derivative 25
in mixed aqueous media, and thus most of the molecules were present in nonaggregated form.
Selectively, in the presence of H2S (in situ generated from Na2S), azido groups are reduced to amino
groups that have less solubility in aqueous media, hence, major fraction of the molecules was pre-
sent in the aggregated form (see Scheme 12.16). The formation of the aggregates restricts the rota-
tions of the molecules and hence, induces the planarization that made the aggregates more rigid
and emissive in mixed aqueous media (see Figure 12.28a–c). The detection limit of the aggregates
of 25 for H2S was found to be 0.33 μM. The 1H NMR studies, time-­resolved studies confirmed that
restriction to rotation is the mechanism of the AEE phenomena. All the change in the emission
behavior was also accompanied by morphological transformation as is clear from SEM images (see
Scheme 12.17). Using the reaction-­based strategy, heterooligomer derivative 26 having four pyridyl
rings as rotors were synthesized  [51]. The derivative 26 formed fluorescent J-­aggregates (see
Figure 12.29a) which were confirmed by temperature-­dependent absorption studies, fluorescence
studies, and concentration-­dependent 1H NMR studies. The SEM image showed the formation of
irregular-­shaped aggregates in the mixed aqueous media (see Figure  12.29b). The fluorescent
aggregates of derivative 26 showed “no quenching” response towards picric acid (PA) because of
the combined aggregation-­induced emission enhancement and excited state intermolecular pro-
ton transfer phenomena. This happens to be the first report of fluorescent materials showing
12.3 ­Guest-­induced Emission Enhancemen 363

Figure 12.28  (a) Time dependent fluorescence (a)


spectra of 25, after incubation with 400 μM H2S in
H2O : DMSO (7 : 3, v/v) mixtures, λex = 300 nm; inset 1000
showing the change in fluorescence (i) before and 400 µM
(ii) after the addition of 400 μM H2S under 365 nm UV 800 16 minutes H2S
light; confocal images showing the change in 400

Intensity (a.u.)
fluorescence intensity of supramolecular aggregates 600 µM
of 25 (b) before and (c) after the addition of H2S,
0
λex = 405 nm in confocal microscope. Source:
400 (i) (ii)
Reproduced from Ref. [50]. Copyright 2015 with
permission from Royal Society of Chemistry.
200

0
350 400 450 500 550 600
Wavelength (nm)
(b) (c)

1 μm

NHNH
1 μm
H2N 2 2
NH2
NH2
N N N
N N N N NH2
N N N N N

Self- N
N
N

assembled N
N
N

into micelles 200 µM 400 µM H2S


50 µM H2S H2S
Micro-rods
Micelles aggregated
Micelles
formed to flower
starts
columnar like
N3 aggregation
aggregates assembly
DMSO : H2O Rod like self- Flower like
Azide groups are
= (7:3, v/v)
reduced to amino
assembly self-assembly
groups
N3

Scheme 12.17  Probable schematic diagram shows H2S sensing mechanism with tuneable self-­assembly
and AEE behavior. Source: Reproduced from Ref. [50]. Copyright 2015 with permission from Royal Society of
Chemistry.

“on–on” response towards PA. The mixed aqueous solution of derivative 26 showed an emission
band at 329 nm due to AEE effect, upon addition of PA intensity of this band is decreased and a
band appeared at 446 nm. The aggregates of derivative 26 interacted with PA and quenched the
emission intensity of the band centered at 339 nm. Eventually, upon light absorption, the basicity
of pyridyl nitrogen is enhanced, which leads to excited state intermolecular proton transfer from
PA to pyridyl nitrogen in the excited state and results in the formation of protonated species, which
emits at 460 nm (see Scheme 12.18). The detection limit of the fluorescent aggregates of derivative
26 for PA was found to be 26  nm. We also prepared terphenyl derivative 27 which exhibited
364 12  Aggregation-enhanced Emission: From Flexible to Rigid Cores

(a) (b)
1.5
N N
1.2
EtOH

Absorbance
0.9 60% H2O in EtOH

0.6
N N
0.3
26 5 μm
0
200 250 300 350 400
Wavelength (nm)

Figure 12.29  (a) UV-­vis absorption spectra of 26 in EtOH and H2O : EtOH (6 : 4). (b) SEM image of
aggregates of derivative 26 in H2O : EtOH (6 : 4). Source: Reproduced from Ref. [51]. Copyright 2014 with
permission from Royal Society of Chemistry.

Picric acid
Picric acid
Excited state intermolecular
proton transfer

λem = 339 nm
λem = 460 nm
J-aggregate

Scheme 12.18  Mechanism of the recognition behavior of aggregates of 26 towards picric acid in
H2O : EtOH (6 : 4). Source: Reproduced from Ref. [51]. Copyright 2014 with permission from Royal Society of
Chemistry.

spermidine-­induced emission enhancement (see Scheme 12.19). Derivative 27 having formyl phe-


nyl groups at periphery formed weakly emissive aggregates in mixed aqueous media [52]. In the
presence of spermidine, Schiff base adduct is generated through interaction between formyl phe-
nyl groups of derivative 27 and amino groups of amine. The assemblies of as prepared Schiff base
28 showed AEE characteristics in mixed aqueous media due to restriction of rotation (see
Scheme 12.19). The detection limit of derivative 27 for spermidine was found to be 46 nm.
Next, we synthesized D–A–D system 29 [53] in which fumaronitrile moiety acts as an acceptor
and HPB core having methoxy groups at its periphery as donors. The HPB derivative 29 exhibited
ICT state in its isolated form. However, upon addition of water to the THF solution of derivative
29, the decrease in emission intensity of band centered at 455 nm was observed due to restriction
of ICT in aqueous media and the formation of a new band at 550 nm was due to aggregation of
derivative 29. Upon addition of water fraction up to 80%, the intensity of emission band at 550 nm
increases gradually (ϕ80% = 0.65), whereas the emission band at 455 nm disappeared completely
(see Figure 12.30a). These studies support that the inhibition of ICT leads to generation of AEE
characteristics upon addition of water. The SEM studies also showed the formation of porous
aggregates (see Figure 12.30b). The derivative 29 formed porous spherical nano-­assemblies in the
mixed aqueous media through various noncovalent interactions such as aromatic π−π stacking,
12.3 ­Guest-­induced Emission Enhancemen 365

OHC
O O O CHO
OHC O O O
CHO
27 H2O

= Benzaldehyde group
Weakly emissive
random aggregates
= Spermidine
Strongly emissive
nanorod aggregates

N O N
O O
HN
O O NH
N O
C
CH H N

28

Scheme 12.19  Schematic representation of aggregation behavior of probe 27 in the presence of


spermidine in different solvent media. Source: Reprinted from Ref. [52]. Copyright 2018, with permission
from Elsevier.

OMe

MeO

(a) H2O fraction (%)


1000 0% 30% 80%
NC AEE
CN
80
800 (b)
0
Intensity (a.u.)

600 40

400 30

ICT
OMe 200

0 100 nm
OMe 400 450 500 550 600 650
29 Wavelength (nm)

Figure 12.30  (a) Fluorescence spectra showing the change in emission intensity of compound 29 in
different H2O : THF mixture, λex = 372 nm. Filled and dashed lines correspond to ICT and AEE states. Inset
showing the photographs with different water fractions under 365 nm UV lamp illumination. (b) SEM image
of derivative 29 in H2O : THF (8 : 2, v/v) mixture. Source: Reprinted with permission from Ref. [53]. Copyright
2018 American Chemical Society.

hydrogen bonding, and hydrophobic interactions. Moreover, solvent evaporation also generated
porous spherical assemblies as indicated by TEM images. Further, fluorescent porous aggregates of
derivative 29 were used to detect triethylamine (TEA) (see Figure 12.31a) via hydrogen bonding
between them leads to deaggregation (see Figure  12.31b). On the contrary, “turn off” response
366 12  Aggregation-enhanced Emission: From Flexible to Rigid Cores

(a)
Et3N (equiv.)
1000 26 550 nm Et3N
450 nm
0
800
0 (i) (ii )
Intensity (a.u.)

600 26

400

200

0
400 450 500 550 600 650
Wavelength (nm)

(b)

TEA
500 nm 500 nm

Figure 12.31  (a) Fluorescence spectra of derivative 29 in addition of aliphatic amines such as


triethylamine (Et3N) in in H2O : EtOH (8 : 2, v/v) buffered with HEPES, pH = 7.05, λex = 372 nm. Inset
showing the photographs of (i) before and (ii) after addition of triethylamine solution to derivative 29
H2O : EtOH (8 : 2) under 365 nm UV lamp illumination. (b) TEM images showing deaggregation of
aggregates of derivative 29 upon addition of TEA. Source: Reprinted with permission from Ref. [53].
Copyright 2018 American Chemical Society.

(see Figure 12.32a) was observed towards aniline because of the formation of bigger aggregates
(see Figure 12.32b). The detection limits of aggregates of derivative 29 for TEA and aniline were
found to be 19.9 and 1.85 nm, respectively.

12.4 ­Conclusion

To summarize, we have discussed a variety of approaches for the synthesis of AEE active building
blocks based on flexible/rigid core. Using these approaches such as increasing the number of
rotors, restriction to π–π intramolecular stacking of aromatic groups, guest-­induced “ordered”
aggregation etc. AEE characteristics are successfully introduced in a variety of molecules.
Restriction to motions, aggregation-­driven growth, nanoaggregation, and J-­aggregation are among
the different mechanisms behind the observed AEE phenomena in these synthesized molecules.
In addition, AEE in combination with ICT/TICT/ESIPT phenomena have also been explored
for  introducing interesting photophysical changes in the molecules upon aggregation. The
  ­Reference 367

(a)
1000 Aniline (equiv.)

800 0
Intensity (a.u.)

600
6
400

200

0
450 500 550 600 650
Wavelength (nm)

(b)

Aniline

500nm
500nm

Figure 12.32  (a) Fluorescence spectra of derivative 29 in addition of aromatic amines such as aniline
(C6H5NH2) in H2O : EtOH (8 : 2, v/v) buffered with HEPES, pH = 7.05, λex = 372 nm. (b) TEM images showing
formation of bigger sized aggregates of derivative 29 upon addition of aniline. Source: Reprinted with
permission from Ref. [53]. Copyright 2018 American Chemical Society.

mechanistic approaches were being presented in this chapter to achieve AEE characteristics that
will help different researchers working in the diverse fields such as material chemistry, supramo-
lecular chemistry, catalysis, and in designing the synthesis of solid emitters.

­Acknowledgment

M. K. and V. B. are thankful to all the coauthors who have contributed to this work. We are thank-
ful to UGC, CSIR, DRDO, DST, and SERB for the financial assistance. We are also thankful to Guru
Nanak Dev University for research facilities.

­References

1 Birk, J. B. (1970). Photophysics of Aromatic Molecules, John Wiley & Sons, Ltd, London.
Qin, A. and Tang, B. Z. ed. (2013). Aggregation-­Induced Emission: Fundamentals and Applications,
2
John Wiley & Sons, Ltd, London.
368 12  Aggregation-enhanced Emission: From Flexible to Rigid Cores

3 Xie, J. Z., Lam, J. W. Y., Cheng, L. et al. (2001). Aggregation-­induced emission of 1-­methyl-­1,2,3,4,5-­
pentaphenylsilole. Chemical Communications 1740–1741.
4 An, B.-­K., Kwon, S.-­K., Jung, S.-­D. et al. (2002). Enhanced emission and its switching in
fluorescent organic nanoparticles. Journal of the American Chemical Society (124):
14410–14415.
5 Dong, Y., Lam, J. W. Y., Qin, A. et al. (2007). Endowing hexaphenylsilole with chemical sensory
and biological probing properties by attaching amino pendants to the silole core. Chemical Physics
Letters (44): 124–127.
6 La, D. D., Bhosale, S. V., Jones, L. A. et al. (2018). Tetraphenylethylene-­based AIE-­active probes for
sensing applications. ACS Applied Materials & Interfaces (10): 12189–12216.
7 Kwok, R. T. K., Leung, C. W. T., Lam, J. W. Y. et al. (2015). Biosensing by luminogens with
aggregation-­induced emission characteristics. Chemical Society Reviews (44): 4228–4238.
8 Mei, J., Leung, N. L. C., Kwok, R. T. K. et al. (2015). Aggregation-­induced emission: together we
shine, united we soar! Chemical Reviews (115): 11718–11940.
9 Bhalla, V., Vij, V., Dhir, A. et al. (2012). Hetero-­oligophenylene-­based AIEE material as a multiple
probe for biomolecules and metal ions to construct logic circuits: application in bioelectronics and
chemionics. Chemistry—­A European Journal (18): 3765–3772.
10 Kumar, M., Vij, V., and Bhalla, V. (2012). Vapor-­phase detection of trinitrotoluene by AIEE-­active
hetero-­oligophenylene-­based carbazole derivatives. Langmuir (28): 12417–12421.
11 Arora, H., Pramanik, S., Kumar, M. et al. (2016). “Not quenched” aggregates of a triphenylene
derivative for the sensitive detection of trinitrotoluene in aqueous medium. New Journal of
Chemistry (40): 3187–3193.
12 Vij, V., Bhalla, V., and Kumar, M. (2013). Attogram detection of picric acid by hexa-­peri-­
hexabenzocoronene-­based chemosensors by controlled aggregation-­induced emission
enhancement. ACS Applied Materials & Interfaces (5): 5373–5380.
13 Zhang, Z., Huang, H., Yang, X. et al. (2018). Tailoring electronic properties of graphene by π–π
stacking with aromatic molecules. Journal of Physical Chemistry Letter (2): 2897–2905.
14 Pérez, E. M. and Martín, N. (2015). π−π Interactions in carbon nanostructures. Chemical Society
Reviews (44): 6425–6433.
15 Bloom, J. W. G. and Wheeler, S. E. (2011). Taking the aromaticity out of aromatic interactions.
Angewandte Chemie International Edition (50): 7847–7849.
16 Karabıyık, H., Sevinçek, R., Karabıyık, H. (2014). π-­Cooperativity effect on the base stacking
interactions in DNA: is there a novel stabilization factor coupled with base pairing H-­bonds.
Physical Chemistry Chemical Physics (16): 15527–15538.
17 Zhao, R. and Zhang, R.-­Q. (2016). New insight into π–π stacking involving remarkable orbital
interactions. Physical Chemistry Chemical Physics (18): 25452–25457.
18 Bhalla, V., Gupta, A., and Kumar, M. (2012). Fluorescent nanoaggregates of pentacenequinone
derivative for selective sensing of picric acid in aqueous media. Organic Letters (14):
3112–3115.
19 Kaur, S., Gupta, A., Bhalla, V. et al. (2014). Pentacenequinone derivatives: aggregation-­induced
emission enhancement, mechanism and fluorescent aggregates for superamplified detection of
nitroaromatic explosives. Journal of Materials Chemistry C (2): 7356–7363.
20 Bhalla, V., Kaur, S., Vij, V. et al. (2013). Mercury-­modulated supramolecular assembly of a
hexaphenylbenzene derivative for selective detection of picric acid. Inorganic Chemistry (52): 4860–4865.
21 Pramanik, S., Bhalla, V., and Kumar, M. (2013). Mercury assisted fluorescent supramolecular
assembly of hexaphenylbenzene derivative for femtogram detection of picric acid. Analytica
Chimica Acta (793): 99–106.
  ­Reference 369

22 Pramanik, S., Bhalla, V., and Kumar, M. (2017). Hexaphenylbenzene-­based fluorescent aggregates
for detection of zinc and pyrophosphate ions in aqueous media: tunable self-­assembly behaviour
and construction of a logic device. New Journal of Chemistry (41): 4806–4813.
23 Kaur, S., Bhalla, V., and Kumar, M. (2014). Fluorescent supramolecular metal assemblies as ‘no
quenching’ probes for detection of threonine in the nanomolar range. Chemical Communications
(50): 9725–9728.
24 Kaur, K., Kataria, M., Kaur, M. et al. (2019). AIEE active supramolecular metallic ensemble as
“turn-­On” platform for detection of creatinine in human blood serum at picomolar level. ACS
Sustainable Chemistry & Engineering (17): 14829–14833.
25 Singh, G., Reja, S. I., Bhalla, V. et al. (2017). Hexaphenylbenzene appended AIEE active FRET
based fluorescentprobe for selective imaging of Hg2+ ions in MCF-­7 cell lines. Sensors and
Actuators B: Chemical (249): 311–320.
26 Kataria, M., Pramanik, S., Kumar, M. et al. (2015). One-­pot multicomponent synthesis of
tetrahydropyridines promoted by luminescent ZnO nanoparticles supported by the aggregates of
6,6-­dicyanopentafulvene. Chemical Communications (51): 1483–1486.
27 Wang, Y. J., Li, Z., Tong, J. et al. (2015). The fluorescence properties and aggregation behavior of
tetraphenylethene–perylenebisimide dyads. Journal of Materials Chemistry C (3): 3559–3568.
28 Kaur, S., Kumar, M., and Bhalla, V. (2015). AIEE active perylene bisimide supported mercury
nanoparticles for synthesis of amides via aldoximes/ketoximes rearrangement. Chemical
Communications (51): 4085–4088.
29 Deol, H., Pramanik, S., Kumar, M. et al. (2016). Supramolecular ensemble of a TICT-­AIEE active
pyrazine derivative and CuO NPs: a potential photocatalytic system for Sonogashira couplings.
ACS Catalysis (6): 3771–3783.
30 Deol, H., Singh, G., Kumar, M. et al. (2019). Metal nanoparticles embedded in AIEE active
supramolecular assemblies: robust, green and reusable nanocatalysts. Dalton Transactions (48):
4769–4773.
31 Pramanik, S., Bhalla, V., and Kumar, M. (2014). A hexaphenylbenzene based AIEE active probe for
the preparation of ferromagnetic α-­Fe2O3 nanoparticles: facile synthesis and catalytic applications.
Chemical Communications (50): 13533–13536.
32 Sharma, K., Kaur, S., Bhalla, V. et al. (2014). Pentacenequinone derivatives for preparation of gold
nanoparticles: facile synthesis and catalytic application. Journal of Materials Chemistry A (2):
8369–8375.
33 Sharma, K., Kumar, M., and Bhalla, V. (2017). Supramolecular ensemble of aggregates of
pentacenequinone derivative and cadmium NPs: a potential catalytic/photocatalytic system for
direct C-­H activation of quinones and azoles. Chemistry Select (2): 680–687.
34 Kataria, M., Kumar, M., and Bhalla, V. (2017). Supramolecular ensemble of
tetraphenylcyclopentadienone derivative and HgO nanoparticles: a one-­pot approach for the
synthesis of quinoline and quinolone derivatives. Chemistry Select (2): 3018–3027.
35 Kaur, S., Kumar, M., and Bhalla, V. (2016). Supramolecular ensemble of PBI derivative and copper
nanoparticles: a light harvesting antenna for photocatalytic C(sp2)–H functionalization. Green
chemistry (18): 5870–5883.
36 Kataria, M., Pramanik, S., Kumar, M. et al. (2016). Ferromagnetic α-­Fe2O3 NPs: a potential catalyst
in Sonogashira Hagihara cross coupling and Hetero-­Diels-­Alder reactions. Green chemistry (18):
1495–1505.
37 Pramanik, S., Bhalla, V., and Kumar, M. (2014). Hexaphenylbenzene based AIEE active probe for
the preparation of ferromagnetic α-­Fe2O3 nanoparticles: facile synthesis and catalytic applications.
Chemical Communications (50): 13533–13536.
370 12  Aggregation-enhanced Emission: From Flexible to Rigid Cores

38 Pramanik, S., Bhalla, V., and Kumar, M. (2015). Hexaphenylbenzene-­stabilized luminescent silver
nanoclusters: a potential catalytic system for the cycloaddition of terminal alkynes with
isocyanides. ACS Applied Materials & Interfaces (7): 22786–22795.
39 Chopra, R., Kumar, M., and Bhalla, V. (2018). Fabrication of polythiophene-­supported Ag@Fe3O4
nanoclusters and their utilization as photocatalyst in dehydrogenative coupling reactions. ACS
Sustainable Chemistry & Engineering (6): 7412–7419.
40 Chopra, R., Kumar, M., and Bhalla, V. (2016). Development of a supramolecular ensemble of an
AIEE active hexaphenylbenzene derivative and Ag@Cu2O core–shell NPs: an efficient
photocatalytic system for C–H activation. Chemical Communications (52): 10179–10182.
41 Sharma, K., Kumar, M., and Bhalla, V. (2015). Aggregates of the pentacenequinone derivative as
reactors for the preparation of Ag@Cu2O core–shell NPs: an active photocatalyst for Suzuki and
Suzuki type coupling reactions. Chemical Communications (51): 12529–12532.
42 Singh, G., Kumar, M., and Bhalla, V. (2018). Ultrafine hybrid Cu2O–Fe2O3 nanoparticles stabilized
by hexaphenylbenzene-­based supramolecular assemblies: a photocatalytic system for the
Ullmann–Goldberg coupling reaction. Green chemistry (20): 5346–5357.
43 Kaur, L., Deol, H., Kumar, M. et al. (2020). Integrating CuO−Fe2O3 nanocomposites and
supramolecular assemblies of phenazine for visible-­light photoredox catalysis. Chemistry: An Asian
Journal (15): 892–898.
44 Kaur, H., Kaur, M., Walia, P. K. et al. (2018). Encapsulating Au-­Fe3O4 nanodots into AIE active
supramolecular assemblies: ambient visible light harvesting “dip-­strip” photocatalyst for C–C/C–N
bond formation reactions. Chemistry: An Asian Journal (14): 809–813.
45 Dadwal, S., Deol, H., Kumar, M. et al. (2019). AIEE active nanoassemblies of pyrazine based
organic photosensitizers as efficient metal-­free supramolecular photoredox catalytic systems.
Scientific Reports (9):11142.
46 Deol, H., Kumar, M., and Bhalla, V. (2018). Smart fluorescent AIEE active supramolecular
polymeric assemblies for selective and quantitative detection of fluoride ions in aqueous media.
Sensors and Actuators B: Chemical (258): 682–693.
47 Bhalla, V., Pramanik, S., and Kumar, M. (2013). Cyanide modulated fluorescent supramolecular
assembly of a hexaphenylbenzene derivative for detection of trinitrotoluene at the attogram level.
Chemical Communications (49): 895–897.
48 Kataria, M., Singh, Z., Kumar, M. et al. (2019). ‘Metal free’ fluorescent anionic assemblies: efficient
platform for detection of electron deficient nitroaromatics and cationic dyes. Dyes and Pigments
(171): 107601.
49 Pramanik, S., Bhalla, V., and Kumar, M. (2014). Hexaphenylbenzene-­based fluorescent aggregates
for ratiometric detection of cyanide ions at nanomolar level: set-­reset memorized sequential logic
device. ACS Applied Materials & Interfaces (6): 5930–5939.
50 Pramanik, S., Bhalla, V., Kim, H. M. et al. (2015). A hexaphenylbenzene based AIEE active two
photon probe for the detection of hydrogen sulfide with tunable self-­assembly in aqueous media
and application in live cell imaging. Chemical Communications (51): 15570–15573.
51 Kaur, S., Bhalla, V., Vij, V. et al. (2014). Fluorescent aggregates of hetero-­oligophenylene derivative
as “no quenching” probe for detection of picric acid at femtogram level. Journal of Materials
Chemistry C (2): 3936–3941.
52 Tejpal, R., Kumar, M., and Bhalla, V. (2018). Spermidine induced aggregation of terphenyl
derivative: an efficient probe for detection of spermidine in living cells. Sensors and Actuators B:
Chemical (258): 841–849.
53 Pramanik, S., Deol, H., Bhalla, V. et al. (2018). AIEE active donor–acceptor–donor-­based
hexaphenylbenzene probe for recognition of aliphatic and aromatic amines. ACS Applied Materials
& Interfaces (10): 12112–12123.
371

13

Room-­temperature Phosphorescence of Pure Organics


Tianwen Zhu, Zihao Zhao, Tianjia Yang, and Wang Zhang Yuan
School of Chemistry and Chemical Engineering, Shanghai Jiao Tong University, Shanghai, China

13.1 ­Introduction

Recently, materials with room-­temperature phosphorescence (RTP) have drawn increasing atten-
tion due to their extended applications in optoelectronic and bioelectronics, such as organic light-­
emitting diodes (OLEDs) [1, 2], information security [3, 4], sensors [5–7] and bioimaging [8, 9]. Up
to date, the reported RTP materials are mainly inorganic and organometallic compounds contain-
ing rare and expensive elements, such as platinum (Pt)  [10], iridium (Ir)  [11], and rhenium
(Re) [12], which brings costly issues and thus hampers their further large-­scale applications. By
contrast, although pure organic materials exhibit distinct advantages on low cost, easy processing,
and varieties of structural designs, the extremely weak constants (<0.1 cm−1) of spin–orbit cou-
pling (SOC) and high sensitivity to temperature and oxygen exist, originating from forbidden tran-
sition of singlet–triplet and triplet–singlet. Therefore, for pure organic phosphors, the
phosphorescence is normally observed in a low temperature (e.g., 77  K) or some other inert
conditions [13].
In general, pure organic luminophores can mainly be categorized into conventional ones and
nonconventional ones based on the difference in conjugation and luminescent mechanisms. For
both of them, there are three steps for achieving persistent RTP (p-­RTP) in pure organics: the first
one is to open up intersystem crossing (ISC) channels for possible transition from singlet to triplet
through introducing aromatic carbonyls [14–19], heavy atoms (bromine, iodine) [20–23], and heter-
oatoms [24, 25] to promote SOC constants, leading to mass generation of triplet excitons. Then, they
will be stabilized by suppressing molecular vibration and rotation via crystallization [14, 16, 26–28],
host–guest doping  [29–35], construction of organic framework (including supramolecular sys-
tems) [36–41], and clustering [42–48]. After deactivation, the triplet excitons will finally jump back
to ground states with RTP, even p-­RTP emission. Despite various reports (Figure  13.1), the RTP
mechanism, particularly for nonconventional luminophores, is still under consideration.
In this chapter, we will firstly focus on the fundamental mechanism of RTP, including the pho-
tophysical process and the factors influencing lifetime and quantum efficiency. Then, recent
advances in organic RTP materials and their emission mechanism are discussed in detail. Finally,
a brief but clear expectation will be put forward to realize further application.

Handbook of Aggregation-Induced Emission: Volume 1 Tutorial Lectures and Mechanism Studies, First Edition.
Edited by Youhong Tang and Ben Zhong Tang.
© 2022 John Wiley & Sons Ltd. Published 2022 by John Wiley & Sons Ltd.
372 13  Room-­temperature Phosphorescence of Pure Organics

Heavy atoms Crystallization

Br 2 8 18 7

Organic framework
Ri
g
I 2 8 18 18 7
C

idi
S
dI

fie
SOC an

d conformatio
Hetero atoms
tive

Clustering
ec

ns
ff

N P E

Aromatic carbonyl Host–guest doping

Figure 13.1  Various strategies to achieve room-­temperature phosphorescence from two main aspects.

13.2  ­Fundamental Mechanism in Organic Phosphorescence

13.2.1  Photophysical Process for Phosphorescence


In theory, according to the Jablonski diagram (Figure 13.2a), the classical photophysical processes
contain several steps, including photon absorption, internal conversion (IC), fluorescence, ISC,
and phosphorescence. After photon excitation, excitons occur and then jump from S0 (ground
states) to singlet electronic states Sn (n ≥ 1), which is spin-­allowed because of the same spin value.
Subsequently, the fast vibration relaxation (VR) within Sn (n ≥ 1) and Sn → S1 IC occurs based on
the well-­known Kasha’s rule  [49, 50]. Consequently, the radiative transition and IC process of
S1 → S0 coexist with fluorescence emission and other energy decays. Moreover, unlike spin-­allowed
singlet–singlet transitions S0 → Sn, the singlet–triplet transitions S0 → Tm are spin-­forbidden only
when Sn → Tm (m, n ≥ 1) ISC channels are opened via SOC. Then, after analogous VR and IC pro-
cesses, excitons transport to the lowest vibrational level of T1 and then deactivate to ground states
with phosphorescence emission and nonradiative transition. In order to evaluate the whole photo-
physical processes, several parameters, such as rate constants, quantum efficiency, and lifetime,
are introduced. For phosphorescent organics, the urgently concerned phosphorescence efficiency
(Φp) and lifetime (τp) can be defined as follows [51]:
13.2  ­Fundamental Mechanism in Organic Phosphorescenc 373

(a) (b)
Sn kisc Allowed 3(n,
1(π, π*) π*)
Tm

S1

T1 Forbidden Forbidden
Absorption

kf
kic kp
knr
Fluorescence
Phosphorescence
1(n,
Allowed
S0 π*) 3(π, π*)

Figure 13.2  (a) Jablonski diagram for the photophysical process in pure organics. (b) Schematic
representation of El-Sayed rule.

p isc kp p
(13.1)

kisc
isc (13.2)
kf kic kisc

1
p (13.3)
kp knr

where Φisc represents the quantum efficiency of ISC process, and k is the constant for describing
the speed rate in every photophysical steps, including IC rate kic for S1 → S0 transition, fluorescence
emission rate kf, ISC rate kisc as well as phosphorescence emission rate kp and nonradiative decay
rate knr of T1 → S0.
Then, after uniting Equations 13.2 and 13.3, Equation 13.1 will be transformed to Equation 13.4:

1
p
kf kic knr
1 1 (13.4)
kisc kp

According to Equations  13.3 and  13.4, evidently, a series of preliminary conclusions can be
drawn. Firstly, to achieve high-­efficient RTP, improving ISC and phosphorescence emission rates
while restraining fluorescence emission and nonradiative decay are extremely essential. Besides,
on pursuing long phosphorescence lifetime, the nonradiative decay rate knr is surely needed to be
reduced, as well as the value of kp, which may lead to lower phosphorescent efficiency. Therefore,
it is reasonable to believe that a point exists for balancing the value of Φp and τp in pure organics.

13.2.2  Theoretical Study on Phosphorescent Process


Based on the foregoing statement, the ISC rate kisc and phosphorescence emission rate kp play cru-
cial roles in the phosphorescent process. It is pretty worthwhile to make a clear understanding of
these parameters combining with the theoretical study. Starting from the “Fermi golden rule”
374 13  Room-­temperature Phosphorescence of Pure Organics

approximation, the ISC rate kisc from the initial state |i⟩ to the final state |f⟩ can be qualitatively
analyzed as follows [52]:
2 2
kisc Hif E i Ef (13.5)
h
where Hif is the matrix elements of Hamiltonian and its value should be much smaller than the
adiabatic energy gap (Ei − Ef ) to ensure the sense of former approximation. The δ mentioned pro-
vides the energy conservation for nonradiative transition.
Moreover, in the phosphorescent process, thanks to SOC, the ISC channel is opened with the
rate constant (kisc). Therefore, the Hif in Equation 13.5 can be replaced by SOC Hamiltonian (Hsoc),
which is calculated as follows [51]:

Ze 2
HSOC LS (13.6)
2m 2c 2r 3
where r is the orbital radius, and L and S represent the orbital and spin angular momentum,
respectively. The e, m, and c are universal constants with values of 1.6 × 10−19 C, 9.1 × 10−31 kg, and
3 × 108 m/s, respectively. Generally, the expectation value of r−3 operator with the Slater atomic
orbitals includes Z3; thus, the Hsoc in Equation 13.6 depends on the nuclear charge (Z) power 4,
which perfectly explains why the ISC is favored by the presence of heavy atoms.
After simplifying Equation 13.5 with Equation 13.6, utilizing the short-­time and high-­temperature
approximation, the kisc is finally summarized as follows [13]:
2
2 2 EST
kisc S HSOC T exp (13.7)
h kBT 4 kBT

where ⟨S|HSOC|T⟩ represents the SOC matrix elements between singlet (S) and triplet (T). ΔEST is
the energy gap between two electronic states mentioned in equation. kB and T are the Boltzmann
constant (1.38 × 10−23 J/K) and ambient temperature, respectively. The reorganization energy λ is
the characteristic parameter for describing the contribution of nuclear change during the ISC
process.
The well-­known El-Sayed rules [53] (Figure 13.2b) claimed that the ISC process between singlet
and triplet would become more possible when involving a change of the orbital type. That is, for
examples, the ISC process from 1(π, π*) to 3(n, π*) is generally quicker than that to 3(π, π*), similar
to 1(n, π*) →  3(π, π*) and 1(n, π*) →  3(n, π*). More importantly, according to previous reports [13,
52], ⟨1(π, π*)|HSOC|3(n, π*)⟩ or ⟨1(n, π*)|HSOC|3(π, π*)⟩ is nearly two orders of magnitude larger than
⟨1(π, π*)|HSOC|3(π, π*)⟩ or ⟨1(n, π*)|HSOC|3(n, π*)⟩. Therefore, it is truly convincing that ISC can be
promoted through enhancing SOC, which is also proved in Equation 13.7. Except for concentrat-
ing on SOC matrix elements, lowering ΔEST and increasing λ also become alternative approaches
to accelerate the ISC process. In essence, ΔEST is directly determined by the electron exchange
energy, which positively linearly depends on the spatial overlap of highest occupied molecular
orbital (HOMO) and lowest unoccupied molecular orbital (LUMO) wavefunctions [54]. In other
words, the low ΔEST will be obtained by introducing charge transfer (CT) structures with big tor-
sion angle to reduce spatial overlap. But the ΔEST should be at least higher than 0.37 eV to avoid
reverse ISC (rISC) process from triplet to singlet, resulting in delayed fluorescence (DF) phenom-
enon [55, 56]. And λ is judged by the frequency of normal mode and displacement between equi-
librium geometries of two electronic states during ISC.
13.3  ­Recent Progress in Organic RTP Material 375

Furthermore, as mentioned in Equations 13.1 and 13.3, the phosphorescence emission rate kp is


required when we calculate both ϕp and τp, and it can be estimated as follows [13]:
64 4 2
kp ET31S0 T1 S0 , (13.8)
3h 4 c 2

where h is the Planck constant (6.626 × 10−34 J∙s), while ET1S0 and T1 S0 represent the energy gap
and transition dipole moment between T1 and S0, respectively. More importantly, the T1 S0 is
defined as follows [57]:
T1 HSOC Sk Tm HSOC S0
T1 S0 Sk S0 Tm T1
(13.9)
3 1 3 1
k E1 Ek m Em E0

where the summation k runs over all singlet states and the summation m runs over all triplet
states. Combining Equations 13.8 and 13.9, evidently, kp shows a positive correlation to ET1S0 and
T1 S0 , whereas T1 S0 is inversely proportional to the energy gaps. In other words, multiple factors
should be concerned when phosphorescence emission rate is discussed, but there is no doubt that
strengthening SOC truly helps for efficient phosphorescence process.

13.3  ­Recent Progress in Organic RTP Materials

In this section, recent advances in pure RTP organics will be introduced and luminous mechanism
will also be discussed in detail.

13.3.1  Crystallization-­induced RTP


In 2010, Yuan et al. [14] discovered the crystallization-­induced phosphorescence (CIP) phenom-
ena in benzophenone (BP) and its derivatives, methyl 4-­bromobenzoate (MBB) and
4,40-­dibromobiphenyl (DBBP′), which offered a new strategy for designing pure organic RTP
materials through crystal engineering. As illustrated in Figure 13.3, these compounds were almost
nonemissive in solutions, poly(methyl methacrylate) (PMMA) films, and thin-­layer chromatogra-
phy (TLC) plates at room temperature, owing to quick quenching by intramolecular motions with
nonradiative transition, but the emission would occur in 77  K. Nevertheless, these compounds
were induced to emit RTP in crystals with high efficiencies (Φp up to ~40%), while the lifetimes of
them range from 19.2 μm to 4.8 ms. Although BP possessed nearly a 100% ISC efficiency, rigidified
conformation was still needed for RTP emission, which could be easily achieved through crystal-
lization. For example, in decafluorobiphenyl (DFBP) crystals (Figure 13.3d), numerous C–H⋯O
(2.724 Å) and C–H⋯F (2.777 and 2.796 Å) short contacts were able to form a firm 3D interaction
network, which remarkably restricted the molecular motions. As a result, combined with high ISC
efficiency of BP, RTP emission would finally be realized as expected.
Generally, the carbonyl group and nonplanar conformation are favorable for effective SOC. Gong
et al. [26] checked the RTP behaviors of 1,4-­bisbenzil (BBZL), (benzoate-­CoA) ligase (BZL), and its
derivatives, and the results are shown in Figure 13.4a–c. Obviously, both BBZL and BZL owned
more carbonyl groups and even more twisted conformations compared to BP, which were more
likely to be CIP-­active. Similar to BP, BBZL, BZL, and its derivatives, all exhibited CIP features with
high RTP efficiencies in crystal states. After examining the single-­crystal structure of BZL, the
researchers found abundant C–H⋯O (2.416, 2.417, 2.482, 2.483 Å) interaction between one BZL
376 13  Room-­temperature Phosphorescence of Pure Organics

(a) (b)
O NH2
O ′

ABP
R R′ O
R = H, R′= H (BP) Br
R = F, R′= F (DFBP) O
MBB
R = Cl, R′= Cl (DCBP)
R = Br, R′= B (DBBP) Br Br
R = Br, R′= H (BBP)
DBBP′
(c) (d)

Figure 13.3  (a) Chemical structures of BP and its derivatives, MBB and DBBP′. (b) Photographs of DFBP in
solutions (from left to right: n-­hexane, tetrahydrofuran (THF), dichloro-­methane (DCM), acetonitrile, ethanol;
c = 0.2 mg/ml) in PMMA films and on TLC plates at room temperature and 77 K under a 365-­nm UV light.
(c) Photographs of BP and its derivatives, MBB and DBBP′ crystals, under a 365-­nm UV light at room
temperature. (d) Fragmental molecular packing and intermolecular interactions in DFBP crystals.
Source: Adapted with permission from Ref. [14] © 2010 American Chemical Society.

molecule and six adjacent structures. Then, the benzene rings and carbonyls of BZL were fully
rigidified by such intermolecular interactions, greatly preventing the conformational changes
through intramolecular rotations, making it possible for the BZL crystal to emit light. In addition,
the researchers also discovered a unique phenomenon called crystallization-­induced dual emis-
sion (CIDE) [16]; that is, simultaneously enhanced fluorescence and phosphorescence via crystal-
lization, were observed in isophthalic acid (IPA), tetrafluoroterephthalic acid (TFTPA), dimethyl
terephthalate (DMTPA), and dimethyl tetrafluoroterephthalate (DMTFTPA) at room temperature.
More surprisingly, even DF was also detected in the IPA crystal. As depicted in Figure 13.4f, the
above results could be rationalized as follows: firstly, while carbonyls readily promoted the ISC
processes, the abundant and effective intermolecular interactions in crystal states were still neces-
sary for rigidifying conformations, leading to RTP emission. Meanwhile, due to small energy gaps
between singlets and triplets, crystallization also triggered DF via the rISC process. Such unique
CIDE mechanism not only provided a model for fundamental research but also paved the way for
developing new materials with RTP and DF for biological and optoelectronic applications.

13.3.1.1  Heavy Atom Effect


According to Equation 13.5, SOC can be greatly enhanced with the presence of halogen atoms,
such as bromine and iodine. For example, Bolton et al. [20] designed a series of halogenated aro-
matic aldehydes (Br6A, BrC6A, BrS6A, and Np6A) and their dibrominated analogues (Br6, BrC6,
BrS6, and Np6) (Figure  13.5). In solutions, these halogenated aromatic aldehydes produced a
strong fluorescent emission but no phosphorescence emission. However, in crystal states, halogen
bonding between these aldehydes delocalized the electrons of the carboxyl oxygen partially toward
13.3  ­Recent Progress in Organic RTP Material 377

Figure 13.4  (a) Chemical structures of BBZL, BZL, and its derivatives. (b) Photographs of BBZL, BZL, and its
derivative crystals under a 365-­nm UV light at room temperature. (c) Fragment intermolecular interactions
of the BZL crystal. Source: Adapted with permission from Ref. [26] © 2013 Science China Press and
Springer-­Verlag Berlin Heidelberg. (d) Chemical structures of TPA, DMTPA, TFTPA, DMTFTPA, and
IPA. (e) Photographs of TPA, DMTPA, TFTPA, DMTFTPA, and IPA crystals under a 365-­nm UV light at room
temperature. (f) Schematic illustration of the CIDE phenomenon. Source: Adapted with permission from
Ref. [16] © 2015 The Royal Society of Chemistry.

the neighboring halogen, leading to enhanced triplet generation and the vibrational freedom of
which was limited by tight packing, allowing triplets to decay emissively with RTP. However, these
homocrystals suffered from a low RTP efficiency (only 2.9% for Br6A) due to the excimer-­induced
self-­quenching. To solve these problems, the researchers utilized dibrominated analogues as host
to isolate the aldehyde molecules and thus prevent their self-­quenching, resulting in improved
RTP efficiency (55%) (Figure 13.5c). More importantly, under a 365-­nm UV light, the RTP emission
color of these mixed crystal could be adjusted from blue (BrC6A-­BrC6), green (BrS6A-­BrS6) to
orange (Np6A-­Np6).
378 13  Room-­temperature Phosphorescence of Pure Organics

(a) Br
Br O

R R
O
R=CHO BrC6A
R=CHO Br6A R=Br BrC6
R=Br Br6
Br
O
Br S

O
S R
R
R=CHO BrS6A R=CHO Np6A
(b) R=Br BrS6 R=Br Np6

BrC6A-BrC6 Br6A-Br6 BrS6A-BrS6 Np6A-Np6

(c)

Figure 13.5  (a) Structures of various hosts/aldehydes and (b) the photographs of their mixed crystals
under a 365-­nm UV light, each containing a 1 wt% aldehyde chromophore and a 99 wt% analogous host
compound. (c) Schematic illustration of designing organic phosphorescence with heavy atom effect.
Source: Adapted with permission from Ref. [20] © 2011 Springer Nature.
13.3  ­Recent Progress in Organic RTP Material 379

Shimizu et al. [22] reported the CIP features of a series of bis(aroyl)benzene derivatives, named
BADBB-­1–5 (Figure 13.6a). All these compounds showed typical RTP features in crystals, along
with the emission color ranging from blue to green due to the effect of substituents (Figure 13.6b).
More importantly, thanks to the existence of bromine atoms, SOC constants were largely enhanced
with high quantum efficiencies of 5–18% at room temperature and 38–67% at 77 K.
Gong et al. [17] designed (4-­(9H-­carbazol-­9-­yl)phenyl)(phenyl)methanone (CZBP) to incorpo-
rate a planar electron donor carbazole with an acceptor BP. Such a donor–acceptor (D–A) structure
would significantly facilitate the CT with low ΔEST and offer more possibilities for the ISC process.
Besides, the heavy bromine atoms ((4-­(3-­bromo-­9H-­carbazol-­9-­yl)phenyl)(phenyl)methanone
[BCZBP], (4-­(3,6-­dibromo-­9H-­carbazol-­9-­yl)phenyl)(phenyl)methanone [DBCZBP]) were also
introduced for comparison (Figure 13.6c, d). The crystals of these compounds with dense packing
and effective intermolecular interactions isolated the triplet excitons from quenching sites, thus
leading to p-­RTP emission according to the CIP mechanism. The results revealed that p-­RTP could
be successfully achieved in crystals with a quantum efficiency of DBCZBP as high as 7.5% owing
to the heavy atom effect.
Mao et al. [23] also designed and synthesized a D–A-­type dipenylsulfone derivative (Cz-­DPS)
with very strong fluorescence emission but a weak RTP property in crystals. After introducing the
iodine atom into the molecule, the ICz-­DPS exhibited a far more strong phosphorescent signal in
crystals at 559 nm compared with Cz-­DPS due to heavy atom effect, the lifetime of which was as
long as 19 ms, as shown in Figure 13.6. This work revealed that halogen bonding could truly help
to induce RTP via strengthening ISC process for generating more contents of triplets.

(a) (b)

(c) (d) CZBP CZBP BCZBP DBCZBP


2 mm 200 μm 5 mm

UV on UV off UV on UV on

(e) (f) Solution (g) 100


Crystal
Fluor. Phos.
Normalized intensity

416 nm 559 nm
10–1
PL intensity

Fluor.
383 nm
10–2
τ = 19 ms

400 450 500 550 600 650 0 20 40 60 80 100


Wavelength (nm) Time (ms)

Figure 13.6  (a) Chemical structures of BADBB derivatives. (b) Photographs of BADBB derivative crystals
under a 365-­nm UV light at room temperature. Source: Adapted with permission from Ref. [22]
© 2016 Wiley-­VCH Verlag GmbH & Co. KGaA. (c) Chemical structures of CZBP, BCZBP, and DBCZBP.
(d) Photographs of CZBP, BCZBP, and DBCZBP crystals under a 365-­nm UV light at room temperature.
Source: Adapted with permission from Ref. [17] © 2015 Wiley-­VCH Verlag GmbH & Co. KGaA. (e) Chemical
structures of compounds Cz-­DPS and ICz-­DPS. (f) Photoluminescence (PL) spectra of ICz-­DPS in chloroform
solution (10−5 mol/l) and crystal state. (g) Time-­resolved emission decay curve of a 559-­nm emission from
the crystal of ICz-­DPS. Source: Adapted with permission from Ref. [23] © 2015 Wiley-­VCH Verlag GmbH &
Co. KGaA.
380 13  Room-­temperature Phosphorescence of Pure Organics

13.3.1.2  Molecular Interaction


Based on foregoing statements, in some cases, materials with AIE property are more possible to
achieve RTP emission owing to the formation of stable triplets via aggregation. Furthermore, the
essence of aggregation can be ascribed to multifarious molecular interactions, including inter-­and
intramolecular interactions [58, 59].
Recently, inspired by a reported ultralong organic phosphorescence (UOP) emitter (p-­Cz), Mao
et al. [60] synthesized an isomer o-­Cz with close spatial proximity between donor and acceptor
units, resulting in extra space charge transfer (SCT) effect, which affords an enhanced CT charac-
ter, as shown in Figure 13.7a. More importantly, the SCT effect was supposed to reduce ΔEST, lead-
ing to highly efficient dual-­channel triplet harvesting (Φp  =  16.6%) with RTP and thermally
activated delayed fluorescence (TADF) through ISC and rISC processes, respectively. The time-­
dependent density functional theory (TD-­DFT) was performed to prove the existence of SCT effect
according to the observation of through-­space orbital overlap between carbazole (donor) and BP
(acceptor) units in o-­Cz. Additionally, the reduced density gradient analysis also revealed the
­significant intramolecular interactions in the folded o-­Cz, whereas they are pretty weak in p-­Cz,
shown in Figure  13.7b. Thanks to the above SCT feature, o-­Cz exhibited an obvious and long
­afterglow after the removal of not only the ultraviolet (UV) light but also near-­infrared (NIR) light
with a four seconds emission (Figure 13.7c). This work provides a new platform for designing NIR-­
excited ultralong organic RTP materials and realizing their potential usages in biological sciences.

13.3.1.3  H-­aggregation
In 2015, inspired by the mechanism of long-­lived afterglow in inorganic materials, An et al. [61]
deduced the existence of energy trap state ( T1* ) in organics via aggregation, which is analogous to
the charge-­trapping center in inorganics, seemingly playing a key role in stabilizing the triplets
with long-­lived excitons for ultralong RTP emission. As a proof of the above principle, the
researchers synthesized 4,6-­diphenyl-­2-­carbazolyl-­1,3,5-­triazine (DPhCzT) and systematically
investigated its photoluminescence properties. By employing TD-­DFT calculation, interestingly,
the researchers found that ISC channels would significantly increase with isolated DPhCzT mon-
omers aggregating into dimers, shown in Figure  13.8a and b. Moreover, the H-­aggregation of
DPhCzT dimers was further confirmed by checking the angle between transition dipoles and the
interconnected axis (80.9°), which was larger than the critical value of 54.7°. The DPhCzT exhib-
ited p-­RTP with a lifetime as long as 1.06 seconds and an efficiency of 1.25%. To further examine
the generality of H-­aggregation in designing RTP organics, a series of analogous molecules named
1,4-­diethoxybenzene (DEOPh), 4,6-­diethoxy-­2-­carbazolyl-­1,3,5-­triazine (DECzT), 2-­carbazolyl-­4,6-­
dichloro-­1,3,5-­triazine (CzDClT), and di(9H-­carbazolyl)-­phenylphosphine (DCzPhP) were syn-
thesized, with lifetimes of 0.23–1.35  seconds and efficiencies of 0.08–2.1%. Meanwhile, the
­emission color of these compounds could be adjusted from green (515  nm) to red (644  nm),
shown in Figure 13.9c. In fact, such H-­aggregation can also be classified into a type of π–π interac-
tion, which not only benefits stabilization of the triplets but also opens up the channels for the
ISC process.
In addition, currently, many other H-­aggregation systems are gradually designed and reported.
For example, Li et al. [62] reported a series of fluoro-­substituted phenylboronic acid derivatives
with UOP feature at room temperature, and the chemical structures are listed in Figure  13.9a.
Notably, the UOP lifetime of 24FBP crystal was up to 2.50 seconds. After examining the crystal
structure of 24FBP, the researchers pointed that such H-­aggregation played a critical role in confin-
ing the molecular rotation and prolonging the UOP lifetime.
(a) O p-Cz O o-Cz UV (b)

UV N Vis
N
IC

SCT
IC

NIR
T
T

p-Cz o-Cz
HOMO LUMO LUMO+3

(c) 365 nm On UV Off (d)


1.0 Steady
Delayed 8 ms
0.8
Intensity (a.u.)

0.6
o-Cz 0s 2s 4s 6s 8s

808 nm On NIR Off 0.4


CT LE
0.2

0.0
o-Cz 0s 1s 2s 3s 4s 400 450 500 550 600 650 700 750
Wavelength (nm)

Figure 13.7  (a) Chemical structure of p-­Cz and o-­Cz; HOMO, LUMO, and LUMO+3 electron cloud distribution of o-­Cz; the red circle indicates the spatial
orbital overlap of the donor and acceptor units. (b) Distribution of intramolecular interaction regions in p-­Cz and o-­Cz. (c) Photographs of o-­Cz under
UV-­light or NIR-­light irradiation and after removing the excitation. (d) Steady-­state and delayed (8 ms) PL spectra of o-­Cz. Source: Adapted with
permission from Ref. [60] © 2019 The Royal Society of Chemistry.
382 13  Room-­temperature Phosphorescence of Pure Organics

Later, Gu et al. [63] reported the emission behavior of 2,4,6-­trimethoxy-­1,3,5-­triazine (TMOT)


(Figure 13.9b). Except the UOP feature at room temperature, tunable afterglow was also observed
when changing the excitation from 300 to 360 nm. After checking the single-­crystal structure of
TMOT, the researchers believed that the mitigation of molecular movement in the same plane and
H-­aggregation were responsible for such tunable UOP phenomena.
Moreover, Yuan et al. [64] designed a series of CN-­substituted phenylcarbazole isomers, named
pCNPhCz, mCNPhCz, oCNPhCz, and DCNPhCz (Figure  13.9c, d), which showed an ultralong
RTP emission with a lifetime of 0.92, 0.81, 0.65, and 0.48 seconds, corresponding to quantum effi-
ciencies of 1.8, 3.7, 2.9, and 8.6%, respectively. After systematic experimental and theoretical stud-
ies, the researchers reckoned that such RTP properties were closely related to the splitting energy
(Δε) of H-­aggregation. More importantly, the Δε with thermally activated reversed-­phase transfor-
mation from the low-­lying dark state to the high-­lying transition-­allowed emission state of
H-­aggregation was identified to summarize the ultralong RTP process. That is, compounds with
larger Δε would lead to longer lifetime, while smaller Δε resulted in short lifetime but higher quan-
tum efficiency. These findings would surely deepen the understanding of the emission behaviors
of organics with H-­aggregation in crystals.

13.3.2  Doping in Rigid Matrix-­induced RTP


For pure organics, to achieve p-­RTP in air, it is also vital to control the nonradiative decay rates of
triplet and isolate the effect of oxygen. As a result, doping some organic luminophores in rigid
environment seemingly is an excellent choice for p-­RTP.

(a) DPhCzT monomer (b) DPhCzT dimer

+0.3 eV T10
H L
T6 H L
H-1 L
T9
H L

S1 T4 S1 H L
H-1 L T7
H-3 L
H-1 L

H L H L
T3 T3 H-3 L
H-3 L
H L H L H-2 L H-2 L
H-1 L H-1 L H-1 L
T2
–0.3 eV H-1 L
T2 H L
H-1 L T1
T1 H-1 L

Figure 13.8  TD–DFT calculation results of DPhCzT monomer (a) and its coplanar dimer (b) with energy
level of singlets (Sn) and triplets (Tn). H and L represent the highest occupied molecular orbital and lowest
unoccupied molecular orbital, respectively. (c) The steady-­state photoluminescence and ultralong
phosphorescence spectra of designed molecules. The insets show the corresponding photographs taken
before and after removing the excitation at room temperature. (d) Single-­crystal packing mode of
DPhCzT. (e) Schematic energy diagram of J-­aggregation (θ < 54.7°) and H-­aggregation (θ > 54.7°).
(f) Possible energy-­transfer processes for fluorescence (Fluo.), phosphorescence (Phos.), and ultralong
phosphorescence (Ultralong phos.) in DPhCzT aggregates. Source: Adapted with permission from Ref. [61]
© 2015 Springer Nature.
(c) Steady-state PL Ultralong phos.

DEOPh 515 nm τ = 0.71 seconds


547 nm τ = 0.69 seconds
O

DECzT 529 nm τ = 1.28 seconds


O N O 574 nm τ = 1.35 seconds

N N
N

530 nm τ = 1.06 seconds


PL intensity (a.u.)

DPhCzT
N 575 nm τ = 1.05 seconds

N N
N

CzDCIT 543 nm τ = 0.47 seconds


Cl N Cl 591 nm τ = 0.49 seconds
N N
N

DCzPhP 587 nm τ = 0.21 seconds

NPN
τ = 0.2
644 nm seconds

300 400 500 600 700

Wavelength (nm)

(d) (f)
Sn
θ = 80.9° Tn
3.38 Å ISC T*n
Stabilizing

Transition dipole moment

Exc.
(e)
Monomer Dimer Ultralong
Fluo. Phos.
phos.
E*

~ns ns–ms >S


J H

S0
0 54.7 90 S0
θ (°)

Figure 13.8  (Continued)
F F
(a) (b) 0.9
F F
360 nm
F F F F
0.7
B B B B 350 nm
HO OH HO OH HO OH HO OH OCH3
2FPB 23FPB 24FPB 234FPB 340 nm
N N 0.5
y 360 nm
Daylight UV off 320 nm
0s 2s 4s 6s 8s H3CO N OCH3 0.3

300 nm
UV on
10 s 12 s 14 s 16 s 18 s 0.1
TMOT
0.0 0.2 0.4 0.6 0.8
X

(c) pCNPhCz mCNPhCz mCNPhCz’ oCNPhCz DCNPhCz


CN τ = 0.92 seconds (d) (e)
Φ = 1.8 %
S1
N T1 E+
T*1
pCNPhCz 0 kRPT 2Δε
CN kPT
PL intensity (a.u.)

τ = 0.81 seconds
Φ = 3.7 %
N
2 E–
Absorption

mCNPhCz
Time (seconds)

τ = 0.65 seconds 4
CN Φ = 2.9 %
N
kEnr- kEr kEnr-
6
oCNPhCz
NC CN τ = 0.48 seconds
Φ = 8.6 % 8
N

S0 E0
DCNPhCz 10
Monomer H-aggregation
350 450 550 650 750
Wavelength (nm)

Figure 13.9  (a) Chemical structures of 2FPB, 23FPB, 24FPB, and 234FPB; photographs of the 24FPB crystal taken before and after removing the
excitation of a 365-­nm UV light. Source: Adapted with permission from Ref. [62] © 2019 Wiley-­VCH Verlag GmbH & Co. KGaA. (b) Chemical
structures of TMOT and CIE coordinate diagrams with the excitation from 300 to 360 nm. Source: Adapted with permission from Ref. [63] © 2019
Springer Nature. (c) Steady-­state PL and RTP spectra of the CN-­substituted molecules excited by a 295-­nm UV light at room temperature with the
corresponding RTP lifetime (τ) and quantum efficiency (Φ). (d) Photographs of pCNPhCz, mCNPhCz, mCNPhCz′, oCNPhCz, and DCNPhCz crystals
taken before and after turning off the excitation (365 nm) at room temperature. (e) Schematic representation of the RTP mechanism in
H-­aggregation. Source: Adapted with permission from Ref. [64] © 2019 The Royal Society of Chemistry.
13.3  ­Recent Progress in Organic RTP Material 385

13.3.2.1  Host–Guest System


For instances, Hirata et  al.  [29] explored the efficient RGB RTP from pure organic amorphous
host–guest systems in air. By using rigid steroidal compounds as a host matrix, for guest species,
the interaction with oxygen and nonradiative relaxation of triplet excitons could be minimized.
More importantly, the deuteration of the guest surely reduced its nonradiative decay rate, while a
secondary amino group substitution also helped to promote ISC process. As a result, the host–
guest systems exhibited both a nonradiative transition from T1 state less than 10−1  s−1 at room
temperature and effective ISC process, which allows persistent RGB RTP emission with a quantum
yield beyond 10% and a lifetime exceeding one second in air. This approach paved a facile way for
achieving amorphous p-­RTP, which could be surely important for the next-­generation imaging
technology and display applications. Later, the researchers [30] doped single aromatic structure
(6EhHBC) into rigid β-­estradiol matrix to form a host–guest system with the content of guest com-
pound accounted for only 0.3 wt%. The ΔEST of the whole system was below 0.5 eV, which made it
possible for DF. As a result, this host–guest system established a blue-­green TADF in 490 nm as
well as red RTP at 628 nm with a lifetime of 3.9 seconds. After superposition, the afterglow of the
host–guest system would change from white to yellow and finally to red. This work proposed a
facile strategy for designing organics with a white afterglow, which was rarely reported before. All
the above statements can be summarized in Figure 13.10.

(a) (b)
Material design
OH
D D
D D CH3OH
C2D5
N D H
C2D5
(i)
D D H H
D D HO
HO
Guest Host
Secondary amino-substituted Amorphous hydroxy steroid
deuterated carbon
(iii) (ii) 6EhHBC β-estradiol
Reducing knr(RT) Reducing kq(RT) (0.3 wt%, guest) (99.7 wt%, host)

(iv) knr(RT) + kq(RT) ⋘ kr < 100 S–1 Under


excitation
0
(v) Efficient persistent RTP
Time (seconds)

Under UV
After stopping UV 4
After ceasing
excitation
8
0
3
6
9 12
Time (seconds) 12

Figure 13.10  (a) Material design for efficient p-­RTP from pure organics in air. Source: Adapted with
permission from Ref. [29] © 2013 Wiley-­VCH Verlag GmbH & Co. KGaA. (b) Chemical structures and emission
characteristics of 0.3 wt% 6EhHBC doping β-­estradiol material. Source: Adapted with permission from
Ref. [30] © 2017 Wiley-­VCH Verlag GmbH & Co. KGaA.
386 13  Room-­temperature Phosphorescence of Pure Organics

Lei et al. [31] reported organic doping systems with CT feature using 4-­(2-­(4-­(diphenylamino)
phenyl)-­2-­oxoethyl)benzonitrile (DOB) as the guest, and TPA and 4-­(cyanomethyl)benzonitrile
(CBN) as donor and acceptor hosts, respectively. Both TPA and CBN exhibited weak fluorescence
and nearly no phosphorescence at room temperature, while enhanced fluorescence accompanying
with long-­lived p-­RTP occurred after host–guest doping due to the intermolecular D–A interaction.
In detail, as shown in Figure 13.11b, for CBN, after doping with DOB, the fluorescent efficiency
dramatically increased from 17 to 66%. After ceasing the excitation, an obvious green afterglow
was observed at room temperature with a lifetime of 317 ms and a high phosphorescent efficiency
of 10%, while the phosphorescent emission of CBN could only be detected in 77 K. Similar finding
also appeared in TPA/DOB doping system with an increased fluorescent efficiency from 10 to 74%
and long-­lived afterglow (lifetime of 278 ms) with a high efficiency of 14%. More interestingly, for
the DOB crystal, the fluorescent and phosphorescent efficiencies were only 37.2 and 5.8%, respec-
tively, with a lifetime of 0.0067 ms at room temperature. Thereby, such facile host–guest doping
had been certainly proved to become effective for obtaining materials with both elevated luminous
efficiencies and lifetimes. To gain deeper insight into this unique phenomena, the Jablonski dia-
gram was employed for assistance. Firstly, for most of the host–guest system, strong intermolecu-
lar interactions would surely exist and rigidify the molecular conformation, thus reducing the
nonradiative transition of both singlet and triplet excitons to boost quantum efficiencies and life-
times. Besides, as illustrated in Figure 13.11c, since the host and guest molecules possessed the
same groups, the guest molecules were readily inserted into the host crystal structures. Therefore,
the donor part of DOB can form CT excited states with the acceptor host, while the acceptor part
of DOB can also form CT excited states with the donor host. The generated CT excited states owned
a small energy gap of singlets to ground state and singlets to triplet, thus offering possible ISC pro-
cess and RTP emission. This work developed a facile approach to obtain p-­RTP materials with high
efficiencies and lifetimes through the host–guest doping; the utilization of the intermolecular CT
effect might inspire further studies for designing doping systems and expanding the scope of
organic phosphorescence applications.
Cyclodextrin (CD) is a kind of cyclic oligosaccharide compounds with a natural cavity, which
can contain a variety of appropriately sized guest molecules and form inclusion complexes in aque-
ous solution. According to the numbers of glucose units in per molecule, the CD can be mainly
divided into three types, namely, α-­CD, β-­CD, and γ-­CD, corresponding to the glucose units of 6, 7,
and 8 per CD molecule, respectively.
Cao et al. [32] synthesized light-­driven β-­CD-­Azo with trans-­and cis-­conformation, which could
easily be transformed to each other under 360 and 430 nm irradiations. In aqueous solution, β-­CD-­
Azo with trans-­and cis-­conformations could be induced with a circular dichroic property, and the
α-­BrNp could also be the guest molecule involved in cis-­β-­CD-­Azo, resulting in RTP emission, as
shown in Figure  13.12a. Combining such a circular dichroic signal with the RTP output of the
above organics, the logic gates were constructed and the potential applications in information
technology were also under expectation.
Later, Li et al. [33] developed a series of amorphous organics with efficient and multicolor RTP
emission through facile modification of phosphor moieties to β-­CD. Although the ordered crys-
tal lattice was absent, the hydrogen bonding between CD derivatives locked the phosphors to
suppress the nonradiative transition and shielded phosphors from quenchers, leading to highly
efficient (16.9%) RTP emission. Due to various structures of branched phosphors, the emission
colors of β-­CD derivatives could be determined from blue to green and even red. Furthermore,
BrNp-­β-­CD was utilized to construct a host–guest system associating with a fluorescent guest
13.3  ­Recent Progress in Organic RTP Material 387

(a) Acceptor host Donor host


Guest (DOB)
(CBN) (TPA)
CN
O
N
NC CN N

D–A interaction D–A interaction


(b) Φ = 17% Φ = 66% Φ = 10%

Doping Turn
off

CBN CBN/DOB

Φ = 10% Φ = 74% Φ = 14%

Doping Turn
off

TPA TPA/DOB 0.1 seconds 0.5 seconds 1.0 seconds 1.5 seconds

(c)
(Host) (Guest) (Host)
D–A interaction D–A interaction
TPA A–D CBN
S1, A
S1,D S1,C CT S1, G CT
C S1,C
IS ISC ∆ = 3.38 eV
T1,C T1,C
Δ = 3.68 eV SC ∆ = 3.06 eV
RI T1, A
Prompt fluo.
Delayed fluo.

Prompt fluo.
Phos.

Phos.
Abs.

T1,D ∆ = 3.07 eV
∆ = 3.00 eV ∆ = 2.85 eV
T1, G
Δ= 2.59 eV
∆ = 2.74 eV ∆ = 2.96 eV
∆ = 2.75 eV

S0,D S0, C S0,G S0, C S0, A

Figure 13.11  (a) Chemical structures of host and guest molecules; (b) photographs of CBN, TPA, CBN/DOB
(0.5 mol%), and TPA/DOB (0.5 mol%) powders under a 365-­nm UV light and after the removal of excitation;
(c) Jablonski diagram for CBN/DOB and TPA/DOB powders. Source: Adapted with permission from Ref. [31]
© 2019 American Chemical Society.

molecule AC, as depicted in Figure 13.12b. Both RTP and fluorescence were obtained and mixed
to white emission, which opened up new research paths to pursue pure organic white-­light-­
emitting materials.
In addition, Chen et al. [34] reported the RTP achieved by host–guest recognition between γ-­CD
and 4-­bromo-­1,8-­naphthalic anhydride polymer (poly-­BrNpA), which could be controlled by the
photoisomerization of the Azo unit of the other polymer (poly-­Azo) in aqueous solution, as illus-
trated in Figure 13.12c.

13.3.2.2  Doping in Polymer Matrix


Moreover, the polymeric matrix can also be well used to create rigid and oxygen barrier environ-
ment. Su et al. [35] obtained PVA composites with excitation-­dependent long-­life luminescence
through doping various pyrene derivatives (PYM, HPY, PCA, and PBA). Undoubtedly, the PVA
matrix could hinder rotation and vibration of these pyrene derivatives via abundant hydrogen
bonding, leading to RTP emission. More surprisingly, when changing the wavelength of excitation,
388 13  Room-­temperature Phosphorescence of Pure Organics

OTs
trans-β-CD-Azo Br
(a) CH2OH CH2OTs
O O (b) Br
OH O OH O
6 a.K2CO3/DMF
O O O
OH OH
b.KOH/MeOH N (E)
c.HCI
N COOH R BrNp-β-CD BrHB-β-CD
HO N (E) R=
Br O
N –OTs Br
COOCH3
430 nm 360 nm N
+N
HOOC HOOC O
Br
O Br O BrBp-β-CD BrNpA-β-CD
N N N N
(Z) (Z) –
Cl O O
O
+
cis-β-CD-Azo/α-BrNp cis-β-CD-Azo N
H2
AC

(c) s
m
ISC T
O n
NH O
H2N
Hydrogen Non-radiative
bonding decay
O N O
Phosphorescence

Br
poly-BrNpA γ-CD

RTP

BrNp-β-CD BrBp-β-CD BrHB-β-CD

Host-guest interaction

BrNpA-β-CD AC AC@BrNP-β-CD

Figure 13.12  (a) Synthetic routes of β-­CD-­Azo and its inclusion phenomenon with α-­BrNp in water. Source:
Adapted with permission from Ref. [32] © 2014 The Royal Society of Chemistry; (b) chemical structures of
host cyclodextrin derivatives and guest molecule AC. Photographs of BrNp-­β-­CD, BrBp-­β-­CD, BrHB-­β-­CD,
BrNpA-­β-­CD, AC, and AC@BrNp-­β-­CD. Source: Adapted with permission from Ref. [33] © 2018 American
Chemical Society; (c) schematic illustration of the CD–RTP system via host–guest interaction between
poly-­BrNpA and γ-­CD. Source: Adapted with permission from Ref. [34] © 2016 The Royal Society of
Chemistry.

the emission color can be adjusted from blue to red, as shown in Figure 13.13a. The mechanism of
excitation-­dependent RTP can be ascribed to the coexistence of two emission centers in the doped
polymer films (Figure 13.13b, c). In detail, one of them belongs to the isolated molecular state,
while the other is from the aggregation state. Under low doping contents, the film mainly exhibited
properties of the isolated state; thus, the photonic channel was primarily from S1 to T1, while the
aggregation state contributed very little. Therefore, the blue afterglow was dominant after removal
of the excitation source. When the doping contents increased, the isolated state gradually decreased
owing to the appearance of aggregation state; meanwhile, the aggregation state gradually increased
to accelerate the ISC pathway. It could be deduced that at high doping contents, the photonic chan-
nel would totally change to S1 and T1 , which is much lower than S1 and T1, respectively.
Consequently, the red afterglow was observed. This work developed a reasonable strategy for
designing excitation-­dependent phosphorescence through modulating the doping contents of pyr-
ene derivatives into PVA, which could be hopefully applied in biosensing, light-­emitting devices,
data storage, and information security.
13.3  ­Recent Progress in Organic RTP Material 389

Doped with PVA matrix Sn


IC S1
PYM HPY PCA PBA S1 S1
ISC S′1 S′1 ISC
T1 T1
T′1 T′1

Phos 1 Phos 1
Phos 2 Phos 2
S0
High-energy excitation Low-energy excitation

Isolated molecule Aggregation

Figure 13.13  (a) Chemical structures of PYM, HPY, PCA, PBA, and PVA; photographs of the doped PVA films
under various excitation wavelengths; (b) schematic diagram for the energy levels of doped polymer films;
(c) schematic diagrams of the isolated molecular state and aggregation state of PYM-­doped PVA polymer.
Source: Adapted with permission from Ref. [35] © 2019 Wiley-­VCH Verlag GmbH & Co. KGaA.

13.3.3  Clustering-­triggered RTP


Unlike the above-­mentioned conventional luminophores with large conjugated π electron blocks,
such as anthracene, phenanthrene, dibenzothiophene, dibenzofuran, carbazole, triphenylamine,
and so on, whose photophysical processes and emission mechanisms are well established, for
those nonconventional ones, they only possess electron-­rich N, O, S, and P atoms and derived
groups like amide, imide, anhydride, ester, etc.; however, their emission mechanism is still under
consideration (Figure 13.14) [65–67].

13.3.3.1  Natural Products


In 2013, Gong et al. [68] observed a fascinating bright blue emission (382, 433 nm) from commer-
cial rice (Figure 13.15a, b). Since rice is a mixture of various natural products, in order to deeply
clarify the mechanism of this interesting phenomenon, the photophysical properties of the main
component of rice and starch were checked. As a result, its aqueous solution and TLC plates were
nonluminescent at room temperature, whereas it exhibited a strong emission at 77 K. More impor-
tantly, starch powders could even emit brightly at room temperature, as shown in Figure 13.15c.
Subsequent research on cellulose, a polysaccharide possessing an analogous structure to starch
based on the same monomer (glucose), showed similar PL behaviors (Figure 13.15d). However, the
former luminous mechanism cannot well explain the emission phenomena of starch and cellu-
lose, which led the researchers to reconsider the relationship between chemical structures and
luminescent properties. For starch and cellulose, apart from hydrogen and saturated carbon, there
were only electron-­rich oxygens existing. Therefore, it came quite natural to ascribe the emission
to oxygen. After examining the single-­crystal structure of their elementary unit (glucose), multiple
O⋯O short contacts are found (Figure 13.15e, f), which promotes electron clouds to be overlapped
with diverse clusters, providing the structures with extended conjugations and rigidified confor-
mations. Moreover, these clustered chromophores could be excited and then radiatively
390 13  Room-­temperature Phosphorescence of Pure Organics

H
S N N O

O
N N Conventional OH
N B building blocks
F F
NH
O O O OH
Luminogens
O O
NH O
Nonconventional
NH O
H chromophores O O
O N NH N
P O O
N
O O O
H2N O O C
H O S N O
N OH C C O
O SO3H
S
C N O C O
O S O COOH

C N C S

Figure 13.14  Example of conventional building blocks and nonconventional chromophores for organic
luminophores. Source: Adapted with permission from Ref. [67] © 2017 Wiley-­VCH Verlag GmbH & Co. KGaA.

deactivated with the help of a strong molecular interaction and effective physical constraints
among polymer chains, which distinctly rigidify the molecular conformations, preventing the non-
radiative deactivations and resulting in bright emission. Such above statements can be summa-
rized to one mechanism, namely, clustering-­triggered emission (CTE), which is a branch of AIE
mechanism, applied in explaining unique emission behaviors of these nonconventional
luminophores.
Additionally, when emission behaviors of nonconventional luminophores, carbohydrates, in
particular, are studied, the generality of the CTE mechanism is still remained to be verified.
Currently, Zhou et al. [42] investigated the emission behaviors of d-­(+)-­xylose (d-­Xyl), pentaeryth-
ritol (PER), d-­fructose (d-­Fru), and d-­galactose (d-­Gal), and the results are summarized in
Figure 13.16. All the compounds exhibited AIE features, which could be well understood using
CTE mechanism. More surprisingly, color-­tunable afterglows were detected in d-­Xyl crystal at
room temperature and also in concentrated solutions of PER, d-­Fru, and d-­Gal at 77 K with vary-
ing excitation wavelengths. To further understand the tunable p-­RTP, single-­crystal structure of
d-­Xyl was detected. Evidently, besides hydrogen bonds, intermolecular short dihydrogen bonds of
O–H⋯H–O, C–H⋯H–O, and C–H⋯H–C were also present, which remarkably restricted the
molecular motions, providing rigid conformations in the crystals. Moreover, abundant intra-­ and
intermolecular O⋯O electronic interactions were found, offering effective 3D through-­space
13.3  ­Recent Progress in Organic RTP Material 391

Starch

2.707, 2.711
2.774, 2.775
2.847
clustering of
nonconventional
chromophores
Emission

Cellulose Electron overlap

Figure 13.15  (a) Photograph of the rice taken under a 365-­nm UV light at room temperature. (b) The
emission spectrum of rice. Photographs of the solid powders of (c) starch and (d) cellulose taken under a
365-­nm UV at room temperature. Source: Adapted with permission from Ref. [68] © 2013 Science China
Press and Springer-­Verlag Berlin Heidelberg. (e) Single-­crystal structure of d-­(+)-­glucose with
intermolecular interactions around one molecule denoted. (f) Illustration of “interconnected” subgroups via
through-­space electronic interactions in the clusters. Source: Adapted with permission from Ref. [67]
© 2017 Wiley-­VCH Verlag GmbH & Co. KGaA.

conjugation and possible ISC process. After clustering, diverse clustered chromophores with vari-
ous emission abilities would finally emit from triplets in response to different excitation wave-
lengths, resulting in multicolor p-­RTP. This work proved the universality of the CTE mechanism
in studying PL emission, even tunable p-­RTP behaviors of nonconventional luminophores, which
might encourage the rational design of single-­component molecules with multicolor afterglows in
a variety of applications in chemistry.
To gain deeper insights into the emission mechanism of polysaccharide, Dou et al. [43] studied
emission behaviors of sodium alginate (SA), a kind of anionic natural products composed of man-
nuronic (M) and guluronic (G) acids, as shown in Figure 13.17a. GMr, for example, was nearly
nonemissive in dilute solutions (i.e., 0.005  wt%, Φ  ≈  0), while further enhanced emission was
detected as concentration increased. Remarkably, for concentrated solutions (i.e., 8 wt%), bright
gray white emissions were observed under a 365-­nm UV light with a Φ value of 3.9%, along with
multiple peaks at 352/417/484  nm. More interestingly, the GMr solid powders and films even
exhibited p-­RTP properties with a light green afterglow after ceasing a 312-­nm UV light. For GMr
solid powders, the blue PL peaks could be identified at 429 and 450 nm, while the RTP could also
be recorded at 518 nm, as depicted in Figure 13.17e. Based on the above results, it is reasonable to
ascribe such unique photophysical properties of SA to the clustering of oxygen atoms and carboxy-
lates in concentrated solutions and solids. Clustering of these nonconventional chromophores is
an indispensable prerequisite; otherwise, the sole conformation could not bring about visible-­light
emission. Meanwhile, clustering also induced more rigidified conformations through electron
cloud overlap. In detail, as illustrated in Figure 13.17f, SA chains were primarily extended in dilute
solutions with isolated oxygen atoms and carboxylates owing to the repulsive force among the
392 13  Room-­temperature Phosphorescence of Pure Organics

254 285 312 365 nm


(a) H (b)

UV on
H H O rt. 365 nm
HO
OH H
H OH
H D-Xyl OH

UV off
1×10–4 0.025 0.05 1M

(c) rt. 254 285 312 365 nm (d)


intra O...O (Å)
inter O...O (Å)
14
HO OH 2.7

41
HO OH

2.8
6 2.71
82 4

2.7
PER 2.

05
2.333
2.768
2.768

2.6
2.676

76
H 2.917
H HO
O
HO
H OH
H H
OH OH
D-Fru

OH OH
HO
H
HO H
H OH
H OH
D-Gal

(e)
Sn
ISC
S1
T1 S′n
ISC
Clustering S′1
S″n ISC
T′1 S″1
Conformation rigidification
RTP

T″1
FL

Cluster A Cluster B Cluster X


Subunits Clustered chromophores S0

Figure 13.16  (a) Photographs of different aqueous d-­Xyl solutions taken under a 365-­nm UV light at room
temperature. (b) Photographs of d-­Xyl crystals taken under various wavelengths of UV light or after stopping
the irradiation at room temperature. (c) Photographs of the crystals of PER, d-­Fru, and d-­Gal taken at room
temperature after stopping UV irradiation at different wavelengths. (d) Single-­crystal structure of d-­Xyl.
(e) CTE mechanism of color-­tunable nonconventional luminophores. Source: Adapted with permission from
Ref. [42] © 2020 The Royal Society of Chemistry Society.

polyanions, which made them nearly nonluminescent. After aggregating, the electrostatic charges
were screened with the overlap of electrostatic droplets; thereby, the polymer chains could
approach one another to induce clustering. More importantly, such a clustering would offer abun-
dant electronic interactions, like O⋯O, C=O⋯C=O, and H–O⋯C=O short contacts among n and
π electrons, leading to enhanced through-­space electronic conjugation along with rigidified con-
formations. At the same time, effective hydrogen bonds of SA also made positive effect on through-­
space conjugation and rigidifying conformations. Consequently, these clustered chromophores
were ready to be excited to produce visible emission under UV light. Besides fluorescence, triplet
excitons involving p-­RTP was also detected because of the promotion of ISC process and highly
rigidified conformation in the solid states.
13.3  ­Recent Progress in Organic RTP Material 393

COO– –OOC OH
–OOC
OH –OOC

(a) O O HO
O OH
O O O OH O HO O
OH O
OH O
–OOC
OH
O
O HO
COO– O OH
O
O
–OOC
OH OH OH OH
OH OH OH –OOC

OH OH G-block M-block M-block

G M GrM M/G = 0.69 GMr M/G = 1.53


Mw = 310 000, PDI = 2.1 Mw = 320 000, PDI = 2.1

(b) 0.005 0.05 0.5 1 2 3 4 6 8 wt% (d) Powders Film Film-Ca2+

6.4% 3.2% 3.5%

ϕ (%)0 3.9 UV on UV off UV on UV off UV on UV off

(c) 800 352


(e) 518
Conc. (wt%) 450
td = 0.1 ms
417 8

Normalized intensity (a.u.)


429
600 6
4
484 487
Intensity

3
400 2 td = 0ms
1
0.5 524
200 0.05
0.005
Water λex = 330 nm
0
320 370 420 470 520 570 620 360 410 460 510 560 610
Wavelength (nm) Wavelength (nm)

(f)
uv

Aggregation Crosslinking

Ca2+

Dilute solution Concentrated solution


or solid Emissive O– O
Nonemissive
O O
O–
– δ+C COO– HO
O– C
O HO
O
C C O O
O

C O
δ–

O
O– δ–
O

O

HO
C
H
O

O +
O δ
O–

HO COO–
O

+ Ca+
O

O– C δ δ– δ– –OCC
OH
O


O

OH
C

O O
O O
O

C – O
δ O
O–
H

C O OH
C

δ+
O

OH –OOC

–O
H

Through space electronic interactions n-π* interaction Dipole-dipole interaction

Figure 13.17  (a) Chemical structure of G and M units and typical chain sequence of SA; (b) photographs
taken under a 365-­nm UV light and (c) emission spectra (λex = 286 nm) of varying aqueous GMr solutions;
(d) photographs of solid powders, cast film, and Ca2+ cross-­linked film of GMr taken under a 312-­nm UV
light or after ceasing the UV irradiation; (e) PL spectra of GMr powders with td of 0 and 0.1 ms
(λex = 330 nm); (f) schematic illustration of SA molecules from isolated to aggregated states and possible
intra-­and intermolecular interactions within the clusters. Source: Adapted with permission from Ref. [43]
© 2018 American Chemical Society.
394 13  Room-­temperature Phosphorescence of Pure Organics

Generally, nonaromatic amino acids are believed to be nonemissive, owing to their lack of
remarkable conjugation in single molecules. However, considering CTE mechanism, it was rea-
sonable to speculate that these amino acids can emit light if various clusters of nonconventional
chromophores (e.g., amino, carbonyl, and hydroxyl) are formed. Based on this contradictions,
Chen et al. [44] checked the PL characteristics of a series of nonaromatic amino acids, such as l-­
lysine (l-­Lys), l-­serine (l-­Ser), l-­aspartic acid (l-­Asp), l-­asparagine (l-­Asn), l-­arginine (l-­Arg),
l-­cysteine (l-­Cys), l-­isoleucine (l-­lle), and glycine (Gly). The chemical structures, PL photographs
under a 365-­nm UV light, and emission efficiencies of these compounds are all shown in
Figure 13.18a. Evidently, all these solid powders exhibited fluorescence emission distributed from
deep blue to light green. To further illustrate the cause of luminescence, the single-­crystal struc-
ture of l-­Ser was examined as an example. Typically, abundant intermolecular interactions were
observed, resulting in the formation of a strong 3D intermolecular interaction network, on the one
hand, helping to rigidify molecular conformations and, on the other hand, more importantly, facil-
itating the through-­space electronic interactions, leading to extended electron delocalization and
possible visible-­light emission. Moreover, the poly(amino acid) (ε-­PLL) exhibited typical AIE fea-
tures and RTP property in solid powders with a Φ value of 7.9%, which could mainly be ascribed to
the CTE mechanism with highly rigid conformations. This work paved new ways for the design
and elaboration of a new class of luminescent biomolecules with promising applications in bio-
medical and optoelectronic fields.

13.3.3.2  Synthetic Compounds


Zhou et al. [45] investigated the photophysical properties of polyacrylonitrile (PAN), which further
confirmed the rationality of the CTE mechanism and deepened the understanding of these

(a) O O
O O
H2N HO H2N
OH HO OH OH OH
NH2 NH2 O NH2 O NH2

L-lysine (L-Lys) (7.4%) L-serine (L-Ser) (4.5%) L-aspartic acid (L-Asp) (0.1%) L-asparagine (L-Asn) (1.7%)

NH O O O O
H2N
H2N N OH HS OH OH OH
H
NH2 NH2 NH2 Glycine (Gly) (1.0%)
L-arginine (L-Arg) (0.5%) L-cysteine (L-Cys) (0.3%) L-isoleucine (L-lle) (3.6%)

(b) O (c)
H
N

ε-PLL NH2 n
0.001 0.01 0.1 0.2 0.5 1 2 2.5 5 7.5 10 15 UV on UV off

Figure 13.18  (a) Chemical structures of nonaromatic amino acids and photographs of them taken under a
365-­nm UV light; photographs of (b) different ε-­PLL aqueous solutions and (c) solid powders taken under a
365-­nm UV light or after ceasing the UV irradiation. Source: Adapted with permission from Ref. [44] © 2017
Science China Press and Springer-­Verlag Berlin Heidelberg.
13.3  ­Recent Progress in Organic RTP Material 395

nonconventional luminophores. As we know, PAN molecules are composed of saturated hydrocar-


bon backbones and electron-­rich cyano groups, without any conventional chromophores. As a
result, the PAN exhibited typical AIE properties (Figure 13.19a, c) with even DF and RTP in the
solid powders and thin films (Figure 13.19b). To understand such unique phenomena, after care-
fully checking the structure of PAN, the researchers deduced the existence of cyano clusters, which
contains several cyano groups. In dilute solutions, the cyano groups were isolated, being unfavora-
ble to be excited. When concentrated or aggregated, however, polymer chains entangled and
approached each other with cyano pendants clustered in close proximity (Figure 13.19d), provid-
ing through-­space electronic communications including electron overlap between lone pairs and
π electrons, dipole–dipole interactions and n–π interactions. As a consequence, the electronic con-
jugation was greatly enhanced, combining with rigidified conformations, which were certainly
beneficial to the light emission, especially for RTP emission. This work perfectly explains the lumi-
nescent phenomena of nonconjugated PAN molecules using CTE mechanism, which will arouse
more interest for the investigation of nonconventional luminophores.

Figure 13.19  (a) Photographs of PAN/DMF solutions with different concentrations taken at room
temperature under a 365-­nm UV light; (b) photograph of the solid powder and film of PAN taken under a
365-­nm UV light; (c) photographs of 1.25 × 10−5 M PAN in DMF and DMF/DCM mixtures with different fDCMs
taken at room temperature under a 365-­nm UV light; (d) schematic illustration of possible electronic
interactions in cyano clusters. Source: Adapted with permission from Ref. [45] © 2016 Wiley-­VCH Verlag
GmbH & Co. KGaA.
396 13  Room-­temperature Phosphorescence of Pure Organics

Furthermore, Zhou et al. [46] predicted that polymers with electron-­rich pendant groups, such
as polyacrylic acid (PAA), polyacrylamide (PAM), and poly(N-­isopropyl acrylamide) (PNIPAM),
could also emit light via strong through-­space electronic conjugation and interaction. To verify the
conjecture, several photophysical experiments were performed and the results are shown in
Figure 13.20. For PAA and PAM solids, p-­RTP emissions were detected, while the phosphorescence
of PNIPAM solids could only be observed under vacuum or in nitrogen atmosphere. These results
could also be well understood using the CTE mechanism (Figure 13.20c). Taking PAM as an exam-
ple, in dilute solutions, isolated chromophores with large energy gaps were hardly excited; thereby,
no emission could be observed even at 77 K. Moreover, in the concentrated solutions, clustering
occurs with effective through-­space electronic communications, such as C=O⋯N, n–π* and dipo-
lar interactions. The various clusters provide diverse energy levels of both singlets and triplets,
which may approach one another or even form an overlap, then making the ISC processes highly
possible, whereas the RTP emission is still difficult to achieve in solution owing to active molecular
motions and external quenching. In solid state, polymer chains of PAA and PAM were tightly
packed to block oxygen. Meanwhile, the interactions, including hydrogen bonds, made the cluster
chromophores rigid, which could significantly help to produce p-­RTP. However, for PNIPAM, due
to the steric hindrance resulting from the existence of the isopropyl group, the permeable oxygen
would quench the RTP emission, so that it could only be observed in 77 K or under vacuum.
Inspirited by previous studies, Zheng et al. [47] reckoned that the efficient RTP in nonconven-
tional luminophores might be achieved via the incorporation and clustering of carbonyl and het-
eroatoms alongside effective electronic interactions, like introducing acidamide groups.
Furthermore, replacing the vinyl group by amino (NH2) in acrylamide (i.e., urea) might afford RTP

(a) H (b) powders

O O
H O N O N PNIPAM PAA
H H
PAM 0 0.2 0.4 s
n n n films

PAA PAM PNIPAM


Mw: 33 200 31 600 22 500 PNIPAM PAA

PDI: 1.3 3.0 1.2 PAM 0 0.6 0.8 s

(c) O H
H
O NH2 N NH2
NH2 δ
O O O
O NH2 δ + δ+ O
O O NH2
O δ
O N NH2
NH2 H H
H2N NH2
Sn O
H2N C=O....N Dipolar interactions n-π*
O
S1
NH2
Sn Sn
Large ΔEST S′n S′n

S1 S′1 S1 S′1
Tn Tn Tn
T′n T′n
Large ΔE ISC ISC
T1 T′1 T1 T′1
T1
Nonrad.
Nonrad.

Nonrad.
Nonrad.
FL

FL

RTP

S0
Dilute solutions Concentrated solutions Solids

Figure 13.20  (a) Structures, Mw, PDI, and features of PAA, PAM, and PNIPAM; (b) photographs of different
powders and films taken under a 312-­nm UV light; (c) schematic illustration and Jablonski diagram of the
polymers (i.e., PAM) from isolated to aggregated states. Source: Adapted with permission from Ref. [46]
© 2019 The Royal Society of Chemistry and the Chinese Chemical Society.
13.3  ­Recent Progress in Organic RTP Material 397

in crystals with stronger intermolecular interactions and multiple hydrogen bonds. More impor-
tantly, the involvement of nitrogen with n electrons residing in the p orbital not only enhanced the
SOC and ISC processes but also formed a four-­center 6-­electron 46 conjugating system, which
would surely help to generate extended delocalization with through-­space conjugation. Besides,
replacing the oxygen in the carbonyl group by sulfur, such SOC constants and subsequent ISC
process could be well promoted due to the heavy atom effect. Consequently, the emission behav-
iors of urea and thiourea (TU) were investigated in detail. As illustrated in Figure  13.21b, urea
solutions were almost nonluminescent when the concentration was below 0.1 M. After the increase
in concentration, they gradually became emissive with the increased efficiency (Φ) and aggregate
sizes varying from approximately 0% and 84 nm (0.1 M) to 2.2% and 347 nm (8 M), which directly
proved the existence of clusters. After ceasing the irradiation, cyan afterglow lasting for several
seconds was observed in the urea crystal, corresponding to p-­RTP emission, whose lifetime and
efficiency were estimated to be 207 ms and 1.0%, respectively. Moreover, significantly enhanced PL
with an efficiency of 16.2% was achieved when the crystals were compacted into tablets
(Figure 13.21c). Then, the researchers obtained two polymorphs of TU (TU-­1 and TU-­2), which all
exhibited a bright green emission at 520 nm with relatively high efficiencies of 19.6% (TU-­1) and
11.6% (TU-­2). Similarly, distinctly boosted PL with efficiencies of 24.5% (TU-­1) and 16.1% (TU-­2)
were achieved via pressurization process, as depicted in Figure 13.21d. Such p-­RTP emission of

(a) O
(b) 30 25
Dm = 84 nm Dm = 347 nm
H2N 25 PDI = 1 20 PDI = 0.56
NH2
20
Intensity (%)

0.1 M 15 8M
15
S 10
NH2 10

H2N 5 5

0 0
1 10 100 101 102 103
(c) Size (nm)

UV on UV off 1 2 3 4 5 6 7 seconds UV on

(d) TU-1 (e) (f) 8


Sn
Excitation energy (eV)

Singlet
S1 7
Triplet
Sn*
Tn*
Difficult

19.6 24.5 S1* 6


Fluo.

T1* –1
Nonrad.

TU-2 5 <S1∣SOC∣T, = 45.9 cm


Phos.

4
S0
Individual <S1∣SOC∣T1 > = 150.6 cm–1
11.6 16.1 Clusters 3
molecules Urea TU-1

Figure 13.21  (a) Structures of urea and thiourea (TU); (b) DLS data and photographs taken under a
312-­nm UV light of 0.1 and 8 M urea solutions; (c) photographs of urea crystals (left) and tablet (right)
taken under a 312-­nm UV light or after ceasing the irradiation; (d) photographs of TU-­1, TU-­2, and their
tablets taken under a 312-­nm UV light; (e) schematic diagram of the CTE mechanism for urea and TU;
(f) energy levels and SOC for the monomers of urea and TU-­1. Source: Adapted with permission from
Ref. [47] © 2020 Wiley-­VCH Verlag GmbH & Co. KGaA.
398 13  Room-­temperature Phosphorescence of Pure Organics

urea and TU could be rationalized by the CTE mechanism. More importantly, the calculated SOC
(S1, T1) values of urea and TU were 45.9 and 150.6 cm−1, respectively, which were much higher
than those of reported conventional phosphors.
As mentioned above, the aromatic carbonyl is an ideal part when designing pure RTP organics
due to its effective ISC process. However, the RTP properties for luminophore with multiple non-
aromatic carbonyls are rarely investigated. Very recently, Zheng et  al.  [48] synthesized N,N -­
carbonylbissuccinimide (CBSI) and N,N -­oxalylbissuccinimide (OBSI) through acylation of
succinimide (SI) in mild conditions (Figure 13.22a). Both CBSI and OBSI were virtually nonlumi-
nescent in dilute solutions, while being brightly emissive in concentrated solutions or crystals,
with an efficiency of CBSI up to 9.0%, demonstrating typical AIE features. Moreover, CBSI single
crystals showed color-­tunable afterglows under different excitation sources, while that of OBSI
could only be found in 77 K (Figure 13.22c, d). To gain deeper insights into the AIE behaviors and
the mechanism of color-­tunable afterglow, single-­crystal structures of OBSI and CBSI were inves-
tigated. As depicted in Figure 13.22e, notably, both compounds were highly distorted in crystals,
which were greatly different from those common aromatic luminophores with large planar conju-
gation. Although large through-­bond conjugation failed to form in OBSI and CBSI, such effective
intermolecular interactions such as C–H⋯O=C, O=C⋯O=C, C–N⋯O=C, and O=C⋯C=O con-
tacts not only helped to rigidify the molecular conformations but also facilitated through-­space

(a) Cl
Cl O Cl
Cl O O O (b)
10–6 10–5 10–4 10–3 10–2 M
Cl O O Cl N N
CBSI

Et3N/DCM
O O O
CBSI
NH O
Cl O
O
Cl O O (d) UV on UV off 0 0.5 1 1.5 2 2.5 3 seconds
O N
N
(n-Hex)3N/DCM O O
O
CBSI

OBSI 312 nm

(c) CBSI 77 k OBSI 77 k


365 nm
312 nm

UV on UV off UV on UV off
OBSI

312 nm
365 nm

365 nm

(e) C–H O=C: 2.453, 2.639, 2.697


(f)
2.702, 2.714 HOMO LUMO HOMO LUMO
O–C O=C: 3.159

Monomer Trimer

C–H O=C: 2.528


C–N O=C: 2.949
O=C C=O: 3.193
O=C O=C: 2.880, 3.035 Dimer Tetramer
H=C O=C: 3.195, 3.111

Figure 13.22  (a) Synthetic routes and chemical structures of CBSI and OBSI; (b) photographs of CBSI/DMF
solutions taken under a 312-­nm UV light; (c) photographs of CBSI and OBSI single crystals taken under
312-­and 365-­nm UV lights or after ceasing the irradiation at 77 K; (d) photographs of CBSI and OBSI single
crystals taken under 312-­and 365-­nm UV lights or after ceasing the irradiation; (e) crystal structure (298 K)
and fragmental molecular packing of CBSI and OBSI; (f) electron densities of HOMO and LUMO levels for
monomer, dimer, trimer, and tetramer of CBSI. Source: Adapted with permission from Ref. [48] © 2020 Wiley-­
VCH Verlag GmbH & Co. KGaA.
13.3  ­Recent Progress in Organic RTP Material 399

conjugation among different moieties with n and π electrons. Therefore, the AIE behaviors could
be explained by restricting intramolecular motions in the aggregates, which significantly sup-
pressed nonradiative transition, thus resulting in bright emission. Furthermore, the presence of
C=O and nitrogen heteroatoms with through-­space conjugation could offer effective ISC process,
thus generating abundant triplets that were further stabilized by rigidified conformations, leading
to p-­RTP emission. In addition, such color-­tunable afterglow could be rationalized by CTE mecha-
nism and electron density calculation. As shown in Figure  13.22f, for CBSI, obvious electron
­delocalization was found in LUMO levels of several aggregate states, including monomer, dimer,
trimer, and tetramer. Thereby, these different types of aggregates would finally form clusters with
various emission abilities, corresponding to tunable p-­RTP when excited by different wavelengths.
This work provided a novel strategy for designing nonaromatic AIEgens with tunable p-­RTP; also,
it deepened the insight into the universal underlying mechanism of such color tunability for both
PL and p-­RTP.

13.3.4  Other Systems


13.3.4.1  Amorphous Organics
In 2016, Chen et al. [69] prepared three different phosphorescent polymers through radical binary
copolymerization, using acrylamide and RTP phosphors with various structures, as shown in
Figure  13.23a, where the polymer products were named as poly-­BrBA, poly-­BrNp, and poly-­
BrNpA. Owing to the variety of pendent phosphors, the emission colors could be adjusted from
light green (510, 520 nm) to orange (580 nm), corresponding to phosphorescent efficiencies of 11.4,
8.1, and 7.4% and lifetimes of 1.15, 5.76, and 5.08 ms, respectively. Moreover, the hydrogen bonding
between amide groups would form a potential 3D crosslinking network, which largely helped to
immobilize the RTP phosphors.
Later, Wang et  al.  [70] reported several amorphous lactam phosphorescent copolymers with
diverse pendent phosphors. Similarly, the emission color of copolymers can be tuned from deep
blue to green, orange and to red. The rigid polymer matrices of poly(N-­vinylpyrrolidone) (PVP)
and poly(N-­vinylcaprolactam) (PNVCL) effectively helped to suppress nonradiative transitions of
phosphors, which significantly enhanced the RTP emission.
Although copolymers with pendent phosphors and other groups offering crosslinking interac-
tions could well achieve the RTP emission, the rotation of pendant phosphors still existed. Kwon
et al. [71] utilized a covalent crosslinking between phosphors and polymer matrices via the Diels–
Alder reaction to maximally restrict molecular motions to promote phosphorescent efficiencies, as
shown in Figure 13.23c. As a result, the Φp of DA1 could reach 26.0%, which was much higher than
the analogous phosphor-­doped polymer system of 15.6%.
Ma et al. [72] also reported a series of amorphous polymers with crosslinked networks between
the polymeric chains. As illustrated in Figure 13.23d and e, taking P3 as an example, the lifetime
of it was 537 ms, with an appreciable quantum yield of 15.39%. Such a facile approach would be
promising for constructing next-­generation amorphous RTP materials applied in molecular imag-
ing, water sensing, and data security.

13.3.4.2  Organic Framework


Recently, researches on organic framework materials are emerging rapidly due to their unique
structures, complex assembling procedures, and potential application in various fields. From the
point of view for designing high-­performance RTP organics, the organic frameworks can surely
provide a rigid environment to shield quenchers, like oxygen, and stabilize both singlet and triplet
400 13  Room-­temperature Phosphorescence of Pure Organics

(a)
o m n m m
o n o n
NH o o NH o
H2N NH H2N H2N

o
o o N o

o Br
Br Φp =11.4% Φp = 8.1% Φp = 7.4%
Br
Poly-BrBA Poly-BrNp Poly-BrNpA

(b) AIBN (c) Ene Phosphorescent Ene


core
Diene Polymer
N
m n unit unit unit backbone
o R N
k EtOH, 60°C
k
H O
O
O n
m n m n m n O
N O O
N N N
o O NH O O N
O O
O O
O Br O

O N O Br DA1 PFMA

COOH
Diels–Alder, 120 °C
O
Br
PVP-BrNpA PVP-BA PVP-BrHexene O

m
O
H O O
m n m n m n O
o N O N O N O O O

n
O NH
O N
O O
N O
O O O
n

O Br
O N O O Br
O
O
Cross–linked DA1
m

COOH O
Br
O
PNVCL-BrNpA PNVCL-BA PNVCL-BrHexene

(d) H2N O (e) UN ON OFF 0.5 s 1.0 s 1.5 s 2.0 s 2.5 s 3.0 s 3.5 s 4.0 s
n
m P1
O O O O
COOH NH2
EDCI/DMAP
P2
+ O
CH2Cl2
P3
AIBN, DMF. 65 °C
COOH OH
P3
O O O O

P4
n m
O NH2 P5
O O O O
O O
O P6
O
COOH COOH O
O O O O P7
1 2 3 4 5

O
P8
O O
O
P9
CHO COOH
COOH OH
O
P10
6 7 8 9 10

Figure 13.23  (a) Chemical structures and photographs of poly-­BrBA, poly-­BrNp, and poly-­BrNpA under UV
light. Source: Adapted with permission from Ref. [69] © 2016 Wiley-­VCH Verlag GmbH & Co. KGaA;
(b) synthetic routes and chemical structures of PVP-­BrNpA, PVP-­BA, PVP-­BrHexene, PNVCL-­BrNpA,
PNVCL-­BA, and PNVCL-­BrHexene. Colors are according to their phosphorescence spectra. Source: Adapted
with permission from Ref. [70] © 2019 Wiley-­VCH Verlag GmbH & Co. KGaA; (c) chemical structures of
designed phosphors and polymers with the description of the strategy. Source: Adapted with permission
from Ref. [71] © 2015 Springer Nature; (d) synthetic routes of the polymers (P) and chemical structures of
compounds 1–10; (e) photographs of P1–P10 powders under a 254-­nm UV light and after the removal of
excitation. Source: Adapted with permission from Ref. [72] © 2018 Wiley-­VCH Verlag GmbH & Co. KGaA.
13.3  ­Recent Progress in Organic RTP Material 401

excitons via suppressing nonradiative transitions. For instance, Mieno et al. [36] reported host–
guest systems based on metal–organic frameworks (MOFs) and small organic molecule (coro-
nene). MOFs were crystalline materials composed of metal ions and organic ligands with excellent
thermal and chemical stabilities and ability to encapsulate guest species in their cavities. As shown
in Figure 13.24a, the commercially available MOF material (ZIF-­8) was employed as a rigid matrix
and coronene was introduced as the guest molecule. Generally, coronene is a typical ACQ

(a) (b)

380 K

BIPA 340 K
300 K
260 K
Zn2+ Zn5
220 K
180 K
Zn-cluster 140 K
Afte
based MOFs
excitation
Zn3

Under UV 0s 15 s 40 s

(c) HO OH
O OH
+ HO
B
B
OH
HO OH HO O
HO OH BZLBA
HTTP

Layer distance = 3.7 Å


Dio⊂BZL-COF H-aggregation

Dio⊂BZL-COF

Activation

Layer distance = 3.4 Å


BZL-COF BZL-COF H-aggregation
(d) (e) 1.0
Intralayer
Normalized intensity (a.u.)

π-π interaction 0.8


UV on UV off
0.6 PHTCz-1
Hydrogen bond O N PhTCz-2
0.4 PhTCz-3
N O
O N 0.2
× 20
0.0 × 20
HOAFs PhTCz HOAFs 400 500 600 700
Wavelength (nm)

Figure 13.24  (a) Structure of coronene@ZIF-­8 and its p-­RTP photographs before and after removing the
UV excitation. Source: Adapted with permission from Ref. [36] © 2016 Wiley-­VCH Verlag GmbH & Co. KGaA;
(b) schematic illustration of temperature-­responsive p-­RTP cluster-­based metal–organic frameworks.
Source: Adapted with permission from Ref. [37] © 2018 American Chemical Society; (c) synthesis routes of
Dio & BZL-­COF and BZL-­COF. Chemical structures, top views, and side views of Dio & BZL-­COF and
BZL-­COF. Source: Adapted with permission from Ref. [38] © 2018, The Royal Society of Chemistry;
(d) schematic illustration of typical hydrogen-­bonded organic frameworks and hydrogen-­bonded aromatic
frameworks (HOAFs) based on PhTCz molecules. (e) PL (dashed lines) and phosphorescence spectra (solid
lines) of PhTCz-­1, PhTCz-­2, and PhTCz-­3 under a 365-­nm UV light at room temperature. Source: Adapted
with permission from Ref. [39] © 2018, Wiley-­VCH Verlag GmbH & Co. KGaA.
402 13  Room-­temperature Phosphorescence of Pure Organics

molecule with unstable singlet and triplet excitons in the solid state. After adding to ZIF-­8, thanks
to the rigidified conformation, the powder of coronene@ZIF-­8 was emissive under UV irradiation
and even detected a bright green-­yellow afterglow lasting for 40 seconds, which was never reported
before. Later, Li et al. [37] explored the construction of cluster-­based MOFs, which paves an effec-
tive way to modulate the fluorescence and RTP emission based on tunable π–π stacking, halogen-­
bonding interaction, and metal-­cluster units. In comparison to the original metal-­free ligand
(5-­bromoisophthalic acid), the Zn5 and Zn3 cluster-­based MOFs exhibit tunable photolumines-
cence. More surprisingly, Zn5 cluster-­based MOFs even showed an obvious temperature-­dependent
luminescence from 140 to 380 K, which could be served as a type of temperature-­responsive phos-
phorescent switch.
Moreover, except for well-­studied MOFs, covalent organic frameworks (COFs) also show abun-
dant interest recently owing to their metal-­free structures, low density, and ease of functionaliza-
tion for a wider range of applications. In 2018, Wang et al. [38] constructed a two-­dimensional
COF (BZL-­COF) by cocondensation of BZLBA and HHTP as the single layer and H-­aggregation for
further layer-­by-­layer assembling, as depicted in Figure 13.24c. After removing solvent molecules,
the layer distance decreased from 3.7 to 3.4 Å, indicating stronger interactions between different
layers to hamper nonradiative transition. Moreover, by incorporating with CIP-­active BZLBA
units, the RTP emission was realized in BZL-­COF. Such COFs provided an ideal research platform
for investigating the RTP behaviors related to ordered intermolecular and polymeric packings.
Cai et al. [39] reported novel hydrogen-­bonded organic aromatic frameworks (HOAFs) based on
PhTCz and were further constructed via intralayer π–π interactions. Three types of HOAFs, namely,
PhTCz-­1, PhTCz-­2, and PhTCz-­3, were obtained with tetragonal pores, homogeneous hexagonal
pores, and hexagonal pores, respectively, containing a chloroform molecule. Among three types of
HOAFs, only PhTCz-­1 crystals demonstrated a yellow afterglow with a lifetime of 79.8 ms at room
temperature, while prolonged RTP of PhTCz-­2 and PhTCz-­3 could only be observed in a N2 atmos-
phere on account of the different pore sizes and contents of oxygen quenchers. This work provided
a platform for expanding the application of a metal-­free porous material, as listed in
Figure 13.24d and e.

13.3.4.3  Supramolecular Organics


Bian et al. [40] designed and synthesized a new class of ultralong organic RTP materials with high
efficiency through self-­assembly of melamine (MA) and aromatic acids (IPA) in aqueous solutions
(Figure 13.25a). Thanks to abundant hydrogen bonds between MA and IPA, and multiple intermo-
lecular interactions (Figure 13.25c), the supramolecular 3D network (MA–IPA) could be formed
with a rigid environment to lock the molecular rotation, resulting in a limited nonradiative decay
of the triplet excitons. Moreover, the carboxyl groups of IPA and effective through-­space conjuga-
tion would promote ISC process. Combining these two points, the MA–IPA synchronously
achieved an ultralong phosphorescent lifetime of up to 1.91 seconds and a high phosphorescent
quantum efficiency of 24.3% at room temperature. Such a designing strategy of supramolecular
RTP material well balanced the conflict between lifetime and efficiency, which might inspire a
series of further studies.
Kuila et al. [41] developed a unique supramolecular hybrid assembly between inorganic laponite
(LP) clay and bromine-­substituted naphthalene diimide (BrNDI) phosphor with a deep red RTP
emission in water and in amorphous film state. This work provided two principles for designing
such a clay–organic hybrid supramolecular material. Firstly, introducing a heavy atom to phosphor
could enhance SOC constant, leading to more effective ISC processes and more generated triplet
excitons. Then, the structural rigidity and oxygen impenetrability of inorganic clay could help to
13.3  ­Recent Progress in Organic RTP Material 403

(a) (b) UV on
(c)

MA
100 μm 100 μm

UV off
IPA
100 μm 100 μm
H2O

(d) 1.0
I
N
I
N (e) No LP
With LP

O N O O N O No LP
Br 0.5
Iem
With LP
Br Laponite (LP)
O N O O N O O
Mg 0.0
25 nm Si 400 480 560 640 720
N I N I Li λ(nm)
pNDI BrNDI H
0.

1.0
(f)
9n

Steady state
m

Gated

pNDI or BrNDI monomers


Water 0.5 No LP
Iem
Co-assembly

or With LP

pNDI-LP BrNDI-LP 0.0


Exfoliated LP 400 480 560 640 720
(excimer) (phosphorescence)
λ(nm)

Figure 13.25  (a) Schematic illustration of supramolecular architecture based on MA and IPA;


(b) photographs of the MA–IPA supramolecular framework with the excitation on and off at room
temperature; (c) single-­crystal structure of MA–IPA supermolecule. Source: Adapted with permission from
Ref. [40] © 2018 American Chemical Society. (d) Chemical structures of pNDI, BrNDI, and laponite (LP).
Preparation routes of pNDI-­LP and BrNDI-­LP; (e) normalized emission spectra of BrNDI-­LP in water excited
by a 380-­nm UV light. (f) Steady-­state and gated emission spectra of BrNDI-­LP excited by a 380-­nm UV
light. Source: Adapted with permission from Ref. [41] © 2018 Wiley-­VCH Verlag GmbH & Co. KGaA.

control molecular rotation and stabilized triplet excitons, which was also crucial for ultimate RTP
emission, as illustrated in Figure 13.25e and f.

13.3.4.4  Hybrid Perovskites


Organic–inorganic hybrid perovskites, incorporating organic and inorganic semiconductors at the
molecular level, have attracted considerable attention recently owing to their advantages of both
high carrier mobility and low-­temperature solution processing. From the principle of designing
RTP materials, organic hybrid perovskites perfectly utilize the rigid framework of inorganic per-
ovskites and emission abilities of doped organic cations. It is reasonable to speculate that once the
generated triplet excitons of organic cations are stabilized by perovskite structures, the possible
phosphorescence emission, even RTP emission, may subsequently occur.
For example, Hu et al. [73] synthesized a 2D-­layer perovskite with alternating [PbX6]−4 octahe-
dra network layers and organic molecular layers. After adjusting the types of organic cations, the
researchers firstly obtained three hybrid perovskite films containing TTMA (TPB), PEA (PPB),
and mixed-­cation (PTPB), as shown in Figure 13.26a–c. Both TPB and PTPB exhibited RTP emis-
sion with lifetimes of 3.04 and 1.25 ms, respectively. Meanwhile, the RTP efficiencies of PTPB
were as high as 11.2% while only being below 1% of TPB. More interestingly, PPB only showed
404 13  Room-­temperature Phosphorescence of Pure Organics

(a) (b) (c) PPB PTPB TPB


S

S Ambient light
NH3+
NH3+

UV light

(d) S
(e) (f)
S CB
NH3+
ET
TTPHMA NH3+
T1
NMA
H3C Fluo.
O Phos. hv
H3C S knr
kp

S0
VB
NH3+ NH3+
MeO-NMA MeT-BA Inorganic TTMA

(g) (h) 5%
– PEPB-NIA5
PEPB-NIA10
NH3 X X = Br, PEAB PEPB-NIA20
Cl, PEAC Pb2+
10%

Intensity
X–
20%

NH3+
Self-assembly

NH3 X PbX2 500 550 600 650 700
400 450
X = Br, NIAB X = Br, PEPB-NIA Wavelength (nm)
Cl, PEPC-NIA Turn off
O N O
Cl, NIAC (i) Turn on

0 second 4 second

Figure 13.26  (a) Schematic illustration of the layered hybrid perovskite (TTMA)2PbBr4 (TPB) with TTMA
cation; (b) schematic illustration of PEA-­based perovskite (PEA)2PbBr4 (PPB) and mixed-­cation perovskite
(PEA1−xTTMAx)2PbBr4 (PTPB); (c) photographs of perovskite thin films under ambient light and UV light;
(d) chemical structure of organic cations for perovskite; (e) photographs of perovskite thin films with
different doped cations and their photophysical properties; (f) mechanism of the optical excitation, energy
transfer, and relaxation paths of excited states in TPB. Source: Adapted with permission from Ref. [73]
© 2018 Wiley-­VCH Verlag GmbH & Co. KGaA; (g) synthesis and structural illustration of the doped
perovskites; (h) photographs under UV light and PL spectra of PEPB-­NIA; (i) photographs of PEPC-­NIA5
powders under UV light and after removal of irradiation. Source: Adapted with permission from [74] © 2018
The Royal Society of Chemistry.

fluorescent emission with an efficiency of 31%. Such obvious difference could be mainly ascribed
to the aggregated morphology of organic cations in the perovskite framework. Specifically, the
tight packing of single-­component cations would lead to phosphorescent quenching due to the
triplet energy transfer or triplet–triplet annihilation (TTA), while hybrid perovskite with mixed-­
cation could perfectly avoid this risk and emit RTP with high efficiency. Moreover, the researchers
revealed the luminous nature of hybrid perovskite, taking TPB as an example in Figure 13.26f.
The inorganic perovskite would be easily excited under irradiation, and excitons jumped from
valence band (VB) to conduction band (CB), while after the energy transfer, the triplet excitons
would decay with phosphorescence and other nonradiative transition. Such a photophysical pro-
cess was quite different from the classical ones, and the matching degree of energy levels was
essential. Furthermore, the researchers doped many other organic cations with various
13.4  ­Conclusions and Perspective 405

chromophores to perovskite, thus affecting their emission colors from blue to red, and the results
are listed in Figure 13.26d and e.
Analogously, Yang et al. [74] doped a common fluorophore (1,8-­naphthalimide [NI]) into PEA
based-­perovskites with various contents and mixed cations to construct organic/inorganic hybrid.
As depicted in Figure 13.26i, PEPC-­NIA5 powders exhibited a white emission under UV light and
an obvious yellow afterglow with an efficiency of 56.1% and a lifetime of 35 ms. Such a unique
performance will meet the application in the field of information technology as security ink and
afterglow LEDs as single luminescent materials.

13.3.5 Applications
Such materials with RTP properties are considered to possess wide applications in many fields,
including anticounterfeiting [3, 4, 75], oxygen detection [7], bioimaging [8, 9], OLEDs [1, 2, 76],
and mechanoluminescent devices [77–79]. For example, Gu et al. [75] found that the prolonged
irradiation of PCzT, BCzT, and FCzT under ambient conditions can activate their ultralong RTP
emission with the lifetimes spanning from 1.8 to 1330 ms. These phosphors can also be deactivated
back to their original states with short-­lived phosphorescence by UV irradiation for three hours at
room temperature or under thermal treatment. Such dynamic RTP emission was applied for a
visual anticounterfeiting application with transformable patterns, like “8,” “H,” “E,” “C,” “I,” and
“11,” as illustrated in Figure 13.27a, which has expanded the scope of smart-­response materials.
Moreover, considering that triplet excitons of pure organics are highly sensitive to oxygen, Zhou
et al. [7] developed a facile strategy for detecting oxygen based on dynamic RTP quenching mecha-
nism in texbrite BBU-­L (TBBU)-­doped films. As shown in Figure 13.27b, obviously, the afterglow
intensity of the doped film is highly susceptible to oxygen, even the RTP lifetimes of which exhibit
excellent linear relation with oxygen fraction. For instance, the RTP lifetime of the doped film
exposed to 2% oxygen is determined to be 103 ms, corresponding to a calculated value of 2.08%
based on the working curve, very close to the real oxygen fraction (2%). Besides, Fateminia et al. [9]
synthesized a red-­emissive RTP molecule (C-­C4-­Br) with a typical alkoxy spacer. The C-­C4-­Br
nanocrystals are applied for the bioimaging of 231 breast cancer cells, and most of them could be
well marked under a confocal laser scanning microscopy (CLSM; Figure 13.27c). Recently, Wang
et  al.  [76] fabricated an OLED with a 10  wt% 2,6-­di(phenothiazinyl)naphthalene (DPTZN)  :
­triazine–benzimidazole (TRZ–BIM) blend film as the emitting layer, which shows a high ­maximum
external quantum efficiency of 11.5%, a current efficiency of 33.8 cd/A, and a power efficiency of
32.6  lm/W. This work truly expands the application of RTP materials and also inspires more
­similar researches.

13.4  ­Conclusions and Perspectives

In summary, with the growing interest in pure organic RTP luminophores due to their fundamen-
tal importance and promising applications, various significant advances have been achieved in the
recent years. Based on the classical theories, the triplets in pure organics are quite unstable and
easily quenched by oxygen and high temperature due to enhancement of molecular collision.
Therefore, to achieve p-­RTP property, two aspects of approaches are realized. One is to accelerate
ISC process via the increment of SOC to ensure the large formation of triplet states, while the other
is to suppress nonradiative transition through crystallization, clustering, organic framework
(including supramolecular systems), and host–guest doping. Molecules in crystal or many other
406 13  Room-­temperature Phosphorescence of Pure Organics

(a) (b) UV flashlight

N Compd. es 1.5 Cell phone


nut ho
mi urs vacuum

Photo-activation region
–OCH3 MCzT 10
N N

Deactivation region
–OC2H5 ECzT
Doped films
N –OC3H7 PCzT N2

–OC4H9 BCzT
–OC5H11 FCzT

2.0 hours
5 minutes
b a: MCzT UV on
a d
b: PCzT
c
c: BCzT
a d UV off 3.
d: FCzT tes 0h
b inu our
s
1m

Pattern model

O
(c) O (d)
Purely organic phoshorescent
N Br
N
O Br 101
Spacer unit
S N EQEmax = 11.5%
C-Br C-C4-Br N S

EQE (%) 100

Radiative
EL
10–1 Transition

T1 S0
500 600 700
10–2
100 101 103 102
104
–2
Luminance (cd/m )

Figure 13.27  (a) Demonstration of multilevel anticounterfeiting using the MCzT, PCzT, BCzT, and FCzT
crystals. Source: Adapted with permission from Ref. [75] © 2018 Wiley-­VCH Verlag GmbH & Co. KGaA;
(b) illustration of oxygen detection using the TBBU-­doped films at various oxygen fractions. Source: Adapted
with permission from Ref. [7] © 2019 Wiley-­VCH Verlag GmbH & Co. KGaA; (c) CLSM images of breast
cancer cells incubated with nanocrystals of C-­C4-­Br. Source: Adapted with permission from Ref. [9]
© 2017 Wiley-­VCH Verlag GmbH & Co. KGaA; (d) schematic diagram of organic light-­emitting diode (OLED)
based on phosphorescent DPTZN. Source: Adapted with permission from Ref. [76] © 2019 American
Chemical Society.

aggregate states are more accessible to form stable triplet excitons, for which RTP along with the
particular AIE feature will be greatly benefited. In this chapter, we comprehensively review the
systems with crystallization, doping in a rigid matrix and others induced RTP, which are widely
found in conventional luminophores. When considering those RTP materials without conven-
tional chromophores, clustering-­triggered emission mechanism is proposed to rationalize the
emission phenomena.
However, for pure organics, firstly, the essence of their RTP emission is still unclear owing to the
complexity of their photophysical processes and plenty of unpredictable factors, like aggregating
types, impurities, and so on, especially for nonconventional luminophores. Although the CTE
mechanism can well explain their RTP behaviors, the universal rules to summarize all these phe-
nomena are remained to be further explored. In addition, more efforts should be made to achieve
long phosphorescent lifetimes associated with high quantum efficiency for pure organics; also,
multicolor emission is ungently desired. As we know, pure organic RTP materials have found their
wide applications in biological imaging, temperature, water or ion sensing, and information secu-
rity. Recently, many stimuli-­responsive materials have sprung up with mechanoluminescence,
  ­Reference 407

which can be applied in smart devices and even detect the human heartbeat [77]. In a word, fur-
ther researches in this area should not only reveal more underlying fundamentals but also offer
new pathways for more intriguing applications.

­References

1 Reineke, S. (2014). Organic light-­emitting diodes: phosphorescence meets its match. Nat.
Photonics, 8(4), 269.
2 Hirata, S., Shizu, K. (2016). Organic light-­emitting diodes: high-­throughput virtual screening. Nat.
Mater., 15(10), 1056.
3 Jiang, K., Wang, Y., Cai, C., Lin, H. (2018). Conversion of carbon dots from fluorescence to
ultralong room-­temperature phosphorescence by heating for security applications. Adv. Mater.,
30(26), 1800783.
4 Su, Y., Phua, S. Z. F., Li, Y. et al. (2018). Ultralong room temperature phosphorescence from
amorphous organic materials toward confidential information encryption and decryption. Sci.
Adv., 4(5), eaas9732.
5 Su, Q., Lu, C., Yang, X. (2019). Efficient room temperature phosphorescence carbon dots:
information encryption and dual-­channel pH sensing. Carbon, 152(152), 609.
6 Huang, L., Chen, B., Zhang, X. et al. (2018). Proton-­activated “off–on” room-­temperature
phosphorescence from purely organic thioethers. Angew. Chem. Int. Ed., 57(49), 16046.
7 Zhou, Y., Qin, W., Du, C. et al. (2019). Long-­lived room-­temperature phosphorescence for visual
and quantitative detection of oxygen. Angew. Chem. Int. Ed., 58(35), 12102.
8 Ni, F., Zhu, Z., Tong, X. et al. (2018). Organic emitter integrating aggregation-­induced delayed
fluorescence and room-­temperature phosphorescence characteristics, and its application in
time-­resolved luminescence imaging. Chem. Sci., 9(9), 6150.
9 Fateminia, S. M. A., Mao, Z., Xu, S. et al. (2017). Organic nanocrystals with bright red persistent
room-­temperature phosphorescence for biological applications. Angew. Chem. Int. Ed., 56(40), 12160.
10 Pomestchenko, I. E., Luman, C. R., Hissler, M. et al. (2003). Room temperature phosphorescence
from a platinum(II) diamine bis(pyrenylacetylide) complex. Inorg. Chem., 42(5), 1394.
11 Maurya, Y. K., Ishikawa, T., Kawabe, Y. et al. (2016). Near-­infrared phosphorescent Iridium(III)
benzonorrole complexes possessing pyridine-­based axial ligands. Inorg. Chem., 55(12), 6223.
12 Chu, W. K., Wei, X. G., Yiu, S. M. et al. (2015). Strongly phosphorescent neutral rhenium(I)
isocyanoborato complexes: synthesis, characterization, and photophysical, electrochemical, and
computational studies. Chem. Eur. J., 21(6), 2603.
13 Ma, H., Lv, A., Fu, L. et al. (2019). Room-­temperature phosphorescence in metal-­free organic
materials. Ann. Phys. (Berlin), 531(7), 1800482.
14 Yuan, W. Z., Shen, X., Zhao, H. et al. (2010). Crystallization-­induced phosphorescence of pure
organic luminogens at room temperature. J. Phys. Chem. C, 114(13), 6090.
15 Zhao, W., He. Z., Lam, J. W. Y. et al. (2016). Rational molecular design for achieving persistent and
efficient pure organic room-­temperature phosphorescence. Chem., 1(4), 592.
16 Gong, Y., Zhao, L., Peng, Q. et al. (2015). Crystallization-­induced dual emission from metal and
heavy atom-­free aromatic acids and esters. Chem. Sci., 6(6), 4438.
17 Gong, Y., Chen, G., Peng, Q. et al. (2015). Achieving persistent room temperature phosphorescence
and remarkable mechanochromism from pure organic luminogens. Adv. Mater., 27(40), 6195.
18 Yang, Z., Mao, Z., Zhang, X. et al. (2016). Intermolecular electronic coupling of organic units for
efficient persistent room-­temperature phosphorescence. Angew. Chem. Int. Ed., 55(6), 2181.
408 13  Room-­temperature Phosphorescence of Pure Organics

19 He, Z., Zhao, W., Lam, J. W. Y. et al. (2017). White light emission from a single organic molecule
with dual phosphorescence at room temperature. Nat. Commun., 8(1), 416.
20 Bolton, O., Lee, K., Kim, H. J. et al. (2011). Activating efficient phosphorescence from purely
organic materials by crystal design. Nat. Chem., 3(3), 205.
21 Kwon, M. S., Lee, D., Seo, S. et al. (2014). Tailoring intermolecular interactions for efficient
room-­temperature phosphorescence from purely organic materials in amorphous polymer
matrices. Angew. Chem. Int. Ed., 53(42), 11177.
22 Shimizu, M., Kimura, A., Sakaguchi, H. (2016). Room-­temperature phosphorescence of crystalline
1,4-­bis(aroyl)-­2,5-­dibromobenzenes. Eur. J. Org. Chem., 2016(3), 467.
23 Mao, Z., Yang, Z., Mu, Y. et al. (2015). Linearly tunable emission colors obtained from a
fluorescent-­phosphorescent dual-­emission compound by mechanical stimuli. Angew. Chem. Int.
Ed., 54(21), 6270.
24 Shimizu, M., Nagano, S., Kinoshita, T. (2020). Dual emission from precious metal-­free
luminophores consisting of C, H, O, Si, and S/P at room temperature. Chem. Eur. J.,
26(23), 5162.
25 Shimizu, M., Shigitani, R., Nakatani, M. et. al. (2016). Siloxy group-­induced highly efficient room
temperature phosphorescence with long lifetime. J. Phys. Chem. C, 120(21), 11631.
26 Gong, Y., Tan, Y. Li, H. et al. (2013). Crystallization-­induced phosphorescence of benzils at room
temperature. Sci. China Chem., 56(9), 1183.
27 Shen, Q. J., Pang, X., Zhao, X. R. et al. (2012). Phosphorescent cocrystals constructed by
1,4-­diiodotetrafluorobenzene and polyaromatic hydrocarbons based on C–I⋯π halogen bonding
and other assisting weak interactions. Crystengcomm., 14(15), 5027.
28 Pang, X., Wang, H., Wang, W. et al. (2015). Phosphorescent π-­hole⋯π bonding cocrystals of pyrene
with halo-­perfluorobenzenes (F, Cl, Br, I). Cryst. Growth Des., 15(10), 4938.
29 Hirata, S., Totani, K., Zhang, J. et al. (2013). Efficient persistent room temperature phosphorescence
in organic amorphous materials under ambient conditions. Adv. Funct. Mater., 23(27), 3386.
30 Hirata, S., Vacha, M. (2017). White afterglow room-­temperature emission from an isolated single
aromatic unit under ambient condition. Adv. Optical Mater., 5(5), 1600996.
31 Lei, Y., Dai, W., Tian, Y. et al. (2019). Revealing insight into long-­lived room-­temperature
phosphorescence of host-­guest systems. J. Phys. Chem. Lett., 10 (20), 6019.
32 Cao, J., Ma, X., Min, M. et al. (2014). INHIBIT logic operations based on light-­driven β-­cyclodextrin
pseudo[1]rotaxane with room temperature phosphorescence addresses. Chem. Commun.,
50(24), 3224.
33 Li, D., Lu, F., Wang, J. et al. (2018). Amorphous metal-­free room-­temperature phosphorescent
small molecules with multicolor photoluminescence via a host-­guest and dual-­emission strategy.
J. Am. Chem. Soc., 140(5), 1916.
34 Chen, H., Xu, L., Ma, X. (2016). Room temperature phosphorescence of 4-­bromo-­1,8-­naphthalic
anhydride derivatives based polyacrylamide copolymer with photo-­stimulated responsiveness.
Polym. Chem., 7(7), 3989.
35 Su, Y., Zhang, Y., Wang, Z. et al. (2020). Excitation-­dependent long-­life luminescent polymeric
systems under ambient conditions. Angew. Chem. Int. Ed., 59(25), 9967.
36 Mieno, H., Kabe, R., Notsuka, N. et al. (2016). Long-­lived room-­temperature phosphorescence of
coronene in zeolitic imidazolate framework ZIF-­8. Adv. Optical Mater., 4(7), 1015.
37 Li, D., Yang, X., Yan, D. (2018). Cluster-­based metal-­organic frameworks: modulated singlet-­triplet
excited states and temperature-­responsive phosphorescent switch. ACS Appl. Mater. Interfaces,
10(40), 34377.
  ­Reference 409

38 Wang, S., Ma, L., Wang, Q. et al. (2018). Covalent organic frameworks: a platform for the
experimental establishment of the influence of intermolecular distance on phosphorescence.
J. Mater. Chem. C, 6(6), 5369.
39 Cai, S., Shi. H., Zhang, Z. et al. (2018). Hydrogen-­bonded organic aromatic frameworks for
ultralong phosphorescence by intralayer π–π interactions. Angew. Chem. Int. Ed., 57(15), 4005.
40 Bian, L., Shi, H., Wang, X. et al. (2018). Simultaneously enhancing efficiency and lifetime of ultralong
organic phosphorescence materials by molecular self-­assembly. J. Am. Chem. Soc., 140(34), 10734.
41 Kuila, S., Rao, K. V., Garain. S. et al. (2018). Aqueous phase phosphorescence: ambient triplet
harvesting of purely organic phosphors via supramolecular scaffolding. Angew. Chem. Int. Ed.,
57(52), 17115.
42 Zhou, Q., Yang, T., Zhong, Z. et al. (2020). A clustering-­triggered emission strategy for tunable
multicolor persistent phosphorescence. Chem. Sci., 11(11), 2926.
43 Dou, X., Zhou, Q., Chen, X. et al. (2018). Clustering-­triggered emission and persistent room
temperature phosphorescence of sodium alginate. Biomacromolecules, 19(6), 2014.
44 Chen, X., Luo, W., Ma, H. et al. (2018). Prevalent intrinsic emission from nonaromatic amino acids
and poly(amino acids). Sci. China Chem., 61(3), 351.
45 Zhou, Q., Cao, B., Zhu, C. et al. (2016). Clustering-­triggered emission of nonconjugated
polyacrylonitrile. Small, 12(47), 6586.
46 Zhou, Q., Wang, Z., Dou, X. et al. (2019). Emission mechanism understanding and tunable
persistent room temperature phosphorescence of amorphous nonaromatic polymers. Mater. Chem.
Front., 3(3), 257.
47 Zheng, S., Hu, T., Bin, X. et al. (2020). Clustering-­triggered efficient room-­temperature
phosphorescence from nonconventional luminophores. ChemPhysChem, 21(1), 36.
48 Zheng, S., Zhu, T., Wang, Y. et al. (2020). Accessing tunable afterglows from highly twisted
nonaromatic organic AIEgens via effective through-­space conjugation. Angew. Chem. Int. Ed.
59(25), 10018.
49 Kasha, M. (1947). Phosphorescence and the role of the triplet state in the electronic excitation of
complex molecules. Chem. Rev., 41(2), 401.
50 Kasha, M. (1950). Characterization of electronic transitions in complex molecules. Discuss.
Faraday Soc., 9(9), 14.
51 Baryshnikov, G., Minaev, B., Ågren, H. (2017). Theory and calculation of the phosphorescence
phenomenon. Chem. Rev., 117 (9), 6500.
52 Gao, X., Bai, S., Fazzi, D. et al. (2017). Evaluation of spin–orbit couplings with linear-­response
time-­dependent density functional methods. J. Chem. Theory Comput., 13(2), 515.
53 El-­Sayed, M. A., Leyerle, R. (1975). Low field Zeeman effect and the mechanism of the
S1→T1 nonradiative process. J. Chem. Phys., 62(4), 1579.
54 Hirata, S. (2017). Recent advances in materials with room-­temperature phosphorescence:
photophysics for triplet exciton stabilization. Adv. Optical Mater., 5(17), 1700116.
55 Uoyama, H., Goushi, K., Shizu, K. et al. (2012). Highly efficient organic light-­emitting diodes from
delayed fluorescence. Nature, 492(7428), 234.
56 Yang, Z., Mao, Z., Xie, Z. et al. (2017). Recent advances in organic thermally activated delayed
fluorescence materials. Chem. Soc. Rev., 46(3), 915.
57 Peng, Q., Niu, Y., Shi, Q. et al. (2013). Correlation function formalism for triplet excited state decay:
combined spin–orbit and nonadiabatic couplings. J. Chem. Theory Comput., 9(2), 1132.
58 Li, Q., Li, Z. (2017). The strong light-­emission materials in the aggregated state: what happens from
a single molecule to the collective group. Adv. Sci., 4(7), 1600484.
410 13  Room-­temperature Phosphorescence of Pure Organics

59 Yang, J., Zhen, X., Wang, B. et al. (2018). The influence of the molecular packing on the room
temperature phosphorescence of purely organic luminogens. Nat. Commun., 9(1), 840.
60 Mao, Z., Yang, Z., Xu, C. et al. (2019). Two-­photon-­excited ultralong organic room temperature
phosphorescence by dual-­channel triplet harvesting. Chem. Sci., 10(10), 7352.
61 An, Z., Zheng, C., Tao, Y. et al. (2015). Stabilizing triplet excited states for ultralong organic
phosphorescence. Nat. Mater., 14(7), 685.
62 Li, M., Ling, K., Shi, H. et al. (2019). Prolonging ultralong organic phosphorescence lifetime to 2.5 s
through confining rotation in molecular rotor. Adv. Optical Mater., 7(10), 1800820.
63 Gu, L., Shi, H., Bian, L. et al. (2019). Colour-­tunable ultra-­long organic phosphorescence of a
single-­component molecular crystal. Nat. Photonics, 13(6), 406.
64 Yuan, J., Wang, S., Ji, Y. et al. (2019). Invoking ultralong room temperature phosphorescence of
purely organic compounds through H-­aggregation engineering. Mater. Horiz., 6(6), 1259.
65 Fang, M., Yang, J., Xiang, X. et al. (2018). Unexpected room temperature phosphorescence from
non-­aromatic, low molecular weight, pure organic molecule through the intermolecular hydrogen
bond. Mater. Chem. Front., 2(2), 2124.
66 Li, Q., Tang, Y., Hu, W. et al. (2018). Fluorescence of nonaromatic organic systems and room
temperature phosphorescence of organic luminogens: the intrinsic principle and recent progress.
Small, 14(38), 1801560.
67 Yuan, W. Z., Zhang, Y. (2017). Nonconventional macromolecular luminogens with aggregation-­
induced emission characteristics. J. Polym. Sci. Part A: Polym. Chem., 55(4), 560.
68 Gong, Y., Tan, Y., Mei, J. et al. (2013). Room temperature phosphorescence from natural products:
crystallization matters. Sci. China Chem., 56(9), 1178.
69 Chen, H., Yao, X., Ma, X. et al. (2016). Amorphous, efficient, room-­temperature phosphorescent
metal-­free polymers and their applications as encryption ink. Adv. Optical Mater., 4(9), 1397.
70 Wang, D., Yan, Z., Shi, M. et al. (2019). Employing lactam copolymerization strategy to effectively
achieve pure organic room-­temperature phosphorescence in amorphous state. Adv. Optical Mater.,
7(23), 1901277.
71 Kwon, M. S., Yu, Y., Coburn, C. et al. (2015). Suppressing molecular motions for enhanced
room-­temperature phosphorescence of metal-­free organic materials. Nat. Commun., 6(6), 8947.
72 Ma, X., Xu, C., Wang, J. et al. (2018). Amorphous pure organic polymers for heavy-­atom-­free
efficient room-­temperature phosphorescence emission. Angew. Chem. Int. Ed., 57(34), 10854.
73 Hu, H., Meier, F., Zhao, D. et al. (2018). Efficient room-­temperature phosphorescence from
organic-­inorganic hybrid perovskites by molecular engineering. Adv. Mater., 30(36), 1707621.
74 Yang, S., Wu, D., Gong, W. et al. (2018). Highly efficient room-­temperature phosphorescence and
afterglow luminescence from common organic fluorophores in 2D hybrid perovskites. Chem. Sci.,
9(9), 8975.
75 Gu, L., Shi, H., Gu, M. et al. (2018). Dynamic ultralong organic phosphorescence by photo-­
activation. Angew. Chem. Int. Ed., 57(28), 8425.
76 Wang, J., Liang, J., Xu, Y. et al. (2019). Purely organic phosphorescence emitter based efficient
electroluminescence devices. J. Phys. Chem. Lett., 10(19), 5983.
77 Wang, C., Yu, Y., Yuan, Y. et al. (2020). Heartbeat-­sensing mechanoluminescent device based on a
quantitative relationship between pressure and emissive intensity. Matter, 2(1), 181.
78 Yu, Y., Wang, C., Wei, Y. et al. (2019). Halogen-­containing TPA-­based luminogens: different
molecular packing and different mechanoluminescence. Adv. Optical Mater., 7(18), 1900505.
79 Li, W., Huang, Q., Mao, Z. et al. (2018). Alkyl chain introduction: in-­situ solar-­renewable organic
colorful mechanoluminescence materials. Angew. Chem. Int. Ed., 57(39), 12727.
411

14

A Global Potential Energy Surface Approach to the Photophysics


of AIEgens: The Role of Conical Intersections
Rachel Crespo-­Otero1 and Lluís Blancafort2
1
 Department of Chemistry, Queen Mary University of London, London, United Kingdom
2
 Institut de Química Computacional i Catàlisi and Departament de Química, Facultat de Ciències, Universitat de Girona, Spain

14.1 ­Introduction

Aggregation-­induced emission (AIE)  [1–3] has become a very valuable tool for the design and
implementation of new emissive materials with potential applications in display technologies,
optical communication, data storage, biological sensing, solid-­state lasing, and others. In AIE, a
molecule that is not or only weakly emissive in solution becomes emissive in aggregate or crystal-
line phases. This property is particularly appealing because these are the phases where most appli-
cations are implemented. AIE circumvents the problem of aggregation-­caused quenching, which
prevents the practical use of molecules that are emissive in solution but become nonemissive in
the aggregate state.
Theory and computations have played an important role in the development of AIE-­related
research by identifying the mechanisms and molecular features behind AIE, providing key princi-
ples for the design of more efficient emitters. The key quantities to rationalize AIE are the fluores-
cence quantum yield, Φfl, and the radiative and nonradiative rates, kr and knr, which are related by
Equation 14.1
kr
(14.1)
knr
fl
kr
The radiative rate is the sum of the fluorescence and phosphorence rates, whereas nonradiative
processes comprise internal conversion and intersystem crossing (ISC) from the emitting states to
the ground state. Enhanced emission can be due to an increase of kr or a reduction of knr, and in
most cases, AIE is associated to a reduction of knr in the aggregate phase.
One can broadly distinguish three complementary approaches to explain AIE: the more
­phenomenological models, such as restriction of intramolecular rotations (RIR) or restriction of
intramolecular motions (RIM); the quantitative evaluation of kr and knr using Fermi’s golden rule
(FGR); and the mechanistic description provided by potential energy surface (PES) studies and
dynamics. The present tutorial chapter for the Handbook of Aggregation-­Induced Emission covers
the contributions of theory using the latter approach, PES studies and dynamics. This has been the

Handbook of Aggregation-Induced Emission: Volume 1 Tutorial Lectures and Mechanism Studies, First Edition.
Edited by Youhong Tang and Ben Zhong Tang.
© 2022 John Wiley & Sons Ltd. Published 2022 by John Wiley & Sons Ltd.
412 14  A Global Potential Energy Surface Approach to the Photophysics of AIEgens: The Role of Conical Intersections

subject of recent minireviews by us [4] and others [5]. Here, we provide a more detailed introduc-
tion to the conceptual and methodological aspects and an update of recent applications not cov-
ered in our previous review. In fact, the reader will find that in little more than a year, there have
been a substantial number of theoretical contributions to AIE research, which testifies the high
activity of the field.
The chapter is structured as follows. In Section 14.2, we discussed the methodological aspects,
starting with a brief overview of the RIR and RIM models and the FGR-­based approaches. We then
discussed the conceptual features of photochemical mechanisms and computational methods for
excited states, including the treatment of large systems and dynamics. In Section 14.3, we reviewed
the literature PES studies, centering on the restricted access to a conical intersection (RACI) model
and the global PES for AIEgens in solution. In Section  14.4, we considered the theoretical
approaches to crystallization-­induced phosphorescence (CIP), and in Section  14.5, we revised
recent theoretical studies on the effect of intermolecular and intramolecular interactions on the
photophysical properties in aggregate phases. In Section 14.6, we introduced recent or less-­studied
developments that represent new challenges for theory, and in Section 14.7, we presented our con-
clusions and outlook.

14.2 ­Methodological Aspects

14.2.1  Intramolecular Restriction Models and the FGR-­based Approach


A first approach to explain AIE is the RIR mechanism [6] and its generalization, RIM [7]. These
mechanisms are based on the idea that the nonradiative decay in solution is due to quenching of
electronic excitation through energy transfer to low-­frequency intramolecular vibrational modes.
These modes are efficient energy acceptors because of the high density of vibrational states. The
intramolecular motions are hindered in aggregate phase, suppressing the quenching and allowing
the molecules to emit. The RIR and RIM models have been used successfully for the design of
numerous AIE emitters [2, 7].
These models have been supported by quantitative calculations of kr and knr using FGR [8–10].
The Huang–Rhys factors and the reorganization energy along the different vibrational modes pro-
vide a mechanistic interpretation of the modes that contribute most to the nonradiative decay and
should be blocked to obtain AIE. FGR calculations have been used successfully to rationalize AIE
in many applications, and the interested reader is referred to a chapter in the present Handbook for
a description of the formalism and the applications.

14.2.2  A PES-­based Description of Photochemical Mechanisms


The main focus of this tutorial review is the description of photochemical mechanisms with the
help of the global PES. Considering the global PES is particularly important to understand the
photophysical performance of molecules in solution because excited-­state relaxation can lead the
molecule far away from the Franck–Condon (FC) region in unconstrained environments. These
cases cannot be covered by FGR approaches, which assume that the relaxation can be described, at
least approximately, with a harmonic model. Moreover, in the last years, it has been recognized for
several AIEgens that excited-­state relaxation in solution does involve passage through a conical
intersection (CI) between the excited and the ground state and that this is the reason for the lack
of fluorescence in solution. This has been the basis for the RACI model [11]. For these AIEgens,
14.2 ­Methodological Aspect 413

emission can be recovered in the aggregate phase if the motion that leads to the CI is restricted.
This provides precise information about the vibrational modes that have to be restricted to gener-
ate emission.
The basic concepts of photochemical mechanisms studied from the PES point of view have
been described in several reviews and book chapters [12–16]. Our present discussion focuses on
the relevant features for AIE. A typical global PES for excited state processes is shown in
Figure 14.1. The PES describes molecular relaxation after excitation and the photochemical and
photophysical processes associated to it. Understanding the global PES provides insight into the
outcome of the excitation and the competing processes, which makes the PES a unique tool for
the rationalization of the photophysical and photochemical processes and the design of more
effective devices.
The shape of the PES depends on the interaction between different electronic states. A typical
organic chromophore usually has several states lying in a relatively narrow range (1–2 eV) above
the lowest electronic excited state, S1. As a consequence, photochemical and photophysical pro-
cesses are not limited to a single excited state, and crossings between excited states may appear.
These crossings are displayed as CIs in Figure 14.1a. In addition, vibronic coupling between the
states, as shown by the shaded area, leads to rather flat potential energy surfaces. On this surface
topography, the excitation usually provides the molecule with enough energy to surmount the bar-
riers on the surface. As a result, excited-­state relaxation can easily bring the molecule far away
from the FC structure.

(a)
(b)
S3

S2

x1
S1
(c)
T1

x1

(d)

S0
x1

Figure 14.1  (a) Global PES involving three singlet excited states (S1–S3, continuous lines), one triplet
excited state (T1, dashed line), and several CIs; examples of sloped (b), peaked (c), and intermediate (d) CI
topography. Possible excited-­state decay paths are illustrated with blue arrows.
414 14  A Global Potential Energy Surface Approach to the Photophysics of AIEgens: The Role of Conical Intersections

A first guideline to understand relaxation is given by Kasha’s rule [17, 18]. It states that lumi-
nescence occurs only from the lowest excited state of a given multiplicity because internal con-
version from higher excited states is always faster than emission. Kasha’s rule has been extended
to the reactivity of excited species, which is assumed to occur with an appreciable yield only
from the lowest excited state [17]. Beyond this approximation, a better description of the mecha-
nism can be obtained calculating the photochemical reaction path [19] on the PES, i.e. the decay
path followed by the molecule after excitation. This approach starts with the computation of the
minimum energy path from the FC structure on the bright excited states, which describes the
initial relaxation and continues with exploration of the paths that ensue from the excited-­state
minima. The resulting decay path on the global PES is displayed as a blue arrowed line in
Figure 14.1a.
Figure 14.1a shows that CIs have a key role in excited-­state relaxation. The CI corresponds to a
point on the PES where the two states cross and their energies are degenerate [20–23]. This maxi-
mizes the probability of internal conversion, which can take place after a few vibrations in the CI
region, in a scale of a few hundreds of fs (1 fs = 10−15 seconds) or less. Accordingly, a CI between
the ground and excited states determines the photophysical performance of the excited molecule,
since it provides a way for it to return to the ground state in a very short time scale. This leaves it
no time to fluoresce and results in an increase of knr. In addition, the passage through the CI can
be associated to the formation of photochemical products and be detrimental for the photostability
of the molecule.
The experimental signature of the passage through the CI is a short excited-­state lifetime, con-
sistent with the large nonradiative decay rates. The lifetime can be measured in time-­resolved
pump–probe experiments, where the molecules are excited with a pump pulse and the excited
state is probed with different types of spectroscopy, such as transient absorption, photoionization,
or fluorescence. Some of the best studied excited-­state processes involving a CI are the photoi-
somerization of the retinal chromophore  [24], which is the molecular basis of vision, and the
photophysics and photochemistry of the DNA bases [25, 26]. The excited-­state lifetimes measured
for these processes are of a few ps (1 ps = 10−12 seconds) or less, and they depend not only on the
time needed to decay from the upper to the lower state but also on the time needed to access the CI
region from the FC region and on the barriers that separate the two.
In the RACI model, knowing the structure of the CI responsible for radiationless decay allows
one to identify the modes that have to be blocked to inhibit the decay and regenerate the emission.
It is important to have a mechanistic understanding of why the CI appears on the PES and how it
is related to the changes in the electronic structure upon excitation. Often this understanding can
be obtained from the molecular orbitals involved in the excitation. The idea here is that relaxation
has to follow a coordinate where structural deformation has a low energetic cost in the excited
state and a high one in the ground state. The energy increase in the ground state favors the appear-
ance of the CI.
This idea is illustrated in Figure 14.2 for an AIEgen that follows the RACI model along with a
double-­bond torsion coordinate [27]. The highest occupied molecular orbital and lowest unoccu-
pied molecular orbital that are involved in the excitation are bonding and antibonding along the
central double bond. This implies that the excitation weakens the double bond and transforms it,
effectively, into a single bond that can rotate almost freely. This is confirmed by the calculated
relaxation coordinate, which does indeed follow the torsion around the central bond (measured by
the dihedral angle φ). The effect of nuclear relaxation on the ground state energy is a steep increase,
since this state retains the double-­bond character, and as a consequence, the CI is reached along
the decay path with virtually no energy barrier.
14.2 ­Methodological Aspect 415

(a) (b)

φ= 17°
FC
3 φ= 54°
φ= 88°
Relative energy (eV)

S1thf

Cl
LUMO
2

O O S1
φ
O O S0
1

0
HOMO
0 10 20 30 40 50
Displacement/bohr.amu1/2

Figure 14.2  (a) TD-­CAM-­B3LYP decay path for TFE [27]; (b) natural transition orbitals involved in the
excitation at the FC geometry. Source: Adapted from Ref. [27] with permission from John Wiley and Sons.

Mathematically, the PES has dimension 3N-­6, where N is the number of atoms of the molecule.
The conical shape of the crossing [23] arises from the fact that there are two directions on the PES
(two nuclear coordinates) that lift the degeneracy linearly, the gradient difference and interstate
coupling vectors (x1 and x2, respectively, see Equations 14.2 and 14.3)

E1 E2
x1 (14.2)
q

∂ Ψ1 Ĥ Ψ 2
x2 = (14.3)
∂q
x1 is the coordinate that maximizes the energy gap arising as one moves away from the CI, and it
can also be understood as the most effective coordinate to approach the CI. In turn, x2 couples the
two states.
The CIs can be classified in terms of the topography along x1 (see Figures 14.1b–d), in particular
according to the relative slopes of the intersecting surfaces [28]. Figure 14.1b shows a sloped topog-
raphy, where the slopes of the two surfaces have the same sign. In this case, the decay is not associ-
ated to any product formation but leads to regeneration of the reactant. This topography is often
encountered, for example, when the decay is associated to a ring puckering coordinate, for which
several examples will be shown below. From the point of view of decay efficiency, it may take the
molecule several oscillations around the S1 minimum to reach the CI and decay to the ground
state. The alternative case is the peaked topography shown in Figure 14.1c, where the two crossing
surfaces have slopes of opposite sign. This topography can be associated to the formation of a pho-
tochemical product, labeled P on the figure, and is encountered for instance in photocycloaddi-
tions, such as those found for tetraphenylethylene (TPE) derivatives. The decay at a peaked
416 14  A Global Potential Energy Surface Approach to the Photophysics of AIEgens: The Role of Conical Intersections

topography is particularly efficient because the gradient of the excited state along the path leads
the molecule directly to the CI. Typically, trajectories that decay at a peaked CI are branched in two
groups, one that decays further to the products and one that reverts back to the reactant. Finally, in
Figure 14.1d, we show an intermediate topography where the CI lies in close vicinity of a minimum
of the excited state. This topography is not associated to reactivity but allows for a very effi-
cient decay.
Another relevant characteristic of the PES associated to the CI is the multidimensional nature of
the CI [22, 23, 29]. In contrast to the x1 and x2 coordinates, along the remaining 3N-­8 degrees of
freedom, the degeneracy is not lifted at first order. Therefore, the CI is not an isolated point on the
PES but forms a multidimensional seam of intersection. This has two main implications for the
photophysical performance. First, the probability of decay at the CI increases because it is not a
single point but an extended space, and second, the extended seam of intersection can be com-
posed of different segments (different regions of the PES) that are associated to different photo-
chemical outcomes.
While the previous considerations are centered on excited states of singlet multiplicity, which
are associated to fluorescence, they can be extended in a straightforward way to triplet states and
phosphorescence. The analog of AIE for triplet states is CIP [30], and the efficiency depends cru-
cially on the population of the triplet from the initially excited singlet state, i.e. ISC. The probability
of ISC depends on spin–orbit coupling (SOC). Molecular systems with heteroatoms and metallic
centers can display significant SOC and large probabilities of ISC. Consequently, a significant pop-
ulation of triplet states can deactivate via phosphorescence. According to El-­Sayed’s rule [31], large
SOCs can also be obtained in organic molecules through electronic transitions involving a change
of orbital angular momentum, such as 1nπ*⟷3ππ* or 1ππ*⟷3nπ* conversions. ISC is also facili-
tated by small singlet-­triplet gaps, which are common in systems with charge-­transfer (CT) states.
Similar to internal conversion to the ground state, the probability of ISC in the vicinity of the FC
geometry can also be calculated with the FGR approach, and it depends on the SOC constant.
However, ISC can also take place at regions of the PES away from the FC geometry where the sin-
glet and triplet states cross. In this case, location of the singlet/triplet crossing on the PES and
trajectory calculations can provide the mechanism behind ISC. However, the difference between a
singlet/triplet crossing and a CI is that the crossing does not have a conical shape because the
length of the x2 vector between states of different multiplicity is zero.

14.2.3  Computational Approaches for Excited States


14.2.3.1  Electronic Structure Methods for Excited States
The study of excited-­state PES is particularly challenging, and the choice of the computational
approach is an important point if one wants to obtain reliable results. AIEgens are quite demand-
ing systems from the computational point of view because of their relatively large size, as they are
frequently composed of a core chromophore with additional aromatic substituents. Therefore, the
choice of the computational approach is usually a compromise between efficiency and accuracy.
The methodological difficulties associated to excited states can be traced down to how electronic
correlation is treated. Two types of electron correlation are usually distinguished, dynamic and
static. Dynamic correlation is important because excited states can be of diverse character, such as
ππ*, nπ*, CT, doubly excited, or excitonic, to mention the most common cases in AIEgens, and the
contribution of dynamic correlation energy can be different. If dynamic correlation is not treated
in a balanced way for all states, artefacts in the PES may appear and distort the mechanistic picture
provided by the calculations. In turn, static correlation is important in regions of the PES where
14.2 ­Methodological Aspect 417

the ground state becomes multireferential or open shell in character. These regions, which are
often found in the vicinity of CIs and are associated to bond breaking/formation or double-­bond
torsion, represent an additional computational challenge.
These issues and the accuracy of the different approaches have been discussed in other refer-
ences (see, for instance [14, 15, 32–37]). In this section, we focus on the methods most frequently
used in the study of AIEgens, namely multireference methods and time-­dependent density func-
tional theory (TD-­DFT).
In multireference methods, static correlation is included by describing the wave function as a
linear combination of Slater determinants. This provides the necessary flexibility to describe
ground states with multireferential or open-­shell character and makes these methods appealing for
the study of excited states, as in principle they can provide correct descriptions of near degenerate
regions and CIs. However, the main disadvantage is that the most common formulations, such as
complete active space self-­consistent field (CASSCF), only give a partial account of dynamic cor-
relation, restricted to the orbitals that form the active space. This deficiency is solved by multirefer-
ence perturbation theory approaches such as complete active space second-­order perturbation
(CASPT2) [38–40], where dynamic correlation is recovered through second-­order perturbation of
the CASSCF wave function. Although CASPT2 and its multistate variant MS-­CASPT2 usually give
a very good description of the excited PES, their use for the systematic exploration of PES is limited
by the fact that they cannot be used routinely for structural optimizations in the typical AIEgens
that contain several aromatic rings. Therefore, a usual approach is to carry out the structural opti-
mizations with CASSCF and recalculate the energies with CASPT2 or MS-­CASPT2. This approach
is called (MS-­)CASPT2//CASSCF.
Although the (MS-­)CASPT2//CASSCF approach has been applied successfully in several studies
of AIEgens [11, 41–49], it has some limitations. One of them is the appearance of differential cor-
relation effects because of the poor recovery of dynamic correlation in CASSCF. This occurs when-
ever dynamic correlation is not provided in a balanced way for all states. Figure 14.3 shows an
example where dynamic correlation for S2, which is shown in red, is not recovered at the CASSCF
level in the same measure than for the other states. The dashed and solid red lines represent S2
calculated at the CASSCF and (MS-­)CASPT2 levels, respectively. The result of the poor description
is an artificial, upward energy shift of S2 at the CASSCF level. This leads to an inversion of the state
order in the FC region, since S2 lies higher in energy than S3 at the CASSCF level. In addition, the
CI with S0 is not located correctly at the CASSCF level and its energy is clearly overestimated.
Clearly, differential correlation can make the study of the PES with CASSCF optimizations prob-
lematic. Another limitation of CASSCF is given by the size of the active space, which should not
exceed 12–14 orbitals to remain practical. Typical AIEgens often have several aromatic rings, and
their treatment at the CASSCF level is challenging because the use of too small active spaces can
also lead to artefacts due to undesired localization of the active orbitals in some fragments of the
molecule. Developments are in progress to overcome this limitation, such as implementation of
nuclear gradients to perform (MS-­)CASPT2 optimizations [50] or methods such as second-­order
perturbation for restricted active space self-­consistent field [51] or the density matrix renormaliza-
tion group [52] for large systems. However, their use has not been generalized yet and there are no
applications to AIE.
In this context, TD-­DFT presents a convenient alternative. It is based on the application of linear
response theory to a DFT wave function. Although the DFT function does not include static cor-
relation, it does include dynamic correlation and this results in many cases in quite an accurate
description of excited states near the Franck–Condon region. This, combined with a higher com-
putational efficiency compared to multireference approaches, has made TD-­DFT very widespread
418 14  A Global Potential Energy Surface Approach to the Photophysics of AIEgens: The Role of Conical Intersections

S2CASSCF

S3 CICASSCF

S2CASPT2

S1

CICASPT2

S0

Figure 14.3  Illustration of differential correlation effects resulting in state inversion and wrong location
of a CI. S2CASPT2 and S2CASSCF represent the correct and the poorly correlated description of the PES of S2; for
the remaining states, it is assumed that the inclusion of dynamic correlation with CASPT2 does not alter
the relative energies significantly.

for the investigation of excited state potential energy surfaces. Many studies of AIE mechanisms
have been performed using these methods  [53–55]. In particular, the hybrid PBE0  [56, 57] and
B3LYP [54–56, 58] and the long-­range corrected CAM-­B3LYP [27, 53] and ωB97X-­D [47, 59] func-
tionals have been widely used in recent studies. In some of these studies, the Tamm–Dancoff
approximation (TDA) is used, which reduces the number of excitations to strictly monoelectronic
ones [60].
In TD-­DFT and TDA-­DFT, a careful selection of the DFT functional is needed to properly
account for different kinds of excited states involved in the excited state pathways. One particular
difficulty is related with the description of CT states, which can be relevant for the description of
the AIE mechanisms in diverse chromophores. The description of CT states can be challenging for
traditional functionals including generalized gradient approximation based and hybrid ones,
where the energy of these states can be underestimated by several eV, causing artefacts on the
PES. The underestimation of the energy of CT states comes from the failure of traditional function-
als to describe the asymptotic behavior of exchange. Long-­range corrected functionals address this
problem by partitioning exchange, dividing the interelectronic distance operator into short-­range
and long-­range components as shown in Equation 14.4, commonly via an error function [61–63]

1 1 erf r12 erf r12


(14.4)
r12 r12 r12
In Equation  14.4, μ determines the separation between short and long ranges. Values of μ
between 0.15 and 0.5 a0−1 work very well for various systems [64]. These two components can be
treated using general functionals and exact HF exchange. Tuning the value of μ has become a suc-
cessful strategy to obtain an optimal description of CT states [65]. In general, long-­range corrected
14.2 ­Methodological Aspect 419

functionals significantly improve the energies of CT states, oscillator strengths, and optical
properties [63].
In the cases where the photochemistry does not involve CT states, hybrids perform relatively
well. This is the case of prototypical propeller-­shaped AIEgens [47, 55, 66], where absorption and
emission energies are fairly well described with B3LYP. However, polar systems where CT states
are relevant might require the use of long-­range corrected functional. This is the case of a tolane-­
based chromophore, where CAM-­B3LYP outperforms PBE0 and B3LYP [53]. For these reasons, it
is highly advisable to benchmark excitation energies vs experimental values to decide which func-
tional is better for the system of interest.
One case that illustrates the relevance of the choice of functional for the investigation of AIE
mechanisms is boron difluorohydrazones (BODIHY). Initially it was suggested that relaxation to S1
(dark state) was hindered in viscous solutions and the solid state, and emission would happen from
states higher in energy [67]. This mechanism, called suppression of Kasha’s rule, was based on
TDA-­PBE calculations. However, these resuIts have recently been revised combining TD-­B3LYP
and TD-­CAM-­B3LYP with CASSCF and XMC-­QDPT2  [49]. With these methods, S1 is a bright
state. Therefore, TDA-­PBE provides a wrong order of the states, and TD-­B3LYP provides the best
agreement of the absorption and emission energies with respect to the experimental values. The
excited state mechanism explored using a combination of CASSCF and XMC-­QDPT2 seems to be
consistent with the RACI model (see Section 14.3.4 for details).
Another relevant issue is the location of CIs with TD-­DFT. TD-­DFT cannot provide a correct
description of the conical shape of CIs because the interstate coupling coordinate (x2, see
Equation 14.3) is ill defined [68, 69]. In some cases, TD-­DFT can provide a reasonable estimation
of geometries and energies of the minimum energy CI, but this has to be carefully tested compared
to multiconfigurational methods [69]. Recent works show the similarities of the CI obtained with
CASSCF and TD-­DFT for proton transfer systems [70–72]. An alternative for the location of CIs
with TD-­DFT is given by spin-­flip TD-­DFT (SF-­TD-­DFT) [73], which has been used successfully
for the location of CIs in some AIEgens  [58, 74]. SF-­TD-­DFT provides a qualitatively correct
description of the CI topography [68, 69, 75] by using a triplet determinant as reference, although
the use of this method is not trivial because it may suffer from spin contamination.
To summarize, TD-­DFT offers a convenient way to study the photophysics of AIEgens, but a
careful choice of the functional is mandatory to obtain reliable results. One should be aware that a
good agreement of the calculated vertical excitations with the experimental spectra does not guar-
antee a good description of other regions of the PES. Therefore, TD-­DFT should be complemented
with higher-­level methods to check the validity of the results, e.g. with single-­point calculations on
the TD-­DFT optimized structures. Another possible strategy is to combine TD-­DFT calculations
near the FC region with multireference methods for the regions of the PES with significant multi-
configurational character.
A possible alternative to multireference methods and TD-­DFT is given by the algebraic diagram-
matic construction (ADC) and approximate coupled cluster n (CCn) methods. Both methods use
linear response theory to obtain excitation energies. They are significantly more computationally
expensive than TD-­DFT and avoid some of the pitfalls described above. ADC works on the basis of
a wave function that contains second-­order perturbation correction, and its accuracy is compara-
ble to that of MP2 in the ground state [76]. The regular variant does not provide the correct dimen-
sionality of CIs, but its spin-­flip formulation does reproduce the correct topology. In turn, CC2 and
CC3 use an approximate coupled cluster wave function as reference [77]. The performance of CC2
and other excited-­state formulations of coupled cluster for the description of CIs has been studied
in several works and it depends on the details of the excited-­state formulation [68, 78–80]. Focusing
420 14  A Global Potential Energy Surface Approach to the Photophysics of AIEgens: The Role of Conical Intersections

on AIE, CC2 has already been used with success for some AIEgens [47, 70, 71], and we expect that
its use in the AIE context will increase in the future.

14.2.3.2  Dynamics Simulations in the Context of AIE


Dynamics simulations provide a direct picture of the time evolution of a molecule on the PES. They
are particularly suitable to simulate the decay of AIEgens in solution involving a CI because of the
short time scales where this usually occurs. Beyond the picture provided by a global PES study,
dynamics provide insights into the excited state lifetimes that can be directly compared to the
results from time-­resolved spectroscopy. In the cases where several relaxation paths are possible,
dynamics can also reveal the preference for one path or another. In principle, the dynamics can
treat the decay at the CI and the nonadiabatic effects associated to it, which is particularly interest-
ing in the cases where there is more than one possible decay path from the CI.
All the simulations carried out up to date for AIEgens have followed the mixed quantum-­classical
(MQC) approach  [81]. In MQC dynamics, the electronic wave function is evaluated quantum
mechanically, calculating the PES on the fly during the propagation, and the nuclei are propagated
classically on the surface. The simulations use the so-­called swarm of trajectories approach, where
the dynamics are calculated for a group of trajectories with different initial conditions that sam-
ple the potential and kinetic energy distribution found for a molecular population in the moment
of the excitation.
Turning to the electronic structure methods, mostly three approaches have been used in the
dynamics simulations: TD-­DFT  [47, 48, 82]; the tight-­binding approximation to TD-­DFT, time-­
dependent tight-­binding density functional theory (TD-­DFTB [83, 84]) [85, 86]; and the semiem-
pirical method OM2 multireference configuration interaction (MRCI) (OM2/MRCI [87]) [41, 42].
The treatment of nonadiabatic effects depends on the electronic structure method. Out of the three
methods mentioned above, only the multireference OM2/MRCI approach can describe these
effects correctly. In this case, the decay from the excited to the ground state is simulated using the
trajectory surface hopping (TSH) approach, where trajectories are allowed to hop to the ground
state in the vicinity of the crossing [88]. This allows one to follow the fate of the trajectory on the
ground state after the decay. In contrast, TD-­DFT and TD-­DFTB cannot simulate the passage
through the CI. In these cases, it is usually assumed that the trajectories decay to the ground state
once the S1/S0 energy gap falls below a certain threshold and only information about the decay
time is obtained. Still, TSH can be applied in a DFT context with the time-­dependent Kohn–Sham
(TD-­KS) approximation, and a combination of TD-­KS with DFTB has been used for the diphenyl
dibenzofulvene (DPDBF) AIEgen [85]. Another possible approach is to parametrize an excited-­
state force field against a quantum mechanics (QM) method. This has been done for the dynamics
of a cyanostilbene derivative, using SF-­TD-­DFT as the QM Ref. [89].

14.2.4  Methods for Large Systems


A common step in the investigation of AIE mechanisms is the analysis of quenching pathways in
the gas phase or dilute solution. This helps identify which reaction coordinates should be blocked
to avoid similar mechanisms happening in the condensed phase. AIE has been witnessed in differ-
ent phases including concentrate solution, amorphous systems, and molecular crystals  [4].
Simulation of excitations in large aggregates and the solid state brings several challenges, and one
fundamental question to address is: Where are the excited states localized? Are they confined to a
single choromophore or spread over several ones [90]?
14.2 ­Methodological Aspect 421

The answer to this question is determined by the electronic structure of the single molecules and
their interactions with the surrounding neighbors. The calculation of exciton couplings can help
determine a priori whether the energy gain by spreading the electronic density over monomers can
overcome the reorganization energy. Typical propeller-­shaped chromophores do not display strong
intermolecular interactions and their molecular crystals have a relatively low density. These mol-
ecules have exciton couplings in the range of 0.2–35  meV, which are much smaller than their
reorganization energies [91]. However, larger exciton couplings can be found in systems displaying
stronger intermolecular interactions [72].
Continuum models are the simplest approach to consider the effect of the environment in
excited states. These methods do not take into account the structure of the solvent but represent it
as a continuum with a specific dielectric constant. The excited AIEgen is placed inside a cavity and
charges are induced on the surface. The interaction between the excited states and the charges is
described with an embedding potential. Different methods that differ on the algorithm to obtain
the surface charges and the embedding potential have been defined and the most common one is
the polarizable continuum model (PCM) [92]. The response of the environment to the excitation
can be also considered using state-­specific methods with the charges calculated for every state of
interest. Continuum models are particularly useful for the description of mechanisms in solution
when specific interactions between the chromophores and the solvent are not important for the
mechanism. These methods have been frequently used to describe excited state processes in
AIEgens [11, 93–96]. If the dielectric constant of the solid is known, continuum models can also
help model excitations in solid state [97]. However, the structure of the environment and its explicit
interactions with the excited density can be essential to provide an accurate description of the
mechanism, and methods considering the nature of the specific interactions may be more appro-
priate for many processes.
Hybrid approaches represent the most common alternative to account for the effect of the envi-
ronment [98]. They combine a higher level of theory to investigate a localized excitation with more
approximate methods for the rest of the system. The low level of theory is commonly a molecular
mechanics scheme (MM). The electrostatic interactions between the regions are simulated by
either using mechanical embedding, where electrostatic interactions are calculated at the lower
level of theory, or more commonly nowadays by using electrostatic embedding that involves
directly including them in the Hamiltonian. A third option, which is becoming more commonly
used, is the use of polarizable force fields, where the charges of the environment are relaxed to take
into account the mutual polarization of the environment due to the excitation  [95]. Hybrid
approaches can be classified as additive or subtractive depending on how the total energies are
obtained. The ONIOM method, proposed by Morokuma’s group [99], is a subtractive approach,
where the interaction energy between the regions is calculated as the difference between the total
energy and the energies of the two regions at the low level of theory. In ONIOM, in addition to MM
force fields, ab initio, DFT or semiempirical methods can also be chosen to describe the environ-
ment. Given its flexibility, ONIOM has been a frequent choice for the study of AIE mechanisms [44,
96, 100].
Periodic boundary conditions methods including periodic TD-­DFT and GW [101] can be used
to describe molecular crystals with highly delocalized excitations. However, the applications of
these methods to explore regions of the PES with a significant multiconfigurational character
such as CIs are somehow limited. The most popular approach is to use cluster models to approxi-
mate the crystalline environment and hybrid methods to explore their excited state mechanisms.
Clusters of several molecules may be considered in the investigation of mechanisms in
AIEgens [27].
422 14  A Global Potential Energy Surface Approach to the Photophysics of AIEgens: The Role of Conical Intersections

Within the hybrid approaches, periodic charge embedding (PCE) methods allow for including
the effect of long-­range interactions from the infinite crystal in a local region [90]. These methods
are commonly used for the simulation of electronic structure defects and nuclear magnetic reso-
nance in molecular crystals  [102]. In particular, the approach proposed by Klintenberg et  al.
defines how to embed a cluster model in a set of point charges to reproduce the electrostatic poten-
tial of the periodic crystal in the central region [103]. A set of artificial charges is optimized to
ensure that the electrostatic potential inside a cluster reproduces the one generated by the infinite
crystal obtained from the Ewald sum as implemented in the Ewald code [103]. PCE methods have
been recently extended to be used in the investigation of excited states in molecular crystals [104].
The simultaneous excitation of all chromophores and their interaction in the crystal can be simu-
lated using a self-­consistent approach as proposed by Ciofini’s group [104, 105]. In the investiga-
tion of the AIE in 2,7-­diphenylfluorenone using the self-­consistent approach, the authors found
that the crystal environment enhanced radiative decay with respect to solution [105].
In the periodic embedding approaches, the description of short-­range non-­Coulombic interac-
tions should be taken into account for the optimization of distorted geometries such as commonly
found in CIs. This can be done by defining ONIOM-­like schemes for the calculation of total ener-
gies and gradients [96]. Different ONIOM QM:QM′ schemes have been implemented in the from-
age program, which is open source and freely available [97]. In these calculations, the short-­range
interactions such as dispersion and exchange are approximated by their values in the ground state,
and the Ewald approach can be used to describe long-­range electrostatic interactions experienced
in the crystal environment.
The method to generate the initial structure for the simulations depends on the phase where AIE
happens. In the case of molecular crystals, the 3D structures can be obtained from specialized
databases such as the Cambridge Structural Database. Then, the crystals are refined using periodic
boundary condition DFT or QM/MM methods. The simulation of AIE in less ordered media such
as solution or amorphous materials might require molecular dynamics simulations to generate the
initial structures [66, 106].
An example is the investigation of the excited state mechanism of dimethoxy-­TPE in water solu-
tion. Molecular dynamics simulations of the aggregates in the nanosecond scale showed that large
aggregates featuring chain-­type structures were stabilized at high concentrations [95]. These clus-
ters mainly presented J-­type aggregation, although the fraction of H types changes with the tem-
perature. The intramolecular rotation was significantly restricted in the clusters. Specific
interactions with water molecules through hydrogen bonds appeared to be relevant for the aggre-
gation process.
A case to illustrate the use of these methods in the simulation of amorphous materials is the
simulation of amorphous DMTPS. Clusters of different sizes were considered with molecular
dynamics with the generalized Amber force field (GAFF). Representative nanoparticles were
extracted from the dynamics to perform the excited state calculations with an additive QM : MM
method considering one molecule in the QM region described at the TD-­B3LYP level of theory and
the GAFF force field for the rest of the system [66]. These calculations showed a significant depend-
ency of the quantum efficiency relative to the size of nanosized DMTPS aggregates.
A concluding example of the potential of computational modeling of AIE processes in different
phases is the recent study of the mechanofluorochromic behavior of difluoroboron of β-­diketonates
by Wilbraham et al. [106] (Figure 14.4). The authors explored the excited state mechanism of poly-
morphs of three derivatives in THF solution (modeled with PCM), amorphous and crystalline
phases, providing an interpretation of the experimental results (Figure 14.5). Excited states of the
polymorphs were simulated using self-­consistent Ewald periodic embedding and those of amor-
phous phases with classical molecular mechanic simulations with the CGenFF force field. The
14.2 ­Methodological Aspect 423

F F
1
B
O O

F F H F F
B 2a 2b B 3a 3b
O O O O

OMe NHBu
O O
O O
(a) P
(b) f
(c) B
0.0197 0.0498 –0.0101
0.0178 0.0450 –0.0090
Excitation energy (eV)

Excitation energy (eV)

Excitation energy (eV)


3.3 0.0159 3.3 0.0403 3.3 –0.0080
3.2 3.2 3.2
3.1 0.0140 3.1 0.0355 3.1 –0.0070
3.0 0.0120 3.0 0.0308 3.0 –0.0060
2.9 2.9 2.9
2.8 0.0101 2.8 0.0260 2.8 –0.0050
2.7 2.7 2.7
0.0082 0.0213 –0.0040
1.0 0.0063 1.0 0.0165 –0.0030
0.5 0.5 1.0
0.0 0.0044 0.0 0.0118 0.5 –0.0020
Å)

Å)
0.0

Å)
–0.4 0.0 –0.5 –0.5
(

(
0.0025 –0.4 0.0 0.0070 –0.4 –0.0010

(
Δy

–0.5

Δy
0.4 0.8

Δy
–1.0 0.4 0.8 –1.0 0.0 0.4
1.2 1.2 0.8 1.2 –1.0
Δx (Å) 0.0006 0.0023 –0.0
Δx (Å) Δx (Å)

(d) (e)
Δy
P×f×B 1 (THF)
1
Normalized intensity (a.u.)

0.0295 1a
0.0266 1 Amorph
3.3
Excitation energy (eV)

0.0237
Δx 3.2
0.0207 0
3.1
3.0
0.0178
400 500 600 700
2.9
2.8 0.0149 1 (THF)
2.7 1
0.0119
1a
0.0090 1 Amorph
1.0
0.5 0.0061
0.0
0
Å)

–0.4 –0.5 0.0032


(

0.0
Δy

0.4
Δx (Å)
0.8
1.2 –1.0 0.0002 350 400 450 500
Wavelength (nm)
(f) (g)
2 (THF) 3 (THF)
Normalized intensity (a.u.)

1 2a 1 3a
Normalized intensity (a.u.)

2b 3b
2 Amorph 3 Amorph

0 0
400 500 600 700 400 500 600 700
2 (THF) 3 (THF)
1 2a
1 3a
2b 3b
2 Amorph 3 Amorph

0 0
350 400 450 500 350 400 450 500
Wavelength (nm) Wavelength (nm)

Figure 14.4  Mechanofluorochromic behavior of difluoroboron β-­diketonates 1-­3 (see top panel). (a)–(d)
Plots of photophysical and energy parameters with respect to scan coordinates for a pair of molecules (1) in
crystalline phase. One molecule (shown in red in (d)) was displaced along the average plane in directions
parallel (Δx) and perpendicular (Δy) to its long axis. (a) Probability of finding a configuration in the
amorphous phase. (b) Oscillator strength. (c) Boltzmann population. (d) Weighted oscillator strength (red
[blue] are regions of higher [lesser] contribution of a given scan coordinate to the calculated emission
spectrum in the amorphous phase). (e)–(g) Experimental (top) and computed (bottom) emission spectra for
1 (e), 2 (f), and 3 (g). Source: Adapted from Ref. [106] with permission from John Wiley and Sons.
424 14  A Global Potential Energy Surface Approach to the Photophysics of AIEgens: The Role of Conical Intersections

(a) (b)

4.91

3.72
3.54 3.01 3.65 3.12
3.10

Solution Crystal

Figure 14.5  Schematic PES of DMTPS in solution (a) and crystalline phase (b) showing the RACI model.
Source: Adapted from Ref. [11] with permission from the Royal Society of Chemistry.

TD-­CAM-­B3LYP functional with the 6-­31G(d) basis set was used to simulate the excitations. For
the amorphous phase, the simulations showed that molecular configurations similar to those
found in the crystalline phase are responsible for the mechanochromic fluorescence, linked to the
weak van der Waals interactions between the slip planes formed by the molecules (Figure 14.5).

14.3 ­CI-­centered Global PES for AIEgens


The study of PES has demonstrated the role of CIs in the decay of AIEgens in solution, providing
the basis for the RACI model. A typical situation exemplifying the RACI model is shown in
Figure 14.5, which represents the PES calculated for the prototypical DMTPS AIEgen [11]. The
molecule has an energetically accessible CI in solution because the barrier that separates the CI
from the FC region (3.54 eV) is lower than the vertical excitation energy (3.72 eV). In the aggregate
phase, the energy of the CI is raised by almost 2 eV and the CI becomes inaccessible, rendering the
molecule emissive. The AIEgens that have been shown to follow a similar behavior are compiled
in Figure 14.6, where they are classified according to the coordinate followed to access the CI. The
atoms involved in the distortions are marked in red in the structural drawings and in turquoise in
the ball and stick structures.

14.3.1  Double-­bond Torsion


Double-­bond torsion is one of the favored coordinates for the appearance of a CI because the situ-
ation described in Figure 14.2 for TFE (Section 14.2.2) is quite general, i.e. the excitation changes
14.3 ­CI-­centered Global PES for AIEgen 425

(a) Double bond torsion (c) HO


H
O

N
N

N
N
ESIPT
O
O O

O
HPIP O
N O
NH NH
BIM N N
H
DPDBF HO O
R4 R3 HPQ
R1
R3 R 4 CN R2 OH O O OH

α
R2 CN R 4 R3
β
R1
N N
R3 R4
α-DCS HC1
OH O O OH

CN NC CN
N N
o
i N
H
HP1
o-DCSP CN-MBE
R

CN
NC
N (d)
NH 2

BST[7]-tBu (R= tBu) C1


BPST[7] (R =Ph)
Si
BDPST[7] (R =diphenyl)
DMTPS
Ring
puckering
(b) TPE Derivatives
bond torsion/cyclization

TPC N
CN
vs F 2B N
N
N
R R

TPE DHP
R
N
CHO
BODIHY
R S
BDAA O O

X N HN NH

N O O
DNCOT
O

R O O

DBU-C (X=CH2,C2H4)
TPE-2R TPE-4mM P4TA
(R =CH3O,F)
(e) Bond stretch
= BH

O O
O O
N N

TFE THBDBA B18H20(NC5H5)2

Figure 14.6  Overview of the AIEgens that follow the RACI model, classified according to the relaxation
coordinate associated to the CI (panels a–e). The atoms involved in the distortions are marked in red in the
structural drawings and in turquoise in the ball and stick structures.

the π-­bonding pattern of the molecule so that torsion around a ground-­state double bond becomes
almost free in the excited state. In many cases, torsion around the double bond is not enough to
reach the degeneracy between the ground and excited state and displacement along an additional
coordinate, usually pyramidalization of one of the sp2 carbons, is needed to reach the CI [107, 108].
This occurs, for example, in symmetric olefins.
The molecules that have been shown to follow the RACI model along the double-­bond torsional
coordinate are shown in Figure 14.6a. They include DPDBF [44, 85], 4-­dimethylamino-­2-­benzylidine
malonic acid dimethyl ester (BIM)  [96], α-­dicyano-­distyryl benzene (α-­DCS)  [43, 94, 109] and
426 14  A Global Potential Energy Surface Approach to the Photophysics of AIEgens: The Role of Conical Intersections

dicyano-­distyryl pyrrole  [58] derivatives, 1-­cyano-­1,2-­bis-­(4′-­methylbiphenyl)ethylene (CN-­


MBE) [89], and a 2,3-­diaminomaleonitrile derivative (C1) [56]. All these molecules have an ener-
getically accessible CI in solution where the dihedral angle around the central double bond is
approximately 90° (see the ball and stick structures in Figure 14.6a). The TPE derivatives can in
principle also follow this mechanism, but they will be discussed separately because they can also
undergo cyclization (see Section 14.3.2).
Of these molecules, DPDBF [110] and BIM [111] are typical AIEgens, i.e. they are only weakly
emissive in solution and become luminescent in the aggregate and crystalline phases. These mol-
ecules have “crowded” double bonds, consistent with the fact that the rotation of the bulky groups
that leads to the CI is hindered in the aggregate phase. The decay of DPDBF has been studied with
PES energy calculations [44] and dynamics. An early TSH study with TD-­KS and DFTB [85] pre-
dicted a lifetime of 1.4 ps in the gas phase, and bond stretching in the DBF ring and ring rotation
around the double bond played an essential role in the nonadiabatic transition. In a more recent
study using long-­range corrected TD-­DFTB [86], a similar lifetime of 0.8 ps was obtained in the gas
phase. In this case, the decay was also simulated in hexane solution by including explicit solvent
molecules in the simulation with a QM/MM approach. The decay in solution slowed down to
8.15 ps and double-­bond torsion remained the preferential decay channel. Experimental support
for the involvement of a CI in the decay of DPDBF is given by recent measurements of the excited
state lifetimes of the related (1,2,3,4)-­tetraphenyl-­(5,6)-­dimethoxycarbonylfulvene  [112], which
are shorter than 1 ps in solution, consistent with the simulated values for DPDBF.
The RACI model has also been useful to understand the photophysics of some compounds that
could not be explained with the RIR or RIM models. This is the case of the α-­DCS compounds
shown in Figure 14.6a. These compounds have AIE behavior, whereas the β-­DCS analogs, which
only differ in the position of the CN substituent in the vinyl group, are luminescent in solution and
in the aggregate phase [94]. In addition, a set of α-­DCS and β-­DCS compounds shows an “inverted
energy gap law” behavior. Thus, the energy gap law predicts that the compounds with a smaller
vertical excitation energy (EFC) should have a larger knr and a smaller Φfl. However, the compound
set shows the inverse correlation, i.e. Φfl increases as EFC decreases (see Figure  14.7)  [109].

100 26
23 18
22
Fluorescence quantum yield ΦF

21 29
30
19 12
10–1 32
33 31
25 7
28 20

94
24 6
10–2 8 13
16
10
27 17
5 15
14 2 1
10–3 3
11

2.5 3.0 3.5 4.0


FC energy EFC,exp (eV)

Figure 14.7  Correlation of experimentally observed quantum yields Φfl vs the FC energies EFC for
a set of 33 DCS compounds. α isomers are displayed in green, and β isomers are displayed in red.
Source: Reproduced from Ref. [109] with permission from the Royal Society of Chemistry.
14.3 ­CI-­centered Global PES for AIEgen 427

Significantly, this correlation is found for a set of 33 derivatives where Φfl varies over more than
three orders of magnitude (0.001–1).
To explain this behavior, it has been proposed that the lack of fluorescence of the α compounds
in solution is due to a CI found along the torsion coordinate around one of the vinyl double bonds.
Such a CI exists both for α and β compounds, but it is separated from the FC region by a barrier,
and the different photophysics in solution depend on the energetic accessibility of the CI. The β
compounds have lower vertical excitation energies (EFC) and therefore have less available energy
to reach the CI than the α compounds. This explains the higher Φfl of the β compounds in solution
and the correlation between Φfl and EFC. The mechanism is also consistent with the Bell–Evans–
Polanyi principle, which states that the barrier of a reaction correlates with the enthalpy of reac-
tion. Here, it is assumed that the enthalpy of reaction corresponds to the energy difference between
EFC and the CI energy and that the energy of the CI is approximately the same for all compounds.
Therefore, the barrier to access the CI correlates with EFC. In turn, Φfl depends on the accessibility
of the CI. These factors result in the correlation between Φfl and EFC. Among the 33 derivatives
there are some outliers, but the correlation can be recovered with an additional descriptor related
to the effect of pyramidalization of the CN-­substituted vinyl carbon on the S1/S0 energy gap [43].
A similar model has been proposed to explain the behavior of the pyrrolic analogs of DCS,
i-­DCSP, and o-­DCSP [58]. In this case, the “outer” o-­DCSP isomer (CN substituent at the o posi-
tion, see Figure 14.6a) is dark in solution and shows AIE, whereas the “inner” i-­DCSP isomer (CN
substituent at the i position) is luminescent in solution and shows aggregation-­caused quenching.
For both compounds, a CI has been found along the double-­bond torsion coordinate, and the dif-
ferent luminescence in solution is explained in terms of the vertical excitation energy, which is
somewhat higher for o-­DCSP, increasing the probability to access the CI and decay in a nonradia-
tive fashion. Significantly, the RACI effect used to design these compounds has been further
exploited using o-­DCSP as a fluorescent anion sensor [113]. In this case, complexation of o-­DCSP
with chlorine anions reduces molecular flexibility in a similar way to AIE and also results in fluo-
rescence by restricting the access to the CI.
In CN-­MBE [114], the Z form represented in Figure 14.6a shows AIE, whereas the E form is
nonemissive both in solution and in the crystalline form (note that we use the IUPAC Z/E nomen-
clature, whereas Refs. [89] and [114] use the inverse terminology, see Ref. [115]). Starting from
both forms, rotation around the central double bond is almost barrierless in the excited state and
leads to a CI with a rotation angle of 90°, which explains the absence of fluorescence in solu-
tion [89]. The different photophysics in the crystalline phase are explained as resulting from differ-
ent packing. The crystallographic structures have been simulated computationally, and the
fractional free volumes of the Z and E forms at 300 K are 22.1 and 24.2%, respectively, i.e. the Z
form is more densely packed than the E form. As a consequence, isomerization of the E form in the
crystal is still possible and the emission is inhibited. This was confirmed by dynamics simulations
with an excited-­state force field parametrized against SF-­TD-­DFT. In contrast to this, the Z form
shows AIE because the torsion is not possible due to the denser packing. The computations also
predict that AIE of the Z form will disappear at 500 K, since at that temperature the free energy
profile for the torsion becomes favorable also for this isomer.
A rational design strategy through control of CI accessibility has been used to design new
stilbene-­based AIEgens [116]. This has been achieved by locking the central double bond with an
alkyl chain that connects one of the central carbons with one of the phenyl rings (structures
BST[7]-­tBu, BPST[7], and BDPST[7] in Figure 14.6a). Both molecules show AIE because the chain
is flexible enough to allow for torsion and CI-­mediated radiationless decay in solution. In contrast,
the compounds with a shorter chain are fluorescent in solution because the torsion is blocked. In
428 14  A Global Potential Energy Surface Approach to the Photophysics of AIEgens: The Role of Conical Intersections

addition, BST[7]-­tBu and its analogs have twisted ground states and a higher vertical excitation
energy than the short-­chain analogs, which favors the access to the CI as seen for the DCS deriva-
tives. BST[7]-­tBu emits in the UV, while BPST[7] and BDPST[7] emit in the visible, thanks to addi-
tional phenyl substituent(s).
The decay in solution has been also associated to double-­bond torsion in a crowded stilbene
AIEgen, (E)-­1,2-­bis-­(2,4,6-­trimetylphenyl)ethene, which was studied with femtosecond transient
absorption spectroscopy [117]. This molecule has an excited-­state lifetime of 2.1 ps in solution and
comparison with the previous example also suggests that here a CI with the ground state may be
involved in the decay in solution. A CI associated to double-­bond torsion has been also shown to
be behind AIE and the mechanochromic behavior of the 2,3-­diaminomaleonitrile derivative
C1 [56] (see Section 14.6.2 for details).

14.3.2  Double-­bond Torsion vs Cyclization in TPE Derivatives


TPE is one of the prototypical AIEgens, and many of its derivatives have been used in AIE-­related
applications  [2]. Based on the previous discussion, torsion around the central double bond is a
potential route for radiationless decay of the excited state. However, TPE [118] and stilbene [119]
have long been known to undergo intramolecular cyclization in the excited state with formation of
an interannular C–C bond, leading to a dihydrophenantrene (DHP) product (see Figure 14.6b). In
stilbene, it was recognized earlier that the formation of the DHP product involved a CI  [120].
Therefore, the photophysics of the TPE derivatives in solution is best understood in terms of a
competition between two decay routes, photocyclization, and double-­bond torsion.
Experimentally, the photophysics in solution of TPE and five derivatives with different substitu-
tion (two with methyl ring substitution and three with locked rings) has been studied with ultrafast
time-­resolved transient absorption (see Figure 14.8) [121]. TPE has a complex relaxation dynamics
with three time constants of 0.39, 1.2, and 18.9  ps and the remaining molecules show similar
dynamics. Three possible relaxation channels were identified: C=C bond twisting with a time con-
stant of 1–2  ps, which was favored for the more flexible structures, including the parent TPE;
ultrafast formation of the photocyclized intermediate, which was favored in two of the ring-­locked
structures; and fluorescence, which was detected with significant intensity for the more sterically
crowded ortho-­tetramethylated derivative TPE-­4oM. These results illustrate the fact that both
double-­bond torsion and cyclization are possible and that the preferred path depends on the
molecular details.
Theoretically, the excited-­state decay of TPE has been simulated with TSH dynamics at the TDA-­
PBE0 level. The preferred decay path in the gas phase is photocyclization (75% out of 60 trajecto-
ries) [82], and it involves passage through a CI seam with a typical C–C distance of 2.0 Å. Only 5%
of trajectories decay to the ground state through the double-­bond rotation mechanism. These results
have been later reproduced simulating the decay in hexane solution with LC-­TD-­DFTB, including
explicit solvent molecules in the calculation [86]. In this medium, the ratio of trajectories that fol-
low cyclization vs double-­bond torsion is 73 : 27. Significantly, the two groups of trajectories have
different average decay times in hexane, 0.3 ps for cyclization and 2.2 ps for torsion, and it is sug-
gested that they correspond to the different time constants measured experimentally. The prefer-
ence for photocyclization is not consistent with the conclusions from the spectroscopic study of
Ref. [121], and it is suggested that the spectroscopic measurements may underestimate the amount
of photocyclization because DHP degrades rapidly, either after light absorption or as a result of the
formation of “hot states” during the dynamics. Another TPE-­based AIEgen that has been suggested
to decay in solution through a CI for cyclization is the thiophene-­substituted P4TA [122].
Sn τ1 τ2 τ3 τ4
τ1 < 0.4 ps

S*
C a b a H H
τ2 τ2 τ2
11.2 ps 1.2 ps <0.4 ps

hv
FL
b H H
τ3
18.9 ps

IM (S0)

τ4
S0 > 1 ms C H
H3C
TPE derivatives
PC + R20R22 (S0)

Figure 14.8  Ultrafast deactivation processes proposed for TPE and five derivatives studied experimentally. Three different pathways for the
decay of the excited state S* are represented by nonradiative decay channels (a) and (b) and a fluorescence decay channel (c). Examples of the
molecules with a dominant decay channel are shown in the box. τ1–τ3 are time constants of motion dominated by C=C bond elongation, quasi C=C
bond twisting coupled with phenyl torsion, and phenyl torsion, respectively. τ4 is the lifetime of the DHP intermediate (IM in Ref. [121]).
Source: Reproduced from Ref. [121] published by the Royal Society of Chemistry.
430 14  A Global Potential Energy Surface Approach to the Photophysics of AIEgens: The Role of Conical Intersections

The alternative between cyclization and double-­bond torsion has been further studied for the TPE-­4mM
and TPE-­4oM derivatives combining CASSCF and CASPT2 calculations of the decay paths with OM2/
MRCI TSH dynamics simulations [42]. TPE-­4mM is a typical AIEgen [2, 42, 123], whereas TPE-­4oM is
fluorescent in solution [124]. In TPE-­4mM, the path from the FC geometry to the cyclization CI is almost
barrierless, whereas the one leading to the twisted CI has a small barrier. Consistent with this, the pre-
ferred decay path in the simulations (88% out of 558 trajectories) is through the cyclization CI. This CI has
a peaked topology (see Figure 14.1c), and the trajectories branch in two groups. Approximately 45% of
trajectories form the cyclized DHP intermediate and the rest reverts back to TPE-­4mM. In contrast, in
TPE-­4oM, the paths to access both types of CI have sizeable barriers and no decay to the ground state is
observed for 568 trajectories run during 1 ps. This is consistent with the fact that TPE-­4oM has a relatively
stable emissive state compared to other TPE derivatives measured in Ref. [121].
Another derivative where cyclization is preferred over double-­bond torsion is THBDBA, where
the phenyl groups on the same end of the double bond are locked by ethylene groups (see
Figure 14.6b) [125]. This molecule was studied combining CASSCF and CASPT2 calculations of
the decay paths with OM2/MRCI TSH dynamics simulations [41]. THBDBA has a CI along the
cyclization coordinate, with an interannular C–C distance of 1.60 Å and one along the double-­
bond twist coordinate. In the dynamics, 57.5% out of 367 trajectories decay to the ground state in
less than 5 ps and all do it at the region of the cyclization CI. Out of these trajectories, one-­third
yields the DHP product after the passage through the CI, whereas two-­thirds return to the reactant.
No trajectories decay at the bond twist CI because the locked THBDBA molecule has not enough
flexibility to avoid clashes of the phenyl substituents during the twist, which precludes it from
twisting. In addition, more than 90% of trajectories staying in S1 also undergo cyclization and after-
ward remain trapped in an S1 minimum of the DHP form. In some trajectories, excited-­state cycli-
zation is followed by decay to the ground-­state through the CI and regeneration of the reactant.
This suggests that long-­time irradiation of a THBDBA solution will result in a series of cycles
where the THBDBA molecules are excited and undergo cyclization; at a later stage during irradia-
tion, the DHP product molecules are also excited and may regenerate the THBDBA in its decay,
which will ultimately lead to a photostationary state (see Figure 14.9). Following this mechanistic
scheme, the yield measured experimentally of 23% corresponds to the photostationary state.

Erel (eV)

5.0
S1

hv

hv

0.0 S0
S0-Min S1-Min (S1/S0)X Cyc-S1-Min Cyc-S0-Min

Figure 14.9  Mechanistic scheme proposed for THBDBA composed of excited-­state cyclization and reversion,
leading to a photostationary state. Source: Reprinted from Ref. [41] with permission from John Wiley and Sons.
14.3 ­CI-­centered Global PES for AIEgen 431

In contrast to the previous cases, the methoxy-­substituted and fluoro-­substituted AIEgens


TPE-­2OMe and TPE-­2F decay in solution through the double-­bond twist mechanism, as shown by
TD-­DFT calculations of the PES [126]. Double-­bond twist is also the radiationless decay mecha-
nism in solution followed by the furan analogs of TPE and TFE [27]. In this case, the intramolecu-
lar cyclization path is disfavored by the fact that the most abundant conformer of the molecule is
the one with the oxygen atoms lying on the inner side of the molecule, which hinders the forma-
tion of new C–C bonds.

14.3.3  Excited-­state Intramolecular Proton Transfer (ESIPT) Compounds


ESIPT compounds have different absorbing and emitting species due to the occurrence of
fast ESIPT before emission. This makes them particularly interesting luminescent devices, since
the ESIPT is associated to a large Stokes shift. In addition, these compounds are less likely to suffer
from ACQ since in a crystal, the emitting molecule has a lower excitation energy than the sur-
rounding molecules that have not undergone the proton transfer. This prevents the formation of
nonemissive delocalized states.
The RACI mechanism has been shown to be valid for several ESIPT AIEgens (see Figure 14.6c).
In this case, excitation first induces ESIPT, and the newly formed tautomer decays through a CI to
the ground state in solution. This mechanism is followed by HPIP [127], which is the first AIEgen
where the involvement of a CI was invoked to explain the lack of fluorescence in solution [46]. In
solution, the PES has a barrierless ESIPT path followed by an almost barrierless path to a CI where
the two rings lie perpendicular to each other. Mechanistically, the appearance of the CI can be
understood following similar ideas to those of the previous section. ESIPT generates a tautomer
with a new double bond connecting the two proton donor and acceptor moieties, and the CI is of
the double-­bond twist type. Such a mechanism is well known from other ESIPT compounds that
function as photoprotectors [128, 129].
Another AIEgen following this mechanism is HPQ [130, 131], where the decay has been investi-
gated combining PES studies at the MS-­CASPT2//CASSCF level of theory with dynamics at the
TD-­PBE0 level [48]. The PES in solution is similar to that described for HPIP. In the trajectories,
ESIPT takes 20 fs in average, and 96% of 188 trajectories reach the twisted CI region in 729 fs. The
RACI mechanism is confirmed by PES calculations in the crystal showing that the ESIPT stays
barrierless, while the ring rotation has a large energy cost and is therefore suppressed.
The case of the chalcones HC1 and HC5 [132] (see Figure 14.10) is somewhat different. HC1 is
an AIEgen, whereas the methoxy-­substituted HC5 derivative is nonemissive in solution and in the
solid phase. These compounds were studied combining CC2 and CASSCF calculations of the PES
with ADC2 excited state dynamics [70, 71]. The compounds have two accessible CIs. One CI is
similar to the ones described for HPQ and HPIP and is reached after ESIPT and torsion about the
newly formed double bond. The corresponding decay path is named the K channel. The other CI
corresponds to the original tautomer and involves torsion around the carbon-­phenyl bond and in
the case of HC1, also around the stilbene-­type central bond, and the corresponding path is the E
channel. Both paths have no barriers on the PES in solution, and the preference for decay along
one or the other channel depends on substitution. For HC1, both paths have similar probability
and the E : K channel splitting in the dynamics is 52 : 48, whereas for HC5 the K channel is favored
(20 : 80). The preference toward proton transfer found in HC5 is related to the higher stability of
the keto form in this tautomer. From the point of view of AIE, calculations in the crystal environ-
ment show that the decay channels are only restricted effectively in HC1, whereas in HC5, the K
channel is still accessible. Thus, the different photophysics of the HC1 and HC5 crystals is due to
the fact that in HC5 not all deactivation channels are restricted.
432 14  A Global Potential Energy Surface Approach to the Photophysics of AIEgens: The Role of Conical Intersections

H
H
O O
O O

HC1 HC5
AIE-active N Non-emissive N

OMe

A2 Dimer

A1 Dimer

P1 Dimer

P2 Dimer

Figure 14.10  Molecular and crystal structures of the chalcones HC1 and HC5. HC1 displays AIE, whereas
HC5 is nonemissive in both aqueous and solid phase. Note the parallel (P) and antiparallel (A)
arrangements in the crystal. Source: Reprinted with permission from Ref. [71]. Copyright 2017 American
Chemical Society.

2-­Hydroxylphenylpropenone derivatives (HP), which are mono-­aryl analogs of chalcones (HC),


also display AIE in the solid state, but with much higher quantum yields (0.72–0.84)
(Table  14.1)  [134]. The CIs, which are inaccessible in the solid state, also feature a rotated-­
pyramidalized structure [72]. In the case of HP1, the CI lies 1.08 eV above the initial excitation
level making it classically inaccessible (MS-­3-­CASPT2(12,11)/6-­31G(d):AMBER//SA-­2-­
CASSCF)/6-­31G(d):AMBER level of theory) (Figure  14.11). Reorganization energies to the K*
minima in the solid state ranged from 0.53 (HC3) to 0.98 eV (HC5) in the HC series and from 1.06
(HP2) to 1.23 eV (HP1) in the HP series. As a consequence, the E* minimum is not stable for the
HP series. The reason for the better quantum yields of these compounds in comparison with HC
derivatives is the higher efficiency of the intramolecular proton transfer, which localized more
effectively the electron density on individual molecules. Thus, competing excitonic pathways
become less likely and the quantum yields of the HP series are larger (see more details in
Section 14.5).

14.3.4  Ring Puckering


Together with double-­bond torsion, ring puckering is one of the most common coordinates leading
a molecule to a CI. Many organic chromophores have aromatic or heteroaromatic ring structures
that are planar in the ground state, and excitation facilitates loss of the planarity in a similar way
14.3 ­CI-­centered Global PES for AIEgen 433

Table 14.1  Emission quantum efficiencies (Φfl) of HC and HP molecular crystals [132–134].

H H
O O O O

R1 N R3
R2 R4

HC HP

R1 R2 Φfl R3 R4 Φfl
HC1 H H 0.32 HP1 H H 0.74
HC2 CH3 H 0.25 HP2 F H 0.84
HC3 OCH3 H 0.26 HP3 H OCH3 0.77
HC4 H CH3 <0.01 HP4 H F 0.72
HC5 H OCH3 <0.01
HC6 F H 0.41 (α)
0.05 (β)a
HC7 H F 0.10
a
 Different polymorphs of HC6.

Crystal
Vacuum

3
Energy (eV)

FC K*min K*MECl

Figure 14.11  Potential energy surface for HP1. Crystal geometries were obtained with ONIOM((TD)-
­ωB97X-­D/6-­31G(d):AMBER) and the energies were calculated at MS-­3-­CASPT2(12,11):AMBER level.
Geometries in vacuum were obtained with (TD)-­ωB97X-­D/6-­311++G(d,p) and energies with MS-­3-­
CASPT2(12,11). Source: Reproduced from Ref. [72] with permission from the Royal Society of Chemistry.

to what we have discussed for double bonds. Therefore, ring distortion becomes possible with low
energy cost in the excited state and a high one in the ground state, which results in the appearance
of energetically accessible CIs. The ring distortion is usually accompanied by bond length inver-
sion in the ring, where short bonds are elongated and long bonds are shortened.
434 14  A Global Potential Energy Surface Approach to the Photophysics of AIEgens: The Role of Conical Intersections

The AIEgens where radiationless decay in solution has been shown to involve a ring-­puckered
CI are shown in Figure 14.6d. DMTPS is a prototypical AIEgen [135], and the excited-­state decay
in solution involves passage through a CI characterized by ring puckering and a flapping motion
of the phenyl substituents. This was first suggested by PES calculations combining CASSCF and
CASPT2  with TD-­DFT (see Figure  14.5)  [11] and it has been recently confirmed by TD-­DFTB
dynamics calculations, where the decay to the ground state occurs in 1.8 ps in the gas phase and
6 ps in hexane solution [86]. The predominant initial motion in the excited-­state dynamics is a
bond length inversion of the single and double bonds in the silole ring. A similar mechanism is
followed by TPC [136]. In this case, the CI was located on the PES at the CASSCF level [47]. It has
a similar structure to that described for DMTPS, although from the electronic structure point of
view there is also significant contribution of a resonance structure with an intra-­annular C2–C5
bond (resonance structure III in Figure 14.12). This results in a shortening of the C2–C5 distance
to 2.09 Å at the CI in solution. The initial phase of the relaxation (500 fs) was studied with TD-­DFT
dynamics in the gas phase, where it was dominated by bond length inversion. This induces a
buildup of biradical character in the ring (see resonance structure II in Figure 14.12) and this in
turn favors a planarization of the phenyl rings that translates into an activation of low-­frequency
torsional modes. Significantly, no activation of the puckering modes that lead to the CI was
observed during the 500 fs dynamics, which suggests that the dynamics follows a two-­phase behav-
ior. The prediction that the decay takes longer than 500 fs is consistent with the lifetimes simulated
for DMTPS [86]. The reorganization energy of the torsional modes, in particular out-­of-­plane rota-
tion of the phenyl rings, is significantly reduced in the crystal (1425 cm−1) with respect to solution
(2493 cm−1). The AIE picture for DMTPS and TPC has been completed by calculations in the crys-
tal, and in both cases, it has been shown that the CI becomes energetically inaccessible, opening
the way for fluorescence.

I
1.6
R2
R1 R3

1.5
II
d =1.46 Å
d (Å)

1.4

III
1.3 d =1.27 Å

1.2
0 20 40 60 80 100
Time (fs)

Figure 14.12  Time evolution of the bond inversion coordinate (d = R1 + R3−R2) during the first 100 fs of
TPC dynamics and relevant resonance structures. Source: Reprinted from Ref. [47] with permission from John
Wiley and Sons.
14.3 ­CI-­centered Global PES for AIEgen 435

A common puckering motif can be found for the BODIHY [49, 67], BDAA [93], and DBU-­C [137,
138] AIEgens shown in Figure 14.6d. In all cases, the distortion leading to the CI in solution is a
pronounced puckering of a ring atom out of the plane of the ring, with the corresponding substitu-
ent being bent by about 90° with respect to the plane. These atoms are highlighted in Figure 14.6d.
The results for BDAA have inspired the design of a pyrene-­derived AIEgen with two dimethyl-
amino substituents introduced in the K region, which stabilize the minimum energy CI in
solution [139].
As discussed in Section  14.2.3.1 the AIE observed in BODIHY was initially assigned to non-­
Kasha emission, i.e. emission from S3, on the basis of TDA-­PBE calculations [67]. However, later
TD-­B3LYP and XMC-­QDPT2 calculations suggest that the fluorescence increase in more rigid
environments actually follows the RACI model [49]. The energetically accessible CI in solution has
a ring-­puckered structure, with the phenyl substituent lying perpendicular to the plane of the
BF2·hydrazone complex. This distortion is described as flip–flop or open–closed umbrella-­
like motion.
Similar motions are involved in the radiationless decay mechanisms of BAA and DBU-­C [138].
In the case of DBU-­C, the identification of the carbon in the pyridone ring as the one that under-
goes the puckering explains the dependence of the photophysics on the side rings attached to the
core. In the case of a seven-­membered or six-­membered side ring, such as DBU-­C, the molecule is
an AIEgen that follows the RACI model, but in the case of a five-­membered side ring, the ring
strain hinders the puckering, so that the CI becomes energetically disfavored and the molecule
emits strongly in solution.
A final example is provided by the DNCOT AIEgen shown in Figure 14.6d [140, 141]. In the
ground state, the COT moiety has a tub shape, and relaxation upon excitation involves ring pla-
narization and folding [74]. The CI is energetically reachable in solution but requires a large distor-
tion, so that the access is restricted in rigid environments and AIE occurs. In contrast to this, the
smaller DPCOT derivative (Figure  14.13) decays in a radiationless way through a different CI
involving puckering of a single CH group. This motion is similar to the ones described above for
BDAA or DBU-­C, but it does not have large steric requirements because there are no large substitu-
ents at the puckered atom. Therefore, it is not restricted in rigid environment, so that DPCOT is
also not emissive in aggregate phases. Finally, the anthracene analog DACOT is not an AIEgen
because, even though it has a similar CI to DNCOT, accessing the CI involves a high barrier and
the molecule is emissive also in solution.

14.3.5  Bond Stretching


One final CI type encountered in AIEgens is associated to bond stretching. The role of such a CI in
the radiationless decay in solution has been identified for the B18H20(NC5H5)2 cluster (Figure 14.6e),

O O

HN NH

O O
DPCOT

Figure 14.13  DPCOT and CI leading to radiationless decay in solution and crystalline phase. The atoms
involved in the distortion to reach the CI are marked in turquoise.
436 14  A Global Potential Energy Surface Approach to the Photophysics of AIEgens: The Role of Conical Intersections

which is composed of two conjoined boron hydride subclusters bearing two pyridine substitu-
ents  [45]. This type of CI is usually found when the excitation involves transfer of an electron
from/to an occupied/empty σ or σ* orbital. Removal of an electron from a bonding orbital or pro-
motion to an antibonding one weakens the bond and induces the stretch. Similar to the previous
cases, the bond stretch has a high energy cost for the ground state, and this leads to the appearance
of a crossing on the PES.
In the B18H20(NC5H5)2 case, the excitation involves a CT from an occupied orbital of the B18H20
cluster, which has bonding character along the B–B bonds, to an empty pyridine orbital. Relaxation
in solution involves torsion of the pyridine rings and finally a stretch of two B–B bonds from 1.848
and 1.973 Å at the FC structure to 2.194 and 2.492 Å (see Figure 14.6e). This also results in a flap-
ping of the pyridine rings. Potentially, this process could be a source of photodegradation of the
molecule, but the calculated energy profile suggests that the CI has a sloped topography, which
would favor the recovery of the intact molecule after the decay. The role of the CT state also
explains the dependence of the photophysics on the polarity of the solvent observed experimen-
tally. However, in the related B18H20(NC5H5) and B16H18(NC5H5)2 clusters, bond stretching is not
accompanied by significant distortions and no fluorescence enhancement in the solid state
occurs  [142]. This shows that the twisting and flapping motions of the pyridine rings are also
important to enhance fluorescence.

14.3.6  A View of AIE Based on the RACI Model and the Global PES
The examples compiled in this section show that the RACI model is a useful tool to understand the
photophysics of AIEgens in solution. There is a wide variety of AIEgens that follows this model.
From a mechanistic point of view, the CIs can be classified into several categories, and the CI can
be explained in terms of the orbital changes that occur during the excitation. More importantly, the
RACI behavior allows to rationalize the role of the different molecular components in the photo-
physics. On the one hand, the distortion leading to the CI is always localized in a specific part of
the molecule such as a double bond, a ring, or even a particular atom of that ring. On the other
hand, the role of the substituents attached to the distorting part can be understood as that of ampli-
fying the molecular distortion, so that it can be blocked in rigid environments. Some examples of
this are the phenyl substituents in DPDBF, DMTPS, or TPC, the alkyl substituents in DBU-­C, or
the pyridine groups in B18H20(NC5H5)2. Significantly, DPCOT provides an example of the opposite
case, where the amplifying substituents are not present in the molecule. In this case, access to the
CI is not blocked in rigid media and the molecule is not an AIEgen.
The previous examples also show how the study of the global PES allows one to understand
some aspects of the photophysics that cannot be understood with the more general models. This is
particularly the case for molecules that have similar structures but remarkably different photo-
physics, such as α-­DCS vs β-­DCS, HC1 vs HC5, the COT derivatives, or DBU-­C vs DBN-­C. In these
cases, the study of the PES explains why some compounds behave as AIEgens, whereas oth-
ers do not.

14.4 ­Crystallization-­induced Phosphorescence

The term CIP was first defined in 2010 to describe the enhancement of phosphorescence observed
in derivatives of benzophenone (BP) due to aggregation [30]. Since then, crystallization has been
exploited as a strategy to engineer organic materials with long-­persistent phosphorescence [143,
14.5 ­Effect of Intermolecular and Intramolecular Interactions on the Photophysics of AIEgens 437

144]. These materials find applications in organic light-­emitting diodes (OLEDs), photovoltaic
cells, solid state lasers, biological sensors photosensitization, and photocatalysis. Phosphorescent
OLEDs can potentially take advantage of 75% of electrically generated excitons, which are triplets.
BP has been utilized in both biomolecular and material science applications. Its excited state
dynamics is highly influenced by the environment. The ISC rates are in the range of 6.5–16  ps
depending on the solvent, and the global value in the crystal is of 25 ps [145]. The solutions of BP
in various nonpolar and polar solvents are transparent at room temperature [30]. However, when
cooled at cryogenic temperature, they are highly emissive and the BP crystal displays a strong blue
emission. Nonadiabatic dynamics simulations provide information on the population of the triplet
state. Calculations of BP in gas phase including the effect of SOC show that the mechanism is
modulated by the mixed nπ*/ππ* features of T1 [146, 147]. The transition from S1 to T1 is very effi-
cient, still its detailed mechanism has been under discussion since the population of T1 can be
obtained either by direct population from S1 or it can involve the population of an intermediate
state S1 ⟶ T2 ⟶ T1 [146–148]. A T2/T1 crossing structure facilitates the second mechanism. T1/
S0 crossings have not been reported so far. Once the triplet state is populated, many deactivation
pathways are possible including energy transfer. In condensed phase, these pathways are ham-
pered, enhancing phosphorescence emission.
Aggregation can also modify SOC, making transitions to triplet states more efficient. This is the
case of isophthalic acid (IPA) and 4,4′-­bis(9H-­carbazol-­9-­yl) in cocrystal form with chloroform,
where the electrostatic interactions change the nature of S1 from nπ* in the gas phase to ππ* in the
solid state [149]. Thus, SOC between S1 and intermediate triplets states of nπ* character is increased,
making ISC more likely. Hydrogen bonds between the C=O and OH groups in IPA and tereph-
thalic acid, whose vibrations can drive deactivation decay, are also blocked in the solid state.

14.5 ­Effect of Intermolecular and Intramolecular Interactions


on the Photophysics of AIEgens

14.5.1  Excitonic Effects in AIE


Specific intermolecular interactions occurring upon aggregation can modulate the excited state
mechanism and determine whether excited states are localized on monomeric units or extended
over two or more molecules. According to Kasha’s exciton theory, dimers are classified depending
on the orientation of their transition dipole moments as J (head-­to-­tail) or H (side-­by-­side) [150].
For J dimers, the excitation energy of the bright state (S1) is shifted to the red and its oscillator
strength is twice the value of the individual chromophores. In the case of H dimers, the bright state
(S2) is shifted to the blue and S1 is dark. J aggregation has been suggested as one of the mechanisms
for AIE. Interestingly, H aggregation has been hypothesized as the reason for CIP and the stabiliza-
tion of long-­lived triplet states [143]. It has also been argued that the term AIE should be reserved
for cases involving intermolecular electronic interactions, whereas the cases involving steric con-
finement should be classified as solid-­state enhanced emission (SLEE) [94].
Kasha’s theory is based on the assumption of long-­range Coulombic interactions between the
monomers. However, close intermolecular interactions are very common in concentrated solu-
tions and crystals. π-­stacked dimers with intermolecular distances smaller than 3.5  Å can have
significant overlaps between the electron densities of the independent molecules [151]. Therefore,
short-­range interactions, including exchange and superexchange interactions, can be of similar
magnitude as of the long-­range ones. For distyrylbenzene crystals, experiments and modeling have
438 14  A Global Potential Energy Surface Approach to the Photophysics of AIEgens: The Role of Conical Intersections

shown that the emission efficiency is the result of a combination of factors, H aggregation that
tends to induce a decrease of kr, and the restriction of nonradiative mechanisms that reduce the
values of knr [94]. The combined effect of excitonic coupling in the aggregated phase, together with
restriction of low-­frequency modes and E/Z isomerization, has been also investigated for star-­
shaped molecules with benzene and s-­triazine cores, and the appearance of AIE or SLEE is sensi-
tive to the molecular details [152].
The impact of intermolecular interactions on the excited states can be measured by the values of
exciton coupling, which gives an idea of how much the energy levels are affected by the interaction
with a neighboring molecule. For the crystals of typical propeller systems, such as DMTPS, TPC, and
TPE (Figure 14.6), where the intermolecular interactions are mainly C–H···π, C–H···C, and H···H, the
values of exciton couplings are in the range of 0.009–0.013 eV [47]. Similarly, they are in the range of
0.027 ± 0.024 eV for TFE [27]. For these systems, the reorganization energies are comparatively larger
than the exciton couplings and the excitations are mostly localized on independent chromophores. No
significant CT between the units is observed. However, work based on FGR has found that knr for both
H and J dimers increases with the exciton couplings, even with small couplings [91].
In recent work, the effect of the 3D structure on the emissive response of the excited state proton
transfer chromophores HC and HP was addressed (Table 14.1) [72]. For proton transfer systems,
excited-­state relaxation can happen in the K* or E* states. Both processes compete with exciton
hopping, which prevents the localization of the electron density. The exciton couplings signifi-
cantly depend on the structure of the dimers, with some of the values in the order of 0.1  eV
(Figure 14.14). However, reorganization energies, in particular for the K* relaxation, are very large,
guaranteeing a high localization of the electron density. The analysis of several dimers from the
molecular crystals shows that herringbone motifs are frequently found in most of the crystals of
the HC series, while T-­shaped dimers are more common in the HP series. However, the low-­
emitting crystals of the HC series, namely HC4, HC5, and HC7, display a large population of
T-­shaped motifs like those found in the high-­emitting HP crystals. Therefore, there is no direct
relationship between the crystal structure and their quantum yields. Additionally, J or H dimers

160 4.53 Å 7.53 Å


Mean coupling
6.22 Å 7.12 Å 3.84 Å
140 5.63 Å 7.82 Å 4.24 Å

120
7.33 Å
5.43 Å
100
Jij (meV)

80
5.35 Å
60

40

20

-1 -2 -3 -4 -5 -6 -7 -1 -2 -3 -4
HC HC HC HC HC HC HC HP HP HP HP

Figure 14.14  Exciton couplings for the HC and HP series showing the structure of the dimer with the
largest coupling. The mean couplings are displayed using cross signs. Source: Reprinted from Ref. [72] with
permission from the Royal Society of Chemistry.
14.6 ­New Challenge 439

were assigned based on the oscillator strengths of the excitations (S1 and S2 states). For most sys-
tems, the population of H and J dimers was very similar, except for HC4, HC5, and HC7, which
have a larger prevalence of H dimers (78, 66, and 100%, respectively). Interestingly, these dimers
are also those with low quantum yields featuring accessible CIs.

14.5.2  Effect of Intramolecular and Intermolecular Interactions on Emission Color


Intermolecular and intramolecular interactions can also tune the wavelength of the emission, and
understanding these effects in crystals is crucial for the design of new luminophores. In polymorph-­
dependent luminescence, different crystal packing leads to fluorescence with different colors. The
ESIPT molecule 6-­CN-­HPIP (see Figure 14.6c for the parent HPIP), which absorbs in the enol form
and emits in the keto form, has three polymorphs with different fluorescence wavelengths. It has
been shown that the emission color depends on the orientation of the π-­stacked arrangements in
the different crystal forms and also on the fact that the emitting keto form and the surrounding
enol form have opposite dipole moments [153].
The combination of through-­space intermolecular and intramolecular interactions with AIE can
also result in large Stokes shifts, shifting the crystal emission to the red. For a series of phenyl-­ring
substituted molecular rotors (Ph3C3H, Ph5C5H, Ph6C6 and Ph7C7H), emission can happen either
from intermolecular or intramolecular π-­dimeric states (Ph5C5H and Ph7C7H)  [154].
1,2-­Diphenylethane (s-­DPE) and its hexamethylated 1,2-­bis(2,4,5-­trimethylphenyl)ethane
(s-­DPE-­TM) derivative have a Stokes shift of approximately 100 nm in the solid state and it has
been suggested that this is due to the formation of excited-­state through-­space complexes with a
large overlap between the phenyl groups of neighboring molecules [155]. The resulting emission is
termed clusteroluminescence [156]. The formation of the excited-­state complexes was confirmed
by transient absorption measurements and it is thought to occur after a large amplitude molecular
motion in the excited state that brings the phenyl groups together. Consistent with this, the clus-
teroluminescence increases at high temperatures that favor molecular motion. Similarly,
1,1,2,2-­tetraphenylethane has a Stokes shift of almost 200 nm in the solid state. This leads to unex-
pected strong fluorescence (Φfl = 70%) in the visible range (400–600 nm) in a molecule with appar-
ently “isolated” phenyl rings, which in this case was explained by intramolecular through-­space
conjugation. In TFE (see Figure  14.2), intramolecular O⋯O interactions lead to a red mor-
phochromism (a red shift of 27 cm−1) when going from amorphous aggregates to crystals. These
interactions in the excited state are strong enough to decrease the intramolecular O–O distance to
2.70 Å, which is below the sum of oxygen van der Waals radii [27]. Analysis based on Bader’s atoms
in molecules theory [157] showed that there are bond critical points between the oxygen atoms,
which indicate significant overlap between the oxygen electron densities. These cases show that,
beyond Kasha’s exciton theory, specific intramolecular and intermolecular interactions in the crys-
talline state can have substantial effects on the photophysical properties.

14.6 ­New Challenges

14.6.1  The Role of Dark States in AIE


The restriction of access to a dark state (RADS) [158] mechanism has been proposed recently as a
new mechanism behind AIE, where the lack of fluorescence in solution is due to population of a
dark, weakly emissive state. In general, emissive properties are determined by the nature of the
440 14  A Global Potential Energy Surface Approach to the Photophysics of AIEgens: The Role of Conical Intersections

emitting state, which is S1 in most cases, according to Kasha’s rule. However, during the excited-­
state relaxation, the molecule can transit through states of different diabatic characters. As the
emission intensity depends on the oscillator strength, the transition to nπ* and CT states, which
have low oscillator strength because they have little superposition with the ground state wave func-
tion, results in weak emission intensity. In RADS, inhibiting the population of these states is an
alternative pathway to achieve AIE. Figure 14.15 exemplifies this idea by showing different sce-
narios involving the population of dark states. In contrast with other excited-­state decay mecha-
nisms such as vibronic coupling (A) and decay through CIs (B), if the dark state is more stable than
the bright state (C), a significant fraction of the excited state population can be trapped in the dark
state (D), leading to a loss of emission. This situation can be reversed by creating a barrier to access
the dark state (E) and/or destabilizing it (F). The investigation of these mechanisms requires a
detailed analysis of potential energy surfaces of different diabatic states and the determination of
crossings between them.
One example of RADS is (9-­anthryl-­methyl)bis(2-­pyridylmethyl)amine (APA), where either
complexation to Zn2+ or solid state aggregation hinders the population of a nonemissive nπ* state,
enhancing the emission. Complexation to Zn2+ is particularly effective with a 99.8% quantum yield
since it blocks the nπ* electron transfer and both S1 and S2 become ππ* states (Figure 14.15b). In
addition to this, the formation of the organometallic complex with the lone pairs of the three nitro-
gen atoms rigidifies the molecule and hinders other nonradiative decay mechanisms. The RADS
mechanism might be the reason for the enhanced emission observed in several organometallic
complexes that have been previously labeled as photoinduced electron transfer systems [159, 160].
In the solid state, relaxation to the dark nπ* state is not possible because intermolecular interac-
tions and exciton splitting raise its energy (case F, Figure 14.15a).

14.6.2  Pressure-­induced Emission Enhancement


Applying pressure on a material can result in the modification of intermolecular interactions
and  drastic changes in the crystal structures. This is the origin of pressure-­induced emission
enhancement (PIEE), a phenomenon related to AIE which contrasts with the more common
pressure-­caused quenching [161]. Some AIEgens such as DMTPS,TPC, 2,3,4,5-­tetraphenylthiophene,
2,3,4,5-­tetraphenylfuran, and carbazole display PIEE in a range of pressures [162–164].
A common behavior for PIEE chromophores is to observe an increase of emission for a range of
pressures due to the restriction of vibrations driving the system to the ground state. At larger pres-
sures, π-­stacked complexes are formed, which accelerates the nonradiative decay and results in
emission quenching. For example, the carbazole crystal features herringbone motifs stabilized by
N−H⋯π and C−H⋯π interactions [164]. This material shows PIEE under 1.0 GPa associated with
the restriction of the N–H vibration. For larger pressures, π⋯π contacts between the molecules are
increased and emission is quenched. This mechanism has been analyzed considering the variation
of crystal structures with pressure and the Hirshfeld surfaces calculated for crystals modeled for
different pressures.
The mechanism behind PIEEE in DMTPS was recently investigated using the thermal vibration
correlation function formalism based on FGR [55]. The structures of the crystals at different pres-
sures were optimized using DFT-­D periodic boundary conditions and subsequent QM/MM with
TD-­B3LYP/6-­31G(d) and GAFF. The quantum efficiency increased in the range from ambient
pressure to 10.9 GPa. Radiative decay constants were not affected by pressure, but the nonradiative
constants decrease with compression until 5.06 GPa and then level off with a saturation of the
quantum efficiency (~80% [0 GPa] to ~97% [5.06 GPa]).
14.6 ­New Challenge 441
(a)
A FC B FC C FC
Dark state Bright state

Nuclear coordination Nuclear coordination Nuclear coordination

Φsoln =1.5% Φsoln =0.2%


N
Φsoln = 0.6%
N

D E Bright F
Bright Dark
state Dark
state Dark state
state state
Bright
Non-rad. state
Non-rad.
Ex. Ex. Fluo. Ex. Fluo.

Nuclear coordination Nuclear coordination Nuclear coordination


(b)
A B
S1 (π, π*)
S2 (π, π*)
(3.827 eV)C S2,min Bright state
D (3.831 eV)
S2,min
(π, π*)
E (3.28 eV)
(3.18 eV)A HOMO LUMO + 1
S1,min
B (2.97 eV)
Bright state
Fluo. S1,min
448nm
(π, π*)
Nuclear coordinate HOMO LUMO
C D

Dark S1 S2,min Bright state


state S2,min
S1,min (π, π*) (π, π*)
(π, π*)
Vertical transition
Energy

HOMO-1 LUMO
3.28 eV

2.46 eV

2.71 eV

Fluo.
504nm
Bright state
S1,min
S0 (π, π*)

Monomer Dimer HOMO LUMO


E

Molecular
motion
RADS RADS

Zinc complex Isolated APA Crystal

Figure 14.15  (a) Illustration of mechanisms associated with different potential energy surfaces showing
the role of dark states. (b) Exemplification of the RADS for APA in the solid state and its complex with Zn2+.
Source: Reprinted from Ref. [158] with permission from John Wiley and Sons.
442 14  A Global Potential Energy Surface Approach to the Photophysics of AIEgens: The Role of Conical Intersections

A detailed investigation of the role of CIs on the PIEE mechanisms is yet to be reported. However,
recent work has addressed pressure-­caused quenching and has established for the first time a rela-
tionship between the RACI model and mechanoluminescent properties, by analyzing the effect of
grinding on two related AIEgen crystals: 2-­((E)-­((9H-­fluoren-­2-­yl)methylene)amino)-­3-­
aminomaleonitrile (C1, see Figure  14.6a) and 2-­amino-­3-­((E)-­(4(diphenylamino)benzylidene)
amino)maleonitrile (C2) [56]. Crystalline powder of C1 shows a strong green emission (Φfl = 7.4%),
which significantly decreases after grinding (Φfl  =  0.73%) or after application of pressure (50%
reduction at 100 MPa). A CI has been associated with these quenching phenomena on the basis of
TD-­PBE0 and CASSCF calculations. The CI lies along the double-­bond torsion coordinate that
connects an E/Z isomer pair, following the model described in Section 14.3.1. Relaxation in the
intact crystal is not possible because closely arranged intermolecular interactions restrict intramo-
lecular rotation, leading to AIE. However, mechanical stimuli destroy these interactions with a
consequent decrease of Φfl. In contrast to this, grinding in C2 induces a bathochromic shift without
significantly affecting Φfl. A CI could not be found for this molecule, which supports the idea that
decay at the CI is related with the pressure-­induced quenching in C1.

14.6.3  AIE in Transition Metal (TM) Compounds


Some luminescent TM compounds also show AIE [165]. Emission arises from the lowest triplet
state because the TM enhances SOC and ISC and the phenomenon has also been termed
aggregation-­induced phosphorescence emission. Four possible reasons for AIE in TM complexes
have been proposed  [166]: (i) reorganization of the metal coordination sphere in the aggregate
state that may enhance SOC and increase radiative deactivation efficiency. (ii) Intermolecular
delocalization of the ground or emitting states. (iii) RIM mechanism. (iv) Restrictions along the
molecular geometry relaxation that, similar to the RIM mechanism, would result in a decrease of
knr. Most mechanistic studies favor the RIM mechanism, and here, we have reviewed some repre-
sentative contributions of theory.
A study for a series of seven Pt(II) complexes showing AIE [167] (Figure 14.16a) proposes an
explanation for AIE called restricted structural distortion in the excited state that is actually remi-
niscent of the RADS mechanism. The complexes have one 2-­phenylpyridine (ppy) and one N^O
ligand and are weakly emissive in solution. This has been explained on the basis of B3LYP calcula-
tions in solution, where T1 is a nonemissive state dominated by mixed N^O and Pt(II)-­centered
transitions. Access to this state involves loss of the quasi-­square-­planar Pt(II) configuration, with
pyramidalization at the Pt(II) center. In the crystal, PBE calculations show that this motion is
restricted and T1 becomes an emissive state dominated by mixed ppy and metal-­centered transitions.
Ir(III) complexes showing AIE have also been extensively studied. For a group of tris-­
cyclometalated cationic complexes (Figure 14.16b), B3LYP calculations have been used to study
whether the lack of emission in solution may be due to nonradiative decay of the triplet state at a
T1/S0 crossing involving a metal-­centered triplet state [168]. However, no differences are found in
the corresponding path for a luminescent complex and two AIEgens, which suggests that this
mechanism is not responsible for the decay. As an alternative explanation, the authors postulate
that the AIE compounds populate a weakly emissive interligand CT in solution, whereas the lumi-
nescent complex emits from a mixed metal–ligand/ligand–ligand charge-­transfer state. Moreover,
the AIE compounds have larger end-­group substituents, leading to larger Huang–Rhys factors and
a larger knr in solution; the RIM mechanism prevents the nonradiative decay in the solid state.
The RIM mechanism has also been favored for other Ir(III) complexes. For a series of 12 mono-­
cyclometalated complexes with nonchromophoric aryl phosphine ligands that show AIE
14.7  ­Conclusions and Outloo 443

(a) (b)
+
F

Ar N
N
N N
F
Pt
Ir
N O F
N N R
N
N N

(c) (d) (e)


P(Ar)xRy +
PPh 3
Ar
N N
H N H N
C
Ir Ir Ir
N H C O
C Cl
N
PPh 3
P(Ar)xRy

Figure 14.16  TM compounds showing AIE introduced in Refs. [167] (a), [168] (b), [169] (c), [166] (d),
and [170] (e).

(Figure 14.16c) [169], B97-­D calculations show that the aryl phosphine rings have different confor-
mations in solution and in the crystal. This suggests that rotation of the rings is hindered in the
crystal, consistent with a RIM explanation of AIE. Calculations for a similar complex series
(Figure 14.16d) [166] also show that T1 in solution has a considerable geometry change compared
to the ground state, namely a bending along the P–Ir–P axis, and this motion is probably also
involved in the RIM mechanism. The same mechanism has been proposed for bis-­cyclometalated
complexes with N^O ligands (Figure 14.16e) [170].

14.7 ­Conclusions and Outlook

Theory and computations have approached AIE in different ways. The RIR and RIM models are
simple, powerful models to understand the basic features of AIE, and they have been very useful
to support the development of a large number of AIEgens with a great structural diversity. FGR
allows for the quantitative treatment of luminescence in the aggregate and crystalline phases,
while the study of PES and the RACI model have been instrumental to uncover the mechanism
behind the lack of fluorescence in solution in a broad variety of AIEgens. These approaches com-
plement each other and have provided us with a good understanding of the structural principles
that determine AIE.
In this tutorial review, we have centered on the global PES and the RACI model. The RACI
model provides specific information about the molecular motions responsible for the decay in
solution. It has recently received strong support from experimental time-­resolved spectroscopy
studies showing the short excited-­state lifetimes that are usually associated to decay at a
CI. Importantly, it has been useful to understand cases where small structural differences lead to
different photophysics and photochemistry and which could not be rationalized with the RIR and
444 14  A Global Potential Energy Surface Approach to the Photophysics of AIEgens: The Role of Conical Intersections

RIM models. A case-­by-­case analysis has allowed us to classify the molecules according to the
nuclear coordinate involved in accessing the CI and analysis of the orbitals involved in the excita-
tion has allowed us to rationalize the mechanisms behind the CI. We believe that this classification
will strengthen the synergy between AIE and other branches of photochemistry, where the role of
CIs has been well established.
As an outlook, we expect that the PES-­based approach will do important contributions to several
areas of AIE research in the future. First, there are still intriguing phenomena at the fundamental
level that can be tackled with the PES approach. Here, we would mention clusteroluminescence,
luminescence from nonconventional fluorophores, PIEE, AIE in TMs, or the role of dark states. In
all these cases, theory has already provided important insights, but a thorough mechanistic descrip-
tion is still missing. A further important goal is being able to predict trends in the photophysical
properties of large compound sets with the help of descriptors, and an interesting step has been
done along this direction with the work on DCS derivatives [43, 109]. Another challenging area is
understanding the role of intermolecular interactions on the photophysics of aggregate and crys-
talline phases and making theory and computations truly predictive for the design of new lumino-
genic phases. The models based on Kasha’s exciton theory are often insufficient to explain the
subtle effects found experimentally and a better understanding is needed to reach the predictive
goal. A final challenge is to model the photophysics of AIEgens in their application media, such as
biological probes embedded in their biological environment. All these developments will hopefully
increase the impact of theory on the development of AIE.

­References

1 He, Z., Ke, C., Tang, B.Z. (2018). Journey of aggregation-­induced emission research. ACS Omega 3:
3267–3277. https://doi.org/10.1021/acsomega.8b00062.
2 Mei, J., Leung, N.L.C., Kwok, R.T.K. et al. (2015). Aggregation-­induced emission: together we shine,
united we soar! Chem. Rev. 115: 11718–11940. https://doi.org/10.1021/acs.chemrev.5b00263.
3 Tang, B.Z., Zhao, Z., Zhang, H. et al. (2020). Aggregation-­induced emission: new vistas at aggregate
level. Angew. Chem. Int. Ed. https://doi.org/10.1002/anie.201916729.
4 Crespo-­Otero, R., Li, Q., Blancafort, L. (2019). Exploring potential energy surfaces for aggregation-­
induced emission—­from solution to crystal. Chem. Asian J. 14: 700–714. https://doi.org/10.1002/
asia.201801649.
5 Kokado, K., Sada, K. (2019). Consideration of molecular structure in the excited state to design new
luminogens with aggregation-­induced emission. Angew. Chem. Int. Ed. 58: 8632–8639. https://doi.
org/10.1002/anie.201814462.
6 Chen, J., Law, C.C.W., Lam, J.W.Y. et al. (2003). Synthesis, light emission, nanoaggregation, and
restricted intramolecular rotation of 1,1-­substituted 2,3,4,5-­tetraphenylsiloles. Chem. Mater. 15:
1535–1546. https://doi.org/10.1021/cm021715z.
7 Leung, N.L.C., Xie, N., Yuan, W. et al. (2014). Restriction of intramolecular motions: the general
mechanism behind aggregation-­induced emission. Chem. Eur. J. 20: 15349–15353. https://doi.
org/10.1002/chem.201403811.
8 Niu, Y., Li, W., Peng, Q. et al. (2018). MOlecular MAterials Property prediction package (MOMAP)
1.0: a software package for predicting the luminescent properties and mobility of organic functional
materials. Mol. Phys. 116: 1078–1090. https://doi.org/10.1080/00268976.2017.1402966.
9 Peng, Q., Shuai, Z. (2016). Theoretical insights into the mechanism of AIE. In: Aggregation-­Induced
Emission: Materials and Applications. Volume 1 (ed. M. Fujiki, B. Liu, B.Z. Tang), 35–59. ACS
Symposium Series, 1226: American Chemical Society, Washington, DC.
 ­Reference 445

10 Shuai, Z., Peng, Q. (2017). Organic light-­emitting diodes: theoretical understanding of highly
efficient materials and development of computational methodology. Natl. Sci. Rev. 4: 224–239.
https://doi.org/10.1093/nsr/nww024.
11 Peng, X.-­L., Ruiz-­Barragan, S., Li, Z.-­S. et al. (2016). Restricted access to a conical intersection to
explain aggregation induced emission in dimethyl tetraphenylsilole. J. Mater. Chem. C 4: 2802–
2810. https://doi.org/10.1039/c5tc03322e.
12 Blancafort, L., Ogliaro F., Olivucci M., Robb M., Beapark M., Sinicropi A. (2005). Computational
investigation of photochemical reaction mechanisms. In: Computational Methods in
Photochemistry (ed. A.G. Kutateladze). Molecular and Supramolecular Photochemistry, 13. Boca
Raton: CRC Press.
13 Klessinger, M., Michl, J. (1995). Excited States and Photochemistry of Organic Molecules. New York,
USA: VCH Publishers, Inc.
14 Lischka, H., Nachtigallová, D., Aquino, A.J.A. et al. (2018). Multireference approaches for excited
states of molecules. Chem. Rev. 118: 7293–7361. https://doi.org/10.1021/acs.chemrev.8b00244.
15 Matsika, S., Krylov, A.I. (2018). Introduction: theoretical modeling of excited state processes.
Chem. Rev. 118: 6925–6926. https://doi.org/10.1021/acs.chemrev.8b00436.
16 Robb, M.A. (2018). Theoretical Chemistry for Electronic Excited States. (ed. J. Hirst, J. Smith,
D. Wei), Theoretical and Computational Chemistry Series, Vol. 12. Croydon: Royal Society of
Chemistry.
17 Braslavsky, S.E. (2007). Glossary of terms used in photochemistry, 3rd edition (IUPAC
Recommendations 2006). Pure Appl. Chem. 79: 293–465. https://doi.org/https://doi.org/10.1351/
pac200779030293.
18 Kasha, M. (1950). Characterization of electronic transitions in complex molecules. Discuss. Farad.
Soc. 9: 14–19. https://doi.org/10.1039/DF9500900014.
19 Giussani, A., Segarra-­Martí, J., Roca-­Sanjuán, D. et al. (2015). Excitation of nucleobases from a
computational perspective I: reaction paths. In: Photoinduced Phenomena in Nucleic Acids I:
Nucleobases in the Gas Phase and in Solvents (ed. M. Barbatti, A.C. Borin, S. Ullrich), 57–97. Cham:
Springer International Publishing.
20 Domcke, W., Yarkony, D.R., Köppel, H., ed. (2004). Conical Intersections: Electronic Structure,
Dynamics & Spectroscopy. Singapore: World Scientific.
21 Domcke, W., Yarkony, D.R., Köppel, H., ed. (2011). Conical Intersections: Theory, Computation and
Experiment. Singapore: World Scientific.
22 Yarkony, D.R. (2001). Conical intersections: the new conventional wisdom. J. Phys. Chem. A 105:
6277–6293.
23 Bernardi, F., Olivucci, M., Robb, M.A. (1996). Potential energy surface crossings in organic
photochemistry. Chem. Soc. Rev. 25: 321–328. https://doi.org/10.1039/CS9962500321.
24 Polli, D., Altoe, P., Weingart, O. et al. (2010). Conical intersection dynamics of the primary
photoisomerization event in vision. Nature 467: 440–443. https://doi.org/10.1038/nature09346.
25 Crespo-­Hernández, C.E., Cohen, B., Hare, P.M. et al. (2004). Ultrafast excited-­state dynamics in
nucleic acids. Chem. Rev. 104: 1977–2019.
26 Improta, R., Santoro, F., Blancafort, L. (2016). Quantum mechanical studies on the photophysics
and the photochemistry of nucleic acids and nucleobases. Chem. Rev. 116: 3540–3593. https://doi.
org/10.1021/acs.chemrev.5b00444.
27 Viglianti, L., Xie, N., Sung, H.H.Y. et al. (2020). Unusual through-­space interactions between
oxygen atoms that mediate inverse morphochromism of an AIE luminogen. Angew. Chem. Int. Ed.
59: 8552–8559. https://doi.org/10.1002/anie.201908573.
28 Atchity, G.J., Xantheas, S.S., Ruedenberg, K. (1991). Potential-­energy surfaces near intersections.
J. Chem. Phys. 95: 1862–1876.
446 14  A Global Potential Energy Surface Approach to the Photophysics of AIEgens: The Role of Conical Intersections

29 Blancafort, L. (2014). Photochemistry and photophysics at extended seams of conical intersection.


ChemPhysChem 15: 3166–3181. https://doi.org/10.1002/cphc.201402359.
30 Yuan, W.Z., Shen, X.Y., Zhao, H. et al. (2010). Crystallization-­induced phosphorescence of pure
organic luminogens at room temperature. J. Phys. Chem. C 114: 6090–6099. https://doi.
org/10.1021/jp909388y.
31 Lower, S.K., El-­Sayed, M.A. (1966). Triplet state and molecular electronic processes in organic
molecules. Chem. Rev. 66: 199–241.
32 Dreuw, A., Head-­Gordon, M. (2005). Single-­reference ab initio methods for the calculation of
excited states of large molecules. Chem. Rev. 105: 4009–4037. https://doi.org/10.1021/cr0505627.
33 González, L., Escudero, D., Serrano-­Andrés, L. (2012). Progress and challenges in the calculation of
electronic excited states. ChemPhysChem 13: 28–51. https://doi.org/10.1002/cphc.201100200.
34 Matsika, S., Krause, P. (2011). Nonadiabatic events and conical intersections. Annu. Rev. Phy.
Chem. 62(1): 621–643. https://doi.org/10.1146/annurev-physchem-032210-103450.
35 Schreiber, M., Silva, M.R.J., Sauer, S.P.A. et al. (2008). Benchmarks for electronically excited states:
CASPT2, CC2, CCSD, and CC3. J. Chem. Phys. 128: 134110. https://doi.org/10.1063/1.2889385.
36 Serrano-­Andrés, L., Merchán, M. (2005). Quantum chemistry of the excited state: 2005 overview.
J. Mol. Struc.-­THEOCHEM 729: 99–108. https://doi.org/10.1016/j.theochem.2005.03.020.
37 Silva-­Junior, M.R., Schreiber, M., Sauer, S.P.A. et al. (2008). Benchmarks for electronically excited
states: time-­dependent density functional theory and density functional theory based
multireference configuration interaction. J. Chem. Phys. 129: 104103. https://doi.
org/10.1063/1.2973541.
38 Andersson, K., Malmqvist, P.A., Roos, B.O. (1992). 2nd-­order perturbation theory with a complete
active space self-­consistent field reference function. J. Chem. Phys. 96: 1218–1226. https://doi.
org/10.1063/1.462209.
39 Pulay, P. (2011). A perspective on the CASPT2 method. Int. J. Quant. Chem. 111: 3273–3279.
https://doi.org/10.1002/qua.23052.
40 Roca-­Sanjuan, D., Aquilante, F., Lindh, R. (2012). Multiconfiguration second-­order perturbation
theory approach to strong electron correlation in chemistry and photochemistry. WIREs Comput.
Mol. Sci. 2: 585–603. https://doi.org/10.1002/wcms.97.
41 Ding, W.-­L., Peng, X.-­L., Cui, G.-­L. et al. (2019). Potential-­energy surface and dynamics simulation
of THBDBA: an annulated tetraphenylethene derivative combining aggregation-­induced emission
and switch behavior. ChemPhotoChem 3: 814–824. https://doi.org/10.1002/cptc.201900112.
42 Gao, Y.-­J., Chang, X.-­P., Liu, X.-­Y. et al. (2017). Excited-­state decay paths in tetraphenylethene
derivatives. J. Phys. Chem. A 121: 2572–2579. https://doi.org/10.1021/acs.jpca.7b00197.
43 Izquierdo, M.A., Shi, J., Oh, S. et al. (2019). Excited-­state non-­radiative decay in stilbenoid
compounds: an ab initio quantum-­chemistry study on size and substituent effects. Phys. Chem.
Chem. Phys. 21: 22429–22439. https://doi.org/10.1039/C9CP03308D.
44 Li, Q., Blancafort, L. (2013). A conical intersection model to explain aggregation induced
emission in diphenyl dibenzofulvene. Chem. Commun. 49: 5966–5968. https://doi.org/10.1039/
c3cc41730a.
45 Londesborough, M.G.S., Dolanský, J., Cerdán, L. et al. (2017). Thermochromic fluorescence from
B18H20(NC5H5)2: an inorganic–organic composite luminescent compound with an unusual
molecular geometry. Adv. Opt. Mater. 5: 1600694. https://doi.org/10.1002/adom.201600694.
46 Shigemitsu, Y., Mutai, T., Houjou, H. et al. (2012). Excited-­state intramolecular proton transfer
(ESIPT) emission of hydroxyphenylimidazopyridine: computational study on enhanced and
polymorph-­dependent luminescence in the solid state. J. Phys. Chem. A 116: 12041–12048. https://
doi.org/10.1021/jp308473j.
 ­Reference 447

47 Stojanović, L., Crespo-­Otero, R. (2019). Understanding aggregation induced emission in a


propeller-­shaped blue emitter. ChemPhotoChem 3: 907–915. https://doi.org/10.1002/
cptc.201900075.
48 Wang, H., Gong, Q., Wang, G. et al. (2019). Deciphering the mechanism of aggregation-­induced
emission of a quinazolinone derivative displaying excited-­state intramolecular proton-­transfer
properties: a QM, QM/MM, and MD study. J. Chem. Theory Comput. 15: 5440–5447. https://doi.
org/10.1021/acs.jctc.9b00421.
49 Zhou, P., Li, P., Zhao, Y. et al. (2019). Restriction of flip-­flop motion as a mechanism for
aggregation-­induced emission. J. Phys. Chem. Lett. 10: 6929–6935. https://doi.org/10.1021/acs.
jpclett.9b02922.
50 Martínez-­Fernández, L., Pepino, A.J., Segarra-­Martí, J. et al. (2017). Photophysics of deoxycytidine
and 5-­methyldeoxycytidine in solution: a comprehensive picture by quantum mechanical
calculations and femtosecond fluorescence spectroscopy. J. Am. Chem. Soc. 139: 7780–7791.
https://doi.org/10.1021/jacs.7b01145.
51 Saurí, V., Serrano-­Andrés, L., Shahi, A.R.M. et al. (2011). Multiconfigurational second-­order
perturbation theory restricted active space (RASPT2) method for electronic excited states: a
benchmark study. J. Chem. Theory Comput. 7: 153–168. https://doi.org/10.1021/ct100478d.
52 Chan, G.K.-­L., Sharma, S. (2011). The density matrix renormalization group in quantum chemistry.
Annu. Rev. Phys. Chem. 62(1): 465–481. https://doi.org/10.1146/annurev-physchem-032210-103338.
53 Le Bras, L., Adamo, C., Perrier, A. (2019). In silico investigation of the aggregation-­caused
quenching: the “tolane-­based molecule” case. ChemPhotoChem 3: 794–803. https://doi.
org/10.1002/cptc.201900117.
54 Le Bras, L., Chaitou, K., Aloïse, S. et al. (2019). Aggregation-­caused quenching versus
crystallization induced emission in thiazolo[5,4-­b]thieno[3,2-­e]pyridine (TTP) derivatives:
theoretical insights. Phys. Chem. Chem. Phys. 21: 46–56. https://doi.org/10.1039/C8CP04730H.
55 Zhang, T., Shi, W., Wang, D. et al. (2019). Pressure-­induced emission enhancement in
hexaphenylsilole: a computational study. J. Mater. Chem. C 7: 1388–1398. https://doi.org/10.1039/
C8TC05162C.
56 Shi, P., Deng, D., He, C. et al. (2020). Mechanochromic luminescent materials with aggregation-­
induced emission: mechanism study and application for pressure measuring and mechanical
printing. Dyes Pigments 173: 107884. https://doi.org/10.1016/j.dyepig.2019.107884.
57 Zhou, Z., Xie, S., Chen, X. et al. (2019). Spiro-­functionalized diphenylethenes: suppression of a
reversible photocyclization contributes to the aggregation-­induced emission effect. J. Am. Chem.
Soc. 141: 9803–9807. https://doi.org/10.1021/jacs.9b04426.
58 Yokoyama, S., Nishiwaki, N. (2019). Fluorescence behavior of bis(cyanostyryl)pyrrole derivatives
depending on the substituent position of cyano groups in solution and in solid state. J. Org. Chem.
84: 1192–1200. https://doi.org/10.1021/acs.joc.8b02517.
59 Rivera, M., Dommett, M., Crespo-­Otero, R. (2019). ONIOM(QM:QM’) Electrostatic embedding
schemes for photochemistry in molecular crystals. J. Chem. Theory Comput. 15: 2504–2516. https://
doi.org/10.1021/acs.jctc.8b01180.
60 Casida, M.E. (2009). Time-­dependent density-­functional theory for molecules and molecular
solids. J. Mol. Struc.-­THEOCHEM 914: 3–18. https://doi.org/10.1016/j.theochem.2009.08.018.
61 Dreuw, A., Head-­Gordon, M. (2004). Failure of time-­dependent density functional theory for
long-­range charge-­transfer excited states: the zincbacteriochlorin-­bacteriochlorin and
bacteriochlorophyll-­spheroidene complexes. J. Am. Chem. Soc. 126: 4007–4016. https://doi.
org/10.1021/ja039556n.
448 14  A Global Potential Energy Surface Approach to the Photophysics of AIEgens: The Role of Conical Intersections

62 Mardirossian, N., Head-­Gordon, M. (2017). Thirty years of density functional theory in


computational chemistry: an overview and extensive assessment of 200 density functionals. Mol.
Phys. 115: 2315–2372. https://doi.org/10.1080/00268976.2017.1333644.
63 Tsuneda, T., Hirao, K. (2014). Long-­range correction for density functional theory. WIREs Comput.
Mol. Sci. 4: 375–390. https://doi.org/10.1002/wcms.1178.
64 Ghosh, S., Verma, P., Cramer, C.J. et al. (2018). Combining wave function methods with density
functional theory for excited states. Chem. Rev. 118: 7249–7292. https://doi.org/10.1021/acs.
chemrev.8b00193.
65 Stein, T., Kronik, L., Baer, R. (2009). Reliable prediction of charge transfer excitations in molecular
complexes using time-­dependent density functional theory. J. Am. Chem. Soc. 131: 2818–2820.
https://doi.org/10.1021/ja8087482.
66 Zheng, X., Peng, Q., Zhu, L. et al. (2016). Unraveling the aggregation effect on amorphous phase
AIE luminogens: a computational study. Nanoscale 8: 15173–15180. https://doi.org/10.1039/
C6NR03599J.
67 Qian, H., Cousins, M.E., Horak, E.H. et al. (2017). Suppression of Kasha’s rule as a mechanism for
fluorescent molecular rotors and aggregation-­induced emission. Nat. Chem. 9: 83–87. https://doi.
org/10.1038/nchem.2612.
68 Gozem, S., Melaccio, F., Valentini, A. et al. (2014). Shape of multireference, equation-­of-­motion
coupled-­cluster, and density functional theory potential energy surfaces at a conical intersection.
J. Chem. Theory Comput. 10: 3074–3084. https://doi.org/10.1021/ct500154k.
69 Levine, B.G., Ko, C., Quenneville, J. et al. (2006). Conical intersections and double excitations in
time-­dependent density functional theory. Mol. Phys. 104: 1039–1051. https://doi.
org/10.1080/00268970500417762.
70 Dommett, M., Crespo-­Otero, R. (2017). Excited state proton transfer in 2′-­hydroxychalcone
derivatives. Phys. Chem. Chem. Phys. 19: 2409–2416. https://doi.org/10.1039/C6CP07541J.
71 Dommett, M., Rivera, M., Crespo-­Otero, R. (2017). How inter-­and intramolecular processes dictate
aggregation-­induced emission in crystals undergoing excited-­state proton transfer. J. Phys. Chem.
Lett. 8: 6148–6153. https://doi.org/10.1021/acs.jpclett.7b02893.
72 Dommett, M., Rivera, M., Smith, M.T.H. et al. (2020). Molecular and crystalline requirements for
solid state fluorescence exploiting excited state intramolecular proton transfer. J. Mater. Chem. C 8:
2558–2568. https://doi.org/10.1039/C9TC05717J.
73 Shao, Y., Head-­Gordon, M., Krylov, A.I. (2003). The spin–flip approach within time-­dependent
density functional theory: theory and applications to diradicals. J. Chem. Phys. 118: 4807–4818.
https://doi.org/10.1063/1.1545679.
74 Suzuki, S., Maeda, S., Morokuma, K. (2015). Exploration of quenching pathways of
multiluminescent acenes using the GRRM method with the SF-­TDDFT method. J. Phys. Chem. A
119: 11479–11487. https://doi.org/10.1021/acs.jpca.5b07682.
75 Minezawa, N., Gordon, M.S. (2009). Optimizing conical intersections by spin−flip density
functional theory: application to ethylene. J. Phys. Chem. A 113: 12749–12753. https://doi.
org/10.1021/jp908032x.
76 Dreuw, A., Wormit, M. (2015). The algebraic diagrammatic construction scheme for the
polarization propagator for the calculation of excited states. WIREs Comput. Mol. Sci. 5: 82–95.
https://doi.org/10.1002/wcms.1206.
77 Christiansen, O., Koch, H., Jorgensen, P. (1995). The 2nd-­order approximate coupled-­cluster
singles and doubles model CC2. Chem. Phys. Lett. 243: 409–418. https://doi.
org/10.1016/0009-­2614(95)00841-­q.
 ­Reference 449

78 Gozem, S., Krylov, A.I., Olivucci, M. (2013). Conical intersection and potential energy surface
features of a model retinal chromophore: comparison of EOM-­CC and multireference methods.
J. Chem. Theory Comput. 9: 284–292. https://doi.org/10.1021/ct300759z.
79 Kjønstad, E.F., Myhre, R.H., Martínez, T.J. et al. (2017). Crossing conditions in coupled cluster
theory. J. Chem. Phys. 147: 164105. https://doi.org/10.1063/1.4998724.
80 Lefrançois, D., Tuna, D., Martínez, T.J. et al. (2017). The spin-­flip variant of the algebraic-­
diagrammatic construction yields the correct topology of S1/S0 conical intersections. J. Chem.
Theory Comput. 13: 4436–4441. https://doi.org/10.1021/acs.jctc.7b00634.
81 Crespo-­Otero, R., Barbatti, M. (2018). Recent advances and perspectives on nonadiabatic mixed
quantum–classical dynamics. Chem. Rev. 118: 7026–7068. https://doi.org/10.1021/acs.chemrev.7b00577.
82 Prlj, A., Došlić, N., Corminboeuf, C. (2016). How does tetraphenylethylene relax from its excited
states? Phys. Chem. Chem. Phys. 18: 11606–11609. https://doi.org/10.1039/C5CP04546K.
83 Humeniuk, A., Mitrić, R. (2015). Long-­range correction for tight-­binding TD-­DFT. J. Chem. Phys.
143: 134120. https://doi.org/10.1063/1.4931179.
84 Seifert, G. (2007). Tight-­binding density functional theory: an approximate Kohn−Sham DFT
scheme. J. Phys. Chem. A 111: 5609–5613. https://doi.org/10.1021/jp069056r.
85 Gao, X., Peng, Q., Niu, Y. et al. (2012). Theoretical insight into the aggregation induced emission
phenomena of diphenyldibenzofulvene: a nonadiabatic molecular dynamics study. Phys. Chem.
Chem. Phys. 14: 14207–14216. https://doi.org/10.1039/C2CP40347A.
86 Tran, T., Prlj, A., Lin, K.-­H. et al. (2019). Mechanisms of fluorescence quenching in prototypical
aggregation-­induced emission systems: excited state dynamics with TD-­DFTB. Phys. Chem. Chem.
Phys. 21: 9026–9035. https://doi.org/10.1039/C9CP00691E.
87 Weber, W., Thiel, W. (2000). Orthogonalization corrections for semiempirical methods. Theor.
Chem. Acc. 103: 495–506. https://doi.org/10.1007/s002149900083.
88 Tully, J.C. (1990). Molecular dynamics with electronic transitions. J. Chem. Phys. 93: 1061–1071.
https://doi.org/10.1063/1.459170.
89 Yamamoto, N. (2018). Mechanisms of aggregation-­induced emission and photo/thermal E/Z
isomerization of a cyanostilbene derivative: theoretical insights. J. Phys. Chem. C 122: 12434–12440.
https://doi.org/10.1021/acs.jpcc.8b02147.
90 Severo Pereira Gomes, A., Jacob, C.R. (2012). Quantum-­chemical embedding methods for treating
local electronic excitations in complex chemical systems. Annu. Rep. Sect. "C" (Phys. Chem.) 108:
222–277. https://doi.org/10.1039/C2PC90007F.
91 Li, W., Zhu, L., Shi, Q. et al. (2017). Excitonic coupling effect on the nonradiative decay rate in
molecular aggregates: formalism and application. Chem. Phys. Lett. 683: 507–514. https://doi.
org/10.1016/j.cplett.2017.03.077.
92 Mennucci, B. (2012). Polarizable continuum model. WIREs Comput. Mol. Sci. 2: 386–404. https://
doi.org/10.1002/wcms.1086.
93 Sasaki, S., Suzuki, S., Sameera, W.M.C. et al. (2016). Highly twisted N,N-­dialkylamines as a design
strategy to tune simple aromatic hydrocarbons as steric environment-­sensitive fluorophores. J. Am.
Chem. Soc. 138: 8194–8206. https://doi.org/10.1021/jacs.6b03749.
94 Shi, J., Aguilar Suarez, L.E., Yoon, S.-­J. et al. (2017). Solid state luminescence enhancement in
π-­conjugated materials: unraveling the mechanism beyond the framework of AIE/AIEE. J. Phys.
Chem. C 121: 23166–23183. https://doi.org/10.1021/acs.jpcc.7b08060.
95 Sun, G., Zhao, Y., Liang, W. (2015). Aggregation-­induced emission mechanism of dimethoxy-­
tetraphenylethylene in water solution: molecular dynamics and QM/MM investigations. J. Chem.
Theory Comput. 11: 2257–2267. https://doi.org/10.1021/ct5009312.
450 14  A Global Potential Energy Surface Approach to the Photophysics of AIEgens: The Role of Conical Intersections

96 Wang, B., Wang, X., Wang, W. et al. (2016). Exploring the mechanism of fluorescence quenching
and aggregation-­induced emission of a phenylethylene derivative by QM (CASSCF and TDDFT)
and ONIOM (QM:MM) calculations. J. Phys. Chem. C 120: 21850–21857. https://doi.org/10.1021/
acs.jpcc.6b07963.
97 Skelton, J.M., Lora da Silva, E., Crespo-­Otero, R. et al. (2015). Electronic excitations in molecular
solids: bridging theory and experiment. Faraday Discuss. 177: 181–202. https://doi.org/10.1039/
C4FD00168K.
98 Senn, H.M., Thiel, W. (2009). QM/MM methods for biomolecular systems. Angew. Chem. Int. Ed.
48: 1198–1229. https://doi.org/10.1002/anie.200802019.
99 Chung, L.W., Sameera, W.M.C., Ramozzi, R. et al. (2015). The ONIOM method and its
applications. Chem. Rev. 115: 5678–5796. https://doi.org/10.1021/cr5004419.
100 Fan, J., Cai, L., Lin, L. et al. (2017). Excited state dynamics for hybridized local and charge
transfer state fluorescent emitters with aggregation-­induced emission in the solid phase: a QM/
MM study. Phys. Chem. Chem. Phys. 19: 29872–29879. https://doi.org/10.1039/C7CP05009G.
101 Reining, L. (2018). The GW approximation: content, successes and limitations. WIREs Comput.
Mol. Sci. 8: e1344. https://doi.org/10.1002/wcms.1344.
102 Weber, J., Schmedt auf der Günne, J. (2010). Calculation of NMR parameters in ionic solids by an
improved self-­consistent embedded cluster method. Phys. Chem. Chem. Phys. 12: 583–603.
https://doi.org/10.1039/B909870D.
103 Klintenberg, M., Derenzo, S.E., Weber, M.J. (2000). Accurate crystal fields for embedded cluster
calculations. Comput. Phys. Commun. 131: 120–128. https://doi.org/10.1016/S0010-­4655(00)
00071-­0.
104 Wilbraham, L., Adamo, C., Labat, F. et al. (2016). Electrostatic embedding to model the impact of
environment on photophysical properties of molecular crystals: a self-­consistent charge
adjustment procedure. J. Chem. Theory Comput. 12: 3316–3324. https://doi.org/10.1021/acs.
jctc.6b00263.
105 Presti, D., Wilbraham, L., Targa, C. et al. (2017). Understanding aggregation-­induced emission in
molecular crystals: insights from theory. J. Phys. Chem. C 121: 5747–5752. https://doi.org/10.1021/
acs.jpcc.7b00488.
106 Wilbraham, L., Louis, M., Alberga, D. et al. (2018). Revealing the origins of mechanically induced
fluorescence changes in organic molecular crystals. Adv. Mater. 30: 1800817. https://doi.
org/10.1002/adma.201800817.
107 Bonačić-­Koutecký, V., Koutecký, J., Michl, J. (1987). Neutral and charged biradicals, zwitterions,
funnels in S1, and proton translocation -­their role in photochemistry, photophysics, and vision.
Angew. Chem. Int. Ed. 26: 170–189.
108 Bonačić-­Koutecký, V., Schoffel, K., Michl, J. (1987). Critically heterosymmetric biradicaloid
geometries of protonated schiff-­bases -­possible consequences for photochemistry and
photobiology. Theor. Chim. Acta 72: 459–474.
109 Shi, J., Izquierdo, M.A., Oh, S. et al. (2019). Inverted energy gap law for the nonradiative decay in
fluorescent floppy molecules: larger fluorescence quantum yields for smaller energy gaps. Org.
Chem. Front. 6: 1948–1954. https://doi.org/10.1039/C9QO00259F.
110 Tong, H., Dong, Y., Hong, Y. et al. (2007). Aggregation-­induced emission: effects of molecular
structure, solid-­state conformation, and morphological packing arrangement on light-­emitting
behaviors of diphenyldibenzofulvene derivatives. J. Phys. Chem. C 111: 2287–2294. https://doi.
org/10.1021/jp0630828.
111 Cariati, E., Lanzeni, V., Tordin, E. et al. (2011). Efficient crystallization induced emissive
materials based on a simple push–pull molecular structure. Phys. Chem. Chem. Phys. 13: 18005–
18014. https://doi.org/10.1039/C1CP22267H.
 ­Reference 451

112 Coluccini, C., Anusha, P.T., Chen, H.-­Y.T. et al. (2019). Tuning of the electro-­optical properties of
tetraphenylcyclopentadienone via substitution of oxygen with sterically-­hindered electron
withdrawing groups. Sci. Rep. 9: 12762. https://doi.org/10.1038/s41598-­019-­49303-­w.
113 Yokoyama, S., Ito, A., Asahara, H. et al. (2019). Anion-­capture-­induced fluorescence
enhancement of bis(cyanostyryl)pyrrole based on restricted access to a conical intersection. Bull.
Chem. Soc. Jpn. 92: 1807–1815. https://doi.org/10.1246/bcsj.20190196.
114 Chung, J.W., Yoon, S.-­J., An, B.-­K. et al. (2013). High-­contrast on/off fluorescence switching via
reversible E–Z isomerization of diphenylstilbene containing the α-­cyanostilbenic moiety. J. Phys.
Chem. C 117: 11285–11291. https://doi.org/10.1021/jp401440s.
115 Chung, J.W., Yoon, S.-­J., An, B.-­K. et al. (2017). Correction to “high-­contrast on/off fluorescence
switching via reversible E–Z isomerization of diphenylstilbene containing the α-­cyanostilbenic
moiety”. J. Phys. Chem. C 121: 26139–26139. https://doi.org/10.1021/acs.jpcc.7b10301.
116 Iwai, R., Suzuki, S., Sasaki, S. et al. (2020). Bridged stilbenes: AIEgens designed via a simple
strategy to control the non-­radiative decay pathway. Angew. Chem. Int. Ed. https://doi.
org/10.1002/anie.202000943.
117 Zhang, H., Liu, J., Du, L. et al. (2019). Drawing a clear mechanistic picture for the aggregation-­
induced emission process. Mater. Chem. Front. 3: 1143–1150. https://doi.org/10.1039/
C9QM00156E.
118 Olsen, R.J., Buckles, R.E. (1979). Substituent effects on the efficiency and regioselectivity of
tetraarylethylene photocyclization. J. Photochem. 10: 215–220. https://doi.
org/10.1016/0047-­2670(79)80009-­X.
119 Sension, R.J., Repinec, S.T., Szarka, A.Z. et al. (1993). Femtosecond laser studies of the cis-­
stilbene photoisomerization reactions. J. Chem. Phys. 98: 6291–6315. https://doi.
org/10.1063/1.464824.
120 Bearpark, M.J., Bernardi, F., Clifford, S. et al. (1997). Cooperating rings in cis-­stilbene lead to an
S0/S1 conical intersection. J. Phys. Chem. A 101: 3841–3847. https://doi.org/10.1021/jp961509c.
121 Cai, Y., Du, L., Samedov, K. et al. (2018). Deciphering the working mechanism of aggregation-­
induced emission of tetraphenylethylene derivatives by ultrafast spectroscopy. Chem. Sci. 9:
4662–4670. https://doi.org/10.1039/C8SC01170B.
122 Xu, B., He, J., Mu, Y. et al. (2015). Very bright mechanoluminescence and remarkable
mechanochromism using a tetraphenylethene derivative with aggregation-­induced emission.
Chem. Sci. 6: 3236–3241. https://doi.org/10.1039/C5SC00466G.
123 Shultz, D.A., Fox, M.A. (1989). Effect of phenyl ring torsional rigidity on the photophysical
behavior of tetraphenylethylenes. J. Am. Chem. Soc. 111: 6311–6320. https://doi.org/10.1021/
ja00198a049.
124 Zhang, G.-­F., Chen, Z.-­Q., Aldred, M.P. et al. (2014). Direct validation of the restriction of
intramolecular rotation hypothesis via the synthesis of novel ortho-­methyl substituted
tetraphenylethenes and their application in cell imaging. Chem. Commun. 50: 12058–12060.
https://doi.org/10.1039/C4CC04241G.
125 Luo, J., Song, K., Gu, F.l. et al. (2011). Switching of non-­helical overcrowded
tetrabenzoheptafulvalene derivatives. Chem. Sci. 2: 2029–2034. https://doi.org/10.1039/
C1SC00340B.
126 Kokado, K., Machida, T., Iwasa, T. et al. (2018). Twist of C═C bond plays a crucial role in the
quenching of AIE-­active tetraphenylethene derivatives in solution. J. Phys. Chem. C 122: 245–251.
https://doi.org/10.1021/acs.jpcc.7b11248.
127 Stasyuk, A.J., Bultinck, P., Gryko, D.T. et al. (2016). The effect of hydrogen bond strength on
emission properties in 2-­(2′-­hydroxyphenyl)imidazo[1,2-­a]pyridines. J. Photochem. Photobiol. A
314: 198–213. https://doi.org/10.1016/j.jphotochem.2015.08.013.
452 14  A Global Potential Energy Surface Approach to the Photophysics of AIEgens: The Role of Conical Intersections

128 Sobolewski, A.L., Domcke, W. (2006). Photophysics of intramolecularly hydrogen-­bonded


aromatic systems: ab initio exploration of the excited-­state deactivation mechanisms of salicylic
acid. Phys. Chem. Chem. Phys. 8: 3410–3417. https://doi.org/10.1039/B604610J.
129 Sobolewski, A.L., Domcke, W., Hättig, C. (2006). Photophysics of organic photostabilizers. Ab
initio study of the excited-­state deactivation mechanisms of 2-­(2′-­hydroxyphenyl)benzotriazole.
J. Phys. Chem. A 110: 6301–6306. https://doi.org/10.1021/jp0574798.
130 Anthony, S.P. (2012). Polymorph-­dependent solid-­state fluorescence and selective metal-­ion-­
sensor properties of 2-­(2-­hydroxyphenyl)-­4(3H)-­quinazolinone. Chem. Asian J. 7: 374–379.
https://doi.org/10.1002/asia.201100832.
131 Gao, M., Li, S., Lin, Y. et al. (2016). Fluorescent light-­up detection of amine vapors based on
aggregation-­induced emission. ACS Sensors 1: 179–184. https://doi.org/10.1021/
acssensors.5b00182.
132 Cheng, X., Wang, K., Huang, S. et al. (2015). Organic crystals with near-­infrared amplified
spontaneous emissions based on 2′-­hydroxychalcone derivatives: subtle structure modification
but great property change. Angew. Chem. Int. Ed. 54: 8369–8373. https://doi.org/10.1002/
anie.201503914.
133 Cheng, X., Zhang, Y., Han, S. et al. (2016). Multicolor amplified spontaneous emissions based on
organic polymorphs that undergo excited-­state intramolecular proton transfer. Chem. Eur. J. 22:
4899–4903. https://doi.org/10.1002/chem.201600355.
134 Tang, B., Liu, H., Li, F. et al. (2016). Single-­benzene solid emitters with lasing properties based on
aggregation-­induced emissions. Chem. Commun. 52: 6577–6580. https://doi.org/10.1039/
C6CC02616H.
135 Tang, B.Z., Zhan, X., Yu, G. et al. (2001). Efficient blue emission from siloles. J. Mater. Chem. 11:
2974–2978. https://doi.org/10.1039/B102221K.
136 Gao, X.-­C., Cao, H., Huang, L. et al. (2003). Comparison of the electroluminescence and its
related properties of two cyclopentadiene derivatives. Appl. Surf. Sci. 210: 183–189. https://doi.
org/10.1016/S0169-­4332(02)01236-­9.
137 Poronik, Y.M., Gryko, D.T. (2014). Pentacyclic coumarin-­based blue emitters – the case of
bifunctional nucleophilic behavior of amidines. Chem. Commun. 50: 5688–5690. https://doi.
org/10.1039/C4CC01106F.
138 Ventura, B., Poronik, Y.M., Deperasińska, I. et al. (2016). How a small structural difference can
turn optical properties of π-­extended coumarins upside down: the role of non-­innocent saturated
rings. Chem. Eur. J. 22: 15380–15388. https://doi.org/10.1002/chem.201603038.
139 Sasaki, S., Suzuki, S., Igawa, K. et al. (2017). The K-­region in pyrenes as a key position to activate
aggregation-­induced emission: effects of introducing highly twisted n,n-­dimethylamines. J. Org.
Chem. 82: 6865–6873. https://doi.org/10.1021/acs.joc.7b00996.
140 Yuan, C., Saito, S., Camacho, C. et al. (2013). A π-­conjugated system with flexibility and rigidity
that shows environment-­dependent RGB luminescence. J. Am. Chem. Soc. 135: 8842–8845.
https://doi.org/10.1021/ja404198h.
141 Yuan, C., Saito, S., Camacho, C. et al. (2014). Hybridization of a flexible cyclooctatetraene core
and rigid aceneimide wings for multiluminescent flapping π systems. Chem. Eur. J. 20: 2193–
2200. https://doi.org/10.1002/chem.201303955.
142 Londesborough, M.G.S., Dolanský, J., Jelínek, T. et al. (2018). Substitution of the laser borane
anti-­B18H22 with pyridine: a structural and photophysical study of some unusually structured
macropolyhedral boron hydrides. Dalton T. 47: 1709–1725. https://doi.org/10.1039/C7DT03823B.
143 An, Z., Zheng, C., Tao, Y. et al. (2015). Stabilizing triplet excited states for ultralong organic
phosphorescence. Nat. Mater. 14: 685–690. https://doi.org/10.1038/nmat4259.
 ­Reference 453

144 Hirata, S. (2017). Recent advances in materials with room-­temperature phosphorescence:


photophysics for triplet exciton stabilization. Adv. Opt. Mater. 5: 1700116. https://doi.org/10.1002/
adom.201700116.
145 Katoh, R., Kotani, M., Hirata, Y. et al. (1997). Triplet exciton formation in a benzophenone single
crystal studied by picosecond time-­resolved absorption spectroscopy. Chem. Phys. Lett. 264:
631–635. https://doi.org/10.1016/S0009-­2614(96)01389-­9.
146 Favero, L., Granucci, G., Persico, M. (2016). Surface hopping investigation of benzophenone excited
state dynamics. Phys. Chem. Chem. Phys. 18: 10499–10506. https://doi.org/10.1039/C6CP00328A.
147 Marazzi, M., Mai, S., Roca-­Sanjuán, D. et al. (2016). Benzophenone ultrafast triplet population:
revisiting the kinetic model by surface-­hopping dynamics. J. Phys. Chem. Lett. 7: 622–626. https://
doi.org/10.1021/acs.jpclett.5b02792.
148 Sergentu, D.-­C., Maurice, R., Havenith, R.W.A. et al. (2014). Computational determination of the
dominant triplet population mechanism in photoexcited benzophenone. Phys. Chem. Chem. Phys.
16: 25393–25403. https://doi.org/10.1039/C4CP03277B.
149 Ma, H., Shi, W., Ren, J. et al. (2016). Electrostatic interaction-­induced room-­temperature
phosphorescence in pure organic molecules from QM/MM calculations. J. Phys. Chem. Lett. 7:
2893–2898. https://doi.org/10.1021/acs.jpclett.6b01156.
150 Kasha, M. (1963). Energy transfer mechanisms and the molecular exciton model for molecular
aggregates. Radiat. Res. 20: 55–70. https://doi.org/10.2307/3571331.
151 Hestand, N.J., Spano, F.C. (2017). Molecular aggregate photophysics beyond the Kasha model:
novel design principles for organic materials. Acc. Chem. Res. 50: 341–350. https://doi.
org/10.1021/acs.accounts.6b00576.
152 Domínguez, R., Moral, M., Fernández-­Liencres, M.P. et al. (2020). Understanding the driving
mechanisms of enhanced luminescence emission of oligo(styryl)benzenes and tri(styryl)-­s-­
triazine. Chem. Eur. J. 26: 3373–3384. https://doi.org/10.1002/chem.201905336.
153 Mutai, T., Shono, H., Shigemitsu, Y. et al. (2014). Three-­color polymorph-­dependent luminescence:
crystallographic analysis and theoretical study on excited-­state intramolecular proton transfer
(ESIPT) luminescence of cyano-­substituted imidazo[1,2-­a]pyridine. CrystEngComm 16: 3890–
3895. https://doi.org/10.1039/C3CE42627K.
154 Sturala, J., Etherington, M.K., Bismillah, A.N. et al. (2017). Excited-­state aromatic interactions in
the aggregation-­induced emission of molecular rotors. J. Am. Chem. Soc. 139: 17882–17889.
https://doi.org/10.1021/jacs.7b08570.
155 Zhang, H., Du, L., Wang, L. et al. (2019). Visualization and manipulation of molecular motion in
the solid state through photoinduced clusteroluminescence. J. Phys. Chem. Lett. 10: 7077–7085.
https://doi.org/10.1021/acs.jpclett.9b02752.
156 Zhang, H., Zhao, Z., McGonigal, P.R. et al. (2020). Clusterization-­triggered emission: uncommon
luminescence from common materials. Mater. Today 32: 275–292. https://doi.org/10.1016/j.
mattod.2019.08.010.
157 Bader, R.F.W. (1991). A quantum theory of molecular structure and its applications. Chem. Rev.
91: 893–928. https://doi.org/10.1021/cr00005a013.
158 Tu, Y., Liu, J., Zhang, H. et al. (2019). Restriction of access to the dark state: a new mechanistic
model for heteroatom-­containing AIE systems. Angew. Chem. Int. Ed. 58: 14911–14914. https://
doi.org/10.1002/anie.201907522.
159 Escudero, D. (2016). Revising intramolecular photoinduced electron transfer (PET) from first-­
principles. Acc. Chem. Res. 49: 1816–1824. https://doi.org/10.1021/acs.accounts.6b00299.
160 Fang, L., Trigiante, G., Crespo-­Otero, R. et al. (2019). Endoplasmic reticulum targeting fluorescent
probes to image mobile Zn2+. Chem. Sci. 10: 10881–10887. https://doi.org/10.1039/C9SC04300D.
454 14  A Global Potential Energy Surface Approach to the Photophysics of AIEgens: The Role of Conical Intersections

161 Fan, X., Sun, J., Wang, F. et al. (2008). Photoluminescence and electroluminescence of
hexaphenylsilole are enhanced by pressurization in the solid state. Chem. Commun. 2008,
2989–2991. https://doi.org/10.1039/b803539c.
162 Gu, Y., Li, N., Shao, G. et al. (2020). Mechanism of different piezoresponsive luminescence of
2,3,4,5-­tetraphenylthiophene and 2,3,4,5-­tetraphenylfuran: a strategy for designing pressure-­
induced emission enhancement materials. J. Phys. Chem. Lett. 11: 678–682. https://doi.
org/10.1021/acs.jpclett.9b03592.
163 Gu, Y., Liu, H., Qiu, R. et al. (2019). Pressure-­induced emission enhancement and multicolor
emission for 1,2,3,4-­tetraphenyl-­1,3-­cyclopentadiene: controlled structure evolution. J. Phys.
Chem. Lett. 10: 5557–5562. https://doi.org/10.1021/acs.jpclett.9b02206.
164 Gu, Y., Wang, K., Dai, Y. et al. (2017). Pressure-­induced emission enhancement of carbazole: the
restriction of intramolecular vibration. J. Phys. Chem. Lett. 8: 4191–4196. https://doi.org/10.1021/
acs.jpclett.7b01796.
165 Alam, P., Climent, C., Alemany, P. et al. (2019). “Aggregation-­induced emission” of transition
metal compounds: design, mechanistic insights, and applications. J. Photochem. Photobiol. C 41:
100317. https://doi.org/10.1016/j.jphotochemrev.2019.100317.
166 Alam, P., Climent, C., Kaur, G. et al. (2016). Exploring the origin of “aggregation induced
emission” activity and “crystallization induced emission” in organometallic iridium(III) cationic
complexes: influence of counterions. Cryst. Growth Des. 16: 5738–5752. https://doi.org/10.1021/
acs.cgd.6b00810.
167 Liu, S., Sun, H., Ma, Y. et al. (2012). Rational design of metallophosphors with tunable
aggregation-­induced phosphorescent emission and their promising applications in time-­resolved
luminescence assay and targeted luminescence imaging of cancer cells. J. Mater. Chem. 22:
22167–22173. https://doi.org/10.1039/C2JM34512A.
168 Wu, Y., Sun, H.-­Z., Cao, H.-­T. et al. (2014). Stepwise modulation of the electron-­donating strength
of ancillary ligands: understanding the AIE mechanism of cationic iridium(III) complexes. Chem.
Commun. 50: 10986–10989. https://doi.org/10.1039/C4CC03423F.
169 Alam, P., Das, P., Climent, C. et al. (2014). Facile tuning of the aggregation-­induced emission
wavelength in a common framework of a cyclometalated iridium(III) complex: micellar
encapsulated probe in cellular imaging. J. Mater. Chem. C 2: 5615–5628. https://doi.org/10.1039/
C4TC00466C.
170 Alam, P., Dash, S., Climent, C. et al. (2017). ‘Aggregation induced emission’ active iridium(III)
complexes with applications in mitochondrial staining. RSC Adv. 7: 5642–5648. https://doi.
org/10.1039/C6RA24792J.
455

15

Multicomponent Reactions as Synthetic Design Tools


of AIE and Emission Solvatochromic Quinoxalines
Lukas Biesen and Thomas J. J. Müller
Institut für Organische Chemie und Makromolekulare Chemie, Heinrich-­Heine-­Universität Düsseldorf, Düsseldorf, Germany

15.1 ­Introduction

The desire in academia and industry for novel synthetic approaches and concepts to highly diversi-
fied functional π-­electron systems is steadily increasing [1, 2]. These π-­electron systems often com-
prise functional organic dyes and are used in a broad variety of applications for instance in
molecular electronics [3, 4], in photoelectronic applications [5–8] such as organic field-­effect tran-
sistors (OFETs)  [9, 10], dye-­sensitized solar cells (DSSCs)  [11–13] and as organic light-­emitting
diodes (OLEDs) [14–17] as well as in biological imaging, such as desoxyribonucleic acid (DNA)
and protein labeling and in bio and environmental analytics [18–21].
The ever-­increasing demand for these functional dyes requires easy access and facile, secure syn-
thetic routes with high yields to the target molecules in a resource-­saving and sustainable manner [22].
Finding syntheses meeting these requirements and generating chromophores with a distinctive pho-
tophysical profile rank among the most crucial tasks for the synthetic organic and material chemists.
Concepts that fulfill these expectations are multicomponent [23–33] and domino reactions [34–37].
Multicomponent reactions (MCRs) represent a reactivity-­based concept [26] and are defined as
a one-­pot transformation of more than two starting materials to form a product that is constituted
of most of the deployed atoms [26, 31, 38–39].
There are three different scenarios for MCRs to ensue. In a MCR following the first class of MCRs,
the domino pathway, all reagents, catalysts, solvents, etc. have to be present from the very beginning
of the sequence. The second class of MCRs encompasses sequential MCRs where the components
are added in a subsequent well-­defined order from step to step, although the reaction conditions
may not be changed. Consecutive MCRs are the last category where the components are added
stepwise as well but the reaction conditions are altered with each step. All three MCR classes offer
high structural and functional diversity and their explorative potential for the establishment and
investigation of functional molecules is vast. Ideally, MCRs should create molecular complexity and
application diversity starting from simple and readily accessible starting materials on a broad scale.
While MCRs are already often used in medicinal chemistry for rapid lead finding [40–42], applica-
tion of MCRs for the synthesis of chromophores and fluorophores is rather adolescent [1, 39].

Handbook of Aggregation-Induced Emission: Volume 1 Tutorial Lectures and Mechanism Studies, First Edition.
Edited by Youhong Tang and Ben Zhong Tang.
© 2022 John Wiley & Sons Ltd. Published 2022 by John Wiley & Sons Ltd.
456 15  Multicomponent Reactions as Synthetic Design Tools of AIE and Emission Solvatochromic Quinoxalines

a m

Scheme 15.1  MCR-­based chromophore approaches to diversity-­oriented fluorophore formation.

Based on a concept developed during our work on consecutive MCRs [1, 43–44], we concluded
that the MCR could be employed as the chromogenic event where the one-­pot process forms the
fluorophore of interest. Consequently, we have named this concept the chromophore approach
(Scheme 15.1).
Novel fluorophores are particularly interesting for a multitude of applications. However, many
promising chromophores that show strong luminescence in solution undergo aggregation-­caused
quenching (ACQ), thus losing their use in applications in the amorphous or crystalline solid state.
This problem is addressed by induction or enhancement of emission upon aggregation. This effect,
known as aggregation-­induced emission (AIE), has been coined by Tang and co-­workers [45–48],
the enhancement of emission via aggregation (aggregation-­induced enhanced emission, AIEE)
has been established by Park and co-­workers [49]. Both conceptual approaches have gained a lot of
interest due to their promising and multifaceted possibilities.
The phenomenon of aggregation working constructively rather than destructively has been dis-
covered in 2001 [50] and structural motifs are still intensively used today in various AIEgens [51–53].
Consequently, novel structural motifs with distinctive and preferably fine-­tunable AIE and solva-
tochromic properties are highly desired. One of these structural motifs that possess strong AIE
properties is quinoxalines [54]. Quinoxalines have been known for numerous applications espe-
cially for their biological and medicinal activity [55–58] but have been employed for their solva-
tochromic and AIE active behavior just recently [54]. In this synopsis, an overview of the different
synthetic strategies for the generation of quinoxalines via MCRs and one-­pot processes are pre-
sented. Furthermore, the solvatochromic and photophysical properties of quinoxalines and their
AIE behavior shall be highlighted and summarized in a flashlight fashion.

15.2 ­Synthetic Approaches to Quinoxalines via Multicomponent


Reactions and One-­Pot Processes

Crucial building blocks for most multicomponent syntheses of quinoxalines are 1,2-­diaminoarenes.
Starting with the formation of benzils via a 1,4-­diazabicyclo[2.2.2]octane (DABCO) catalyzed oxi-
dation of desoxybenzoins using air oxygen as an oxidant in dimethylformamide and addition of
15.2  ­Synthetic Approaches to Quinoxalines via Multicomponent Reactions and One-­Pot Processe 457

1,2-­diaminoarenes Chen et al. could obtain the 2,3-­substituted quinoxalines in good to excellent
yields. The one-­pot process starts with the formation of a carbanion via deprotonation of the
ketone, a radical chain process initiated by triplet oxygen generates a hydroperoxyl ketone anion
which undergoes elimination to give the intermediate 1,2-­diketone. Then, the quinoxaline is
formed by a Hinsberg reaction starting with a condensation reaction to an imine followed by
DABCO-­mediated oxidation with air oxygen as oxidant and subsequent intramolecular
cyclization [59].
Recently, another synthetic approach via a concluding Hinsberg reaction to give quinoxalines has
been presented by Niesobski et  al. In a resource-­saving process, it was possible to combine two
Pd(II)/Cu(II)-­catalyzed processes. At first, Sonogashira coupling of terminal alkynes and aryl
(pseudo)halides furnish internal alkynes which are transformed by Pd(II)/Cu(II)-­catalyzed Wacker-­
type oxidation using dimethylsulfoxide (DMSO) and oxygen as dual oxidants to give 19 examples of
symmetrically and unsymmetrically substituted 1,2-­diketones. The addition of 1,2-­diaminoarenes
to the reaction mixtures leads finally to the formation of substituted quinoxalines [60].
Hashmi proposed a gold(I)-­catalyzed process for the formation of 2-­substituted quinoxalines
using a pyridine N-­oxide starting from arylacetylenes. The intermediary dicarbonyl product is
obtained via a nucleophilic addition of pyridine N-­oxide to the gold-­coordinated alkyne leading to
vinyl gold intermediate. Nucleophilic attack of another molecule of pyridine N-­oxide and a gold-­
assisted fragmentation releases the dicarbonyl compound. After the addition of diaminoarene, the
2,3-­substituted quinoxalines are formed [61].
Another attempt to gain access to 2,3-­substituted quinoxalines is achieved via a PdCl2/CuCl2
catalyzed reaction on PEG-­400  in the presence of water using internal alkynes as a starting
­material [62]. Starting from α-­bromo ketones, HClO4·SiO2 could be employed as heterogeneous
recyclable catalysts for the synthesis of quinoxalines and dihydropyrazines at room temperature.
Furthermore, it is possible to recycle the catalyst system several times [63].
Quinoxalines can also be synthesized in a continuous flow reactor starting from acyl chlorides
and trimethylsilyl (TMS)-­protected diazomethane to give diazoketones. To transform the TMS-­
protected diazoketone to the corresponding diazoketone, polystyrene-­tetraalkylammonium fluo-
ride salt was used. The subsequent reaction to give quinoxalines was performed by the addition of
the diamionarene in a third stream and then passing through a cartridge containing a sequential
plug of Cu(OTf)2 and polystyrene-­thiourea as well as polystyrene-­tosylhydrazine or polystyrene-­
isocyanate to quench residual diazoketone [63].
Furthermore, implementing a Boc-­protected diaminoarene in the synthetic procedure, an
­Ugi-­based MCR can be applied to generate quinoxaline derivatives by adding an aldehyde and an
isocyanate. The intermediate can be oxidized with air oxygen to give the quinoxaline. Another
­Ugi-­based protocol encompasses the usage of hydrochloric acid in methanol to give the dihydro-
quinoxaline. After treating with 2,3-­dichloro-­5,6-­dicyano-­1,4-­benzoquinone, aromatization
­proceeds to give the carboxyquinoxalines. The same class of quinoxalines is obtained with trifluoro
acetic acid (TFA) followed by the addition of manganese(IV)oxide [64–70].
Mono boc-­protected diaminoarene can also be used for a two-­step Petasis-­deprotonation-­
cyclodehydration-­oxidation reaction with glyoxal derivatives and boronic acids in a microwave reac-
tor. The internal nucleophile of the Petasis product is deprotected, cyclization with the carbonyl group
and oxidation to the quinoxaline are simultaneously triggered by TFA in dichloroethane. With this
protocol, a large variety of quinoxalines is easily-­accessible [71, 72]. Another approach to substituted
quinoxalines involves apart from o-­phenlyenediamine aromatic aldehydes and p-­toluenesulfonylmethyl
isocyanide (TosMIC). The mechanism can be rationalized as follows. The initial step is the formation
of an imine which is nucleophilically attacked by the activated TosMIC. Subsequent ring closure and
eliminating the tosyl moiety leads to the tetrahydroquinoxaline intermediate which is finally oxidized
458 15  Multicomponent Reactions as Synthetic Design Tools of AIE and Emission Solvatochromic Quinoxalines

R2
R1 N R1 R1–X
N O +
R Coupling-
R oxidation R2
R2 N DABCO
Ar N sequence
air
DMF Pd(PPh3)4,CuI
TFA/DCE TosMIC- Iodine NEt3, O2, DMSO
O Based catalysis O
H MCR
N R2
R1 R1
R Ar
R2 O
BocHN ArCHO
TosCH2NC I2
O DMSO
open air Hinsberg
+ R3B(OH)2 Petasis-
R1 reaction
reaction
O
Protection
H2N H2 N Pd/Cu-catalyzed
R reaction R1 N
R
Ugi-based BocHN R
H2 N
reaction and + O R1 R2
PdCl2/CuCl2 R2 N
oxication – N R 2
H
O + PEG-400/H2O
R1
N2 R1
R1 H O2 Gold-catalyzed
reaction [Au]
R 1
N Cu(OTf) 2 R2
1
R O
R
R2 R 1, R 2 =Ph pyridine N-oxide
N N
H Br

HClO 4 · SiO 2
Copper-catalyzed
synthesis
Heterogenous
flow chemistry
catalysis
Ugi-based R1 N
3CR H
R
R1 N N
O R2 N
R + +
H – N R2 H O
N R1 +
O

Scheme 15.2  Overview of synthetic approaches to substituted quinoxalines [59–65, 71–73, 75–79].

to the corresponding ­quinoxalines  [73]. Similarly substituted quinoxalines can be obtained via o-­
nitroanilines and phenyethylamines in an iron/sulfur catalyzed redox condensation process where
sulfur and iron(III)chloride hexahydrate form the reactive cluster [74]. Another feasible catalyst for
the generation of quinoxalines is iodine. The synthesis can either be started by the formation of
1,2-­diketones originating from an alkyne or directly with a diketone and diaminoarenes in DMSO to
gain access to quinoxalines (Scheme 15.2) [75, 76].
Valuable building blocks for the formation of quinoxalines are ynediones. Due to their densely
functionalized electrophilic nature, ynediones are powerful precursors for the generation of differ-
ent heterocycles  [80–81]. To fathom the exploratory potential of ynediones  [82], we developed
several methods to access ynediones, for example by in situ activation of carboxylic acids with
oxalyl chloride [83], by a glyoxylation/Stephens-­Castro coupling sequence starting from electron-­
rich heteroaryl systems [84]. It is noteworthy to mention that Sonogashira coupling (Pd/Cu cataly-
sis) leads to the extrusion of one equivalent of carbon monoxide [85]. Based on this knowledge, we
established two complementary multicomponent one-­pot syntheses towards ethynylquinoxalines.
Starting from heteroaryls, a four-­component glyoxylation-­alkynylation-­cyclocondensation (GACC)
sequence was introduced and an activation-­alkynylation-­cyclocondensation (AACC) sequence
starting from glyoxylic acid, lead both to intermediary generated ynediones by using glyoxylic acid
chloride, copper(I)iodide, alkynes, and triethylamine. The subsequent Hinsberg reaction of the
15.2  ­Synthetic Approaches to Quinoxalines via Multicomponent Reactions and One-­Pot Processe 459

HO O

R1 H
1) 1.00 equiv (COCl)2 R1 O
2) 5.00 mol% CuI
1.00 equiv R2
Cl O 2.20 equiv NEt 3 Cl O
via via
R1 O R2 R1 O
O

R1 O

3) 1.00 equiv 1,2-diaminoarene

GACC AACC
R2
sequence sequence
N
R3
24 examples, Four examples,
28 – 85% R1 N X 25–85%

Scheme 15.3  One-­pot syntheses of ethynylquinoxalines via four-­component-­GACC sequence and


three-­component-­AACC sequence [86–88].

ynediones and 1,2-­diaminoarenes produces the quinoxalines without affecting the triple bond.
Although the GACC sequence for the formation of 3-­ethynylquinoxalines is limited to certain
electron-­rich π-­nucleophiles, the AACC sequence starting from 2-­(hetero)aryl glyoxylic acid is
valuable complementary access to other electronic substitution patterns in the 2-­position of qui-
noxalines. (Scheme 15.3) [86–88].
This one-­pot procedure can even be extended by a concluding aza-­Michael-­addition which can
be either conducted by TMS-­protected or deprotected ethynylquinoxalines. The deprotection
works quantitatively by using potassium fluoride. For the aza-­Michael addition an excess of pyr-
rolidine is added to the reaction mixture so that a second electron-­donating, aminovinyl substitu-
ent is established on the quinoxaline core [86–88].
The combination of pyrrolidine and methanol creates sufficiently basic conditions to cleave
the  TMS group in situ so that a domino desilylation-­addition-­sequence gives rise to
­pyrrolidinovinylquinoxalines even exceeding the combined yields of the stepwise process.
Substituted 2-­(phenylethynyl)quinoxalines were also transformed into 2-­aminovinyl-­quinoxalines.
In the  ­presence of weak acids, the hydrolysis of the enamine was observed yielding
2-­hydroxyvinylquinoxalines. An intramolecular hydrogen bond ensures the (Z)-­configuration of
both 2-­hydroxyvinylquinoxalines. Even in solution, only the enol form can be observed. By using
strong acids, 2-­aminovinylquinoxalines are protonated resulting in the formation of vinamidinium
ions, however, hydrolysis of 2-­aminovinylquinoxalines was not detected. According to density func-
tional theory (DFT)-­calculations, protonation occurs at one of the quinoxaline nitrogen atoms [86,
88–89]. Starting from the deprotected 3-­ethynylquinoxalines another glyoxylation-­Stephens-­Castro
coupling sequence could be established so that three examples of alkynedione derivatives were
obtained.
The formal addition of different hydrogen halides yielded 2-­halovinylquinoxalines. By using
TMS-­substituted 3-­ethynylquinoxalines a one-­pot reaction transforms intermediary
2-­halovinylquinoxalines into quinoxaline based ene-­ynes by Sonogashira coupling. In contrast to
460 15  Multicomponent Reactions as Synthetic Design Tools of AIE and Emission Solvatochromic Quinoxalines

R2
H 2N X
R1 H
Sonogashira-
or + N coupling N
H 2N
HO O
R2 R1 N R1 N
R1 O X = Cl, Br, I
Five examples, Three examples,
GACC-
36–56%
or 65–88%
AACC- (E/Z)-ratios 8:1 - 14:1
sequence
Halo- Michael- O
addition Glyoxylation-
R2 alkynylation- R1
N Deprotection N sequence N
O

R1 N R 2 = TMS R1 N R1 N
Four examples, Three examples,
81%-quant. 43–48 %
Aza-Michael
R 2 = Ph Aza-Michael-
addition
addition
= TMS
R2
3 3 3
(R )2N Ph (R )2N (R )2N
H
N N Protonation N
X
R1 N R1 N R1 N
Five examples, Nine examples, Nine examples,
80–93 % 53 – 89% quant.
(E/Z-ratios) 14:1–33:1 X =Cl, BF 4, SO 3(p-Tol)

Hydrolysis Hydrolysis

Ph OH OH

N N

N N
N N
Me Me
65% 78%

Scheme 15.4  Overview of the chemical transformations of 3-­ethynylquinoxalines [87–89].

2-­halovinylquinoxalines, only trans-­configurated ene-­ynes were obtained with the subsequent


Sonogashira coupling (Scheme 15.4).
The AACC and GACC protocols gave quick and easy access to a huge variety of quinoxalines.
Therefore, we became interested in creating a more sophisticated dye system with enhanced and
novel properties. With this respect, 5-­bromothienyl-­3-­ethynyl quinoxaline was identified as a
building block for extending the π-­system via different cross-­coupling methodologies [90]. Suzuki
coupling of 5-­bromothienyl-­3-­ethynyl quinoxaline with the corresponding boronic acid or pinaco-
lyl boronate leads to 14 examples with aryl substituents ranging from electron-­donating to electron-­
withdrawing. As a standard catalyst Pd(PPh3)4 and sodium carbonate acting as a base were
employed in the reaction sequence. Additionally, bisboronic acids or bispinacolyl boronates were
used for generating dumbbell-­shaped thiophene linked bisquinoxalines under the same Suzuki
conditions [54, 90].
15.2  ­Synthetic Approaches to Quinoxalines via Multicomponent Reactions and One-­Pot Processe 461

R2
3) 1.00 equiv 1,2- R2
HO O 1) 1.00 equiv (COCl) 2 O
diaminoarene N
R3
R1 O 2) 5.00 mol% CuI R1 O
R1 N
1.00 equiv R2
2.10 equiv NEt3
24 examples
10.0 mol% Pd(PPh 3)4 35% - quant.
TIPS 2.50 equiv Na2CO3
N 1.10 equiv B(OR) 2-R1
R3
S
R1 N 1,4-Dioxane/MeOH (2:1)
100 °C, 16 hours
14 examples Suzuki- TIPS
66 – 85% coupling N
R1
S
Br N
10.0 mol% Pd(PPh 3)4
N N
N 2.50 equiv Na2CO3
N
S S 1.10 equiv (B(OR) 2)2-Phn 6.00 mol% Pd(dba) 2
( π )n
5.00 mol% [(tBu)3PH]BF 4
1,4-Dioxane/MeOH (2:1) Buchwald-
n = 1,2 1.10 equiv HNR2R3
100°C, 16 hours Hartwig-
TIPS Two examples, 29 – 50% TIPS coupling 1.15 equiv NaOtBu
1,4-Dioxane
100°C, 16 hours

TIPS
N
R2 R1
S
N N
R3

Two examples
88 – 89%

Scheme 15.5  Synthesis of 3-­ethynylquinoxalines via three-­component activation-­alkinylation-­


cyclocondensation sequence (AACC sequence) and Suzuki and Buchwald–Hartwig-­coupling of 2-­(5-­
bromothiophen-­2-­yl)-­3-­((triisopropylsilyl)ethynyl)quinoxaline [90].

Furthermore, 5-­bromothienyl 3-­ethynyl quinoxaline can be readily aminated under Buchwald–


Hartwig coupling conditions and furnished two examples of aminated 5-­(hetero)aryl-­thienyl-­2-­
substituted ethynyl quinoxaline in excellent yields (Scheme 15.5) [54, 90].
The already established GACC and AACC sequences were expanded even further by a Cu-­
catalyzed alkyne-­azide cycloadditon (CuAAC) to give access to 3-­triazolylquinoxalines in a one-­
pot fashion. The modular nature of the CuAAC-­sequence allowed for the introduction of a variety
of azides like electron-­rich or electron-­poor benzylic substituents, aromatic, sterically hindered,
aliphatic, secondary substituted, and ester-­ and amide-­functionalized azides. This opens avenues
for bioconjugation of peptides to suitable quinoxaline fluorophores. In a resource-­saving manner,
the same catalyst source and loading as in the GACC and AACC sequences can be used as well as
a small amount of sodium ascorbate. Advantageously, both sequences allow the use of the compo-
nents in almost equimolar amounts and the average yields per bond-­forming step range at approxi-
mately 90%. By analyzing the crystallographic packing a centrosymmetric pairwise arrangement
by strong π–π interactions between the triazolyl rings was revealed (Scheme 15.6) [91].
MCRs and one-­pot methodologies offer fast and direct access to highly diversified and complex
molecules starting from simple reactants and thus bearing a large explorative potential. MCRs
encompass several advantages like recycling precious catalysts, economizing work-­up routines due
to unneeded isolation of intermediates and drastically reducing overall reaction times. However, it
has to be noted that MCRs can lead to the formation of undesired byproducts and for establishing
462 15  Multicomponent Reactions as Synthetic Design Tools of AIE and Emission Solvatochromic Quinoxalines

GACC-CuAAC AACC-CuAAC
sequence sequence

HO O
R1 H
R1 O
1) 1.00 equiv (COCl) 2
2) 5.00 mol% CuI
1.00 equiv TMS
2.20 equiv NEt3
3) 1.00 equiv 1,2-diaminoarene

TMS TMS
N N
R2
R1 N R1 N

4) 2.00 equivs KF
3
5) 1.10 equiv R N3
10.0 mol% NaAsc
N N N N
R3 N N R2 N N
R2
R1 N R1 N

18 examples, Three examples,


38 – 83% 62 – 70%

Scheme 15.6  Synthetic approach to 3-­triazolylquinoxalines via five-­component glyoxylation-­alkynylation-­


cyclocondensation-­CuAAC sequence (GACC-­CuAAC sequence, left), and four-­component activation-­
alkynylation-­cyclocondensation-­CuAAC-­sequence (AACC-­CuAAC sequence, right) [91].

novel MCRs and one-­pot procedures lengthy and careful optimization studies are required. By
harvesting concepts of MCR and one-­pot methodology, novel routes to even more complex
quinoxaline-­based structural motifs can be developed directly accessing complex and specifically
designed molecules for applications.

15.3 ­Photophysical Properties and Emission Solvatochromicity


of Quinoxalines

Investigation of the electronic properties of 2-­heteroaryl-­substituted 3-­ethynylquinoxalines reveals


three to four distinct, broad maxima with the longest wavelength band appearing between λmax =
393 and 418  nm and with molar absorption coefficients ε ranging from 8 000 to 14 600  l/mol/
cm  [87, 88]. The remote arylethynyl substituent proves to have only a minor influence on the
absorption behavior. For the substituent on the 2-­position the effect is more pronounced, especially
electron-­rich substituents like 1-­phenylpyrrol-­2-­yl substituents result in a blue-­shift of the absorp-
tion band. According to DFT calculations, for all investigated highest occupied molecular orbitals
(HOMOs) the coefficient density is predominantly located on the quinoxaline moiety and on the
2-­substituent. The longest wavelength absorption bands result from dominant contributions of the
HOMO– lowest unoccupied molecular orbital (LUMO) transitions.
15.3  ­Photophysical Properties and Emission Solvatochromicity of Quinoxaline 463

Figure 15.1  Emission solvatochromism of methoxyphenylethynyl methylindol quinoxaline (recorded


in cyclohexane, toluene, THF, EtOAc, acetone, CH2Cl2, MeCN, iPrOH, EtOH, and MeOH [from left to right];
T = 293 K; c = 10−4 m; λexc = 356 nm, hand-­held UV lamp) [87].

A similar effect can be observed concerning the emission spectra, where the shortest wavelength
emission band varies between λmax = 474 and 520 nm. With the increasing electron-­withdrawing
character of the arylethynyl substituent, the emission maximum is red-­shifted. The Stokes shifts
range between Δν̃ = 4400 and 4800 cm–1 and the quantum yields ΦF between 0.15 and 0.32 for the
1-­methylindole substituted derivatives. The dipolar nature of the excited state can be explored in
more detail by a solvatochromism study with the donor-­substituted 1-­methylindole derivative.
While the absorption spectra display only a minor positive solvatochromism, the influence of the
solvent polarity on the emission band is significant and more pronounced; the emission bands
range from λmax = 468 (cyclohexane) to 525 nm (methanol) (Figure 15.1).
The decrease of solvent polarity leads to an increase of the quantum yield ΦF. The highest quan-
tum yield is detected in toluene with ΦF = 0.30 and the lowest in protic solvents such as alcohols.
This specific solvent effect may be rationalized via a deactivation pathway from the excited state
triggered by intermolecular solvent/solute hydrogen bonding [92].
Additionally, according to the Lippert–Mataga equation, the large positive solvatochromism cor-
relates with a considerable increase of the dipole moment upon excitation. Plotting the solvent-­
dependent Stokes shifts against the orientation polarizability Δf results in excellent goodness of fit
which clearly indicates dominant importance of a general solvent effect. Upon applying the
Lippert–Mataga equation the change of dipole moment from ground to excited state can be calcu-
lated. With computed values for the transition dipole moment Δμ, by DFT calculation in the gas
phase the dipole radius is constituted to 0.51 nm, which accounts to 0.42 nm for the unrelaxed
structure and 0.40 nm for the relaxed structures in dichloromethane [87, 88].
Upon introduction of a strong donor, namely N,N-­dimethylaniline in 2-­position the emission
solvatochromism of 3-­ethynylquinoxalines can be enhanced to cover the complete visible spectral
range from blue-­green to deep red-­orange with a single chromophore in a relatively narrow polar-
ity window. X-­ray crystal structure analysis showed a significant π-­stacking and a sizeable overlap
of the aryl-­planes. The consanguineous series of the 2-­(p-­N,N-­dimethylanilino)-­3-­(trimethyl)
silylethynyl-­substituted quinoxalines reveals the direct correlation of the red-­shift of the longest
wavelength absorption band and the increasing electron-­withdrawing character of the quinoxaline
core. All 2-­(p-­N,N-­dimethylanilino)-­3-­(trimethyl)silylethynyl-­substituted quinoxa­lines are highly
luminescent in solution while the quantum yields ΦF of these compounds add up to 0.89. The
Stokes shifts are generally higher for p-­dimethylamino substituted compounds and the trend of
increasing Stokes shifts correlates with the donor capacities. By enhancing the push–pull character
464 15  Multicomponent Reactions as Synthetic Design Tools of AIE and Emission Solvatochromic Quinoxalines

using strong donors and strong acceptors, the emission in dichloromethane is completely
quenched. The direct correlation of the nonradiative decay of the excited state and the red-­shift of
the emission bands, as laid out in the energy gap law [92, 93], is supported by the occurrence of
fluorescence of these compounds in less polar solvents like n-­pentane, cyclohexane or toluene.
Consequently, these quinoxalines may be applied as polarity sensors.
Correlating the longest wavelength absorption bands, the emission bands, and the Stokes
shifts with σp+ Hammett parameters  [94] establishes linear structure–property relationships.
The negative slope of the correlation of the Stokes shifts indicates a pronounced degree of posi-
tive emission solvatochromism originating from the polar excited state. Furthermore, the larger
slope of the emission correlation in comparison to the absorption correlation underlines the
fairly polar nature of the vibrationally excited state, which can be attributed from a significant
charge-­transfer character of the S1 state. The Hammett correlations allow the conclusion that
the remote substituent effect is transferred by resonance mechanisms with charge-­transfer
­characteristics [54, 88].
The already pronounced solvatochromism of the 3-­ethynylquinoxalines can be enhanced by
increasing the donor capacity while the most dominant effect is observed with the for N,N-­
dimethylaminophenyl derivatives, covering nearly the whole visible spectrum upon variation of
the solvent polarity from n-­pentane to acetonitrile (Figure 15.2) [88].
While the absorption solvatochromism is only minor and ranges between λmax = 430 and 447 nm,
the emission maximum is shifted bathochromically and ranges from greenish-­blue luminescence
(λmax = 478  nm) in cyclohexane to deep red fluorescence in N,N-­dimethylformamide (λmax =
755 nm). This emission solvatochromism correlates with a change in the dipole moment of the
chromophore upon excitation and dipolar relaxation of the surrounding molecules [92].
Consequently, the Lippert–Mataga model was used to treat this effect quantitatively [95]. The
excellent goodness of fit when plotting Stokes shifts against Lippert–Mataga orientation polariza-
bility Δf underlines the importance of a general solvent effect. According to the Lippert–Mataga
equation, the change in dipole moment from the electronic ground state to the vibrationally relaxed
excited state Δμ adds up to 26D by employing an Onsager radius of 12.9 × 10−9 m. This change in
dipole moment corresponds to a charge separation upon charge-­transfer excitation.
Investigating the electronic structure using time-­dependent density functional theory (TD-­DFT)
calculations, the HOMO coefficient densities are predominantly localized on the p-­dimethylamino
substituents and much lower coefficient densities on the quinoxaline cores. The coefficient densi-
ties of the LUMOs show an inverse distribution. The Kohn–Sham frontier molecular orbitals on
the quinoxaline moieties are in good agreement with a significant overlap for the Franck–Condon
transitions resulting in pronounced transition dipole moments and oscillator strength. Additionally,
the predominant charge-­transfer π–π* character of the longest wavelength absorption bands
increases. The distinct solvatochromism of dimethylaniline 3-­ethynyl dichloroquinoxaline is in
good accordance with the theoretical calculations where the change in dipole moment upon exci-
tation to the vibrationally relaxed S1 state correlates with a dominant charge-­transfer
character [88].
Solvatochromism properties are also observed for 2-­[(1E)-­2-­(1H-­pyrrol-­2-­yl)ethenyl]-­quinoxaline
(PQX). PQX exhibits solvatochromic fluorescence behavior starting from deep blue fluorescence in
n-­hexane (λmax = 430 nm) to red fluorescence in an aqueous buffer (λmax = 607 nm) [96].
In n-­hexane, solvation relaxation can be excluded while in other solvents the excited state is
stabilized by solvent relaxation. Upon correlation of the Reichardt’s values ( ETN ) [97] a stabilizing
role of polar substituents for the excited singlet state is revealed. A weaker correlation between
acceptor or donor numbers of the solvents and the emission maxima hints that the excited state of
15.3  ­Photophysical Properties and Emission Solvatochromicity of Quinoxaline 465

Wavelength λ (nm)
800 750 700 650 600 550 500 450 400

1.0 1.0
Normalized fluorescence (a.u.)

Normalized absorbance (a.u.)


0.8 0.8

0.6 0.6

Abs. toluene
Em. toluene
Abs. Cyclohexane
0.4 Em. Cyclohexane
Abs. THF
0.4
Em. THF
Abs. diethyl ether
Em. diethyl ether
Abs. DCM
Em. DCM
0.2 Abs. EtOAc 0.2
Em. EtOAc
Abs. DMF
Em. DMF
Abs. MeCN
Em. MeCN
0.0 0.0
14 000 16 000 18 000 20 000 22 000 24 000

Wavenumber v (cm–1)

TMS
N Cl

N Cl

Me2N

n-Pentane Cyclohexane Toluene 1,4-Dioxane Diethylether Ethyl acetate CH2Cl2 DMF Acetonitrile

Figure 15.2  UV/Vis absorption (solid line) and emission (dashed lines) spectra of TMS-­protected dichloro
dimethylaniline 3-­ethynyl quinoxaline in various solvents with increasing polarity (recorded at T = 293 K,
c = 10–5 m) (above) and fluorescence of this chromophore with increasing solvent polarity (left to right:
n-­pentane, cyclohexane, toluene, 1,4-­dioxane, diethylether, ethyl acetate, dichloromethane, N,N-­dimethyl
formamide, and acetonitrile) under UV-­light (λexc = 365 nm, c = 10–7 m) (below) [54, 88].

PQX having charge transfer characteristics is likely to be zwitterionic. The charge transfer charac-
ter can be rationalized upon molecular orbital calculations which indicate that the HOMO elec-
tron of the pyrrole moiety is transferred to the LUMO of the quinoxaline site by light absorption.
In protic solvents, the quantum yield of PQX decreases drastically while the largest quantum yields
were determined in DMSO. The absorption spectra show no dependence on solvent polarity.
Additionally, the fluorescence lifetimes of PQX are considerably longer in DMSO (τ = 4.1 ns) than
in benzene (τ = 0.19 ns) or in methanol (τ = 0.024 ns). The ratios of rate constants for nonradiative
decay and fluorescence emission indicate that out of the four possible rotational isomers only two
isomers are energetically favored in the ground state and consequently only two rotational isomers
are converted to the excited singlet state upon photoirradiation (Figure 15.3) [96].
466 15  Multicomponent Reactions as Synthetic Design Tools of AIE and Emission Solvatochromic Quinoxalines

n-Hexane Benzene Ethyl acetate Acetone DMSO MeOH Buffer

Hexane
HN
N Benzene
Ethyl acetate
N
Normalized FL intensity

Acetone
DMSO

MeOH
Buffer

400 450 500 550 600 650 700 750


Wavelength (nm)

Figure 15.3  Images of pyrrolvinyl quinoxaline in different solvents (from left to right: n-­hexane benzene,
ethyl acetate acetone, DMSO, methanol, and phosphate puffer at pH = 7.4) under UV-­light (λexc = 365 nm and
fluorescence spectra of this compound in various solvents. (Source: Reprinted from Ref. [96]. Copyright
(2011), with permission from Royal Society of Chemistry).

Using PQX as a polarity sensor for the binding pocket of bovine serum albumin (BSA), the fluo-
rescence is significantly blue-­shifted from λmax = 608 nm in a sole buffer to λmax = 530 nm upon
binding to the pocket. Furthermore, the fluorescence is intensified by 200-­fold. Thus, PQX can be
utilized as a fluorescent polarity sensor for binding sites in proteins and host molecules [96].
Achelle underlined this solvotochromic behavior of vinyl quinoxalines by expanding the
π-­system of vinyl quinoxalines and thus enhancing the fluorescence intensity of the solvatochro-
mic properties. These push–pull systems are readily synthesized by Suzuki cross-­coupling and
show large Stokes shifts up to Δν̃ = 5300 cm–1. Furthermore, the quinoxalines possess enhanced
applicable nonlinear optical properties (NLO) due to an extended π-­spacer. Vinyl quinoxalines also
display halochromic properties upon protonation by the addition of TFA complete quenching of
fluorescence accompanied by a massive bathochromic shift of the absorption maxima is observed.
This color change and fluorescence quenching is completely reversible upon neutralization with a
base [98].
Divinyl quinoxalines are potent D-­π-­A chromophores and exhibit either positive or negative
solvatochromism upon changing the substitution pattern on the vinyl moieties. This behavior can
be rationalized by two different possible resonance forms, a nonionic form and a zwitterionic form
can be considered to be a highly polar “quinoid” structure. Positive solvatochromism can be attrib-
uted to an increase of the dipole moment upon excitation and a predominance of the low polar
15.3  ­Photophysical Properties and Emission Solvatochromicity of Quinoxaline 467

form in the ground state. The inversion of solvatochromism (positive solvatochromism predating
negative solvatochromism) indicates the increase of quinoidization induced by intramolecular
charge transfer (ICT). With increasing solvent polarity the contribution of the charge-­separated
resonance structure in the ground state can be enhanced [99, 100].
A common photophysical property of quinoxalines is the acidochromicity behavior due to the
protonation of the nitrogen atoms. For selected arylated quinoxalines, fluorescence can be induced
by protonation with TFA which, in turn, can be reversed by adding a base to the solution. The
­addition of different acids also leads to varying fluorescence with shifted emission maxima.
Therefore, 2,3-diarylated quinoxalines can be utilized as selective sensors for detecting and identi-
fying protonic acids in solution and in a vapor state [101].
Benzo[g]indolo[2,3-­b]quinoxaline derivatives also show halochromism and protonation-­induced
assembly. This thiophene-­linked bisquinoxaline can be protonated at each quinoxaline moiety
separately leading to a bathochromic shift of the absorption bands. Upon adding more acid, a
tetraprotonated species is obtained which can be observed by a more distinct bathochromic shift of
absorption maxima accompanied by precipitation of the dye. Thin films of the diprotonated
benzo[g]indolo[2,3-­b]quinoxaline can be generated by drop-­casting or spin-­coating giving rise to a
solid-­state acid vapor sensor that changes color upon exposition to protic acid vapors [102].
2,3-­Bispyridino quinoxaline exhibits acidochromic behavior as well, where emission is hypsoch-
romically shifted from greenish to blue fluorescence. Additionally, using picric acid instead of
other acids a new, red-­shifted absorption maximum can be observed, as well as fluorescence
quenching by increasing the amount of picric acid. Consequently, 2,3-­bispyridino quinoxaline
can  be used as a selective sensor for picric acid. This quinoxaline derivative additionally is
solvatochromic [103].
Phenylcarbazol substituted quinoxalines reveal charge-­transfer character and solvatochromicity.
The magnitude of solvatochromic emission shift is small for anthracenylamine derivatives and
larger for pyrenyl and fluorenylamine derivatives which also possess higher quantum yields. This
trend of quantum yields and solvent dependence of emission is mainly related to the decay process
due to oxidative quenching by the amino moieties [104]. The decay process is fostered in polar
solvents [105].
Condensation and Buchwald–Hartwig coupling procedures gave access to 2,3-­di(thiophen-­2-­yl)
quinoxaline amine derivatives which possess clearly positive solvatochromic properties in emis-
sion as well with quantum yields ranging from ΦF = 0.01 to 0.26 [106].
The photophysical investigations of condensation products of 1,2-­diaminoarenes with diketo-­
substituted indoles, namely indole-­substituted quinoxalines, reveal that the degree of conjugation
of the substituent is more important than the electronic effect, allowing a selective fine-­tuning of
the fluorescence properties. Indole-­substituted quinoxalines show no absorption solvatochromic-
ity but positive solvatochromism in the emission spectra. This behavior may be explained by
enhanced intermolecular solvent/solute hydrogen bonding which could trigger deactivation of the
molecular excited state [107, 108].
T-­shaped donor–acceptor–donor type tetraphenylethylene substituted quinoxaline derivatives
display an ICT band due to their D–A–D structure. The ICT transitions are highly responsive to
solvent polarity thus exhibiting solvatochromic behavior. Tetraphenylethylene substituted quinox-
aline undergoes a bathochromic shift of absorption and emission maxima with increasing solvent
polarity. Due to the weak absorption solvatochromism, it can be stated that the ground-­state elec-
tronic structures are not affected by the solvent polarity. The notable emission solvatochromism is
attributed to more polarized excited states which can be reasoned with the enhanced charge sepa-
ration and a higher dipole moment caused by charge transfer transitions. A Lippert–Mataga plot
468 15  Multicomponent Reactions as Synthetic Design Tools of AIE and Emission Solvatochromic Quinoxalines

reveals a linear correlation underlining the higher polarization of the excited state compared to the
ground state [109].
Quinoxaline-­based cross-­conjugated luminophores with triphenylamine and aromatic substitu-
ents on 2-­and 3-­position establish a promising class of D–A–D chromophores [110]. The electron-­
donating character of the triphenylamines and the electron-­withdrawing quinoxaline accounts for
ICT transitions. In good accordance to the small dipole moments as a consequence of ICT and
solvent polarity independent electronic structures of the ground states, the absorption spectra
show nearly no difference in solvents of different polarities. On the other hand, the emission spec-
tra differ significantly, thus showing strong positive fluorescence solvatochromic behavior.
Additionally, with increasing solvent polarity the fluorescence intensity decreases. Using the
Lippert–Mataga equation a linear correlation of the Stokes shifts and the orientational polarizabil-
ity becomes apparent. This indicates the polarization of the excited state and a larger dipole
moment in the excited than in the ground state. The charge transfer from electron-­donating triph-
enylamine to the electron-­withdrawing quinoxaline occurs in polar solvents and leads to an
enhanced excited state dipole moment. More polar solvents stabilize the polarized excited state
and cause a bathochromic shift of the fluorescence maximum. The reduced fluorescence intensity
in more polar solvents accounts for the decay of ICT by nonradiative channels [111].
Bunz introduced hitherto almost unknown heteroacene-­derivatives namely alkynylated diazadi-
oxaacenes which also embed quinoxaline moieties. These unusual structural motifs were synthe-
sized via ruthenium catalysis followed by a Sonogashira coupling. While the absorption maxima
do not change by varying the solvent from n-­hexane to methanol, the emission spectra are sensitive
to solvent polarity and exhibited a bathochromic shift from blue to green emission. No specific
interactions between solvent and dye can be proclaimed except for the polarizability as stated by a
Lippert–Mataga equation [92]. The value of methanol is omitted in the Lippert–Mataga analysis to
exclude superposing hydrogen bonding effects. Quantum yields for these compounds range from
ΦF = 0.31 to 0.52 [112].
Bipolar N,N-­diphenylaminoaryl-­substituted hexatriphenylenes, tetraazaphenanthrene, and qui-
noxalines show strong fluorescence in different solvents and in solid states. The fluorescence max-
ima are red-­shifted in dependence of core size, length of aromatic spacer, and polarity increase of
the solvents. For quinoxalines with a symmetric discotic structure the Stokes shifts add up to
9000 cm–1. Fluorescence colors range from green (λmax = 480 nm) to red (λmax = 630 nm) and the
highest quantum yields are observed in toluene with ΦF = 0.85 and decrease with increasing sol-
vent polarity. Without N,N-­diphenylaminoaryl groups, these quinoxalines are not affected by sol-
vent polarity and fluoresce only weakly [113].

15.4 ­AIE Characteristics and Effects of Quinoxalines

The aforementioned pyrrolidinylvinylquinoxaline (PVQ) which were established in our research


group, were intensively studied in organic solvents and solvent–water mixtures, as dye aggregates,
solids, and encapsulated in polystyrene particles (PSP). Crystal structures reveal a coplanarity of
the quinoxaline ring and the conjugated vinylic double bond whereas the torsion angle between
the quinoxaline plane and the 2-­substituent varies significantly in size and steric demands. The
different torsion angles affect the packing of the PVQ derivatives, namely phenyl-­, pyrrol-­, indolyl-­,
and thiophene-­substituted PVQs. Despite their large torsion angles phenyl-­and pyrrol-­substituted
dyes form despite their large torsion angles only pair packing. Indolyl-­and thiophene-­substituted
dyes, due to their smaller torsion angles, show centrosymmetric or acentric column packing,
15.4  ­AIE Characteristics and Effects of Quinoxaline 469

which is not as dense as for pair packing. The influence of molecular packing on the fluorescence
properties is related to two opposing effects. For instance, with an increasing packing distance
between the dye molecules, π–π stacking interactions are reduced and should consequently lead to
an enhanced emission compared to π-­stacked dyes. Smaller distances of the dye molecules in the
crystals restrict intramolecular rotations leading to the diminishing of nonradiative decay rates
and consequently favoring the radiative deactivation pathways. The extent of the observed AIE is
determined by these two competing effects [114].
The longest fluorescence lifetime is observed for the indolyl-­substituted PVQ with τ = 6.03 ns
and it possesses the most intense fluorescence with a quantum yield of ΦF = 0.20, which suggests
that for this example, rotational motions are particularly restricted. As mentioned earlier, PVQs
show charge-­transfer-­type emission, strong positive solvatochromism, and hydrogen bonding-­
induced fluorescence quenching by protic solvents such as alcohols or upon protonation [86, 114].
Examining fluorescence in ethanol/water mixtures a strong emission is revealed (Figure 15.4, mid-
dle). By adding surfactants like cetyltrimethylammonium bromide (CTAB) or sodium dodecyl sul-
fate (SDS) at a concentration below the critical micelle concentration the initiation of crystallization
and even the formation of different crystal phases is observed [115–118]. This immediately leads to
the formation of heterogeneous highly fluorescent micrometer-­sized crystals with surfactant shape
control. The usage of SDS gives rise to needle-­like structures with the length of 10–30 μm, whereas
with CTAB, needle-­like and hexagonal platelets are obtained (Figure 15.4, above) [114].
Absorption spectra of PVQ indicate aggregation due to an increase in baseline and background
which is typical for scattering effects. A red-­shift in absorption because of the increased polarity of
dye microenvironment and a broadening of the detected absorption band clearly hints at the het-
erogeneity of the microenvironment. These effects are predominantly observed for the indolyl-­
substituted PVQs due to extended conjugation.
An enhanced polarity of the dye’s microenvironment additionally results in a red-­shifted emis-
sion. At higher water fractions of 80–90%, the emission is blue-­shifted and the fluorescence inten-
sity is enhanced. The thiophenyl-­substituted PVQ undergoes a continuous red-­shift which accounts
to the fact that with decreasing torsion angle and increasing dye–dye stacking distance, the shield-
ing of dye molecules in the inner core of dye aggregates is diminished. With increasing water frac-
tion, and thus the beginning of dye aggregation, quantum yields and fluorescence lifetimes strongly
increase for all dyes by factors of 4.5–6 up to 8 for quantum yields and by factors between 6.6 and
23 for fluorescence lifetimes, respectively, until the aggregate formation is completed. This alludes
to the aggregation-­induced restriction of intramolecular flexibility and rotations. The aforemen-
tioned polarity and bonding-­induced fluorescence quenching are prevented by increasing aggre-
gate size, which automatically raises the number of dye molecules in a less polar environment in
the core of the dye aggregate and may be exploited as an analytical tool. An increasing torsion
angle leading to an increasing extent of the AIE effect consequently underlines the structural con-
trol of the torsion angle and the stacking of the PVQ dyes imposed by the substitution pattern.
Comparing the quantum yields and lifetimes of aggregates and crystalline solids, the aggregates
exhibit lower quantum yields and shorter lifetimes. This is attributed to dye molecules at the aggre-
gate surface, where the emission is probably partially quenched by the high polarity of the sur-
rounding water molecules and by hydrogen bonding [114].
The dried aggregate particles of pyrrolyl-­substituted PVQ have additionally been studied by
transmission electron microscopy (TEM) and reveal almost spherical aggregates. Hydrodynamic
diameters from different dynamic light scattering (DLS) measurements state that the hydrody-
namic diameters range between 190 and 430 nm with a polydispersity index between 0.1 and 0.3.
The torsion angle influences the nucleation process and the packing of the molecules during
470 15  Multicomponent Reactions as Synthetic Design Tools of AIE and Emission Solvatochromic Quinoxalines

20 μm 1 CTAB 20 μm 3 CTAB
1 3
Water fraction (%) 610 580
λ ex, max (nm)

λ ex, max (nm)


3 0 600 3
50 590 570
80
580
90
570 560
2 95 0 20 40 60 80 100 2 0 20 40 60 80 100
98 Water fraction (%) Water fraction (%)

99

1 1

0 0

N N

N N

N N
N
80% water 98% water 80% water Me 98% water

Figure 15.4  Confocal laser scanning microscopy (CLSM) images (λexc = 488 nm) of crystal of phenyl (left)
and methylindole pyrrolidinvinylquionoxalines (right) (above), area-­normalized, absorption-­weighted
emission spectra in ethanol–water mixtures of different water contents excited in the corresponding
maxima (middle) and images in ethanol–water solutions of 20/80 (v/v) and 2/98 (v/v) under UV-­light
(λexc = 365) (below) [86, 114].

aggregation. Smaller torsion angles lead to the smallest aggregates and larger torsion angles to
larger aggregates. Upon increasing the water content, the hydrodynamic diameters of all dye
aggregates drop. A faster aggregate nucleation for hydrophobic dye molecules with increasing
polarity is the consequence, giving access to more aggregate nuclei. This leads to fewer dye mole-
cules during the growth process, yielding smaller aggregates. Upon prolongation of the growth
times, macroscopic structures are formed [114].
15.4  ­AIE Characteristics and Effects of Quinoxaline 471

The phenyl-­substituted PVQ with the most pronounced AIE effect was furthermore incorpo-
rated into 1 μm sized carboxy-­functionalized polystyrene particles (PSP) using a swelling proce-
dure [119]. Using integrating sphere spectroscopy and fluorescence microscopy in dispersion on a
single-­particle level for investigation of the dye-­stained particles it is possible to assess the rele-
vance of the steric restrictions of intramolecular rotations of the PSP-­entrapped AIE dye molecules
in nonpolar solvents. As expected, the less polar polystyrene matrix leads to a blue-­shift of the
excitation and emission spectra of the phenyl-­substituted PVQ in comparison to the spectra of dye
in ethanol and the dye aggregates in ethanol–water mixtures. The quantum yield is enhanced upon
encapsulation in PSP and the lifetimes are prolonged and thus combining the influence of the
reduced polarity of the dye microenvironment and partial restriction of molecular motions. As
proved by microscopic studies of single particles a homogenous dye staining of the particle is
observed. The fluorescence lifetime determined via fluorescence lifetime imaging microscopy
(FLIM) adds up to τ = 6.10 ± 0.10 ns and has been averaged over ten single particles and matches
with the data of the ensemble studies (Figure 15.4) [86, 114, 119].
The previously mentioned 5-­(hetero)aryl-­thien-­2-­yl substituted 3-­ethynyl quinoxalines also pos-
sess strong AIE properties and were therefore closely investigated. The most dominant effects are
observed for the 5-­diphenyl-­thien-­2-­yl substituted 3-­ethynyl quinoxalines and, especially, the 5-­di
methoxytriphenylamine-­thien-­2-­yl substituted 3-­ethynyl quinoxalines. Investigations were carried
out in acetonitrile/water mixtures with ratios ranging from 0 to 95%. Fluorescence cannot be
detected up to a water fraction of 50%. At water fractions above 50%, the formation of dye aggre-
gates is induced and therefore the dye starts to fluoresce. At water fractions higher than 80% the
binary solvent mixture simultaneously increases the polarity of the microenvironment and an
opalescence of the emission color can be observed (Figure 15.5) [54, 90].

Water fraction (%)


0 20 40 60 80 100
8 × 10–5

7 × 105

6 × 105
0% 10% 20% 30% 40% 50% 60% 70% 80% 90% 95%

5 × 105
PL intensity (a.u.)

4 × 105
Nonemissive Emissive aggregate
3 × 105 molecular solution suspension
MeO
TIPS
2 × 105 N

N S
N
1 × 105

MeO
0
0 20 40 60 80 100
Water fraction (%)

Figure 15.5  Plot of fluorescence intensity versus composition of acetonitrile/water mixtures of


triisopropylethynyl triphenylaminothiophene substituted quinoxaline (λexc = 430 nm, c = 1.43 × 10−7 m) and
fluorescence images of acetontrile/water mixtures with different water contents (increasing from left to
right) under UV-­light (λexc = 365 nm) [54, 90].
472 15  Multicomponent Reactions as Synthetic Design Tools of AIE and Emission Solvatochromic Quinoxalines

To explore rotational dynamics and restrictions two derivatives were prepared using Buchwald–
Hartwig amination, namely 5-­diphenyl-­thien-­2-­yl substituted 3-­ethynyl quinoxaline and
5-­phenothiazinyl-­thien-­2-­yl substituted 3-­ethynyl quinoxaline. In the case of the latter, the phenyl
rings are covalently locked by the thiazine-­sulfur bridge. Comparing these two examples with
respect to their AIE properties, the phenothiazinyl-­substituted quinoxaline shows no changes in
its emission characteristics. It can be reasoned that upon aggregation intramolecular motions like
rotation and vibration of freely rotatable aryl substituents, which favor radiationless deactivation,
become more restricted. For the 5-­diphenyl-­thien-­2-­yl substituted 3-­ethynyl quinoxaline, fluores-
cence increases above a water fraction of 60%, and AIE effects predominate, but upon further
increasing the water fraction, partial ACQ effect occurs. Up to a water fraction of 60%, the environ-
mental polarity of the dye is decreased with concomitant restriction of intramolecular rotation
(RIR) taking over. Additionally, a minor fluorescence was detected in pure acetonitrile [54, 90].
Kamble could prove that phenylmethanone quinoxalines are powerful D–A molecules with
­distinct AIE properties. These structural motifs were generated by condensation of
3,4-­diaminobenzophenone and dibromobenzil, which were then treated under Buchwald–Hartwig
conditions to yield the amine phenyl quinoxaline-­phenylmethanones  [120]. Quantum yields of
these molecules are in the range of ΦF = 0.15–0.52 and exhibit a positive solvatochromism. High
dipole moments and a highly polarized excited state have been identified and an intense fluores-
cence in solid films is observed. Therefore, AIE studies using tetrahydrofuran (THF)/water mix-
tures were employed and by adding increasing amounts of water to the solution nanoaggregates
were obtained. In pure THF the dyes fluoresce despite ICT from donor amines to acceptor quinoxa-
line moiety. The largest fluorescence is observed for molecules with bulky groups like naphthyl
phenyl in THF as a solvent. The bisbenzo-­fused five-­membered ring in the carbazole hinders the
intramolecular rotations leading to an enhanced fluorescence due to conformational rigidity. The
fluorescence is weakened if an intramolecular twisting by rotatable phenyl rings of a diphe-
nylamine moiety is possible. Up to water fractions of 50% fluorescence is nearly completely
quenched. Upon raising the water fraction above 60% aggregation occurs and the mixture starts
again to fluoresce intensively. Only at a water fraction of 100% no fluorescence can be detected
caused by the insolubility of these organic dyes in water. Due to physical constraint the nonradia-
tive pathway is blocked and activates radiative channels, thus enabling the molecules to emit
strongly and hinting to applications in OLEDs and OFETs (Figure 15.6) [121].
Emission colors in aggregated form range from yellow-­green to blue-­green at water fractions
from 60 to 90%. A possible explanation for this emission color shift is morphological changes in
aggregates caused by different conformations and packing modes  [122]. The AIE behavior was
additionally investigated via absorption spectroscopy and clearly shows scattering effects caused
by the formed nanoparticles during aggregation [123].
Field emission gun scanning electron microscopy (FEG-­SEM) was used to determine the size
and morphology of nanoparticles for the diphenylamine substituted phenylmethanone quinoxa-
line. These nanoparticles were obtained from a THF/water mixture with a water fraction of 70%
giving spherical shapes with a mean diameter of 30–50 nm. Furthermore, the nanoparticle suspen-
sion formed upon increasing water fractions to more than 60% is transparent and stable for longer
periods without precipitation. According to previous reports, the carbonyl groups may be a reason
for the observed behavior due to induced electronic repulsion preventing the particle to flocculate
or coalesce in water [49].
By ruthenium-­catalyzed oxidative annulation of different 2-­phenylquinoxalines and diphenyla-
cetylenes various quinoxalinium salts can be obtained. Upon photophysical characterization, two
distinct absorption bands are observed and a prominent charge transfer band is detected and
15.4  ­AIE Characteristics and Effects of Quinoxaline 473

N
O
N

800
800 800
0% 2 Intensity
700
10% 2 Wavelength
20%
Emission intensity (a.u.)
Emission intensity (a.u.)

600 600 600


30%

Wavelength (nm)
40%
500
50%
60% 400
400 400
70%
80%
300
90%
200 200
200

100
0 0
0
450 500 550 600 650 700 750 0 20 40 60 80 100
Wavelength (nm) Water fractions (Vol%)

Figure 15.6  Plot images of bisdiphenylaminophenylquinoxaline-­phenylmethanone in THF–water


mixtures with increasing water fractions (c = 10−5 m) by 10% under UV-­light (λexc = 365 nm) (above) and
fluorescence spectra (c = 10−5 m) in THF–water mixtures with increasing water fractions (fw) (below, left) and
plots of intensity and peak wavelength versus fw (below, right) (Source: Reprinted from Ref. [120]. Copyright
(2019), with permission from Elsevier).

confirmed by TD-­DFT calculations. Furthermore, depending on the substitution pattern the qui-
noxalinium salts fluoresce strongly with emission maxima ranging from λmax = 488 to 645 nm
with large Stokes shifts. Due to the donor-­π-­acceptor configuration of the quinoxalinium salts
fluorescence intensity is high. For salts dissolved in dimethylformamide/water mixtures, an
AIEE-­active luminogenic character is obvious upon an increasing water fraction. The formation
of nanoaggregates leads to a 6.36-­fold increase in fluorescence intensity. On one hand, in the
solid-­state the two phenyl rings of the diphenylacetylenes are twisted out of the quinoxaline core
giving rise to a diminished π-­conjugation and green emission. On the other hand, in aggregated
form planarization extends the π-­conjugation and causes a bathochromic shift to yellowish
emission [124].
Time-­resolved photoluminescence reveals that at higher water fractions the fluorescence life-­
times are clearly prolonged, stating an increase of excited state stability. At higher concentrations,
crystalline aggregates are formed and at lower concentrations, distorted spherical aggregates are
474 15  Multicomponent Reactions as Synthetic Design Tools of AIE and Emission Solvatochromic Quinoxalines

obtained. Freezing the dimethylformamide solution to 0 °C leads to a fluorescence enhancement


with thermochromic behavior and a positive solvatochromism can be observed [124].
Selected expanded quinoxaline luminogens show simultaneously twisted intramolecular charge
transfer (TICT) and AIE characteristics and are applicable as turn ON fluorescent probes for glu-
tathione. These properties are obtained by covalently linking triphenylethylene and pyridine moie-
ties to the quinoxaline core (QUPY). QUPY has a donor–acceptor structure and dominantly
possesses AIEE characteristics in DMSO/water mixtures. A locally-­excited (LE) state and a (TICT)
can be observed. In pure DMSO, the LE-­state is active and the emission maximum is located at
λmax = 473 nm. The emission of the LE-­state is deactivated at water fractions of above 30% and the
emission maximum is bathochromically shifted to λmax = 519 nm as a synergetic TICT–AIEE effect.
Upon aggregation, the fluorescence quantum yield increases from ΦF = 0.04 to 0.15. In good
accordance with the characteristics of nanoparticles, the UV/vis spectrum in the aggregated state
is broader with level-­off tails and baseline drift. Similar behavior has been observed in acetonitrile/
water mixtures [125].
Another example of QUPY was obtained by functionalizing the pyridinium group with a dinitro-
phenoxy benzyl moiety. These modified QUPY forms small and compact aggregates at higher
water fractions up to 70% and the increase of fluorescence compared to pure DMSO is even higher.
At water fractions exceeding 70%, the molecules agglomerate to give rise to less emissive amor-
phous particles with smaller sizes [126, 127]. The increased fluorescence upon the incorporation
of glutathione can also be applied as the fluorescent light-­up probe  [125]. The previously men-
tioned sequence of condensation and Buchwald–Hartwig amination can also be used to generate
2,3-­di(thiophen-­2-­yl)quinoxaline amine derivatives [106]. Absorption spectroscopy reveals an ICT
from peripheral thiophene to the electron-­deficient quinoxaline moiety and from the amine to the
quinoxaline core [128].
While only a weak solvatochromic effect is detectable in the absorption spectra, a strong positive
solvatochromic behavior can be observed in emission. Expectedly, solvent polarity affects the ICT
transition. Furthermore, the bathochromic shift in emission suggests that polarized excited states
of dyes are stabilized in more polar solvents due to a decrease in energy of the charge transfer. The
fact that some 2,3-­di(thiophen-­2-­yl)quinoxaline amine derivatives exhibit solid-­state fluorescence
lead to the conclusion to employ AIE studies in THF/water mixtures. The strong green or blue
fluorescence is quenched at lower water fractions, but the emission is turned on at water fractions
above 70% and then decreases again. A clear red-­shift of the emission upon aggregation can also be
observed. The fluorescence in pure THF may be explained due to steric and orbital localization
effects [129, 130]. The enhancement of emission can be rationalized with the formation of nanoag-
gregate and thus reducing unfavorable interactions of water and dye particles. As a consequence,
the dipole–dipole interaction of aggregates dominates over solvent-­aggregate interactions  [131,
132]. The size and morphology of the nanoparticles were measured and indicate the formation of
nanometer-­scale particles, which are spherical with a diameter of 15–30 nm [128].
T-­shaped donor–acceptor–donor type tetraphenylethylene substituted quinoxaline derivatives
not only show solvatochromism but also AIE properties due to the prototypical AIE system namely
the tetraphenylethene moiety. Using THF/water mixtures the AIE properties were investigated. At
low water fractions, up to 40% donor–acceptor–donor type tetraphenylethylene substituted qui-
noxaline derivatives fluoresce intensively because of the ICT from the donor to the acceptor moi-
ety. Further raising the water fraction leads to fluorescence quenching, probably caused by
stabilization of charge transfer upon increasing solvent polarity. At water fractions above 50%, the
formation of aggregates is initiated while simultaneously enhancing the fluorescence. The forma-
tion of nanoaggregates was confirmed by light scattering experiments. Additionally, the
15.4  ­AIE Characteristics and Effects of Quinoxaline 475

donor–acceptor–donor type tetraphenylethylene substituted quinoxaline derivatives exhibit strong


mechanochromic properties [109].
Liu introduced another quinoxaline-­based NIR AIE luminogen consisting of two tetraphenylethene
units used as electron donors and molecular rotators and  [1,2,5]thiadiazolo[3,4-­g]quinoxaline
(BTPETQ). AIE investigations carried out in THF/water mixtures reveal an emission maximum at
λmax = 700 nm. Upon addition of water up to a water fraction of 30%, the emission is decreased possibly
attributed to a twisted intramolecular transfer [133, 134]. Upon increasing water fraction above 30% the
fluorescence increases drastically indicating the formation of nanoaggregates. BTPETQ dots can be
fabricated by nanoprecipitation and can be used for detecting deep cerebral and tumor vasculatures by
NIR-­excited intravital two-­photon microscopy exploiting the AIE characteristics of BTPETQ [135].
The picric-­acid sensitive 6,7-­dimethyl-­2,3-­bis(2-­pyridyl)-­quinoxaline (BPQ) also shows AIEE
behavior in THF/water mixtures. With increasing water fraction the BPQ aggregates and the emis-
sion is intensified. The emission maximum is reached at a water fraction of 90% with a 475-­fold
intensified fluorescence at λmax = 398 nm compared to emission in pure THF. At a water fraction
of 99.9% the emission intensity drops most commonly caused by the possible formation of amor-
phous particles and precipitation. At lower water fractions, BPQ can most possibly assemble in a
more ordered fashion giving rise to more emissive, crystalline aggregates. Following the spectral
evolution of aquatic solutions of BPQ the fluorescence intensity reaches its maximum 20 minutes
after its preparation due to particle formation and partial crystallization. Furthermore, fluores-
cence intensifies with the increasing concentration of BPQ [103].
5,12-­Diacetyl-­5,12-­dihydroquinoxalino[2,3-­b]quinoxalines were obtained by condensation of
phenylendiamine and 2,3-­dichloroquinoxaline, the dihydroquinoxalinoquinoxalines were then
acylated by heating in acetic anhydride [136]. All acylated dihydroquinoxalinoquinoxalines exhibit
fluorescence in solid-­state and upon aggregation. The AIE studies were performed in
dichloromethane/n-­hexane mixtures because using water leads to precipitation of the compound.
With increasing n-­hexane fraction the fluorescence increased five times with emission bands
resembling the emission maxima in solid-­state [137].
Another AIE system based on a quinoxaline core is 2,3-­diphenylquinoxaline-­6,7-­dicarboimide.
These molecules were synthesized via condensation of 1,2-­diphenyl-­1,2-­ethanedione and benzil to
generate the quinoxaline core. This core was modified by reaction with copper(I)cyanide in
dimethylformamide yielding the 2,3-­diphenylquinoxaline-­6,7-­dicarboimide. AIE studies were per-
formed in THF/water mixtures and show an increase of fluorescence intensity starting from a
quantum yield of ΦF = 0.0035 to 0.075. The quantum yields in the solid and aggregated states are
not as high as shown by other contemporary AIE dyes, such as hexaphenylsilole or tetraphenyle-
thene. For further increasing ΦF the introduction of more functional, favorably aromatic groups
which can act as free rotators whose motion is restricted upon aggregation represents a promising
strategy to higher ΦF of the aggregates. Another possible approach may be triggered by the forma-
tion of bichromophores with quinoxaline dyes acting as one of the components. Energy transfer
between chromophores and intertwining fluorescence and aggregation properties offer an auspi-
cious route to interesting new functional materials. DLS and absorption measurements support
the occurrence of AIE. Absorption spectra reveal a level-­off tail in the visible region, which can be
attributed to light scattering of aggregate suspension [138, 139].
X-­ray crystallography states typical propeller-­like structures of the two phenyl groups indicating
that AIE behavior occurs because of an RIR process. These 2,3-­diphenylquinoxaline-­6,7-­dicarbo-im
ides are also used for the formation of dimeric neutral mercury complexes thus changing the pho-
tophysical properties of this compound profoundly and can thereby be used as a highly selective
sensor for mercury ions [140].
476 15  Multicomponent Reactions as Synthetic Design Tools of AIE and Emission Solvatochromic Quinoxalines

15.5 ­Conclusion

In recent years quinoxaline-­based structural motifs gained ever-­growing attention not only due to
their medicinal and pharmaceutical applications but also for their photophysical properties.
Quinoxaline proved to be a versatile chromogenic system with distinctively tunable AIE and solva-
tochromic characteristics depending on the substitution pattern of the quinoxaline core. The elec-
tronic effects of quinoxalines are mainly determined by their high electron affinity which can be
easily exploited in various ways. The facile and elegant synthetic access to quinoxalines offers a
great variety of AIE genic structures. Especially MCRs and one-­pot methodologies have become
powerful tools for the development of novel quinoxaline-­based fluorophores. MCRs enable a man-
ifold and diversity-­oriented exploration of solvatochromic and AIE emissive quinoxalines and will
further support establishing substituted and anellated quinoxalines as common motifs for AIE
chromophores with immense potential for a multitude of applications as outlined in this contribu-
tion especially in analytics and sensing applications of ions and acids.

­Acknowledgments

The authors gratefully acknowledge the continuous support by the Fonds der Chemischen
Industrie.

­References

1 Müller, T.J.J. and Bunz, U.H. (2007). Functional organic materials: syntheses, strategies and
applications. Weinheim: John Wiley & Sons.
2 Müller, T.J.J. (2016). Multicomponent and domino syntheses of AIE chromophores, in Aggregation
Induced Emission: Materials and Applications. (ed. M. Fujiki, B. Liu, B. Z. Tang), 85–112, Oxford:
Oxford University Press.
3 Petty, M.C., Bryce, M.R. and Bloor, D. (1995). An introduction to molecular electronics. Oxford:
Oxford University Press.
4 Müllen, K. and Wegner, G. (2008). Electronic materials: the oligomer approach. Weinheim: John
Wiley & Sons.
5 Carroll, R.L. and Gorman, C.B. (2002). The genesis of molecular electronics. Angew. Chem. Int. Ed.
41: 4378–4400. https://doi.org/10.1002/1521-­3773(20021202)41:23<4378::
AID-­ANIE4378>3.0.CO;2-­A.
6 Walzer, K., Maennig, B., Pfeiffer, M. and Leo, K. (2007). Highly efficient organic devices based on
electrically doped transport layers. Chem. Rev. 107: 1233–1271. https://doi.org/10.1021/cr050156n.
7 Coropceanu, V., Cornil, J., da Silva Filho, D.A., Olivier, Y., Silbey, R. and Brédas, J.-­L. (2007).
Charge transport in organic semiconductors. Chem. Rev. 107: 926–952. https://doi.org/10.1021/
cr050140x.
8 Shirota, Y. and Kageyama, H. (2007). Charge carrier transporting molecular materials and their
applications in devices. Chem. Rev. 107: 953–1010. https://doi.org/10.1021/cr050143+.
9 Torsi, L., Magliulo, M., Manoli, K. and Palazzo, G. (2013). Organic field-­effect transistor sensors: a
tutorial review. Chem. Soc. Rev. 42: 8612–8628. https://doi.org/10.1039/C3CS60127G.
10 Kymissis, I. (2008). Organic field effect transistors: theory, fabrication and characterization.
New York: Springer Science & Business Media.
  ­Reference 477

11 Mishra, A., Fischer, M.K. and Bäuerle, P. (2009). Metal-­free organic dyes for dye-­sensitized solar
cells: from structure: property relationships to design rules. Angew. Chem. Int. Ed. 48: 2474–2499.
https://doi.org/10.1002/anie.200804709.
12 Grätzel, M. (2003). Solar cells to dye for. Nature. 421: 586. https://doi.org/10.1038/421586a.
13 Grätzel, M. (2001). Photoelectrochemical cells. Nature. 414: 338. https://doi.org/10.1038/35104607.
14 Müllen, K. and Scherf, U. (2006). Organic light emitting devices: synthesis, properties and
applications. Weinheim: John Wiley & Sons.
15 Sun, S.-­S. and Dalton, L.R. (2016). Introduction to organic electronic and optoelectronic materials
and devices. New York: CRC Press.
16 Ma, B., Knowles, D.B., Brown, C.S., Murphy, D. and Thompson, M.E. (2004). Organic light
emitting materials and devices. US Patent 6, 687, 266 B1, filed 8 November 2002 and issued 3
February 2004.
17 Li, Z., Li, Z.R. and Meng, H. (2006). Organic light-­emitting materials and devices. New York:
CRC Press.
18 Chen, C.-­T., Wagner, H. and Still, W.C. (1998). Fluorescent, sequence-­selective peptide detection by
synthetic small molecules. Science. 279: 851–853. https://doi.org/10.1126/science.279.5352.851.
19 Nilsson, D., Kugler, T., Svensson, P.-­O., and Berggren, M. (2002). An all-­organic sensor-­transistor
based on a novel electrochemical transducer concept printed electrochemical sensors on paper.
Sens. Actuators B Chem. 86: 193–197. https://doi.org/10.1016/S0925-­4005(02)00170-­3.
20 Alford, R., Simpson, H.M., Duberman, J., Hill, G.C., Ogawa, M., Regino, C., Kobayashi, H. and
Choyke, P.L. (2009). Toxicity of organic fluorophores used in molecular imaging: literature review
Mol. Imag. 8: 7290. https://doi.org/10.2310/7290.2009.00031.
21 Guo, Z., Park, S., Yoon, J. and Shin, I. (2014). Recent progress in the development of near-­infrared
fluorescent probes for bioimaging applications. Chem. Soc. Rev. 43: 16–29. https://doi.org/10.1039/
C3CS60271K.
22 Gaich, T. and Baran, P.S. (2010). Aiming for the ideal synthesis. J. Org. Chem. 75: 4657–4673.
https://doi.org/10.1021/jo1006812.
23 Posner, G.H. (1986). Multicomponent one-­pot annulations forming 3 to 6 bonds. Chem. Rev. 86:
831–844. https://doi.org/10.1021/cr00075a007.
24 Zhu, J., Wang, Q. and Wang, M. (2015) Multicomponent reactions in organic synthesis. Weinheim:
John Wiley & Sons.
25 Levi, L. and Müller, T.J.J. (2016). Multicomponent syntheses of fluorophores initiated by metal
catalysis. Eur. J. Org. Chem. 2016: 2902–2918. https://doi.org/10.1002/ejoc.201600409.
26 Müller, T.J.J. (2014). Science of synthesis series. Stuttgart: Georg Thieme Verlag KG.
27 Ugi, I., Dömling, A. and Hörl, W. (1994). Multicomponent reactions in organic chemistry.
Endeavour 18: 115–122. https://doi.org/10.1016/S0160-­9327(05)80086-­9.
28 Dömling, A. and Ugi, I. (2000). Multicomponent reactions with isocyanides. Angew. Chem. Int. Ed. 39:
3168–3210. https://doi.org/10.1002/1521-­3773(20000915)39:18<3168::AID-­ANIE3168>3.0.CO;2-­U.
29 Dömling, A. (2006). Recent developments in isocyanide based multicomponent reactions in
applied chemistry. Chem. Rev. 106; 17–89. https://doi.org/10.1021/cr0505728.
30 Ugi, I., Dömling, A. and Werner, B. (2000). Since 1995 the new chemistry of multicomponent
reactions and their libraries, including their heterocyclic chemistry. J. Heterocyclic. Chem. 37:
647–658. https://doi.org/10.1002/jhet.5570370322.
31 Zhu, J. and Bienaymé, H. (2006). Multicomponent reactions. Weinheim: John Wiley & Sons.
32 Ruijter, E., Scheffelaar, R. and Orru, R.V. (2011). Multicomponent reaction design in the quest for
molecular complexity and diversity. Angew. Chem. Int. Ed. 50: 6234–6246. https://doi.org/10.1002/
anie.201006515.
478 15  Multicomponent Reactions as Synthetic Design Tools of AIE and Emission Solvatochromic Quinoxalines

33 Orru, R.V.A. and de Greef, M. (2003). Recent advances in solution-­phase multicomponent


methodology for the synthesis of heterocyclic compounds. Synthesis. 2003: 1471–1499. https://doi.
org/10.1055/s-­2003-­40507.
34 Tietze, L.F. (1996). Domino reactions in organic synthesis. Chem. Rev. 96: 115–136. https://doi.
org/10.1021/cr950027e.
35 Tietze, L.F. and Gordon Brasche, K.M. (2007). Domino reactions in organic synthesis. Chem. Listy.
101: 185–187.
36 Tietze, L.F. (2014). Domino reactions: concepts for efficient organic synthesis. Weinheim: John
Wiley & Sons.
37 Tietze, L.F. and Beifuss, U. (1993). Sequential transformations in organic chemistry: a synthetic
strategy with a future. Angew. Chem. Int. Ed. 32: 131–163. https://doi.org/10.1002/
anie.199301313.
38 Bienayme, H., Hulme, C., Oddon, G. and Schmitt, P. (2000). Maximizing synthetic efficiency:
multi-­component transformations lead the way. Chem. Eur. J. 6: 33213329. https://doi.org/10.100
2/1521-­3765(20000915)6:18<3321::AID-­CHEM3321>3.0.CO;2-­A.
39 Levi, L. and Müller, T.J.J. (2016). Multicomponent syntheses of functional chromophores. Chem.
Soc. Rev. 45: 2825–2846. https://doi.org/10.1039/C5CS00805K.
40 Zhang, W. and Cue, B.W. (2018). Green techniques for organic synthesis and medicinal chemistry.
Weinheim: John Wiley & Sons.
41 Weber, L. (2002). The application of multi-­component reactions in drug discovery. Curr. Med.
Chem. 9: 2085–2093. https://doi.org/10.2174/0929867023368719.
42 Weber, L., Illgen, K. and Almstetter, M. (1999). Discovery of new multi component reactions with
combinatorial methods. Synlett. 1999: 366–374. https://doi.org/10.1055/s-­1999-­2612.
43 Müller, T.J.J. and D’Souza, D.M. (2008). Diversity-­oriented syntheses of functional π-­systems by
multicomponent and domino reactions. Pure Appl. Chem. 80: 609–620. https://doi.org/10.1351/
pac200880030609.
44 Müller, T.J.J. (2012). Synthesis of carbo-­and heterocycles via voupling-­izoerisation reactions.
Synthesis. 44: 159–174. https://doi.org/10.1055/s-­0031-­1289636.
45 Hong, Y., Lam, J.W. and Tang, B.Z. (2011). Aggregation-­induced emission. Chem. Soc. Rev. 40:
5361–5388. https://doi.org/10.1039/C1CS15113D.
46 Hong, Y., Lam, J.W. and Tang, B.Z. (2009). Aggregation-­induced emission: phenomenon,
mechanism and applications. Chem. Commun. 2009: 4332–4353. https://doi.org/10.1039/
B904665H.
47 Mei, J., Leung, N.L., Kwok, R.T., Lam, J.W. and Tang, B.Z. (2015). Aggregation-­induced emission:
together we shine, united we soar! Chem. Rev. 115: 11718–11940. https://doi.org/10.1021/acs.
chemrev.5b00263.
48 Hu, R., Leung, N.L. and Tang, B.Z. (2014). AIE macromolecules: syntheses, structures and
functionalities. Chem. Soc. Rev. 43: 4494–4562. https://doi.org/10.1039/C4CS00044G.
49 An, B.-­K., Kwon, S.-­K., Jung, S.-­D. and Park, S.Y. (2002). Enhanced emission and ist switching in
fluorescent organic nanoparticles. J. Am. Chem. Soc. 124: 14410–14415. https://doi.org/10.1021/
ja0269082.
50 Luo, J., Xie, Z., Lam, J.W., Cheng, L., Chen, H., Qiu, C., Kwok, H.S., Zhan, X., Liu, Y. and Zhu,
D. (2001). Aggregation-­induced emission of 1-­methyl-­1,2,3,4,5-­pentaphenylsilole. Chem. Commun.
2001: 1740–1741. https://doi.org/10.1039/B105159H.
51 Huang, J., Yang, X., Wang, J., Zhong, C., Wang, L., Qin, J. and Li, Z. (2012). New tetraphenylethene-­
based efficient blue luminophors: aggregation induced emission and partially controllable
emittiong color. J. Mater. Chem. 22: 2478–2484. https://doi.org/10.1039/C1JM14054J.
  ­Reference 479

52 Zhan, Y., Yang, Z., Tan, J., Qiu, Z., Mao, Y., He, J., Yang, Q., Ji, S., Cai, N. and Huo, Y. (2020).
Synthesis, aggregation-­induced emission (AIE) and electroluminescence of carbazole-­benzoyl
substituted tetraphenylethene derivatives. Dyes Pigm. 173: 107898. https://doi.org/10.1016/j.
dyepig.2019.107898.
53 Bhosale, R. S., Aljabri, M., La, D. D., Bhosale, S.V., Jones, L.A. and Bhosale, S.V. (2019) in Principles
and Applications of Aggregation-­Induced Emission. Basel: Springer, 223–264.
54 Merkt, F.K. and Müller, T.J.J. (2018). Solid state and aggregation induced emissive chromophores
by multi-­component syntheses. Isr. J. Chem. 58: 889–900. https://doi.org/10.1002/ijch.201800058.
55 Pereira, J.A., Pessoa, A.M., Cordeiro, M.N.D., Fernandes, R., Prudêncio, C., Noronha, J.P. and
Vieira, M. (2015). Quinoxaline, its derivatives and applications: a state of the art review. Eur.
J. Med. Chem. 97: 664–672. https://doi.org/10.1016/j.ejmech.2014.06.058.
56 Ajani, O.O. (2014). Present status of quinoxaline motifs: excellent pathfinders in therapeuthic
medicines. Eur. J. Med. Chem. 85: 688–715. https://doi.org/10.1016/j.ejmech.2014.08.034.
57 Montana, M., Mathias, F., Terme, T. and Vanelle, P. (2018). Antitumoral activity of quinoxaline
derivatives: a systematic review. Eur. J. Med. Chem. 163: 136–147. https://doi.org/10.1016/j.
ejmech.2018.11.059.
58 Sharma, P.C., Kaur, G., Pahwa, R., Sharma, A. and Rajak, H. (2011). Quinazolinone analogs as
potential therapeutic agents. Curr. Med. Chem. 18: 4786–4812. https://doi.
org/10.2174/092986711797535326.
59 Qi, C., Jiang, H., Huang, L., Chen, Z. and Chen, H. (2011). DABCO-­catalyzed oxidation of
deoxybenzoins to benzils with air and one-­pot synthesis of quinoxalines. Synthesis. 2011: 387–396.
https://doi.org/10.1055/s-­0030-­1258375.
60 Niesobski, P., Martínez, I.S., Kustosz, S. and Müller, T.J.J. (2019). Sequentially Pd/Cu-­catalyzed
alkynylation-­oxidation synthesis of 1,2-­diketones and consecutive one-­pot generation of
quinoxalines. Eur. J. Org. Chem. 2019: 5214–5218. https://doi.org/10.1002/ejoc.201900783.
61 Shi, S., Wang, T., Yang, W., Rudolph, M. and Hashmi, A.S.K. (2013). Gold-­catalyzed synthesis of
glyoxals by oxidation of terminal alkynes: one-­pot synthesis of quinoxalines. Chem. Eur. J. 19:
6576–6580. https://doi.org/10.1002/chem.201300518.
62 Chandrasekhar, S., Reddy, N.K. and Kumar, V.P. (2010). Oxidation of alkynes using PdCl2/
CuCl2 in PEG as a recylable catalytic system: one-­pot synthesis of quinoxalines. Tetrahedron Lett.
51: 3623–3625. https://doi.org/10.1016/j.tetlet.2010.05.006.
63 Das, B., Venkateswarlu, K., Suneel, K. and Majhi, A. (2007). An efficient and convenient protocol
for the synthesis of quinoxalines and dihydropyrazines via cyclization-­oxidation processes using
HClO4·SiO2 as a heterogenous recyclable catalyst. Tetrahedron Lett. 48: 5371–5374. https://doi.
org/10.1016/j.tetlet.2007.06.036.
64 Azuaje, J., El Maatougui, A., García-­Mera, X. and Sotelo, E. (2014). Ugi-­based approaches to
quinoxaline libraries. ACS Comb. Sci. 16: 403–411. https://doi.org/10.1021/co500036n.
65 Krasavin, M., Shkavrov, S., Parchinsky, V. and Bukhryakov, K. (2009). Imidazo[1,2-­a]quinoxalines
accessed via two sequential isocyanide-­based multicomponent reactions. J. Org. Chem. 74:
2627–2629. https://doi.org/10.1021/jo900050k.
66 Kalinski, C., Umkehrer, M., Ross, G., Kolb, J., Burdack, C. and Hiller, W. (2006). Highly substituted
indol-­2-­ones, quinoxalin-­2-­ones and benzodiazepin-­2,5-­diones via a new Ugi (4CR)-­Pd assisted
N-­aryl amidation strategy. Tetrahedron lett. 47: 3423–3426. https://doi.org/10.1016/j.
tetlet.2006.03.069.
67 Krasavin, M. and Parchinsky, V. (2008). Expedient entry into 1,4-­dihydroquinoxalines and
quinoxalines via a novel variant of isocyanide-­based MCR. Synlett. 2008: 645–648. https://doi.
org/10.1055/s-­2008-­1032106.
480 15  Multicomponent Reactions as Synthetic Design Tools of AIE and Emission Solvatochromic Quinoxalines

68 Heravi, M.M., Baghernejad, B. and Oskooie, H.A. (2009). A novel three-­component reaction for the
synthesus of N-­cyclohexyl-­3-­aryl-­quinoxaline-­2-­amines. Tetrahedron Lett. 50: 767–769. https://doi.
org/10.1016/j.tetlet.2008.11.123.
69 De Moliner, F. and Hulme, C. (2012). A Van Leusen deprotection-­cyclization strategy as a fast entry
into two imidazoquinoxaline families. Tetrahedron lett. 53: 5787–5790. https://doi.org/10.1016/j.
tetlet.2012.08.072.
70 Gunawan, S., Nichol, G. and Hulme, C. (2012). Concise route to a series of novel 3-­(tetrazol-­5-­yl)
quinoxalin-­2 (1H)-­ones. Tetrahedron lett. 53: 1664–1667. https://doi.org/10.1016/j.tetlet.2012.01.080.
71 Ayaz, M., Dietrich, J. and Hulme, C. (2011). A novel route to synthesize libraries of quinoxalines
via Petasis methodology in two synthetic operations. Tetrahedron lett. 52: 4821–4823. https://doi.
org/10.1016/j.tetlet.2011.06.115.
72 Lian, M., Li, Q., Zhu, Y., Yin, G. and Wu, A. (2012). Logic design and synthesis of quinoxalines via
the integration of iodination/oxidation/cyclization sequences from ketones and 1,2-­diamines.
Tetrahedron. 68: 9598–9605. https://doi.org/10.1016/j.tet.2012.09.056.
73 Neochoritis, C., Stephanidou-­Stephanatou, J. and Tsoleridis, C.A. (2009). Heterocyclizations via
TosMIC-­based multicomponent reactions: a new approach to one-­pot facile synthesis of substituted
quinoxaline derivatives. Synlett. 2009: 302–305. https://doi.org/10.1055/s-­0028-­1087518.
74 Nguyen, T.B., Retailleau, P. and Al-­Mourabit, A. (2013). A simple and straightforward approach to
quinoxalines by iron/sulfur-­catalyzed redox condensation of o-­nitroanilines and phenethylamines.
Org. lett. 15: 5238–5241. https://doi.org/10.1021/ol402435c.
75 Bhosale, R.S., Sarda, S.R., Ardhapure, S.S., Jadhav, W.N., Bhusare, S.R. and Pawar, R.P. (2005). An
efficient protocol for the synthesis of quinoxaline derivatives at room temperature using molecular
iodine as the catalyst. Tetrahedron lett. 46: 7183–7186. https://doi.org/10.1016/j.tetlet.2005.08.080.
76 Viswanadham, K.D.R., Reddy, M.P., Sathyanarayana, P., Ravi, O., Kant, R. and Bathula, S.R. (2014).
Iodine-­mediated oxidative annulation for the one-­pot synthesis of pyrazines and quinoxalines
using a multipathway coupled domino stategy. Chem. Commun. 50: 13517–13520. https://doi.
org/10.1039/C4CC05844E.
77 Meshram, H., Kumar, G.S., Ramesh, P. and Reddy, B.C. (2010). A mild and convenient synthesis of
quinoxalines via cyclization-­oxidation process using DABCO as catalyst. Tetrahedron lett. 51:
2580–85. https://doi.org/10.1016/j.tetlet.2010.01.107.
78 Martin, L.J., Marzinzik, A.L., Ley, S.V. and Baxendale, I.R. (2011). Safe and reliable synthesis of
diazoketones and quinoxalines in a continuous flow reactor. Org. lett. 13: 320–323. https://doi.
org/10.1021/ol1027927.
79 Ayaz, M., Martinez-­Ariza, G. and Hulme, C. (2014). A robust protocol for the synthesis of
quinoxalines and 5H-­benzo[e][1,4]diazepines via acidless Ugi reaction. Synlett, 25: 1680–1684.
https://doi.org/10.1055/s-­0033-­1339135.
80 Zhang, Z. and Jiang, X. (2014). Oxidative coupling of terminal alkyne with α-­hydroxy ketone: an
expedient approach toward ynediones. Org. Lett. 16: 4400–4403. https://doi.org/10.1021/ol502298a.
81 Liu, Y., Liu, M., Guo, S., Tu, H., Zhou, Y. and Gao, H. (2006). Gold-­catalyzed highly efficient access
to 3(2H)-­furanones from 2-­oxo-­3-­butynoates and related compounds. Org. Lett. 8: 3445–3448.
https://doi.org/10.1021/ol061059z.
82 Gers-­Panther, C.F. and Müller, T.J.J. (2016). Multicomponent syntheses of heterocycles initiated by
catalytic generation of ynones and ynediones, in Advances in Heterocyclic Chemistry (ed. Eric.
F. V. Scriven, Christopher A. Ramsden), 120, 67–98. Amsterdam: Elsevier.
83 Boersch, C., Merkul, E. and Müller, T.J.J. (2011). Catalytic syntheses of N-­heterocyclic ynones and
ynediones by in situ activation of carboxylic acids with oxalyl chloride, Angew. Chem. Int. Ed. 50:
10448–10452. https://doi.org/10.1002/anie.201103296.
  ­Reference 481

84 Merkul, E., Dohe, J., Gers, C.F., Rominger, F. and Müller, T.J.J. (2011). Three-­component synthesis
of ynediones by a glyoxylation/stephens-­castro coupling sequence. Angew. Chem. Int. Ed. 123:
3023–3026. https://doi.org/10.1002/anie.201007194.
85 Merkul, E., Oeser, T. and Müller, T.J.J. (2009). Consecutive three-­component synthesis of ynones
by decarbonylative sonogashira coupling. Chem. Eur. J. 15: 5006–5011. https://doi.org/10.1002/
chem.200900119.
86 Gers-­Panther, C.F., Fischer, H., Nordmann, J., Seiler, T., Behnke, T., Würth, C., Frank, W., Resch-­
Genger, U. and Müller, T.J.J. (2017). Four-­and five-­component syntheses and photophysical
properties of emission solvatochromic 3-­aminovinylquinoxalines. J. Org. Chem. 82: 567–578.
https://doi.org/10.1021/acs.joc.6b02581.
87 Gers, C.F., Nordmann, J., Kumru, C., Frank, W. and Müller, T.J.J. (2014). Solvatochromic
fluorescent 2-­substituted 3-­ethynyl quinoxalines: four-­component synthesis, photophysical
properties, and electronic structure. J. Org. Chem. 79: 3296–3310. https://doi.org/10.1021/
jo4025978.
88 Merkt, F.K., Höwedes, S. P., Gers-­Panther, C.F., Gruber, I., Janiak, C. and Müller, T.J.J. (2018).
Three-­component activation/alkynylation/cyclocondensation (AACC) synthesis of enhanced
emission solvatochromic 3-­ethynylquinoxalines. Chem. Eur. J. 24: 8114–8125. https://doi.
org/10.1002/chem.201800079.
89 Gers, C.F. (2014). Synthese, reaktivität und photophysikalische eigenschaften von
3-­ethinylchinoxalinen. Dissertation. Heinrich-­Heine University Düsseldorf.
90 Merkt, F.K. and Müller, T.J.J. (2018). Synthesis and electronic properties of expanded 5-­(hetero)
aryl thien-­2-­yl substituted 3-­ethynyl quinoxalines with AIE charateristics. Sci. China Chem. 61:
909–924. https://doi.org/10.1007/s11426-­018-­9295-­4.
91 Merkt, F.K., Pieper, K., Klopotowski, M., Janiak, C. and Müller, T.J. (2019). Sequential Cu-­
catalyzed four-­and five component syntheses of luminescent 3-­triazolylquinoxalines. Chem. Eur.
J. 25: 9447–9455. https://doi.org/10.1002/chem.201900277.
92 Lakowicz, J.R. (2013). Principles of fluorescence spectroscopy. New York: Springer Science &
Business Media.
93 Umeda, M., Shimada, T., Aruga, T., Niimi, T. and Sasaki, M. (1993). Energy gap dependence of the
photocarrier generation efficiency in layered organic photoreceptors. J. Phys. Chem. 97: 8531–8534.
https://doi.org/0022-­3654/93/2097-­8531 $04.00/0.
94 Hansch, C., Leo, A. and Taft, R. (1991). A survey of Hammett substituent constants and resonance
and field parameters. Chem. Rev. 91: 165–195. https://doi.org/10.1021/cr00002a004.
95 Mataga, N., Kaifu, Y. and Koizumi, M. (1956). Solvent effects upon fluorescence spectra and the
dipolemoments of excited molecules. Bull. Chem. Soc. Jpn. 29: 465–470. https://doi.org/10.1246/
bcsj.29.465.
96 Kudo, K., Momotake, A., Kanna, Y., Nishimura, Y. and Arai, T. (2011). Development of a
quinoxaline-­based fluorescent probe for quantitative estimation of protein bindung site polarity.
Chem. Commun. 47: 3867–3869. https://doi.org/10.1039/C1CC10183H.
97 Reichardt, C. (1994). Solvatochromic dyes as solvent polarity indicators. Chem. Rev. 94: 2319–2358.
https://doi.org/10.1021/cr00032a005.
98 Achelle, S., Baudequin, C. and Plé, N. (2013). Luminescent materials incorporating pyrazine or
quinoxalune moieties. Dyes Pigm. 98: 575–600. https://doi.org/10.1016/j.dyepig.2013.03.030.
99 Kalinin, A.A., Sharipova, S.M., Burganov, T.I., Dudkina, Y.B., Khamatgalimov, A.R., Katsyuba,
S.A., Budnikova, Y.H. and Balakina, M.Y. (2017). Push–pull isomeric chromophores with vinyl-­
and divinylquinoxaline-­2-­one units as π-­electron bridge: synthesis, photophysical, thermal and
electro-­chemical properties. Dyes Pigm. 146: 82–91. https://doi.org/10.1016/j.dyepig.2017.06.062.
482 15  Multicomponent Reactions as Synthetic Design Tools of AIE and Emission Solvatochromic Quinoxalines

100 Kalinin, A.A., Sharipova, S.M., Burganov, T.I., Levitskaya, A.I., Dudkina, Y.B., Khamatgalimov,
A.R., Katsyuba, S.A., Budnikova, Y.H. and Balakina, M.Y. (2018). High thermally stable D-­π-­A
chromophores with quinoxaline moieties in the conjugated bridge: synthesis, DFT calculations
and physical properties. Dyes Pigm. 156: 175–184. https://doi.org/10.1016/j.dyepig.2018.04.002.
101 Gupta, S. and Milton, M.D. (2019). Design and synthesis of novel V-­shaped AIEE active
quinoxalines for acidochromic applications. Dyes Pigm. 165: 474–487. https://doi.org/10.1016/j.
dyepig.2019.02.038.
102 Black, H.T., Pelse, I., Wolfe, R.M. and Reynolds, J.R. (2016). Halochromism and protonation-­
induced assembly of a benzo[g]indolo[2,3-­b]quinoxaline derivative. Chem. Commun. 52:
12877–12880. https://doi.org/10.1039/C6CC06443D.
103 Mazumdar, P., Maity, S., Shyamal, M., Das, D., Sahoo, G.P. and Misra, A. (2016). Proton triggered
emission and selective sensing of picric acid by the fluorescent aggregates of 6,7-­dimethyl-­2,3-­bis-­
(2-­pyridyl)-­quinoxaline. Phys. Chem. Chem. Phys. 18: 7055–7067. https://doi.org/10.1039/
C5CP05827A.
104 Hilczer, M., Traytak, S. and Tachiya, M. (2001). Electric field effects on fluorescence quenching
due to electron transfer. J. Chem. Phys. 115: 11249–11253. https://doi.org/10.1063/1.1421364.
105 Justin Thomas, K., Velusamy, M., Lin, J.T., Chuen, C.-­H. and Tao, Y.-­T. (2005). Chromophore-­
labeled quinoxaline derivatives as efficient electroluminescent materials. Chem. Mater. 17:
1860–1866. https://doi.org/10.1021/cm047705a.
106 Mahadik, S.S., Chacko, S. and Kamble, R.M. (2019). 2,3-­Di (thiophen-­2-­yl) quinoxaline amine
derivatives: yellow-­blue fluorescent materials for applications in organic electronics. Chem Select.
4: 10021–10032. https://doi.org/10.1002/slct.201902109.
107 Marini, A., Munoz-­Losa, A., Biancardi, A. and Mennucci, B. (2010). What is solvatochromism?.
J. Phys. Chem. B 114: 17128–17135. https://doi.org/10.1021/jp1097487.
108 Zhang, Z., Dai, Z. and Jiang, X. (2015). Copper-­catalyzed aerobic oxidative dicarbonylation of
indoles toward solvatochromic fluorescent indole-­substituted quinoxalines. Asian J. Org. Chem. 4:
1370–1374. https://doi.org/10.1002/ajoc.201500332.
109 Ekbote, A., Jadhav, T. and Misra, R. (2017). T-­shaped donor–acceptor–donor type
tetraphenylethene substituted quinoxaline derivatives: aggregation-­induced emission and
mechanochromism. New J. Chem. 41: 9346–9353. https://doi.org/10.1039/C7NJ01531C.
110 Lu, X., Fan, S., Wu, J., Jia, X., Wang, Z.-­S. and Zhou, G. (2014). Controlling the charge transfer in
D–A–D chromophores based on pyrazine derivatives. J. Org. Chem. 79: 6480––6489. https://doi.
org/10.1021/jo500856k.
111 Chen, Y., Ling, Y., Ding, L., Xiang, C. and Zhou, G. (2016). Quinoxaline-­based cross-­conjugated
luminophores: charge transfer, piezofluorochromic, and sensing properties. J. Mater. Chem. C. 4:
8496–8505. https://doi.org/10.1039/C6TC02945K.
112 Schaffroth, M., Lindner, B.D., Vasilenko, V., Rominger, F. and Bunz, U.H. (2013). Alkynylated
diazadioxaacenes: syntheses and properties. J. Org. Chem. 78: 3142–3150. https://doi.org/10.1021/
jo400092r.
113 Hirayama, T., Yamasaki, S., Ameku, H., Ishi-­i, T., Thiemann, T. and Mataka, S. (2005).
Fluorescent solvatochromism of bi-­polar N,N-­diphenylaminoaryl-­substituted
hexaazatriphenylenes, tetraazaphenanthrene, and quinoxalines. Dyes Pigm. 67: 105–110. https://
doi.org/10.1016/j.dyepig.2004.09.023.
114 Nirmalananthan, N., Behnke, T., Hoffmann, K., Kage, D., Gers-­Panther, C.F., Frank, W., Müller,
T.J.J. and Resch-­Genger, U. (2018). Crystallization and aggregation-­induced emission in a series
of pyrrolidinylvinylquinoxaline derivatives. J. Phys. Chem. C. 122: 11119–11127. https://doi.
org/10.1021/acs.jpcc.8b01425.
  ­Reference 483

115 Kumar, S., Chawla, G. and Bansal, A.K. (2008). Role of additives like polymers and surfactants in
the crystallization of mebendazole. Yakugaku Zasshi. 128: 281–289. https://doi.org/10.1248/
yakushi.128.281.
116 Canselier, J. (1993). The effects of surfactants on crystallization phenomena. J. Disp. Sci. Technol.
14: 625–644. https://doi.org/110.1080/01932699308943435.
117 Rodríguez-­Hornedo, N. and Murphy, D. (2004). Surfactant-­facilitated crystallization of dihydrate
cabamazepine during dissolution of anhydrous polymorph. J. Pharm. Sci. 93: 449–460. https://
doi.org/10.1002/jps.10496.
118 Wei, H., Shen, Q., Zhao, Y., Zhou, Y., Wang, D. and Xu, D. (2004). Effect of anionic surfactant–
polymer complexes on the crystallization of calcium carbonate. J. Cryst. Growth 264: 424–429.
https://doi.org/10.1016/j.jcrysgro.2004.01.001.
119 Behnke, T., Würth, C., Hoffmann, K., Hübner, M., Panne, U. and Resch-­Genger, U. (2011).
Encapsulation of hydrophobic dyes in polystyrene micro-­and nanoparticles via swelling
procedures. J. Fluoresc. 21: 937–944. https://doi.org/10.1007/s10895-­010-­0632-­2.
120 Kanekar, D.N., Chacko, S. and Kamble, R.M. (2019). Quinoxaline based amines as blue-­orange
emitters: effect of modulating donor system on optoelectrochemical and theoretical properties.
Dyes Pigm. 167: 36–50. https://doi.org/10.1016/j.dyepig.2019.04.005.
121 Yang, W., Li, C., Zhang, M., Zhou, W., Xue, R., Liu, H. and Li, Y. (2016). Aggregation-­induced
emission and intermolecular charge transfer effect in triphenylamine fluorophores containing
diphenylhydrazone structures. Phys. Chem. Chem. Phys. 18: 28052–28060. https://doi.org/
10.1039/C6CP04755F.
122 Liu, W., Ying, S., Zhang, Q., Ye, S., Guo, R., Ma, D. and Wang, L. (2018). New multifunctional
aggregation-­induced emission fluorophores for reversible piezofluorochromic and nondoped
sky-­blue organic light-­emitting diodes. Dyes Pigm. 158: 204–212. https://doi.org/10.1016/j.
dyepig.2018.05.030.
123 Li, H., Chi, Z., Zhang, X., Xu, B., Liu, S., Zhang, Y. and Xu, J. (2011). New thermally stable
aggregation-­induced emission enhancement compounds for non-­doped red organic light-­emitting
diodes. Chem. Commun. 47: 11273–11275. https://doi.org/10.1039/C1CC14642D.
124 Ghosh, S., Pal, S., Rajamanickam, S., Shome, R., Mohanta, P.R., Ghosh, S.S. and Patel,
B.K. (2019). Access to multifunctional AEEgens via Ru (II)-­catalyzed quinoxaline-­directed
oxidative annulation. ACS Omega. 4: 5565–5577. https://doi.org/10.1021/acsomega.9b00274.
125 Cui, M., Li, W., Wang, L., Gong, L., Tang, H. and Cao, D. (2019). Twisted intramolecular charge
transfer and aggregation-­enhanced emission characteristics based quinoxaline luminogen:
photophysical properties and a turn-­on fluorescent probe for glutathione. J. Mater. Chem. C. 7:
3779–3786. https://doi.org/10.1039/C8TC05360J.
126 Dong, Y., Lam, J.W.Y., Qin, A., Sun, J., Liu, J., Li, Z., Sun, J., Sung, H.H., Williams, I. D. and Kwok,
H.S. (2007). Aggregation-­induced and crystallization-­enhanced emissions of 1,2-­diphenyl-­3,4-­
bis(diphenylmethylene)-­1-­cyclobutene. Chem. Commun. 2007: 3255–3257. https://doi.
org/10.1039/B704794K.
127 Zheng, X., Peng, Q., Zhu, L., Xie, Y., Huang, X. and Shuai, Z. (2016). Unraveling the aggregation
effect on amorphous phase AIE luminogens: a computational study. Nanoscale. 8: 15173–15180.
https://doi.org/10.1039/C6NR03599J.
128 Triantou, D., Soulis, S., Asaftei, C. and Janietz, S. (2015). Effect of the acceptor moiety on the
electrochemical and electrochromic properties of donor–acceptor–donor polymer films. Int.
J. Electrochem. Sci. 10: 3458–3477. https://doi.org/10.1088/1741-­4326/aa6ec1.
129 Moshkina, T.N., Nosova, E.V., Lipunova, G.N., Valova, M.S. and Charushin, V.N. (2018). New
2,3-­bis(5-­arylthiophen-­2-­yl)quinoxaline derivatives: synthesis and photophysical properties.
Asian J. Org. Chem. 7: 1080–1084. https://doi.org/10.1002/ajoc.201800217.
484 15  Multicomponent Reactions as Synthetic Design Tools of AIE and Emission Solvatochromic Quinoxalines

130 Kwon, O., Barlow, S., Odom, S.A., Beverina, L., Thompson, N.J., Zojer, E., Brédas, J.-­L. and
Marder, S.R. (2005). Aromatic amines: a comparison of electron-­donor strengths. J. Phys. Chem.
A. 109: 9346–9352. https://doi.org/10.1021/jp054334s.
131 Huang, J., Sun, N., Yang, J., Tang, R., Li, Q., Ma, D. and Li, Z. (2014). Blue aggregation-­induced
emission luminogens: high external quantum efficiencies up to 3.99% in LED device, and
restriction of the conjugation length through rational molecular design. Adv. Funct. Mater. 24:
7645–7654. https://doi.org/10.1002/adfm.201401867.
132 Qin, W., Yang, Z., Jiang, Y., Lam, J.W., Liang, G., Kwok, H.S. and Tang, B.Z. (2015). Construction
of efficient deep blue aggregation-­induced emission luminogen from triphenylethene for
nondoped organic light-­emitting diodes. Chem. Mater. 27: 3892–3901. https://doi.org/10.1021/acs.
chemmater.5b00568.
133 Liu, J., Evrard, M., Cai, X., Feng, G., Tomczak, N., Ng, L.G. and Liu, B. (2018). Organic
nanoparticles with ultrahigh quantum yield and aggregation-­induced emission characteristics for
cellular imaging and real-­time two-­photon lung vasculature imaging. J. Mater. Chem. B. 6:
2630–2636. https://doi.org/ 10.1039/C8TB00386F.
134 Qin, W., Li, K., Feng, G., Li, M., Yang, Z., Liu, B. and Tang, B.Z. (2014). Bright and photostable
organic fluorescent dots with aggregation-­induced emission characteristics for noninvasive
long-­term cell imaging. Adv. Funct. Mater. 24: 635–643. https://doi.org/10.1002/adfm.201302114.
135 Wang, S., Liu, J., Goh, C.C., Ng, L.G. and Liu, B. (2019). NIR-­II-­excited intravital two-­photon
microscopy distinguishes deep cerebral and tumor vasculatures with an ultrabright NIR-­I AIE
luminogen. Adv. Mater. 31: 1904447. https://doi.org/10.1002/adma.201904447.
136 Fisher, A. Lusi, M.H. and Egerton, J.R. (1977). Anthelmintic dihydroquinoxalino
[2,3-­b]quinoxalines. J. Pharm. Sci. 66: 1349–1352. https://doi.org/10.1002/jps.2600660942.
137 Miura, Y. and Yoshioka, N. (2018). 5,12-­Diacetyl-­5,12-­dihydroquinoxalino[2,3-­b]quinoxalines:
solid-­state fluorescence, AIE properties, and orbital switching by substituent effect. Chem. Asian
J. 13: 1683–1687. https://doi.org/10.1002/asia.201800502.
138 Chen, J., Peng, H., Law, C.C., Dong, Y., Lam, J.W., Williams, I.D. and Tang, B.Z. (2003).
Hyperbranched poly(phenylenesilolene)s: synthesis, thermal stability, electronic conjugation,
optical power limiting, and cooling-­enhanced light emission. Macromolecules 36: 4319–4327.
https://doi.org/10.1021/ma034012r.
139 Tong, H., Dong, Y., Hong, Y., Häussler, M., Lam, J.W., Sung, H.H.-­Y., Yu, X., Sun, J., Williams,
I.D. and Kwok, H.S. (2007). Aggregation-­induced emission: effects of molecular structure,
solid-­state conformation, and morphological packing arrangement on light-­emitting behaviors of
diphenyldibenzofulvene derivatives. J. Phys. Chem. C. 111: 2287–2294. https://doi.org/10.1021/
jp0630828.
140 He, Y., Li, Y., Su, H., Si, Y., Liu, Y., Peng, Q., He, J., Hou, H. and Li, K. (2019). An o-­phthalimide-­
based multistimuli-­responsive aggregation-­induced emission (AIE) system. Mater. Chem. Front. 3:
50–56. https://doi.org/10.1021/jp0630828.
485

16

Aggregation-­induced Emission Luminogens with Both


High-­luminescence Efficiency and Charge Mobility
Ying Yu, Zheng Zhao, and Ben Zhong Tang
Shenzhen Institute of Aggregate Science and Technology, School of Science and Engineering, The Chinese University of Hong Kong,
Shenzhen, Guangdong, China

16.1 ­Introduction

Inorganic semiconductors such as silicon materials are approaching their theoretical limit in the
component size and device performance. Inorganic materials usually need high temperature to
process, have only finite types, and can just be applied to several substrates, which makes them
suffer the disadvantages of fragile, high cost, and severe pollution during the process  [1].
Furthermore, because they lack flexibility and are not biocompatible, they are difficult to be
applied in the flourishing bioelectronic area such as wearable flexible device. As a contrast, organic
semiconductors (OSCs) has attracted great attention and experienced a rapid development in the
past decades. In comparison with their inorganic counterparts, OSCs have tremendous advantages
such as light weight, flexibility, diversity, and solution processability, which enable their promising
applications in electronic and photonic devices. Furthermore, the excellent solution processability
endows them with room temperature low-­cost mass production. Their compatibility with flexible
solid supports as substrates, such as glass, plastic, metal oxide, and others, makes organic elec-
tronic devices amenable to diverse solution-­based fabrication methods, including interfacial
assembly, casting, spinning, layer-­by-­layer assembly, or roll-­to-­roll protocols such as inkjet and
screen printing [2, 3].
OSCs generally are some π-­expanded molecules that could be used for organic optoelectronics with
the capabilities of charge transport or light emission. According to the type of charge they transport,
they can be categorized as p-­type (hole transport), n-­type (electron transport), and ambipolar type
(with both hole and electron transport), while according to whether they can emit light or not, they
can be categorized as light-­emitting semiconductors or nonemissive ones. In addition, they can be
categorized as OSCs for organic field-­effect transistors (OFETs), OSCs for organic light-­emitting
diodes (OLEDs), OSCs for organic photovoltaics, and OSCs for thermoelectricity according to their
applications. Among the varied optoelectronic devices, OSCs are the active layer of the devices and
also the most significant components of the devices.
OLEDs have been successfully commercialized in lighting and flexible display. They have
superiorities in low power consumption, fast response time, wide viewing angle, and high reso-
lution. In particular, flexible screen displays based on OLEDs are widely thought as the most

Handbook of Aggregation-Induced Emission: Volume 1 Tutorial Lectures and Mechanism Studies, First Edition.
Edited by Youhong Tang and Ben Zhong Tang.
© 2022 John Wiley & Sons Ltd. Published 2022 by John Wiley & Sons Ltd.
486 16  Aggregation-­induced Emission Luminogens with Both High-­luminescence Efficiency and Charge Mobility

promising candidates of next generation of display  [4, 5]. Typical OLED display such as the
active-­matrix OLED display panels, where field-­effect transistors (FETs) are integrated with
OLED devices to control each OLED pixel on/off through controlling the current density.
Therefore, the integration of OLED and FET into one device is inspired by commercial interest
since it may simplify the device structure and processing procedure, thus reducing the device
cost. The devices integrated the modulation function of a transistor with light emission of an
OLED is called organic light-­emitting transistors (OLETs). The research of OLETs are of impor-
tance both in fundamental study and technological applications. They are excellent test systems
to investigate physical and electronic properties, such as charge injection, transport, and electro-
luminescence (EL) in OSCs. They are also promising for the development of highly integrated
organic optoelectronic devices, such as active-­matrix full-­color EL displays. Moreover, their
development can also lead to electrically driven organic lasers. Unfortunately, although the
device performance of single OLED and OFET have been well-­developed, respectively, the
development of OLET device is still rather lagged due to the challenges of both materials and
device fabrication [6–8].
Figure 16.1 shows the basic device structures of OLED, OFET, and OLET. In OLEDs, holes and
electrons are injected from anode and cathode, respectively, and recombined in the emitting layer
to form excitons that emit light (Figure 16.1a). Usually, additional layers such as electron transport
and hole transport layers are needed to balance the charge carriers to achieve better device perfor-
mance. External quantum efficiency is an important parameter to evaluate the electrical-­to-­light
energy conversion capacity, which is defined as the number of emitted photons per injected charge
carrier. In the OLEDs, charge carriers commonly only need to travel tens of nanometers to meet
the opposite charge carriers to recombine. Therefore, the EL materials do not require a high charge
carrier mobility (𝜇) of more than 10−4 to 10−2 cm2/V/s. In Figure 16.1b, the basic working principle
of OFET is shown by a hole-­transporting OFET. A typical OFET device contains OSC, organic or
inorganic insulator, and metal electrodes, which were used as the active channel materials, dielec-
tric layers and source, drain and gate electrodes, respectively. During the operation, charge carriers
induced by a particular gate bias will accumulate at the dielectric/semiconductor interface to form
a conductive channel between the source and drain contacts. After applying an appropriate source-­
drain voltage, a current flow is formed, which could be modulated with on/off function by gate
voltage [9]. In OFETs, charge carriers are required to travel a longer distance (a few hundred of
nanometers to a few micrometers) compared with that in OLEDs. To ensure the stable work of the
device, semiconductors with as high as possible charge mobility is favored. According to the charge
transporting type of the carriers in the semiconducting layer, OFET was usually classified as p-­type,
n-­type, and ambipolar type.

(a) (b) (c)


Cathode Organic semiconductor Organic semiconductor
ETL Source Drain Source Drain
Emitting layer Gate dielectric Gate dielectric
HTL Gate
Gate VGS
Anode VGS
VDS VDS
Transparent substrate A
A

Figure 16.1  Schematic representation of the device structures and the main optoelectronic processes
occurred in an OLED (a), OFET (b), and OLET (c). VGS = gate-source voltage, VDS = drain-source voltage
16.2 ­p-­Type OSCs 487

OLET device actually combines the switching functionality of a FET and the light emission
property of an OLED, which is more like one unique type of OFET using light-­emitting OSC as the
active layer (Figure 16.1c). Correspondingly, the requirements of OSCs for OLETs include both
high charge mobility and high-­luminescence efficiency. However, it is still a big challenge to
develop OSCs with both high charge mobility and quantum yield because strong π–π interactions
are usually favorable for affording high charge mobility, which however, were thought as a quench-
ing effect for the emission of luminogens. Some star molecular semiconductors such as rubrene
and pentacene generally show high mobility of up to 10–20 cm2/V/s but exhibit no emission in
their single crystal (SC) [10, 11]. Actually, many of the luminogens with planar and rigid structure
exhibit bright emission in their dilute solutions but dim emission after aggregation or in the solid
state. This is a commonly encountered phenomenon, named as concentration quenching effect or
aggregation caused quenching (ACQ) effect, that has been reported as early as 1954 by Föster [12].
Reasons proposed to explain the ACQ problem includes the formation of excimers, singlet fission,
and strong vibronic coupling between ground state and excited state [13, 14], all of which basically
originated from strong intermolecular π–π interactions in the solid state or crystal state. However,
the introduction of strong intermolecular π–π interactions is the commonly utilized strategy to
improve the charge carrier mobility. Therefore, it seems that there is a big conflict between the
high-­charge mobility and luminescent quantum yield [15, 16].
In 2001, Tang and coworkers found a kind of propeller-­shaped molecules that showed nonemis-
sion in solution state but much bright emission in the solid state, which is contrary to the tradi-
tional ACQ phenomenon  [17]. Thus, they coined this unconventional phenomenon as
aggregation-­induced emission (AIE). The highly emissive property of AIE luminogens (AIEgens)
in the solid state actually provides an excellent platform to develop new luminescent functional
materials and sheds new light on the molecular designs of organic electronic materials. On the
platform of AIE, all kinds of chemical and physical properties or behaviors such as chirality, self-­
assembly, liquid crystal, and stimuli-­responsive property could be integrated with AIE to develop
new functional optical materials, such as chiral luminogens, light-­emitting liquid crystals, and
stimuli-­responsive luminogens [18, 19]. Furthermore, AIE may also provide an excellent strategy
to overcome the conflict between high charge mobility and luminescent quantum yield to develop
high-­mobility light-­emitting materials. In this chapter, the exploration of materials with both high
charge mobility and luminescence efficiency in the solid state will be reviewed and summarized.
In particular, the role of AIE materials in endowing charge transporting materials with high lumi-
nescence efficiency would be highlighted. In terms of the charge carrier transported, they are also
divided into three categories: p-­type luminogens, n-­type luminogens, and ambipolar luminogens
(Figure 16.2). Since the charge transporting properties and luminescent properties correlate closely
with the crystal stacking, film morphology, and the energy levels of the semiconductor molecules,
the influence of these factors on the device performance thus will be mainly discussed.

16.2  ­p-­Type OSCs

p-­Type OSCs are some electron-­rich molecules with highest occupied molecular orbitals (HOMO)
energy level within a range of –5.5 to –5.0 eV. Most of the reported luminogens generally are p-­type.
Tetraphenylethylene (TPE) and its derivatives are the mostly investigated luminogens with aggre-
gate/solid emission, however, its charge transporting properties have scarcely been studied  [20].
Zhao and coworkers reported a TPE derivative of p-­BTPATPE by attaching two triphenylamine (TPA)
units to the TPE core (Figure  16.3a) and investigated its luminescence and charge transporting
488 16  Aggregation-­induced Emission Luminogens with Both High-­luminescence Efficiency and Charge Mobility

p-Type n-Type Ambipolar

N C10H21 C10H21 N
C10H21 Ar =
N
C8H17
O O C8H17 O C8H17
N N N ArAr
O O
O H
S
p-BTPATPE n
1-X
S X (Z)-oBCaPTPE
Chem. Commun. 2011, 47, 6924 H Adv. Mater. 2011, 23, 5430
O O O N
N N C8H17 O
O C8H17 O
C8H17
C10H21 C10H21
C10H21

PNV PNV-DT7
DPA Chem. Commun. 2017, 53, 4934 Macromolecules 2019, 52, 8332
Nat. Commun. 2015, 6, 10032 dNaAnt
J. Am. Chem. Soc. 2017, 139, 17261
C6H13

R O O S
N
HPVAnt O N O R R
S
Angew. Chem. Int. Ed. 2012, 51, 3837 S
R=
BDPV2T
S R Adv. Funct. Mater. 2018, 28, 1802454
O N O
R O N O
NC
27PyE TNDI-20: R = C20H41
TriPE-3PDI
Chem. Mater. 2017, 29, 3580 TNDI-6: R = C6H13
Adv.Funct. Master. 2018, 28, 1705609 CN
Chem. Asian J. 2014, 9, 3207

N β-CNDSB
N Chem. Commun. 2016, 52, 2370

1, 6-DTEP
Adv. Mater. 2020, 1907791

Figure 16.2  AIE materials for OLET application.

properties [21].p-­BTPATPE emitted at 509 nm (Figure 16.3b) with solid-­state fluorescence efficiency


close to unity. The twisted structure of p-­BTPATPE avoids the quenching factors resulted from strong
intermolecular π–π interactions and the restriction of intramolecular rotation in the solid state con-
tributed a lot to the high photo luminescence quantum yield (PLQY). The cyclic voltammetry (CV)
analysis of p-­BTPATPE showed its HOMO value as −5.15 eV, close to the work function (WF) of gold
(−5.1 eV), which was suitable for hole transportation. Time-­of-­flight (TOF) transient photocurrent
technique was used to investigate its hole mobility, as shown in Figure 16.3c. The measured sample
was prepared by p-­BTPATPE-­doped poly(styrene) (PS) films (10 mm) with C60 sandwiched between
two indium tin oxide (ITO)-­coated glass slides. The introduction of C60, which had a strong electron-­
accepting ability, was to facilitate the exciton dissociation and increase the efficiency for the charge
carrier generation. The hole mobility was estimated to be 5.2 × 10−4 cm2/V/s, which was higher than
that of N,N′-­diphenyl-­N,N′-­bis(3-­methylphenyl)-­1,10-­biphenyl-­4,40-­diamine, a well-­known hole-­
transporting material under the same measurement conditions [22]. The FET using p-­BTPATPE as
the active layer was also fabricated by depositing 60-­nm thick p-­BTPATPE films onto octadecyl-
trichlorosilane (OTS)-­treated silicon wafers. A layer of gold was deposited on the top of the active
layer through a shadow mask to form the top-­contact source and drain electrodes. X-­Ray diffraction
(XRD) suggested the p-­BTPATPE film was amorphous, which was also confirmed by the smooth
morphology observed under the atomic force microscopy (AFM). Different samples of p-­BTPATPE
with a varied channel length (50, 100, and 150 mm) and width (1 and 2 mm) were tested in the air
under ambient conditions. Results showed that p-­BTPATPE functioned as a p-­type semiconductor,
16.2 ­p-­Type OSCs 489

(b)
(a)

PL intensity (a.u.)
N
N

p-BTPATPE

400 500 600 700


(c) Wavelength (nm)
(d)
0.05 0.0004
10–1 t = 4.3 μs 10–7
Photocurrent (a.u.)

0.04 10–2 10–8 0.0003


Photocurrent (a.u.)

(–IDS)½ (A)½
0.03 10–3
10–9
–IDS (A)

0.0002
10–4
0.02 10–10
1 10 100
Time (μs) 0.0001
0.01 10–11

0.00 10–12 0.0000


0 20 40 60 0 –10 –20 –30 –40 –50
Time (μs) VGS (V)

Figure 16.3  (a) Structure of p-­BTPATPE; (b) PL spectrum of the thin films of p-­BTPATPE; (c) transient
photocurrent for hole transport in p-­BTPATPE/PS/C60 composite (50/48.5/1.5 wt%) under an applied electric
field of 60 V/mm. Inset: logarithmic plot of the photocurrent versus time; (d) transfer curves of a p-­BTPATPE
based FETs (VDS = 50 V, L = 100 μm, W = 1 mm). IDS = drain-source current. Source: Adapted from Ref. [21]
with permission; Copyright Royal Society of Chemistry.

and the thin film (TF) deposited at 80 °C gave the best device performance. Figure 16.3d showed the
transfer I–V curve of p-­BTPATPE obtained under optimized conditions, which gave a hole mobility
of 2.6 × 10−3 cm2/V/s. On/off ratio of the drain current (Ion/off ) obtained between 0 and −50 V gate
bias from the transfer I–V curve was larger than 2 × 104. The performance of p-­BTPATPE was high
for amorphous TPA-­based OSCs.
Although the quantum yield of p-­BTPATPE is very high, the charge mobility of its FET device
is not satisfying. This is possibly ascribed to the relatively twisted structure of TPE moiety despite
its effect in guaranteeing the high solid-­state PLQY. J-­aggregation has been known with long-­
range ordered molecular packing and partially but effective π electron overlap, which not only
promotes the efficient charge transporting but also avoids the luminescence quenching derived
from strong face-­to-­face π–π stacking [23]. Therefore, rationally designed semiconductors with
J-­aggregation alignment is a potential strategy to improve the charge transporting properties and
luminogens. Liu et  al.  [24] have designed and synthesized a derivative of anthracene
2,6-­diphenylanthracene (DPA) by substituting the 2,6-­position of anthracene with phenyl groups
(Figure 16.4a). SC structure analysis indicated that the anthracene core was flat while the sub-
stituted phenyls were slightly out of the plane with a torsion angle of 20.05°, which helped to
reduce the face-­to-­face π–π interactions. More importantly, the nonplanar DPA molecules
490 16  Aggregation-­induced Emission Luminogens with Both High-­luminescence Efficiency and Charge Mobility

(a) (b)

e e

DPA

e e

(c) (d)
0.003 VDS = –60 V VDS = –60 V
0.004

10–6 10–6

0.002
(–IDS)½ (A)½

(–IDS)½ (A)½
–IDS (A)

–IDS (A)
10–8 0.002 10–8

0.001
10–10 10–10

μ = 30 cm2/Vs 0.000 μ = 22.4 cm2/Vs


0.000
10–12 10–12
–40 –20 0 –40 –20 0
VGS (V) VGS (V)

Figure 16.4  (a) Molecular structure of DPA; (b) charge transport along b and c axes in DPA crystals;
(c) typical transfer curve along b axis, L = 16.5 μm, W = 3.3 μm; (d) typical transfer curve along c axis,
L = 26.4 μm, W = 12.1 μm. Source: Adapted from Ref. [24] with permission; Copyright Springer Nature.

adopted J-­aggregation in its crystals that worked effectively to mitigate the fluorescence quench-
ing of DPA in the solid state. It is worthy to note that the multi-­/strong intermolecular C–H⋯π
interactions (with molecular distance only 2.84–2.86  Å) also helped to rigidify the molecular
conformation to benefit the fluorescence. And the compact molecular packing that observed in
DPA crystals enabled the charge transporting through a two-­dimensional (2D) network trans-
porting pathway (Figure  16.4b). As a result, the quantum yields of DPA crystals and powder
were up to 41.2 and 48.4%, respectively, both with blue emission. Bottom-­gate top-­contact tran-
sistors based on DPA micrometer-­sized crystals were fabricated to evaluate the charge transport
properties of DPA. The anisotropy charge transporting property of the SC device was also inves-
tigated as shown in Figure 16.4c and d. Among the examined 100 devices, over 95% of devices
showed mobility larger than 10  cm2/V/s along the b axis, and an ultrahigh mobility of up to
34 cm2/V/s could be achieved. However, only 75% of devices exhibited mobility over 10 cm2/V/s
along the c axis with the maximum mobility of up to 22.4 cm2/V/s. OLED devices based on DPA
showed that a maximum brightness of up to 6627  cd/m2 peaked at 445  nm. Moreover, DPA
OLED arrays could be driven successfully by DPA-­based OFET arrays, which demonstrated the
preliminary integration of optoelectronic devices based on the same OSC, which was the first
time that optoelectronic devices based on the same OSC were integrated together. Their work
suggests the great potential of DPA in real OLET device application.
16.2 ­p-­Type OSCs 491

Besides the crystal engineering to tune the molecular packing of anthracene derivatives to
improve their electronic and photophysical properties, various molecular structure modifications
have also been explored to generate highly emissive semiconductors based on anthracene deriva-
tives. Styrylacenes have emerged as a promising subclass of OSC, which could extend the molecu-
lar π-­conjugation through the styrene group. In Perepichka’s work [25], they used alkyl substituted
phenyl group to modify anthracene and synthesized a structurally asymmetric anthracene deriva-
tive named 2-­(4-­hexylphenylvinyl)anthracene (HPVAnt). It was surprising that HPVAnt exhibited
a PLQY of 55% in solution state but an enhanced PLQY of 70% in the crystal state, which was like
the luminogens with aggregation-­enhanced emission (AEE) property. However, it was also worthy
to note that the powder of HPVAnt only show a PLQY of 20% in the TF state, which actually sug-
gested that there were some quenching factors may exist in the TF state (Figure  16.5a). It was
proposed that the high PLQY of HPVAnt crystals might result from the cessation of the singlet-­
fission process, since the time-­dependent density functional theory (DFT) calculation indicated
that the energy of singlet exciton S1 of HPVAnt (3.04 eV) was insufficient to generate two triplets
(T1 = 1.74 eV). Triplet is generally thought as a quenching factor for fluorescence. Although the
explanation looks reasonable, it does not explain why the TF show lower PLQY than that of both

(a) C6H13 (b)


Au HPVAnt Au

SiO2 (200 nm)


Substrate (n-doped Si)
HPVAnt
(c) (d)
0.02 10–3
0.12
10–5
(IDS)½ (A)½

0.08
IDS (mA)

0.01
IDS (A)

10–7

0.04
10–9

0.00
10–11 0.00
–60 –30 0 0 –30 –60
VGS (V) VGS (V)
(e)
Increasing drain-source bias

Figure 16.5  (a) Structure of HPVAnt; (b) schematic top-­contact TF-­FET fabricated with HPVAnt; (c) transfer
curves of the biased device. VDS = −75 V; (d) output curves of the biased device. VGS = −75 V (blue), −60 V
(teal), −45 V (pink), −30 V (black), −15 V (red); (e) shift of the EL zone upon increasing bias in a SC-­FET
configuration. Source: Adapted from Ref. [25] with permission; Copyright Wiley-­VCH.
492 16  Aggregation-­induced Emission Luminogens with Both High-­luminescence Efficiency and Charge Mobility

the solution and crystal. Considering excited-­state molecular motion correlated directly with the
luminescence efficiency of luminogens, we suggest that the lower PLQY of the film of HPVAnt
possibly results from some loose packing dimers, or oligomers were formed in the films, which
have some translational motion to quench the emission. While in the crystal state, this kind of
intermolecular translational motion was effectively suppressed, leading to the enhanced emission
of crystals. Actually, many previously reported acenes such as anthracene and perylene also show
the same trend of PLQY (for example, the PLQY of the crystal, solution, and powder of anthracene
are 64, 28, and 18%, respectively) [26]. CV measurement gave the HOMO energy level of −5.47 eV
for HPVAnt, which was suitable for hole transportation. The charge transporting properties of the
HPVAnt were then studied in TFs and SCs top-­contact bottom-­gate device configurations
(Figure 16.5b). Remarkably, without any device optimization, TF-­FET based on HPVant had an
average hole mobility value of (1.14  ±  0.2) cm2/V/s with the highest value of 1.5  cm2/V/s
(Figure 16.5c, d). At the same time, the devices based on the poly(methyl methacrylate) (PMMA)-­
coated substrate showed excellent stability after the storage of the unencapsulated devices under
ambient conditions (the 𝜇 value dropped from 1 to 0.9 and 0.84 cm2/V/s after 3-­ and 12-­month
storage, respectively). The high mobility in TF-­FET benefits from the almost perfect layer-­by-­layer
growth at room temperature, which lead to continuous OSC films at a low nominal thickness.
Meanwhile, XRD analysis of the vacuum-­deposited films suggested that the molecules packed in a
usual herringbone fashion with the anthracene groups of adjacent molecules align with one
another to form sheets that were perpendicular to the long molecular axis (monoclinic unit cell).
Such packing maximized the 2D π interactions between the molecules and benefited the charge
transportation. The EL property was also investigated, the EL spectrum closely resembled the PL
spectrum with a little broadened, peaked at 490 nm. Figure 16.5e showed optical images of EL of
a device operating at a fixed gate voltage (−200 V) while sweeping the drain-­source voltage (VDS)
between 0 and −200 V. As increasing the electron injection at higher VDS values progressively, the
light emission zone extended from the drain electrode to the source electrode. It is worth mention-
ing that unlike other single-­layer OLETs, the emission of HPVAnt was also not subjected to a
“fatigue” effect [27], which meant the emission would not fade out during continuous operation.
As clearly demonstrated, intermolecular interactions and orientations are critical to generate
solid-­state OSCs with strong emission and efficient charge transport simultaneously. Pyrene is
one kind of rigid building block with broader π-­conjugation than anthracene and is highly emis-
sive both in solution and solid states. Therefore, pyrene derivatives are one kind of promising
candidates to develop light-­emitting semiconductors. Zhang incorporated 𝛽-­styryl units into the
different substitution positions of pyrene core to tune the electronic properties and intermolecu-
lar interactions of pyrene derivatives (Figure 16.6a) [28]. By employing 1,6-­dibromopyrene and
2,7-­dibromo pyrene as substrates, respectively, they obtained two pyrene derivatives named
16PyE and 27PyE. Both 16PyE and 27PyE are brightly emissive in either solution state or crystal
state, suggesting that pyrene is an excellent building block to construct light-­emitting semicon-
ductors. It was very interesting that from solution to crystal, 16PyE exhibited ACQ property and
27PyE exhibited AEE property. The SC structure analysis suggested that both 16PyE and 27PyE
adopted E-­conformation in their crystals, and their phenyl groups were not coplanar with the
respective pyrene rings, with a dihedral angle of 18.46° for 16PyE and 32.42° for
27PyE. Additionally, the styryl substitution at 2,7-­positions generally is thought to have less per-
turbation on the electronic properties of pyrene since nodal planes pass through the 2,7-­positions
in both the highest occupied molecular orbitals (HOMO) and lowest unoccupied molecular
orbitals (LUMO) orbitals of pyrene. 16PyE thus showed better π-­conjugation than 27PyE, which
was consistent with the fact that 16PyE exhibited better planarity than 27PyE. The worse planar-
ity and conjugation of 27PyE than that of 16PyE suggested that 27PyE had more active excited
16.2 ­p-­Type OSCs 493

(a) (b)

27PyE

(c) (d)
0.010 6 × 10–5
10–4 –40 V
VDS = –100 V 0.008
–50 V
10–6 –60 V

(–IDS)½ (A)½
4 × 10–5 –70 V
0.006
–IDS (A)

–IDS (A)
–80 V
10–8
0.004
10–10 2 × 10–5
0.002
10–12
0.000 0
–80 –60 –40 –20 0 –80 –60 –40 –20 0
VGS (V) VGS (V)

Figure 16.6  (a) Molecular structures and dihedral angles between the respective benzene ring and the
pyrene core of 27PyE; (b) intermolecular arrangements of 27PyE; output (c) and transfer characteristics
(d) of OFETs with thin films of 27PyE deposited on OTS-­modified SiO2/Si substrate at 50 °C; W = 1440 nm,
L = 40 nm. Source: Adapted from Ref. [28] with permission; Copyright American Chemical Society.

intramolecular motion than 16PyE in the excited state, which possibly caused its lower quantum
yield in solution (QY = 15.2%) than that of 16PyE (QY = 81.4%). Notably, in the crystals, multi-
ple intermolecular interactions helped rigidify the molecular conformation of 27PyE, resulting
in its enhanced emission (QY = 27.4%). Although the dense packing of 16PyE kind of quenched
its solid emission but still, its crystal was bright emissive with a quantum yield of QY = 28.8%
since the neighboring pyrene units did not show a face-­to-­face arrangement (Figure 16.6b). It is
worthy to note that the relative compact packing of both 16PyE and 27PyE are much favorable
for efficient charge transport. The charge mobility property of 16PyE and 27PyE were explored
by fabricating the bottom-­gate/bottom-­contact OFETs with their films deposited on OTS modi-
fied SiO2/Si substrate at different temperatures. As applying the negative VGS, current IDS
increased, both TFs of 16PyE and 27PyE showed typical p-­type characteristics. In comparison
with 16PyE, the device of 27PyE annealing at 50 °C exhibited much higher mobility of 1.66 cm2/
V/s (Figure 16.6c, d) than that of 16PyE (0.17 cm2/V/s) at the same conditions. The TFs of 16PyE
and 27PyE were characterized by AFM and XRD, and the results indicated that the film of 27PyE
showed high crystallinity with diffractions up to fourth orders after annealing. It was commonly
accepted that annealing could help molecules to recrystallize and form a certain degree-­ordered
arrangements. However, as a contrast, the film of 16PyE showed no diffractions at different
annealing temperature, which was consistent with its lower charge mobility. Furthermore, the
AFM images of 27PyE showed that layer structures were formed, and the grain size obviously
increased when annealing at 50 °C. While for 16PyE, no layer structures were formed during the
annealing process although the grain size also increased. This may also explain the variation of
mobilities of 16PyE and 27PyE. These results indicated that molecular semiconductors with AIE
property have great potential in constructing high-­mobility light-­emitting devices.
494 16  Aggregation-­induced Emission Luminogens with Both High-­luminescence Efficiency and Charge Mobility

Except for styrylacenes (double bond linkage), triple bond is also a commonly used conjugation
linker to extend the π-­conjugation of semiconductors. The linear structure of the triple bond does
not undermine the molecular coplanarity while introducing certain flexibility into the rigid large
π-­conjugated molecule. Zhao et al. utilized multiple triple bonds to construct the polyyne-­bridged
AIEgen, which can not only broaden the π-­conjugation length of AIEgens but also avoid the emis-
sion quenching resulted from too large π-­extension  [29]. Tian and coworkers employed triple
bonds as linkers to bridge pyrene with the spatial twisted TPA group and constructed two mole-
cules of 1,6-­DTEP and 2,7-­DTEP [30]. The regular millimeter crystals of 1,6-­DTEP and 2,7-­DTEP
were obtained by slow solvent evaporation method in the ambient atmosphere. In 1,6-­DTEP’s SC,
the two dihedral angles between two triple bonds linking benzene rings of TPA groups and pyrene
rings were minimal (6.25 and 6.25°) as shown in Figure 16.7a. Therefore, the benzene rings of
TPA groups and pyrene rings were almost coplanar, improving the conjugation length of
1,6-DTEP molecules. However, the dihedral angles in 2,7-­DTEP were 38.40 and 39.38°. Thus,
1,6-DTEP had tighter molecular packing than that of 2,7-­DTEP (Figure 16.7b). Moreover, there
were more intermolecular interactions in 1,6-­DTEP crystals, which helped lock the molecular
­conformation and benefited both the luminescence and the charge transport properties. The pho-
tophysical spectra investigation of crystals indicated that both 1,6-­DTEP and 2,7-­DTEP form
J-­aggregation in their crystals. The crystals of 1,6-­DTEP and 2,7-­DTE showed much high PLQY of
32 and 35%, respectively. According to the CV measurement, HOMO energy levels of 1,6-­DTEP
and 2,7-­DTE were calculated to be −5.30 and −5.33 eV, which were suitable for the application of
p-­type charge transportation. To investigate the carrier transport properties of the two molecules,
bottom-­gate top-­contact SC-­FETs were fabricated on OTS-­modified silicon wafers, using Au as
the source and drain electrodes. The anisotropy of the carrier transport of 1,6-­DTEP was studied
by using SC-­OFETs along the c-­axis and b-­axis. The highest mobility along the c-­axis was up to
2.1 cm2/V/s with an Ion/off ratio of 106. Moreover, the highest mobility along the b-­axis was about
0.4 cm2/V/s with an Ion/off ratio of 105 (Figure 16.7c). For 2,7-­DTE, the highest mobility was just
0.025 cm2/V/s due to the loose packing. These results suggest that simply alternation of the con-
jugated linkers possibly a feasible and effective strategy to tune the photophysical and charge
transporting properties of OSCs.

(a) (c) 10–6 VGS = 60 V 0.0012


6.25°
c axis
6.25°
b axis
1,6-DTEP 10–8 0.0008
(–IDS)½ (A)½
–IDS (A)

(b)

10–10 0.0004

10–12 0.0000
–60 –40 –20 0
VGS (V)

Figure 16.7  (a) The single-­crystal structure of 1,6-­DTEP; (b) the packing of 1,6-­DTEP single crystal; (c) the
transfer curves of 1,6-­DTEP SC-­OFETs (c-­axis and b-­axis). Source: Adapted from Ref. [30] with permission;
Copyright Wiley-­VCH.
16.3 ­n-­Type OSCs 495

16.3  ­n-­Type OSCs

Compared with p-­type OSCs, the design of n-­type OSCs is more challenging since electron can
easily be attracted by water, oxygen, and various impurities in ambient conditions. In addition, the
commonly used gold electrode has a WF of approximately −5.1 eV, which also need a matched
LUMO energy level of the materials to complete the electron injection. Therefore, a low LUMO
energy ( −3.7 eV) level is usually needed to facilitate the electron injection and stabilize the elec-
tron transport. To lower the LUMO energy level of the molecule, the commonly used strategy is the
introduction of electron-­deficient building blocks. However, in comparison with the electron-­rich
building blocks, the variety of electron-­deficient building blocks is rather limited. Until now, the
studied emissive n-­type semiconductors mostly are based on the classic n-­type semiconductors,
such as naphthalene diimide (NDI), perylene diimide (PDI), and their derivatives. Tan et  al.
employed brominated NDI and the organometallic reagent of vinyl linkage as monomers
(Figure  16.8a) to do the Stille coupling polymerization and fabricated a red emissive polymer
poly(naphthalene diimide) vinylene (PNV) with n-­type charge transport property in OFETs [31].
In the molecular design, long-­branched alkyl chains were introduced into the nitrogen of the diim-
ide group to enhance the solubility of polymer, and PNV thus exhibited good solubility in chloro-
form and was suitable for spin-­coating process method to fabricate TF devices. PNV emitted at
550 nm in chloroform solution and peaked at 632 nm in a TF with PLQY of 28.3 and 33.4%, respec-
tively. The largely red-­shift and broadened emission in the solid state might have resulted from the
aggregation of PNV molecules. PNV did not suffer from the emission quenching problem, which
possibly was ascribed to the long alkyl chains that inhibited the compact π–π stacking in the solid
state [32]. The fluorescence spectra of PNV in the mixture of CHCl3/hexane with varied fraction of
hexane showed that the peak at 550  nm gradually decreased and a new emission peak around
620 nm emerged with the increase of the hexane fraction. This phenomenon suggested that the
longer wavelength emission in the TF of PNV should be contributed by the enhanced intermolecu-
lar interactions. An oligomer with two NDI cores and one vinyl linkage was chosen for the DFT
calculation, and the alkyl chains were replaced by methyl to simplify the calculation process
(Figure 16.8b). The result suggested that noncovalent C–H⋯O interactions were formed between
the carbonyl group and the vinyl linkage, which helped rigidify the molecular conformation.
Meanwhile, the H-­bonds in PNV had a slightly twisted angle of about 29° from the H on vinyl to
the NDI plane due to the steric hindrance from the neighboring H atom. This led to a slightly
twisted polymer backbone, resulting in a weak intermolecular interaction and a low crystallinity
to facilitate the emission in the solid state. The LUMO of PNV according to CV measurement was
calculated to be −3.8  eV, beneficial for stabilizing the electron charge carriers. To evaluate the
charge transporting property of PNV, bottom-­gate/bottom-­contact OFET devices based on the TF
of PNV were fabricated by spin coating technique. The optimized device exhibited an average
mobility of 0.9 × 0−3 cm2/V/s with the maximum electron mobility of up to 1.5 × 10−3 cm2/V/s. To
better understand the luminescence and charge transporting property of PNV polymer, grazing-­
incidence X-­ray diffraction (GI-­XRD) and AFM were carried out. In general, the 2D-­GI-­XRD pat-
tern manifested a relatively long π–π stacking distance, which benefited the strong solid-­state
emission but unfavorable for efficient electron hopping. In addition, the AFM image showed a
small grain size and a low roughness morphology in the TF that suggested the low crystallinity of
the film and possibly correlated with the relatively low electron mobility of PNV compared to those
of the highly crystalline NDI based donor–acceptor conjugated polymers [33].
To further enhance the intermolecular interactions and improve the charge transporting proper-
ties of PNV, Tan et al. tried to introduce thiophene into the polymer backbone of PNV through a
496 16  Aggregation-­induced Emission Luminogens with Both High-­luminescence Efficiency and Charge Mobility

H-bond
(a) (b)
C10H21
109.5°
C8H17 2.75 A
O
N
O H

n
H
O
N Twisted construction
O
C8H17 29°

C10H21 31°
PNV
(c) (d)
10–6
VDS = 80 V VGS
0.5
0V
0.0006 10 V
0.4 20 V
10–8 30 V
40 V
|IDS|½ (A)½

50 V
0.3
IDS (μA)
|IDS| (A)

0.0004 60 V

10–10 0.2
0.0002
0.1
10–12
0.0000 0.0

–20 0 20 40 60 0 20 40 60 80
VGS (V) VGS (V)

Figure 16.8  (a) Structure of PNV; (b) demonstration of the C–H⋯O hydrogen bond and twisted
construction for the calculated oligomer (n = 2) structure of PNV from the front and side views, respectively;
transfer (c) and output (d) characteristics of OFET devices (W = 1440 μm, L = 50 μm) based on PNV
measured under ambient conditions. Source: Adapted from Ref. [31] with permission; Copyright Royal
Society of Chemistry.

simple random polymerization of NDI, bithiophene, and vinylene monomers and produced a
series of polymers named PNV-­DTn  [34]. Basically, the charge carrier mobility in conjugated
­polymers depends on both the intrachain and interchain charge transport along the polymer
­backbone. It needs at least partial intermolecular hopping through short π–π bridges. Therefore,
different ratios of bithiophene units were embedded into the PNV backbone to tune the intramo-
lecular π-­conjugation and intermolecular π interactions, affording polymer PNV-­DT1, PNV-­DT3,
PNV-­DT5, PNV-­DT7 with varied bithiophene fragment. The emission maximums of the polymers
from PNV-­DT1 to PNV-­DT7 were 711, 717, 721, and 730 nm in solid states, correspondingly with
PL efficiencies of 16.6, 14.0, 13.8, and 11.2%. The gradually red-­shifted emission suggested that the
charge transfer or electron communication enhanced with the increasing ratio of bithiophene. In
addition, although the quantum yield of these polymers slightly decreased with the ratio of bithio-
phene unit increased, the values were still much impressive in consideration of their near-­infrared
16.3 ­n-­Type OSCs 497

emission. These results suggest that the randomized polymerization strategy might be useful in
designing conjugated polymers with near-­infrared solid emission. The LUMO energy levels derived
from the CV measurement were all relatively close, at around −3.7 eV. Top-­gate/bottom-­contact
OFET devices based on the four polymers were fabricated through a conventional spin coating
method. PNV-­DT7 gave the best performance with electron mobility (𝜇e) of up to 0.011 cm2/V/s
and an average 𝜇e of 8.5 × 10−3 cm2/V/s (Figure 16.9). Devices using PNV-­DT5 as the active layer
exhibited similar 𝜇e as that of PNV-­DT7, with a highest mobility of 9.2  ×  10−3  cm2/V/s. For
PNV-­DT1 and PNV-­DT3, mobilities were only around 0.003 cm2/V/s, but still higher than that of
the PNV devices. The improved OFET performance of PNV-­DTn demonstrated that introduction of
bithiophene fragment into the PNV polymer backbone was an effective way to enhance the charge
transport. AFM and GI-­XRD were carried out further to understand the influence of bithiophene
units on the property of the polymers. Although the AFM results indicated that the TFs of the four
polymers all exhibited relatively smooth pattern, without obviously crystallinity change, while
GI-­XRD indicated that intermolecular π–π interaction was obviously strengthened along the rising
bithiophene ratio. Overall, introduction of bithiophene with limited ratio could enhance the local-
ized π–π intermolecular interactions while retaining the overall low crystallized TF for these poly-
mers and therefore balanced the charge carrier mobility and the solid emission.

(a) C10H21 C10H21

C8H17 0.011 cm2/Vs 730 nm


O C8H17 O
N N φ = 11.2%
O O

S
S X
1-X

O O N
N O
C8H17 O C8H17

C10H21 C10H21
PNV-DT7
(b) (c)
10–5
0.6 VGS
VDS = 80 0.0012
0V
0.5 10 V
0.0010
10–7 20 V
0.4 30 V
0.0008 40 V
(IDS)½ (A)½

IDS (μA)

50 V
IDS (A)

0.3
10–9 0.0006

0.2
0.0004

10–11 0.1
0.0002

0.0
0.0000
10–13
–20 0 20 40 60 0 20 40 60 80
VGS (V) VGS (V)

Figure 16.9  (a) Structure of PNV-­DT7; transfer (b) and output (c) characteristics of OFET devices
(W = 1000 mm, L = 40 mm) based on PNV-­DT7 measured under ambient conditions. Source: Adapted from
Ref. [34] with permission; Copyright American Chemical Society.
498 16  Aggregation-­induced Emission Luminogens with Both High-­luminescence Efficiency and Charge Mobility

In general, the crystallinity of the TF of the polymer has a big influence on the performance of
the polymer-­emitting semiconductors; however, when really applied to the optoelectronic devices,
defects and kinks in the TFs of polymers possibly trap the charge carriers, and even annihilate the
excitons, which will impair the efficiency of the devices to a large degree. Compared with poly-
mers, organic small molecules have definite structures and their SC is easy to obtain, which is
better for analysis of the relationship between molecular structure and photophysical properties.
Zhang and his coworkers reported a series of arylacetylene-­substituted NDIs with both
­light-­emitting and semiconducting functions (Figure 16.10a) [35]. In their work, linear ethynylene
(–C≡C–) was used as a conjugation bridge to link the NDI and different donor units. The triple
bonds extended the π-­conjugation while introducing a certain degree of flexibility by reducing the
steric hindrance. The absorption spectra of the TF of NDI derivatives showed a red shift in com-
parison with that in solution, which suggests the enhanced intermolecular interactions in the
TF. The absorption of NDIs with –C6H13 alkyl chains exhibited larger red-­shift than those with
longer alkyl chain –C20H41. By changing the different functional groups, the emission colors could
be tuned from green to red. Furthermore, all the six compounds showed AIE properties with
higher PLQY in the solid state compared with that of the solution. NDIs with shorter alkyl chains
showed relatively higher PLQY and longer red-­shifted emission wavelength than those with
–C20H41 chain. This is because that the bulky alkyl chains will cause weaker intermolecular
interactions due to the spatial isolation of the molecules resulting from the flexible alkyl chains
(Figure 16.10b). Meanwhile, due to the existence of the long alkyl chains and the resulted spatial
isolation effect, the excited-­state internal rotations of the triple bond and donor units could not be

(a) R
(b)
O N O
S

S
O N O
R

TNDI-20: R = C20H41
TNDI-6: R = C6H13
(c) (d)
4 4
10–5 10–4
VDS = 60 V VDS = 60 V
3 10–5 3
10–7
10–6
(IDS)½ (A)½

(IDS)½ (A)½
IDS (A)
IDS (A)

2 2
10–9 10–7
1 10–8 1
Vth = 7.5 V Vth = 21 V
10–11 μ = 0.035 cm2/Vs μ = 0.075 cm2/Vs
10–9
Ion/Ioff = 106 0 Ion/Ioff = 105 0
10–13 10–10
–20 0 20 40 60 –20 0 20 40 60
VGS (V) VGS (V)

Figure 16.10  (a) Structure of TNDI-­20 and TNDI-­6; (b) illustration of the intermolecular arrangements of
TNDI-­20 and TNDI-­6; (c) typical transfer curves of OFETs with thin films of TNDI-­20; and (d) TNDI-­6
(W = 1440 μm, L = 50 μm [TNDI-­20], L = 40 μm [TNDI-­6]). Source: Adapted from Ref. [35] with permission;
Copyright Wiley-­VCH.
16.3 ­n-­Type OSCs 499

effectively restricted even in the solid state. Consequently, the PLQY of these molecules was rela-
tively low (PLQY < 10%). The reduction potentials detected from CV measurements provided the
LUMO energy levels of NDI derivatives, in which the values of TNDI-­20 and TNDI-­6 were both
−3.98  eV. The semiconducting behavior of these compounds was evaluated by characterizing
bottom-­gate bottom-­contact OFETs based on their TFs. After thermal annealing at 100 °C, their
performances got improved. The electron mobility of TNDI-­20  increased from 5.4  ×  10−3 to
0.035 cm2/V/s and mobility of TNDI-­6 increased from 7.0 × 10−4 to 0.075 cm2/V/s (Figure 16.10c, d).
XRD measurement indicated that the crystallinity of TFs was obviously improved upon thermal
annealing, which was consistent with the improved electron mobility. AFM measurements showed
that morphologies were also changed after annealing at 100 °C. For instance, in the TF of TNDI-­20,
the sizes of the crystalline domains increased to about 250–350 nm, and the formation of intercon-
nected microplates was observed in the TF of TNDI-­6, which was beneficial for charge transport.
Besides, arylacetylene-­substitute NDIs with longer alkyl chains showed better air stability. This
could be explained as follows: long alkyl chain may be beneficial to exclude oxygen and water mol-
ecules, which would otherwise hamper the stability of the respective anions in air.
PDI is another commonly used building block for the construction of n-­type semiconductors
because of its easy commercial availability, versatile synthetic modification, high electron mobility,
and excellent chemo-­/photostability. However, the sizeable rigid π-­framework and strong π–π
interaction of PDI is harmful to its solid emission due to the ACQ problem. Zhao et al. reported a
series of triphenylethylene (TriPE) derivatives with a varying number of perylenediimide (PDI)
units, in which, TriPE-­3PDI showed near-­infrared emission with high solid QY and optimized
electron mobility of over 0.01 cm2/V/s (Figure 16.11) [36]. The modified molecule, TriPE-­3PDIL,
with a longer alkyl chain, even exhibited high mobility of up to 0.04 cm2/V/s in the annealed OFET
device. Decorating conventional ACQ luminogens with typical AIEgens, such as TPE, is a common
way to construct new molecules with high solid emission and improved photophysical properties.
Tang and coworkers had modified PDI at the bay area with TPE and produced a series of deep-­red
emissive light-­emitting semiconductors in previous work; unfortunately, although the ACQ effect
of PDI was successfully suppressed, the TPE-­substituted PDIs exhibited reduced charge mobil-
ity [37]. That might be ascribed to the fact that the highly twisted structure of the TPE unit inhib-
ited the efficient charge transport. In Zhao’s molecular design, PDIs were attached to the peripheral

(a) (b)
10–6
VDS = 100 V 0.8
O N O 10–7
R R 0.6
(IDS)½ (A)½

10–8
IDS (A)

R= 0.4

10–9 Vth = 8 V
R 0.2
μ = 0.0042 cm2 V–1 s–1
10–10 Ion/Ioff = 9 × 103
O N O 0.0
0 20 40 60 80 100
TriPE-3PDI
VGS (V)

Figure 16.11  (a) Structure of TriPE-­3PDI; (b) transfer characteristic of OFET device of TriPE-­3PDI annealed
at 140 °C. Source: Adapted from Ref. [36] with permission; Copyright Wiley-­VCH.
500 16  Aggregation-­induced Emission Luminogens with Both High-­luminescence Efficiency and Charge Mobility

side of the TriPE to strengthen the intermolecular interactions of the PDI unit and promote the
intermolecular charge transport. TriPE was chosen as the luminogenic core because of its less
crowded structure compared with TPE, which could endow the molecule with higher planarity.
The achieved TriPE-­1PDI, TriPE-­2PDI, and TriPE-­3PDI all showed deep-­red or NIR emission in
both solution and solid state. Among them, TriPE-­3PDI emitted at 680 nm with a PLQY of 30% in
the solid state. The LUMO energy levels calculated by CV were around −3.8 eV, suitable for elec-
tron transport. Bottom-­gate/top-­contact OFETs devices were fabricated through a solution process
method under a nitrogen atmosphere. TriPE-­3PDI showed the highest electron mobility of
0.01 cm2/V/s with average mobility of 4.2 × 10−3 cm2/V/s. TriPE-­3PDIL had similar optical proper-
ties as TriPE-­3PDI, while the PLQY in the solid state was only 6%. The longer alky chain endowed
the molecule with better solubility and improved the film morphology to achieve better charge
transporting performance. The annealed device of TriPE-­3PDIL exhibited high mobility of up to
0.04 cm2/V/s with average mobility of 0.02 cm2/V/s. AFM and XRD indicates that the film crystal-
linity improved upon annealing, which benefited the electron mobility. The results were among
the best of NIR emitters with AIE characteristics.

16.4  ­Ambipolar OSCs

The unipolar OLETs devices, using only one type of semiconductors, were commonly considered
to be performed with low luminescence due to the frequently occurring metal contact-­induced
exciton quenching in the area closing to the electrodes, from which the minority charge carriers
are injected. Ambipolar OLETs, which have a balanced charge carrier injection and a spatially
controlled recombination zone within the lateral channel, are promising candidates to have better
device performance. Different protocols have been applied to construct ambipolar OLETs such as
the use of asymmetric contacts, shorting the channel length, and modification of the dielectric
layer. Ambipolar OLETs based on a TF of a single-­component OSC ought to be the most facilely
method to construct ambipolar OLET devices. Unfortunately, in comparison with the flourishing
developed unipolar semiconductors, the development of high-­performance ambipolar semicon-
ductors with balanced hole and electron transport are still much lagged.
Zhao and coworkers developed a kind of TPE derivative featured with cross-­shaped (chiasmatic)
molecular conformations [38]. Interestingly, although these molecules did not contain typical electron
withdrawing groups to facilitate the electron transport, they exhibited ambipolar charge transport
properties. Figure  16.12a showed the structure of one of the derivatives, (Z)-­o-­BCaPTPE. (Z)-­o-­
BCaPTPE emits at 497 nm in THF solution with a PLQY of 43% and 477 nm in the TF with enhanced
PLQY of 100%, suggesting its AEE property. The rigid chiasmatic conformation of (Z)-­o-­BCaPTPE
partially inhibited the intramolecular rotation that thus explained why its emission was brightly emis-
sive (Figure 16.12b). While in the solid state, intramolecular π–π interaction and multiple C–H⋯π
hydrogen bonds were found in the crystal, helping rigidify the molecular conformation and achieving
unity PLQY. Different from many organic materials, in which the charge transport was dominated by
either holes or electrons, (Z)-­o-­BCaPTPE possessed bipolar charge transport capability; although lack
of strong electron withdrawing group to facilitate the electron injection and stabilize the electron
transport, they ascribed the bipolar charge transporting property to its special chiasmatic conforma-
tion. TOF transient photocurrent technique was used to evaluate the charge mobility of these unique
AEE active semiconductors. (Z)-­o-­BCaPTPE-­doped poly(styrene) films (10 μm) with C60 were pre-
pared and sandwiched between two ITO-­coated glass slides for the measurement. As shown as in
Figure  16.12c and d, the hole mobility was calculated to be 4.9  ×  10−4  cm2/V/s and the electron
16.4  ­Ambipolar OSC 501

(a) (b)

N
Ar =

Ar Ar d = 3.274 Å

(Z)-o-BCaPTPE
(c) (d)
10–1
10–1
Photocurrent (a.u.)

Photocurrent (a.u.)
10–2

10–2
t+ = 3.4 μs, 10–3 t– = 3.8 μs,
μ+ = 4.9 × 10–4 cm2/Vs μ– = 4.3 × 10–4 cm2/Vs

10–3 10–4
1 2 4 6 8 10 1 2 4 6 8 10
Time (μs) Time (μs)

Figure 16.12  (a) Structure of (Z)-­o-­BCaPTPE; (b) ORTEP drawings of (Z)-­o-­BCaPTPE (CCDC 825621);
log–log plot of the photocurrent as a function of time in (c) hole transport and (d) electron transport
configurations for the film (10 μm) of (Z)-­o-­BCaPTPE:PS: C60 (50 : 48.5 : 1.5 wt%) composite. Source: Adapted
from Ref. [38] with permission; Copyright Wiley-­VCH.

𝜇e was more than one order of magnitude higher than that of tris(8-­hydroxyquinolinolato)aluminum
mobility was calculated to be 4.3 × 10−4 cm2/V/s at an applied electric field of 60 V/μm. The value of

(Alq3), a commonly used electron-­transporting material. OLEDs devices using (Z)-­o-­BCaPTPE as an


emitting layer were fabricated. Without any electron-­or hole-­transporting layer, OLED device based
on (Z)-­o-­BCaPTPE still showed good OLED performance, demonstrating the good electron-­and hole-­
transporting capabilities of (Z)-­o-­BCaPTPE in addition to excellent light-­emitting properties. As a con-
trast, the linear counterparts, p-­BCaPTPE had hardly any electron-­transporting property, indicating
the importance of conformation regulation in generating electron transporting semiconductors.
Although (Z)-­o-­BCaPTPE exhibits ambipolar charge transporting property, the relatively low
charge mobilities (~10−4 cm2/V/s) are of more fundamental research importance rather than real
application. The design of molecules with J-­aggregation has been turned out to be an effective
strategy to balance the molecular packing and luminescence process, thus are promising in achiev-
ing molecules with both high mobility and luminescence. Ma et al. reported a butterfly molecule
BDPV2T whose SC adopted J-­aggregation alignment (Figure 16.13) [39]. The molecular backbone
(the bithiophene core and E-­form phenyl groups) exhibited a slightly twisted structure with an
angle of 32.7°, which did not affect the whole π-­conjugation along the main backbone. Consequently,
the electron can be well-­delocalized over the molecular backbone and extended electron cloud
distribution over the whole conjugated backbone was critical to the efficient charge transport. In
comparison, the phenyl groups of Z-­form BDPV2T were nearly vertical to the conjugated plane
with a large dihedral angle of 78.9°, exhibiting a “butterfly-­like” structure that inhibited the strong
intermolecular π–π stacking and is beneficial for the strong photoluminescence. Multiple strong
502 16  Aggregation-­induced Emission Luminogens with Both High-­luminescence Efficiency and Charge Mobility

(a) (b)

S Au/MoO3 Ca
S
SiO2
Si
BDPV2T

(c) 3.939 Å

31°
a

(d) (e)
1.2 1.2
VDS = –100 V 100 VDS = 100 V 100
1.0 1.0

0.8 0.8
(–IDS)½ (μA)½

(–IDS)½ (μA)½

10–1
–IDS (μA)

–IDS (μA)
10–1
0.6 0.6

0.4 0.4
10–2
0.2 10–2 0.2

0.0 0.0
10–3
–100 –80 –60 –40 –20 0 20 –20 0 20 40 60 80 100
VGS (V) VGS (V)

Figure 16.13  (a) Structure of BDPV2T; (b) schematic device structure of single-­crystal OLETs; (c) crystal
alignment of BDPV2T molecules; (d and e) Ambipolar transfer curves of BDPV2T single-­crystal OLETs.
Source: Adapted from Ref. [39] with permission; Copyright Wiley-­VCH.

C⋯C and C–H⋯π interactions between adjacent molecules in two dimensions benefited the self-­
assembly of BDPV2T to form 2D crystals. In the horizontal crystal plane, the BDPV2T molecules
adopted a herringbone molecular packing pattern with partially molecular conjugation overlap
along the c axis as well as strong intermolecular interactions between the neighboring molecules,
which were beneficial for efficient charge transport and high-­carrier mobility. Additionally, due to
the steric hindrance and strong intermolecular interactions induced by the peripheral phenyls, the
BDPV2T adopted slide stacking with a pitch angle (α) of 31° and perpendicular distance (d) of
3.939 Å. Such typical J-­aggregation with a pitch angle of less than 54.7° had been demonstrated to
be an ideal stacking mode for integrating optical and electronic properties in organic
16.4  ­Ambipolar OSC 503

molecules [40]. The crystals of BDPV2T had a high PLQY of 30% peaked at 540 nm, higher than
that of the solution (11%), showing AIE characteristics. The HOMO and LUMO energy levels
obtained from its CV curves were −5.21 and −2.71 eV, respectively, suggesting the hole-­transporting
capability of BDPV2T. Bottom-­gate top-­contact SC-­OFETs on OTS-­treated Si/SiO2 substrates with
gold films as the source and drain electrodes were fabricated. The calculated average charge mobil-
ity was about 0.5 cm2/V/s, with the highest mobility of 1.0 cm2/V/s. It is worth mentioning that
SC-­OFETs based on BDPV2T exhibited high stability when kept under ambient conditions, and
the transfer curve scarcely changed compared to the first scan after 60 days. It is worthy to note
that by electrode selection or modification, ambipolar charge transport OLET can also be achieved
by using BDPV2T as active layer. An asymmetric device structure was used for BDPV2T SC-­based
OLETs with a high WF of Au/MoO3 (−5.3 eV) as the hole injection electrode and a low WF of Ca
(−2.8 eV) as the electron injection electrode. Relatively, balanced ambipolar charge transport with
a hole and electron carrier mobility of 0.14 and 0.2  cm2/V/s, respectively, were afforded. The
obtained EL spectra from the devices were quite similar to the PL of the crystal. Moreover, the EL
spectra of the OLET devices were stable against the gate voltage. With the increase of the gate volt-
age, the intensity of the EL spectra became stronger with no peak shift observed, suggesting the
reliable operation of BDPV2T OLETs.
Li et al. reported an anthracene derivative, dNaAnt, which showed a PLQY of 29.2% in the solid
state with high-­charge mobility [41]. This work basically is an extension of their previous work in
which they reported the molecule DPA with both strong blue emission and high hole mobility [14].
In this work, they prolonged conjugation length DPA by changing the phenyl ring with naphtha-
lene group, which was expected to tune the energy levels and modify the packing arrangements for
the charge injection and transport. From the Ultraviolet-­visible (UV–Vis) absorption and ultravio-
let photoelectron spectroscopy, the HOMO and LUMO energy levels were estimated to be −5.33
and −2.60 eV, respectively, corresponding well with the CV results. In the bottom-­gate top-­contact
SC-­FETs, average mobility of 7.0 cm2/V/s was obtained. The highest mobility was up to 12.3 cm2/
V/s with a threshold voltage around −10  V and Ion/off about 3.9  ×  107. Similarly, by employing
asymmetrical electrodes (Au/MoO3: Ca/CsF) to achieve better energy matching, ambipolar OLETs
based on the SC of dNaAnt were fabricated (Figure  16.14a). From the transfer curve, balanced

(a) (b)
10–4
VDS = 60 V

dNaAnt
–2
2.5
|IDS (A)|

2.6 10–5
Ca/CsF 2.8 LUMO
–3
Energy level (eV)

Ca
dNaAnt

–4
VDS = –60 V
10–6
–5 5.1 5.3
HOMO
5.3 Au Au/MoO3
–100 –50 0 50 100
–6
VGS (V)

Figure 16.14  (a) Energy level alignment for efficient charge injection; (b) typical ambipolar transfer curves
of the organic light-­emitting transistors. Source: Adapted from Ref. [41] with permission; Copyright American
Chemical Society.
504 16  Aggregation-­induced Emission Luminogens with Both High-­luminescence Efficiency and Charge Mobility

ambipolar charge mobility was calculated to be 𝜇h = 1.10 cm2/V/s and 𝜇e = 0.87 cm2/V/s, which is


among the best performances for ambipolar OLETs (Figure 16.14b).
Besides device manipulation, to enhance the mobility of emissive OSCs especially their ambipo-
lar charge mobility, effectively molecular packing or arrangement integrated with suitable HOMO
and LUMO energy levels is indispensable. Ma and coworkers introduced the cyano groups into
trans-­1,4-­distyrylbenzene (trans-­DSB) and produced the molecule 1,4-­bis(2-­cyano-­2-­phenylethenyl)
benzene (β-­CNDSB) with AIE activity (Figure  16.15a)  [42]. In the SCs, the molecule exhibited
nearly planar conformation and adopted J-­aggregation packing mode, resulting in the high lumi-
nescence efficiency. The formation of 2D SC was thought to be driven by the strong π–π interac-
tions and multiple hydrogen bonds and beneficial for charge carrier transporting. In the slice-­like
SCs, the intense green light with a peak around 520 nm showed high PLQY up to 75%. For the β-­
CNDSB SC, the simultaneous injection and transportation of both electron and holes had been
obtained from the OFETs with symmetric gold electrodes in the previous work [43]. However, the
electro-­driven light emission has not been observed from the devices. The high injection barrier
between LUMO level and the WF of Au might limit the electron injection and transport in the
channel. To reduce the injection barrier for electron without interfering the hole injection effi-
ciency, OFETs with calcium/gold asymmetric electrodes geometry was fabricated. Ca’s WF
(−2.87 eV) was closer to the LUMO level of β-­CNDSB (−3.26 eV) (Figure 16.15b). As a result, the
accumulated threshold voltage for the N-­channel in these devices reduces from more than 30 V in

(a) NC (b)
–2.87
–3 Ca –3.26

CN LUMO
Energy level (eV)

β-CNDSB –4
β-CNDSB

Ca Au –5.10
Crystal –5
Au
Dielectric

Si (Gate) –6.01
–6
HOMO

(c) (d)
10–5 VDS = 130 V 10–5 VDS = –110 V
3 × 10–3 3 × 10–3
10–6
10–6
(IDS)½ (A)½

(IDS)½ (A)½

2 × 10–3 10–7 2 × 10–3


IDS (A)

IDS (A)

10–7
10–8
1 × 10–3 1 × 10–3
10–8
10–9

10–9 0
0
0 30 60 90 120 0 –30 –60 –90 –120 –150
VGS (V) VGS (V)

Figure 16.15  (a) Schematic representation of the β-­CNDSB single-­crystal device; (b) energy level diagram
of β-­CNDSB, comparing Au and Ca’s work function; (c) transfer curve of N-­channel; (d) transfer curve of the
P-­channel. W = 100 μm, L = 50 μm. Source: Adapted from Ref. [42] with permission; Copyright Royal Society
of Chemistry.
  ­Reference 505

the symmetric gold electrodes system to less than 20 V when the asymmetric Ca–Au electrodes
system was employed. Though the energy map suggested that it might have ohmic contact in the
crystal/metal interface, the curves show nonlinear I–V relation as the VDS was low, which indicated
there still existed some injection barrier. This barrier might result from the slightly oxidation of the
calcium. The results showed the highest electron and hole mobility of 2.50 and 2.10  cm2/V/s,
respectively. Among the 26 working devices, the average mobility is 0.526 and 0.232 cm2/V/s for
electron and hole, respectively (Figure 16.15c, d). This work demonstrated that AIE active materi-
als could not only achieve high luminescence, but also be used in FETs and achieve very high
mobility.

16.5  ­Conclusion and Perspective

Despite the great achievement of materials with high PL quantum yields and materials with high
charge mobilities, materials that exhibit both semiconducting and light-­emissive properties are
still rare. However, such kinds of dual-­functional materials are not only important for fundamen-
tal research, but also have significant applications in advanced optoelectronics such as OLETs
and organic lasers. Strong intermolecular interactions and dense packing are commonly thought
as the requirements to generate high charge mobility, which however, usually cause severe lumi-
nescence quenching effect of luminogens. To circumvent this dilemma, lots of strategies were
explored including manipulation of the molecular arrangements to form J-­aggregation, introduc-
tion of suitable conjugation linkers such as double bonds or triple bonds to extend the π-­
conjugation while avoiding strong π–π interactions, introduction of alkyl chains to avoid strong
π–π stacking, and introduction of luminogens with AIE property. Among these strategies, lumi-
nogens with AIE property provide an excellent choice to constructing highly efficient light-­
emitting semiconductors. The strong solid emission of AIEgens is particularly suitable to
construct solid-­emitting optical devices. In addition, as a general platform to develop light-­
emitting materials, AIE is well-­compatible with other molecular design strategies to develop
novel functional materials. In this chapter, we presented some typical molecular semiconductors
with both high charge mobility and luminescence efficiency, in particular, the potential of AIE in
constructing materials with both semiconducting and luminescence properties is highlighted. In
addition, the applications of these materials in OFET and OLET have been introduced and the
factors that influence the device performance have also been discussed. In general, for materials
to rationally tune the intermolecular interactions and arrangements is still challenging. Although
the performance of p-­type light-­emitting semiconductors has made great achievements in recent
years, the performance of n-­type and ambipolar is still much lagged. With the development of
strategies to regulate molecular alignments or self-­assembly, the performance of materials still
has large room to improve. And for devices, through morphological control, interface modifica-
tion and optimizing the SC and TF fabrication techniques, there is bigger room to improve the
performance of the devices.

­References

1 Huang, W., Mi, B. X., Gao, Z. Q. (2010). Organic Electronics. Beijing: Science Press.
Liu, Z., Zhang, G., Zhang, D. (2016). Molecular materials that can both emit light and conduct
2
charges: strategies and perspectives. Chem. Eur. J. 22 (2): 462–471.
506 16  Aggregation-­induced Emission Luminogens with Both High-­luminescence Efficiency and Charge Mobility

3 Zhang, X., Dong, H., Hu, W. (2018). Organic semiconductor single crystals for electronics and
photonics. Adv. Mater. 30 (44): 1801048 (1–34).
4 Zhong, C., Duan, C., Huang, F. et al. (2011). Materials and devices toward fully solution
processable organic light-­emitting diodes. Chem. Mater. 23 (3): 326–340.
5 Ostroverkhova, O. (2016). Organic optoelectronic materials: mechanisms and applications. Chem.
Rev. 116 (22): 13279–13412.
6 Muccini, M. (2006). A bright future for organic field-­effect transistors. Nat. Mater. 5: 605–613.
7 Xie, L., Li, W., Huang, W. et al. (2013). Progress of organic light-­emitting transistors. Chin. Sci. Bull.
58 (19): 1817–1832.
8 Zhang, C., Chen, P., Hu, W. (2016). Organic light-­emitting transistors: materials, device
configurations, and operations. Small 12 (10): 1252–1294.
9 Lou, Z., Du, W., Qi, J. et al. (2008). Research progress of organic light emitting transistors.
Semiconductor or Optoelectronics 29: 313–318.
10 Menard, E., Podzorov, V., Hur, S. H. et al. (2004). High-­performance n-­and p-­type single-­crystal
organic transistors with free-­space gate dielectrics. Adv. Mater. 16 (23): 2097–2101.
11 Takeya, J., Nishikawa, T., Takenobu, T. et al. (2004). Effects of polarized organosilane self-­assembled
monolayers on organic single-­crystal field-­effect transistors. Appl. Phys. Lett. 85 (21): 5078–5080.
12 Forster, T., Kasper, K. (1955). Ein konzentrationsumschlag der fluoreszenz des pyrens.
Z. Elektrochem. Angew. Phys. Chem. 59: 976–980.
13 Najafov, H., Lee, B., Zhou, Q. et al. (2010). Observation of long-­range exciton diffusion in highly
ordered organic semiconductors. Nat. Mater. 9 (11): 938–943.
14 Smith, M. B., Michl, J. (2010). Singlet fission. Chem. Rev. 110: 6891–6936.
15 Cicoira, F., Santato, C. (2007). Organic light emitting field effect transistors: advances and
perspectives. Adv. Funct. Mater. 17 (17): 3421–3434.
16 Sirringhaus, H., Zaumseil, J. (2007). Electron and ambipolar transport in organic field-­effect
transistors. Chem. Rev. 107: 1296–1323.
17 Luo, J., Xie, Z., Lam, J. W. Y. et al. (2001). Aggregation-­induced emission of 1-­methyl-­1,2,3,4,5-­
pentaphenylsilole. Chem. Commun. 381: 1740–1741.
18 Mei, J., Leung, N. L., Kwok, R. T. et al. (2015). Aggregation-­induced emission: together we shine,
united we soar! Chem. Rev. 115: 11718–11940.
19 Nie, H., Huang, J., Zhao, Z. et al. (2016). Aggregation-­Induced Emission Luminogens (AIEgens) for
Non-­Doped Organic Light-­Emitting Diodes. Aggregation-­Induced Emission: Materials and
Applications. Washington, DC: American Chemical Society.
20 Dong, Y., Lam, J. W. Y., Qin, A. et al. (2007). Aggregation-­induced emissions of tetraphenylethene
derivatives and their utilities as chemical vapor sensors and in organic light-­emitting diodes. Appl.
Phys. Lett. 91 (1): 011111 (1–3).
21 Zhao, Z., Li, Z., Lam, J. W. Y. et al. (2011). High hole mobility of 1,2-­bis[4′-­(diphenylamino)
biphenyl-­4-­yl]-­1,2-­diphenylethene in field effect transistor. Chem. Commun. 47: 6924–6926.
22 Maldonado, J. L., Ramos-­Ortíz, G., Meneses-­Nava, M. A. et al. (2007). Effect of doping with C60 on
photocurrent and hole mobility in polymer composites measured by using the time-­of-­flight
technique. Opt. Mater. 29 (7): 821–826.
23 Xie, Z., Yang, B., Liu, L. et al. (2013). J-­type dipole stacking and strong π–π interactions in the
crystals of distyrylbenzene derivatives: the crystal structures, high luminescence properties and
prediction of high mobility. J. Mol. Eng. Mater. 1 (03): 1340002 (1–13).
24 Liu, J., Zhang, H., Dong, H. et al. (2015). High mobility emissive organic semiconductor. Nat.
Commun. 6: 10032 (1–8).
  ­Reference 507

25 Dadvand, A., Moiseev, A. G., Sawabe, K. et al. (2012). Maximizing field-­effect mobility and
solid-­state luminescence in organic semiconductors. Angew. Chem. Int. Ed. 51 (16): 3837–3841.
26 Higgins, R. W. T., Monkman, A. P., Nothofer, H. G. et al. (2001). Effects of singlet and triplet energy
transfer to molecular dopants in polymer light-­emitting diodes and their usefulness in chromaticity
tuning. Appl. Phys. Lett. 79 (6): 857–859.
27 Stegmaier, K., Fleissner, A., Janning, H. et al. (2011). Influence of electrical fatigue on hole
transport in poly(p-­phenylenevinylene)-­based organic light-­emitting diodes. J. Appl. Phys. 110 (3):
034507 (1–9).
28 Ju, H., Wang, K., Zhang, J. et al. (2017). 1,6-­and 2,7-­trans-­β-­styryl substituted pyrenes
exhibiting both emissive and semiconducting properties in the solid state. Chem. Mater. 29 (8):
3580–3588.
29 Zhao, Z., Su, H., Zhang, P. et al. (2017). Polyyne bridged AIE luminogens with red emission:
design, synthesis, properties and applications. J. Mater. Chem. B. 5 (8): 1650–1657.
30 Tao, J., Liu, D., Qin, Z. et al. (2020). Organic UV-­sensitive phototransistors based on
distriphenylamineethynylpyrene derivatives with ultra-­high detectivity approaching 1018. Adv.
Mater. 32: 1907791 (1–9).
31 Liang, X., Tan, L., Liu, Z. et al. (2017). Poly(naphthalene diimide) vinylene: solid state red emission
and semiconducting properties for transistors. Chem. Commun. 53 (36): 4934–4937.
32 Zhao, Z., Chen, C., Wu, W. et al. (2019). Highly efficient photothermal nanoagent achieved by
harvesting energy via excited-­state intramolecular motion within nanoparticles. Nat. Commun. 10
(1): 768 (1–11).
33 Guo, Y., Li, Y., Awartani, O. et al. (2016). A vinylene-­bridged perylenediimide-­based polymeric
acceptor enabling efficient all-­polymer solar cells processed under ambient conditions. Adv. Mater.
28 (38): 8483–8489.
34 Chen, Y., Liang, X., Yang, H. et al. (2019). Strong near-­infrared solid emission and enhanced n-­type
mobility for poly(naphthalene diimide) vinylene by a random polymerization strategy.
Macromolecules 52 (21): 8332–8338.
35 Li, Y., Zhang, G., Zhang, W. et al. (2014). Arylacetylene-­substituted naphthalene diimides with
dual functions: optical waveguides and n-­type semiconductors. Chem. Asian. J. 9 (11):
3207–3214.
36 Zhao, Z., Gao, S., Zheng, X. et al. (2018). Rational design of perylenediimide-­substituted
triphenylethylene to electron transporting aggregation-­induced emission luminogens (AIEgens)
with high mobility and near-­infrared emission. Adv. Funct. Mater. 28 (11): 1705609 (1–9).
37 Zhao, Q., Zhang, S., Liu, Y. et al. (2012). Tetraphenylethenyl-­modified perylene bisimide:
aggregation-­induced red emission, electrochemical properties and ordered microstructures.
J. Mater. Chem. 22 (15): 7387–7394.
38 Zhao, Z., Lam, J. W. Y., Chan, C. Y. et al. (2011). Stereoselective synthesis, efficient light emission,
and high bipolar charge mobility of chiasmatic luminogens. Adv. Mater. 23 (45): 5430–5435.
39 Ma, S., Zhou, K., Hu, M. et al. (2018). Integrating efficient optical gain in high-­mobility organic
semiconductors for multifunctional optoelectronic applications. Adv. Funct. Mater. 28 (36):
1802454 (1–8).
40 Hestand, N. J., Spano, F. C. (2017). Molecular aggregate photophysics beyond the kasha model:
novel design principles for organic materials. Acc. Chem. Res. 50 (2): 341–350.
41 Li, J., Zhou, K., Liu, J. et al. (2017). Aromatic extension at 2,6-­positions of anthracene toward an
elegant strategy for organic semiconductors with efficient charge transport and strong solid state
emission. J. Am. Chem. Soc. 139 (48): 17261–17264.
508 16  Aggregation-­induced Emission Luminogens with Both High-­luminescence Efficiency and Charge Mobility

42 Deng, J., Xu, Y., Liu, L. et al. (2016). An ambipolar organic field-­effect transistor based on an
AIE-­active single crystal with a high-­mobility level of 2.0 cm2 V−1 s−1. Chem. Commun. 52 (11):
2370–2373.
43 Deng, J., Tang, J., Xu, Y. et al. (2015). Cyano-­substituted oligo(p-­phenylene vinylene) single-­crystal
with balanced hole and electron injection and transport for ambipolar field-­effect transistors. Phys.
Chem. Chem. Phys. 17 (5): 3421–3425.
509

17

Morphology Modulation of Aggregation-­induced Emission:


From Thermodynamic Self-­assembly to Kinetic Controlling
Kaizhi Gu, Chenxu Yan, Zhiqian Guo, and Wei-­Hong Zhu
Shanghai Key Laboratory of Functional Materials Chemistry, Key Laboratory for Advanced Materials and Institute of Fine Chemicals,
Joint International Research Laboratory of Precision Chemistry and Molecular Engineering, Feringa Nobel Prize Scientist Joint Research
Center, School of Chemistry and Molecular Engineering, East China University of Science & Technology, Shanghai, China

17.1 ­Introduction

As a branch of biophotonics research, optical imaging is an ideal technique for real-­time and non-
invasive detection of important biogenic species in vitro and in vivo, providing information about
biological processes, pathological pathways, therapeutic effects, etc. [1–9]. In fact, luminescence-­
based technique has been made a perfect choice for biomedical and preclinical applications
because it enjoys high-­temporal-­spatial resolution, excellent sensitivity, simplicity, and low
cost [10–13]. To perform high-­quality fluorescence imaging, a large variety of exogenous contrast
agents, including organic dyes, organic-­dye-­doped nanoparticles (NPs), organic semiconducting
materials, rare-­earth metal ion-­doped upconversion materials, quantum dots, and fluorescent pro-
teins, have been developed and used [14–23]. Evidently, in terms of synthetic flexibility, structural
modification, chromatic tunability, and biocompatibility, organic dyes, and organic-­dye-­doped
NPs have been at the center of attention.
However, traditional organic dyes often suffer from enrichment-­quenching effect because of the
intense intermolecular π–π stacking, denoted as aggregation-­caused quenching (ACQ), which has
been documented in Birks’ classic book on Photophysics of Aromatic Molecules [24]. This intrinsic
effect greatly hinders their application in biosensing and bioimaging. Hydrophobic organic dyes
are incompatible with water and naturally form aggregates in a hydrophilic medium, making them
unsuitable for high-­performance biosensing and optical imaging in the living system. Furthermore,
ACQ dyes usually cannot be used at high concentrations, and dye molecules at such diluted con-
centrations could easily be photobleached under excitation light irradiation, compromising their
ability for long-­term imaging.
Aggregation-­induced emission (AIE), a concept proposed by Tang et al. in 2001, is a unique pho-
tophysical phenomenon shown by a group of luminogenic materials that are nonemissive in dilute
solution but become highly luminescent in poor solvents or solid state as aggregates/clusters [25].
The restriction of intramolecular motions (RIM) has been proven to be the main cause of the AIE
phenomenon (Scheme 17.1) [26–28]. The isolated molecules of AIE-­active luminogens (AIEgens)
in a good solution can undergo active RIM, including rotations and vibrations, which consume the

Handbook of Aggregation-Induced Emission: Volume 1 Tutorial Lectures and Mechanism Studies, First Edition.
Edited by Youhong Tang and Ben Zhong Tang.
© 2022 John Wiley & Sons Ltd. Published 2022 by John Wiley & Sons Ltd.
510 17  Morphology Modulation of Aggregation-­induced Emission: From Thermodynamic Self-­assembly to Kinetic Controlling

Aggregation

Restriction of
intramolecular
rotation (RIR)
Dynamic rotation
TPE in solution; Diphenylmethylene
(DPM) intermediate RIR in aggregate;
non-emissive Restriction of
highly emissive
intramolecular
motion (RIM)

Aggregation

Restriction of
intramolecular
Active vibration vibration (RIV)
and rotation Ethylene-tethered and rotation (RIR)
THBA (1) in solution; DPM intermediate RIV and RIR in aggregate;
non-emissive (* = ·, +, –) highly emissive

Scheme 17.1  Schematic illustration showing the mechanisms of the AIE phenomenon. Source: Ref. [26].
Reproduced with permission of American Chemical Society.

excited-­state energy via nonradiative decay. In aggregate state, however, the RIM is greatly
restricted, which blocks the radiationless decay channel and populates the radiative decay path-
way. In addition, the twisted three-­dimensional configurations of AIEgens also inhibit π–π interac-
tion effectively. In this regard, AIEgens are suitable as contrast agents in biomedical imaging
applications: (i) once bound with biomolecules or influenced by the surrounding bioenvironment,
the intramolecular motion of AIEgens is greatly restricted, giving bright fluorescent signals from
united AIE molecules and enabling visualization and imaging of bioanalytes and biological pro-
cesses [29, 30]; (ii) AIEgens are “lit up” upon spontaneously aggregated in a hydrophilic bioenvi-
ronment or artificially synthesized to fluorescent NPs, in which the AIE-­featured NPs’ brightness
can be controllably tuned by changing the loading ratio or aggregation degree of AIEgens inside to
improve the contrast and fidelity of imaging [31–33]; and (iii) due to the effective protection from
the shell of NPs, the encapsulated AIEgens shows appreciable resistance to photobleaching under
light irradiation, permitting the long-­term tracking of dynamic biological processes.
Although AIE bioprobes and bioimaging have experienced an exponential growth of research
interest in the field of biomedical engineering and life science  [34–36], a huge challenge still
remains in controllable aggregation and morphology modulation of AIE molecules, as the size and
shape of AIE materials play a vital role on their performance in applications. Mapping the nature
of AIEgens to finely control the morphologies and sizes of organic-­aggregated nanostructures is
uncharted territory, especially the influence of substituents on the shape, and therefore, the excel-
lent optical properties for biosensing and bioimaging. Here, we focus on the bottlenecks and reso-
lutions of AIEgens in aggregation and morphology modulation of AIE molecules, and this chapter
is arranged in four sections: aggregation modulation of AIE bioprobes via hydrophilicity improve-
ment, thermodynamic self-­assembly of AIE materials, morphology tuning of AIE nanoaggregates,
and kinetic-­driven preparation of AIE NPs. Some representative works, such as enzyme activity
monitoring, protein tracking, and in vivo tumor-­specific bioimaging are also illustrated. Finally,
we summarize the challenges and opportunities for AIE research in the future.
17.2  ­Aggregation Modulation of AIE Bioprobes via Hydrophilicity Improvemen 511

17.2 ­Aggregation Modulation of AIE Bioprobes


via Hydrophilicity Improvement

An ideal fluorescent probe for biosensing and bioimaging should provide high signal-­to-­noise ratio
in a reasonable time period with the least possibility to interfere with the biological sample [37–39].
Unfortunately, hydrophobic organic AIE-­active probes are incompatible with water and naturally
form aggregates under physiological conditions. This undesirable initial aggregation before the
AIE-­active probes binding to the target of interest inevitably induces a “false-­positive” fluorescent
signal, making them unsatisfactory for biosensing and bioimaging with high signal-­to-­noise ratio.
Therefore, new strategies are needed for improving water-­solubility of AIE bioprobes so that they
can maximize the accuracy of sensing signal and the quality of bioimaging.

17.2.1  Molecular Modification


Incorporating a hydrophilic functional group, such as sulfonate group, saccharide, phosphonate
group  [40], and peptide, into AIEgen via molecular modification is an effective approach to
improve water solubility of AIE-­active probes. In a seminal work in this vein, Tang’s group designed
and synthesized a fluorescent off–on probe Ac-­DEVDK-­TPE with AIE characteristics for a real-­
time cell apoptosis imaging and drug screening (Figure 17.1) [41]. The probe is comprised of three
parts: (i) a hydrophilic Asp-­Glu-­Val-­Asp (DEVD) peptide that can be specifically cleaved by
caspase-­3/-­7 and able to improve the probe’s water solubility; (ii) a hydrophobic and AIE-­active
tetraphenylethene (TPE) chromophore as reporting unit; (iii) an alkyne-­functionalized lysine unit
as a linker to connect DEVD and TPE moieties. Ac-­DEVDK-­TPE is highly water-­soluble and dis-
plays almost nonfluorescent in an aqueous solution. After being internalized into apoptotic cells
and activated by caspase-­3/-­7, highly specific and efficient enzymatic catalysis would release the
hydrophobic TPE residues. Then, the molecular aggregation will give rise to strong emission oper-
ated by the AIE mechanism. Notably, the significant fluorescence off–on switch enables real-­time
monitoring of caspase-­3/-­7 activities both in solutions and in living cells with a high signal-­to-­
noise ratio, and further allows in situ screening and quantification of apoptosis-­inducing agents.
This design strategy of AIE probe may open a new way for construction of bioprobes with high
signal-­to-­noise ratio.
Incorporating sulfonate unit as a hydrophilic functional group to organic fluorophores is a typical
strategy to realize efficiently water-­soluble bioprobes. Herein, Zhu et al. reported two hydrophilic AIE-­
active probes, EDS and EDPS (Figure 17.11a), by decorating AIE-­building block quinoline-­malononitrile
(QM) with sulfonate group in different substitution positions (Figure  17.2)  [43]. Apart from the
improvement of hydrophilicity for probes, the sulfonate unit is also served as a negatively charged car-
rier for electrostatic interaction. The substitution position effect of sulfonate group plays essential roles
in the hydrophilicity, emitting color, and specific AIE characteristics. EDPS exhibits a typical AIE prop-
erty with increasing water fractions in a mixture of water/ethanol system. Surprisingly, EDS is almost
nonemissive in aqueous solution. However, the emission of EDS increases quickly and monotonously
when the volume fraction of ethanol is higher than 70 vol%. The dynamic light scattering and transmis-
sion electron microscopy (TEM) results reveal that EDS is capsule-­like with loosely packing character-
istic in pure water, which still has a large internal free volume to consume energy, resulting in nearly no
fluorescence. EDS shows strong red emission in the presence of bovine serum albumin (BSA) due to
the formation of tightly packing aggregates between EDS and BSA. Specifically, EDS is successfully
explored as a label-­free real-­time AIE fluorescent assay for trypsin detection.
Figure 17.1  (a) Illustration of Ac-­DEVDK-­TPE for caspase activities study; (b) emission spectra of TPE-­N3 and Ac-­DEVDK-­TPE in DMSO/water (v/v = 1 : 199);
(c) emission spectra of Ac-­DEVDK-­TPE upon treatment with caspase-­3 and -­7 in the presence and absence of inhibitor 5-­[(S)-­(+)-­2-­(methoxymethyl)
pyrrolidino]sulfonylisatin; (d) fluorescence microscope images of Ac-­DEVDK-­TPE-­pre-­incubated MCF-­7 cells (human breast cancer cells) upon treatment with
1 μM each of DMSO, sodium ascorbate (Na asb), cisplatin, and staurosporine (STS). Source: Ref. [41]. Reproduced with permission of American Chemical Society.
17.2  ­Aggregation Modulation of AIE Bioprobes via Hydrophilicity Improvemen 513

Figure 17.2  (a) Molecular structure of EDS and EDPS; (b) emission spectra and I/I0 plot of EDS in mixture
of water/ethanol with different ethanol fractions; (c) emission spectra and I/I0 plot of EDPS in mixture of
ethanol/water with different water fractions; (d) emission spectra of EDS (10.0 μM) in the presence of

room temperature. λex = 460 nm; (e) proposed mechanism for the interaction of EDS and BSA, and its
different concentration of BSA (0.0–800.0 μg/ml) in PBS buffer solution (10.0 mM, pH 7.0) incubated at

disassembly in the presence of trypsin. Source: Ref. [43]. Reproduced with permission of Royal Society of
Chemistry.

Water-­soluble fluorescent AIE-­active materials are in great demand for high-­contrast imaging of
mitochondrial dynamic activity. However, the real-­time monitoring of the mitochondria in living cells
is restricted by the lack of high-­fidelity molecular sensors. Zhu et al. present a rational design strategy
of a novel AIE molecular motif, investigate the structure–function relationship of substituent groups,
and focus on how to modulate the hydrophilicity and the aggregated state for realizing “off–on” fluo-
rescence, as well as matching the charge density to go across the cell membrane for the mitochondria
targeting [44]. Optimized by substitution positions of hydrophilic triphenylphosphine cation unit on
AIE-­active red-­emitter building block of tricyano-­methylene-­pyridine (TCM), the ortho-­substituted
TCM-­1 probe exhibits the unprecedented high-­fidelity spatial-­temporal tracking of mitochondria in
three-­dimensional localization imaging, “off–on” near-­infrared (NIR) property, high-­targeting capac-
ity, good biocompatibility, and excellent photostability (Figure  17.3). Furthermore, the aggregated
TCM-­1 in cancer cell mitochondria can generate a high concentration of reactive oxygen species in
situ for image-­guided photodynamic therapy.
514 17  Morphology Modulation of Aggregation-­induced Emission: From Thermodynamic Self-­assembly to Kinetic Controlling

Figure 17.3  (a) Molecular structure of TCM-­1, TCM-­2, and TCM-­3; (b) emission spectra and I/I0 plot of
TCM-­1, TCM-­2, and TCM-­3 in 99% water under the same excitation conditions; (c) emission spectra and I/I0
plot of TCM-­1 in mixture of water/THF with different THF fractions; (d) confocal images of HeLa cells
incubation with TCM-­1 and MTG (mito-­tracker green), white arrow indicates the aggregation of TCM-­1 in

MTG, green channels of MTG (λex = 488 nm, λem = 500–550 nm); red channels of TCM-­1 (λex = 561 nm,
mitochondria; (e) cell slices of HeLa cells by Leica TCS SP8 (63 × oil lens) with incubation of TCM-­1 and

λem = 640–720 nm). Source: Ref. [44]. Reproduced with permission of John Wiley & Sons, Inc.
17.2  ­Aggregation Modulation of AIE Bioprobes via Hydrophilicity Improvemen 515

In another AIE-­active probe QM-­βgal [45], Zhu et al. combined a highly hydrophilic galactose
moiety and a hydrophobic QM fluorogen (QM-­OH), in which the galactose moiety not only endows
the probe with water solubility but also acts as the specifically enzyme-­responsive trigger
(Figure 17.4). QM-­βgal is highly hydrophilic and shows negligible fluorescence in aqueous media
because of the consumption of excitonic energy through intermolecular motions, which can mini-
mize the interference of the background. Furthermore, the intact probe is molecular status in a
physiological environment and displays good cell membrane permeability. But, when activated by
endogenous β-­gal (a biomarker of primary ovarian cancers and cell senescence) in the ovarian
cancer cells, the glycosidic bond will be specifically cleaved. The hydrophobic AIEgen QM-­OH will
be released, and then highly fluorescent nanoaggregates are in situ generated according to the AIE
mechanism, permitting on-­site sensing of β-­gal activity in the living cells. Significantly,
the  improved intracellular retention of QM-­OH nanoaggregates allows long-­term tracking of
β-­gal-­overexpressing ovarian cancer cells with high temporal–spatial resolution. These results may
indicate that the exocytosis rate of NPs is slower than that of the small molecules and decreases
with the increasing size of the NP. This study provides a powerful tool for investigating the biologi-
cal functions of enzymes and an enzyme-­regulated liberation strategy for the development of AIE
probes with high fidelity.
Recently, based on the traditional laser dye dicyanomethylene-­4H-­pyran (DCM), Zhu et  al.
adopted a “step-­by-­step” rational design strategy to exploit a NIR AIE-­active probe, QM-­FN-­SO3,
for meeting the high-­fidelity requirements to detect amyloid-­β (Aβ) plaques in vivo (Figure 17.5) [46].
This molecular design strategy is mainly embodied in three parts: (i) introducing a lipophilic π-­
conjugated thiophene-­bridge for extending the emission to the NIR region with the enhancement
of blood–brain barrier penetrability; (ii) replacing the heterocyclic oxygen atom with the nitrogen
atom, achieving the transformation from ACQ to AIE building block; and (iii) decorating the AIE
system with the sulfonate group to increase the hydrophilicity of probe, guaranteeing fluorescence-­
off state before binding to Aβ protein. In vitro and in vivo results demonstrated that QM-­FN-­SO3
successfully addresses the AIE dilemma between the lipophilic requirement for NIR emission and
aggregation behavior from water to protein fibrillogenesis, and enables in situ mapping of Aβ
plaques with ultrasensitivity and high-­fidelity, thus making a breakthrough in high-­fidelity feed-
back on in vivo detection of Aβ plaques with remarkable binding affinity. This work paves a new
avenue for insights into protein fibrillogenesis in vivo and provides a promising strategy for the
design of high-­performance bioprobes.

17.2.2  Polymerization with Hydrophilic Matrix


A complementary method taken by the chemists for increasing water solubility is to graft AIEgens
into hydrophilic matrix through covalent linkage. AIE-­grafted polymeric materials show some
superior advantages in biosensing and bioimaging, such as superb processability, high-­cell perme-
ability, and labeling performance, excellent solubility and appreciable biocompatibility [47, 48]. As
an example, Zhan et  al. fabricated a water soluble AIE-­active perylene-­functionalized polymer
P(PMVP-­co-­MVP) by quaternization of poly(4-­vinylpyridine) with bromomethyl-­perylene and
methyl iodide (Figure  17.6)  [49]. Practically no fluorescence signal is collected from the dilute
aqueous solution of P(PMVP-­co-­MVP). With the increase of THF fractions, however, its fluores-
cence intensity is gradually enhanced because of the aggregation of the polymer. Operated by the
intense electrostatic interactions, the polycation P(PMVP-­co-­MVP) and polyanion single-­stranded
DNA (ssDNA) can generate a complex probe that allows for fluorescence detection of DNA hybrid-
ization. Upon adding noncomplementary ssDNA, the fluorescence of the complex probe obviously
Figure 17.4  (a) Schematic illustration of an enzyme-­regulated liberation strategy for on site sensing and long-­term tracking; (b) molecular structure of
QM-­βgal; (c) time-­dependent fluorescence spectra of QM-­βgal (10 mM) with 6 U β-­gal; (d) time-­dependent confocal fluorescence images of SKOV-­3 cells
incubated with QM-­βgal and control probe DCM-­βgal. Source: Ref. [45]. Reproduced with permission of Royal Society of Chemistry.
17.2  ­Aggregation Modulation of AIE Bioprobes via Hydrophilicity Improvemen 517

Figure 17.5  (a) Rational design of NIR AIE-­active probes for Aβ deposition; (b) in vivo mapping of Aβ
deposition in AD model (APP/PS1 transgenic) mice and ex vivo fluorescence observation of brain slices
from APP/PS1 mice after injection of probe QM-­FN-­SO3. Source: Ref. [46]. Reproduced with permission of
American Chemical Society.
518 17  Morphology Modulation of Aggregation-­induced Emission: From Thermodynamic Self-­assembly to Kinetic Controlling

Figure 17.6  Conceptual illustration of the complex probe for DNA detection based on AIE effect.
Source: Ref. [49]. Reproduced with permission of American Chemical Society.

increase because of the AIE mechanism. Upon adding complementary ssDNA, there will be a little
change of fluorescence due to the combined effects of AIE and duplex-­quenching induced by the
intercalation of perylene into the duplex. This makes the AIE-­active polymer promising candidate
materials as a novel optical biosensor for applications in bioassays.
For the most traditional fluorescent dyes utilized as contrast agents, the incubated cells need to
wash several times before cell imaging because the dyes dispersed in cell culture medium will
cause a strong background interference. To overcome this limitation, Zhao et al. reported a novel
water-­soluble polymer probe P3 that possesses wash-­free cellular imaging capacity based on AIE
characteristics (Figure 17.7) [50]. P3 is facilely prepared via conjugation of TPE and poly(acryloyl
ethylene diamine) (a kind of polyelectrolyte), in which AIE-­active TPE serves as the fluorescent
unit and polyelectrolyte as hydrophilic matrix. This polymer probe shows pH-­responsive and Cd2+-­
responsive photophysical properties simultaneously based to molecular aggregation principle, and
enables wash-­free cellular imaging with high signal-­to-­noise ratio when the work concentration of
probe is up to 50 μg/ml, demonstrating its robust biosensing and bioimaging applications.

ClH3N
OH
COOH
O
NH
n

S O
O S
n

HN
O
HOOC
OH
NH3Cl

Hydrophilic matrix AIE unit Hydrophilic matrix

Figure 17.7  Molecular structure of P3.


17.4 ­Morphology Tuning of AIE Nanoaggregate 519

17.3 ­Thermodynamic Self-­assembly of AIE Materials

Micelle encapsulation by the self-­assembly of AIE molecules and surfactants or amphiphilic block
copolymers is one of the major methods to construct AIE NPs for biomedical imaging. For exam-
ple, Liu’s group described the strategy of using the amphiphilic block copolymers to encapsulate
AIEgens [39]. These obtained AIE nanorods, but not the nanodots, could be selectively internal-
ized into cancer cells, which show better tumor accumulation, higher tumor penetration, and
more efficient in vivo cancer cell uptake. In 2012, Tang’s group proposed an effective strategy that
utilizes a nanostructure of far-­red/NIR AIEgens TPE-­TPA-­DCM encapsulated in BSA as contrast
agents for in vitro and in vivo imaging applications (Figure 17.8) [51]. BSA is utilized as the encap-
sulation matrix for preparing NPs because albumin is natural, biocompatible, nonantigenic, and
clinically approved. The obtained protein NPs show uniform nanosize, bright emission, and low
cytotoxicity, which is beneficial for in vivo and in vitro bioimaging applications. In the fabrication
process of NPs, TPE-­TPA-­DCM molecules aggregate and entangle with the hydrophobic sections of
BSA to afford fluorogen-­loaded BSA NPs, which are further treated with glutaraldehyde to lock the
BSA matrix. After intravenous injection of BSA-­encapsulated NPs, the significant optical signal is
predominately collected on the tumor site due to the enhanced permeability and retention (EPR)
effect, successfully demonstrating the feasibility of encapsulation strategy.
In another work, He et al. expanded this strategy to develop biocompatible and photostable AIE
dots by encapsulating TTF AIEgens within amphiphilic phospholipids carrying a poly(ethylene
glycol) chain (DSPE-­mPEG5000) for in vivo two-­photon bioimaging (Figure 17.9) [52]. In aqueous
solution, the hydrophobic lipid segments of block copolymer tend to be embedded in the hydro-
phobic TTF core, while the hydrophilic PEG chain extends into the aqueous phase, rendering
nanodots with the superior water dispersibility. Moreover, the AIE dots possess excellent biocom-
patibility since the PEG chains have the function of improving the long-­time circulation of NPs
and helping them avoid capture/degradation by reticuloendothelial systems. Capitalizing on their
outstanding two-­photon property, the reported AIE dots enable three-­dimensional dynamic imag-
ing with high resolution in blood vessels of mouse brain by two-­photon fluorescence microscopy.

17.4 ­Morphology Tuning of AIE Nanoaggregates

AIEgens as a kind of promising bioimaging contrast have received considerable attention because
of their fascinating light-­up and bright emission in the aggregated state. However, AIE molecules
tend to randomly aggregate into various irregular shapes with broad particle size distrubution in

NC CN

N N

Figure 17.8  Chemical structure of TPE-­TPA-­DCM.


520 17  Morphology Modulation of Aggregation-­induced Emission: From Thermodynamic Self-­assembly to Kinetic Controlling

Figure 17.9  (a) Molecular structures of the AIEgen TTF and amphiphilic block copolymer; (b) schematic for
the preparation of AIE dots; (c) imaging of ear blood vessels of a mouse using AIE dots: (A) Bright field,
(B) one-­photon confocal and (C) two-­photon luminescence images, (D) merged image; (E) a line scan along
capillaries of a mouse brain was used to observe RBC flow and determine its instantaneous velocity (dx/dT)
with two-­photon confocal microscope. (F) blood circulation kinetics of AIE-­dots in mice. Source: Ref. [52].
Reproduced with permission of Springer Nature.

water so that it is difficult to precise control over the morphology and size of AIE nanoaggregates.
Given that the size, shape, and orientation of dye aggregates play a vital role on the rate, efficiency,
and endocytic mechanism of cellular uptake as well as tumor-­targeting imaging performance [53,
54], the ability to systematically tune the size and morphology of AIE nanoaggregates for optimum
tumor uptake and targeting is highly demanded.
To provide an ideal AIEgen organic nanoprobe for tumor-­targeting bioimaging, Zhu et al. sys-
tematically investigated the shape-­specific effects with various substituent groups in AIE mole-
cules and contributed a series of quinolone-­malononitrile (QM)-­based far-­red and NIR AIE-­active
nanoprobes by essentially tailoring π-­bridge and donor unit in the molecular structures
(Figure  17.10)  [55]. Introduction of strong donor groups in AIE-­active QM building block can
extend molecules’ emission to far-­red or even to NIR region. Importantly, the alteration of π-­bridge
from the thiophene unit to the 3,4-­ethylene-­dioxythiophene (EDOT) unit can controllably change
the nanostructure morphologies from rod like to spherical shape. Considering QM-­2 and QM-­5 are
typical representatives with rod like and spherical shapes, respectively, two nanoprobes are further
examined about their potential shape effects on bioimaging in vitro and in vivo. Cells imaging veri-
fies that compared with commercial ICG contrast agent, QM-­2 and QM-­5 enable track the HeLa
17.4 ­Morphology Tuning of AIE Nanoaggregate 521

Figure 17.10  QM-­based nanoprobes and AIE characteristics: (a) molecular structures and the single crystal
configurations of QM-­1, QM-­2, and QM-­3. Emission spectra and plots of the relative fluorescence intensity

fractions (fw); λex = 480 nm. Inset: Fluorescent images in pure THF solvent and THF/H2O solution of
of QM derivatives: (b) QM-­1, (c) QM-­2, (d) QM-­4, and (e) QM-­5 in THF/H2O mixtures with different water

(b) QM-­1 (fw = 90%), (c) QM-­2 (fw = 70%), (d) QM-­4 (fw = 70%), and (e) QM-­5 (fw = 90%) under 365 nm
illumination. (f) Normalized fluorescent spectra of QM-­1 (fw = 90%), QM-­2 (fw = 70%), QM-­3 (fw = 90%), QM-­4
(fw = 70%), QM-­5 (fw = 90%), and QM-­6 (fw = 90%) in THF/H2O solution (g) Normalized fluorescent spectra of
QM derivatives in the solid state. Source: Ref. [55]. Reproduced with permission of John Wiley & Sons, Inc.
522 17  Morphology Modulation of Aggregation-­induced Emission: From Thermodynamic Self-­assembly to Kinetic Controlling

cells with bright far-­red/NIR fluorescence as long as four passages because QM aggregates with
different shape and size is beneficial to retain themselves in the cells. Moreover, kinetics of the cell
uptake results measured by flow cytometric manifest the initial uptake rate of QM-­5  which is
about five-­fold more that of QM-­2 after incubation at three hours. Notably, the spherical
QM-­5 nanoaggregates with NIR emission show excellent tumor-­targeted bioimaging performance
due to EPR effect, while the rod-­like QM-­2  nanoaggregates do not exhibit any tumor-­targeting
properties. This work provides a promising platform for tumor-­specific bioimaging and guides the
development of AIE-­active nanomaterials with ideal biological geometries for imaging in vivo.
Next, to optimize the morphology of rod-­like QM-­2 aggregates, Zhu and Li et al. adopted encap-
sulation technique to obtain spherical AIEgen-­based hybrid NPs (QM-­2@PNPs) [42]. As shown in
Figure  17.11, QM-­2@PNPs are prepared by encapsulating AIEgen QM-­2  within hybrid micelles
during the self-­assembly of block copolymer polystyrene-­b-­poly (acrylic acid), and stabilized the
micelles by the hydrolysis of organic silane, 3-­mercaptopropyl-­trimethoxysilane (MPTMS). Then,
functionalization with biocompatible methoxypolyethylene glycol maleimide (mPEG-­MAL) onto
the surface of NPs via Michael addition contributes to increase the colloidal stability and reduce the
phagocytic capture of NPs by the immune system. This essential encapsulation approach success-
fully makes the rod like, microsized bare QM-­2 aggregates transform into spherical QM-­2@PNPs

Figure 17.11  (a) Schematic diagram for the fabrication of QM-­2@PNPs. (i) Formation of oil-­in water
hybrid micelles with hydrophobic QM-­2 encapsulated in the cores and the subsequent shell cross-­linking
with MPTMS, introducing thiol groups on the shells (ii) Functionalization with mPEG-­MAL; (b) molecular
structure of QM-­2; (c) QM-­2@PNPs and (d) bare QM-­2 via intravenous injection for 0.5, 2, and 24 hours and
ex vivo fluorescence images of organs after injected with QM-­2@PNPs or QM-­2 for 24 hours; (e) average
fluorescence intensity for tumor and internal organs from mice sacrificed at 24 hours postinjection with
QM-­2@PNPs and bare QM-­2. Source: Ref. [42]. Reproduced with permission of John Wiley & Sons, Inc.
17.5 ­Kinetic-­driven Preparation of AIE NP 523

with uniform nanosize, monodispersion, high stability, and negligible cytotoxicity. Cell-­imaging
experiments suggest that the as prepared QM-­2@PNPs are capable of labeling live MCF-­7 cells as
long as five days with high resolution, corroborating that the spherical QM-­2@PNPs possess excel-
lent retention ability in live cells. Impressively, after delivery into tumor-­bearing mice via tail vein
injection, QM-­2@PNPs display preferential tumor-­targeting ability with brightness fluorescence for
in vivo bioimaging due to the efficient morphology and size tuning from microsized rods to nano-
sized spheres. Therefore, this study provides a versatile route to modulate the morphology and size
of nanoprobes to improve their cellular uptake and tumor-­targeting ability in bioimaging.
Recently, a novel engineering strategy described as flash nanoprecipitation (FNP) was intro-
duced by Zhu and Guo et  al. to finely tune the morphology and size of AIE nanoaggregates
(Figure 17.12) [56]. Specifically, the FNP technique is a kinetically controlled process to prepare
NPs within only one second by using a multi-­inlet vortex mixer system [57]. In this study, hydro-
phobic quinolinemalononitrile derivative ED and QM-­2 are selected as AIEgen-­imaging agents
and different amphiphilic block copolymers, namely, dextran-­b-­polylactide (Dex-­b-­PLA) and
dextran-­b-­polycaprolactone (Dex-­b-­PCL) as polymer matrix. Specifically, AIE-­active molecules are
dissolved in THF in one stream simultaneously mixing with other three streams of water to obtain
bare organic NPs. For example, if without using any polymer by simply mixing a solution of ED in
THF with a large amount of water, it yields rod-­like ED NPs. These bare ED rods prepared by FNP
can be stably preserved for about one week. When different polymers are introduced in one of
water stream with alteration of concentration, NPs with diverse morphology and size (character-
ized by TEM and DLS) are prepared due to the interactions between the AIE molecules and the
hydrophobic segment of polymer.
Cellular uptake results demonstrate that as the aspect ratio of the length and width is about 5,
the nanorods decorated with dextran show better cellular uptake and targeting capability in the
cancer cells. In a zebrafish model, compared with the nanorod without the polymer shell and
spherical NPs with polymer brush, the nanorod particles with dextran shell possess superb tumor-­
targeting capability in vivo. All these results therefore illustrate that the FNP technique together
with optimizing of amphiphilic polymers turns out to be a universal strategy to systematically
modulate the morphology and size of AIE materials, promising for potential applications in pre-
cise bioimaging at specific tumor sites.

17.5 ­Kinetic-­driven Preparation of AIE NPs

The most common method to prepare AIEgen-­based NPs mainly relies on the thermodynamic-­
driven self-­assembly between AIE molecules and amphiphilic polymers [58]. However, this self-­
assembly method often takes a long time to form stable NPs and faces some inherent defects: (i) the
degree of aggregation and morphology is not controllable during the spontaneous process; (ii) the
formed NPs usually exhibit a large size with wide size distribution; and (iii) it is extremely difficult
to scaled-­up preparation of the NPs with the batch-­to-­batch reproducibility, which greatly limit the
biomedical imaging applications of AIEgen NPs. Thus, it is highly desirable to develop a novel way
to scalable fabrication of AIEgen NPs with good reproducibility.
To this end, Zhu and Guo et  al. demonstrated that the kinetically controlled FNP technique,
which is a directed or engineered self-­assembly driven by externally applied forces with vortex
mixer equipment (Figure 17.13), is also suitable for scalable and fast preparation of AIE NPs with
good reproducibility, apart from its favorable function of morphology and size regulation com-
bined with screening of amphiphilic polymers [59]. On the basis of the FNP method, AIE-­active
Figure 17.12  Illustration of morphology and size tuning by FNP and their effects on tumor cell imaging of zebrafish model. Various morphologies and sizes:
(a) bare NP; (b) nanorod with a polymer shell; (c) microrod; and (d) nanosphere. Source: Ref. [56]. Reproduced with permission of American Chemical Society.
17.5 ­Kinetic-­driven Preparation of AIE NP 525

Water
(a) stream 2

Block copolymer

Organic solute
in organic solvent
stream 1
Water
stream 3

Water
AIE-active NPs
stream 4
(b)

20 nm 50 nm

50 nm
100 nm

Figure 17.13  (a) Schematic diagram of the FNP method and (b) TEM images of AIE-­active NPs. Source:
Ref. [59]. Reproduced with permission of American Chemical Society.

fluorescent NPs with a small polydispersity index is facilely constructed in a scale-­up mode, in
which the hydrophobic AIE molecules and the hydrophobic unit of the amphiphilic block copoly-
mers are encapsulated in the core of the core/shell-­type NPs. Changing the formulation parame-
ters in FNP, the size of AIE-­active NPs can be controllably geared from 20 to 60  nm with
high-­performance photophysical properties, which is confirmed by TEM and spectroscopic char-
acterization. Clearly, the kinetic FNP technique is a promising approach to facile and scalable
preparation of AIE NPs with a tunable, uniform, narrow, and submicrometer size distribution.
Recently, Zhu and Guo et al. reported a novel type of saponin-­encapsulated NIR and AIE-­active
NPs with enhanced cell compatibility and permeability by combining rational design strategy with
the scalable FNP method (Figure 17.14) [60]. Based on the late-­model AIE building block of TCM,
NIR-­emitting, and AIE-­active TCMN-­5 having high-­solid fluorescent quantum yield, large Stokes
shift, and superior photostability is obtained by fine tuning the π-­conjugation bridge and electron-­
donating group. Choosing the amphiphilic saponins such as α-­hederin that occurs extensively in
nature as the engineering template and using the FNP technology, the authors massively prepared
526 17  Morphology Modulation of Aggregation-­induced Emission: From Thermodynamic Self-­assembly to Kinetic Controlling

Figure 17.14  (a) Rational design of red/NIR-­emitting AIE molecules based on TCM building block by
alteration of π-­conjugated bridges and donor groups. In D–π–A system, TCM with electron-­withdrawing
tricyano units can be alternatively attached with different bridges (phenyl and thiophene) and electron
donors (triphenylamine and methoxytriphenylamine) to obtain the NIR AIE molecules; (b) preparation of
stable saponin-­based AIE NPs by employing FNP technology for efficiently cellular uptake. THF solution
(stream 1) of AIE molecules and saponin, and water (streams 2, 3, and 4) were mixed in the multi-­inlet
vortex mixer system to prepare AIEgen NPs. (c) The size and polydispersity index of saponin-­based AIE NPs
can be well-­controlled by tuning flow rates of water and THF. Source: Ref. [60]. Reproduced with permission
of American Chemical Society.

the saponin-­encapsulated AIEgen NPs, which exhibit uniform spherical morphology, excellent
colloidal stability, and highly photostability. Moreover, these saponin-­functionalized NPs display
efficient cellular uptake and enable rapid cell imaging because α-­hederin performs biocompatible
amphiphilicity and specific membrane penetration mechanism. Definitely, the kinetic-­oriented
 ­Reference 527

FNP technology cannot only tailor the diameter size and morphology, but also significantly
improve the colloidal stability of NPs when compared to the conventional thermodynamic self-­
assembly approach, which proved to be an effective methodology for scaled-­up fabrication of
nanomaterials in bioimaging applications.

17.6 ­Conclusion and Outlook

The convergence of chemistry, materials science, and biomedicine has facilitated the development of
a large number of versatile AIEgens for biosensing and bioimaging applications. The emergence of
AIEgens has thoroughly overcome the ACQ effect of traditional fluorescent dyes and provided new
opportunities for in situ sensing and in vivo imaging. By integrating recognition or target-­responsive
motifs, AIEgens have been developed into smart activatable bioprobes to correlate signals with the
level or activity of disease-­specific biomarkers. Furthermore, artificially synthesizing AIE NPs with
bright emission and biocompatibility is highly suitable for biomedical imaging. Here, we mainly
focused on aggregation modulation of AIE bioprobes via hydrophilicity improvement, thermody-
namic self-­assembly of AIE materials, morphology tuning of AIE nanoaggregates, and kinetic-­driven
preparation of AIE NPs and demonstrated with representative examples. It is hoped that this chapter
will provide insight into how high-­performance AIE materials can be developed combined with vari-
ous strategies such as structural modifications, micelle encapsulation, and FNP.
In the future, there is still much room to improve performances of AIEgens to promote the devel-
opment of AIE probes to a new height. We conceive that the following aspects but not limited
could be considered: (i) Extending absorption and emission wavelength of AIEgens. Owing to
advances in reducing photon scattering, light absorption, and autofluorescence, NIR or second
NIR (NIR-­II) photon affords deeper-­tissue bioimaging, which is conducive for high-­imaging reso-
lution and contrast. (ii) Boosting fluorescent quantum efficiency of AIEgens. Highly bright
AIEgens or AIE NPs is favorable for performing high-­quality bioimaging, and much efforts should
be devoted to achieve this. (iii) Endowing the AIE-­based bioprobes with multi-­modality. The cur-
rent AIE systems are mainly limited to dual-­modality, such as fluorescence imaging/magnetic
resonance imaging (MRI) and fluorescence imaging/photoacustic (PA). The AIEgens-­based multi-­
modality by combination of fluorescence imaging with PA, MRI, or positron-­emission tomography
still need to be exploited. (iv) Utilizing AIEgens’ room temperature phosphorescence (RTP) for
lifetime imaging. The significant features of lifetime imaging lie in its anti-­interference and zero
background. Lifetime is the inherent nature of fluorophores and is immune to environmental fac-
tors, such as probe concentration and light intensity. Moreover, the autofluorescence from tissues
has shorter lifetime than that of AIEgens with RTP, which can easily be eliminated through “time-­
gate” technology. Definitely, more bran-­new bioprobes based on AIEgens will be developed and act
as a powerful tool to reveal mysterious life processes and address more healthcare issues.

­References

1 Weissleder, R. (2006). Molecular imaging in cancer. Science 312 (5777): 1168–1171.


Guo, Z., Park, S., Yoon, J. et al. (2014). Recent progress in the development of near-­infrared
2
fluorescent probes for bioimaging applications. Chemical Society Reviews 43 (1): 16–29.
3 Aron, A. T., Ramos-­Torres, K. M., Cotruvo, J. A. et al. (2015). Recognition-­and reactivity-­based
fluorescent probes for studying transition metal signaling in living systems. Accounts of Chemical
Research 48 (8): 2434–2442.
528 17  Morphology Modulation of Aggregation-­induced Emission: From Thermodynamic Self-­assembly to Kinetic Controlling

4 Li, J., Pu, K. (2019). Development of organic semiconducting materials for deep-­tissue optical
imaging, phototherapy and photoactivation. Chemical Society Reviews 48 (1): 38–71.
5 Li, B., Lu, L., Zhao, M. et al. (2018). An efficient 1064 nm NIR-­II excitation fluorescent molecular
dye for deep-­tissue high-­resolution dynamic bioimaging. Angewandte Chemie International Edition
57 (25): 7483–7487.
6 Liu, K., Kong, X., Ma, Y. et al. (2017). Rational design of a robust fluorescent probe for the
detection of endogenous carbon monoxide in living zebrafish embryos and mouse tissue.
Angewandte Chemie International Edition 56 (43): 13489–13492.
7 Yin, J., Kwon, Y., Kim, D. et al. (2014). Cyanine-­based fluorescent probe for highly selective
detection of glutathione in cell cultures and live mouse tissues. Journal of the American Chemical
Society 136 (14): 5351–5358.
8 Antaris, A. L., Chen, H., Cheng, K. et al. (2016). A small-­molecule dye for NIR-­II imaging. Nature
Materials 15: 235–242.
9 Gu, Y., Guo, Z., Yuan, W. et al. (2019). High-­sensitivity imaging of time-­domain near-­infrared light
transducer. Nature Photonics 13 (8): 525–531.
10 Chen, X., Zhou, Y., Peng, X. et al. (2010). Fluorescent and colorimetric probes for detection of
thiols. Chemical Society Reviews 39 (6): 2120–2135.
11 Chan, J., Dodani, S. C., Chang, C. J. (2012). Reaction-­based small-­molecule fluorescent probes for
chemoselective bioimaging. Nature Chemistry 4: 973–984.
12 Gu, K., Zhu, W., Peng, X. (2019). Enhancement strategies of targetability, response and
photostability for in vivo bioimaging. Science China Chemistry 62 (2): 189–198.
13 Liu, H., Chen, L., Xu, C. et al. (2018). Recent progresses in small-­molecule enzymatic fluorescent
probes for cancer imaging. Chemical Society Reviews 47 (18): 7140–7180.
14 Owens, E. A., Henary, M., El Fakhri, G. et al. (2016). Tissue-­specific near-­infrared fluorescence
imaging. Accounts of Chemical Research 49 (9): 1731–1740.
15 Wu, D., Sedgwick, A. C., Gunnlaugsson, T. et al. (2017). Fluorescent chemosensors: the past,
present and future. Chemical Society Reviews 46 (23): 7105–7123.
16 Chen, H., Gu, Z., An, H. et al. (2018). Precise nanomedicine for intelligent therapy of cancer.
Science China Chemistry 61 (12): 1503–1552.
17 Zhou, J., Liu, Z., Li, F. (2012). Upconversion nanophosphors for small-­animal imaging. Chemical
Society Reviews 41 (3): 1323–1349.
18 Grimm, J. B., English, B. P., Chen, J. et al. (2015). A general method to improve fluorophores for
live-­cell and single-­molecule microscopy. Nature Methods 12 (3): 244–250.
19 Wu, Y., Huang, S., Wang, J. et al. (2018). Activatable probes for diagnosing and positioning liver injury
and metastatic tumors by multispectral optoacoustic tomography. Nature Communications 9 (1): 3983.
20 Liu, J., Liu, Y., Bu, W. et al. (2014). Ultrasensitive nanosensors based on upconversion
nanoparticles for selective hypoxia imaging in vivo upon near-­infrared excitation. Journal of the
American Chemical Society 136 (27): 9701–9709.
21 Guo, Z., Ma, Y., Liu, Y. et al. (2018). Photocaged prodrug under NIR light-­triggering with dual-­
channel fluorescence: in vivo real-­time tracking for precise drug delivery. Science China Chemistry
61 (10): 1293–1300.
22 Miao, Q., Xie, C., Zhen, X. et al. (2017). Molecular afterglow imaging with bright, biodegradable
polymer nanoparticles. Nature Biotechnology 35 (11): 1102–1110.
23 Zhu, S., Zhang, J., Qiao, C. et al. (2011). Strongly green-­photoluminescent graphene quantum dots
for bioimaging applications. Chemical Communications 47 (24): 6858–6860.
24 Birks, J. B. (1970). Photophysics of Aromatic Molecules, Wiley, London.
 ­Reference 529

25 Luo, J., Xie, Z., Lam, J. W. Y. et al. (2001). Aggregation-­induced emission of 1-­methyl-­1,2,3,4,
5-­pentaphenylsilole. Chemical Communications (18): 1740–1741.
26 Mei, J., Leung, N. L. C., Kwok, R. T. K. et al. (2015). Aggregation-­induced emission: together we
shine, united we soar! Chemical Reviews 115 (21): 11718–11940.
27 Hong, Y., Lam, J. W. Y., Tang, B. Z. (2009). Aggregation-­induced emission: phenomenon,
mechanism and applications. Chemical Communications (29): 4332–4353.
28 Leung, N. L. C., Xie, N., Yuan, W. et al. (2014). Restriction of intramolecular motions: the general
mechanism behind aggregation-­induced emission. Chemistry -­A European Journal 20 (47):
15349–15353.
29 Mei, J., Hong, Y., Lam, J. W. Y. et al. (2014). Aggregation-­induced emission: the whole is more
brilliant than the parts. Advanced Materials 26 (31): 5429–5479.
30 Mei, J., Huang, Y., Tian, H. (2017). Progress and trends in AIE-­based bioprobes: a brief overview.
ACS Applied Materials & Interfaces 10 (15): 12217–12261.
31 Chen, S., Wang, H., Hong, Y. et al. (2016). Fabrication of fluorescent nanoparticles based on AIE
luminogens (AIE dots) and their applications in bioimaging. Materials Horizons 3 (4): 283–293.
32 Nicol, A., Kwok, R. T. K., Chen, C. et al. (2017). Ultrafast delivery of aggregation-­induced emission
nanoparticles and pure organic phosphorescent nanocrystals by saponin encapsulation. Journal of
the American Chemical Society 139 (41): 14792–14799.
33 Wu, W., Mao, D., Hu, F. et al. (2017). A highly efficient and photostable photosensitizer with
near-­infrared aggregation-­induced emission for image-­guided photodynamic anticancer therapy.
Advanced Materials 29 (33): 1700548.
34 Kwok, R. T. K., Leung, C. W. T., Lam, J. W. Y. et al. (2015). Biosensing by luminogens with
aggregation-­induced emission characteristics. Chemical Society Reviews 44 (13): 4228–4238.
35 Qian, J., Tang, B. Z. (2017). AIE luminogens for bioimaging and theranostics: from organelles to
animals. Chem 3 (1): 56–91.
36 Chong, K. C., Hu, F., Liu, B. (2019). AIEgen bioconjugates for specific detection of disease-­related
protein biomarkers. Materials Chemistry Frontiers 3 (1): 12–24.
37 Lyu, Y., Pu, K. (2017). Recent advances of activatable molecular probes based on semiconducting
polymer nanoparticles in sensing and imaging. Advanced Science 4 (6): 1600481.
38 Shi, H., Liu, J., Geng, J. et al. (2012). Specific detection of integrin αvβ3 by light-­up bioprobe with
aggregation-­induced emission characteristics. Journal of the American Chemical Society 134 (23):
9569–9572.
39 Feng, G., Mao, D., Liu, J. et al. (2018). Polymeric nanorods with aggregation-­induced emission
characteristics for enhanced cancer targeting and imaging. Nanoscale 10 (13): 5869–5874.
40 Ji, S., Gao, H., Mu, W. et al. (2017). Enzyme-­instructed self-­assembly leads to the activation of
optical properties for selective fluorescence detection and photodynamic ablation of cancer cells.
Journal of Materials Chemistry B 17 (6): 2566–2573.
41 Shi, H., Kwok, R. T. K., Liu, J. et al. (2012). Real-­time monitoring of cell apoptosis and drug
screening using fluorescent light-­up probe with aggregation-­induced emission characteristics.
Journal of the American Chemical Society 134 (43): 17972–17981.
42 Li, Y., Shao, A., Wang, Y. et al. (2016). Morphology-­tailoring of a red AIEgen from
microsized rods to nanospheres for tumor-­targeted bioimaging. Advanced Materials 28 (16):
3187–3193.
43 Shao, A., Guo, Z., Zhu, S. et al. (2014). Insight into aggregation-­induced emission characteristics of
red-­emissive quinoline-­malononitrile by cell tracking and real-­time trypsin detection. Chemical
Science 5 (4): 1383–1389.
530 17  Morphology Modulation of Aggregation-­induced Emission: From Thermodynamic Self-­assembly to Kinetic Controlling

44 Zhang, J., Wang, Q., Guo, Z. et al. (2019). High-­fidelity trapping of spatial-­temporal mitochondria
with rational design of aggregation-­Induced emission probes. Advanced Functional Materials
29 (16): 1808153.
45 Gu, K., Qiu, W., Guo, Z. et al. (2019). An enzyme-­activatable probe liberating AIEgens: on-­site
sensing and long-­term tracking of β-­galactosidase in ovarian cancer cells. Chemical Science 10 (2):
398–405.
46 Fu, W., Yan, C., Guo, Z. et al. (2019). Rational design of near-­infrared aggregation-­induced-­
emission-­active probes: in situ mapping of amyloid-­β plaques with ultrasensitivity and high-­
fidelity. Journal of the American Chemical Society 141 (7): 3171–3177.
47 Hu, R., Kang, Y., Tang, B. Z. (2016). Recent advances in AIE polymers. Polymer Journal 48:
359–370.
48 Hu, Y. B., Lam, J. W. Y., Tang, B. Z. (2019). Recent progress in AIE-­active polymers. Chinese Journal
of Polymer Science 37 (4): 289–301.
49 Wang, G., Zhang, R., Xu, C. et al. (2014). Fluorescence detection of DNA hybridization based on
the aggregation-­induced emission of a perylene-­functionalized polymer. ACS Applied Materials &
Interfaces 6 (14): 11136–11141.
50 Qian, Y., Liu, H., Tan, H. et al. (2017). A novel water-­soluble fluorescence probe with wash-­free
cellular imaging capacity based on AIE characteristics. Macromolecular Rapid Communications
38 (10): 1600684.
51 Qin, W., Ding, D., Liu, J. et al. (2012). Biocompatible nanoparticles with aggregation-­induced
emission characteristics as far-­red/near-­infrared fluorescent bioprobes for in vitro and in vivo
imaging applications. Advanced Functional Materials 22 (4): 771–779.
52 Wang, D., Qian, J., Qin, W. et al. (2014). Biocompatible and photostable AIE dots with red emission
for in vivo two-­photon bioimaging. Scientific Reports 4 (1): 4279.
53 Geng, Y., Dalhaimer, P., Cai, S. et al. (2007). Shape effects of filaments versus spherical particles in
flow and drug delivery. Nature Nanotechnology 2: 249–255.
54 Chithrani, B. D., Chan, W. C. W. (2007). Elucidating the mechanism of cellular uptake and removal
of protein-­coated gold nanoparticles of different sizes and shapes. Nano Letters 7 (6): 1542–1550.
55 Shao, A., Xie, Y., Zhu, S. et al. (2015). Far-­red and near-­IR AIE-­active fluorescent organic
nanoprobes with enhanced tumor-­targeting efficacy: shape-­specific effects. Angewandte Chemie
International Edition 54 (25): 7275–7280.
56 Wang, M., Xu, Y., Liu, Y. et al. (2018). Morphology tuning of aggregation-­induced emission probes
by flash nanoprecipitation: shape and size effects on in vivo imaging. ACS Applied Materials &
Interfaces 10 (30): 25186–25193.
57 Akbulut, M., Ginart, P., Gindy, M. E. et al. (2009). Generic method of preparing multifunctional
fluorescent nanoparticles using flash nanoprecipitation. Advanced Functional Materials 19 (5):
718–725.
58 Prud Homme, R. K., Johnson, B. K. (2003). Mechanism for rapid self-­assembly of block copolymer
nanoparticles. Physical Review Letters 91 (11): 118302.
59 Wang, M., Yang, N., Guo, Z. et al. (2015). Facile preparation of AIE-­active fluorescent nanoparticles
through flash nanoprecipitation. Industrial & Engineering Chemistry Research 54 (17): 4683–4688.
60 Zhang, J., Wang, Q., Liu, J. et al. (2019). Saponin-­based near-­infrared nanoparticles with
aggregation-­induced emission behavior: enhancing cell compatibility and permeability. ACS
Applied Bio Materials 2 (2): 943–951.
531

18

AIE-­active Polymer
Rong Hu1, Anjun Qin1, and Ben Zhong Tang1,2,3
1
State Key Laboratory of Luminescent Materials and Devices, Guangdong Provincial Key Laboratory of Luminescence from Molecular
Aggregates, Center for Aggregation-­Induced Emission, South China University of Technology, Guangzhou, China
2
Department of Chemistry, Hong Kong Branch of Chinese National Engineering Research Centre for Tissue Restoration and
Reconstruction, Institute for Advanced Study, and Department of Chemical and Biological Engineering, The Hong Kong University of
Science & Technology, Clear Water Bay, Kowloon, Hong Kong, China
3
Shenzhen Institute of Aggregate Science and Technology, School of Science and Engineering, The Chinese University of Hong Kong,
Shenzhen, Guangdong, China

18.1 ­Introduction

Luminescent polymers, especially conjugated polymers with well light-­harvesting ability and
amplification effect, have attracted significant attention owing to their broad application in elec-
tronics devices, sensors, and medical applications, etc. However, the majority of the conventional
luminescent polymers always suffer from aggregation-­caused quenching (ACQ) effect, since poly-
mers with the hydrophobic skeleton and planer structures will prefer to form aggregate in an aque-
ous environment [1]. Thus, ACQ effect of conventional polymers will result in low sensitivity and
limit their further applications in practical uses. Since the first report in 2003, polymers with
aggregation-­induced emission (AIE) characteristic have been well developed and widely used in
related applications [2–10]. AIE-­active polymers always possess the unique properties, such as the
intense photoluminescence (PL), well photostability, large stokes-­shift, and outstanding synergis-
tic effect as compared to the traditional ones. Moreover, polymers with AIE feature own additional
advantages of easy modification, multifunction, intense adsorption, and amplification effect as
compared to the low-­mass AIE luminogens (AIEgens) [11–20]. These properties endow AIE-­active
polymers great advantages for various applications in sensing, electronic, and biomedical filed.
The AIE-­active polymers could be prepared by incorporating AIEgens into the skeleton or as the
pendants via polymerization and modification strategies. The synthetic methods for AIE-­active
polymer preparation have been discussed in detail and summarized by several exciting reviews.
Their applications in optoelectronic devices, chemo-­/biosensing, and biological applications have
also been introduced systematically [21–25]. These reviews and chapters have provided well plat-
form for designing and constructing AIE-­active polymers with objective functions and further uti-
lizing for practical applications. Nevertheless, the advantages of AIE-­active polymers over the
low-­mass molecules have rarely been summarized and discussed. Herein, this chapter will focus
on introducing the unique properties and advantages of AIE-­active polymer as compared to the

Handbook of Aggregation-Induced Emission: Volume 1 Tutorial Lectures and Mechanism Studies, First Edition.
Edited by Youhong Tang and Ben Zhong Tang.
© 2022 John Wiley & Sons Ltd. Published 2022 by John Wiley & Sons Ltd.
532 18  AIE-­active Polymer

low-­mass AIEgens, including the photophysical properties and practical applications. Moreover,
the trends of further development will be discussed as well.

18.2 ­Photophysical Properties

18.2.1  Quantum Yield


Compared to the low-­mass molecules, the construction of polymer can efficiently change the mor-
phology and extend the conjugation to a certain degree, which will have effect on the photophysi-
cal properties of the polymer. Luminogens with high-­PL quantum yield possesses high-­sensitivity
and good performance for real-­world applications, such as chem-­/biosensing, biological applica-
tions, and optoelectronic devices. According to the AIE mechanism of restriction of intramolecu-
lar motion (RIM), both the introduction of AIEgens to the backbone and the sidechain will partly
restrict the molecular motion by the steric hindrance, resulting in the PL enhancement both in
solution and solid states [26–31]. As a result, it has been widely reported that AIE-­active polymers
always exhibit aggregation-­enhanced emission (AEE) characteristic, owing to their weak emission
in the solution state [32–34].
Hong and coworkers have studied the influence of molecular weight on the AIE activity, which
was rarely evaluated before  [35]. As shown in Figure  18.1, vinyl polymers with low-­molecular
weight of 37 300 g/mol (P1) and high-­molecular weight of 525 400 g/mol (P2) with narrow poly-
mer dispersity index (PDI) were prepared, respectively. The properties of the model compound of
2,4,6-­triphenylpyridine and two polymers were investigated, in which P2 showed the highest spec-
tral stability without any reduction on emission until heating to higher than 200 °C. By observing
their AIE activity, the results revealed that they showed varied AIE behaviors owing to the differ-
ent conformations. For the model compound, nonemission was observed in solution state, and the
fluorescence enhanced gradually with the formation of the aggregate. The emissive behavior for
P1 was characterized as AEE, since the weak emission for the solution state was detected. While,
the solution emission of P2 was quite different from the model compound and P1, in which there
was only a little difference before and after the formation of aggregate, and a small peak shift from
375 to 380 nm was observed. This phenomenon was owing to a higher degree of aggregate for the
backbone of P2 than those of P1 and 2,4,6-­triphenylpyridine. Thus, they proposed that the model
compound would form excimer in the solid state. Meanwhile, fluorphoric triphenylpyridine in
linear P1 with low-­molecular weight would be separated dividedly in solution states and result in
weak emission, and the densely-­packing of triphenylpyridine would be formed upon aggregation.
While, for coil-­like P2, the highly curled region would cause the packing of triphenylpyridine to
form excimer with long wavelength emission in both solution and solid states (Figure  18.1b).
Moreover, by polymerization and enhancing the molecular weight, the quantum yield of fluores-
cence in aggregate states was increased from the model compound (0–32%) to P1 (25–78%) and
then P2 (82–84%). The combined results revealed that the construction of polymer would vary the
intermolecular interaction to influence the photophysical properties.
Based on the difference in PL intensity between the low-­mass molecule and polymer, Tang’s
group has utilized this characteristic to monitor the polymerization process, because the mecha-
nism of RIM makes AIEgens highly sensitive to the variation of environment [36]. As exhibited in
Figure 18.2a–d, reversible addition fragmentation chain transfer (RAFT) initiator (M1) with AIE
features was designed and prepared to construct AIE-­activity P3. This TPE-­containing initiator has
nonemission both in solution and solid states originated from the quenching effect of the carbonyl
sulfur moiety. While the PL gradually increased with the progressing of the RAFT polymerization
by naked eyes. Relatively weak emission was observed at the conversion of below c. 34% as
18.2 ­Photophysical Propertie 533

(b)
(a)
scc-Bu CH2 CH H
n
linear
water

O
P1: low molecular N
weight N aggregation
N

P2: high molecular coil-like


weight

water

(c)
1.0 (d) 1.0

0.8
Relative quantum yield (Φ1)

Relative quantum yield (Φ1)

0.9

0.6
Precipitation 0.8
Precipitation
0.4 Solution quantum yield Solution quantum yield
P1 P2
0.7
0.2 Solid quantum yield Solid quantum yield
P1 : 84.1
84.1%% P2 : 82.5
82.5%%

0.0 0.6
0 20 40 60 80 100 0 20 40 60 80 100
Water fraction (vol%) Water fraction (vol%)

Figure 18.1  (a) Chemical structures of P1 and P2. (b) Intermolecular approaches within the aggregated
nanoparticles for the more linear P1 and the coil-­like P2 chains. (c) The summarized solution quantum yield
of P1 vs. solution composition. (d) The summarized solution quantum yield of P2 vs. solution composition.
(a–d) Source: Adapted with permission [35]. Copyright 2012, Royal Society of Chemistry.

insufficient restriction induced by the low viscosity, and the PL intensity enhanced rapidly because
of the significant enhancement in viscosity. Moreover, this in-­situ monitoring polymerization pro-
cess could not be realized by using ACQ luminophore. This approach provided a useful platform
for the investigation of the polymerization process.
In order to construct artificial light-­harvesting systems, Xing and coworkers prepared an AIE-­
active supramolecular polymer (P4) by mixing cucurbit[8]uril (CB[8]) and M2 (Figure 18.2e–g) [37].
The supramolecular polymer exhibited enhanced fluorescence intensity compared to M2, resulted
by the restriction of molecular rotation of M2. By using sulfocyanine5 carboxylic acid (Cy5) as
energy acceptor and supramolecular polymer as energy donor, the efficient energy transferred
between the supramolecular polymer and Cy5. Thus, this AIE-­active supramolecular polymer
shows great potential as light-­harvesting system.
Except for previously typical ones, it has been widely reported that nonconventional AIE-­active
polymers without aromatic moiety also showed intense emission in their concentrated solution and
solid states, while, their low-­mass analogs always showed nonemission in both the solution and
solid states. Even though the luminescence mechanism for nonconventional polymers is still under
debatable and obscure [38–41]. You and coworkers have prepared the nonconventional polymers
with nonfluorescent monomers in 2012, and the fluorescence increased gradually with the
534 18  AIE-­active Polymer

O O
(a) S S

S N S N
N n N

O
M1 P3

(b)
0% 34% 47% 61% 70% 84% 95%

(c)
70 000 (d) 100
Conv (%) 20
95
56 000
84 18
70
80
Conversion (%)

61
PL intensity

42 000 16

Mn (kg/mol)
47
34

28 000
0 14
60
12
14 000

10
0 40
400 450 500 550 600 0 20 000 40 000 60 000
Wavelength (nm) PL intensity

(e)
N N
N N
CB [8]

M2 P4
(f) (g)
Acceptor emission
M2
P4 Donor emission 0
0.08%
0.21%
0.36%
Intensity

Intensity

0.76%
1.5%
2.0%
2.5%

500 550 600 650 700 750 800 500 550 600 650 700 750 800
Wavelength (nm)
Wavelength (nm)

Figure 18.2  (a) Illustration of the reaction process of P3. (b) Fluorescent photos of P3 solutions at
different conversion taken under 365 nm UV irradiation from a hand-­held UV lamp. (c) PL spectra of the
polymerization mixtures at different conversion. (d) The exponential relationship of conversion and Mn
with PL intensity. (a–d) Source: Adapted with permission [36]. Copyright 2018, Wiley-­VCH. (e) Illustration of
the formation of supramolecular polymer P4 through self-­assembly of M2 and CB8. (f) Fluorescence spectra
of M2 (60.2 mM, black) and P4 (60.2 mM, red). The inset photographs show the corresponding fluorescence
changes of M2 and P4. (g) Fluorescence spectra of P4 in aqueous solution with different concentrations of
Cy5; λex = 450 nm. (e–g) Source: Adapted with permission [37]. Copyright 2018, Elsevier.
18.2 ­Photophysical Propertie 535

proceeding of RAFT polymerization [42]. Tang and coworkers have developed series oligomers of
oligo(maleic anhydride)s and polymers of poly[(maleic anhydride)-­alt-­(2,4,4-­trimethyl-­1-­pentene)],
whose structures have been illustrated in Figure 18.3 [43]. By observing the PL spectra, both the
monomer (MAh, M3) and the repeating units (succinic anhydride, M4) showed nonemission in
solution and solid states. While P6 without any conventional chromophore possessed strong emis-
sion in its bulk solution and solid states with the emission peaks at 430 and 500 nm. Interestingly,
alternative polymer P5 with MAh moiety separated by the bulky t-­butyl units have relatively weak
fluorescence under the same conditions. To investigate the mechanism, density functional theory
was employed to optimize the conformations of P5 and P6, and revealed that the distance for adja-
cent MAh moieties was among 2.84~3.18 Å for P6. Meanwhile, the distance ranged from 4.90 to
5.37 Å for P5, which was farther than that of P6. These results demonstrated that the existence of
n–π* interaction for P6 would facilitate intra-­/interchain noncovalent interaction of the carbonyl

(a) O O O O
O O
O
O O

O O O O O
O O O O O
O O O O O
M3 M4 P5 P6
(b) (c)
3.6 P6 (365 nm)
P6 P6 (458 nm)
P5 P5

2.4
PL intensity (a.u.)
Absorptivity

1.2

0.0

200 300 400 500 600 700 400 500 600 700
Wavelength (nm) Wavelength (nm)

(d)
UV light

Clustering

O O O
Clusteroluminescence

Figure 18.3  (a) Chemical structures of M3, M4, P5, and P6. (b) UV–Vis spectra of P5 and P6 in THF. (c) PL
spectra of P5 (λex = 365 nm) and P6 (λex = 365 and 458 nm) in THF. (d) Illumination of the process of cluster
formation. (a–d) Source: Adapted with permission [43]. Copyright 2017, Royal Society of Chemistry.
536 18  AIE-­active Polymer

groups, resulting in the formation of molecular cluster. However, the carbonyl groups of P5 would
be separated by the deviation of the distance and angles without the formation of molecular cluster
and fluorescence emission. The combined results demonstrated the mechanism of clusterolumines-
cence of nonconventional polymers, and further provide a platform for design of the luminescence
polymers without conjugated structures.

18.2.2 Photosensitization
Therapy strategy based on reactive oxygen species (ROS) is coined as photodynamic therapy (PDT),
which possesses the unique properties of high selectivity, good spatial-­temporal control, minimal
invasiveness, and low-­drug resistance [44, 45]. For PDT, the transformation oxygen into ROS is the
fundamental element for practical applications. However, it has been widely reported that traditional
photosensitizers always suffered from the limitation of reduced ROS generation induced by the for-
mation of aggregate [46]. Photosensitizer with AIE characteristic could utilize the spontaneous pro-
cess of aggregate formation to enhance the ROS production [47]. Moreover, it has been demonstrated
that the polymerization will enhance the generation of ROS. Liu and coworkers revealed that conju-
gated polymers would show higher photosensitization effect than their small molecule counter-
parts  [48]. As displayed in Figure  18.4, three AIE-­active polymers and their low-­mass molecule
analogs were well designed and constructed to investigate the structure–property relationship. By
observing their ability of singlet oxygen (1O2) generation, the AIE-­active polymers (P7, P8, and P9)
exhibited 5.06-­, 5.07-­and 1.73-­fold higher 1O2 production efficiency than their small-­molecule ana-
logs (M5, M6, and M7), respectively. Time-­dependent density functional theory (TD-­DFT) study
revealed that with the increasing of the repeated units, the energy-­level distribution would become
denser, which could significantly improve energy transition channels. Thus, the intersystem crossing
(ISC) from the singlet excitation state to the triplet excitation state could be improved significantly,
which was essential to the enhanced 1O2 generation efficiency. Moreover, AIE-­active polymers with
conjugated backbone showed stronger light-­harvesting ability than the low-­mass molecules by
extending the conjugation length. The improved ISC process and light-­harvesting ability will endow
AIE-­activity conjugated polymer enhanced 1O2 generation. P7 possessed the strongest 1O2 generation
efficiency, and had been successfully utilized for waste decomposition, organic oxidation and PDT
based on the generated 1O2, which were better than those obtained by M5.
Tang’s group also proposed the method to enhance the efficiency of photosensitization by
polymerization-­facilitated strategy [49]. As shown in Figure 18.5a–d, the conjugation was increased
from M8 to M9 and then to P10, the ROS generation efficiency increased from 3.8 to 8.9% and then
to 14%, followed with the reducing in the PLQY from 87 to 15% and then to 7%. These results sug-
gested that the increase of conjugation would enhance the ROS generation by increasing the ISC. In
order to confirm this hypothesis, different polymers (P11 and P12) and their low-­mass molecule
analogs (M10 and M11) were prepared, and the similar trend was observed, so that the nanoparti-
cles were made with different preparation strategies. Moreover, by preparing different size and
shape of nanoparticles, there was ignorable change observed in the efficiency of ROS production.
Further, they also demonstrated that the ROS generation would be enhanced by increasing the
molecular weight from 3 800 to 11 000 g/mol. The nanoparticles of the conjugated polymer were
used to label cells with the enrichment in mitochondria, in which the high Pearson coefficient of
0.92 with MitoTracker Green was observed. Cell apoptosis was detected induced by the dysfunction
of mitochondria based on the generated ROS. Compared to the low-­mass molecule, the enhanced
ROS generation of the conjugated polymer demonstrated that polymerization was a promising
strategy to increase photosensitization for practical applications. Meanwhile, Tang and coworkers
(a) OCH3
H3CO OCH3
O
S
H3CO OCH3 N N

O N
M5 M6 H3CO
NC CN M7

R1O OR1 R 1O OR1


N S N
R1 =
N
n
O

n N
P7 P8
n P9
NC O
CN
(b)
No.
a
Mw (g/mol–1) Mw /Mn
a b
λmax (nm) εc/ (l/mol–1/cm–1)
d
QY (%)
1 O Generation
2
e
(c)
efficiency (nmol)
Sn Sn
M5 (–) (–) 318 2.59 × 104 12.1 4.02 ISC IS
C
420 1.16 × 104 S1 S
∆ES1-T1 ∆ES-T
P7 17 000 1.45 335 5.50 × 104 6.2 20.3
T1
Fluorescence

Fluorescence
M6 (–) (–) 335 3.02 × 104 1.2 0.27 1O
Excitation

Excitation

2
1O 1O
2 2

435 9.01 × 103 1O


2

P8 19 500 1.66 340 3.60 × 104 ~0 1.37


O2
435 1.02 × 104 O2 O2
M7 (–) 315 2.28 × 104
S0 S0 O2
(–) 39.2 2.5
485 1.11 × 104
P9 9 500 1.92 320 4.00 × 104 5.5 4.3 Low-mass molecule
Conjugated polymer photosensitizer
485 1.47 × 104 photosensitizer

Figure 18.4  (a) The chemical structures of small molecules M5–M7 and conjugated polymers P7–P9. (b) Comparison between conjugated
polymer photosensitizers and low-­mass molecule photosensitizers. a. Determined by gel permeation chromatography in THF on the basis of
polystyrene calibration. b. Absorption peaks of photosensitizers in THF/water = 1/99 (v/v). c. Molar absorption coefficients of photosensitizers at
the absorption peaks in THF/water = 1/99 (v/v). d. Quantum yield of photosensitizers in THF/water = 1/99 (v/v) with 4-(dicyanomethylene)-2-
methyl-6- (4-dimethylaminostyryl)-4H-pyran (43% in methanol) as reference. e. Tested with ABDA as an indicator, the values are based on the
consumed amount of ABDA by 10 nmol of different photosensitizers upon white-light irradiation (60 mW/cm2, 400–700 nm) for 1 min. (c) The
different photosensitization processes of low-­mass molecule photosensitizer and conjugated polymer photosensitizer. (a–c) Source: Adapted
with permission [48]. Copyright 2018, Elsevier.
538 18  AIE-­active Polymer

(a) N
S
N
O
CN O
N N
N O O
NC
O
M8
O
M10
N
S
N M11
C6H13-n
N N n-C8H17
CN O
N N O
M9 S
N n
n
NC
S
N N O
P11
N P12
n
O
C8H17-n
P10
(b) (c) (d) n-C6H13

16 100 8 12
Φo 10.1
Φf 80 6.2
12 6
8
60
Φo (%)
Φf (%)
Φo (%)

Φo (%)

Φo (%)
8 4
40
4
4 2
20 0.9 1.5

0 0 0 0
M8 M9 P10 M10 P11 M11 P12

(e) (f)
S
35
N N S
N N 30
Blank
O M12
O
25 P13
N Chlorin E6
N 20
O Br
O Br
I/I0

O F 15
F
N O F F
O N3
10
Br F N
F F O
O F Br N3 5
F F F
N3 F
O
O N3 0
F F
O
F F 0 1 2 3 4 5
M12 Time of light irradiation (minutes)
P13

Figure 18.5  (a) Chemical structures of M8, M9, M10, M11, P10, P11, and P12. (b) 1O2 quantum yield and
fluorescence quantum yield of M8, M9, and P10. (c) 1O2 quantum yield of M10 and P11. (d) 1O2 quantum
yield of M11 and P12. (a–d) Source: Adapted with permission [49]. Copyright 2018, Wiley-­VCH. (e) Chemical
structure of M12 and P13. (f) Relative fluorescence intensity of 2′,7′-­dichlorofluorescein (DCFH) with
addition of P13, M12, and Ce 6 upon exposure to white light. (e–f) Source: Adapted with permission [50].
Copyright 2020, Wiley-­VCH.

also utilized the polymerization strategy to improve the ROS generation by constructing an AIE
conjugated polymer P13 [50]. This AIE-­active polymer with donor–π–acceptor structure displayed
relatively ultra-­strong ROS generation ability, which was 13-­and 11-­fold higher than that of the low-­
mass molecules of Ce 6 and the model compound M12. By forming aggregate in PBS buffer, the ROS
production could be further enhanced. These works demonstrated the construction of AIE-­active
polymer was a promising strategy for enhanced ROS generation.

18.2.3  Two-­photon Absorption and Emission


The development of luminogens with long wavelength absorption and emission can overcome
the problem of limited penetration, thus, imaging and therapy based on two-­and multiphoton
can increase the penetration depth and enhance the therapeutic treatment effect. It is worth
18.2 ­Photophysical Propertie 539

mentioning that the enhanced ICS usually induces the limited conjugation, which is not
favorable for the strong two-­photon absorption cross sections. Thus, simultaneous improve-
ment in both the 1O2 generation and two-­photon absorption cross sections is not easy to real-
ize, which is a challenge not limited to luminogens with AIE activity. Lu and coworkers
prepared a homopolymer including pyrazoline and 1,8-­naphalimide moieties by atom transfer
radical polymerization  [51], and the structure of the monomer and polymer are shown in
Figure 18.6a–c. The photophysical properties of M13 and P14 were investigated and compared

(a) (b) (c)


O
Br
500 P-toluene
1000
O p-THF

TPA cross section (GM)


TPL emission (a.u.)

P-DMF
400 Toluene 800
O m-Cyclohexane
O THF
m-Toluene
300 DMF 600 m-THF
m-DMF
200 400
N
N
N N 100 200
O O 0
N N Et N Et 0
Et N 500 550 600 650
nBu
O nBu Et 760 780 800 820 840
O Wavelength (nm)
M13 P14 Wavelength (nm)

(d) H3CO OCH3 BrC6H12O OC6H12Br OC6H12Br

n BrC6H12O n

M14 NC CN
P15 NC CN P16 NC CN

P16

(e) Nonoprecipitation TAT-SH


THF: Water = 1:10 (v/v) Click reaction

DSPE-PEG-Mal
P16 dots
P16-TAT dots
(f) (g) (h)
1.0 100 7.5 M14
P15
2PACS (x105 GM)

0.8 6.0 P16


80
Absorbance

A/A0 (%)
PL (a.u.)

4.5
0.6 60
Only ABDA 3.0
0.4 40
Ce6
M14 M14 1.5
0.2 20
P15 P15
P16 P16 0.0
0.0 0
300 400 500 600 700 800 0 30 60 90 120 800 850 900 950 1000 1050
Wavelength (nm) Time (seconds) Wavelength (nm)

Figure 18.6  (a) Chemical structure of M13 and P14. (b) Two photofluorescence spectra of P14. (c) Two
photoabsorption spectra of M12 and P13 in different solutions at the same concentration. (a–c) Source:
Adapted with permission [51]. Copyright 2013, Royal Society of Chemistry. (d) Chemical structures of M14,
P15, and P16. (e) Schematic illustration for the synthesis of AIE PS dots and AIE-­TAT PS dots using P16 as an
example. (f) Normalized absorption and photoluminescence (PL) spectra of AIE PS dots in aqueous media.
(g) Normalized degradation percentages of ABDA in the presence of PS dots in aqueous media upon white
light irradiation (400−700 nm, 50 mW/cm2). [PS dots] = 10 μM based on dye; [ABDA] = 50 μM. (h) Two-­
photon absorption cross-­section spectra of PS dots in aqueous solution. (d–h) Source: Adapted with
permission [52]. Copyright 2013, American Chemical Society.
540 18  AIE-­active Polymer

in this work. The study of the ground-­state geometry of M13 revealed that there was a charge–
transfer interaction between the donor and acceptor moieties, thus, the solvent effect was
observed for M13 and P14. The results demonstrated that P14 was more insensitive to the polar
of the solvent than that of M13, which was understandable that the chromophore was strictly
anchored and intramolecular rotation was inhibited for P14. Moreover, compared to that of
M13, P14 exhibited a larger Stokes shift in two-­photon emission and more obvious enhance-
ment of emission, in which the two-­photon fluorescence spectrum was similar to that of the
one-­photon fluorescence. Moreover, P14 enjoyed a much larger two-­photon absorption cross-­
section of 983  GM than that observed for M13 (22  GM) in the same condition, which was
related to the intramolecular charge transfer of P14. Furthermore, the two-­photon emission of
P14 could be enhanced upon the formation of aggregate. These results demonstrated the
amplification effect of the polymer, and polymerization could efficiently enhance the two-­
photon absorption.
Liu’s group reported a polymerization strategy for enhancing two-­photon absorption and photo-
sensitization [52]. As exhibited in Figure 18.6d–g, a pair of conjugated polymers (P15 and P16)
with the model compound (M14) were designed and prepared. The introduction of donor–accep-
tor structure was designed for raising two-­photon absorption cross sections and 1O2 generation,
and the adjustment of linkage position between the donor and acceptor could enhance the conju-
gation of the polymer. As mentioned earlier, both the donor–acceptor structure and the conjugated
backbone could improve the production of 1O2, as the efficiency of 1O2 generation for P15 (548 and
637%) was much higher than that of low-­mass molecules of M14 and Ce 6. The TD-DFT based on
the three model compounds revealed the good HOMO–LUMO separation and the smaller energy
level gap, which could favor the generation of 1O2. Further, by investigating the two-­photon absorp-
tion cross sections, the maximal values from M14 to P15 and then P16  were measured to be
1.13 × 105, 3.56 × 105, and 7.36 × 105 GM, respectively. TD-DFT also indicated that the torsional
angels between different units of P16 were smaller than those of M14 and P15, demonstrating the
enhanced conjugation of P16 to facilitate a larger two-­photon absorption cross sections. Moreover,
the photostability of P16-­dots was verified to be better than the low-­mass molecules of Qtracker
625 and Alexa Fluor 647. This work highlights a useful strategy for the design of efficient photo-
sensitizer based on two-­and multiphoton.

18.2.4  Circularly Polarized Luminescence


As the emission analog of circular dichroism, CPL is the selective emission of right-­ and left-­
handed circularly polarized light from the chiral chromophore. With strong solid-­state emission
and mechanical flexibility, AIE polymeric materials with CPL activity have attracted great atten-
tion in high-­tech photonic devices [53]. Zhu and coworkers have constructed a chiral polymer with
strong and tunable CPL dissymmetry factor based on a chiral amino acid pendant and a TPE core
(Figure  18.7)  [54]. Meanwhile, two types of model compounds were prepared (M15 and M16).
Thanks to the chiral signal amplification of polymer, the specific rotation value of P17 was +400
(c = 0.2 g/100 ml, THF), which was much higher than that of chiral M16 (+25, c = 0.2 g/100 ml,
THF). Moreover, P17 exhibited stronger fluorescence than that of M15 with the red-­shifted emis-
sion of 27 nm. CPL dissymmetry factor of this polymer could be regulated in the range of 0.44 to
0.08 by varying the water fraction in THF, demonstrating the superiorities of AIE polymer in CPL-­
related applications over low-­mass molecules.
18.3 ­Application 541

(a)
OC8H17-n n
Br

OC8H17-n
COOCH3

HN O COOCH3
Br
HN O
O
M15 M16 O P17
(b) (c)
10
fw (vol%) 0.45
9 0
30 0.40
8
60
7 0.35
CPL intensity (∆I)

80
6 90 0.30
95
5 glum
0.25
4 0.20
3
0.15
2
0.10
1
0 0.05
450 500 550 600 0 20 40 60 80 100
Wavelength (nm) Water fraction (vol%)

Figure 18.7  (a) Chemical structures of M15, M16, and P17. (b) CPL spectra of P17 in THF-­water mixtures
with different fractions of water (fw). (c) CPL dissymmetry factor glum versus water fraction for P17,
λex = 371 nm. (a–c) Source: Adapted with permission [54]. Copyright 2014, American Chemical Society.

18.3 ­Applications

18.3.1  Chem-­sensor
AIE-­active polymers, especially the conjugated ones, possess an amplification effect and a large
contact area in aggregate state compared to low-­mass analogs, which endows its high sensitivity
for sensing applications. Moreover, the AIE mechanism of RIM make polymers with AIE activity
highly sensitive towards the variation of microenviroment, making it highly desirable for the sens-
ing applications. A variety of AIE-­active polymers have been developed for the sensing of ions,
pollutants, explosive, pH, and multiplex with high sensitivity, and many of them exhibited the
unique advantages over the low-­mass molecules [55–60].
Tang and coworkers have developed an anionic-­conjugated polytriazole with good solubility via
azide-­alkyne click polymerization  [61]. Thanks to the introduction of TPE, this polymer shows
AIE behavior with enhanced fluorescence in aggregate state for 19.5 times higher than that of the
solution state. The responsibility towards 14  metal cations of Na+, K+, Ag+, Zn2+, Mg2+, Cu2+,
Co2+, Ca2+, Ni2+, Pb2+, Mn2+, Al3+, Fe3+, and Cr3+ was carried out, and only Al3+ could boost the
fluorescence of P18, demonstrating its high specificity (Figure 18.8). The results showed that its
detection limit of Al3+ was as low as 31 ppb, which was much lower than the regulated level of
0.2 mg/l in drinking water documented by the World Health Organization (WHO). Interestingly,
the model compound exhibited no fluorescence response with the addition of Al3+, indicating the
specificity towards Al3+ was related to the skeleton of the polymer. They had proposed the mecha-
nism based on photoinduced electron transfer (PET) and the excited-­state intramolecular proton
transfer. The 1H-­NMR spectra were employed to verify the strong interaction between Al3+ and the
nitrogen atom of the triazole rings. Thus, this interaction would induce the closer of two polymer
542 18  AIE-­active Polymer

(a)
SO3Na

N N
SO3Na N N
N N
P18
n
(b) 6
700
Metal cations
AI3+
5 600 Cr3+
Blank
K'
500 Ca2+
PL intensity (a.u.)

Mn2+
4 NI2+
400 Ag+
Mg2+
Pb2+
Zn2+
3 300
(I–I0)/I0

Na+
Co2+
Cu2+
200 Fe3+

2
100

1 0
400 460 520 580 640
Wavelength (nm)

0
n 2+

g 2+
k

a 2+

o 2+
u 2+
2+

2+
r 3+
+

i 2+

a+
+
an
AI 3

K+

Ag

Pb
Zn
Bl

M
C

N
C
C

3+
Fe

–1
Metal cations

(c)

Water-soluble unit AIEgen unit

Tied Up by CuAACP Aggregated by AI3+

Figure 18.8  (a) Chemical structure of P18. (b) The change of the relative PL intensity of P18 upon the
addition of metal cations into the DMSO/water mixture with a water fraction (fw) of 5%. (c) Proposed
schematic sensing mechanism of P18 for Al3+. (a–c) Source: Adapted with permission [61]. Copyright 2016,
Royal Society of Chemistry.

chains to restrict the intermolecular motion of TPE units to realize the turn-­on fluorescence of
P18. This anionic-­conjugated polytriazole displayed synergistic effect over low-­mass molecules,
and can be used for drinking water purity control of Al3+ content.
An explosive detection is an international event due to the concerns on environmental protec-
tion and global security, and ultraselective and great-­sensitive detection of explosives is highly
demanded for civilian safety and antiterrorism operation. AIE-­active polymers have been well
designed and developed for explosive detection with superamplified-­quenching effect. These AIE-­
active polymers possess the following unique advantages: (i) extended electronic conjugation
structure and signal-­amplifying ability. (ii) The presence of plentiful molecular cavities endows the
18.3 ­Application 543

high-­binding ability with explosives. (iii) The excitonic migration with multidimensional path-
way [62]. As a result, kinds of hyperbranched polymer with AIE features have been developed by
Tang’s group [63]. By introducing spring-­like flexible spacers, two AIE-­active hyperbranched poly-
mers (P19 and P20) were produced with the quantum yields (ΦF) value measured to be 100% in the
solid state [64]. Picric acid (PA) and 2,4,6-­trinitrotoluene (TNT) were selected as model explosives,
and the results showed that the fluorescence decreased obviously with no peak shift, which is
desirable for detection applications. Meanwhile, the sensitivity towards PA was much better than
that of TNT, which was owing to the energy transfer from polymers to PA, while the mechanism
for fluorescence quenching of TNT was charge transfer. Moreover, the Stern–Volmer plots of these
hyperbranched polymers upon explosive addition exhibited bend upward, demonstrating the
superamplified-­quenching effect (Figure 18.9a–c). Hyperbranched poly(2,5-­silole)s was also well
designed and synthesized with the degree of branching of 0.55 [67]. It had been demonstrated that
the substituents at the 2,5-­positions had a great effect on the energy gap to the photophysical prop-
erties, and the emission could be regulated by varying the morphologies and concentrations. This
hyperbranched poly(2,5-­silole) was employed to test TNT and PA, and the sensitivity of the aggre-
gate state was much higher than that of the polymer solution was obtained, in which obvious
superamplified effect with the quenching constant of 33  333  l/mol was realized. Moreover, the
quenching constant is much higher than that of linear polymers (<20 000 l/mol). Another hyper-
branched poly(silylenephenylene) with a rigid aromatic scaffold of three-­dimensional (3D) topo-
logical structure was reported, and exhibited superamplification effect for PA detection with the
concentration as low as 1  ppm in a static-­quenching mechanism  [68]. The superamplification
effect proposed that the PA in the inner core and around the outer shells would bind with the
hyperbranched polymer cooperatively with the facilitation of energy/electron transfer.
Scherf’s group first developed two polycarbazoles (P21 and P22) with the AIE active moiety in
the side chain for 1,3,5-­Trinitrobenzene (TNB) detection in 2014  [65]. The fluorescence of this
polycarbazole decreased apparently with no change in spectra. The amplified quenching was also
observed by fitting the data with the Stern–Volmer equation, which was attributed to the improved
exciton diffusion along the backbone to the quenching sites based on the 3D topology of the nano-
aggregates. Meanwhile, the vapor mode and solution tests of TNB detection were carried out
(Figure 18.9d–f), and indicated the potential for the fabrication of sensors for explosive detection
with great sensitivity.
An interesting through space conjugated polymer (P23) containing folded TPE units in the back-
bone was developed by Tang and coworkers (Figure 18.9g–i) [66]. It was found that this polymer
showed obvious AIE property and high-­decomposition temperatures (340–436  °C). This AIE-­
active polymer was successfully employed to sensitize the explosive with the quenching constant
of 552 000 M−1, which was relatively higher than that in THF solution. What is more interesting,
the quenching constant of the through-­space conjugated polymer was higher than that of the con-
ventional AIE-­active polymer with the similar repeated unit structure, since the presence of the
folded moiety would provide more molecular cavities to interact with analytes efficiently. Further,
the theoretical calculation revealed that the energy level of the repeated unit was higher than that
of PA, resulting in the PET to realize the PA test with high efficiency.

18.3.2 Bioimaging
Compared to the conventional fluorophores with ACQ effect, polymers with AIE features exhib-
ited the advantages of superior photostability, strong fluorescence, high sensitivity, and well
biosafety, making it possess technological and academic superiority for imaging and related
544 18  AIE-­active Polymer

(a) (b) (c)


200
200
[PA] / mM
PA
0 150
150
100

50
0.45

I0/I
I
100
0
0.00 0.25 0.50
50

TNT
0
390 440 490 540 590 640 0 2 4 6 8
R = –(CH2)6 (hb-P5a) P19 λ (nm) [Q] (mM)
–(CH2)4 (hb-P5b) P20
(d) (e) (f)
a c a c e

b d b d f

P21 : x = y = 0.5
P22 : x = 0,y = 1
(g) (h) (i) (j)
[PA] mM [PA] mM 120 THF/H2O (v/v 2:8)
0 0 THF
PL intensity (a.u.)

PL intensity (a.u.)

90 KSV, I = 1.84 × 104 M–1


KSV, II = 1.40 × 105 M–1
III
0.69 KSV, III = 5.52 × 105 M–1
0.57
I0/I–1

60
KSV, I = 4.83 × 103 M–1
KSV, II = 1.83 × 104 M–1

30 KSV, III = 7.82 × 104 M–1

P23 II
III
I
II
0
380 430 480 530 580 630 380 430 480 530 580 630 0.0 0.2 0.4 0.6
Wavelength (nm) Wavelength (nm) [PA] (mM)

Figure 18.9  (a) Chemical structure of P19 and P20. (b) PL spectra of P19 in THF/water mixture containing
different amounts of PA. (c) Stern–Volmer plots of I0/I versus [PA] in THF/water with fw = 90% and [TNT] in
THF/methanol with fm = 90%. Inset: enlarged plot of I0/I versus [PA]. (a–c) Source: Adapted with
permission [64]. Copyright 2011, Royal Society of Chemistry. (d) Chemical structure of P21 and P22. (e)
Paper strip tests. Vapor-­mode detection of TNB: test strips before (a and b) and after (c and d) placing the
strips on top of a glass vial containing solid TNB for five minutes; for P21 (a and c) or P22 (b and d),
respectively. (f) Solution-­mode detection of TNB: test strips before (a and b) and after dipping the strips into
pure THF (c and d) and into a 10-­4 M TNB solution in THF (e and f); for P21 (a, c, and e) or P22 (b, d, and f),
respectively. (d–f) Source: Adapted under a Creative Commons Attribution 3.0 Unported Licence [65].
Copyright 2014, Royal Society of Chemistry. (g) Chemical structure of P23. Photoluminescence (PL) spectra
of P23 in (h) THF/water mixture (2 : 8 v/v), and (i) in THF containing different amounts of PA. (j) Stern–
Volmer plots of (I0/I − 1) versus [PA] in THF/water mixture with fw = 80% and THF, λex = 328 nm. (g–j)
Source: Adapted with permission [66]. Copyright 2018, Royal Society of Chemistry.

applications. Great endeavors have been paid for the development of AIE-­active polymer in the
biological imaging and diagnosis [69]. For example, the discrimination and detection viable cells
are quite important for providing the information on therapeutic effect during in vitro treatment.
However, it is not easy to discriminate the viable cells over the apoptosis ones, especially the cells
in the early apoptosis stage. Qin and coworkers designed an AIE-­active polymer with conjugated
structure (P24) [70]. As shown in Figure 18.10, by introducing oligo(ethylene glycol)s on the side
(a) O
(b) 120
O
O 100
O

Cell viability (%)


Br O 80
Calcein AM
60
n
40
O Br 20
O O
M16 P24 0
O O 0 10 20 30 40 50 60 70
Concentration (μM)
(c) P24 Propidium iodie Annexin V-FITC Merged

M16 Propidium iodie Annexin V-FITC Merged

Calcein AM Propidium iodie Phase contrast Merged

(d) Day 1 Day 2 Day 3 Day 8

P24

Day 1 Day 2 Day 3


0.9 Calcein. AM

0.6
Calcein.AM
I/Ig

0.3

0.0
1 2 3 4 5 6 7 8
Time (day)

Figure 18.10  (a) Chemical structure of M16 and P24. (b) Cell viability of HeLa cells incubated with
different concentration of P24 and Calcein AM incubated for 24 hours. (c) CLSM images of HeLa cells
incubated with P24, M16, and Calcein AM after treated by 1000 μM H2O2 for six hours, and followed by
staining with Annexin V-­FITC and PI. (d) CLSM images of HeLa cells incubated with P24 for one, two, three,
and eight days or Calcein AM for one, two, and three days, respectively. And histogram of the changes in
fluorescence intensity with different incubation time. (a–d) Source: Adapted with permission [70]. Copyright
2019, Elsevier.
546 18  AIE-­active Polymer

chain of the conjugated backbone, this polymer could selectively bind with viable cells over apop-
tosis and dead cells even microbes. Moreover, thanks to the high-­molecular weight and superam-
plification effect, this AIE-­active polymer exhibited better selectivity towards viable cells over the
low-­mass molecule of the model compound M16 and commercial probe of Calcien-­AM. By inves-
tigating the uptake mechanism, it was figured out that the side chain of oligo(ethylene glycol)s and
the endocytosis process both have critical effect on the selectivity toward viable cells, which could
not be realized by the model compound. Meanwhile, by investigating the long-­term tracing behav-
ior, the intense fluorescence of the polymer could be observed inside cells after incubation for eight
days, while, relatively weak signa of Calcien-­AM was detected at the third day, demonstrating the
stronger tracing ability of this AIE-­active polymer. This phenomenon was attributed to the strong
retention ability of polymer. This work verified the superiority of AIE polymers for bioimaging
applications over low-­mass molecules.
Unnatural reactions inside living cells as a powerful tool are highly demanded for the biological
research. Tang and coworkers developed a novel intracellular polymerization based on spontane-
ous amino-­yne click polymerization for turn-­on cell imaging and enhanced antitumor, which
could not be achieved by the monomers (Figure 18.11) [71]. Herein, carbonyl group activated ter-
minal diyne (M18) could successfully polymerize with TPE-­containing primary diamine (M17)
both in organic and water, and the reaction efficiency would be enhanced in water, demonstrating
the obvious “on-­water” effect. By using confocal laser scanning microscope (CLSM), the intense
fluorescence inside cells was observed after sequential incubation with two monomers, while,
quite weak fluorescence could be detected only upon the treatment of TPE-­containing primary
diamine. Thus, turn-­on imaging of HeLa cells was realized that was attributed to the immobiliza-
tion of TPE-­containing primary diamine by intracellular polymerization. Moreover, compared to
the individual treatment with the monomer, enhanced antitumor was realized by destroying the
structure of tubulin and actin based on the spontaneous click polymerization inside cells, which
would induce the necrocytosis of cells. The FT–IR spectra was utilized to verify the structure of
AIE-­active polymer obtained inside cells with the disappearance of ≡C–H and C≡C stretching
vibrations at 3225 and 2078 cm−1 and with the appearance of C=C stretching vibration at 1603 cm−1.
The molecular weight was evaluated to be 7300 by APC. This intracellular polymerization shows
great promise for bioimaging and therapy in biomedical applications.

18.3.3  Therapy Applications


Thanks to the diversity structure and AIE feature, AIE polymers possess high-­fluorescence effi-
ciency and outstanding sensitivity in the aggregate state. As a result, AIE-­active polymers have
been utilized widely in imaging-­guided therapy and related applications [23, 72–76]. Biomaterials
with favorable antibacterial activity and good biocompatibility are highly demanded for infection
therapy. Because the infection therapy based on PDT is not easy to meet the problem of multidrug
resistance, theranostic agents with PDT effect have attracted enormous attention and been well
developed for the clinical treatment of bacterial infections. Tang and coworkers have well
designed and produced a unique conjugated polymer feathering AIE property (P12) that shows
outstanding advantages over low-­mass molecules for infection diagnose and therapy
(Figure  18.12a, b)  [50]. What different from the previously conjugated polymer with the AIE
effect is that this polymer could selectively bind with microbes over mammalian cells, and showed
no cytotoxicity toward the mammalian one. Thus, effective bacterial eradication both in vitro and
in vivo was realized with outstanding biosafety. However, the model compound M11 could bind
with both the microbes and mammalian cells effectively, and the cell viability test revealed that it
18.3 ­Application 547

(a) NH2
O
O H H
N N
+
O O
O O
H2N
n
M17 M18 P25
(b) (c) (d)
120

90 90
90
Cell viability (%)

Cell viability (%)

Cell viability (%)


60 60
60

30 30 30

0 0 0
0.5 1 2 4 8 16 32 64 1 2 4 8 16 32 64 0.5 1 2 4 8 16 32 64
Concentration (μM) Concentration (μM) Concentration (μM)

PL Channel M17 M18 Intracellular


polymerization

Merged

Figure 18.11  Cell viability of HeLa cells incubated with (a) M17, (b) M18, and (c) intracellular amino-­yne
click polymerization (P25) with the monomer’s concentration ranging from 0 to 64 μM. (d) CLSM images of
HeLa cells incubated with M17, M18, and intracellular polymerization, respectively. (a–d) Source: Adapted
with permission [71]. Copyright 2019, Springer Nature.

exhibited serious cytotoxicity toward HeLa cells. It was proposed that the great selectivity and
biocompatibility of this conjugated polymer was originated from the positively charged side chain
to bind with microbes via electrostatic interaction, while, the balanced hydrophobicity and hydro-
philicity of this polymer would inhibit the binding with mammalian cells. The stronger ROS
548 18  AIE-­active Polymer

(a) P13 M12 (b)


1.2
HeLa HeLa+Microbes HeLa M12
P13
1.0

0.8

Cell vialbility (%)


0.6

0.4

0.2

0.0
0 0.5 1 2 4 8 16 32
Concentration (μM)

(c)
AIE-TAT dots
Live cell Two-photon laser
Irradiation area
Dead cell

Cytoplasm

1
O2
1
O2 1
O2 Nucleus

1
O2

(d)
20 scans 40 scans 60 scans 80 scans

M14

P15

P16

Figure 18.12  (a) CLSM images of HeLa cells and mixed sample (HeLa cells+ microorganisms) incubated
with PTB-­APFB or MTB-­APFB, respectively. (b) Cell viability of HeLa cells incubated with P13 or M12 for
24 hours. (a–b) Source: Adapted with permission [50]. Copyright 2020, Wiley-­VCH. (c) Schematic illustration
of the AIE-­TAT PS dots for in vitro 2PE cell PDT. (d) Confocal live/dead cell images of HeLa cells incubated
with M14-­TAT dots, P15-­TAT dots, and P16-­TAT dots subjected to 2PE-­PDT after different scanning numbers.
(c–d) Source: Adapted with permission [52]. Copyright 2013, American Chemical Society.
18.4  ­Conclusion and Perspectiv 549

generation, better selectivity, and biosafety, demonstrating the synergetic effect of the polymer
over low-­mass molecules.
Given to the better 1O2 generation of P16 than that low-­mass molecule, M14 and the polymer
P15, their two-­photon photosensitizations were evaluated intracellularly  [52]. The intracellular
1
O2 production of the three nanoparticles based on M14, P15, and P16 were investigated, respec-
tively. The brightest fluorescence of ROS probe inside cells could be detected after the treatment of
P16-­dots. Thus, the cell ablation of these nanoparticles was also evaluated on the basis of the
PDT. As shown in Figure 18.12c and d, by using propidium iodide and fluorescein diacetate (PI/
FDA) staining for live/dead cell imaging, almost all cells were killed efficiently for P16-­dots treated
groups, and the majority of cells were survived for both M14-­dots and P15-­dots. The better PDT
performance of P16-­dots than those of nanoparticles based on M14 and P15 indicates the great
precision and high efficiency of P16-­dots.

18.4  ­Conclusion and Perspective

With the establishment of AIE concept, AIE polymeric materials have been well developed and
widely used in chemo-­/biosensing, biological applications, and optoelectronic devices. AIE-­active
polymers not only exhibit the unique properties of efficient emission, high photostability, large
Stokes shift, and good biocompatibility as compared to the conventional polymeric luminogens,
but also possesses the outstanding advantages of structure diversity and topologies, amplification
effect, and multifunctional properties over the low-­mass AIEgens. We have discussed and sum-
marized the effect of the construction of AIE-­active polymer on the photophysical properties,
including the quantum yield, CPL, TPA, and photosensitization, compared to their low-­mass ana-
logs. Moreover, the applications of AIE polymeric materials in chem-­sensing, bioimaging, and
therapy were introduced in details, and the advantages over the low-­mass molecules were also
illustrated in this chapter.
Inspired by the remarkable achievements of AIE polymeric materials, continuous efforts should
be paid to explore more possibilities and breakthrough in this field. First, the design and prepara-
tion of newly developed AIE-­active polymers with exciting properties and inspiring functions are
still highly desirable. The fabrication of AIE polymers with diverse structures and topologies are
also demanded to meet the various demands of practical applications. Second, the relationship of
the polymer structure with the property should be studied more systematically, which could facili-
tate the design and development of AIE polymeric materials with desirable characteristics.
Persistent efforts should be devoted to observing the PL mechanism for nonconventional polymers
with AIE behavior with the support of experimental evidence and theoretical simulations. Third,
the combination of AIE activity with stimuli responsiveness is an exciting direction for construct-
ing multifunctional polymers to be applied in sensing and theranostic fields. Meanwhile, AIE poly-
mers owing high-­PL efficiency, great biosafety, strong stability, tunable emission as well as room
temperature photoluminescence, and NI emission is a major challenge in related applications.
Finally, the advantages of AIE-­active polymers over the low-­mass molecules in properties and
applications should be further investigated. Last but not the least, it is not possible to address all
the opportunities here. With such prospects, we hope this chapter could inspire the exploration of
AIE-­active polymers possessing fascinating structures and properties for the innovative high-­tech
applications.
550 18  AIE-­active Polymer

­Acknowledgments

This work was financially supported by the National Natural Science Foundation of China
(21788102, 21525417, 21907034 and 51620105009), the Natural Science Foundation of Guangdong
Province (2019B030301003, 2018A030313763 and 2016A030312002), the Fundamental Research
Funds for the Central Universities (D2191530), and the Innovation and Technology Commission
of Hong Kong (ITC-­CNERC14S01).

­References

1 Mei, J., Leung, N. L. C., Kwok, R. T. K. et al. (2015). Aggregation-­induced emission: together we
shine, united we soar! Chemical Reviews 115 (21): 11718–11940.
2 Wang, Z., Chen, S., Lam, J. W. Y. et al. (2013). Long-­term fluorescent cellular tracing by the
aggregates of AIE bioconjugates. Journal of the American Chemical Society 135 (22):
8238–8245.
3 Ishiwari, F., Hasebe, H., Matsumura, S. et al. (2016). Bioinspired design of a polymer gel sensor for
the realization of extracellular Ca2+ imaging. Scientific Reports 6: 24275.
4 Zhang, H., Zheng, X., Kwok, R. T. K. et al. (2018). In situ monitoring of molecular aggregation using
circular dichroism. Nature Communications 9:4961.
5 Zhang, N., Chen, H., Fan, Y. et al. (2018). Fluorescent polymersomes with aggregation-­induced
emission. ACS Nano 12 (4): 4025–4035.
6 Shao, L., Sun, J., Hua, B. et al. (2018). An AIEE fluorescent supramolecular cross-­linked polymer
network based on pillar 5 arene host-­guest recognition: construction and application in explosive
detection. Chemical Communications 54 (38): 4866–4869.
7 Suenaga, K., Yoshii, R., Tanaka, K. et al. (2016). Sponge-­type emissive chemosensors for the protein
detectionbased on boron ketoiminate-­modifying hydrogels with aggregation-­induced blueshift
emission property. Macromolecular Chemistry and Physics 217 (3): 414–421.
8 Mi, S., Wu, J., Liu, J. et al. (2015). AIEE-­active and electrochromic bifunctional polymer and a device
vcomposed thereof synchronously achieve electrochemical fluorescence switching and
electrochromic switching. ACS Applied Materials & Interfaces 7 (49): 27511–27517.
9 Li, W., Ding, Y., Tebyetekewa, M. et al. (2019). Fluorescent aggregation-­induced emission (AIE)-­
based thermosetting electrospun nanofibers: fabrication, properties and applications. Materials
Chemistry Fronters 3 (11): 2491–2498.
10 Song, L., Zhu, T., Yuan, L. et al. (2019). Ultra-­strong long-­chain polyamide elastomers with
programmable supramolecular interactions and oriented crystalline microstructures. Nature
Communications 10: 1315.
11 Zhan, R., Pan, Y., Manghnani, P. N. et al. (2017). AIE polymers: synthesis, properties, and
biological applications. Macromolecular Bioscience 17 (5): 1600433.
12 Hu, R., Leung, N. L. C., Tang, B. Z. (2014). AIE macromolecules: syntheses, structures and
functionalities. Chemical Society Reviews 43 (13): 4494–4562.
13 Guo, Y., Shi, D., Luo, Z.-­W. et al. (2017). High efficiency luminescent liquid crystalline polymers
based on aggregation-­induced emission and “Jacketing” effect: design, synthesis, photophysical
property, and phase structure. Macromolecules 50 (24): 9607–9616.
14 Qin, A., Lam, J. W. Y., Tang, B. Z. (2012). Luminogenic polymers with aggregation-­induced
emission characteristics. Progress in Polymer Science 37 (1): 182–209.
 ­Reference 551

15 Kang, C. W., Lee, D. H., Shin, Y. J. et al. (2018). Conjugated macro-­microporous polymer films
bearing tetraphenylethylenes for the enhanced sensing of nitrotoluenes. Journal of Materials
Chemistry A 6 (36): 17312–17317.
16 Morishima, K., Ishiwari, F., Matsumura, S. et al. (2017). Mesoscopic structural aspects of Ca2+-­
triggered polymer chain folding of a tetraphenylethene-­appended poly(acrylic acid) in relation to
its aggregation-­induced emission behavior. Macromolecules 50 (15): 5940–5945.
17 Shan, Y., Yao, W., Liang, Z. et al. (2018). Reaction-­based AIEE-­active conjugated polymer as
fluorescent turn on probe for mercury ions with good sensing performance. Dyes and Pigments
156: 1–7.
18 Li, J., Li, Y., Chan, C. Y. K. et al. (2014). An aggregation-­induced-­emission platform for direct
visualization of interfacial dynamic self-­assembly. Angewandte Chemie-­International Edition 53
(49): 13518–13522.
19 Dai, C., Yang, D., Fu, X. et al. (2015). A study on tunable AIE (AIEE) of boron ketoiminate-­based
conjugated polymers for live cell imaging. Polymer Chemistry 6 (28): 5070–5076.
20 Lei, L., Ma, H., Qin, Y. et al. (2017). AIE-­active florescent polymers: The design, synthesis and the
cell imaging application. Polymer 133: 151–159.
21 Hu, R., Qin, A., Tang, B. Z. (2020). AIE polymers: synthesis and applications. Progress in Polymer
Science 100: 101176.
22 Qiu, Z., Liu, X., Lam, J. W. Y. et al. (2019). The marriage of aggregation-­induced emission with
polymer science. Macromolecular Rapid Communications 40 (1): 1800568.
23 Hu, Y. B., Lam, J. W. Y., Tang, B. Z. (2019). Recent progress in AIE-­active polymers. Chinese Journal
of Polymer Science 37 (4): 289–301.
24 Zhou, S.-­Y., Wan, H.-­B., Zhou, F. et al. (2019). AIEgens-­lightened functional polymers: synthesis,
properties and applications. Chinese Journal of Polymer Science 37 (4): 302–326.
25 Chen, F., Hu, J., Wang, D. et al. (2019). Synthesis and electroluminescent properties of through-­
space charge transfer polymers containing acridan donor and triarylboron acceptors. Frontiers in
Chemistry 7: 854.
26 Li, Y., Liu, S., Han, T. et al. (2019). Sparks fly when AIE meets with polymers. Materials Chemistry
Frontiers 3 (11): 2207–2220.
27 Kwok, R. T. K., Leung, C. W. T., Lam, J. W. Y. et al. (2015). Biosensing by luminogens with
aggregation-­induced emission characteristics. Chemical Society Reviews 44 (13): 4228–4238.
28 Wei, Q., Poetzsch, R., Liu, X. et al. (2016). Hyperbranched polymers with high transparency and
inherent high refractive index for application in organic light-­emitting diodes. Advanced
Functional Materials 26 (15): 2545–2553.
29 Liu, Q., Xia, Q., Wang, S. et al. (2018). In situ visualizable self-­assembly, aggregation-­induced
emission and circularly polarized luminescence of tetraphenylethene and alanine-­based chiral
polytriazole. Journal of Materials Chemistry C 6 (17): 4807–4816.
30 Qin, A., Lam, J. W. Y., Tang, B. Z. (2010). Click polymerization: progresses, challenges, and
opportunities. Macromolecules 43 (21): 8693–8702.
31 Gu, P.-­Y., Zhang, Y.-­H., Chen, D.-­Y. et al. (2015). Tuning the fluorescence of aggregates for end-­
functionalized polymers through varying polymer chains with different polarities. RSC Advances 5
(11): 8167–8174.
32 Huang, D., Liu, Y., Qin, A. et al. (2018). Recent advances in alkyne-­based click polymerizations.
Polymer Chemistry 9 (21): 2853–2867.
33 Shi, Y., Sun, J. Z., Qin, A. (2017). Click polymerization: the aurora of polymer synthetic
methodology. Journal of Polymer Science Part A-­Polymer Chemistry 55 (4): 616–621.
552 18  AIE-­active Polymer

34 Liang, G., Weng, L.-­T., Lam, J. W. Y. et al. (2014). Crystallization-­induced hybrid nano-­sheets of
fluorescent polymers with aggregation-­induced emission characteristics for sensitive explosive
detection. ACS Macro Letters 3 (1): 21–25.
35 Lai, C. T., Hong, J. L. (2012). Influence of molecular weight on the aggregation-­induced emission
enhancement and spectral stability of vinyl polymers containing the fluorescent
2,4,6-­triphenylpyridine pendant groups. Journal of Materials Chemistry 22 (19): 9546–9555.
36 Liu, S., Cheng, Y., Zhang, H. et al. (2018). In situ monitoringof RAFT polymerization by
tetraphenylethylene-­containing agents withaggregation-­induced emission characteristics.
Angewandte Chemie International Edition 57 (21): 6274–6278.
37 Qiao, F., Zhang, L., Lian, Z. et al. (2018). Construction of artificial light-­harvesting systems in
aqueous solution: supramolecular polymers based on host-­enhanced pi-­pi interaction with
aggregation-­induced emission. Journal of Photochemistry and Photobiology A-­Chemistry 355:
419–424.
38 Zhou, Q., Cao, B., Zhu, C. et al. (2016). Clustering-­triggered emission of nonconjugated
polyacrylonitrile. Small 12 (47): 6586–6592.
39 Zhao, E., Lam, J. W. Y., Meng, L. et al. (2015). Poly (maleic anhydride)-­alt-­(vinyl acetate): a pure
oxygenic nonconjugated macromolecule with strong light emission and solvatochromic effect.
Macromolecules 48 (1): 64–71.
40 Shang, C., Wei, N., Zhuo, H. et al. (2017). Highly emissive poly(maleic anhydride-­alt-­vinyl
pyrrolidone) with molecular weight-­dependent and excitation-­dependent fluorescence. Journal of
Materials Chemistry C 5 (32): 8082–8090.
41 Ma, J., Wang, Y., Li, X. et al. (2018) Aggregation-­induced CPL response from chiral binaphthyl-­
based AIE-­active polymers via supramolecular self-­assembled helical nanowires. Polymer 143:
184–189.
42 Yan, J.-­J., Wang, Z.-­K., Lin, X.-­S. et al. (2012). Polymerizing nonfluorescent monomers without
incorporating any fluorescent agent produces strong fluorescent polymers. Advanced Materials 24
(41): 5617–5624.
43 Zhou, X., Luo, W., Nie, H. et al. (2017). Oligo(maleic anhydride)s: a platform for unveiling the
mechanism of clusteroluminescence of non-­aromatic polymers. Journal of Materials Chemistry C 5
(19): 4775–4779.
44 Yang, C., Ren, C., Zhou, J. et al. (2017). Dual fluorescent-­and isotopic-­labelled self-­assembling
vancomycin for in vivo imaging of bacterial infections. Angewandte Chemie-­International Edition
56 (9): 2356–2360.
45 Ge, J., Lan, M., Zhou, B. et al. (2014). A graphene quantum dot photodynamic therapy agent with
high singlet oxygen generation. Nature Communications 5: 4596.
46 Jiang, L., Bai, H., Liu, L. et al. (2019) Luminescent, oxygen-­supplying, hemoglobin-­linked
conjugated polymer nanoparticles for photodynamic therapy. Angewandte Chemie-­International
Edition 58 (31): 10660–10665.
47 Wang, D., Su, H., Kwok, R. T. K. et al. (2018). Rational design of a water-­soluble NIR AIEgen, and
its application in ultrafast wash-­free cellular imaging and photodynamic cancer cell ablation.
Chemical Science 9 (15): 3685–3693.
48 Wu, W., Mao, D., Xu, S. et al. (2018). Polymerization-­enhanced photosensitization. Chem 4 (8):
1937–1951.
49 Liu, S., Zhang, H., Li, Y. et al. (2018). Strategies to enhance thephotosensitization: polymerization
and the donor-­acceptor even-­odd effect. Angewandte Chemie-­International Edition 57 (46):
15189–15193.
 ­Reference 553

50 Zhou, T., Hu, R., Wang, L. et al. (2020). AIE conjugated polymer with ultra-­strong ROS generation
ability and great biosafety for efficient therapy of bacterial infection. Angewandte Chemie-­
International Edition DOI: 10.1002/anie.201916704.
51 Gu, P.-­Y., Lu, C.-­J., Hu, Z.-­J. et al. (2013). The AIEE effect and two-­photon absorption (TPA)
enhancement induced by polymerization: synthesis of a monomer with ICT and AIE effects and
its homopolymer by ATRP and a study of their photophysical properties. Journal of Materials
Chemistry C 1 (14): 2599–2606.
52 Wang, S., Wu, W., Manghnani, P. et al. (2019). Polymerization-­enhanced two-­photon
photosensitization for precise photodynamic therapy. ACS Nano 13 (3): 3095–3105.
53 Sanchez-­Carnerero, E. M., Moreno, F., Maroto, B. L. et al. (2014). Circularly polarized
luminescence by visible-­light absorption in a chiral O-­BODIPY dye: unprecedented design of CPL
organic molecules from achiral chromophores. Journal of the American Chemical Society 136 (9):
3346–3349.
54 Liu, X., Jiao, J., Jiang, X. et al. (2013). A tetraphenylethene-­based chiral polymer: an AIE
luminogen with high and tunable CPL dissymmetry factor. Journal of Materials Chemistry C 1 (31):
4713–4719.
55 Hu, Y., Han, T., Yan, N. et al. (2019). Visualization of biogenic amines and in vivo ratiometric
mapping of intestinal pH by AIE-­active polyheterocycles aynthesized by metal-­free
multicomponent polymerizations. Advanced Functional Materials 29 (31): 1902240.
56 Wei, G., Jiang, Y., Wang, F. (2018). A new click reaction generated AIE-­active polymer sensor for
Hg2+ detection in aqueous solution. Tetrahedron Letters 59 (15): 1476–1479.
57 Hu, Y., Liang, X., Zhuang, Z. et al. (2019). Cell-­penetrating peptide modified AIE polymeric
nanoparticles by miniemulsion polymerization and application for cell fluorescence imaging.
Polymer Chemistry 10 (30): 4220–4228.
58 Zhou, Y., Zhang, L., Gao, H. et al. (2019). Rapid detection of aromatic pollutants in water using
swellable micelles of fluorescent polymers. Sensors and Actuators B-­Chemical 283: 415–425.
59 Wang, K., Lu, H., Liu, B. et al. (2018). Multi-­stimuli-­responsive fluorescence of AEE polyurethane
films. European Polymer Journal 101: 225–232.
60 Wang, K., Wang, M., Lu, H. et al. (2019). Mechano-­fluorochromic behavior of AEE polyurethane
films and their high sensitivity to halogen acid gas. RSC Advances 9 (17): 9517–9521.
61 Dong, W., Wu, H., Chen, M. et al. (2016). Anionic conjugated polytriazole: direct preparation,
aggregation-­enhanced emission, and highly efficient Al3+ sensing. Polymer Chemistry 7 (37):
5835–5839.
62 Gao, M., Wu, Y., Chen, B. et al. (2015). Di(naphthalen-­2-­yl)-­1,2-­diphenylethene-­based conjugated
polymers: aggregation-­enhanced emission and explosive detection. Polymer Chemistry 6 (44):
7641–7645.
63 Wu, Y., Qin, A., Tang, B. Z. (2017). AIE-­active polymers for explosive detection. Chinese Journal of
Polymer Science 35 (2): 141–154.
64 Wang, J., Mei, J., Yuan, W. et al. (2011). Hyperbranched polytriazoles with high molecular
compressibility: aggregation-­induced emission and superamplified explosive detection. Journal of
Materials Chemistry 21 (12): 4056–4059.
65 Liu, J., Zhong, Y., Lu, P. et al. (2010). A superamplification effect in the detection of explosives by a
fluorescent hyperbranched poly(silylenephenylene) with aggregation-­enhanced emission
characteristics. Polymer Chemistry 1 (4): 426–429.
66 Zhang, Y., Shen, P., He, B. et al. (2018). New fluorescent through-­space conjugated polymers:
synthesis, optical properties and explosive detection. Polymer Chemistry 9 (5): 558–564.
554 18  AIE-­active Polymer

67 Liu, J., Zhong, Y., Lam, J. W. Y. et al. (2010). Hyperbranched conjugated polysiloles: synthesis,
structure, aggregation-­enhanced emission, multicolor fluorescent photopatterning, and
auperamplified detection of explosives. Macromolecules 43 (11): 4921–4936.
68 Liu, J., Zhong, Y., Lu, P. et al. (2010). A superamplification effect in the detection of explosives by a
fluorescent hyperbranched poly(silylenephenylene) with aggregation-­enhanced emission
characteristics. Polymer Chemistry 1 (4): 426–429.
69 Wang, K., Zhang, X., Zhang, X. et al. (2015). Fabrication of cross-­linked fluorescent polymer
nanoparticles and their cell imaging applications. Journal of Materials Chemistry C 3 (8):
1854–1860.
70 Hu, R., Zhou, T., Li, B. et al. (2020). Selective viable cell discrimination by a conjugated polymer
featuring aggregation-­induced emission characteristic. Biomaterials 230: 119658.
71 Hu, R., Chen, X., Zhou, T. et al. (2019). Lab-­in-­cell based on spontaneous amino-­yne click
polymerization. Science China-­Chemistry 62 (9): 1198–1203.
72 Wang, Z., Wang, C., Fang, Y. et al. (2018). Color-­tunable AIE-­active conjugated polymer
nanoparticles as drug carriers for self-­indicating cancer therapy via intramolecular FRET
mechanism. Polymer Chemistry 9 (23): 3205–3214.
73 Zhao, J., Dong, Z., Cui, H. et al. (2018). Nanoengineered peptide-­grafted hyperbranched polymers
for killing of bacteria monitored in real time via intrinsic aggregation-­induced emission. ACS
Applied Materials & Interfaces 10 (49): 42058–42067.
74 Chen, Y., Han, H., Tong, H. et al. (2016). Zwitterionic phosphorylcholine-­TPE conjugate for
pH-­responsive drug delivery and AIE active imaging. ACS Applied Materials & Interfaces 8 (33):
21185–21192.
75 Long, Z., Mao, L., Liu, M. et al. (2017). Marrying multicomponent reactions and aggregation-­
induced emission (AIE): new directions for fluorescent nanoprobes. Polymer Chemistry 8 (37):
5644–5654.
76 Wang, Z., Wang, C., Gan, Q. et al. (2019). Donor-­acceptor-­type conjugated polymer-­based
multicolored drug carriers with tunable aggregation-­induced emission behavior for self-­
illuminating cancer therapy. ACS Applied Materials & Interfaces 11 (45): 41853–41861.
555

19

Liquid-­crystalline AIEgens: Materials and Applications


Kyohei Hisano, Supattra Panthai, and Osamu Tsutsumi
Department of Applied Chemistry, College of Life Sciences, Ritsumeikan University, Kusatsu, Japan

19.1 ­Introduction

Light-­emitting liquid crystals (LCs) are one of the key state-­of-­the-­art materials in a significant variety
of applications such as organic light-­emitting diodes (OLEDs), light-­emitting LC displays, polarized
organic lasers, and optical sensors  [1–3]. The luminescence behavior of light-­emitting LCs can be
controlled by designing the molecular structure and/or molecular orientational structure in an aggre-
gate. Among the advances in LC technology, aggregation-­induced emission (AIE), firstly discovered
by Tang et al., has attracted much attention for developing a novel chromophore because the AIE
luminogen (AIEgen) makes it possible to obtain a bright emission in aggregates, unlike the case of
conventional luminogens with aggregation-­caused quenching (ACQ)  [4]. However, in contrast to
AIEgens, AIE-­active LCs may not enjoy a compatible growth. This is because the general strategy for
designing and synthesizing AIE-­active LCs is currently still under development. In this chapter, we
highlight some recent design examples of AIE-­active LCs and their emerging applications, focusing
on organic AIEgen-­based organic mesogens, metallomesogens, and polymeric materials. We hope
that this will help develop a strategy for the design of AIE-­active LCs and a wide variety of high-­
performance potential applications, for example, linearly polarizing luminescence LCs. These are
promising materials in LC displays with simpler device design, cost-­effectiveness, lower power con-
sumption, higher brightness and color contrast, and wider viewing angles.
LCs have both liquid-­like fluidity and crystal-­like orientational order, described as the fourth
state of the matter, which is a thermodynamically stable mesophase [5]. The unique feature of LC
molecules is their ability to communicate with each other, which is called the molecular coopera-
tive effect, and the reversibility of the molecular orientational order when manipulated by external
stimuli such as temperature, shear, topographic effects, and electro-­magnetic fields [6]. For exam-
ple, one of the most advanced technologies for inscribing the molecular orientational order of LCs
is the photoalignment method, where the orientation of LCs with the incorporation of a small
amount of photochromic dopants, for example, azobenzene derivatives, can be manipulated along
the polarization state of an incident light [7–9]. Owing to the drastic development of these meth-
ods, various LC-­based materials with a controlled molecular orientation, not only in low-­molecular-­
weight LCs but also in polymeric and supramolecular systems, have been achieved. Thus, they
have enabled numerical applications, from LC displays, now the pervasive technology in our daily

Handbook of Aggregation-Induced Emission: Volume 1 Tutorial Lectures and Mechanism Studies, First Edition.
Edited by Youhong Tang and Ben Zhong Tang.
© 2022 John Wiley & Sons Ltd. Published 2022 by John Wiley & Sons Ltd.
556 19  Liquid-­crystalline AIEgens: Materials and Applications

life, to next-­generation applications such as smart glasses, smart windows, flexible sensors, and
soft actuators with the desired molecular orientational patterns [10–13]. Nevertheless, the synthe-
sis of light-­emitting LCs and their practical applications in devices are still scientifically challeng-
ing because of the ACQ effect of luminophores, even in host viscous LCs. To overcome this issue,
the development of light-­emitting LCs with AIE properties is highly desired.
Typical organic LCs are molecules with both a mesogenic core, which is a rigid anisotropic subu-
nit of a rod-­like (calamitic LC) or disk-­like shape (discotic LC), and a flexible chain, for example,
alkoxy and alkyl chains. Such LCs can form various mesophases depending on the molecular
structure at an appropriate temperature (thermotropic LC) and/or solvent concentration (lyotropic
LC). For example, the most common mesophase of LCs is the nematic phase, where LCs have a
one-­dimensional orientational order with a director parallel to the molecular axes. An emerging
AIE-­active luminogen is a propeller-­shaped conjugated molecule with a unit of hexaphenylsilole
(HPS) or tetraphenylethene (TPE). In the aggregates of such molecules, an internal molecular
motion, which activates the nonradiative decay of the excited state, is restricted, and π–π stacking
interaction, which causes the nonradiative decay, is prevented because of the steric hindrance of
their bulky substituents. Therefore, the luminescence quantum yield is dramatically enhanced in
those aggregates. Recently, the incorporation of flexible alkoxy or alkyl chains into rigid AIE-­active
units acting as mesogens has been found to be an advanced concept for designing AIE-­active
LCs [14–17]. Because of the geometric anisotropy arising from the AIE unit, the AIE-­active LCs
frequently show a columnar phase or a smectic phase. To generate a variety of mesophases, another
approach was also developed: amalgamating AIEgens with mesogens  [18]. This allows one to
design the mesophase and AIE activity separately, and some AIE-­active LCs show a nematic phase
with a very high quantum yield. Furthermore, d 10 transition-­metal complexes such as gold(I) com-
plexes are of considerable interest because they possess inherently high luminescence intensity in
their condensed phases and unique luminescence behavior depending on the aggregate structures,
which is attributed to the interactions between metals, termed as metallophilic interactions [19,
20]. Such metal complexes also act as mesogen units, called metallomesogens, and exhibit the
same mesophase as that of typical organic LCs, and thus enable AIE-­active LCs. Such enrichment
in the variety of AIE-­active LCs would open the pathway for achieving practical applications, espe-
cially with polarizing emitting behaviors and/or stimuli-­responsive color change behaviors.
The goal of this chapter is to review and highlight, through selected recent examples of AIE-­
active LCs based on the emerging AIEgens and metallomesogens, some significant examples of
unique luminescence properties such as color tuning, color “on–off” switching, polarizing emis-
sion, and some twists in the development of potential applications. Some important reviews on
AIEgens have been reported in the past years  [1–3], which we recommend for a more detailed
understanding. We start by outlining some recent advances in AIE-­active LCs for exploring the
general design concept and then propose some key applications of these materials, finally sum-
marizing the remaining challenges and future perspectives.

19.2 ­Materials: Molecular Design

19.2.1  Discotic LC AIEgen


The development of AIE-­active LCs has been pioneered by a series of studies of low-­molecular-­
weight organogelators, which cause self-­assembly of small gelator molecules and form supramolecu-
lar networks through noncovalent interactions, such as π–π stacking and/or hydrogen bonding, and
van der Waals interactions  [21–23]. Li et  al. reported in 2007 that some organogelators with
19.2 ­Materials: Molecular Desig 557

bisurea-­functionalized naphthalene units can form a columnar LC mesophase, showing a strong


fluorescence emission depending on the aggregation of the fluorophore and exhibiting sensitivity
under temperature and chemical stimuli [24]. Soon after, Lai et al. reported two novel organogelators
based on 2,3,4,5-­tetraphenylsilole functionalized with long-­chain alkoxydiacylamido (1, 2) with both
AIE behavior and columnar mesophase over a wide temperature range (Figure 19.1) [25]. However,
such tetraphenylsilole derivatives have a noncoplanar geometry, which prevents aggregation and
generally makes the formation of LC phases and/or gels difficult. The drawbacks were overcome by
incorporating amide groups into the tetraphenylsilole units, inducing intermolecular hydrogen
bonding. Numerous studies have revealed that mesogenic molecules with some intermolecular inter-
actions, such as dipole–dipole interaction, hydrogen-­bonding, halogen-­bonding, and metallophilic
interactions, act as organogelators and, in some cases, lead to the formation of smectic or discotic LC
phases. Following these pioneering studies including metallogens reported in the very early
stage [26], many other AIE-­active organogels are developed, as reviewed in 2014 [27].
A wide variety of AIE-­active LCs have been explored, but the molecular design is still constrained
by a few types of AIEgen, such as the disk-­like shape of TPE and HPS (forming smectic and/or
columnar mesophases), rod-­like cyanostilbene and tolane (forming smectic and/or nematic meso-
phases), and metallomesogens (forming all mesophases). In previous studies, AIE-­active LCs have
been mainly designed using AIEgen as a mesogenic unit and incorporating them with flexible
alkyl/alkoxy chains, as summarized in Figure 19.2. For example, Cho et al. investigated the effect
of peripheral nonpolar and polar exterior chains on the AIE properties of bulk TPE derivatives
(11–13, Figure 19.3) [28]. They found that the bulk TPE derivatives with nonpolar dodecyl and
polar di(ethylene oxide) chains are in columnar LC phase and polycrystalline phases, respectively,
at room temperature (RT). Interestingly, the difference in structural mobility, arising from the
number of peripheral flexible chains and their intrinsic polarity, might lead to the numerous crys-
tallographic defects and exhibit the undesirable nonemissive electronic state. This means that the
structural mobility due to a flexible chain is crucial for high AIE efficiency for minimizing the
number of undesirable nonemissive local sites in aggregates of the TPE derivatives. In addition to
such research regarding the effect of the peripheral flexible chains on an existing AIE core, we
believe that further exploration of other AIE core design strategies in AIE-­active LCs may give us
numerous opportunities to develop novel applications.
To pursue a new AIE unit for generating the LC phase because most of the existing methods for
generating AIE-­active LCs require complicated and harsh synthetic conditions with relatively low
yields, a unique approach is proposed by using ionic self-­assembly (ISA). ISA is a facile and efficient
mechanism to construct novel supramolecular structures such as supramolecular columnar LCs
(14) by binding oppositely charged tectonic discotic units together via electrostatic interactions. In

R1 = (CH2)11CH3
RO O O
OR R2 = (CH2)15CH3
Si
N N
RO H H OR

RO OR
1,2

Figure 19.1  A representative molecular structure showing both AIE behavior and columnar mesophase
over a wide temperature range.
558 19  Liquid-­crystalline AIEgens: Materials and Applications

CN O
C12H25O H3CO
OC12H25 O O
3 4

R 5O OR5 OR6
R7

7 R7
R 5O OR5 OR6
5 6

R5 = C12H24O C5H11

R6 = C11H22O O

R7 = O CN
O
N

R 8O OR8

R8 O O O OR8
O O
R 8O OR8

N N
N N
Ir+
N N

R8 = C6H12O

OC12H25 C12H25O OC12H25


C12H25O

OC12H25
C12H25O C12H25O
OC12H25 O
O
C12H25O O O
OC12H25

OC12H25
S
C12H25O O
OC12H25
C12H25O S O OC12H25
C12H25O 9
O
C12H25O O OC12H25
10
C12H25O OC12H25

Figure 19.2  Some representative molecular structures of AIE-­active LCs have been reported in the
literatures. Source: Reproduced from Ref. [17] with permission from the Royal Society of Chemistry.
R1 R1
R2 R2
N N
N N
R3 N N R3 2.0 11 12 13
Relative fluorescence intensity

at blue edge
1.5 at center
at red edge
at red tail

1.0
N N N N
R1 N N R1

R2 R2
R3 R3 0.5

11: R1 =R2 =R3 =O(CH2)11CH3


12: R1 =R2 =O(CH2)11CH3, R3=H 50 100 150 200 50 100 150 200 50 100 150 200
13: R1 =R2 =O(CH2)2O(CH2)2OCH3, R3=H Temperature (°C)

Figure 19.3  Molecular structure of the TPE derivatives (11–13) for investigation of the effect of peripheral flexible chains on LC
behavior (left) and temperature-­dependent relative emission intensities at various detection wavelengths from the blue-­to-­red-­edge
region of the emission bands of the TPE derivatives (right). Source: Adapted with permission from Ref. [28]. Copyright (2016)
American Chemical Society.
560 19  Liquid-­crystalline AIEgens: Materials and Applications

2016, Ren et al. first proposed the ISA process for creating AIE-­active LCs by using a TPE unit as the
AIE core and a dimethyldioctadecylammonium bromide as the ionic element (Figure 19.4) [29].
The resultant materials successfully formed a hierarchical supramolecular structure showing a
columnar mesophase and excellent luminescent properties, as determined by differential scanning
calorimetry (DSC) measurements, polarized optical microscopy (POM) observations, and one-­and
two-­dimensional X-­ray diffraction (XRD) experiments. The AIE activity of the ISA complex was
quantitatively investigated by measuring the fluorescence quantum yield in both the solution and
solid states. This revealed that the complex clearly exhibited AIE activity, where the quantum yield
in DMF was 0.68%, but the value in the solid film was up to 46%, meaning that the AIE factor (equal
to the quantum yield in the film divided by that in the solution) is 67.6. This is the highest level of
AIE-­active LCs ever reported.
Very recently, a novel discotic AIE-­active LC has been proposed by Yang et  al., where they
employed another material design concept for separating the LC-­forming units (using a porphyrin
derivative) and an AIE unit (using cyanostilbene) [30]. It is well known that the planar-­conjugated
aromatic structures of porphyrins enable one to generate a columnar mesophase forming excellent
supramolecular structures merely by incorporating alkyl/alkoxy chains on the porphyrin unit.

(a)
C18H37 + C18H37
N N+
C18H37 COO COO
C18H37

O O


O O
14

C18H37 + C18H37
N COO COO N+
C18H37 C18H37

(b) (c)
800
fw(vol%)
0%
600 20%
40%
60%
PL intensity

400 80%
90%

200

0
400 450 500 550 600 650 700
Wavelength (nm)

Figure 19.4  (a) A molecular structure (left) and a schematic diagram (right) of a TPE derivative (14). (b) A
model of the helical column in the lattice of the molecule via ISA process. For clarity, the peripheral alkyl
chains are not shown in the lattice, and the four benzene rings of the TPE core are drawn in different colors.
(c) Photoluminescence spectra of the molecule in the DMF and DMF/water mixtures. The inset graph is a
film of the molecule taken under a UV illumination. Concentration = 1.0 × 10−5 M. Excitation
wavelength = 365 nm. Source: Adapted with permission from Ref. [29]. Copyright (2016) American Chemical
Society.
19.2 ­Materials: Molecular Desig 561

Unfortunately, as with other typical discotic luminescent LCs, porphyrin LCs also exhibit ACQ but
no AIE effect. As first reported, a porphyrin LC (15), which forms a hexagonal columnar meso-
phase between 70 and 120 °C with strong fluorescence in both the solution and aggregated states,
was designed and synthesized by introducing polyglycol–diphenylacrylonitrile units onto the por-
phyrin skeleton (Figure 19.5). Here, the polyglycol unit acts as a flexible chain and is key for the
manifestation of LC properties. The diphenylacrylonitrile unit is a classical AIE group with an
emission wavelength of approximately 425 nm in the aggregated state, which acts as a photoan-
tenna for inducing fluorescence resonance energy transfer (FRET) to porphyrin with an absorp-
tion wavelength in the same region. This cooperative mechanism of AIE and FRET provided the
great photophysical properties with a large pseudo-­Stokes shift of ~210 nm and a high quantum
yield of 0.12 in the solid state. Owing to the near-­infrared luminescence with a large Stokes shift
and the hydrophilicity arising from the polyglycol unit, this AIE-­active material exhibited low bio-
toxicity and excellent fluorescence bioimaging ability in living cells.
Another emerging field is the use of metallomesogens [19, 20]. One of the fascinating features of
the introduction of a metallic fragment in organic compounds is that it enables the modification of
the physical properties, such as conductivity, color, luminescence, and magnetism. This expands
the possibility of designing new materials for developing novel applications where the coordina-
tion geometry of complexes plays a crucial role. Among various advanced metallomesogens, Au(I)
has a strong affinity for linear coordination and metal–metal interactions, allowing us to generate
supramolecular structures and form both calamitic and discotic mesogenic structures, resulting in
various LC mesophases. Therefore, molecules containing Au(I) ions are examples of AIEgens that
exhibit strong photoluminescence in condensed phases, and some Au(I) complexes exhibit AIE-­
active LCs.
In 2018, Gimeńez et al. reported a series of cyclic trinuclear Au(I)-­pyrazole complexes (16) [31].
The complex dispersed in a poly(methyl methacrylate) film exhibited a bright-­red/deep-­red RT
phosphorescence with exceptionally high quantum yields of 90% and a large Stokes shift of
~400 nm (Figure 19.6). The designed ligand, 4-­hexyl-­3,5-­dimethylpyrazolate, bears only one short
hexyl chain in the 4-­position of the pyrazole ring. This imparts good solubility properties for mix-
ing with polymer matrices and for film processing, and it also induces columnar LC phases.
Furthermore, the Au(I) complex showed thermochromism from red to blue under cooling and
changed the color response to silver ions, indicating its applicability as a sensing material. In addi-
tion, other metal complexes (Ag and Cu) have been synthesized. A detailed crystallographic analy-
sis and the thermal and luminescence properties revealed that the RT phosphorescence arises from
metallophilic interaction and the stronger metallophilic interaction (Ag > Au > Cu) tends to form
stable mesophases (the phase-­transition temperature to the columnar phase for Au is 145 °C and
for Ag is 193  °C; there is no mesophase for Cu). Such color tunability and “on/off” switching
behavior are very useful in various sensing applications such as bioimaging and chemosensors.

19.2.2  Calamitic LC AIEgens


Calamitic LC is a class of LCs with rod-­like molecular shapes and typically exhibits smectic and/or
nematic phases. In particular, a nematic phase is very common and important for LC applications
such as LC displays, and nematic AIE-­active LCs have a great potential to manifest linearly and
circularly polarized luminescence (CPL). Unfortunately, considering the geometrical restriction of
conventional AIEgens with propeller-­like molecular shapes such as TPE and HPS, the synthesis of
calamitic AIE-­active LCs is scientifically challenging. Thus, one of the promising ways for achiev-
ing an AIE-­active materials with the nematic LC phase is the use of a mixture of a very small
562 19  Liquid-­crystalline AIEgens: Materials and Applications

Figure 19.5  (a) Molecular structure and synthetic route of a porphyrin derivative (15). (b) Proposed
molecular stacking mode for the hexagonal columnar mesophase of the porphyrin derivative. (c)
Fluorescence spectra of the porphyrin derivative (1 mM) in the THF/water mixture with different water
fractions (excitation at 420 nm). (d) Fluorescence spectra of the porphyrin derivative (1 mM) in the THF/
water mixture with different water fractions (excitation at 350 nm). Inset: the line plot of fluorescence
intensity changes from 0 to 90% water fractions of the porphyrin derivative and fluorescence photos (under
UV-­365 light) of the porphyrin derivative in THF/water mixtures at fw = 10, 50, and 90%. Source: Reproduced
from Ref. [30] with permission from the Royal Society of Chemistry.
19.2 ­Materials: Molecular Desig 563

Figure 19.5  (Continued)

portion of AIEgen (< a few wt%) with a host nematic LC. Alternatively, the development of a
single-­component system of calamitic AIE-­active LCs showing nematic and/or smectic phases is
also important for enabling high-­performance applications such as light-­emitting LC displays, but
it has rarely been reported. For circumventing the issue, in 2013, Tang et al. proposed a new syn-
thetic strategy of applying AIE phenomenon to the molecular design of LCs, where they employed
a typical LC molecule of tolane as a mesogenic unit  [32]. In a series of researches, they have
564 19  Liquid-­crystalline AIEgens: Materials and Applications

(a)
O Semicarbazide C6H13
O
hydrochloride, H2O
N
N
C6H13 H
16-1 16-2

C6H13 C6H13
N Au N
N N
Au Au
[AuCl(S-(CH3)2)], KOH, Acetone N N

C6H13

16
(b) (c)
1,0
Relative emission intensity

0,8

0,6

0,4

0,2

0,0
200 300 400 500 600 700 800
Wavelength (nm)

Figure 19.6  (a) Molecular structure and synthetic route of Au(I) trinuclear complex (16). (b) POM
microphotographs (crossed polarizers) of the texture of complex 16 at 142 °C. (c) Powder sample of the
Au(I) complex. Excitation spectra at RT (red line) and at 77 K (orange line). Emission spectra at RT (dark blue
line) and 77 K. Source: Reproduced from Ref. [31] with permission from the Royal Society of Chemistry.

succeeded in synthesizing the tolane-­based AIE-­active LCs, showing a nematic phase and/or a
smectic phase  [33, 34]. According to the AIE phenomena, the tolane-­based luminogen can be
viewed as a propeller-­like molecule with two blades (phenyl groups): thus, the aggregation enables
us to restrict the intramolecular rotations of the two blades around the ethynyl unit, which is the
origin of ACQ, and leads to AIE properties.
Another fascinating approach is the introduction of a metallomesogen unit. As mentioned above, the
geometry of an Au(I) complex is linear, indicating that the system has the potential to form calamitic
molecules exhibiting a nematic phase even with a single component without any dopants. Recently, our
group designed and synthesized a set of Au(I) complexes: to improve the intermolecular aurophilic
interaction, a crucial factor for exhibiting AIE activity, the steric hindrance due to bulky ligands around
Au(I) ions is minimalized by employing the simpler coordinating ligands of an ethynylbenzene group
conjugated to various lengths of alkoxy chains and n-­alkyl isocyanide (18–21) [15, 35]. The detailed
investigation of the LC behavior, crystallographic data, and photophysical properties revealed that the
19.2 ­Materials: Molecular Desig 565

complexes easily formed dimers through an aurophilic interaction between molecules in any phase,
even in the isotropic phase (Figure 19.7). These dimeric structures acted as unit mesogens in the LC
phase (nematic and smectic phases), observed at a relatively low temperature (~100  °C), which is

(a)

O Au N

N Au O

N Phase

O Au N

N Au O

Smectic

(b) (c)
(f)
2000
Intensity (Norm.)
Emission intensity (Arb.)

In crystal
1500

(d) (e) 1000


400 500 600
Wavelength (nm)

500

In solution
0
400 450 500 550 600
Wavelength (nm)

Figure 19.7  (a) (Top) Structural model of a mesogen unit formed by the aurophilic interaction in the
condensed phase. (Bottom) Schematic illustration of molecular packing in LC phases. (b–e) POM micrographs
of a series of Au(I) complexes 18–21 with alkyl chains of different lengths of –CnH2n + 1– with 5 (nematic), 6
(nematic), 7 (nematic), and 8 (smectic), respectively. (f) Photoluminescence spectra of the Au(I) complex with a
chain length of 5 in the crystal (red; excitation, 340 nm) and CH2Cl2 solution (blue, 2.7 × 10−4 mol/l;
excitation, 330 nm). The normalized photoluminescence spectra in both the crystal and the solution are
shown in the inset. Source: Reproduced from Ref. [15] by permission of The Royal Society of Chemistry.
566 19  Liquid-­crystalline AIEgens: Materials and Applications

preferable for designing light-­emitting LCs because high temperature enhances the nonradiative transi-
tion of an excited state, meaning thermal deactivation and causing a drastic decrease in quantum yield.
Interestingly, all Au(I) complexes showed RT phosphorescence at a peak wavelength of ~450 nm, and
the highest quantum yield was 50%. According to other studies of metallomesogens, our proposed
material design can be applied to other metals, with the potential to be a platform for generating­
various calamitic AIE-­active LCs, and it paves the way for developing novel polarizing luminescence
applications.

19.2.3  Polymeric LC AIEgens


Polymeric AIE-­active LCs have the potential to exhibit an upgraded performance rather than a
low-­molecular-­weight material from the viewpoint of film processing, design of other functionali-
ties, hierarchical orientational control, and so on. A wide variety of studies have been conducted
and have achieved many polymeric systems by incorporating an AIEgen into polymeric net-
works [36, 37]. Among them, LC elastomers are one of the emerging polymeric LC systems. LC
elastomers, where the polymer network structure is strongly coupled with a molecular orienta-
tional order and exhibits a fascinating reversible shape change coincident with the orientational
order, change under the application of external stimuli  [10, 11]. A representative example is a
polydomain to a monodomain transition under external stimuli. The pristine LC elastomer with a
polydomain structure with macroscopically randomly oriented mesogens and polymer networks
became unidirectionally oriented as the uniaxial strain imposed on the elastomer increased.
Simultaneously, after the removal of the strain, the elastomer gradually returned to its initial shape
and molecular orientation. This unique reversible phenomenon opens up a new window of next-­
generation applications of stimuli-­responsive materials, for example, sensors and actuators.
Considering that the photophysical properties of AIEgen are very sensitive to the external environ-
ment as well as its aggregate structure, the AIE-­active LC elastomer exhibits unique mechano-­
optical properties.
In 2018, Yang et  al. first fabricated AIE-­active LC elastomers, where AIEgens of a TPE
derivative (22) directly bonded to an LC polymer network as a cross-­linker (concentration:
10  mol%) through an in situ two-­step polymerization/crosslinking method in bis-­acrylic
materials (Figure  19.8)  [38]. They fabricated monodomain AIE-­active LC elastomers and
investigated the thermal effect on both shape morphing, monodomain-­to-­polydomain transi-
tion upon heating, and photophysical property switching. DSC and XRD measurements
revealed that the AIE-­active LC elastomers exhibit phase-­transition behavior: SmecticC-­89 °
C-­nematic-­110  °C-­isotropic (on heating) and Isotropic-­103  °C-­nematic-­82  °C-­smectic (on
cooling). Unfortunately, the result indicated that the change in photophysical properties was
predominately determined by the temperature effect, and the luminescence intensity
decreased as the temperature increased, even with the macroscopic shape deformation of the
AIE-­active LC elastomer. Basically, the heating causes LC to isotropic phase transition, lead-
ing to the aggregation of AIE units by relaxation of the polymer network. Thus, it is expected
to increase the luminescent intensity; however, no effect on AIE activity was observed in the
report. Nevertheless, these experimental results and hypotheses help us to open a new design
concept of thermomechanically controllable fluorescent soft actuators. For example, AIE-­
active main-­chain-­LC is one of the promising candidates where the relaxation of the polymer
network directly affects the relaxation of AIEgens, with the potential to make them more
aggregate.
19.3 ­Applications of LC AIEgen 567

Figure 19.8  (a) Chemical composition of the AIE-­active LC elastomer film. (b and c) Photographs of the
thermally induced actuation behaviors of the elastomer film exposed to (b) a UV 365-­nm radiation and (c)
ambient light (scale bar = 5 mm). Source: Reproduced with permission from Ref. [38]. Copyright (2018)
American Chemical Society.

19.3  ­Applications of LC AIEgens

19.3.1  Linearly Polarized Luminescence


As described in the previous section, AIE-­active LCs from discotic to calamitic and polymeric sys-
tems have been extensively studied by many research groups and are becoming a platform for
generating next-­generation applications such as linearly polarized luminescence, CPL, and sens-
ing materials in life sciences. The most explored application is an LC display, which is now a per-
vasive technology in our daily life but still has some drawbacks: low brightness and low energy
efficiency because the output light passes through a commercially available polarizer from white
unpolarized backlight, leading to other issues of low contrast ratio and viewing-­angle dependence.
Alternatively, the use of AIE-­active LCs is a trend [32, 39–41]. In contrast to a conventional back-
light, AIE-­active LCs exhibiting a nematic phase show linearly polarized luminescence only at the
localized region where a voltage is applied. Thus, this drastically reduces the consumption of
energy and increases the brightness, contrast ratio, and wide viewing angle. Furthermore, indus-
trial fabrication also becomes cost-­effective.
Recently, Tang et al. developed a new pathway for the fabrication of AIE-­active LCs by the amalga-
mation of a mesogenic unit, ranging from small molecules to polymeric systems, and the AIE unit of
the TPE core. Owing to the incorporation of an LC unit, such material systems are easily miscible
with host LCs. The material first synthesized was TPE with tolane derivatives (termed TPE-­PPE, 17),
exhibiting thermotropic smectic phase behavior and AIE properties (Figure 19.9) [32]. By doping the
568 19  Liquid-­crystalline AIEgens: Materials and Applications

C 3H 7 C 3H7

C 3H 7 17
C 3H7

Eoff Eon

Figure 19.9  Molecular structure of TPE-­PPE (17) and photographs of the light-­emitting LCD device in the
electric field-­off and field-­on states using a light-­emitting LC mixture of 17 and a host nematic LC. Source:
Adapted from Ref. [32] with permission from Wiley.

material into a host nematic LC, the mixture showed a nematic phase. More interestingly, this mix-
ture could overcome the ACQ effect of conventional luminescent LCs. Furthermore, thanks to the
manifestation of the nematic LC phase, TPE-­PPE can be used as a basic component for achieving
linearly and/or circularly polarized luminescent applications, which will be described later in detail.
This strategy is simple and enables the creation of various AIE-­active LCs merely by slight modifica-
tion of the AIE core. Shortly after, they applied the material for developing a novel light-­emitting
liquid crystal display (LCD) system (Figure  19.9)  [41]. First, they employed the ultraviolet (UV)-­
sensitive LC photoalignment material sulfonic-­dye-­1 (SD1) coated over substrates of a cell to realize
the photocontrolled alignment and inscribing a molecular orientational pattern. By irradiation of the
SD1 photoalignment layer with a polarized UV light through a photomask, the molecules located at
the irradiated region were aligned perpendicular to the UV light polarization. Then, the remaining
area was exposed to a polarized UV light in the direction perpendicular to the polarization. After that,
the mixture of the host LC and TPE-­PPE was filled into the empty cell. As a result, a number pattern
was generated in this luminescent LCD device under a polarizer. By rotating the polarizer, the paral-
lel or perpendicular area of the device to the polarizer direction is altered. The light-­emitting and
dark regions were then switched. When an electric field is applied to the LC cell, the LC mixture will
be vertically aligned, and no patterned image can be seen. These results clearly indicated that the
resultant system exhibited a linearly polarized luminescence and developed a light-­emitting LC
device. As mentioned above, the material system still has a low quantum yield (~21%), but this does
not fundamentally limit the practical application of light-­emitting LC displays. Further increase of
the quantum yield of AIE-­active LCs will develop future light-­emitting LC displays, polarizing lasers.

19.3.2  Circularly Polarized Luminescence


The interest in circularly polarized light is drastically increasing in the modern era, as it enables us
to develop various advanced optical and photonic applications such as three-­dimensional displays,
19.3 ­Applications of LC AIEgen 569

information communication, and polarization sensors  [42–44]. Thus, CPL is highly desired.
Theoretically, we can define the purity of right CPL or left CPL by the luminescence dissymmetry
factor, glum. This represents the ratio of the intensity difference between the right and left CPL
divided by the average of the total luminescence intensity. More specifically, this value can be
expressed by the relationship between the electric dipole transition moment μgn and the imaginary
magnetic dipole transition moment mgn, as shown below [42]
gn
m gn
glum 2 2 (19.1)
gn
m gn

According to the equation, a larger glum is achieved when the electric dipole transition is forbid-
den but the magnetic one is allowed. However, typical luminescent materials have a much smaller
magnetic dipole transition moment than electric materials; thus, the luminescence intensity dras-
tically decreases when a large glum is obtained. This is the reason why most CPL organic materials
show very low glum in the range of 10−5 to 10−2 only by designing the molecular structure. Therefore,
the development of materials with both strong CPL signals and high glum is a hot research field
from the viewpoint of hierarchical structure from nano-­to microlevels.
Recently, many studies on CPL organic materials with a controlled hierarchical structure have
been conducted, and various material systems have been presented, for example, chiral supramo-
lecular self-­assembly. Akagi et al. first found that highly ordered helically π-­stacked structures of
conjugated polymeric systems allow the generation of CPL with much higher glum on the order of
10−1 [45]. Soon after, Liu et al. in 2012 proposed AIE-­active tetraphenylsilole derivatives with chiral
units (mannose-­containing periphery chains) forming a supramolecular structure that showed a
strong CPL with a glum in the range of −0.08 and −0.32, keeping the bright emission intensity due
to the AIE activity with the fluorescence quantum efficiency by 136-­fold (quantum yield of 0.6% in
the solution versus that of 81.3% in the solid state) [46]. The number of studies on the use of cho-
lesteric LC phase (chiral nematic phase) is increasing enormously because of the possibility of
manifesting an extremely high glum. This is partially explained by the selective reflection of circu-
larly polarized light arising from the helically twisted molecular orientation in a cholesteric LC,
which acts as a pseudo-­one-­dimensional periodic structure. Therefore, CLCs can reflect an inci-
dent light due to Bragg’s law and thus the reflection peak wavelength of CLCs when the angle of
the incident light is normal. This can be explained by Equation 19.2:
nave P (19.2)
peak

where P is the helical pitch of the orientational structure of LCs, and nave is the average of the
refractive index of the LCs. The initial discovery of the phenomenon drastically enhancing glum in
the cholesteric LC medium was discovered by Chen et al. in 1999, when an achiral luminophore
was doped in the cholesteric LC film. Interestingly, in 2016, Tang et  al. demonstrated the great
influence of AIE-­active LCs on a highly emissive CPL material in the LC phase by employing the
aforementioned TPE-­PPE dissolved in a cholesteric LC medium [47]. Most recently, Cheng et al.
demonstrated that achiral AIE-active dyes (guest 2) doped in a cholesteric LC medium showed
very high glum values, as high as +1.42/−1.39, and the luminescence peak wavelength can be tuned
from 403 to 601 nm merely by changing the doped AIEgen (Figure 19.10) [48]. In addition, the
quantum yield of the CPL reached 20.4%, which is much higher than that of conventional organic
luminophores. This indicates that the cholesteric LC medium acted not only as a simple circularly
polarized light reflector but also as an effective chirality transfer.
570 19  Liquid-­crystalline AIEgens: Materials and Applications

(a) R
CN
1 = –OC6H13-n
CN
N
2= N

OC6H12O
OC6H12O R

N S 3= N
CN CN
O 4= –Br
O
Chiral dopant enantiomers
(Guest 1)
AIE-active dyes (Guest 2)

(b)

Mixing Doping

N-LC(Host-E7)

1 mole ratio Guest 2

1wt% Guest 1 N*-LCs AIE-N*-LCs

Figure 19.10  (a) Molecular structure of AIE-active LC systems. (b) Schematic diagram of self-­assembly in
the cholesteric LC medium. Source: Adapted from Ref. [48] by permission of The Royal Society of Chemistry.

An interesting new development in the design of AIE-­active LC with CPL property is the drive from
the system of doping AIE-­active LCs into commercially available cholesteric LCs to a rational system
of single AIE-­active cholesteric LCs. Tang et al. proposed a new molecular design strategy to incorpo-
rate a chiral LC unit directly into AIEgen via covalent bonds (e.g. dicyanodistyrylbenzene-­cholesterol
derivatives), instead of mixing several components that are with potential of phase separation leading
to the degradation of material performance. This strategy allows us to develop the elaborated CPL
material with a high glum value (>10−1), a large quantum yield (up to ~70%), without phase separation,
and/or CPL switching behavior depending on their self-­assembled aggregate structures [49, 50].
The rational design of AIE-­active LCs with both high quantum yield and polarization is still scien-
tifically challenging due to the limited number of available AIEgen cores such as TPE or HTP for disc-­
like shape and tolane or stilbene for calamitic shape. However, the recent development of AIE-­active
LCs shed light on the further development of the field not only by modifying a molecular structure but
also by designing mesoscopic structures, e.g. supramolecular fiber, cholesteric LC phase, and lamellar
assembly, formed via intermolecular interactions including metallomesogens. This enables us to
manipulate the polarization of luminescence without the degradation of luminescence intensity.
 ­Reference 571

19.4  ­Conclusion

The design and synthesis of novel AIE-­active LCs are an emerging trend in the modern era for
the development of various optics and photonics applications, such as OLEDs, light-­emitting
LC displays, sensors, and stimuli-­responsive optical elements. The most important feature for
the development of AIE-­active LCs is not only the generation of polarizing luminescence but
also the functionalization of AIE-­based materials, for example, introducing reversible stimuli-­
responsive behavior such as color tunability and color switching “on/off” effects. However, the
general design strategy for synthesizing AIE-­active LCs is still unclear. In this chapter, we intro-
duced the key current research for exploring the general concept of designing and creating
AIE-­active LCs. The most simple and powerful method is the incorporation of peripheral subu-
nits, such as flexible alkyl/alkoxy chains, on the conventional AIEgens of TPE and HPS deriva-
tives. Meanwhile, some studies have focused on the development of new AIEgen cores suitable
for manifestation of LC nature by using various synthetic approaches: introducing some inter-
molecular interactions such as hydrogen-­bonding, halogen-­bonding, metallophilic interac-
tions, and/or electrostatic interactions. The concepts have one common strategy: the separation
of the AIEgen core from the flexible units and the increase in interactions between AIEgen
cores. After the drastic maturing of the design strategy and the significant increase in the num-
ber of synthesized AIE-­active LCs, numerous applications possessing high-­performance com-
pared with that of the typical AIEgens are now possible, such as linearly and CPL materials.
Continued efforts toward the development of AIE-­active LCs can pave the way for the next
generation of applications.

­References

1 Tang, B. Z., Zhao, Z., Zhang, H., and Lam, J. W. (2020). Aggregation-­induced emission: new vistas at
the aggregate level. Angew. Chem. Int. Ed. 59 (25): 9888–9907.
2 Mei, J., Leung, N. L., Kwok, R. T., Lam, J. W., and Tang, B. Z. (2015). Aggregation-­induced emission:
together we shine, united we soar! Chem. Rev. 115 (21): 11718–11940.
3 Mei, J., Hong, Y., Lam, J. W., Qin, A., Tang, Y., and Tang, B. Z. (2014). Aggregation-­induced emission:
the whole is more brilliant than the parts. Adv. Mater. 26 (31): 5429–5479.
4 Luo, J. D., Xie, Z. L., Lam, J. W. Y., Cheng, L., Chen, H. Y., Qiu, C. F., Kwok, H. S., Zhan, X. W., Liu,
Y. Q., Zhu, D. B., and Tang, B. Z. (2001). Aggregation-­induced emission of 1-­methyl-­1,2,3,4,5-­
pentaphenylsilole. Chem. Commun. 18: 1740–1741.
5 Goodby, J. W., Mandle, R. J., Davis, E. J., Zhong, T., and Cowling, S. J. (2015). What makes a liquid
crystal? The effect of free volume on soft matter. Liq. Cry., 42 (5–6): 593–622.
6 Tschierske, C. (2007). Liquid crystal engineering–new complex mesophase structures and their
relations to polymer morphologies, nanoscale patterning and crystal engineering. Chem. Soc. Rev. 36
(12): 1930–1970.
7 Goodby, J. W., Collings, P. J., Kato, T., Tschierske, C., Gleeson, H. F., and Raynes, P. (2014).
Handbook of Liquid Crystals. Weinheim, Germany: Wiley-­VCH.
8 Chigrinov, V. G., Kozenkov, V. M., and Kwok, H.S. (2008). Photoalignment of Liquid Crystalline
Materials: Physics and Applications. New York: Wiley.
9 Ikeda, T., and Tsutsumi, O. (1995). Optical switching and image storage by means of azobenzene
liquid-­crystal films. Science 268 (5219): 1873–1875.
572 19  Liquid-­crystalline AIEgens: Materials and Applications

10 White, T. J., and Broer, D. J. (2015). Programmable and adaptive mechanics with liquid crystal
polymer networks and elastomers. Nat. Mater. 14 (11): 1087–1098.
11 Finkelmann, H., Kim, S. T., Munoz, A., Palffy-­Muhoray, P., and Taheri, B. (2001). Tunable
mirrorless lasing in cholesteric liquid crystalline elastomers. Adv. Mater. 13 (14): 1069–1072.
12 Xie, M., Hisano, K., Zhu, M., Toyoshi, T., Pan, M., Okada, S., Tsutsumi, O., Kawamura, S., and
Bowen, C. (2019). Flexible multifunctional sensors for wearable and robotic applications. Adv.
Mater. Technol. 4 (3): 1800626.
13 O’Neill, M., and Kelly, S. M. (2003). Liquid crystals for charge transport, luminescence, and
photonics. Adv. Mater. 15 (14): 1135–1146.
14 Castillo-­Vallés, M., Martínez-­Bueno, A., Giménez, R., Sierra, T., and Ros, M. B. (2019). Beyond
liquid crystals: new research trends for mesogenic molecules in liquids. J. Mater. Chem. C 7 (46):
14454–14470.
15 Fujisawa, K., Kawakami, N., Onishi, Y., Izumi, Y., Tamai, S., Sugimoto, N., and Tsutsumi, O. (2013).
Photoluminescent properties of liquid crystalline gold(I) isocyanide complexes with a rod-­like
molecular structure. J. Mater. Chem. C 1 (34): 5359–5366.
16 Wan, J. H., Mao, L. Y., Li, Y. B., Li, Z. F., Qiu, H. Y., Wang, C., and Lai, G. Q. (2010). Self-­assembly
of novel fluorescent silole derivatives into different supramolecular aggregates: fibre, liquid crystal
and monolayer. Soft Matter 6 (14): 3195–3201.
17 Guo, L. X., Xing, Y. B., Wang, M., Sun, Y., Zhang, X. Q., Lin, B. P., and Yang, H. (2019).
Luminescent liquid crystals bearing an aggregation-­induced emission active tetraphenylthiophene
fluorophore. J. Mater. Chem. C 7 (16): 4828–4837.
18 Yuan, W. Z., Yu, Z. Q., Lu, P., Deng, C., Lam, J. W., Wang, Z., Chen, E.-­Q., Ma, Y., and Tang,
B. Z. (2012). High efficiency luminescent liquid crystal: aggregation-­induced emission strategy and
biaxially oriented mesomorphic structure. J. Mater. Chem. 22 (8): 3323–3326.
19 Binnemans, K. (2009). Luminescence of metallomesogens in the liquid crystal state. J. Mater.
Chem. 19 (4): 448–453.
20 Bardají, M. (2014). Gold liquid crystals in the XXI century. Inorganics 2(3): 433–454.
21 Kishimura, A., Yamashita, T., and Aida, T. (2005). Phosphorescent organogels via “metallophilic”
interactions for reversible RGB–color switching. J. Am. Chem. Soc.127 (1): 179–183.
22 Sagara, Y., Mutai, T., Yoshikawa, I., and Araki, K. (2007). Material design for piezochromic
luminescence: hydrogen-­bond-­directed assemblies of a pyrene derivative. J. Am. Chem. Soc. 129
(6): 1520–1521.
23 Sagara, Y., and Kato, T. (2009). Mechanically induced luminescence changes in molecular
assemblies. Nat. Chem. 1 (8): 605–610.
24 Yang, H., Yi, T., Zhou, Z., Zhou, Y., Wu, J., Xu, M., Li, F., and Huang, C. (2007). Switchable
fluorescent organogels and mesomorphic superstructure based on naphthalene derivatives.
Langmuir 23 (15): 8224–8230.
25 Wan, J.-­H., Mao, L.-­Y., Li, Y.-­B., Li, Z.-­F., Qiu, H.-­Y., Wang, C., and Lai, G.-­Q. (2010). Self-­assembly
of novel fluorescent silole derivatives into different supramolecular aggregates: fibre, liquid crystal
and monolayer. Soft Matter 6 (14): 3195–3201.
26 Hirst, A. R., Escuder, B., Miravet, J. F., and Smith, D. K. (2008). High-­tech applications of self-­
assembling supramolecular nanostructured gel-­phase materials: from regenerative medicine to
electronic devices. Angew. Chem. Int. Ed. 47 (42): 8002–8018.
27 Babu, S. S., Praveen, V. K., and Ajayaghosh, A. (2014). Functional π-­gelators and their applications.
Chem. Rev. 114 (4): 1973–2129.
28 Bui, H. T., Kim, J., Kim, H. J., Cho, B. K., and Cho, S. (2016). Advantages of mobile liquid-­crystal
phase of AIE luminogens for effective solid-­state emission. J. Phys. Chem. C 120 (47): 26695–26702.
 ­Reference 573

29 Jing, H., Lu, L., Feng, Y., Zheng, J. F., Deng, L., Chen, E. Q., and Ren, X. K. (2016). Synthesis,
aggregation-­induced emission, and liquid crystalline structure of tetraphenylethylene–surfactant
complex via ionic self-­assembly. J. Phys. Chem. C 120 (48): 27577–27586.
30 Guo, H., Zheng, S., Chen, S., Han, C., and Yang, F. (2019). A first porphyrin liquid crystal with
strong fluorescence in both solution and aggregated states based on the AIE-­FRET effect. Soft Mat.
15 (41): 8329–8337.
31 Cored, J., Crespo, O., Serrano, J. L., Elduque, A., and Giménez, R. (2018). Decisive influence of the
metal in multifunctional gold, silver, and copper metallacycles: high quantum yield
phosphorescence, color switching, and liquid crystalline behavior. Inorg. Chem. 57 (20):
12632–12640.
32 Zhao, D., Fan, F., Cheng, J., Zhang, Y., Wong, K. S., Chigrinov, V. G., Kwok, H. S., Guo, L., and
Tang, B. Z. (2015). Light-­emitting liquid crystal displays based on an aggregation-­induced emission
luminogen. Adv. Opt. Mater. 3 (2): 199–202.
33 Chen, Y., Lin, J., Yuan, W., Yu, Z., Lam, J. W., and Tang, B. Z. (2013). 1-­((12-­Bromododecyl)
oxy)-­4-­((4-­(4-­pentylcyclohexyl) phenyl)ethynyl) benzene: liquid crystal with aggregation-­induced
emission characteristics. Sci. China Chem. 56 (9): 1191–1196.
34 Tong, J., Wang, Y. J., Wang, Z., Sun, J. Z., and Tang, B. Z. (2015). Crystallization-­induced emission
enhancement of a simple tolane-­based mesogenic luminogen. J. Phys. Chem. C 119 (38):
21875–21881.
35 Yamada, S., Yamaguchi, S., and Tsutsumi, O. (2017). Electron-­density distribution tuning for
enhanced thermal stability of luminescent gold complexes. J. Mater. Chem. C 5 (31): 7977–7984.
36 Hu, Y. B., Lam, J. W., and Tang, B. Z. (2019). Recent progress in AIE-­active polymers. Chinese
J. Polym. Sci. 37 (4): 289–301.
37 Hu, R., Leung, N. L., and Tang, B. Z. (2014). AIE macromolecules: syntheses, structures and
functionalities. Chem. Soc. Rev. 43 (13): 4494–4562.
38 Liu, L., Wang, M., Guo, L. X., Sun, Y., Zhang, X. Q., Lin, B. P., and Yang, H. (2018). Aggregation-­
induced emission luminogen-­functionalized liquid crystal elastomer soft actuators.
Macromolecules 51 (12): 4516–4524.
39 Koo, J., Lim, S. I., Lee, S. H., Kim, J. S., Yu, Y. T., Lee, C. R., Kim, D. Y., and Jeong, K. U. (2019).
Polarized light emission from uniaxially oriented and polymer-­stabilized AIE luminogen thin
films. Macromolecules 52 (4): 1739–1745.
40 Mochizuki, N., and Morita, R. (2018). Development of transparent emissive LCD using novel
polarized-­light-­emitting film. J. Soc. Inf. Display 26 (11): 670–674.
41 Zhao, D., Fan, F., Chigrinov, V. G., Kwok, H. S., and Tang, B. Z. (2015). Aggregate-­induced emission
in light-­emitting liquid crystal display technology. J. Soc. Inf. Display 23 (5): 218–222.
42 Sánchez-­Carnerero, E. M., Agarrabeitia, A. R., Moreno, F., Maroto, B. L., Muller, G., Ortiz, M. J.,
and de la Moya, S. (2015). Circularly polarized luminescence from simple organic molecules.
Chem. Euro. J. 21 (39): 13488–13500.
43 Roose, J., Tang, B. Z., and Wong, K. S. (2016). Circularly-­polarized luminescence (CPL) from chiral
AIE molecules and macrostructures. Small 12 (47): 6495–6512.
44 Han, J., Guo, S., Lu, H., Liu, S., Zhao, Q., and Huang, W. (2018). Recent progress on circularly
polarized luminescent materials for organic optoelectronic devices. Adv. Opt. Mater. 6 (17):
1800538.
45 San Jose, B. A., Matsushita, S., and Akagi, K. (2012). Lyotropic chiral nematic liquid crystalline
aliphatic conjugated polymers based on disubstituted polyacetylene derivatives that exhibit high
dissymmetry factors in circularly polarized luminescence. J. Am. Chem. Soc. 134 (48):
19795–19807.
574 19  Liquid-­crystalline AIEgens: Materials and Applications

46 Liu, J., Su, H., Meng, L., Zhao, Y., Deng, C., Ng, J. C., Lu, P., Faisal, M., Lam, J. W. Y., Huang, X.,
Wu, H., Wong, K. S., and Tang B. Z. (2012). What makes efficient circularly polarised luminescence
in the condensed phase: aggregation-­induced circular dichroism and light emission. Chem. Sci. 3
(9): 2737–2747.
47 Zhao, D., He, H., Gu, X., Guo, L., Wong, K. S., Lam, J. W., and Tang, B. Z. (2016). Circularly
polarized luminescence and a reflective photoluminescent chiral nematic liquid crystal display
based on an aggregation-­induced emission luminogen. Adv. Opt. Mater. 4(4): 534–539.
48 Li, X., Hu, W., Wang, Y., Quan, Y., and Cheng, Y. (2019). Strong CPL of achiral AIE-­active dyes
induced by supramolecular self-­assembly in chiral nematic liquid crystals (AIE-­N*-­LCs). Chem.
Commun. 55 (35): 5179–5182.
49 Song, F., Cheng, Y., Liu, Q., Qiu, Z., Lam, J. W., Lin, L., Yang, F., and Tang, B. Z. (2019). Tunable
circularly polarized luminescence from molecular assemblies of chiral AIEgens. Mater. Chem.
Front. 3 (9): 1768–1778.
50 Wu, Y., You, L. H., Yu, Z. Q., Wang, J. H., Meng, Z., Liu, Y., Li, X.-­S., Fu, K., Ren, X.-­K., and Tang,
B. Z. (2020). Rational design of circularly polarized luminescent aggregation-­induced emission
luminogens (AIEgens): promoting the dissymmetry factor and emission efficiency synchronously.
ACS Mat. Lett. 2 (5): 505–510.
575

20

Push–Pull AIEgens
Andrea Nitti and Dario Pasini
Department of Chemistry, University of Pavia and INSTM Research Unit, Pavia, Italy

20.1 ­Introduction

Organic conjugated chromophores emissive in solid state may represent attractive solutions for organic-­
based devices such as organic light-­emitting diodes (OLEDs), polymer light-­emitting diodes (PLEDs),
and solid-­state lasers. Organic push–pull chromophores have π-­systems that incorporate electron donor
(D) and electron acceptor (A) units. Their D–A structures allow easy tunability of the molecular back-
bones, so that it is possible to control the optoelectronic and emissive properties in solid state.
Intramolecular or intermolecular charge transfer (ICT) is a unique quantum effect that occurs in these
molecules, which leads to distinct optoelectronic properties. For example, the facile excitation of the
electrons within the molecular orbitals by visible light makes such molecules generally colored [1].
Typical push–pull luminogens suffer from the aggregation-­caused quenching (ACQ) phenome-
non, which prevent them from achieving the best emissive performance in the solid or aggregate
state [2]. In dilute solutions, these luminophores often emit strongly. However, upon aggregation,
their emission is suppressed, promoting decay to the ground state via nonradiative channels. For
example, 4-­(dicyanomethylene)-­2-­methyl-­6-­(4-­dimethylaminostyryl)-­4H-­pyran, which is widely
used as a standard red dopant in OLEDs, shows a weak emission at 665 nm with an emission quan-
tum yield (PLQY) of 0.05 in neat films [3].
Diametrically opposite to ACQ behavior is the emission enhancement after molecular aggrega-
tion given by aggregation-­induced emission (AIE). The concept of AIE was first proposed in 2001
by B. Z. Tang and coworkers on the observation done on 1-­methyl-­1,2,3,4,5-­pentaphenylsilole
(MPPS) [4] and widely explored in the past years as one of the best solutions to enhance the emis-
sive properties of solid-­state chromophores. These luminogens, named AIEgens, present the
intriguing characteristic to be weakly emissive in solution but strongly emissive in the aggregate
state. Typical AIEgens possess propeller-­like or rotor structures, which have spatial conformations
in the solid state able to avoid the formation of close π–π stacking and thus suppress the major
cause of nonradiative decay, favoring the emission processes [5].
By incorporating the molecular design principle of AIEgens and of push–pull chromophores, a
novel class of solid-­state-­emissive materials having benefit from both strategies can be realized.
Using the push–pull molecular design, a fine control of highest occupied molecular orbital/lowest
occupied molecular orbital (HOMO)/LUMO levels is possible in order to obtain specific

Handbook of Aggregation-Induced Emission: Volume 1 Tutorial Lectures and Mechanism Studies, First Edition.
Edited by Youhong Tang and Ben Zhong Tang.
© 2022 John Wiley & Sons Ltd. Published 2022 by John Wiley & Sons Ltd.
576 20  Push–Pull AIEgens

optoelectronic properties such us high molar absorption, wavelength emission, and activation of
delay fluorescence (DF). By incorporating AIE principle design in the push–pull structure, it is
possible to obtain materials with superior emissive properties, avoiding the problem of ACQ, with
high PLQY values in the solid state. Besides, the synergic incorporation of both push–pull and AIE
concepts to the design of novel materials has opened new exciting frontiers of research in recent
years, bringing to light new mechanisms for the enhanced emission of solid-­state materials, such
as aggregation-­induced delay fluorescence (AIDF) or excimer-­induced enhanced emission (EIEE)
to cite a few.
Push–pull AIEgens have started to find fertile ground in the field of manufacturing electronic
devices. Most OLED emitters exhibit strong fluorescence and phosphorescence in dilute solution but
are often affected by ACQ in the solid state, which greatly reduce their device performance. On the
contrary, push–pull AIEgens are capable of offering strong and bright emissions in films with PLQY
up to unity. Recently, a series of unsymmetrical AIDF emitters with a wide range of color have been
used as active layers for third-­generation non-­doped OLEDs and white-­light OLEDs (WOLEDs),
showing low turn-­on voltages and high luminescence [6, 7]. Apart from OLED technology, push–pull
AIEgens have found applications as acceptor materials for organic solar cells (organic photovoltaics
[OPVs]) [8]. Using push–pull AIEgens is possible to reduce the domain sizes and increases electron
mobility of acceptors in order to achieve better morphology of the active layers. Recently, push–pull
AIEgens have also been used as chromophores for luminescent solar concentrators (LSCs) [9].
Mechanochromism is a phenomenon where a material undergoes a noticeable and reversible
emission color change in response to a variety of mechanical stimuli such as stretching, grinding,
and pressing. Push–pull AIEgens are promising candidates for mechanoresponsive chromophores
due to their unique properties. For instance, pH-­responsive and vapor-­responsive push–pull
AIEgens have been developed [10], as well as white-­light-­emissive AIEgens with tricolor emission
switching properties at room temperature find application in anticounterfeiting applications [11].
Water solubility with red/near-­infrared (NIR) emission is particularly appealing in nanomedi-
cine devices, where low autofluorescence interference, deep penetration depth, and low phototox-
icity to heathy tissue are required. For example, red/NIR AIEgens have been applied for in vivo
bioorthogonal fluorescence turn-­on tumor labeling  [12]. Most AIEgens with absorption in the
short-­wavelength spectrum are used in multiphoton imaging to reduce the photon scattering.
Therefore, multiphoton imaging using AIEgens has been demonstrated to offer through-­skull
imaging of mouse brain vasculature and mapping lipid in mouse atherosclerotic plaques [13].
In order to gradually drive the reader more and more in-­depth, we have divided this chapter into
three paragraphs. In the first paragraph, we introduce the basic concepts of solid-­state exciton
theory, fundamental concepts to understand the solid-­state emission world, followed by an over-
view on the basic concepts of AIE molecular engineering and push–pull design. In the second and
third paragraphs, several examples that make use of such design principles will be illustrated.

20.2  ­Basic Concept of Molecular Design

20.2.1  Photophysical Excited States in Aggregates


Supramolecular aggregates are scaffolds in which molecules are weakly bonded by van der Waals
forces. Due to their weak bonding, the properties of individual molecules are retained in a solid.
Therefore, to understand the emissive properties of aggregate, it is important to understand the
properties of the molecules themselves. The photophysical destiny of excited molecules is effi-
ciently summarized by a Jablonski diagram like that shown in Figure  20.1. Excited molecules,
H-aggregate J-aggregate HJ-aggregate
ISC

IC
S1 J0 > 0 J0 < 0 J0 < 0
ΔEST
Single molecules

T1 E(k) E(k) E(k)


Solid-state

RISC
F NRD DF
P

S0

[M*··M] [M··M*]
D* D
LUMO
Förster D D*
RD D
mechanism
(FRET)
HOMO RD D D*
M* M M M*
D* D
[M*··M] [M··M*] RD
D D* D
LUMO RD
Dexter NRD
mechanism RD D D* Energy
(DET) NRD migration
HOMO
RD
M* M M M*
NRD

Figure 20.1  From left to right: on the top are schematized the Jablonski diagram for a single molecule, and the energy level diagram for H-­, J-­, and
HJ-­aggregate arrays, while on the bottom are schematized the mechanisms of energy transfer according to Forster and Dexter models and the energy
migration process (RD, radiative decay; NRD, nonradiative decay).
578 20  Push–Pull AIEgens

generated following the absorption of light, can undergo several possible relaxation pathways from
the singlet or triplet state. One possibility to come back to the ground state is emitting light by fluo-
rescence (F) process. In addition to the radiative transition, the molecules may lose energy in non-
radiative pathway, either by internal conversion (IC) or by adiabatic intersystem crossing (ISC)
process to a triple state (Tn). Although S1 → Tn transitions are not theoretically allowed by spin
selection rule, ISC may happen if strong spin–orbit coupling occurs between S1 and Tn states. In
this case, with a population of the T1 state, the phosphorescence (P) process may occur. In addi-
tion, a radiative transition with longer lifetime than fluorescence decay is observed from S1 state
when the excited molecules at room temperature undergo a reverse intersystem crossing (RISC)
from T1 to S1, which is known as thermally activated delayed fluorescence (TADF). The fundamen-
tal photophysical parameters that govern these phenomena, such as emission wavelength (λF, λP,
or λDF), the spin–orbit coupling, and ΔEST values are closely dependent from molecular design, so
that emission can be dramatically changed by making small structural changes.
In the solid state, molecules can form aggregates whose structure affects emissive properties
such as emission wavelength (λem), lifetime (τ), and photoluminescence quantum yield (PLQY).
Comprehensive discussion of exciton dynamics in aggregate materials can be found in Refs. [14,
15]. The exciton model, proposed by Kasha [16], describes the excited state of aggregates in terms
of resonance between weakly interacting excited molecular states. The exciton model consists of
(i) definition of exciton as excited local state and (ii) the energy migration into aggregates through
energy transfer. According to Kasha’s model, a system of N molecules positioned close to one
another by weak interactions can assume three types of 3D-­arrays, as shown in Figure 20.1: (i) the
H-­aggregate, which involves parallel alignment of molecules; (ii) the J-­aggregate, in which mole-
cules are aligned in a head-­to-­tail manner; and (iii) HJ-­aggregate, which is a mix of the aforemen-
tioned alignment.
The simplest approach to describe optical excitation of aggregate organic materials is to consider
a linear array of N identical molecules i (i = 1, 2, 3, . . ., N) coupled via electrostatic interaction. The
resolution of the corresponding Hamiltonian equation gives rise to N spitted excited state with
energies: E(k) = Ei + K + J0 cos (ka), where Ei is the excitation energy of noninteracting molecules,
K is the repulsion energy contribution, J0 is the energy of interaction between neighboring mole-
cules, a is the distance between nearest neighbors, and k = 0, ±2π/Na, . . ., ±π/a. The description
of the energy diagram for H-­, J-­ and HJ-­aggregates coming from E(k) equation is illustrated in
Figure 20.1.
According to the selection rule for the excitation of the aggregate array, only the excitation
energy corresponding to k  =  0 can be populated and the transitions to the other levels are not
allowed. This implies that in J-­and HJ-­aggregates, the lower-­level exciton is allowed, and fluores-
cence is red-­shifted compared to that observed in a single molecule. On the other hand, fluores-
cence in the H-­aggregate is blue-­shifted compared to that of the monomer because the higher-­energy
level is allowed.
Two processes of energy transfer may occur depending on the degree of interaction between
the molecules: Förster resonant energy transfer (FRET) and Dexter energy transfer (DET). The
FRET mechanism shown in the blue box of Figure 20.1 is based on a dipole–dipole approxima-
tion and consists of a simultaneous relaxation of an electron located in the LUMO of a molecule
to the HOMO and excitation of an electron located in the HOMO of its neighbor giving rise to a
pair of a ground-­state and excited-­state molecular dimers. It is a long-­range process, occurring in
a range of 10–100  Å, typically relevant for the singlet-­exciton transport. At shorter distance
between molecules (<10  Å), the dipole approximation breaks down, and the intermolecular
orbital overlap is large enough to affect electronic communication, so that an energy transfer
20.2  ­Basic Concept of Molecular Desig 579

through electron exchange (DET) becomes possible. During the energy transfers, the population
of excitons does not change through the migration. Hence, if the photophysical process of exci-
tons is limited only to fluorescence and energy migration, the fluorescence efficiency should be
unitary. Contrary, if nonradiative pathways are operative in addition to fluorescence, energy
migration amplifies the opportunity for energy loss. Excitons could be converted in two triplet
excitons through a singlet fission (SF) process. The SF mechanism provides the internal conver-
sion of pairs S0 and S1 into a pair of coherently coupled triplet states, named multiexcitons. The
SF process, S0 + S1 → 1(TT) → T1 + T1, is spin-­allowed and can occur on a subpicosecond time-
scale making it competitive with other processes. If the SF process is active, the exciton can
decay through phosphorescence or delayed fluorescence.

20.2.2  Fundamental Molecular Design to Achieve Push–Pull AIEgens


Regarding molecular design to build efficient AIEgens, one must always keep mind the following
sentence: Organic photoluminescent materials can be obtained through a precise design of molec-
ular structures and the control of solid-­state intermolecular interactions. The rational design of
push–pull AIEgens requires the fusion of the basic design concepts of AIE-­active luminophores
and push–pull structures. The following discussion aims to illustrate these basic concepts.
Most popular chromophores suffer from the ACQ effect caused by a strong intermolecular π–π
stacking interaction occurring between neighbors in the condensed state. On the contrary, AIEgens
are molecules that do not emit light in solution but that do it in aggregate state. Nowadays, we
know that the basic principle of AIE phenomenon is the restriction of intermolecular motion
(RIM) of the twisted or stator structures in condensed phases [17]. In solution, AIEgens exist in a
free state without restriction imposed on the intramolecular motions. Therefore, these molecules
are not luminescent due to depopulation of the excited state mainly by nonradiative decay.
However, in condensed phases (aggregate, film, or powder), the packing geometry freezes the
molecular motion (rotational and vibrational motions) in a nonplanar geometry through weak
intermolecular interactions. As a consequence, the nonradiative decay is deactivated favoring the
population of radiative relaxation pathway. The fundamental molecular feature to promote the
radiative decay is the conservation of twisted geometry in a condensed phase in order to generate
an interplanar molecular distance large enough to ideally suppress the π–π stacking interaction.
Originally, the strong solid-­state-­emissive behaviors were correlated with the capability to form
J-­aggregates, with notoriously reduced π–π stacking interaction [17–19]. However, the strong emis-
sion observed in molecules with H-­aggregate features has made to doubt such a correlation [20,
21]. Generally, on the basis of the literature analysis, we can enunciate that tight supramolecular
packing through weak intermolecular interactions to enhance the molecular rigidity is a crucial
factor in highly luminescent solids based on twisted AIEgens; likewise, molecular rigidity is ben-
eficial to the high luminescence efficiency of a single molecule. This concept matches well with
the novel experimental evidence of excimer-­induced enhance emission (EIEE) due to pairwise
anthracene or perylene stacking in crystal [22, 23]. Emission from excimer states is usually avoided
because the excimer species drastically quench the emission. However, stabilizing the excimer
interaction through spatial isolation of dimer units in the crystal structure and avoiding the forma-
tion of lower-­energy excimer clusters give rise to a single emissive species with high PLQY and
long lifetime in solid state.
Organic π-­systems end-­capped with an electron donor (D) and an electron acceptor (A) are
known as push–pull systems (D–π–A). The intriguing properties of these molecules are given by
their ability to favor the D–A interaction, or the intramolecular charge-­transfer (ICT) processes.
580 20  Push–Pull AIEgens

The ICT generally occurs in the photoexcited state of molecules following the absorption of light.
The photoexcitation makes possible the electron transfer from side to side of a molecule, generat-
ing an asymmetric charge distribution in the excited state markedly different from that of the
ground state.
As shown in Figure 20.2a, in molecules with D and A groups linked by a π-­bridge, the ICT occurs
by through-­bond (TBICT), while in molecules where the crosstalking is denied but D and A have
a favorable position for charge transfer, an intramolecular through-­space charge transfer (TSICT)
is allowed. It is known that excited molecules with ICT effect may give rise to dual emission
(Figure 20.2b): a blue-­shifted emission arising from a locally excited state (LE) and a red-­shifted
emission generated by ICT excited state [25]. Generally, the ICT state has lower energy than the LE
state, so considering that the LE state corresponds to the LUMO level in a HOMO–LUMO excita-
tion, the ICT state is virtually a lower LUMO level than the starting one. One of the important
effects of ICT is to lower the LUMO level in push–pull molecules and therefore they are generally
colored. The ICT is also responsible for the strong polarization of the push–pull luminophores and
generation of a great molecular dipole in the excited state that affects their Stokes shift values in
the emission spectra. The ICT emission generally occurs in polar solvents as a consequence of
solvent stabilization of the polar excited state. On the contrary, apolar solvents stabilize the LE
emission due to its nonpolar nature. In molecules in which the activation energy between the LE
state and the ICT state is low at room temperature, the charge-­transfer process is favored and well
characterized by the solvatochromic effect.
Many experimental and theoretical studies have been devoted to investigating the charge-­
transfer mechanism in through-­bond D–A systems, and they support the twisted intramolecular
charge-­transfer state (TICT). In the framework of the TICT model, following the excitation toward
the LE state, the ICT state is accessible only by the rotational motion around the bond that con-
nects the D and A units. If there is no energy barrier between LE and ICT states, the excited relaxa-
tion can occur from the ICT state only. In such systems, where through-­bond crosstalking is denied
and charge transfer occurs only spatially, be it intramolecular or intermolecular, the charge trans-
fer follows a DET mechanism.

(b)
(a)
S1
Through-bond CT IC kNR
IC
TICT

Through-space CT kR1 kR1 k


hν hν R2

(c)
S0 S0

r0 r1 r0 r1 rnct
Conformation Conformation
coordinate coordinate

Figure 20.2  (a) Schematic representation of the ICT process in a model system and type of intramolecular
(b) and intermolecular (c) charge transfer. Source: Partially reproduced with permission from Ref. [24]
published by the PCCP Owner Societies.
20.3 ­Push–Pull AIEgens from Rotor Structur 581

(a) (b) (c)


D1 D2 D3 A1 A2 A3 π1 π2 π3

L
U
M
O

H
O
M
O

Length
(d)

Composition

Conjugation/
Conformation

Figure 20.3  Effect of (a) HOMO engineering, (b) LUMO engineering, and (c) π unit on the HOMO and
LUMO levels, as well as on the Eg values. (d) Possibilities for tuning and matching the A and D units.

ICT is not only a feature that makes a push–pull luminophore intriguing. Another aspect that
have conferred them great success is a structure modularity thanks to which a fine-­tuning of the
optoelectronic properties is possible  [26]. In fact, according to the quantum theory mixing, the
molecular orbital of conjugated units D, A, and π yield to new hybridized molecular orbitals strictly
related to the molecular orbitals of the single units. Figure 20.3 shows the effect of HOMO/LUMO
engineering and π system on the Eg values. For instance, increasing the donating ability of D
(D1  <  D2  <  D3) gradually pushes up the HOMO level (Figure  20.3a), while an increase of the
electron-­withdrawing ability of A (A1 < A2 < A3) affects the depth of LUMO levels (Figure 20.3b).
Because the π-­conjugated bridge is an integral part of the D–π–A system, it participates in the con-
struction of the hybrid molecular orbital and influence the optoelectronic properties in equal man-
ner of D and A (Figure  20.3c, d). The π-­bridge consists of a combination of multiple bonds,
aromatic, and heteroaromatic rings positioned on the crosstalking direction or periphery. The shift
of the HOMO–LUMO orbitals in the function of π unit (π1, π2, and π3) is not immediate, affecting
the shift of both HOMO and LUMO levels, and depends on their length, composition, conjugation,
and spatial arrangement. Acting on the HOMO–LUMO magnitudes through the D, π, and A
nature, we can tune the molecular Eg and then modify the fundamental parameters that character-
ize a luminophore such as λabs and absorptivity, as well as emissive features such as λem, Stokes
shift, and the emission nature (F, DF, or P).

20.3 ­Push–Pull AIEgens from Rotor Structure

Rotor structures or propeller-­like structures represent the most typical twisted π structures for
AIEgens. The typical rotor structure is based on: (i) a stator, the stationary part of the rotating
molecular system, which are typically double bond, nitrogen atoms and benzene or heterocycles;
582 20  Push–Pull AIEgens

and (ii) a rotor, a moving part of the rotating molecular system, which are typically benzene rings
or derivatives.
We report the most popular examples of AIEgens with rotor structures. All examples are subdi-
vided according to their stators in: (i) double bond, (ii) point-­restricted rotors from atoms or func-
tional groups, and (iii) aromatic rings.

20.3.1  Double Bond Stator


Tetraphenylethene (TPE) is the most studied AIE chromophore with the double bond as a stator.
The central olefin stator of this molecule is surrounded by four peripheral benzene rotors. Its high
solid-­state PLQY (40%) and its relatively simple synthesis have prompted its inclusion in a wide
variety of complex architectures [27]. The cyano group is a typical acceptor with a simple structure
widely used to design and create many AIEgens. For example, Park and coworkers first reported a
cyano-­containing AIE compound of 1-­cyano-­trans-­1,2-­bis-­(40-­methylbiphenyl)ethylene in
2002 [18]. Although the cyano group is small in size, its steric effect still leads to a twisted confor-
mation due to its rigid structure. In addition to its steric effect, it also possesses strong electron-­
withdrawing properties, which, combined with donors, can generate simple D–A systems.
Examples of these classes of push–pull AIEgens are shown in Figure 20.4.
Decorating TPE with ylidenemalononitrile acceptors and methoxy donors generates compounds
1–4 with diverse emission colors  [28]. These compounds are weakly emissive in good solvents
(PLQY around 0.1%) but display strong fluorescence in the aggregate and solid states, displaying
typical AIE activity. Furthermore, when the number of methoxy units increases, the D–A interac-
tion becomes stronger with a consequently gradual red-­shifting of the emission wavelength. The
aggregate dispersions emit strongly in the yellow-­green, orange, and red lights with PLQY values
of 54, 61, 52, and 48%, respectively. Besides, the position of D/A also plays an important role in
determining the emission wavelength. By changing the ylidenemalononitrile unit from the para-­
to meta-­position in compound 5, the emission (595 nm) is blue-­shifted of 33 nm in comparison
with that of 3, due to its weak D–A intramolecular interaction. The introduction of steric effects in
ylidenemalononitrile groups has an important impact on the Eg and consequently on the emission
wavelength. From 6 to 8, owing to the increasing steric effects from the substituents of the vinyl
group compared with 3, the torsional angle between the TPE moiety and dicyano groups is
increased, resulting in a reduced π conjugation and shortened π length. The reduced conjugation
of the acceptor increases the LUMO energy level, leading to a hypochromic shift of the emission
maximum [29, 30]. However, in compounds 8 and 9, where the substituents are a benzene and
thiophene ring, respectively, the emission maxima recover to 632 and 636 nm. Although the tor-
sional angle between MTPE and the ylidenemalononitrile unit is still large for 8 and 9, the intro-
duction of conjugated rings forms more π-­extended acceptor systems, leading to similar optical
properties as that of 3. When thiophene instead of benzene is located between TPE and the dicy-
anovinyl moiety, 11 displays much red-­shifted emission maximum than 10  [31]. In addition,
extending the acceptor conjugation, such as 12 and 13, is possible to obtain AIE-­active chromo-
phores with emission wavelengths of 700 and 750  nm, respectively, for bioimaging and cancer
cell-­tracking [32].
Push–pull AIEgens 14 and 15, having ylidenemalononitriles as acceptor units and N,N-­
dimethylaniline units as donors in tri-­and tetrasubstituted ethene molecules, have been prepared
by our group [33]. 14 and 15 show solvatochromic behavior related to their strong D–A interac-
tions, indicating that the excited state is more polar than the ground state. A low PLQY of 0.1% is
recorded in solution, but by freezing of solution or adding a nonsolvent to the solution, an increase
20.3 ­Push–Pull AIEgens from Rotor Structur 583

MeO OMe

TPE MTPE

CN CN
Y X Y X
MTPE MTPE
CN CN
t-Bu
A 6 7

Z A CN CN
MTPE MTPE
CN CN
1 X=Y=Z=H 5 X = Y = OMe
2 X = Z = H, Y = OMe CN S
3 Z = H, X = Y = OMe A=
4 X = Y = Z = OMe CN
8 9

NC CN NC CN
NC CN NC CN
NC O
S
MTPE MTPE
MTPE MTPE
10 11 12 13

NC CN NC CN F3COC COCF3 R′OOC COOR′

R
N N N N N
R
17 R = Me R′ = Me
14 15 16 18 R = Me R′ = Et
19 R = Et R′ = Me

Figure 20.4  AIEgens structure with double bond stator.

of PL intensity is observed. The PLQYs of their powder are 3 and 11%, respectively. Such evidence
confirms the AIE activity by the RIR effect of these molecules. The crystal structure of 14 presents
a pseudolayered structure without the presence of π–π interaction due to long interplanar distance
between neighboring units. The absence of strong π–π stacking relieves the charge transfer by DET
mechanism and then the exciton annihilation. At the same time, the presence of too weak inter-
molecular interaction between neighbors (such as C–H⋯N and C–H⋯π interactions) is not
enough to suppress the nonradiative decay affecting the PLQY value. Contrary to 14, the cruciform
conformation of 15 with the out-­of-­molecular plane of one phenyl ring and one CN group is the
cause of mitigation of π–π interaction between neighbors. The stronger intermolecular interaction
584 20  Push–Pull AIEgens

acting in 15 leads to a tighter packing compared to that in 14, resulting in a higher value of
PLQY. Interestingly, 15 shows four kinds of crystals [34]. The crystals are characterized by different
morphologies, absorption, and emission colors. In particular, orange-­emissive crystals from pris-
matic shape (A), yellow emissive from rod-­like crystals, (B) and green emissive from needles (C)
and plate (D). A distinctive feature of the molecular structures of A–D is their twisted conforma-
tion due to sterical hindrance both between the CN and the dimethylamino-­phenyl substituents,
as well as between the phenyl rings. The authors have found three geometrical factors that play in
a concerted way to reduce such a hindrance: (i) the (N)C–C=C–C(Ph) torsion angle; (ii) the recip-
rocal tilting of the phenyl rings; and (iii) the significantly elongated central double bond denoting
a high conjugation degree in 15. Small differences of these parameters affect the degree of conjuga-
tion of 15 into the polymorphic crystals resulting in their different fluorescence behaviors.
Substituting a cyano acceptor unit with the trifluoroacetyl group, Pasini et al. prepared 16 exhibit-
ing both TICT and AIE activities [24]. The PLQY are measured in solution revealing very low val-
ues for all nonviscous solvents (around 0.02%), while higher values are measured for the viscous
ones (7% for the PEG solution) and in the solid state (11% for the powders). Combining ultrafast
pump–probe experiment on viscous and nonviscous systems of 16, as well as DFT calculations, the
TICT mechanism is an effective elucidate. Compounds 17–19 having an archetypal structure of 14
with ester acceptor in the substitution of a cyano group are prepared [35].These compounds with
nonemissive molecules are intriguing, which are induced for emitting by crystallization but not by
amorphization, a behavior that has been defined as crystallization-­induced emission (CIE). CIE is
particularly relevant since crystalline films usually exhibit a higher charge carrier mobility than
their amorphous counterpart. In solutions 17–19 are weak emissive (PLQY < 0.1%), while their
crystalline powders emit strongly at 468, 473, and 522 nm with a PLQY of 38, 38, and 1%, respec-
tively. The fluorescence drastically decreases in amorphous films: for example, 17 shows only a
weak blue-­greenish emission at room temperature with a PLQY value of 1%. The very low emission
of 17 is related to exciton nonradiative relaxation processes favored by molecular torsional mobility
acting in the amorphous and solution phases, which opposites are not observed in the crystal state
where the tight packaging of molecules reduces the exciton annihilations. The much higher emis-
sion intensity displayed by 17 and 18 with respect to 19 is explained on the basis of differences in
their crystalline state. 17 and 18 share a similar packing of the aromatic rings, while 18 shows a
markedly different torsional angle between (Ph)C–C(vinyl) bonds, which affects the interplanar
distance between adjacent arenes that result in much longer in 19 than those of 17 and 18. This
difference confers to 17 and 18 a J-­aggregate responsible for the very high PLQY, while 19 shows
H-­aggregate responsible for its poor emission. By means of an ultrafast pump−probe spectroscopy
combined with DFT calculations, the authors show direct evidence of intramolecular rotation in
19 [36]. The spectral evolution of the stimulated emission band of the chromophore in the first
45 ps after photoexcitation is fully consistent with the presence of a torsional relaxation toward the
equilibrium geometry of the excited state, taking place on timescales that depend on the solvent
viscosity. The structural features of the excited state fully account for the different photolumines-
cence efficiencies observed in solvents with different viscosities.

20.3.2  Point-­restricted Rotors from Atoms or Functional Groups


Because of their electron-­donating properties and propeller molecular structures, triphenylamine
(TPA) and its derivatives have been used for creating functional AIE molecules [37]. In virtue of
their electron-­donating nature, combining TPA or its derivatives with electron-­withdrawing groups
can produce D–A interaction, resulting in long-­wavelength emission (see Figure 20.5). Besides, as
20.3 ­Push–Pull AIEgens from Rotor Structur 585

OMe

S
N N N
N N
MeO
TPA MTPA DFA Cbz TPZ

CN TPA TPA
CHO CN
TPA
CN

N N
CN
Y X Y X X
20 X = Y = Cbz
23 X = Y = Cbz 25 X = H
21 X = DPA, Y = H 27
24 X = DPA Y = H 26 X = NO2
22 X = Y = DPA X

R Y Z
X
R
CN CN
Y
N
Z N
N
NC N Z
28 R = Cbz
R Y
29 R = PTZ NC
30 R = DPA
R X
Z Y

CN 31 X = H Y=H Z=H
N 32 X = F Y=H Z=H
N 33 X = CF3 Y=H Z=H
NC 34 X = H Y = CF3 Z=H
35 X = H Y = CF3 Z = tBu
36

Cbz DPA N TPE


CN CN CN

CN CN TPE CN
Cbz DPA N
37 38 39

Figure 20.5  Structure of twisted TPA-­based AIEgens.


586 20  Push–Pull AIEgens

mentioned above, the propeller star-­burst structures of TPAs are in favor of the construction of
AIE materials. The AIEgens incorporating TPA and derivatives are shown in Figure 20.5.
Incorporating diphenylamine (DPA) or carbazole (Cbz) as sterically-­hindered donor groups and
aldehydes or ylidenemalononitriles as acceptor units with a TPA core, the series of twisted TPA-­
based AIEgens 20–24 have been prepared [38].
Although their molecular structures are similar, their emission behaviors in the solid-­state are
different. In particular, 21–24 are AIE-­active showing a weak fluorescence in solution (PLQY value
of 0.2%), but strong emission when a large amount of nonsolvent is added and aggregates are
formed (PLQY of 31, 12, 10, and 15%, respectively). On the opposite, 20 is AIEE-­active showing
moderate fluorescence with a PLQY of 9.7% in solution and an enhanced emission in aggregates
with a PLQY value of 75%. Despite different emission behaviors, these luminophores can effec-
tively prohibit the ACQ effect and demonstrate a high PLQY in the solid state. The big difference
between the PLQY of 20 and 22 in the organic solvent may be a result of their difference in rota-
tional freedoms, which helps to reduce the nonradiative decay processes and then causes emission
in solution. 20–24 have a “pyramid-­mimicked” configuration imposed by the TPA core, which can
prevent the π–π interaction in the solid state. The PLQY of 22 is higher than that of 21 in an aggre-
gated state because the presence of steric hindrance on both sides of TPA core decreases the π–π
stacking interactions. The emission colors of these compounds can be changed from blue to deep
red acting on the D/A units and then on their D–A interactions. Upon increasing the acceptor abil-
ity of the electron-­withdrawing units and the donor ability of the electron-­donating units, the
emission spectra of these compounds are gradually red-­shifted from blue to the red region on the
order of 20 (490 nm), 21 (560 nm), 22 (640 nm), and 23 (725 nm).
R. Lu et al. prepared D–A compounds 25–27 by merging TPA moieties with phenylacrylonitrile
group [39]. These compounds are weakly emissive in solutions with PLQY values of 2, 1, and 3%,
respectively. However, the PLQY values for their solids are up to 33, 9, and 67%, respectively. The
emission colors of these AIEgens can be tuned acting on their ICT band by changing the conjuga-
tion length and the acceptor ability of phenylacrylonitrile. 25 emits a green fluorescence (506 nm)
in the solid state, while derivative 26 decorated with a nitro group shows a strong red fluorescence
(624 nm) in the solid state. Decorating 25 with two TPA units, linked in the para-­position using an
ethene bridge, the conjugation is extended and the resultant compound 27 emits an orange light
(596 nm) in the solid state.
By linking D units DPA, Cbz, or phenothiazine (PTZ) groups with electron-­accepting dicy-
anobenzene moieties via TPA bridges, a number of multibranched TPA-­based derivatives 28–30
were prepared [40]. All these compounds exhibit strong AIE behavior. When nonsolvent reaches
90%, the fluorescence intensities are increased by 50-­, 70-­and 230-­fold than their solutions. Upon
increasing the electron-­donating ability of the substituents, the suspensions of 28–30 in aggregates
show deep yellow, orange, and red fluorescence, respectively. Owing to the strong D–A effect, these
compounds also demonstrate a 2PA performance with 2PA cross sections as large as 814, 1484, and
1016 GM, respectively.
H. Gao et al. synthesized D–A AIEgens 31–35, which emit an intensive red fluorescence in their
solid states [41]. In these compounds, the DPAs are used as D groups, while phenylacrylonitrile
units play as A group. In addition to their inductive effect, these two groups also maintain the
twisted structures of these compounds, resulting in their AIE properties. 31–35 are nonemissive in
solution but highly emissive when aggregates are formed. In their solid states, 31–35 display a
strong emission peaked at 602, 644, 667, 676, and 686 nm, with PLQY values of 87, 99, 47, 49, and
43%, respectively. The problematic ACQ phenomenon is more serious for red fluorophores having
the elongated conjugation with large aromatic rings. It is so remarkable that 34 and 35 in the solid
20.3 ­Push–Pull AIEgens from Rotor Structur 587

state display strong red/NIR emissions. 32 exhibits higher PLQY value, owner of a dihedral angle
between the central benzene plane and the phenyl group of 35° and an optimal distance between
neighbors of 6.8 Å, which gives it a superior tight intermolecular packing as demonstrated by the
higher radiative constant rate value (kr) and a much lower value of nonradiative constant rate (knr).
By altering the substituents in the phenyl and diarylamine units, the emission colors of the com-
pounds 31–35 are changed from orange to red/NIR. For example, without substituents in the phe-
nyl and diarylamine units, 31 shows an orange fluorescence. However, when –CF3 units are
introduced into the phenyl rings at the meta-­position, and –tBu groups are linked to the diar-
ylamine units, the resulting compound emits NIR light.
T. Chen et al. prepared red-­emitting AIEgens 36 decorating biphenylfumaronitrile acceptor core
with arylamino units [42]. 36 is slightly emissive when it is dissolved in common organic solvents
due to the motion of the nonplanar phenyl and naphthalenyl rings and the antiparallel dipolar
arrangement of the biphenylfumaronitrile groups, but it shows a strong orange-­red fluorescence in
the solid state due to the RIM process. As a consequence of the strong ICT interaction, 36 shows a
pronounced red-­shifted emission centered at 616 nm. Replacing N-­naphthyl-­N-­phenylamine with
Cbz, DPA, and DPE with TPE in the substitution of a phenyl ring, compounds 37–39 are prepared.
D. C. Neckers et al. reported compound 37 with AIEE-­activity [43]. In THF, the PLQY value for 37
is 2%, while its solids possess a high PLQY value of 72% with an emission maximum at 608 nm.
Interestingly, 37 shows both positive and negative solvatochromisms with emission peaks being
shifted from 580 to 608 nm and to 520 nm by altering the solvent. Also, 37 exhibits an AIEE behav-
ior, while 39 is AIE-­active. In THF, 38 shows weak fluorescence peaked at 652 nm with a PLQY of
2.32%, while 39 exhibits a red-­shifted emission at 660 nm with PLQY values of 0.59%. Their solid
emissions are enhanced 18-­fold (PLQY of 42.5%) and 89-­fold (PLQY of 52.5%).

20.3.3  Aromatic Rotors


Figure 20.6 shows the structure of the examples reported in discussion. In 2009, B. Z. Tang and
coworkers reported the AIE activity of 5,6-­diphenylpyrazine-­2,3-­dicarbonitrile (DCDPP, 40)
caused by the RIM effect [44]. When molecularly dissolved in good solvents, 40 is nonemissive
with a PLQY of <0.1% but its fluorescence is 25-­fold intensified after the addition of nonsolvents
due to the formation of aggregates. During the formation of aggregate, 40 reveals a red-­shifting of
emission wavelength when “small” amounts of nonsolvents are added into its solution but shifts
toward blue when the percentage of nonsolvent exceeds those of the good solvents.
An emission wavelength of 415 nm is observed, which is in agreement with the emission wave-
length recorded in the solid state. This trend and the high dipole momentum determined by calcu-
lations, as well as the shifting of emission wavelength with changing of the solvent polarity, suggest
that the solvatochromic effect in 40 is caused by ICT. Introducing donor units with different donat-
ing abilities in the para-­position of phenyl rings is a successful strategy to increase the ICT effect.
The HOMO engineering actuate modifying the donating ability of D units, affect the Eg value of 41
and 44, and consequently shift the emission wavelength toward the visible region of spectrum. 41
and 42 show typical AIE activity: they are nonemissive in solution, but strongly emissive in the
aggregate state [45]. In the powder state, 41 and 42 show intense green and yellow emissions (emis-
sion maxima at 577 and 527 nm, respectively) with a PLQY of 20 and 12.8%, respectively. Both
compounds also reveal a reversible mechanochromic effect after grinding. The authors ascribe
mechanochromism to the conformational planarization and potentially excimer formation upon
grinding, which is associated with the conversion from crystalline (twisted) to amorphous (planar-
ized) states.
588 20  Push–Pull AIEgens

NC CN

N N
40 R = H R= N N O
41 R = DPA
42 R = Cbz

R R 43 44

NC CN NC CN NC CN

N N N N N N

TPA TPA TPA TPA TPA TPA


45 46 47

tBu
N

N N R= N S N N
R
O
tBu
48 49 50

Figure 20.6  Structure of AIEgens with aromatic stators.

In a D–A system, the CT state can be stabilized when the π system adopts a twisted geometry,
thereby lowering the excited-­state energy level and so that lowering the LUMO level and Eg value.
The stabilization of the TICT in the solid state also affects ΔEST. In fact, the twisted conformation
of the excited state in the D–A structure leads to a spatial separation of HOMO and LUMO orbitals,
confined on the D moiety and A moiety, respectively, decreasing the overlap-­integral J0 and conse-
quently the ΔEST values. As a consequence of these features, RISC is activated causing an AIDF-­
emissive behavior. For instance, the para-­substitution of phenyl rings with more donating
9,9-­dimethylacridan (Ac) and phenoxazine (Px) leads to compounds 43 and 44 with AIDF behav-
iors [46]. The strong steric hindrance between the bulky donating units Ac and Px with the periph-
eral phenyl rings of DCDPP acceptor causes dihedral angle values of 78–90°, and 43 and 44 assume
a cruciform conformation. The cruciform conformation imposes a spatial separation of the HOMO
orbital (localized on the donor units) from the LUMO orbital (localized on the DCDPP unit), which
lead to ΔEST values of 0.09 and 0.04, respectively. In the solid state, 43 and 44 exhibit an enhanced
delay fluorescence peaked at 570 and 610 nm with a PLQY of 67 and 15%, respectively. The differ-
ence in the emission wavelength is attributed to the stronger donating ability of Px compared with
that of Ac, which increase the energy of the HOMO level in 44 with a consequent reduction of its
Eg value. Different from HOMO engineering observed in 41–44, LUMO engineering reduces the Eg
acting on the LUMO levels of the acceptor DCDPP unit. By introduction of the TPA in the DCDPP
core, the enhanced molecular extension led to red-­shifted emission from 616 nm for 41 to 632 nm
for 45  [47]. From 45 to 46, by introducing the fused ring structure of dibenzoquinoxaline-­2,3-­
dicarbonitrile (DCDPNP), the ICT state of 46 is enhanced as compared to that of 45, leading to a
20.4 ­Push–Pull AIEgens from ACQ Chromophore 589

significant red shift of 75 nm in emission (λem = 706 nm) [48]. This indicates that DCDPNP is a
good acceptor because of its large and rigid π-­conjugated structure with a strong electron-­
withdrawing capability. Compound 47 is composed of an acenaphthopyrazine-­8,9-­dicarbonitrile
acceptor (APDC) core and TPA donor units as peripheries [49]. It is found that APDC possesses a
stronger electron-­accepting character than DCDPP because the central fluoranthene core with six
π electrons tends to form an aromatic cyclopentadienyl structure. As a result of the stronger accep-
tor system, 47 exhibits a red-­shifted emission, peaked at 756 nm, in comparison with 46. The opti-
mized geometry also showed that the dihedral angle between APDC and TPA units in 47 forms an
orthogonal conformation that favors the ICT state, making more lower the LUMO energy level
than those of previous compounds. Furthermore, the bulkier conformation of 47 not only prohib-
its the π–π stacking of the planar central core but also restricts the intramolecular motion of the
phenyl rings, both of which contributes to the high PLQY 17% of 47. This value is unusual for any
organic compound with an emission maximum above 750 nm.
The ICT process through-­bond is not only the way of achieving the charge-­transfer effect in
D–π–A AIEgens with aromatic stator. T. Swager et  al. reported aggregation-­induced enhanced
delay fluorescence (AIEDF) compounds 48–50 with through-­space ICT [50]. 48–50 presents a π
unit of xanthene substituted in 4-­ and 5-­positions with a 2,4-­diphenyl-­1,3,5-­triazine as acceptor
and TPZ, Cbz, and tBu-­Cbz as donors. All compounds exhibit a spatial confinement of HOMO on
the donor unit and LUMO on the acceptor unit and then low ΔEST values, which activate the DF
prompt. Emission spectra of 48–50 show strong solvatochromic effect; the maximum of the 48
emission shifts from 524 nm in cyclohexane to 645 nm in acetone, which is indicative of a large
increase in polarity in the excited state. Notably, the 48 emission exhibited a remarkably large
Stokes shift from 214 nm in cyclohexane to 335 nm in acetone (524 and 645 nm emissions relative
to the 310 nm) as a result of its strong through-­space ICT character. 48–50 are purely emissive in
solution but strong DF in film-­cast and aggregate, demonstrating AIEDF activity.

20.4  ­Push–Pull AIEgens from ACQ Chromophores

The fluorescence behavior of small ACQ luminophores can be thoroughly changed from ACQ to
AIE by decorating them with AIE-­active units, which confer them sufficient twisted structures
able to avoid strong π–π stacking interaction in the solid state. By introducing electron donors and
acceptors and AIE activators into the conjugated structures, many red/NIR AIEgens can be pre-
pared. This modular synthesis typical of the push–pull strategy allows a tuning of the molecular
structures that lead to novel AIEgens with tailored properties depending on the purposes. This
synthetic strategy results are extremely powerful when the number of known ACQgens is consid-
ered. We report the most popular examples of this strategy. All examples are subdivided into clus-
ters with the acceptor moiety as the common denominator. We have preferred to use the acceptor
moiety as the base-­of-­classification in order to underline the progress coming by application of
basic principles on a specific prefixed molecular structure.

20.4.1  BT-­based AIEgens


Benzothiadiazole is one of the most typical low-­band-­gap acceptor, and its excellent photo/ther-
mostable properties have been applied broadly in the optoelectronics and biological imaging [51].
The red-­emitting AIE performances of BT–π–D compounds can be induced by the twisted molecu-
lar structures and D–A effects: the intermolecular π–π interactions are restrained by the twisted
590 20  Push–Pull AIEgens

flank units (TPE, MTPE, TPA, MTPA, etc.) resulting in a strong emission of their solid states, while
their Eg values are reduced by the strong D–A effects and thus long-­wavelength emission can be
observed. To achieve red-­emitting BT-­based AIEgens with greater PLQY, a balance of the twisted
molecular conformation and push–pull behaviors is fundamental. The importance of the BT core
for the construction of push–pull AIEgens is confirmed by several modifications of the BT scaffold
proposed in order to increase its acceptor and solubility features such as BTP, BI, BBTP, and QBT
cores. The following examples illustrate the strategy adopted to make AIE-­active BT-­based chromo-
phores. The molecular structures of examples are shown in Figures 20.7 and 20.8 and their photo-
physical properties are summarized in Table 20.1.
A series of TPE-­modified BT AIEgens 51–53 were prepared by B. Z. Tang et  al.  [52]. TPE-­
substituted heterocyclic compounds 51–53 are composed of BT or BT-­thiophenes as the core and
TPE as peripheries. Compounds 51–53 present good PLQY in solution. The reasonably strong
emission of the present luminogens in solutions suggests that the excitons are mainly localized on
the BT and BT-­thiophene cores, whose decay process is less influenced by the RIM process of TPE
units. Increasing the extension of acceptor by the introduction of thiophene rings, the fluorescent
colors of compounds 51–53 can be tuned from green to red in their solid states, respectively.
However, the introduction of thiophene rings decreases the molecular rigidity with a correspond-
ing decrease of the PLQY by activation of nonradiative processes. Compound 53 is used to fabricate
nondoped OLEDs, which display a red electroluminescence peak at 668  nm with a maximum
luminance of 1640 cd/m2 and a maximum EQE of 1.0%.
BT chromophores are well known to give ICT effects. ICT effects can play an essential role in the
emission of BT–π–D molecules, especially when the ICT emission rather than locally excited (LE)
emission is dominant. One way to red-­shift ICT emission is to design molecules with more twisted
D–A structure to lower the energy level of the ICT state.
Ishi-­i et al. prepared a series of D–A AIEgens 54–57 with efficient red emission in their solid states
by connecting BT to TPA units by a π spacer [53]. Compound 54 has a BT core directly linked to TPA
units, while compounds 55–57 present vinylene, ethynylene, and aryl ring groups as π spacers,
respectively. Compounds 54–57 exhibit both TICT and red-­emitting AIEE properties. We take 54 as

Figure 20.7  Molecular structure of BBTD-­based, QTD-­based, and BT-­based AIEgens.


20.4 ­Push–Pull AIEgens from ACQ Chromophore 591

S
N N
Ph TPE N N N N
N N
S S
TPE N Ph TPE TPE TPE TPE

62 63 64

S S S S
N N N N N N N N

R R R R MTPA MTPA TPA TPA

N N N N N N
S S
65 R = TPE 67 R = TPE
69 70
66 R = MTPE 68 R = MTPE

Figure 20.8  Molecular structure of BTP-­/BI-­/BBTD-­based and QTD-­based compounds.

Table 20.1  Photophysical properties of BT-­based AIEgens.

Solution Aggregate Film

Compound λabs (nm) λem (nm) PLQY λem (nm) PLQY λem (nm) PLQY

51 418 538 61 —­ —­ 539 89.0


52 464 592 37 —­ —­ 600 55.0
53 510 623 25 —­ —­ 661 —­
54 460 614 71 610 80 —­ —­
55 500 640 —­ 647 15 —­ —­
56 465 615 —­ 622 28 —­ —­
57 440 618 —­ 588 44 —­ —­
58 471 616 —­ 609 —­ 617 48.8
59 470 613 —­ 604 —­ 617 63.0
60 522 666 32 —­ —­ 674 44.0
61 520 658 32 —­ —­ 658 30.0
62 —­ —­ —­ —­ —­ 656   7.25
63 470 588 —­ 593 —­ 617 56.6
64 568 701 —­ —­ —­ 767   0.50
65 —­ —­ —­ —­ —­ 704 36.9
66 —­ —­ —­ —­ —­ 761   8.2
67 612 —­ —­ 787 3 —­ —­
68 632 —­ —­ 857 6.4 —­ —­
69 —­ —­ —­ —­ —­ 1050   2.9
70 582 —­ —­ —­ —­   696 60
592 20  Push–Pull AIEgens

an example to describe the TICT–AIE behavior of these compounds. 54 is highly emissive in the
THF solution at 614 nm with a PLQY value of 71%, but when water is gradually added, emission
progressively decreased and it becomes bathochromically shifted until 639 nm, where 8% of PLQY
is recorded. However, the emission becomes stronger and hypsochromically shifts when the per-
centage of water is over 90%, recording a PLQY value of 80% and the emission shifted to 610 nm.
The emissive behavior is demonstrated to be fine-­tuned by the nature of the spacers. When adding
a vinyl group between the D and A units, it leads to a more planar conformation and thus more red-­
shifted absorption and emission (55 vs. 54). In contrast, when the linker is an ethynyl group (56), the
optical range slightly changes in comparison with 54. Intriguingly, changing the π linker to a ben-
zene ring (57) disrupts the communication between D and A, which significantly blue-­shifts the
optical range. In all cases, the PLQY values are affected by the introduction of a spacer because the
molecular rigidity in nanoaggregates is weakened through the higher distance between neighbors
with consequent activation of the nonradiative channel. Qin et al. prepared two D–A compounds 58
and 59, which are composed of electron-­donating arylamine groups, electron-­accepting BT core,
and AIE-­active TPE moieties [54]. Both compounds 58 and 59 show TICT and AIE behaviors in the
THF/water mixed solvents. In THF, they show an orange-­red fluorescence with emission peaks at
616 and 613  nm, respectively. When water fraction is increased, the emissions of 58 and 59 are
gradually weakened and red-­shifted, after that they intensify and blue-­shifts at 616 and 613 nm,
respectively, with a further increase of water fraction. Similar to previous cases, compounds 60 and
61 proposed by B. Z. Tang and collaborators show both ICT and AIE features [55]. Compared to
previous 58 and 59, which possess similar chemical structures to 60 and 61 but with two additional
phenyl rings between the BT and amine units, 60 and 61 showed much redder absorptions and
emissions (522/520 and 666/658 nm). The red shift of the absorptions and emission of 60 and 61
should be ascribed to their stronger ICT effect as compared to those of 58 and 59 because their
electron-­donating nitrogens are spatially separated from the electron-­accepting BT moiety by the
additional phenyl rings; more phenyl ring substitutions mitigate the electron density of the nitro-
gen, thus weakening the ICT. This indicates that the direct attachment of the electron donor to the
electron acceptor is a more efficient approach to extend both absorption and emission due to the
more effective ICT effect. Slight changes in the emission properties of 60 and 61 from solution to
thin film are observed, implying that the molecular deformation of 60 and 61 in the excited state is
inactive, possibly due to the good electron conjugation and the large steric hindrance between the
arylamine and the BT unit, which limit the rotors from free rotation or vibration. Consequently, the
nonradiative decay pathway of the exciton deactivation was greatly suppressed and thus 60 and 61
fluoresced brightly. Theoretical calculations confirm that 60 and 61 possess twisted configurations
with similar torsion angles of 33.44–54.62° between the BT unit and the phenyl rings of the adjacent
arylamines. These twisted conformations of the two compounds were derived from the intrinsic
twisting structure of the TPE segments and the large steric hindrance between the BT unit and the
phenyl rings of the adjacent arylamine. Such bulky conformations and crowded structures of the
two molecules not only prohibit close intermolecular packing but also forbid intramolecular motion,
both of which contribute to the bright solid emission.
In comparison with BT, thiadiazolo[3,4-­c]pyridine (BTP) has a stronger electron-­withdrawing
ability. Jiang et  al. prepared a D–A-­type molecule 62 by using arylamines as electron-­donating
units, BTP as electron-­accepting groups, and TPE as AIE activators  [56]. 62 results from being
weakly emissive in the DMF solution but shows a 32-­fold emission enhancement in aggregates. In
the solid state, it emits a red fluorescence peaked at 656 nm with a PLQY value of 7.25%. Although
62 and 58 have similar structures, the emission maximum of 62 is bathochromically shifted to
39 nm in comparison with that of 58 due to the stronger electron-­withdrawing ability of BTP in 62
20.4 ­Push–Pull AIEgens from ACQ Chromophore 593

than of BT in 58. Another intriguing modification to the BT core for the construction of AIEgens
is proposed by B. Z. Tang et al., which substitute the BT core with the more electron-­withdrawing
2H-­benzo[d]imidazole core (BI) [57]. The BI-­based AIEgen 63 is composed of an AIE-­active TPE
unit as the donor, a strong electron-­withdrawing BI moiety as the core, and steric cyclohexane as
bulky substitution against the intermolecular π–π stacking and increasing molecular solubility. BI-­
based AIEgen 64, the owner of thiophene spacers between the BI core and TPE units, exhibits NIR
emission for in  vivo bioimaging applications. Both molecules are AIE-­active. Compared to 51,
compound 63 exhibits (i) red-­shifted absorption and emission in solution, resulting in the stronger
electron-­withdrawing capability of the BI core to diminish the Eg; (ii) better AIE performance certi-
fied by a threefold emission enhancement than 51 in aggregate; and (iii) brighter orange-­red solid
luminescence at a 617  nm wavelength with 56.6% of absolute quantum efficiency, which was
attributed to the synergistic interaction of the electron-­deficient BI acceptor and the intermolecu-
lar antistacking between spiro-­cyclohexane substituents. The antistacking behavior of spiro-­
cyclohexane substituents on the BI core played a significant role in the suppression of nonradiative
pathway to enhance solid luminescence of 63 and 64. According to the single crystal and packing
structures, the bulky cyclohexane substituted on the BI core and its peripheral TPE segments pos-
sess the required steric hindrance, enabling us to separate each molecule. Besides, TPE segments
mainly play the role of molecular rotators to consume the energy of the excited molecule, leading
to the weak emission in solution. Such intramolecular motions are restricted by multiple intermo-
lecular electrostatic interactions in the crystalline state; at the same time, the molecular conforma-
tions are also rigidified.
[1,2,5]-­Thiadiazolo-­[3,4-­g]quinoxaline (QTD) and benzobis(thiadiazole) (BBTD) are typical
modifications of the BT core with strong electron-­withdrawing abilities; as a result, it has been
widely used to tune the electronic structure of conjugate compounds  [58]. Du et  al. reported a
series of NIR-­emitting molecules 65–68, which are composed of electron-­withdrawing QTD or
BBTD moieties and electron-­donating TPE or MTPE groups and their fluorescence behavior has
been comparatively studied [59]. Compound 68 is AIE-­active because it shows a weak emission
with a PLQY value of 0.20% in solution and an enhanced emission in aggregate dispersion (6.39%).
Surprisingly, when nonsolvent is progressively added into the solution of 67, the fluorescence
intensity of 67 is gradually decreased. The PLQY value of 67 is 13.0% in solvent but decreases to
0.26% in aggregate. This ACQ phenomenon may be attributed to the strong D–A effect of 67, which
makes the structure of 67 planar, allowing π–π interactions between neighbors in the solid state.
On the contrary, both compounds 65 and 66 exhibit AIE behaviors showing low emissions in solu-
tion, with a PLQY of 10.1 and 0.28% respectively, and enhanced NIR emission in aggregates with a
PLQY of 36.9 and 8.2%. Comparison of 65 and 66 evidences the effect of the donor moiety onto the
photophysical emissive properties. 66 displays an emission peak at 780 nm in solution, demon-
strating a red shift of 80 nm in comparison with that of 65, owing to the stronger electron-­donating
ability of MTPE in 66 compared with those of TPE in 65. In addition, the electron-­accepting ability
of the acceptors also affects the fluorescence behavior of these compounds. Because of the stronger
electron-­withdrawing ability of BBTD compared with that of QTD, the emission peak of 68 is
857 nm, which is bathochromically shifted by 77 nm in comparison with that of 66. NIR-­emitting
AIE molecule 69 presents a BBTD moiety flanked by two MTPA units and an electron-­donating
unit stronger than those of compounds 68 [60]. 69 shows a very weak luminescence with a PLQY
value of 0.1% in solution, but the NIR emission is gradually enhanced upon increasing the nonsol-
vent fraction with a PLQY value increase of up to 2.9%, exhibiting a typical AIE performance. The
solid films of 69 show bright NIR emission with a peak at 1050  nm. Both the rigidity and π-­
extension of 70 contribute strongly on the formation of the ICT state as well as red shift of the
594 20  Push–Pull AIEgens

emission wavelength. Compared to 54, the rigidity of 70 is increased by steric hindrance of the
hydrogen atoms. 70 can form highly twisted D–A conformation in the excited state, which favors
the formation of the ICT state to yield a PLQY of 60%, one of the higher PLQY values for AIEgens
with a wavelength of over 650 nm.

20.4.2  Cyanine and DCM-­based AIEgens


Cyanine dyes demonstrate large molar absorptivity, high PLQY, and good biocompatibility. Their
fluorescence spectra can be well tuned from the visible region to the NIR region by altering the
length of the conjugated chain. However, these dyes display a small Stokes shift and ACQ effect,
which limits their practical applications [61]. The push–pull AIEgens incorporating a cyanine and
hemicyanine dyes are reported in Figure 20.9.
Decoration of N-­alkylated indolinum group with the TPE unit, B. Z. Tang et al. prepared AIE-­
active hemicyanine dye 71 [62]. 71 is weakly fluorescent in solution with a maximum wavelength of
625 nm, a PLQY value of 0.15%, and a very large Stokes shift of 185 nm. With the progressive addi-
tion of nonsolvents, the fluorescence intensity is increased gradually until stabilizing at an amplifi-
cation value of 15 times greater compared with that of its solution. The PLQY value for the solid
films of 71 is as high as 55.8%, showing a 372-­fold enhancement compared with that in solution.
Besides its AIE activity, 71 is yet to be pH-­responsive, which makes it a pH sensor. At a pH
value below 5, 71 displays a red fluorescence with unchanging intensity, but when the pH is
above 5, the fluorescence intensity decreases gradually until the critical threshold pH-­value of
10, where the emission intensity is almost zero. At a pH above 10, 71 shows a blue emission
peak at 480 nm with strong intensification. Going from acid to neutral conditions, gradually
deprotonation decreases the hydrophobicity of 71, changes the aggregation mode and so the

TPE
TPE R

S
N Et
N
(CH2)4SO3
S
N 73 R = H
(CH2)4SO3
71 72 74 R = OC6H13
75 R = OC12H25

NC CN

TPE TPE
O

N N
76

Figure 20.9  Molecular structure of cyanine-­based and DCM-­based AIEgens.


20.4 ­Push–Pull AIEgens from ACQ Chromophore 595

fluorescence intensities. In basic buffer solutions, 71 breaks its conjugation system, resulting in
short-­wavelength emission because its hydrophobicity is enhanced due to the neutralization of
nitrogen ions. Analogue substitution with the TPE unit of benzothiazolium generates hemicya-
nine dye 72 with a red-­emitting AIE activity [63]. 72 is weakly emissive in solution with a peak
at 663 nm and a PLQY value of 3.66%. With the progressive addition of nonsolvents, the emis-
sion intensity is first reduced and red-­shifted by the TICT phenomenon, then enhanced more
than threefold, recording a PLQY value of 13.12%, and blue-­shifted for the coming of the AIE
effect. Interestingly, the crystal of 72 grown from different solvent mixtures possesses different
crystal structures and fluoresces (green, yellow, and orange, with the corresponding emission
peaks at 565, 578, and 591 nm) with different intensities. Hemicyanine dyes 73–75 are also red-­
emitting molecules [64]. Owing to their enhanced D–A interactions, 73–75 show strong deep
red emission peaks at around 710  nm in the aggregation state. Unlike 74 and 75, which are
AIE-­active, 73 shows a red emission peak at 710 nm with a PLQY value of 3.4% in solution and
the emission intensity is gradually enhanced in the aggregate state, exhibiting an AIEE activity.
Without an alkoxy chain, 73 shows AIEE activity because its excited energy in solution cannot
be completely consumed by the rotations of the phenyl rings. On the contrary, the excited ener-
gies of 74 and 75 can be completely dissipated by the active rotations of the TPE moieties and
alkoxy chains, resulting in the AIE activities of 74 and 75.
Dicyanomethylene-­4H-­pyran (DCMs) are typical red-­emissive dyes with a strong electron-­
accepting ability; thus, the attachment of an electron donor to the DCM core can effectively
decrease the Eg. However, their D–π–A structures show serious ACQ effects. B. Z. Tang et al. intro-
duced two TPE units to both sides of a TPA-­modified DCM core to generate AIE-­active compound
76 [65]. 76 emits a red light peaked at 633 nm in solution. With the addition of nonsolvents, the
fluorescence intensity decreases and the emission is progressively blue-­shifted, and then later
gradually increases bathochromically shifting at 660 nm. These features indicate that 76 is both
TICT-­ and AIE-­active. With excellent properties such as high quantum yield, red emission, and
AIE activity, molecule 76 has been widely used in the field of bioimaging.

20.4.3  QM-­based AIEgens


Replacing the oxygen atom in quinoline-­malononitrile (QM) by N-­ethyl, its planar conformation
can be destroyed and introducing aromatic ring rotors is possible to generate a series of red-­
emitting AIE molecules 77–82. Their properties are summarized in Table 20.2, while their struc-
tures are shown in Figure 20.10.
Compound 77 prepared by H. Tian and coworkers presents a QM core as A unit and N,N-­
dimethylaniline as D unit [66]. 77 shows a weak fluorescence centered at 594 nm and with a
PLQY of 0.1% in solution. When a certain amount of nonsolvent is added, its fluorescence is
gradually enhanced and red-­shifted. The solid-­state PLQY of 77 reaches 14.7%. Besides its AIE
activity, 77 shows other excellent properties, such as the ability to self-­assemble into ordered
microstructures, optical waveguide properties, and mechanochromic effects. By incorporating a
sulfonate unit into the QM core, water-­soluble red-­emitting molecule 78 is prepared [67]. 78 is
nonemissive in water, while its aggregates formed in water/ethanol mixtures emit a strong red
fluorescence with a maximum PLQY value of 27.5%, 275-­fold amplified than that of its pure
water solution. Owing to its AIE properties, red emission, and water miscibility, 78 is used as
fluorescent bioprobes for trypsin and cell tracking. Substituting N,N-­dimethylaniline with
5-­methylthiophene donor unit and TPA donor unit, compounds 79 and 80 were prepared by Wei-­
Hong Zhu et al. [68, 69]. 79 and 80 exhibit typical AIE activity in the solution/aggregate system,
596 20  Push–Pull AIEgens

Table 20.2  Photophysical properties of compounds 77–82.

Solution Aggregate Film

Compounds λabs (nm) λem (nm) PLQY (%) λem (nm) PLQY (%) λem (nm) PLQY (%)

77 430 594 0.1 614 4.1 604 14.7


78 415 —­ 0.1 610 27.5 —­ —­
79 452 589 0.04 612 7.1 651 6.1
80 430 584 0.1 584 42.9 625 8.6
81 452 626 0.06 615 11.0 655 5.9
82 466 672 0.1 705 5.1 673 4.0

NC CN
NC CN NC CN

N S
N N TPA
R
N

77 R = Et
79 80
78 R = (CH2)3SO3

NC CN NC CN

S S
N TPA N MTPA

81 82

Figure 20.10  QM-­based AIEgen structures.

showing an enhanced PLQY of aggregate (7.1 and 42.9%) compared with those of respective
solution (0.04 and 0.1%). However, 80 has a PLQY higher than that of 79, attributing to a more
twisted TPA unit able to deactivate the nonradiative channels more efficiently. Moreover, the
fluorescence emission of 80 shows a significant blue shift with decreasing polarity, indicating
that it has ICT effect. In the solid state, 80 showed an intensive fluorescence emission with about
a 40-­nm red shift in maximum emission from solution to solid powder, which is caused by the
fact that 80 aggregates are more tighter in solid than in solution.
Furthermore, the effects of electron-­donating groups and thiophene-­based π-­bridges on the
emission wavelength and the morphology of the organic nanostructures of red-­emitting AIE-­
active QM derivatives 81 and 82 are investigated  [61]. The emission peaks are progressively
red-­shifted when stronger D units are introduced into the thiophene moieties. The emission
peak of 81 is at 655 nm, but by substituting TPA with a stronger D unit MTPA, the resultant
molecule 82 shows an emission maximum red shift of 18 nm. Thanks to their bright aggregate-­
state red fluorescence, 81 and 82 are used as bioprobes for cell tracking, which show a good
long-­term cell tracking ability.
20.4 ­Push–Pull AIEgens from ACQ Chromophore 597

20.4.4  DPP-­based AIEgens


The DPP-­based AIEgens are reported in Figure  20.11, while their photophysical properties are
reported in Table  20.3. Diketopyrrolopyrrole (DPP) is one of the most investigated red-­emitting
materials for OLEDs and bioimaging due to its narrow Eg. The Eg and the emission colors of DPP
acceptors can be easily tuned by the attachment of an electron donor to the DPP core. However,
owing to the well-­conjugated structures of DPP and its derivatives, they are prone to form aggre-
gates in concentrated solutions and possess heavy ACQ effects.
B. Z. Tang and coworkers were the first to investigate fluorescence behaviors in DPP-­based deriv-
atives 78–81 (Figure  20.11), and their photophysical data are summarized in Table  20.3  [70].
Attaching two TPE units to the DPP core directly and with phenyl unit as spacers, the authors have
synthetized adducts 78 and 79, respectively. Molecule 77 has a high PLQY in solution (95.2%), but
it drastically decreases in poor solvent system (0.95%) and film cast (18.3%), showing a heavy ACQ
effect. Just like 77, structure 78 is also ACQ-­active, demonstrating a fluorescence in solution with
a PLQY value (13.6%) higher than those recorded in aggregate (1.27%) and film (11.8%). These
PLQY’s trends suggest that DPP chromophores cannot be transformed into an AIEgen by the direct
coupling of the neutral TPE unit to the DPP core. The reason for the ACQ effect of 78 is to be
ascribed to a higher conjugation between the DPP core and direct-­linked phenyl moieties of the

R C8H17 C8H17
TPE
D π N O N N
TPE O O

O π D O TPE O
N N N
R TPE
C8H17 C8H17
DPP 78 77 R = Br
79 R = TPE

C8H17 R1 C8H17
R1
N O N O

R1 = OMe
O O
N N S CHO
N Br C8H17
C8H17

80 81

OMe R1 C8H17 R2 C8H17


N O O
O

O N O
R1 R2
R2 = C8H17 C8H17

83 R = C4H9
84 R = C8H17
82
85 R = benzyl
86 R = (CH2)6NEt2

Figure 20.11  Structure of AIEgens incorporating the DPP core as the acceptor unit.
598 20  Push–Pull AIEgens

Table 20.3  Photophysical properties of DPP-­based AIEgens in solution and aggregate.

Solution Aggregate Film

Compound λabs λem PLQY λem PLQY λem PLQY [%]

77 480 543 95.2 537   0.95 558 18.3


78 503 572 13.6 568   1.27 595 11.8
79 496 568 43.0 592   4.10 594 30.2
80 500 —­  0 670 29 —­ —­
81 520 —­  0 685  9 —­ —­
82 515 —­  0 655 24 —­ —­
83 507 597   0.17 615 —­ 651   4.8
84 509 597   0.09 615 —­ 654 56.0
85 507 580   0.36 615 —­ 640 57.0
86 507 620   2.18 588 13.7 —­ —­

TPE unit, which imposes a planar structural conformation. In order to avoid this, the authors used
a phenyl spacer in a D–π–A structure, resulting in product 79. In solution, molecule 79 presents a
yellow fluorescence, hypsochromically shifted than the fluorescence of 78 because of inefficient
conjugation extension by the introduction of phenyl spacers and an enhanced PLQY when com-
pared with those of 78. In aggregates and cast films, the red fluorescence at 592 and 594 nm with
PLQY values of 4.1 and 30.2%, respectively, are recorded. The PLQY values for 79 are different from
78 and 77 in both solutions and films. The PLQY change is ascribed to the intramolecular rotations,
which can exhaust the energy of the excited state. For 77, there are no extra rotational phenyls
around the DPP core; it has the highest PLQY in solution (95.2%) but low PLQY (18.3%) in the film.
For 78, the rotational phenyls are closely linked to the DPP core; its PLQY steeply decreases in solu-
tion (13.6%) and film (11.8%). For 79, the insertion of phenyl bridges between DPP and TPE moie-
ties disturbs the intermolecular π–π interaction in the solid state, which leads to the highest PLQY
among all molecules (30.2%), but active in the rotation-­caused energy dissipation in solution
(43.0%). Despite an enhancement of solid-­state-­emissive behaviors observed from 77 to 79, the
strategy involving the decoration of DPP core with neutral TPE resulted in nonperforming to
obtain AIEgen.
The incorporation of triphenylamine (TPA) derivatives, the owner of their twisted conformation
on nitrogen atom, in the DPP scaffold was proposed as a successful strategy to achieve red-­emitting
AIEgens. J. Hua et al. [71] reported compounds 80, 81, and 82 with two-­photon absorption (2PA)
behaviors, which present a DPP core flanked by one TPA unit (80), one TPA unit and one
thiophene-­2-­carbaldehyde (81), and two TPA units (82), respectively.
All compounds are AIE-­active, exhibiting weakly fluorescence in solution but red fluores-
cence emissive in the solid state with interesting PLQY for compounds 80 (29%) and 82 (24%).
The twisted conformation of phenyl rings in the TPA units in these compounds is able to dissi-
pate the excited energy in the solution state, leading to weak emission. Besides, in the aggrega-
tion state, the RIM switch on the radiative channel avoiding fluorescence quenching by the
increase of the distance between neighboring molecules in crystal lattice and reducing the inter-
molecular π–π stacking interactions. In this case, the nature of molecular aggregation is con-
trolled by a delicate balance of various weak intermolecular interactions at the conjugated
20.4 ­Push–Pull AIEgens from ACQ Chromophore 599

backbone level. However, the force and nature of such intermolecular interactions can be
fine-­tuned also from remote functional groups, which are not in conjugation. Therefore, both
variations in the structure of π-­conjugated backbones and the remote functional groups are uti-
lized to fine-­tune the solid-­state photophysical AIE properties.
D. Cao et al. have synthetized a series of red-­emissive DPP-­based AIE molecules 83–85 with dif-
ferent N-­alkyl substituents and are shown in Figure 20.11, using anthranone as flanker units and
ethane bridge as π units [72, 73]. Compounds 83–85 are nonemissive in dilute solution but they
show typical AIE curves in the solvent/nonsolvent mixtures and solid films emit a bright red fluo-
rescence. As can be seen in Table 20.3, compounds 83–85 show different AIE characteristics origi-
nating from the difference in their alkyl side chains due to the different strengths of intermolecular
π–π interactions when the groups were altered from butyl to N,N-­diethylhexyl.
As mentioned in paragraph 20.2.1 of this chapter, the energy of excitons and their energy migra-
tions are related to the distance a between identical neighbors in the aggregate lattice. According
to the exciton’s model, by acting on the aforementioned parameter a, we can simultaneously tune
the energy of the excited states, the emission wavelength, the exciton transfer, and migration pro-
cesses, from which the PLQY values depend. The low PLQY of 83 can be attributed to the faint
steric effect of butyl substituents, which leads to strong intermolecular π–π interactions, opening
the nonradiative channel of the exciton migration process. Contrarily, in 84 and 85, the distance a
between neighbors results in optimal to minimize the loss fluorescence by nonradiative decay, and
their PLQY is enhanced. Compound 86 shows a pH response behavior due to the reactivity of
diethylamino groups to H+ and OH−, which induces dissolution and aggregation in the acidic and
basic aqueous media. When the pH is lower than 9, 86 is nonemissive because the ammonium salt
is formed and dissolved in solution. When the pH is higher than 10, the emission is progressively
increased because the ammonium salt is decomposed and thus aggregates are formed.

20.4.5  Rylene-­based AIEgens


1,4,5,8-­Naphthalene diimide (NDI) dyes are electron-­deficient scaffolds that can be fine-­tuned to
make them AIE-­active by linking with donor units at the core positions. Li et al. prepared a series
of NDIs 87–92 with different aryl groups at the two flanking sides for OFET applications
(Figure 20.12) [74].
87–92 are AIE-­active; they are weakly emissive in their solutions but display intensive emission in
the aggregate and solid states. The emission colors can be changed from green to red by increasing
the electron-­donating capacity of the flanking groups. For instance, compound 87, the owner of two
pyridine flanking groups, shows a green fluorescence at 541 nm. Changing with thiophene groups,
compounds 90 and 91 exhibit a red fluorescence at 708 and 630  nm, respectively. Furthermore,
modifying substituents on the nitrogen atoms, the authors have fine-­tuned the strength of the inter-
molecular π–π stacking interactions with beneficial effects on PLQY values and red-­shifted emis-
sion. For example, 88 and 89 with hexyl side chains exhibit higher PLQY values and longer emission
wavelengths than those of 90 and 91 that have C20-­alkyl side chains because the intramolecular
rotations in NDIs with C20-­alkyl groups cannot be completely restricted. Aggregation-­induced mul-
ticolor emission by side-­chain engineering was recently studied by P. K. Das et al. [75]. The authors
prepared a series of NDIs 93–95 AIE-­active compounds differing from their side chains. Side chains
are based on ʟ-­aspartic acid linked at amphiphilic molecules having a benzyl ester group at both
terminals with varying substituents electron-­reach and electron-­poor. All compounds are nonemis-
sive or weakly emissive in the solution, but strong emissive upon the increase of nonsolvent by the
formation of J-­type aggregates. 94–97 exhibit excimer-­induced enhanced multicolor emission
600 20  Push–Pull AIEgens

R
N S S
O N O
R1 R1 =

R = C6H13 87 88 89
R1
O N O
R R = C20H41 90 91 92

R NO2 Cl
O O O O
O 93 94
O N N O
O OMe
O O O O 97
R
95 96

C4H9
O N O O N O O N O

R1 R R R
R2 R

O N O O N O O N O
C4H9

98 R1 = R2 = Br 101 R = TPE 103 R = TPE


102 R = OTPE
99 R1 = TPE R2 = H
100 R1 = R2 = TPE

Figure 20.12  Structure of AIEgens incorporating perylene bisimide core as the acceptor unit.

through the variation of the electron-­reach and electron-­poor aromatic substitutions at the terminal
of NDI derivatives [76]. The nature of aromatic substituents facilitate the intramolecular charge
transfer to the NDI core resulting in a change of the wavelength emission of the excimer band as
well as the emission color from cyan blue to bright orange.
Perylene bisimides (PBI) possess an extended conjugated system, which have been used as red-­
emissive materials in electronic and optical fields [77–79], due to their remarkable thermal and
photochemical stabilities, near-­unity PLQY, and high electron affinity. However, PBIs show a
strong ACQ effect, the owner of the powerful π–π stacking interactions among neighboring mole-
cules, which greatly limits their practical applications.
Tang et al. firstly prepared red-­emissive AIE molecules by introducing TPE moieties onto the
PBI cores and systematically investigated the fluorescence performances of the resulting TPE–PBI
adducts summarized in Table 20.4. Firstly, the effects of the number of TPE on the fluorescence
20.4 ­Push–Pull AIEgens from ACQ Chromophore 601

Table 20.4  Photophysical properties of PBI-­based AIEgens in solution and aggregate.

Solution Aggregate

Compound λabs λem PLQY λem PLQY

  98 478 540 95 603   2.2


  99 478 538 2.2 638   9.0
100 478 —­ 0.07 706 18.9
101 478 —­ 0.07 668   6.3
102 488 —­ 0.03 670 13.0
103 478 —­ 0.07 640 29.7

performances of these adducts are comparatively investigated [80]. The ACQ behavior of non-­TPE-­
substituted precursor 98 results from its strong green light fluorescence in dilute solution (540 nm)
with high PLQY (92%) and very weak red fluorescence in aggregates. The PBI derivative 99 with
one TPE unit in solution shows a typical AIEE performance with an emission with a PLQY of 2.2%,
much lower than those of 98, and a brightened red emission (λem = 638 nm) with a PLQY value of
9.0% in aggregate. Contrary, structure 100 results to be nonemissive in dilute solution, but in aggre-
gate, it shows an intensive NIR emission with a high PLQY of 18.9%. The above results indicate
that 98 is only partially converted into the AIE emitter 99 owing just one TPE unit, but it can be
completely transformed into an AIEgen (compound 100) by modification with two TPE units at its
1,7-­positions. The above results indicate that 99 and 100 display red fluorescence in their aggregate
and solid states because their twisted structures hinder the π–π stacking interactions among neigh-
bors and at the same time the crowded space restricts the rotations of phenyl rings in the TPE units.
Secondly, the effects of the relative positions of the TPE units in the PBI cores on the fluores-
cence performances of the TPE–PBI adducts have also been comparatively investigated  [81].
Attaching TPE groups at the 1,7-­ and 1,6-­positions to the PBI cores leads to two isomers 101 and
103 that are AIE-­active with some differences in the fluorescence behaviors. For instance, they are
nonemissive in dilute solution (PLQY  =  0.07%); however, their aggregates display a strong red
emission centered at 670 and 640 nm with PLQY values of 22.8 and 29.7%, respectively. These data
suggest that 101 has a better AIE performance than 103 due to its more symmetrical 1,7-­substituted
pattern. Although TPE units are bulky and twisted, both 101 and 103 can form ordered crystal
structures with strong red emission and good waveguide performance.
Considerable differences in the photophysical properties of aggregates 100 and 101 are reported
in Table 20.4. Structure 100 exhibits a red-­shifted fluorescence with a greater value of PLQY if com-
pared with structure 102, only for a different alkyl side chain on the nitrogen atoms of the PBI scaf-
fold. Since the crystalline structures of 99 are not reported, a direct comparison is not possible and
no consideration can be done. However, a possible cause can be attributed to the different mor-
phologies of the aggregate states. Contrarily, a direct comparison can be done between structures
101 and 102 differing for the oxygen atom that link both TPE units at the 1,7-­position of PBI scaf-
fold [82]. Structure 102 is AIE-­active, showing a strong red emission with a PLQY value of 13.0% in
the solid state, a 433-­fold augmentation in comparison with that of its solution. These data demon-
strate that the oxygen bridge cannot obstruct the transformation of PBI from an ACQ-­phore to an
AIE-­gen. With efficient fluorescence in its solid state, the red-­emitting dye 102 is encapsulated into
nanoparticles (NPs) with the DSPE–PEG2000 and DSPE–PEG5000 for the application in bioimaging.
602 20  Push–Pull AIEgens

20.5 ­Concluding Remarks

We have introduced the basic concepts of solid-­state exciton theory, and we have given an overview
on the basic concepts of AIE molecular engineering when combined with a push–pull design. We
have summarized recent advances in the world of molecular AIEgens with a push–pull structure.
Several examples that make use of such design principles have been illustrated.

­References

1 Misra, R., and Bhattacharyya, S. P. (2018). Intramolecular Charge Transfer Theory and Applications,
Wiley-­VCH, Weinheim.
2 (a) Vogelsang, J. (1993). Extended mechanism of excimer formation. J. Chem. Soc. Faraday
Trans. 89: 15.
3 (a) Zhong, G. Y., Xu, Z., He, J., Zhang, S. T., Zhan, Y. Q., Wang, X. J., Xiong, Z. H., Shi, H. Z., and
Ding, X. M. (2002). Aggregation and permeation of 4-­(dicynomethylene)-­2-­methyl-­6-­(p-­
dimethylaminostyryl)-­4H-­pyran molecules in Alq3. Appl. Phys. Lett. 81: 1122–1124.(b) Chiang,
C. L., Wu, M. T., Dai, D. C., Wen, Y. S., Wang, J. K., and Chen, C. T. (2005). Red-­emitting fluorenes
as efficient emitting hosts for non-­doped, organic red-­light-­emitting diodes. Adv. Funct. Mater. 15:
231–238. https://doi.org/10.1002/adfm.200400102
4 Luo, J., Xie, Z., Lam, J. W. Y., Cheng, L., Chen, H., Qiu, C., Kwok, H. S., Zhan, X., Liu, Y., Zhu, D.,
and Tang, B. Z. (2001). Aggregation-­induced emission of 1-­methyl-­1,2,3,4,5-­pentaphenylsilole.
Chem. Commun.: 1740–1741. https://doi.org/10.1039/B105159H
5 Mei, J., Leung, N. L., Kwok, R. T., Lam, J. W., and Tang, B. Z. (2015). Aggregation-­induced
emission: together we shine, united we soar! Chem. Rev. 115: 11718–11940. https://doi.
org/10.1021/acs.chemrev.5b00263
6 Guo, J., Zhao, Z., and Tang, B. Z. (2018). Purely organic materials with aggregation-­induced
delayed fluorescence for efficient nondoped OLEDs. Adv. Optical Mater. 6 (15): 1800264. https://
doi.org/10.1002/adom.201800264
7 Chen, B., Liu, B., Zeng, J., Nie, H., Xiong, Y., Zou, J., Ning, H., Wang, Z., Zhao, Z., and
Tang, B. Z. (2018). Efficient bipolar blue AIEgens for high-­performance nondoped blue
OLEDs and hybrid white OLEDs. Adv. Funct. Mater. 28: 1803369. https://doi.org/10.1002/
adfm.201803369
8 Liu, Y., Mu, C., Jiang, K., Zhao, J., Li, Y., Zhang, L., Li, Z., Lai, J. Y. L., Hu, H., Ma, T., Hu, R., Yu,
D., Huang, X., Tang, B. Z., and Yan, H. (2015). A tetraphenylethylene core-­based 3D structure
small molecular acceptor enabling efficient non-­fullerene organic solar cells. Adv. Mater. 27:
1015–1020. https://doi.org/10.1002/adma.201404152
9 Nisi, F. D., Francischello, R., Battisti, A., Panniello, A., Fanizza, E., Striccoli, M., Gu, X., Leung, N.,
Tang, B. Z., and Pucci, A. (2017). Red-­emitting AIEgen for luminescent solar concentrators. Mater.
Chem. Front. 1: 1406–1412. https://doi.org/10.1039/C7QM00008A
10 Huang, G., Jiang, Y., Yang, S., Li, B. S., and Tang, B. Z. (2019). Multistimuli response and
polymorphism of a novel tetraphenylethylene derivative. Adv. Funct. Mater. 29 (16): 1900516.
https://doi.org/10.1002/adfm.201900516
11 Li, J. A., Zhou, J., Mao, Z., Xie, Z., Yang, Z., Xu, B., Liu, C., Chen, X., Ren, D., Pan, H., Shi, G.,
Zhang, Y., and Chi, Z. (2018). Transient and persistent room-­temperature mechanoluminescence
from a white-­light-­emitting AIEgen with tricolor emission switching triggered by light. Angew.
Chem. 130: 6559–6563. https://doi.org/10.1002/ange.201800762
 ­Reference 603

12 Hu, F., Mao, D., Cai, X., Wu, W., Kong, D., and Liu, B. (2018). A light-­up probe with
aggregation-­induced emission for real-­time bio-­orthogonal tumor labeling and image-­
guided photodynamic therapy. Angew. Chem., Int. Ed. 57 (32): 10182–10186. https://doi.
org/10.1002/anie.201805446
13 Situ, B., Gao, M., He, X., Li, S., He, B., Guo, F., Kang, C., Liu, S., Yang, L., Jiang, M., Hu, Y. Tang,
B. Z., and Zheng, L. (2019). A two-­photon AIEgen for simultaneous dual-­color imaging of
atherosclerotic plaques. Mater. Horiz. 6: 546–553. https://doi.org/10.1039/C8MH01293H
14 Ostroverkhova, O. (2016). Organic optoelectronic materials: mechanisms and applications. Chem.
Rev. 116 (22): 13279–13412. https://doi.org/10.1021/acs.chemrev.6b00127
15 Hestand, N. J., and Spano, F. C. (2017). Molecular aggregate photophysics beyond the Kasha
model: novel design principles for organic materials. Acc. Chem. Res. 50: 341–350. https://doi.
org/10.1021/acs.accounts.6b00576
16 Kasha, M. (1963). Energy transfer mechanisms and the molecular exciton model for molecular
aggregates. Radiat. Res. 20 (1): 55–70. https://www.jstor.org/stable/3571331
17 Ren, Y., Lam, J. W. Y., Dong, Y., Tang, B. Z., and Wong, K. S. (2005). Enhanced emission efficiency
and excited state lifetime due to restricted intramolecular motion in silole aggregates. J. Phys.
Chem. B 109 (3): 1135–1140. https://doi.org/10.1021/jp046659z
18 An, B.-­K., Know, S.-­K., Jung, S.-­D., and Park, S. Y. (2002). Enhanced emission and its switching in
fluorescent organic nanoparticles. J. Am. Chem. Soc. 124: 14410–-­14415. https://doi.org/10.1021/
ja0269082
19 Cariati, E., Lanzeni, V., Tordin, E., Ugo, R., Botta, C., Giacometti Schieroni, A., Sironi, A., and
Pasini, D. (2011). Efficient crystallization induced emissive materials based on a simple push–pull
molecular structure. Phys. Chem. Chem. Phys. 13: 18005–-­18014. https://doi.org/10.1039/
C1CP22267H
20 Xie, Z., Yang, B., Li, F., Cheng, G., Liu, L., Yang, G., Xu, H., Ye, L., Hanif, M., Liu, S., Ma, D., and
Ma, Y. (2005). Cross Dipole Stacking in the Crystal of Distyrylbenzene Derivative: The Approach
toward high solid-­state luminescence efficiency. J. Am. Chem. Soc. 127 (41): 14152–14153. https://
doi.org/10.1021/ja054661d
21 Lucenti, E., Forni, A., Botta, C., Carlucci, L., Giannini, C., Marinotto, D., Previtali, A., Righetto, S.,
and Cariati, E. (2017). H-­Aggregates granting crystallization-­induced emissive behavior and
ultralong phosphorescence from a pure organic molecule. J. Phys. Chem. Lett. 8: 1894−1898.
https://doi.org/10.1021/acs.jpclett.7b00503
22 Liu, H., Yao, L., Li, B., Chen, X., Gao, Y., Zhang, S., Li, W., Lu, P., Yang, B., and Ma, Y. (2016).
Excimer-­induced high-­efficiency fluorescence due to pairwise anthracene stacking in a
crystal with long lifetime. Chem. Commun. 52: 7356–7359. https://doi.org/10.1039/
C6CC01993E
23 Shen, Y., Zhang, Z., Liu, H., Yan, Y., Zhang, S., Yang, B., and Ma, Y. (2019). Highly efficient
orange-­red/red excimer fluorescence from dimeric π–π stacking of perylene and its nanoparticle
applications. J. Phys. Chem. C 123: 13047−13056. https://doi.org/10.1021/acs.jpcc.9b02447
24 Mróz, M. M., Benedini, S., Forni, A., Botta, C., Pasini, D., Cariati, E., and Virgili, T. (2016).
Long-­living optical gain induced by solvent viscosity in a push–pull molecule. Phys. Chem. Chem.
Phys. 18: 18289–18296. https://doi.org/10.1039/C6CP02988D
25 Tanaka, H., Shizu, K., Nakanotani, H., and Adachi, C. (2014). Dual intramolecular charge-­transfer
fluorescence derived from a phenothiazine-­triphenyltriazine derivative. J. Phys. Chem. C 118 (29):
15985−15994. https://doi.org/10.1021/jp501017f
26 Bureš, F. (2014). Fundamental aspects of property tuning in push–pull molecules. RCS Adv. 4:
58826–58851. https://doi.org/10.1039/C4RA11264D
604 20  Push–Pull AIEgens

27 Zhao, Z., Lam, J. W. Y., and Tang, B. Z. (2012). Tetraphenylethene: a versatile AIE building block
for the construction of efficient luminescent materials for organic light-­emitting diodes. J. Mater.
Chem. 22: 23726–23740. https://doi.org/10.1039/C2JM31949G
28 Gu, X., Yao, J., Zhang, G., Zhang, C., Yan, Y., Zhao, Y., and Zhang, D. (2013). New electron-­donor/
acceptor-­substituted tetraphenylethylenes: aggregation-­induced emission with tunable emission
color and optical-­waveguide behavior. Chem. Asian J. 8: 2362–2369. https://doi.org/10.1002/
asia.201300451
29 Yuan, Y., Zhang, C. J., Gao, M., Zhang, R., Tang, B. Z., and Liu, B. (2015). Specific light-­up bioprobe
with aggregation-­induced emission and activatable photoactivity for the targeted and image-­
guided photodynamic ablation of cancer cells. Angew. Chem. Int. Ed. 54: 1780–1786. https://doi.
org/10.1002/anie.201408476
30 Xu, S., Yuan, Y., Cai, X., Zhang, C.-­J., Hu, F., Liang, J., Zhang, G., Zhang, D., and Liu, B. (2015).
Tuning the singlet-­triplet energy gap: a unique approach to efficient photosensitizers with
aggregation-­induced emission (AIE) characteristics. Chem. Sci. 6: 5824–5830. https://doi.
org/10.1039/C5SC01733E
31 Wu, W., Mao, D., Xu, S., Ji, S., Hu, F., Ding, D., Kong, D., and Liu, B. (2017). High performance
photosensitizers with aggregation-­induced emission for image-­guided photodynamic anticancer
therapy. Mater. Horiz. 4: 1110–1114. https://doi.org/10.1039/C7MH00469A
32 Hu, F., Mao, D., Kenry, Wang, Y., Wu, W., Zhao, D., Kong, D., and Liu, B. (2018). Metal–organic
framework as a simple and general inert nanocarrier for photosensitizers to implement activatable
photodynamic therapy. Adv. Funct. Mater. 28: 1707519. https://doi.org/10.1002/adfm.201707519
33 Nitti, A., Villafiorita-­Monteleone, F., Pacini, A., Botta, C., Virgili, T., Forni, A., Cariati, E., Boiocchi,
M., and Pasini, D. (2017). Structure–activity relationship for the solid state emission of a new
family of “push–pull” π-­extended chromophores. Faraday Discuss. 196: 143–161. https://doi.
org/10.1039/C6FD00161K
34 Botta, C., Benedini, S., Carlucci, L., Forni, A., Marinotto, D., Nitti, A., Pasini, D., Righetto, S., and Cariati,
E. (2016). Polymorphism-­dependent aggregation induced emission of a push–pull dye and its multi-­
stimuli responsive behavior. J. Mater. Chem. C 4: 2979–2989. https://doi.org/10.1039/C5TC03352G
35 Cariati, E., Lanzeni, V., Tordin, E., Ugo, R., Botta, C., Giacometti-­Schieroni, A., Sironic, A., and
Pasini, D. (2011). Efficient crystallization induced emissive materials based on a simple push–pull
molecular structure. Phys. Chem. Chem. Phys. 13: 18005–18014.
36 Virgili, T. Forni, A., Cariati, E., Pasini, D., and Botta, C. (2013). Direct evidence of torsional motion
in an aggregation-­induced emissive chromophore. J. Phys. Chem. C 117: 27161−27166. https://doi.
org/10.1021/jp4104504
37 Chen, G., Li, W., Zhou, T., Peng, Q., Zhai, D., Li, H., Yuan, W. Z., Zhang, Y., and Tang, B. Z. (2015).
Conjugation-­induced rigidity in twisting molecules: filling the gap between aggregation-­caused
quenching and aggregation-­induced emission. Adv. Mater. 27: 4496. https://doi.org/10.1002/
adma.201501981
38 Ning, Z., Chen, Z., Zhang, Q., Yan, Y., Qian, S., Cao, Y., and Tian, H. (2007). Aggregation-­induced
emission (AIE)-­active starburst triarylamine fluorophores as potential non-­doped red emitters for
organic light-­emitting diodes and Cl2 gas chemodosimeter. Adv. Funct. Mater. 17: 3799–3807.
https://doi.org/10.1002/adfm.200700649
39 Zhao, X., Xue, P., Wang, K., Chen, P., Zhang, P., and Lu, R. (2014). Aggregation-­induced emission
of triphenylamine substituted cyanostyrene derivatives. New J. Chem. 38: 1045–1051. https://doi.
org/10.1039/C3NJ01343J
40 Wang, B., Wang, Y., Hua, J., Jiang, Y., Huang, J., Qian, S., and Tian, H. (2011). Starburst
triarylamine donor–acceptor–donor quadrupolar derivatives based on cyano-­substituted
 ­Reference 605

diphenylaminestyrylbenzene: tunable aggregation-­induced emission colors and large two-­photon


absorption cross sections. Chem. Eur. J. 17: 2647–2655. https://doi.org/10.1002/chem.201002821
41 Lu, H., Zheng, Y., Zhao, X., Wang, L., Ma, S., Han, X., Xu, B., Tian, W., and Gao, H. (2016). Highly
efficient far red/near-­infrared solid fluorophores: aggregation-­induced emission, intramolecular
charge transfer, twisted molecular conformation, and bioimaging applications. Angew. Chem. 128:
163–167. https://doi.org/10.1002/anie.201507031
42 Yeh, H. C., Yeh, S. J., and Chen, C. T. (2003). Readily synthesised arylamino fumaronitrile for
non-­doped red organic light-­emitting diodes. Chem. Commun.: 2632–2633. https://doi.org/10.1039/
B309780C
43 Palayangoda, S. S., Cai, X., Adhikari, R. M., and Neckers, D. C. (2008). Carbazole-­based donor−
acceptor compounds: highly fluorescent organic nanoparticles. Org. Lett. 10: 281–284. https://doi.
org/10.1021/ol702666g
44 Qin, A. Lam, J. W. Y., Mahtab, F., Jim, C. K. W., Tang, L., Sun, J., Sung, H. H. Y., Williams, I. D., and
Tang, B. Z. (2009). Pyrazine luminogens with “free” and “locked” phenyl rings: understanding of
restriction of intramolecular rotation as a cause for aggregation-­induced emission. Appl. Phys. Lett.
94: 253308. https://doi.org/10.1063/1.3137166
45 Wang, Y., He, Z., Chen, G., Shan, T., Yuan, W., Lu, P., and Zhang, Y. D-­A structured high efficiency
solid luminogens with tunable emissions: molecular design and photophysical properties. Chin.
Chem. Lett. 28: 2133–2138. https://doi.org/10.1016/j.cclet.2017.09.054
46 Park, I. S., Lee, S. Y., Adachi, C., and Yasuda, T. (2016). Full-­color delayed fluorescence materials
based on wedge-­shaped phthalonitriles and dicyanopyrazines: systematic design, tunable
photophysical properties, and OLED performance. Adv. Funct. Mater. 26: 1813–1821. https://doi.
org/10.1002/adfm.201505106
47 Chen, S., Xu, X., Liu, Y., Yu, G., Sun, X., Qiu, W., Ma, Y., and Zhu, D. (2005). Synthesis and
characterization of n-­type materials for non-­doped organic red-­light-­emitting diodes. Adv. Funct.
Mater. 15: 1541–1546. https://doi.org/10.1002/adfm.200500105
48 Wang, S., Yan, X., Cheng, Z., Zhang, H., Liu, Y., and Wang, Y. (2015). Highly efficient near-­
infrared delayed fluorescence organic light emitting diodes using a phenanthrene-­based charge-­
transfer compound. Angew. Chem. Int. Ed. 54: 13068–13072. https://doi.org/10.1002/
anie.201506687
49 Yuan, Y., Hu, Y., Zhang, Y.-­X., Lin, J.-­D., Wang, Y.-­K., Jiang, Z.-­Q., Liao, L.-­S., and Lee,
S.-­T. (2017). Over 10% EQE near-­infrared electroluminescence based on a thermally activated
delayed fluorescence emitter. Adv. Funct. Mater. 27: 1700986. https://doi.org/10.1002/
adfm.201700986
50 Tsujimoto, H., Ha, D.-­G., Markopoulos, G., Chae, H. S., Baldo, M. A., and Swager, T. M. (2017).
Thermally activated delayed fluorescence and aggregation induced emission with through-­space
charge transfer. J. Am. Chem. Soc. 139: 4894−4900. https://doi.org/10.1021/jacs.7b00873
51 Selected examples: (a) Nitti, A., Osw, P., Abdullah, M. N., Galbiati, A., and Pasini, D. (2018).
Scalable synthesis of naphthothiophene-­based D-­π-­D extended oligomers through cascade direct
arylation processes. Synlett 29: 2577–2581. https://doi.org/10.1055/s-­0037-­1610331 (b) Osw, P.,
Nitti, A., Abdullah, M. N., Etkind, S. I., Mwaura, J., Galbiati, A., and Pasini, D. (2020). Synthesis
and evaluation of scalable D-­A-­D π-­extended oligomers as p-­type organic materials for bulk-­
heterojunction solar cells. Polymers 12: 720. https://doi.org/10.3390/polym12030720
52 Zhao, Z., Deng, C., Chen, S., Lam, J. W. Y., Qin, W., Lu, P., Wang, Z., Kwok, H. S., Ma, Y., Qiu, Y.,
and Tang, B. Z. (2011). Full emission color tuning in luminogens constructed from
tetraphenylethene, benzo-­2,1,3-­thiadiazole and thiophene building blocks. Chem. Commun. 47:
8847–8849. https://doi.org/10.1039/C1CC12775F
606 20  Push–Pull AIEgens

53 Ishi-­i, T., Ikeda, K., Kichise, Y., and Ogawa, M. (2012). Red-­light-­emitting system based on
aggregation of donor–acceptor derivatives in polar aqueous media. Chem. Asian J. 7: 1553–1557.
https://doi.org/10.1002/asia.201200136
54 Qin, W. Li, K., Feng, G., Li, M., Yang, Z., Liu, B., and Tang, B. Z. (2014). Bright and photostable
organic fluorescent dots with aggregation-­induced emission characteristics for noninvasive
long-­term cell imaging. Adv. Funct. Mater. 24: 635–643. https://doi.org/10.1002/adfm.201302114
55 Lee, W. W. H., Zhao, Z., Cai, Y., Xu, Z., Yu, Y., Xiong, Y., Kwok, R. T. K., Chen, Y., Leung, N. L. C.,
Ma, D., Lam, J. W. Y., Qin, A., and Tang, B. Z. (2018). Facile access to deep red/near-­infrared
emissive AIEgens for efficient non-­doped OLEDs. Chem. Sci. 9: 6118–6125. https://doi.
org/10.1039/C8SC01377B
56 Jiang, T., Qu, Y., Li, B., Gao, Y., and Hua, J. (2015). Tetraphenylethene end-­capped
[1,2,5]thiadiazolo[3,4-­c]pyridine with aggregation-­induced emission and large two-­photon
absorption cross-­sections. RSC Adv. 5: 1500–1506. https://doi.org/10.1039/C4RA09789K
57 Ni, J.-­S., Zhang, P., Jiang, T., Chen, Y., Su, H., Wang, D., Yu, Z.-­Q., Kwok, R. T. K., Zhao, Z., Lam,
J. W. Y., and Tang, B. Z. (2018). Red/NIR-­emissive benzo[d]imidazole-­cored AIEgens: facile
molecular design for wavelength extending and in vivo tumor metabolic imaging. Adv. Mater. 30,
1805220. https://doi.org/10.1002/adma.201805220
58 Selected examples: (a) Ono, K., Tanaka, S., and Yamashita, Y. (1994). Benzobis(thiadiazole)s
containing hypervalent sulfur atoms: novel heterocycles with high electron affinity and short
intermolecular contacts between heteroatoms. Angew. Chem., Int. Ed. 33: 1977–1979. https://doi.
org/10.1002/anie.199419771.(b) Karikomi, M., Kitamura, C., Tanaka, S., and Yamashita, Y. (1995).
New narrow-­bandgap polymer composed of benzobis(1,2,5-­thiadiazole) and thiophenes. J. Am.
Chem. Soc. 117: 6791–6792. https://doi.org/10.1021/ja00130a024.(c) Kitamura, C., Tanaka, S., and
Yamashita, Y. (1996). Design of narrow-­bandgap polymers. syntheses and properties of monomers
and polymers containing aromatic-­donor and o-­quinoid-­acceptor units. Chem. Mater. 8: 570–578.
https://doi.org/10.1021/cm950467m
59 Du, X., Qi, J., Zhang, Z., Ma, D., and Wang, Z. (2012). Efficient non-­doped near infrared organic
light-­emitting devices based on fluorophores with aggregation-­induced emission enhancement.
Chem. Mater. 24: 2178–2185. https://doi.org/10.1021/cm3008733
60 Qian, G., Dai, B., Luo, M., Yu, D., Zhan, J., Zhang, Z., Ma, D., and Wang, Z. Y. (2008). Band gap
tunable, donor−acceptor−donor charge-­transfer heteroquinoid-­based chromophores: near
infrared photoluminescence and electroluminescence. Chem. Mater. 20: 6208–6216. https://doi.
org/10.1021/cm801911n
61 Kiyose, K., Aizawa, S., Sasaki, E., Kojima, H., Hanaoka, K., Terai, T., Urano, Y., and Nagano,
T. (2009). Molecular design strategies for near-­infrared ratiometric fluorescent probes based on the
unique spectral properties of aminocyanines. Chem. – Eur. J. 15: 9191–9200.
62 Chen, S., Liu, J., Liu, Y., Su, H., Hong, Y., Jim, C. K. W., Kwok, R. T. K., Zhao, N., Qin, W., Lam,
J. W. Y., Wong, K. S., and Tang, B. Z. (2012). An AIE-­active hemicyanine fluorogen with stimuli-­
responsive red/blue emission: extending the pH sensing range by “switch + knob” effect. Chem.
Sci. 3: 1804–1809. https://doi.org/10.1039/C2SC01108E
63 Zhao, N., Yang, Z., Lam, J. W. Y., Sung, H. H. Y., Xie, N., Chen, S., Su, H., Gao, M., Williams, I. D.,
Wong, K. S., and Tang, B. Z. (2012). Benzothiazolium-­functionalized tetraphenylethene: an AIE
luminogen with tunable solid-­state emission. Chem. Commun. 48: 8637–8639. https://doi.
org/10.1039/C2CC33780K
64 Jayaram, D. T., Ramos-­Romero, S., Shankar, B. H., Garrido, C., Rubio, N., Sanchez-­Cid, L., Gòmez,
S. B., Blanco, J., and Ramaiah, D. (2016). in vitro and in vivo demonstration of photodynamic
 ­Reference 607

activity and cytoplasm imaging through TPE nanoparticles. ACS Chem. Biol. 11: 104–112. https://
doi.org/10.1021/acschembio.5b00537
65 Qin, W., Ding, D., Liu, J., Yuan, W. Z., Hu, Y., Liu, B., and Tang, B. Z. (2012). Biocompatible
nanoparticles with aggregation-­induced emission characteristics as far-­red/near-­infrared
fluorescent bioprobes for in vitro and in vivo imaging applications. Adv. Funct. Mater. 22: 771–779.
https://doi.org/10.1002/adfm.201102191
66 Shi, C., Guo, Z., Yan, Y., Zhu, S., Xie, Y., Zhao, Y. S., Zhu, W., and Tian, H. (2013). Self-­assembly
solid-­state enhanced red emission of quinolinemalononitrile: optical waveguides and stimuli
response. ACS Appl. Mater. Interfaces 5: 192−198. https://doi.org/10.1021/am302466m
67 Shao, A., Guo, Z., Zhu, S., Zhu, S., Shi, P., Tian, H., and Zhu, W.-­H. (2014). Insight into aggregation-­
induced emission characteristics of red-­emissive quinoline-­malononitrile by cell tracking and
real-­time trypsin detection. Chem. Sci. 5: 1383–1389. https://doi.org/10.1039/C3SC52783B
68 Shao, A., Xie, Y., Zhu, S., Guo, Z., Zhu, S., Guo, J., Shi, P., James, T. D., Tian, H., and Zhu,
W.-­H. (2015). Far-­red and near-­IR AIE-­active fluorescent organic nanoprobes with enhanced
tumor-­targeting efficacy: shape-­specific effects. Angew. Chem. Int. Ed. 54: 7275–7280. https://doi.
org/10.1002/anie.201501478
69 Wu, Y., Jin, P., Gu, K., Shi, C., Guo, Z., Yu, Z.-­Q., and Zhu, W.-­H. (2019). Broadening AIEgen
application: rapid and portable sensing of foodstuff hazards in deep-­frying oil. Chem. Commun. 55:
4087–4090. https://doi.org/10.1039/C9CC01172B
70 Shen, X. Y., Wang, Y. J., Zhang, H., Qin, A., Sun, J. Z., and Tang, B. Z. (2014). Conjugates of
tetraphenylethene and diketopyrrolopyrrole: tuning the emission properties with phenyl bridges.
Chem. Commun. 50: 8747–8750. https://doi.org/10.1039/C4CC03024A
71 Wang, B., He, N., Li, B., Jiang, S., Qu, Y., Qu, S., and Hua, J. (2012). Aggregation-­induced emission
and large two-­photon absorption cross-­sections of diketopyrrolopyrrole (DPP) derivatives. Aust.
J. Chem. 65: 387–394. https://doi.org/10.1071/CH11410
72 Jin, Y., Xu, Y., Liu, Y., Wang, L., Jiang, H., Li, X., and Cao, D. (2011). Synthesis of novel
diketopyrrolopyrrole-­based luminophores showing crystallization-­induced emission enhancement
properties. Dyes Pigm. 90: 311–318. https://doi.org/10.1016/j.dyepig.2011.01.005
73 Wang, L., Yang, L., and Cao, D. (2015). Probes based on diketopyrrolopyrrole and anthracenone
conjugates with aggregation-­induced emission characteristics for pH and BSA sensing. Sens.
Actuators B 221: 155–166. https://doi.org/10.1016/j.snb.2015.06.074
74 Li, Y., Zhang, G., Zhang, W., Wang, J., Chen, X., Liu, Z., Yan, Y., Zhao, Y., and Zhang, D. (2014).
Arylacetylene-­substituted naphthalene diimides with dual functions: optical waveguides and
n-­type semiconductors. Chem. Asian J. 9: 3207–3214. https://doi.org/10.1002/asia.201402768
75 Choudhury, P., Sarkar, S., and Das, P. K. (2018). Tunable aggregation-­induced multicolor emission
of organic nanoparticles by varying the substituent in naphthalene diimide. Langmuir 34:
14328−14341. https://doi.org/10.1021/acs.langmuir.8b02996
76 Liu, H., Yao, L., Li, B., X. Chen, Gao, Y., Zhang, S., Li, W., Lu, P., Yang, B., and Ma, Y. (2016).
Excimer-­induced high-­efficiency fluorescence due to pairwise anthracene stacking in a crystal
with long lifetime. Chem. Commun. 52: 7356–7359. https://doi.org/10.1039/C6CC01993E
77 Würthner, F., Kaiser, T. E., and Saha-­Möller, C. R. (2011). J-­aggregates: from serendipitous
discovery to supramolecular engineering of functional dye materials. Angew. Chem. Int. Ed. 50:
3376–3410. https://doi.org/10.1002/anie.201002307
78 Weil, T., Vosh, T., Hofkens, J., Peneva, K., and Müller, K. (2010). The rylene colorant family:
tailored nanoemitters for photonics research and applications. Angew. Chem. Int. Ed. 49:
9068−9093. https://doi.org/10.1002/anie.200902532
608 20  Push–Pull AIEgens

79 Zhan, X., Facchetti, A., Barlow, S., Marks, T. J., Ratner, M. A., Wasielewski, M. R., and Marder,
S. R. (2011). Rylene and related diimides for organic electronics. Adv. Mater. 23: 268−284. https://
doi.org/10.1002/adma.201001402
80 Zhao, Q., Zhang, X. A., Qiang, W., Wang, J., Shen, X. Y., Qin, A., Sun, J. Z., and Tang, B. Z. (2012).
Tetraphenylethene modified perylene bisimide: effect of the number of substituents on AIE
performance. Chem. Commun. 48: 11671−11673. https://doi.org/10.1039/C2CC36060H
81 Zhao, Q., Zhang, S., Liu, Y., Mei, J., Chen, S., Lu, P., Qin, A., Ma, Y., Sun, J. Z., and Tang,
B. Z. (2012). Tetraphenylethenyl-­modified perylene bisimide: aggregation-­induced red emission,
electrochemical properties and ordered microstructures. J. Mater. Chem. 22: 7387−7394. https://
doi.org/10.1039/C2JM16613E
82 Zhao, Q., Li, K., Chen, S., Qin, A., Ding, D., Zhang, S., Liu, Y., Liu, B., Sun, J. Z., and Tang,
B. Z. (2012). Aggregation-­induced red-­NIR emission organic nanoparticles as effective and
photostable fluorescent probes for bioimaging. J. Mater. Chem. 22: 15128–15135. https://doi.
org/10.1039/C2JM31368E
609

Index

a Bond length alternation (BLA)  179


Acceptor (A)  274, 575 Boron‐difluorohydrazone (BODIHY)  16, 419
Activation‐alkynylation‐cyclocondensation Bovine serum albumin (BSA)  335, 466, 511
(AACC) sequence  458
Aggregation‐caused quenching (ACQ)  27, 55, c
99, 153, 177, 221, 273, 311, 333, 456, 487, 509, Calamitic LC  556
531, 555, 575 Calf thymus DNA (CT‐DNA)  74
Aggregation‐enhanced emission (AEE)  333, Cation‐exchange capacity (CEC)  203
491, 532 Cetyltrimethylammonium bromide (CTAB)  469
Aggregation‐induced delayed fluorescence Charge‐separated (CS)  302
(AIDF)  222, 314, 576, 589 Charge transfer (CT)  14, 92, 194, 274, 374, 416
Aggregation‐induced emission (AIE)  1, 27, 56, Chondroitin (Chs)  258
99, 153, 177, 204, 221, 251, 273, 311, 333, 411, Circular dichroism (CD)  69
456, 487, 509, 531, 555, 575 Circularly polarized luminescence (CPL)  69,
Aggregation‐induced emission enhancement 540, 561
(AIEE)  333, 456 Clusterization‐triggered emission (CTE)  17, 390
AIE luminogens (AIEgens)  177, 487, 509, 531 Clusteroluminescence  166
Algebraic diagrammatic construction Commission Internationale de l’Éclairage
(ADC)  419 (CIE)  142
Amyloid‐β (Aβ)  515 Complete active space second‐order perturbation
Anisotropy of the induced current density (CASPT2)  417
(ACID)  4 Complete active space self‐consistent field
(9‐anthrylmethyl) bis(2‐pyridylmethyl) amine (CASSCF)  417
(APA)  14 Compound cyanoacetic acid (CAA)  142
Anti‐Kasha emission  312 Concentration quenching (CQ)  55
Arene–perfluoroarene (AP)  274 Conduction band (CB)  404
Atomic force microscopy (AFM)  488 Confocal laser scanning microscopy
Au(I)–thiolate  166 (CLSM)  405, 546
Average lifetime ( τ )  290 Conical intersections (CIs)  12, 78, 412
Conjugated microporous polymer
b (CMP)  161, 264
Benzophenone (BP)  375, 436 Coupled cluster n (CCn) methods  419
Benzylidene methyloxazolone (BMO)  92 Covalent organic frameworks (COFs)  402

Handbook of Aggregation-Induced Emission: Volume 1 Tutorial Lectures and Mechanism Studies, First Edition.
Edited by Youhong Tang and Ben Zhong Tang.
© 2022 John Wiley & Sons Ltd. Published 2022 by John Wiley & Sons Ltd.
610 Index

Crystal‐crosslinked gel (CCgel)  262 Duschinsky rotation effect (DRE)  9, 29


Crystallization‐induced dual emission Duschinsky rotation matrix (DRM)  11
(CIDE)  376 Dye‐sensitized solar cells (DSSCs)  455
Crystallization‐induced emission enhancement Dynamic light scattering (DLS)  340, 469
(CIEE)  177
Crystallization‐induced phosphorescence e
(CIP)  375 Electric dipole transition moment (μgn)  569
Cu‐catalyzed alkyne‐azide cycloadditon Electroluminescence (EL)  222, 486
(CuAAC)  461 Electron acceptor (A)  131
Cucurbit[7]uril (CB7)  256 Electron donor (D)  131
Cucurbit[8]uril (CB[8])  533 Electron mobility (μe)  497
Cu nanocluster (CuNC)  257 Electron‐rich atoms  275
1‐cyano‐1,2‐bis(4’‐methylbiphenyl)ethylene Electron‐vibration coupling (EVC)  2
(CN‐MBE)  56 El‐Sayed rules  374
Cyano‐naphthyridine (CND)  239 Emission wavelength (λF, λP, λem or λDF)  578
Cyclic voltammetry (CV)  488 Emitting layer (EML)  225
Cyclization‐induced emission (CIE)  253, 412 Energy transfer (ET)  283
Cyclodextrin (CD)  256, 259, 386 Energy‐transfer rate constant (kET)  290
Cyclooctatetrathiophene (COTh)  4, 35 Enhanced permeability and retention
Cyclooctotetraene (COT)  35 (EPR)  519
ET efficiency (EET)  289
d Ethylenic bond rotation (φ)  86
Decafluorobiphenyl (DFBP)  375 Excimer‐induced enhanced emission (EIEE) 
Degree of charge transfer  278 576, 579
Delayed fluorescence (DF)  222, 374, 576 Excitation wavelength‐dependent emission
Density function theory (DFT)  6, 29, 283 (EWDM)  156
Desoxyribonucleic acid (DNA)  455 Excited single state (S1)  133
Detection limit  258 Excited‐state absorption (ESA)  72
Dexter energy transfer (DET)  222, 578 Excited‐state intramolecular proton transfer
Dibenzofulvene (DBF)  183 (ESIPT)  78, 431
2,3‐Dicyano‐5,6‐diphenylpyrazine (DCDPP)  37 Excited‐state through‐space aromatic dimers
Dicyanomethylenated acridones (DCNAC)  186 (ESTSAD)  162
Dicyanopentafulvene (DCF)  353 Excited‐state through‐space complex (ESTSC) 
Differential scanning calorimetry (DSC)  21, 162
277, 560 Excited triplet state (T*)  131
Dimethylformamide (DMF)  322 External quantum efficiency (EQE, ηext) 
Dimethyl tetraphenylsilole (DMTPS)  12 222, 223
Diphenyl dibenzofulvene (DPDBF)  89, 420 E/Z isomerization (EZI)  64, 438
Discotic LC  556 E–Z photoisomerization  79
9,10‐Distyrylanthracene (DSA)  37
Ditolyldibenzofulvene (DTDBF)  102 f
Donor (D)  274, 575, 579 Femtosecond transient absorption (fs‐TA)  21
Donor‐acceptor (D‐A)  222, 241, 311, 575, 579 Fermi’s golden rule (FGR)  27, 373, 411
Donor–Acceptor–Donor’(D‐A‐D’)  222, 236, 379 Field‐effect transistors (FETs)  486
Donor–acceptor–donor (D–A–D)  340, 467, 579 Field emission gun scanning electron
Donor‐π‐acceptor (D‐π‐A)  466 microscopy (FEG‐SEM)  472
Double‐bond torsion  424 First excited state (S1)  31, 79
Index 611

First‐principle calculations  29 i
Fish sperm DNA (FS‐DNA)  74 Indium tin oxide (ITO)  488
Flash nanoprecipitation (FNP)  523 Induced photon upconversion (iPUC)  87
Fluorescence lifetime imaging microscopy Infrared (IR)  274
(FLIM)  471 Intermolecular charge transfer (ICT)  575, 579
Fluorescence resonance energy transfer (FRET)  Intermolecular excitonic coupling (IEC)  45, 116
323, 561 Intermolecular interactions  115, 116, 128, 158,
Förster resonance energy transfer (FRET)  285, 182, 278
351, 578 Internal conversion (IC)  28, 72, 224, 314,
Fourier transform infrared (FTIR)  277 371, 578
Franck–Condon approximation  38 Internal conversion rate (kIC)  224
Franck–Condon (FC) configuration  72 Internal quantum efficiency (IQE)  222
Franck–Condon factor  213 Intersystem crossing (ISC)  28, 87, 116, 224,
Franck–Condon region  48, 412 273, 311, 371, 411, 536, 578
Free energy (FE)  86 Intersystem crossing rate (kISC)  224
Intramolecular McMurry coupling reaction  76
g Ionic self‐assembly (ISA)  557
General Amber force field (GAFF)  29, 422 Isotope effect (IE)  40
Glass‐transition temperature (Tg)  260
Glyoxylation‐alkynylation‐cyclocondensation j
(GACC)  458 J‐aggregates  164, 191, 206, 333, 437, 489, 579
Glyoxylation‐Stephens‐Castro coupling  458
Grazing‐incidence X‐ray diffraction k
(GI‐XRD)  495 Kasha’s rule  311, 414
Ground state (S0)  31, 79 Kohn–Sham frontier molecular orbitals  464

h l
H‐aggregates  164, 191, 437, 579 Ligand‐to‐metal charge transfer (LMCT)  167
H‐aggregation  60, 131, 313 Light‐emitting diode (LED)  55
Halogen‐to‐ligand charge transfer (XLCT)  35 Limit of detection (LOD)  264
Herzberg–Teller (HT) effect  22 Linear interpolation in internal coordinates
Hetero‐aggregation‐induced Tunable Emission (LIIC)  13
(HAITE)  273 Linear polyethyleneimine (LPEI)  156
Hexaphenylbenzene (HPB)  242 Linear range  258
Hexaphenylsilole (HPS)  2, 56, 99, 177, Lippert–Mataga equation  468
334, 556 Liquid crystal display (LCD)  568
High‐density Data Storage  197 Liquid crystals (LCs)  555
Highest occupied molecular orbital (HOMO)  Local‐excited (LE)  14, 314, 355, 474
182, 222, 230, 274, 374, 462, 492, 575, 581 Lowest unoccupied molecular orbital (LUMO) 
High‐pressure liquid chromatography 182, 222, 230, 265, 274, 374, 462, 492, 575, 581
(HPLC)  69 Luminescent solar concentrators (LSCs)  576
Host–Guest System  385 Lyotropic LC  556
Huang–Rhys factors  442
Hydrogen‐bonded organic aromatic frameworks m
(HOAFs)  402 Magnetic dipole transition moment (mgn)  569
Hydrogen bonding acceptor (HBA)  265 Magnetic resonance imaging (MRI)  527
Hydrogen bonds  142 Marcus–Levich–Jortner theory  317
612 Index

Maxima current efficiency (ηC)  223 o


Maximum emission wavelength (λem)  156 Organic field‐effect transistors
Mechanochromism (MC)  100, 197 (OFETs)  455, 485
Mechanoluminescence (ML)  100 Organic light‐emitting diodes (OLEDs)  177,
Melting temperature (Tm)  257 196, 221, 371, 455, 485, 555, 575
Membrane‐associated glycoprotein mucin 1 Organic photovoltaics (OPVs)  576
(MUC1)  257 Organic semiconductors (OSCs)  485
Metal–ligand/ligand–ligand charge‐transfer
state  442 p
Metal–organic framework (MOF)  262, 401 Parrinello–Rahman barostat  30
Metal‐to‐ligand charge transfer (MLCT)  35, 167 Periodic charge embedding (PCE) methods  422
Methyl 4‐bromobenzoate (MBB)  375 Persistent RTP (p‐RTP)  371
Methylcyclohexane (MCH)  312 Phenothiazine (PTZ)  586
Methylene blue (MB)  206 Phosphorescence efficiency (Φp)  372
Microwire (MW)  295 Phosphorescence lifetime (τp)  372
Minimum energy conical intersection Phosphorescent organic light‐emitting diodes
(MECI)  251 (PhOLEDs)  102
Minimum energy path (MEP)  7, 81 Photoacustic (PA)  527
Minimum‐energy point of the conical Photodynamic therapy (PDT)  536
intersection (MECI)  86 Photoinduced electron transfer (PET)  59, 541
Mixed quantum classical (MQC) approach  420 Photoluminescence (PL)  100, 222, 531
Molecular dynamics (MD) simulations  29 Photoluminescence quantum yield (PLQY) 
Molecular mechanics scheme (MM)  421 102, 287, 575, 578
Molecular Uniting Set Identified Characteristic Picric acid (PA)  362, 543
(MUSIC)  100 π‐conjugation  31
Multicomponent reactions (MCRs)  455 π–π stacking  48, 60, 101, 179, 259, 273, 333,
Multireference configuration interaction 395, 487, 489, 556, 575, 579
(MRCI)  13, 420 Polarizable continuum model (PCM)  29
Polarized optical microscopy (POM)  560
n Poly(acrylic acid) (PAA)  262
Nanoclusters (NCs)  166 Poly(dimethyl siloxane) (PDMS)  262
Nanoparticle (NP)  295, 509, 601 Poly(methyl methacrylate) (PMMA)  375
Nanorod (NR)  295 Poly(N‐hydroxysuccinimidyl methacrylate)
Naphthyridine (ND)  239 (PNHSMA)  19
Near‐infrared (NIR)  380, 513, 576 Poly(N‐isopropyl acrylamide) (PNIPAM)  396
N‐heterocyclic carbene (NHC)  255 Poly(propylene imine) (PPI)  156
Nonadiabatic coupling (NAC)  9 Poly(vinyl alcohol) (PVA)  262
Nonadiabatic electronic coupling (NAEC)  9 Polyacrylamide (PAM)  396
Nonlinear optical properties (NLO)  466 Polyacrylamide gel electrophoresis (PAGE)  166
Nonradiative decay rate constant (knr)  28, 587 Polyacrylonitrile  18
Nontraditional intrinsic luminescence Polycyclic aromatic hydrocarbons (PAH)  295
(NTIL)  154 Poly[(maleic anhydride)‐alt‐(vinyl acetate)]
Nuclear charge (Z) power  374 (PMV)  164
Nuclear magnetic resonance (NMR)  251 Polymer dispersity index (PDI)  532
Nuclear Overhauser effect spectroscopy Polymerization‐induced emission (PIE)  19, 164
(NOESY)  256 Polymer light‐emitting diodes (PLEDs)  575
Nucleus‐independent chemical shift (NICS)  4 Potential energy (PE)  81
Index 613

Potential energy surfaces (PES)  7, 28, 81, 411 Silacyclopentadines (Siloles)  31, 251
Powder X‐ray diffraction (PXRD)  107 Single crystal (SC)  487
Power efficiency (ηP)  223 Single‐stranded DNA (ssDNA)  515
Pressure‐induced emission enhancement Singlet fission (SF)  292, 579
(PIEE)  44 Singlet oxygen (1O2)  536
Pump‐induced absorption (PIA)  292 Singlet–triplet annihilation (STA)  224
SOC Hamiltonian (Hsoc)  374
q Solid‐state enhanced emission (SLEE)  437
Quantum‐classical (MQC) approach  420 Spin‐flip TD‐DFT (SF‐TD‐DFT)  419
Quantum efficiency (ΦF)  31, 411 Spin–orbit coupling (SOC)  16, 116, 300, 314,
Quantum mechanics and molecular mechanics 371, 416
(QM/MM)  9, 29, 143, 314 Spin‐vibronic coupling (SVC)  22
Quantum mechanics (QM) method  420 Steepest‐descent (SD) pathways  83
Quantum yield (QY, ΦF)  9, 68, 102, 158, 178, Stern–Volmer equation  291
222, 259, 493, 543 Stern–Volmer quenching constant (KSV)  291
Quinolone‐malononitrile (QM)  520 Stokes shift (κ)  37, 463
Structure resembling effect (SRE)  213
r Suppression of Kasha’s rule (SOKR)  16
Radiative decay rate (kF)  224, 587 Surface‐fixation induced emission (S‐FIE)  203
Radiative decay rate constant (kr)  28
Reactive oxygen species (ROS)  536 t
Reorganization energy (RE, λ)  9, 31 Tamm–Dancoff approximation (TDA) 
Resonance Raman spectroscopy (RRS)  38 79, 418
Restricted access to conical intersection (RACI)  Terephthalic acid (TPA)  33
78, 412 Tetra(ethylene glycol) (TEG)  260
Restriction of access to the dark state (RADS)  Tetrafurylethene (TFE)  414
15, 439 Tetrahydrofuran (THF)  14, 335
Restriction of intramolecular motion (RIM)  2, Tetrahydropyrimidines (THPs)  101
27, 57, 177, 221, 411, 509 1,2,3,4‐Tetraphenylbutadiene (TPBD)  9
Restriction of intramolecular rotation (RIR)  2, cis,cis‐1,2,3,4‐Tetraphenyl‐1,3‐butadiene
27, 57, 251, 333, 411 (TPBD)  37
Restriction of intramolecular vibration (RIV)  Tetraphenylethylene (TPE)  2, 56, 99, 178, 251,
2, 27, 57 334, 556, 582
Reverse intersystem crossing (RISC)  222, Thermally activated delayed fluorescence
297, 578 (TADF)  100, 297, 317, 380, 578
Reverse intersystem crossing (RISC) rate Thermal vibration correlation function (TVCF) 
(kRISC)  224, 374 8, 27
Reversible addition fragmentation chain transfer Thermotropic LC  556
(RAFT)  532 Thin film (TF)  489
Room‐temperature phosphorescence (RTP)  22, Thin‐layer chromatography (TLC)  375
100, 140, 300, 371, 527 Through‐bond ICT (TBICT)  580
Through‐space conjugation (TSC)  17, 155
s Through‐space ICT (TSICT)  580
Scanning electron microscope (SEM)  277 Time‐dependent density functional theory
Scanning tunneling microscope (STM)  277 (TD‐DFT)  29, 79, 312, 380, 417, 464, 536
Second NIR (NIR‐II)  527 Time‐dependent Kohn–Sham (TD‐KS)
Selected area diffraction (SAED)  277 approximation  420
614 Index

Time‐dependent tight‐binding density functional Ultraviolet (UV)  4, 64, 380, 568


theory (TD‐DFTB)  420 Ultraviolet‐visible (UV–Vis)  503
Trajectory surface hopping (TSH)  13, 420 United force field (UFF)  29
Transition Metal (TM)  442
Transmission electron microscopy (TEM)  277, 511 v
Triboluminescence (TL)  106 Valence band (VB)  404
Tricyano‐methylene‐pyridine (TCM)  513 Velocity rescaling thermostat  30
Triethylamine (TEA)  365 Vibration relaxation (VR)  372
2,4,6‐Trinitrotoluene (TNT)  543 Volatile Organic Compounds (VOCs)  196
Triphenylamine (TPA)  487, 584, 598
Triplet–triplet annihilation (TTA)  224, 404 w
Twisted intramolecular charge transfer (TICT)  White light emission (WLE)  283
355, 474, 580 White‐light OLEDs (WOLEDs)  576
World Health Organization (WHO)  541
u
Ultrafast pump–probe spectroscopy  584 x
Ultralong organic phosphorescence (UOP)  380 X‐ray diffraction (XRD)  277, 488, 560

You might also like