Solvent-Free Methods in Nanocatalysis

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 347

Solvent-Free Methods in Nanocatalysis

Solvent-Free Methods in Nanocatalysis

From Catalyst Design to Applications

Edited by Rafael Luque, Manoj B. Gawande, Esmail Doustkhah,


and Anandarup Goswami
Editors All books published by WILEY-VCH are carefully
produced. Nevertheless, authors, editors, and
Prof. Rafael Luque publisher do not warrant the information
Universidad de Córdoba contained in these books, including this book,
Departamento de Química Orgánica to be free of errors. Readers are advised to keep
Carretera Nacional IV-A, Km. 396 in mind that statements, data, illustrations,
Edificio C-3 procedural details or other items may
14014 Córdoba inadvertently be inaccurate.
Spain
Library of Congress Card No.: applied for
Prof. Manoj B. Gawande
Department of Industrial and British Library Cataloguing-in-Publication Data
Engineering Chemistry A catalogue record for this book is available
Institute of Chemical Technology from the British Library.
Mumbai
Marathwada Campus Bibliographic information published by
Aurangabad Road the Deutsche Nationalbibliothek
Jalna-431213, Maharashtra The Deutsche Nationalbibliothek lists
India this publication in the Deutsche
Nationalbibliografie; detailed bibliographic
Dr. Esmail Doustkhah data are available on the Internet at
Koç University Tüpraş Energy Center <http://dnb.d-nb.de>.
(KUTEM)
Koç University © 2023 WILEY-VCH GmbH, Boschstr. 12,
34450 Istanbul 69469 Weinheim, Germany
Turkey
All rights reserved (including those of
Prof. Anandarup Goswami translation into other languages). No part of
Vignan’s Foundation for Science this book may be reproduced in any form – by
Technology and Research (VFSTR) photoprinting, microfilm, or any other
Department of Chemistry, School of means – nor transmitted or translated into a
Applied Science and Humanities machine language without written permission
Vadlamudi from the publishers. Registered names,
522 213 Guntur trademarks, etc. used in this book, even when
India not specifically marked as such, are not to be
considered unprotected by law.

Print ISBN: 978-3-527-34874-9


ePDF ISBN: 978-3-527-83145-6
ePub ISBN: 978-3-527-83147-0
oBook ISBN: 978-3-527-83146-3

Cover Image Esmail Doustkhah


Cover Design FORMGEBER, Mannheim,
Germany

Typesetting Straive, Chennai, India


v

Contents

Preface xi

1 Introduction: Scope of the Book 1


Anil Kumar Nallajarla and Anandarup Goswami
1.1 Introduction: Green Chemistry, Solvent-free Synthesis, and
Nanocatalysts 1
1.2 Topics Covered in this Book 5
1.3 Solvent-Free Synthesis of Nanocatalysts 6
1.4 Solvent and Catalyst-Free Organic Transformations 8
1.5 Solvent-Free Reactions Using NCs 10
1.5.1 Different Metal Oxides as a Catalyst/Support in Solvent-Free
Reaction 12
1.5.1.1 Titanium Oxide 12
1.5.1.2 Tin Oxide 13
1.5.1.3 Manganese Oxide (MnOx ) 14
1.5.1.4 Zinc Oxide 14
1.5.1.5 Aluminum Oxide 15
1.5.1.6 Iron Oxide 15
1.5.2 Silica-Based Materials as Catalysts/Supports in Solvent-Free Organic
Reactions 15
1.5.3 Carbon-Based Materials as Catalysts/Supports in Solvent-Free Organic
Reactions 16
1.5.4 Nitride-Based Materials as Catalysts/Supports in Solvent-Free Organic
Reactions 17
1.5.5 Ionic Liquid-Based Materials as Catalysts/Supports in Solvent-Free
Organic Reactions 18
1.6 Present Status and Future Direction 20
References 22

2 Strategies for the Preparation of Nanocatalysts and Supports


Under Solvent-Free Conditions 31
Rahul P. Gaikwad, Arun D. Kute, and Manoj B. Gawande
2.1 Introduction 31
vi Contents

2.2 Mechanochemistry 32
2.2.1 Ball Milling 33
2.2.2 Mortar and Pestle 38
2.3 Thermal Treatment 44
2.3.1 Simple Thermal Treatment 44
2.3.2 Thermal Decomposition 50
2.3.3 Microwave Heating Energy 51
2.4 Plasma-Assisted Methods 52
2.4.1 Thermal Plasma Method 53
2.4.2 Cold Thermal Plasma Method 56
2.5 Deposition Method 59
2.5.1 Atomic Layer Deposition (ALD) Method 59
2.5.2 Chemical Vapor Deposition (CVD) Method 63
2.6 Conclusion and Future Perspective 64
Acknowledgments 65
References 65

3 Solvent- and Catalyst-Free Organic Transformation 69


Fatemeh Majidi Arlan and Farzad Zamani
3.1 Introduction 69
3.2 Solvent- and Catalyst-Free Organic Transformations 69
3.2.1 Mechanochemistry 70
3.2.2 Microwave Irradiation 80
3.2.3 Classical Heating 91
3.2.4 Ultrasound Irradiation 103
3.3 Conclusion 109
References 109

4 Metal Oxides as Catalysts/Supports in Solvent-Free Organic


Reactions 121
Manishkumar S. Tiwari, Saravana Kumaran, and Haresh G. Manyar
4.1 Introduction 121
4.2 Different Metal Oxides as a Catalyst/Support in Solvent-Free
Reactions 123
4.2.1 Titanium Dioxide-Based Catalysts 123
4.2.2 Tin Oxide-Based Catalysts 127
4.2.3 Manganese Oxide-Based Catalysts 132
4.2.4 Zinc Oxide-Based Catalysts 135
4.2.5 Aluminum Oxide-Based Catalysts 137
4.2.6 Iron Oxide-Based Catalysts 142
4.2.6.1 Fe3 O4 -Based Catalyst/Support 143
4.2.6.2 Fe2 O3 -Based Catalyst/Support 148
4.3 Conclusion 150
References 150
Contents vii

5 Silica-Based Materials as Catalysts or Supports in


Solvent-Free Organic Reactions 163
Gianvito Vilé and Christophe Len
5.1 Solvent-Free Reactions Over Silica Gel 163
5.2 Silica Nanoparticles and its Applications 168
5.3 Zeolites and Hierarchical Zeolite Structures 171
5.4 Conclusion 175
References 175

6 Carbon-Based Materials as Catalysts/Supports in Solvent-Free


Organic Reactions 181
Nagy L. Torad, Ahmed A.M. El-Amir, Fathy M. Hassan, Amira A. El-Maddah,
Wael A. Amer, and Mohamad M. Ayad
6.1 Introduction 181
6.2 Solvent-Free Catalysis Using Carbon-Based Materials 182
6.2.1 Activated Carbons (ACs) 183
6.2.1.1 Acetylation Reactions 183
6.2.1.2 Oxidation of Cyclohexane 184
6.2.2 Carbon-Based Solid Acid (CBSA) Catalysts 185
6.2.2.1 Cross-Aldol Condensation of Ketones with Aromatic
Aldehydes 185
6.2.2.2 Substituted Imidazoles 188
6.2.2.3 Amidoalkyl Naphthols 188
6.2.2.4 Reductive Amination of Aldehydes and Ketones 190
6.2.2.5 Xanthenes and Dibenzoxanthenes 192
6.2.2.6 Dihydropyrimidinone Compounds (Biginelli Reaction) 192
6.2.2.7 Acylation, Acetalization, Thioacetalization of Aldehydes 194
6.2.3 Carbon Nanotubes (CNTs) 195
6.2.3.1 Esterification of Alcohols 195
6.2.3.2 Benzyl Alcohol Oxidation 197
6.2.3.3 Phenol Derivatives Antioxidants 199
6.2.3.4 Acrylonitrile Derivatives 200
6.2.4 Graphene Oxide (GO) 200
6.2.4.1 Alkylaminophenols Derivatives 201
6.2.4.2 N-Arylation Reactions 202
6.2.4.3 Oxidation of Benzylic Alcohols 203
6.2.4.4 Aldol and Konevenagel Condensation Reaction 204
6.2.4.5 Oxidation of Cyclohexene 205
6.2.4.6 Oxidation of Hydrazide and Pyrazole Derivatives 207
6.2.5 Porous Carbon Materials 208
6.2.5.1 Oxidation of Alcohol and Hydrocarbons 209
6.2.5.2 Coupling of Amines 210
6.3 Summary and Future Perspectives 213
References 214
viii Contents

7 Nitride-Based Nanostructures for Solvent-Free Catalysis 229


Fatemeh Majidi Arlan, Ahmed A.M. El-Amir, Amira A. El-Maddah,
Nagy L. Torad, and Esmail Doustkhah
7.1 Carbon Nitride 229
7.1.1 Introduction 229
7.1.2 Synthesis of Carbon Nitride 232
7.1.3 Modification of Carbon Nitrides 232
7.1.4 Solvent-Free Catalysis with Carbon Nitrides 234
7.2 Boron Nitride 245
7.2.1 Introduction 245
7.2.2 Synthesis and Modification of Boron Nitride 246
7.3 Molybdenum Nitride 247
7.3.1 Introduction 247
7.3.2 Synthesis of Molybdenum Nitride 247
7.3.3 Solvent-Free Catalytic Application of Molybdenum Nitride 248
7.4 Aluminum Nitride 249
7.4.1 Introduction 249
7.4.2 Synthesis of Aluminum Nitride 249
7.4.2.1 Solvent-Free Synthesis 250
7.4.3 Solvent-Free Application of Aluminum Nitride 250
7.5 Conclusion 251
References 251

8 Supported Ionic Liquids for Solvent-Free Catalysis 261


Vahid Khakyzadeh and Sahra Sheikhaleslami
8.1 Introduction 261
8.2 Supported Ionic Liquids 263
8.3 Building Blocks of SILs 264
8.3.1 Ionic Segment 265
8.3.2 Solid-Support Segment 265
8.3.2.1 Silica Gels 267
8.3.2.2 Ordered Mesoporous Silicas 267
8.3.2.3 Carbon Nanotubes (CNTs) 268
8.3.2.4 Silica-Coated Magnetic Nanoparticles (SMNPs) 269
8.4 SIL Catalytic Systems 269
8.5 Supported IL Solvent-Free Catalysis 272
8.6 Solvent-Free Hydrogenation of Olefins 272
8.7 Solvent-Free Heck Reaction 273
8.8 Solvent-Free Multicomponent Reactions 275
8.8.1 Synthesis of Pyran-Based Heterocycles 275
8.8.2 Synthesis of 1,4-Dihydropyridine (Hantzsch Reaction) 279
8.8.3 Synthesis of 3,4-Dihydropyrimidine-2(1H)-One/Thiones
(Biginelli Reaction) 280
8.8.4 Synthesis of 1-Amidoalkyl Naphthol 282
8.8.5 Miscellaneous Solvent-Free Multicomponent Reactions 283
Contents ix

8.9 Solvent-Free Condensation Reactions 285


8.9.1 Solvent-Free Friedländer Condensation 285
8.9.2 Solvent-Free Knoevenagel Condensation 286
8.9.3 Esterification 286
8.10 Solvent-Free CO2 Conversion Reactions 287
8.11 Solvent-Free Oxidation Reactions 290
8.12 Miscellaneous Solvent-Free Organic Reactions 291
8.13 Conclusion 296
References 296

9 Present Status and Future Outlook 305


Elka Kraleva, Maria L. Saladino, and Izabela S. Pieta
9.1 Summary 305
9.2 Future Outlook 322
Acknowledgments 323
References 323

Index 329
xi

Preface

This book is a collection of recent developments in the area of solvent-free synthesis


and catalysis. This book can be a suitable collection for students and researchers
of green chemistry. We are witnessing a tremendous effort from green chemistry
scientists in developing solvent-free organic reactions since solvent-free methods
are assumed to be a promising approach to greening chemical transformations.
Solvent-free reactions are not only a green approach but also, in some cases, it is the
only highly efficient and straightforward strategy and the last resort that chemists
can take it to have an facile and highly efficient synthesis. In this regard, coupling
techniques such as microwave, ultrasonic irradiation, and ball-milling techniques
have been the new gates in the development of solvent-free techniques. Therefore,
we have invited the outstanding researchers and professors of the field to contribute
to this book to build a comprehensive collection of green chemistry to review and
discuss the reports from very old to very recent.
In this era, we believe that chemists should move on to developing solvent-free
methods of catalysts syntheses. Although catalysts in some cases are being synthe-
sized using solvent-free methods – also known as a solid-state synthesis methods –
they are not genuinely highlighted in the green chemistry synthesis class. Hence,
this part of the catalyst synthesis needs more attention and highlight.
Here, in this book, we have collected the solvent-free methods that can be clas-
sified through the catalyst type, e.g., carbon or silica, and besides, all the solvent-
free and catalyst-free approaches together in a separate chapter (Chapter 3). This
classification helps chemists and researchers to understand the efficiency and nature
of these catalysts based on the chemical structure of the catalysts. As the last word,
we hope this book could be a small portion in confronting the environmental threats
coming from the inevitable chemical processes. We wish this book could attract the
students and researchers more toward green chemistry of catalysts synthesis and the
reactions.

Esmail Doustkhah
Koç University, Turkey
Manoj B. Gawande
Institute of Chemical Technology, India
Anandarup Goswami
Vignan’s Foundation for Science, India
Rafael Luque
Universidad de Córdoba, Spain
1

Introduction: Scope of the Book


Anil Kumar Nallajarla and Anandarup Goswami
Vignan’s Foundation for Science, Technology and Research (VFSTR deemed to be University),
Department of Chemistry, School of Applied Science and Humanities, Vadlamudi,
Guntur 522 213, Andhra Pradesh, India

1.1 Introduction: Green Chemistry, Solvent-free


Synthesis, and Nanocatalysts
Since the realization that the future of chemical/industrial processes primarily
depends on their sustainable quotient, the path of modern science has shifted
toward the improvement of processes/products following green chemistry princi-
ples [1, 2]. Green chemistry as a branch of chemistry primarily deals with developing
chemical processes using environmentally benign protocols, including inexpensive
renewable less-toxic precursors. In that respect, the “12 rules of green chemistry,”
first formally introduced by Anastas and Warner in their book “Green Chemistry:
Theory and Practices” [3], play a pivotal role in identifying the areas that should
be focused to achieve the expected sustainability goals [4]. For chemical processes,
the main aim of introducing these 12 principles is to save the environment and
society by reducing the usage of toxic and hazardous chemicals and solvents
without affecting the product yield/selectivity. While the 12 principles are quite
self-explanatory, as depicted in Figure 1.1, the emphasis in certain areas often
depends on the convenience of implementing them for specific protocols and the
outcome. Of these 12 principles, this book focuses on the historical and recent
developments of strategies that minimize solvent use in a chemical process, often
termed “solvent-free synthesis,” and the associated catalytic procedures involving
nanomaterials.
The use of a solvent in a reaction creates a homogeneous solution phase where
the reactants can interact effectively. While ideally, any liquid can be used as a
solvent, the focus is largely on solvents based on their polarity and protic nature
(e.g. methanol, ethanol, chloroform, dichloromethane, dimethylformamide [DMF],
dimethyl sulfoxide [DMSO], toluene). The primary reasons can be attributed to
their ability to solubilize various organic reactants/products as well as to control
the stability of the transition state/intermediates (leading to modifications of the
thermodynamic and kinetic reaction parameters) [5]. However, with the growth of
Solvent-Free Methods in Nanocatalysis: From Catalyst Design to Applications, First Edition.
Edited by Rafael Luque, Manoj B. Gawande, Esmail Doustkhah, and Anandarup Goswami.
© 2023 WILEY-VCH GmbH. Published 2023 by WILEY-VCH GmbH.
2 1 Introduction: Scope of the Book

12 Principles of Green Chemistry

Atom economy Less hazardous chemical synthesis

Waste prevention Designing safer chemicals

Safer chemistry for accident Safer solvents and auxiliaries


prevention Recycle catalysts
Reuse chemicals
Reduce solvents
Realtime pollution prevention Design for energy efficiency

Design for degradation Use of renewable feedstocks

Catalysis Reduce derivatives

Figure 1.1 12 principles of green chemistry.

industrial chemical processes along with growing interest in developing sustainable


protocols, emphasis has been shifted toward choosing the best solvent based on
not only their solubilizing power but also their abundance, cost, and, last but not
least, short- and long-term impacts on environment [6]. In that context, water has
long been considered a sustainable choice. However, the poor solubility of organic
species in aqueous solution has limited its wide use primarily at the industrial scale
[7, 8]. While some of the current choices, such as ionic liquids (ILs), often come as
rescue options, their selection has remained an area of concern for the processes
related to bulk scale production of materials [9]. Around this debate regarding
the choice of solvent, the idea of “no solvent is best solvent” was also considered.
However, it did not receive its due because of the lack of initial appreciation,
especially for industrial purposes. However, with the advent of eco-friendly and
greener approaches, the idea of “solvent-free synthesis” resurfaced, and presently,
it is being explored as one of the viable options for the synthesis of chemicals as
well as various materials (Figure 1.2a) [10]. Initially, synthesis under solvent-free
conditions was associated with the solid-phase synthesis where the reactants
were made to react in the solid phase. However, recent advances in the area of
materials syntheses (which include thermal treatment, plasma etching, etc.) have
extended its scope significantly [11, 12]. Modern-day scientific and technological
developments are primarily governed by the utilization of materials for specific
purposes. Thus, the choice of their synthetic strategies is often determined by the
type of materials, their subsequent use, and their sustainable quotient. Considering
these, the class of “nanomaterials” has emerged as a crucial player. Hence, a brief
introduction of nanomaterials with special emphasis on their catalytic applications
seems timely before moving to a detailed discussion on their solvent-free synthetic
procedures.
Nanomaterials are the class of materials size that falls under 1–100 nm in at least
one dimension (Figure 1.2b) [13, 14]. The exceptional growth in the development
Figure 1.2 Introduction to (a) solvent-free synthesis, (b) nanomaterials, (c) catalysts, and (d) nanocatalysis. CNTs = Carbon nanotubes.
4 1 Introduction: Scope of the Book

of nanomaterials can be attributed to high surface area, quantum confinement


effect, and the possibility of fine-tuning their surface properties utilizing relatively
straightforward methods. While the natural origin of nanomaterials can be traced
back to the time of big-bang, Prof. Feynman, in his great lecture series “There is
plenty of room at the bottom,” first introduced the enormous potential of nanoma-
terials [15, 16]. Since then, progress in nanomaterials has been significant, and in
the modern world, it is nearly impossible not to encounter nanomaterials in daily
lives [17, 18]. Among the various fields in which nanomaterials have been explored,
one of the areas involves their catalytic applications due to their unique size and
shape-dependent surface properties [19, 20].
The term “catalyst” (or “catalysis”) was first introduced in 1835 by Swedish
chemist J.J. Berzelius [21, 22]. Since then, the catalyst is defined as a substance/
material that improves the reaction rate by minimizing the activation energy of
the process without being consumed during the process (Figure 1.2c). The initial
developments in catalysis were concentrated on relatively expensive transition
metal–derived systems primarily due to their intrinsic catalytic properties. However,
the need for sustainable results has allowed the recent advancements to focus on
less-toxic, low-cost, and highly abundant and recyclable catalytic alternatives
[23, 24]. Though the categorization of catalysts may vary depending on the clas-
sification criteria, catalysts are commonly divided into two classes: homogeneous
and heterogeneous [25]. Homogeneous catalysts are the compounds that remain in
the same phase as reactants/products during the catalytic reactions (mostly in the
presence of a suitable solvent). While homogeneous catalysts often exhibit higher
catalytic activity, poor separation and recyclability appear significant challenges.
In contrast, heterogenous systems involve different reactants/products and catalyst
phases and ideally have better separation and reusability. However, due to various
mass transfer and diffusion limitations, the catalytic activity of heterogeneous
catalysts remains inferior to their homogeneous counterparts. Thus, a combination
of advantageous factors for both systems is essential to overcome the existing
challenges of both sides to achieve the desired goals. In that context, the utilization
of nanomaterials (often termed “nanocatalysts,” [NCs], Figure 1.2d) either as
catalysts or as support materials for various homogeneous/heterogeneous catalytic
entities has opened newer avenues as they often exhibit the potential to overcome
the respective challenges in homo- and heterogenous catalysts [26].
The synthesis of NCs does not deviate too much from the synthesis of nano-
materials. It hence can primarily be classified into “top-down” and “bottom-up”
approaches [27, 28], each of which can further be divided based on specific tech-
niques. In “top-down” approaches, NCs are prepared from the bulk using various
“cutting” techniques, whereas the “bottom-up” approaches involve synthesis of
NCs from their atomic and/or molecular precursors. Both approaches have their
own advantages and disadvantages, and often, the choice of synthetic methods is
dictated by the NCs’ specific properties and applications. For instance, while various
“top-down” strategies are preferred for carbon-based nanomaterials (e.g. graphene,
nanotubes), metal-oxide nanoparticles (NPs) are generally synthesized using
1.2 Topics Covered in this Book 5

“sol–gel” techniques [29]. Irrespective of the synthetic processes, “solvent-free”


methods are always preferred as they can be directly related to the goals of green
and sustainable transformations.
This brief introduction provides a general idea about various related topics inter-
linked with a common theme of sustainability and hopefully allows the readers to
have a smooth transition in the remaining parts of the chapter.

1.2 Topics Covered in this Book


The chapters in this book are carefully designed to provide a very in-depth idea
about NCs and their applications using solvent-free methods. In Chapter 2,
Gawande and coworkers provide an illustrative overview of the syntheses of
nanocatalysts using solvent-free methods. The idea of introducing this chapter is
to make readers aware of various synthetic techniques and their advantages and
disadvantages that can be used for the solvent-free preparation of nanocatalysts.
The remaining part of the book (except the conclusion chapter, Chapter 9) primarily
deals with catalytic applications of nanocatalysts using solvent-free methods. In
Chapter 3, Prof. Zamani introduce the topic of solvent and catalyst-free organic
transformations with specific examples of academic and industrial importance to
set up the stage for the next chapters. While Manyar and coworkers describe various
solvent-free organic transformations, catalyzed by metal/metal-oxide nanocatalysts
in Chapter 4, a separate chapter is dedicated to silica-based nanomaterials as
catalysts/support for solvent-free organic reactions (Chapter 5). In Chapter 6,
Prof. Torad emphasizes the importance of carbon-based nanomaterials either
as supports or as nanocatalysts for solvent-free organic reactions, focusing on
doped and functionalized nanocarbons. In Chapters 7 and 8, concentration has
been deliberately steered toward current developments in the areas of solvent-free
reactions using nitride-based and ionic liquid-based nanocatalytic systems,
respectively, because they have been explored recently as potential sustainable
choices in comparison to the existing ones. In the book’s concluding chapter
(Chapter 9), Prof. Pieta summarizes the present status of solvent-free synthesis
of nanomaterials and solvent-free nanocatalytic transformations, emphasizing
their relevance with green chemistry and sustainability. The author also pro-
vides a brief account of the challenges related to current approaches and some
possible solutions as an outlook. Starting from the introductory chapter to the
concluding one, the primary focus of this book has been on providing readers with
insights into the background, recent advances, and future possibilities regarding
solvent-free synthesis of nanocatalysts as well as solvent-free nanocatalytic meth-
ods. We, as contributors, strongly feel that this book will be helpful to students
and researchers who want to gain knowledge about these topics and pursue their
research in those areas. Following that thought, the next sections highlight some
solvent-free methods used in the synthesis of NCs, followed by their specific catalytic
applications.
6 1 Introduction: Scope of the Book

1.3 Solvent-Free Synthesis of Nanocatalysts


As the name suggests, solvent-free methods for preparing uniform, monodisperse
solid NCs essentially follow the synthetic protocols that do not use any solvent
[27, 30]. The size and shape of the synthesized NCs vary depending on the methods
and the reaction conditions. While the approaches used in these syntheses can
also be considered as a part of traditional “top-down” or “bottom-up” based
classification, the present literature examples tend to categorize them in terms of
specific procedures. In that context, primarily four types of procedures along with
their subclassifications have been reported: (i) mechanochemical, (ii) thermal, (iii)
plasma-assisted, and, last but not least, (iv) deposition techniques (Figure 1.3).
The mechanochemical process for the synthesis of nanomaterials primarily
involves grinding of bulk precursors into nanoscale materials using mechanical
force [31, 32]. In that category, the traditional “ball-milling” approach has widely
been used [33–35]. For example, Barcellos and coworkers synthesized CuO NCs
using high-energy ball milling, and the synthesized NCs were used for nitroarene
reduction under aqueous media [36]. Recently, solid-phase synthesis that involves
grinding the precursor materials using a mortar and pestle has also generated much
interest owing to its simple operational procedure [37]. For instance, gram-scale
synthesis of Au/chitosan was reported by Reddy et al. recently using a mortar and
pestle. The nanocatalyst was used for catalytic homocoupling of phenylboronic acid
and the aerobic oxidation of benzyl alcohol in water [38].
One of the most widely used and well-studied approaches for the synthesis of NCs
includes the preparation of NCs using heat as an energy source, and this strategy
(often termed “thermal treatment”) primarily involves heating of molecular precur-
sors at high temperatures under an oxidative or reductive environment to obtain
nanomaterials [39–41]. For example, various metal oxides are routinely synthesized
from their precursors using heat treatment under an oxidative environment (e.g.
air, O2 ) [42]. However, for synthesizing carbon-based nanomaterials, precursors are
often pyrolyzed under an inert atmosphere (N2 , Ar, etc.). The product yield, extent
of doping, and degree of graphitization depend on the nature of precursors, tem-
perature, heating rate, etc. [43, 44]. Goswami et al. showed that a hydrogen-assisted
thermal treatment could be used to convert metal precursors into metal NPs [45]. As
opposed to conventional heating treatment, to make the synthetic process greener
and more sustainable, alternative energy sources (such as microwave (MW) or ultra-
sound) or bio-derived precursors/processes have also been used to prepare metal
nanoparticles NPs [46–48].
With the advent of instrumental developments, several sophisticated “top-down”
approaches have been developed to synthesize NCs. Among them, plasma-assisted
strategies have shown great promise due to their environmentally benign nature,
no additional requirements of solvent or stabilizing agents, etc. [49]. In this case,
“feed materials” are transformed into atoms or molecules through vaporization with
the help of plasma and thus NPs/NCs are formed. The size, morphology, and prop-
erties of the final materials are generally dependent on the absolute temperature
of plasma, the kinetics of plasma formation, quenching process, and the size and
Figure 1.3 Representative solvent-free methods used for the synthesis of NCs. NPs = Nanoparticles.
8 1 Introduction: Scope of the Book

composition of feed materials. Based on the internal energy of electrons used for
plasma generation, this can be classified into two categories: (i) thermal plasma and
(ii) cold-plasma methods [50]. While plasma is used to generate an intense heat
source in both cases, the significant difference between these two processes lies in
the operating temperature. In the case of thermal plasma, feed materials are atom-
ized at an operating temperature of a few thousand degrees. Nanomaterials/NCs
are formed during the cooling process. For example, NiO nanocubes were prepared
from bulk Ni metal utilizing thermal plasma with oxygen as a carrier gas [51]. On the
other hand, cold plasma uses a low-pressure, low-temperature method where NCs
can be synthesized even at room temperature [52]. This energy-efficient method is
primarily used to synthesize noble-metal NPs (e.g. Pd, Ag, Au), as exemplified by
recent reports [53–55]. Recently, Haye et al. have also utilized this method to syn-
thesize non–noble-metal-based FeNPs (more precisely Fe3 N) embedded on carbon
support [56].
Another solvent-free method for the synthesis of nanomaterials/nanocatalysts
that gained tremendous attention is deposition techniques (more precisely, vapor
deposition techniques) [57, 58]. In this method, nanomaterials/nanocatalysts are
deposited on a substrate in the form of thin films from their atomic precursors.
Depending on the nature of deposition, they can further be classified into physical
and chemical vapor deposition (CVD), and among them also several subclassifica-
tions are made. In the context of solvent-free synthesis of NPs, CVD techniques [59]
and atomic layer deposition (ALD) techniques [60] have been widely explored. For
instance, carbon-based support materials such as carbon nanotubes (CNTs) and
graphene are routinely synthesized using the CVD method [61–63]. In addition,
metal NPs have been embedded onto a carbon matrix using the CVD method
[64, 65]. On the other hand, the precursors copper(II)-hexafluoroacetylacetonate
[Cu(hfac)2 ] and diethylzinc [DEZ, (C2 H5 )2 Zn] were used to synthesize Cu/ZnO-50
nanocatalysts using the ALD method [66].
The aforementioned examples are the only selected ones chosen from the vast pool
of synthetic strategies employed for the preparation of NPs while these are primar-
ily representative of the solvent-free protocols; a detailed discussion on this topic is
included in Chapter 2.

1.4 Solvent and Catalyst-Free Organic Transformations


To have a smooth transition from the solvent-free synthesis of NCs to solvent-free
processes utilizing NCs; as an intermediate, a separate section has been devoted
to solvent and catalyst-free organic transformations (Figure 1.4). As opposed to
the use of relatively toxic catalytic entities and solvents, solvent- and catalyst-free
organic transformations can be considered among the classes of reactions that
aim to follow sustainability goals [78]. While these straightforward protocols have
shown great promise, the desired success of these procedures is often restricted
Figure 1.4 Representative examples of solvent and catalyst-free organic transformations. Source: Adapted from the references given in parentheses.
10 1 Introduction: Scope of the Book

due to poor yield, significant energy investment, etc. However, in recent years,
the advancement of alternative energy sources and increasing knowledge of
fundamental reaction mechanisms enable them to get closer to the expected ideal
outcome.
Among other processes, mechanochemical processes hold a special place because
of the utilization of simple mechanical force/energy to drive the reactions forward
and its underexplored potential for large-scale production of organic compounds
without complicated purification steps. Different types of mechanical processes
have been developed depending on the types of reactions and their outcome,
among which ball-milling, twin-screw, mortar and pestle methods have become
very popular. For example, thiourea derivatives were synthesized using ball-milling
within 10–20 minutes [67], and various multicomponent reactions (MCRs) such
as Ugi, Biginelli have been carried out using twin-screw extruder (TSE) [68].
Additionally, mortar and pestle method has also been exploited for the synthesis
of fused heterocycles [69]. These are representative of the enormous possibility of
mechanochemical synthesis.
Conventional thermal heating under solvent- and catalyst-free conditions is often
considered the classical way of achieving the desired product for any chemical trans-
formations. Despite the challenges related to the product selectivity and energy effi-
ciency that have impacted the long-term use of the process negatively, several het-
erocycles (including imidazoles [70], pyrazoles [71]), phosphonates [72] have been
synthesized under solvent- and catalyst-free thermal heating methods.
To tackle the challenges related to the conventional thermal heating process, alter-
native energy sources such as MW and ultrasound have given a fresh impetus to
make the synthetic methods more energy efficient and greener [73, 74]. The use
of MW under solvent- and catalyst-free conditions has been explored to synthesize
spiro compounds [75], N-containing heterocycles [79], etc. The employed proto-
cols have shown faster kinetics, better reactivity, higher selectivity, and broader sub-
strate scope. The utilization of ultrasound irradiation to drive a chemical reaction
is also considered one of the green approaches. In this context, a combination of
solvent- and catalyst-free method with ultrasound irradiation stands unique because
of safer energy inputs, waste reduction, higher yields/selectivity, etc., compared to
other traditional approaches [80]. Starting from simple formylation [76] or protec-
tion of amines [77] to multicomponent coupling of heterocycles [81], the use of
ultrasound irradiation has proven to be highly efficient, as exemplified in recent
examples.
The aforementioned methods provide a glimpse of solvent- and catalyst-free
approaches that have shown great promise compared to traditional ones. More
details about these procedures/approaches are provided in Chapter 3.

1.5 Solvent-Free Reactions Using NCs


Most of the solvent-free reactions using NCs are largely focused on three major
classes: (i) metal or metal-oxide NPs (Figure 1.5), (ii) silica-based NCs (Figure 1.6),
Figure 1.5 Representative examples of solvent-free nanocatalytic procedures using metal-oxide NPs as catalysts/supports. Source: Adapted from the
references given in parentheses. MNPs = Magnetic nanoparticles.
12 1 Introduction: Scope of the Book

Figure 1.6 Representative examples of silica-supported nanocatalytic systems for organic


transformations. Source: Adapted from the references given in parentheses.

and (iii) carbon-based nanosystems as catalyst/support (Figure 1.7). Additionally,


nitride-based (Figure 1.8) and ionic liquid (IL)-based NCs (Figure 1.9) have recently
been reported. While such topics will be discussed in detail in subsequent chapters,
several such reactions are highlighted to exhibit a variety of the responses that can
be performed using these NCs.

1.5.1 Different Metal Oxides as a Catalyst/Support in Solvent-Free


Reaction
1.5.1.1 Titanium Oxide
Titanium oxide (TiO2 ) and titanium oxide–supported catalysts are used for organic
reactions, including solvent-free methods, because of their unique properties and
catalytic activity. In 2007, M. Hosseini-Sarvari et al. reported TiO2 as a new and
reusable nanocatalyst for the Knoevenagel condensation reaction, which exhibited
good to excellent yields [100]. For efficient conversion of cyclohexylamine into
cyclohexanone oxime (78.4%) with high selectivity (89.1%) under solvent-free
1.5 Solvent-Free Reactions Using NCs 13

Figure 1.7 Saluted examples of carbon-based nanocatalysts for organic transformations


under solvent-free conditions. Source: Adapted from the references given in parentheses.

conditions, Liu et al. used mobil composition of matter (MCM)-41-supported


TiO2 NCs [82]. After using for five cycles, no significant change was observed
in catalytic efficiency, and characterization data showed the hydroxyl groups on
titania acted as catalytic sites. Recently, Amoozadeh and coworkers synthesized
nickel(II) Schiff base complex supported on nano-titanium dioxide, and the sup-
ported NCs were used for the synthesis of 3,4-dihydopyrimidin-2(1H)-ones through
Biginelli reaction under solvent-free conditions [83]. The catalytic activity was
found to be superior to previous catalytic systems and showed diverse substrate
scope.

1.5.1.2 Tin Oxide


Tin oxide nanoparticles (more specifically SnO2 NPs) exhibit excellent catalytic
properties in various organic reactions to synthesize organic compounds. In
addition to the high catalytic activity/selectivity of SnO2 , the possibility of easy
separation (or reusability) has made this oxide unique as exemplified in several
instances, including ones that use solvent-free conditions. SnO2 NPs as NCs were
utilized to synthesize 2H-indazolo[2,1-b]phthalazine-triones using an MCR of
aromatic aldehydes 1,3-cyclohexanedione and phthalhydrazide under solvent-free
conditions [84]. The catalytic results showed that the final products can be obtained
in good yields with high selectivity. While initial research was more focused on the
catalyst and yields, with more advancements in the catalysis, recyclability came
into the picture. For example, 1,2,4,5-tetra substituted imidazoles were prepared
using silica-supported tin oxide under solvent-free conditions [85]. In addition to
14 1 Introduction: Scope of the Book

high catalytic performance (yield: 84–97%), the catalyst can be separated easily
from the reaction mixture and recycled up to five cycles without any significant
change in the catalyst’s activity and/or composition. In another example, Ahmed
et al. synthesized nanocrystalline sulfated tin oxide. They used them as NCs to
synthesize coumarin derivatives under solvent-free conditions using acetoacetate
and m-cresol as reactants [101].

1.5.1.3 Manganese Oxide (MnOx )


Manganese metal is widely used in catalysis because it is inexpensive, is easily avail-
able, is less toxic, and shows variable oxidation states ranging from +II to +VII. Its
various oxides are also available in different forms such as 3D structure, chain-like
structures, layered or sheet structures. These materials are routinely used as hetero-
geneous catalysts for numerous chemical transformations. In that respect, graphene
oxide–supported manganese oxide (GO/MnO2 ) was used as a catalyst for synthe-
sizing chalcogens under solvent and solvent-free conditions [86]. The NCs showed
superior catalytic activity than their counterparts, confirming the synthesis between
each compound. In another example, manganese oxide–doped magnesium oxide
(MnO2 /MgO) was employed as a NC to prepare ethyl cinnamate (mostly known as
a flavoring agent) using a Witting reaction between benzaldehyde and tri-phenyl
phosphonium salts following a green mechanochemical solvent-free approach [87].

1.5.1.4 Zinc Oxide


Zinc oxide (ZnO) NPs have also been routinely explored in various applica-
tions, including optoelectronic field, photoluminescence devices, and solar cells.
In addition, the recent research efforts are also engaged in tuning the shape,
morphology, and properties of ZnO NPs for their utilization in electronic and
antibacterial applications. In the case of nanocatalysis (especially for organic
transformations), ZnO NPs have also been exploited as novel and reusable catalysts.
In the present context of solvent-free reactions, ZnO-NPs-catalyzed Biginelli
reaction was reported to synthesize dihydropyrimidinones using aromatic alde-
hydes, urea, or thiourea and acetoacetic esters under solvent-free conditions [88].
The catalyst showed excellent catalytic activity (94–97% of yields). In another
study, a wide range of chloroesters were synthesized from cyclic ethers and acyl
chlorides using ZnO NPs as nanocatalysts under solvent-free conditions at room
temperature [102]. The catalyst can provide good yields (87–95%) and be recycled
easily up to three cycles. The ZnO nanoflowers, derived from the peel of Musa
balbisiana and zinc nitrate, were used as catalysts for synthesizing chalcones
via the Claisen–Schmidt condensation reaction using MW irradiation under
solvent-free conditions [89]. The ZnO-decorated GO nanocomposite acts as a highly
efficient reusable catalyst for synthesizing xanthenedione from 1,3-dicarbonyl
compounds and aromatic aldehydes under neat reaction conditions [103]. The final
products were obtained in excellent yields, and the catalyst can be used up to five
cycles.
1.5 Solvent-Free Reactions Using NCs 15

1.5.1.5 Aluminum Oxide


Aluminum oxide (Al2 O3 , commonly known as alumina) has been extensively
employed in the separation and purification of organic compounds primarily
due to the insolubility of aluminum oxide in both water and organic solvents
and variable interactions between the alumina surface and the eluting com-
pounds with different polarities. Moreover, due to alumina’s high Lewis acidic
nature, catalytic applications of aluminum oxide and supported aluminum
oxide-based materials have also been explored for various organic transformations.
For example, alumina NPs were used for the MCR to synthesize dihydropyrim-
idinones under solvent-free synthesis through the Biginelli reaction [90]. ZrO2 -
stabilized aluminum oxide was also reported as recyclable NCs (up to six
cycles) for O-methoxymethylation reaction between substituted alcohols and
dimethoxymethane under solvent-free conditions [104]. In another case,
Shetarian et al. synthesized phosphoric acid–supported aluminum oxide
(H3 PO4 /Al2 O3 ), which could be used as an efficient catalyst for MCRs to syn-
thesize 2H-indazolo[2,1-b]phthalazinetriones, 2,3-dihydroquinazoline4(1H)-ones,
and benzo[4,5]imidazo[1,2-a]pyrimidines under solvent-free conditions [91].

1.5.1.6 Iron Oxide


Iron oxide NPs exist in various forms depending on the oxidation states of iron,
oxygen vacancy, etc., among which maghemite, magnetite, and hematite are
the prevalent ones. Because of the magnetic nature of some of the iron oxide
NPs, the magnetic separation of the catalyst from the reaction mixture becomes
easy, resulting in better catalyst recovery for recyclability. In 2015, Habibi et al.
used iron oxide NPs as catalysts for the synthesis of benzoxanthenes using the
reaction between aryl aldehydes, dimedone, and 2-naphthol [92]. The catalyst
worked effectively with excellent yields (80–95%) and was reusable up to 20 cycles.
Magnetic iron oxide NPs as support were also explored extensively for anchoring
catalytic entities on them. For example, Abbasi et al. synthesized CuO@γ-Fe2 O3
(copper oxide–supported magnetic NPs) to synthesize substituted guanidines via
the simple addition of amines to carbodiimides [93]. The separation of the catalyst
was straightforward by applying an external magnetic field, and it was confirmed
that the presence of copper improved the activity of the catalyst and produced
the desired products in good yields (60–97%) under solvent-free conditions.
The catalyst could be reusable for four cycles without any significant loss of its
activity.

1.5.2 Silica-Based Materials as Catalysts/Supports in Solvent-Free


Organic Reactions
The use of silica and/or silica-based nanomaterials as either NCs or supports mainly
stems from their exceptional thermal and chemical stability, high surface area,
straightforward synthetic protocols, and the possibility of surface functionalization
to incorporate various organic and/or inorganic functionalities [94, 105]. Among
16 1 Introduction: Scope of the Book

the different silica materials, two major types are extensively used for catalytic
applications either as catalysts or as support: (i) nonporous and (ii) porous silica.
While nonporous silica is historically important, the high surface area of porous
silica offers a significant advantage due to the enhanced accessibility of catalytically
active sites. As the catalytic activity of silica-based materials depends on the
functionalities present on the surface, controlling the amount, distribution, and
nature of surface functionalities is essential. Irrespective of the materials, most of
the synthetic and/or functionalization strategies utilize “sol–gel” methods. Hence,
the synthesis of silica-based nanocatalysts hardly follows solvent-free protocols
[95]. However, post-synthetic modifications are often performed under solvent-free
conditions to introduce and control surface functionalities [106]. In terms of
catalytic activity, surface silanol groups with their Lewis acidic character assist
the active catalysts whenever needed. In addition, surface functionalization, i.e.
incorporating catalytically functional groups, becomes a pivotal step in developing
better NCs. In general, organic functionalities are grafted on the silica surface using
organosilanes, and they have been used as recyclable organocatalysts for organic
transformations [107]. These supported NCs have been used in various organic
transformations, including solvent-free ones (Figure 1.6). For example, a sulfonic
acid-functionalized silica nanosphere (SAFSNS) catalyst was prepared and used
under solvent-free conditions to synthesize carboxylic acid ester [96]. To incorporate
metal/metal-based NPs on silica surface, suitable organic functionalities are grafted
that act as anchoring sites for the metal precursors. Depending on the requirements,
metals can be further reduced to their nanoparticle forms or oxidized to oxide forms.
In that respect, selective oxidation of toluene was achieved using silica-supported
Au NPs where the performance of the nanocatalysts improved through a mild
reductive deprotection strategy [108].

1.5.3 Carbon-Based Materials as Catalysts/Supports in Solvent-Free


Organic Reactions
Carbon-based NCs are often considered highly promising due to their high thermal
stability, higher surface area, and low cost. The traditional “organocatalysts” are
excluded in this category as they have already been considered under “small
molecule” homogeneous catalysis rather than a heterogeneous one. In most cases,
acid- or base-functionalized carbon-based materials can be used as catalysts.
Alternatively, carbon materials can also act as support for anchoring other cat-
alytic entities. Recent progress in this area indicates that carbon materials can be
synthesized by pyrolysis of biodegradable waste materials such as corn, coconut
shells at higher temperatures [43, 109]. The amphiphilic carbon has been used as
a recyclable catalyst for the reaction between glycerol and 2-propanone to form
solketal under solvent-free conditions [97]. Zali et al. used a carbon-based solid
acid catalyst for the solvent-free aldol condensation of aromatic aldehydes with
ketones to obtain the desired products in good yields; the catalyst can be reusable
up to five times without a decrease in yields [98]. The carbon-based catalyst
has also been explored for various MCRs. For instance, Tavakoli-Hoseini et al.
1.5 Solvent-Free Reactions Using NCs 17

reported a carbon-based solid acid catalyst for the synthesis of tetra-substituted


imidazoles via one-pot reaction using benzil, primary amines, aromatic aldehyde,
and ammonium acetate as starting materials [99]. Figure 1.7 represents some
examples where carbon-based nanocatalysts have been explored under solvent-free
applications.

1.5.4 Nitride-Based Materials as Catalysts/Supports in Solvent-Free


Organic Reactions
The functionalized or modified graphitic carbon nitrides (g-C3 N4 ) are used as cat-
alysts to synthesize several organic compounds under solvent-free green synthetic
methods. In 2021, Azizi and coworkers synthesized xanthene derivatives using
sulfonic acid–functionalized graphitic carbon nitride under solvent-free condition
using the ball-milling method (Figure 1.8) [110]. The catalyst showed excellent
catalytic activity with high yields and short reaction time and can be easily separated
and reused up to four cycles.
Sonogashira coupling reaction is well known for the formation of the carbon–
carbon bond. Commonly, palladium-based catalytic systems are used for
Sonogashira coupling reactions. However, efforts have been devoted to finding a sus-
tainable solution to replace expensive and scarce palladium-based catalysts. In that
context, Akhlaghinia group reported a new catalyst, i.e. cobalt oxide (Co3 O4 ) embed-
ded in a triplet shelled carbon nitride (TSCN) for the Sonogashira–Hagihara cross-
coupling reaction between aryl halides and substituted acetylenic compounds
under solvent-free conditions (Figure 1.8) [111]. Though some of the reactions
took place in water, the catalyst exhibited good performance under solvent-free
conditions. The catalyst can be reusable up to five cycles without significant change
in their conversion and rate.
In 2021, Liu and coworkers reported hexagonal boron nitride (h-BN) nanoflakes
and a titanium dioxide hybrid photocatalytic system for the oxidation of cyclohexane
with oxygen to form cyclohexanone with better selectivity than cyclohexanol under
solvent-free conditions. The catalyst showed excellent catalytic activity up to four
cycles without any change in the yield and selectivity (Figure 1.8) [112].
For controlling environmental pollution, conversion of carbon dioxide (CO2 ) into
useful products is an efficient way of mitigating the hazardous impact of CO2 . Chand
et al. developed a catalytic system for converting carbon dioxide into cyclic carbon-
ates at atmospheric pressure using boron-doped graphitic carbon nitride as a catalyst
under solvent-free conditions [113]. The catalyst was synthesized using thermal con-
densation method (Figure 1.8). The catalyst’s performance with respect to yield,
selectivity, and turn-over number (TON) was superior to the previously reported cat-
alytic systems. The catalyst can be reusable for eight cycles without any change in
the activity and efficiency. A similar work was previously reported by Yin’s group in
2016 (Figure 1.8) [114]. They synthesized phosphorous-modified carbon nitride and
used it as a catalyst (in the presence of tetra butyl ammonium bromide [Bu4 NBr] as
cocatalyst) for the cycloaddition reaction between carbon dioxide and epoxides to
obtain the cyclic carbonates in excellent yields. The catalyst was readily separated
18 1 Introduction: Scope of the Book

Figure 1.8 Representative examples of nitride-based materials as catalysts/supports in


solvent-free organic reactions. Source: Adapted from the references given in parentheses.
TSCN = triplet shelled carbon nitride.

from the reaction mixture via centrifugation. After washing and drying, the catalyst
could be recycled up to five cycles with above 90% yields and 100% selectivity of the
product.

1.5.5 Ionic Liquid-Based Materials as Catalysts/Supports


in Solvent-Free Organic Reactions
Ionic liquids (ILs) are organic salts that exist under ambient conditions. Generally,
they are made of polyatomic inorganic anions and organic cations and used
1.5 Solvent-Free Reactions Using NCs 19

Figure 1.9 Representative examples of ionic liquid (IL)-based materials as


catalysts/supports in solvent-free organic reactions. Source: Adapted from the references
given in parentheses.

as solvents because of their wide range of solubility. Recently, they (either in


their original forms or as functionalized derivatives) have been used as catalysts,
especially under solvent-free conditions. For example, Dadhania et al. prepared iron
oxide nanoparticle-supported acidic ionic liquid catalyst to synthesize quinolines
and fused polycyclic quinolines by the Friedlander reaction under solvent-free
conditions (Figure 1.9) [115]. The catalysts can be used up to six cycles without any
loss of efficiency.
In recent years, synthesis of diphenyl carbonates has become extremely impor-
tant due to their diverse applications, including the preparation of polycarbonates.
Recently, Wang et al. reported that diphenyl carbonates could be successfully
20 1 Introduction: Scope of the Book

synthesized using a metal-free, Santa Barbara Amorphous (SBA)-15 IL hybrid


catalyst under solvent-free conditions [116]. By changing the functional groups,
they prepared 14 different catalysts, among which [SBA-15-IL-OH]Br with
hydroxyl-terminated groups showed superior performance than the previously
reported transition metal–based catalysts. The catalyst was reused up to the sixth
cycle without significant changes in conversion, yield, and selectivity.
The oxidation of hydrocarbons plays a crucial role in organic synthesis. Instead
of using metal-based compounds as catalysts (Co, Mn, etc.), Dobras et al. syn-
thesized a catalyst using commercially available silica gel immobilized with
N-hydroxyphthalimide (NHPI), followed by a coating with ILs [117]. The cata-
lysts have been used for the oxidation of ethyl benzene (EB) under solvent-free
conditions, resulting in three different products, namely, acetophenone (AP), ethyl-
benzene hydroperoxide (EBOOH), and 1-phenylethanol (PEOH). The selectivity of
the product changes with the reaction conditions. To the best of their knowledge,
this was the first time this class of catalyst was used under solvent-free conditions
and recycled for three cycles [117].
IL-catalyzed Knoevenagel condensation reaction between different types of aro-
matic aldehydes and malononitrile or ethyl cyanoacetate under solvent-free condi-
tions was reported by Yue et al. in 2008 [118]. With a minimal amount of the catalyst,
the reactions were completed within less than an hour, and the catalysts exhibited
high reactivity/selectivity and substrate scopes. The easy separation of the catalysts
via simple filtration allowed the NCs to be recycled up to five cycles without any
noticeable changes in the activity [118].

1.6 Present Status and Future Direction

Rapid progress in NCs can be attributed to their shape- and size-dependent


properties (which are very different from their bulk counterparts), tunable surface
functionality, higher surface area, ease of synthetic protocols, and a better under-
standing of their reaction mechanisms due to the rapid growth of sophisticated
instrumentation. During the developmental phase of nanocatalysis, the main focus
was devoted to making them more efficient and selective, and, in that process,
sustainability parameters were often overlooked. However, since the formulation
of “12 principles of green chemistry” to make processes more sustainable, the
field of nanocatalysis has also reorganized its importance and started to develop
accordingly. In that respect, both solvent-free synthesis of catalyst and solvent-free
catalytic processes using NCs have taken up a central stage.
The idea of a solvent-free process undoubtedly pushes the boundaries of the
existing synthetic strategies to achieve the required goal of sustainability. While
the recent thrust in this area accompanies delivering catalytically improved active
species, the term “solvent-free” refers to any solventless process either during the
preparation of NCs or during the catalytic processes. Since both parts are recognized
1.6 Present Status and Future Direction 21

as equally important for developing sustainable procedures, the chapters in this


book emphasize both aspects.
The question of how solvent-free procedures can be applied to the synthesis
of NCs depends on the structure of nanomaterials and the target of use. In that
respect, any form of energy that can replace the conventional thermal meth-
ods has the potential of making the synthetic steps more energy efficient and
economically viable. The processes ranging from mechanochemical approaches
to relatively sophisticated plasma-based strategies have been exploited for the
solvent-free preparation of NCs. In addition, the synthesis of NCs has also been
widely investigated using energy-efficient alternative energy sources. Moreover,
for the catalytic systems where the primary synthetic strategies do not necessarily
involve solvent-free approaches (e.g. silica), several post-synthetic solvent-free
modifications are adopted to incorporate additional functionalities to further
modify surface properties. The advancement of solvent-free deposition techniques
has been beneficial for preparing various 1D and 2D nanomaterials. Briefly, the
present status of solvent-free methods of NCs preparation indicates a brighter future
where more precise control can be achieved.
As far as the solvent-free catalytic processes are concerned, the potential of NCs
has been investigated in various reactions. Though gas-phase catalytic reactions are
beyond the scope of the present discussion, it is worth mentioning that significant
development has been observed. This progress can be attributed to the drive from
industrial sectors to utilize green-house gases to produce value-added chemicals and
also to convert the stored chemical energies into other forms of energies [119]. Nano-
materials’ involvement in catalyzing organic reactions has been extensively studied
because NCs offer the advantages of both heterogeneous and homogeneous cata-
lysts [120]. The solvent-free organic reactions utilizing NCs often deliver further
assistance to improve the catalytic activity and selectivity by providing a conducive
reaction environment. In this respect, metal-oxide-based NCs are routinely explored
for solvent-free catalytic organic transformations, ranging from simple Knoevenagel
and Michael condensation reactions to the complicated synthesis of heterocycles or
multicomponent reactions with a particular emphasis on bioactive target molecules.
Other classes of NCs that have also gained tremendous attention include silica-based
and carbon-based nanosystems. Different types of oxidation, reduction, coupling,
multicomponent reactions are catalyzed using these materials under solvent-free
conditions. While both are predominantly used to anchor/support other catalyti-
cally active species onto their surface to improve catalytic stability and recyclability,
the presence of surface functional groups (e.g. silanols for silica and various organic
functionalities for carbon-based nanomaterials) has also been reported to act as cat-
alytic entities. In that context, “Carbocatalysis” has also emerged as an active area
of research where defects and surface functionalities play an essential role in deter-
mining the catalytic outcome [121, 122].
Despite the significant attention the solvent-free methods (either NCs synthesis or
nanocatalytic organic transformations) have received in recent years, the expected
22 1 Introduction: Scope of the Book

growth is often limited due to the various challenges. Some of the existing hurdles
that need to be overcome to make the field more acceptable are as follows:

(1) The solvent-free methods for the synthesis of NCs are still at their develop-
mental phase due to a lack of knowledge of energy-transfer processes. Thus,
to make these processes more economically viable and universally acceptable,
more experimental and theoretical studies are essential for a fundamental
understanding of the mechanism under solvent-free conditions.
(2) Despite several literature precedents where nanomaterials are effectively
used as catalysts, barring energy industries, the industrial uses of NCs have
been limited, which is particularly pronounced in pharmaceutical indus-
tries where hardly any step in the synthesis of potential drug candidates
uses NCs. The difficulties in scaling up the catalysts, existing well-established
protocols for homogeneous catalysts, inferior activity of nanocatalysts compared
to the homogeneous ones, poor understanding of nanocatalytic mechanism,
etc. are identified as the reasons behind the lack of interest from such industries.
Thus, more serious efforts are needed to explore the possibility of using similar
NCs for industrially relevant challenging transformations.
(3) Till date most studies involving NCs under solvent-free conditions focus on find-
ing out the efficiency of NCs and establishing the superiority of the solvent-free
methods in terms of reactivity and selectivity. However, other parameters includ-
ing short-term and long-term effects of NCs on the health and environment, the
compatibility between the highly concentrated reactants and NCs, the reasons
for catalyst poisoning are often ignored and hence need to be thoroughly inves-
tigated.

Undoubtedly, these challenges are multifaceted and require sincere efforts from
every discipline. Fortunately, the researchers who work in these areas have already
acknowledged these gaps and have actively been engaged tackling these issues. More
details about these concerns and their potential solutions are discussed in the sub-
sequent chapters. To the best of our knowledge, no available books have dealt with
these issues yet. We do hope readers, especially those who want to know the present
status and future opportunities of solvent-free nanocatalytic processes, will find this
book essential and relevant. It will motivate them to pursue their research in the
exciting multidisciplinary area.

References

1 Roschangar, F., Sheldon, R.A., and Senanayake, C.H. (2015). Overcoming


barriers to green chemistry in the pharmaceutical industry – The Green
Aspiration LevelTM concept. Green Chem. 17 (2): 752.
2 Phan, T.V.T., Gallardo, C., and Mane, J. (2015). GREEN MOTION: a new and
easy to use green chemistry metric from laboratories to industry. Green Chem.
17 (5): 2846.
References 23

3 Anastas, P.T. and Warner, J. (1998). Green Chemistry: Theory and Practice.
London: Oxford University Press.
4 Anastas, P. and Eghbali, N. (2010). Green chemistry: principles and practice.
Chem. Soc. Rev. 39: 301.
5 Riddick, J.A. and Bunger, W.B. (1986). Organic Solvents: Physical Properties and
Methods of Purification, 4e. Nashville, TN: Wiley.
6 Gani, R., Jiménez-González, C., and Constable, D.J.C. (2005). Method for
selection of solvents for promotion of organic reactions. Comput. Chem. Eng. 29
(7): 1661.
7 Lubineau, A. and Augé, J. (1999). Water as solvent in organic synthesis. In:
Modern Solvents in Organic Synthesis (ed. P. Knochel), 1. Berlin, Heidelberg:
Springer Berlin Heidelberg.
8 Li, C.-J. and Chen, L. (2006). Organic chemistry in water. Chem. Soc. Rev. 35
(1): 68.
9 Chanda, A. and Fokin, V.V. (2009). Organic synthesis “on water”. Chem. Rev.
109 (2): 725.
10 Tanaka, K. and Toda, F. (2000). Solvent-free organic synthesis. Chem. Rev. 100
(3): 1025.
11 Metzger, J.O. (1998). Solvent-free organic syntheses. Angew. Chem. Int. Ed. Engl.
37 (21): 2975.
12 Li, C.-J. (2000). Water as solvent for organic and material synthesis. In: ACS
Symposium Series (ed. P.T. Anastas, L.G. Heine and T.C. Williamson), 62.
Washington, DC: American Chemical Society.
13 Ghoranneviss, M., Soni, A., Talebitaher, A., and Aslan, N. (2015). Nanomaterial
synthesis, characterization, and application. J. Nanomater. 2015: 1.
14 Bayda, S., Adeel, M., Tuccinardi, T. et al. (2019). The history of nanoscience
and nanotechnology: from chemical-physical applications to nanomedicine.
Molecules 25 (1): 112.
15 Feynman, R.P. (2011). There’s plenty of room at the bottom: an invitation to
enter a new field of physics. Resonance 16 (9): 890.
16 Toumey, C. (2009). Plenty of room, plenty of history. Nat. Nanotechnol. 4 (12):
783.
17 Gupta, R. and Xie, H. (2018). Nanoparticles in daily life: applications, toxicity
and regulations. J. Environ. Pathol. Toxicol. Oncol. 37 (3): 209.
18 Nasrollahzadeh, M., Sajadi, S.M., Sajjadi, M., and Issaabadi, Z. (2019). Applica-
tions of nanotechnology in daily life. Interface Sci. Technol. 28: 113.
19 Alayoglu, S. (2017). Achievements, present status, and grand challenges of
controlled model nanocatalysts. In: Morphological, Compositional, and Shape
Control of Materials for Catalysis (ed. P. Fornasiero and M. Cargnello), 85.
Elsevier.
20 Polshettiwar, V. and Varma, R.S. (2010). Green chemistry by nanocatalysis.
Green Chem. 12 (5): 743.
21 Berzelius, J.J. (1835). Sur un Force Jusqu’ici Peu Remarquée qui est
Probablement Active Dans la Formation des Composés Organiques, Section
on Vegetable Chemistry. Jahres-Bericht 14: 237.
24 1 Introduction: Scope of the Book

22 Wisniak, J. (2010). The history of catalysis. From the beginning to Nobel prizes.
Educ. quím. 21 (1): 60.
23 Ludwig, J.R. and Schindler, C.S. (2017). Catalyst: sustainable catalysis. Chem 2
(3): 313.
24 Turner, N.J. (2016). Sustainable catalysis. Beilstein J. Org. Chem. 12: 1778.
25 Li, C. and Liu, Y. (ed.) (2014). Bridging Heterogeneous and Homogeneous
Catalysis: Concepts, Strategies, and Applications, 1e. Weinheim: Wiley-VCH
Verlag.
26 Astruc, D., Lu, F., and Aranzaes, J.R. (2005). Nanoparticles as recyclable
catalysts: the frontier between homogeneous and heterogeneous catalysis.
Angew. Chem. Int. Ed. 44 (48): 7852.
27 Dhand, C., Dwivedi, N., Loh, X.J. et al. (2015). Methods and strategies for the
synthesis of diverse nanoparticles and their applications: a comprehensive
overview. RSC Adv. 5 (127): 105003.
28 Saravanan, A., Kumar, P.S., Karishma, S. et al. (2021). A review on biosynthesis
of metal nanoparticles and its environmental applications. Chemosphere 264
(Pt 2): 128580.
29 Niederberger, M. and Pinna, N. (2009). Aqueous and nonaqueous sol–gel
chemistry. In: Metal Oxide Nanoparticles in Organic Solvents (ed. B. Derby), 7.
London: Springer London.
30 Panigrahi, S., Kundu, S., Ghosh, S. et al. (2004). General method of synthesis
for metal nanoparticles. J. Nanopart. Res. 6 (4): 411.
31 Tsuzuki, T. and McCormick, P.G. (2004). Mechanochemical synthesis of
nanoparticles. J. Mater. Sci. 39 (16): 5143–5146.
32 Xu, C., De, S., Balu, A.M. et al. (2015). Mechanochemical synthesis of advanced
nanomaterials for catalytic applications. Chem. Commun. 51 (31): 6698.
33 Fecht, H.J., Hellstern, E., Fu, Z., and Johnson, W.L. (1990). Nanocrystalline
metals prepared by high-energy ball milling. Metall. Trans. A 21 (9): 2333.
34 Kumar, M., Xiong, X., Wan, Z. et al. (2020). Ball milling as a mechanochemical
technology for fabrication of novel biochar nanomaterials. Bioresour. Technol.
312: 123613.
35 Piras, C.C., Fernández-Prieto, S., and De Borggraeve, W.M. (2019). Ball milling:
a green technology for the preparation and functionalisation of nanocellulose
derivatives. Nanoscale Adv. 1 (3): 937.
36 Lucchesi, S.A., Farias Soares, M.R., Machado, G., and Barcellos, T. (2021).
Improved mechanochemical fabrication of copper(II) oxide nanoparticles with
low E-factor. Efficient catalytic activity for nitroarene reduction in aqueous
medium. ACS Sustainable Chem. Eng. 9 (29): 9661.
37 Moores, A. (2018). Bottom up, solid-phase syntheses of inorganic nanomaterials
by mechanochemistry and aging. Curr. Opin. Green Sustainable Chem. 12: 33.
38 Reddy, K.P., Meerakrishna, R.S., Shanmugam, P. et al. (2021). Rapid gram-scale
synthesis of Au/chitosan nanoparticles catalysts using solid mortar grinding.
New J. Chem. 45 (1): 438.
39 Navaladian, S., Viswanathan, B., Viswanath, R., and Varadarajan, T. (2006).
Thermal decomposition as route for silver nanoparticles. Nanoscale Res. Lett.
2 (1): 44.
References 25

40 Salavati-Niasari, M. and Davar, F. (2009). Synthesis of copper and copper(I)


oxide nanoparticles by thermal decomposition of a new precursor. Mater. Lett.
63 (3, 4): 441.
41 Goodarz, N.M., Saion, E.B., Ahangar, H.A. et al. (2011). Synthesis and
characterization of manganese ferrite nanoparticles by thermal treatment
method. J. Magn. Magn. Mater. 323 (13): 1745–1749.
42 Kim, B., Kim, J., Baik, H., and Lee, K. (2015). Large-scale one pot synthesis of
metal oxide nanoparticles by decomposition of metal carbonates or nitrates.
CrystEngComm. 17 (27): 4977.
43 Zahid, M.U., Pervaiz, E., Hussain, A. et al. (2018). Synthesis of carbon
nanomaterials from different pyrolysis techniques: a review. Mater. Res Exp.
5 (5): 052002.
44 Zhang, S., Jiang, S.F., Huang, B.C. et al. (2020). Sustainable production of
value-added carbon nanomaterials from biomass pyrolysis. Nat. Sustain. 3 (9):
753.
45 Goswami, A., Kadam, R.G., Tuček, J. et al. (2020). Fe(0)-embedded thermally
reduced graphene oxide as efficient nanocatalyst for reduction of nitro
compounds to amines. Chem. Eng. J. 382: 122469.
46 Bilecka, I. and Niederberger, M. (2010). Microwave chemistry for inorganic
nanomaterials synthesis. Nanoscale 2 (8): 1358.
47 Xu, H., Zeiger, B.W., and Suslick, K.S. (2013). Sonochemical synthesis of
nanomaterials. Chem. Soc. Rev. 42 (7): 2555.
48 Huang, J., Lin, L., Sun, D. et al. (2015). Bio-inspired synthesis of metal
nanomaterials and applications. Chem. Soc. Rev. 44 (17): 6330.
49 Palma, V., Cortese, M., Renda, S. et al. (2020). A review about the recent
advances in selected non-thermal plasma assisted solid-gas phase chemical
processes. Nanomaterials 10 (8): 1596.
50 Balasubramanian, C. (2020). Thermal plasma processes and nanomaterial
preparation. In: Nanotechnology for Energy and Environmental Engineering
(ed. L. Ledwani and J.S. Sangwai), 73. Cham: Springer International Publishing.
51 Hou, G., Du, Y., Cheng, B. et al. (2018). Enhanced capacity of NiO nanocubes
with high dispersion and exposed facets reinforced by thermal plasma. ACS
Appl. Nano Mater. 1 (11): 5981.
52 Di, L., Zhang, J., and Zhang, X. (2018). A review on the recent progress,
challenges, and perspectives of atmospheric-pressure cold plasma for
preparation of supported metal catalysts. Plasma Processes Polym. 15 (5):
1700234.
53 Adak, D., Chakrabarty, P., Majumdar, P. et al. (2020). Pd nanoparticle-decorated
hydrogen plasma-treated TiO2 for photoelectrocatalysis-based solar energy
devices. ACS Appl. Electron. Mater. 2 (12): 3936.
54 Hamood Al-Masoodi, A.H., Goh, B.T., Farhanah Binti Nazarudin, N.F. et al.
(2020). Efficiency enhancement in blue phosphorescent organic light emitting
diode with silver nanoparticles prepared by plasma-assisted hot-filament
evaporation as an external light-extraction layer. Mater. Chem. Phys. 256:
123618.
26 1 Introduction: Scope of the Book

55 Izadi, A. and Anthony, R.J. (2019). A plasma-based gas-phase method for


synthesis of gold nanoparticles. Plasma Processes Polym. 16 (7): e1800212.
56 Haye, E., Soon, C.C., Dudek, G. et al. (2019). Tuning the magnetism of
plasma-synthesized iron nitride nanoparticles: application in pervaporative
membranes. ACS Appl. Nano Mater. 2 (4): 2484.
57 Ahmad, R., Wolfbeis, O.S., Hahn, Y.B. et al. (2018). Deposition of
nanomaterials: a crucial step in biosensor fabrication. Mater. Today Commun.
17: 289.
58 Charitidis, C.A., Georgiou, P., Koklioti, M.A. et al. (2014). Manufacturing
nanomaterials: from research to industry. Manuf. Rev. (Les Ulis). 1: 11.
59 He, C.N., Zhao, N.Q., Shi, C.S., and Song, S.Z. (2009). Fabrication of carbon
nanomaterials by chemical vapor deposition. J. Alloys Compd. 484 (1): 6.
60 Zhang, Z., Zhao, Y., Zhao, Z. et al. (2020). Atomic layer deposition-derived
nanomaterials: oxides, transition metal dichalcogenides, and metal–organic
frameworks. Chem. Mater. 32 (21): 9056.
61 Manawi, Y., Ihsanullah, S.A., Al-Ansari, T., and Atieh, M. (2018). A review
of carbon nanomaterials’ synthesis via the chemical vapor deposition (CVD)
method. Materials 11 (5): 822.
62 Liu, Z., Lin, L., Ren, H., and Sun, X. (2017). CVD synthesis of graphene. In:
Thermal Transport in Carbon-Based Nanomaterials (ed. G. Zhang), 19. Elsevier.
63 Esteves, L.M., Oliveira, H.A., and Passos, F.B. (2018). Carbon nanotubes as
catalyst support in chemical vapor deposition reaction: a review. J. Ind. Eng.
Chem. 65: 1.
64 Choi, D.S., Robertson, A.W., Warner, J.H. et al. (2016). Low-temperature
chemical vapor deposition synthesis of Pt–Co alloyed nanoparticles with
enhanced oxygen reduction reaction catalysis. Adv. Mater. 28 (33): 7115.
65 Choi, D.S., Kim, C., Lim, J. et al. (2018). Ultrastable graphene-encapsulated
3 nm nanoparticles by in situ chemical vapor deposition. Adv. Mater. 30 (51):
e1805023.
66 Ren, Q.H., Zhang, Y., Wang, T. et al. (2019). Facile synthesis and
photoluminescence mechanism of ZnO nanowires decorated with cu
nanoparticles grown by atomic layer deposition. ACS Appl. Electron. Mater.
1 (8): 1616.
67 Ould, M.M., Alshammari, A.G., and Lemine, O.M. (2016). Green high-yielding
one-pot approach to Biginelli reaction under catalyst-free and solvent-free ball
milling conditions. Appl. Sci. 6 (12): 431.
68 Ali El-Remaily, M.A., Soliman, A.M., and Elhady, O.M. (2020). Green method
for the synthetic Ugi reaction by twin screw extrusion without a solvent and
catalyst. ACS Omega 5 (11): 6194.
69 Bhat, S.I., Choudhury, A.R., and Trivedi, D.R. (2012). Condensation of
malononitrile with salicylaldehydes and o-aminobenzaldehydes revisited:
solvent and catalyst free synthesis of 4H-chromenes and quinolines. RSC Adv.
2 (28): 10556.
70 Zhao, H.Y., Wu, F.S., Yang, L. et al. (2018). Catalyst- and solvent-free approach
to 2-arylated quinolines via [5+1] annulation of 2-methylquinolines with
diynones. RSC Adv. 8 (9): 4584.
References 27

71 Khan, M.M., Shareef, S., Saigal, S., and Sahoo, S.C. (2019). A catalyst- and
solvent-free protocol for the sustainable synthesis of fused 4H-pyran derivatives.
RSC Adv. 9 (45): 26393.
72 Trofimov, B., Artem’ev, A., Malysheva, S. et al. (2014). Catalyst- and
solvent-free stereoselective addition of secondary phosphine chalcogenides
to alkynes. Synthesis 47 (02): 263.
73 Shanmugavelan, P., Sathishkumar, M., Nagarajan, S. et al. (2014). The
first solvent-free, microwave-accelerated, three-component synthesis of
thiazolidin-4-ones via one-pot tandem Staudinger/aza-Wittig reaction.
J. Heterocycl. Chem. 51 (4): 1004.
74 Jida, M., Soueidan, O.M., Deprez, B. et al. (2012). Racemic and
diastereoselective construction of indole alkaloids under solvent-and
catalyst-free microwave-assisted Pictet–Spengler condensation. Green Chem.
14 (4): 909.
75 Yuvaraj, P., Manivannan, K., and Reddy, B.S.R. (2015). Microwave-assisted effi-
cient and highly chemoselective synthesis of oxazolo[5,4-b]quinoline-fused
spirooxindoles via catalyst- and solvent-free three-component tandem
Knoevenagel/Michael addition reaction. Tetrahedron Lett. 56 (1): 78.
76 Claudio-Catalán, M.Á., Pharande, S.G., Quezada-Soto, A. et al. (2018).
Solvent-and catalyst-free one-pot green bound-type fused bis-heterocycles
synthesis via Groebke–Blackburn–Bienaymé reaction/SNAr/ring-chain
azido-tautomerization strategy. ACS Omega 3 (5): 5177.
77 Dar, B., Singh, A., Sahu, A. et al. (2012). Catalyst and solvent-free, ultrasound
promoted rapid protocol for the one-pot synthesis of α-aminophosphonates at
room temperature. Tetrahedron Lett. 53 (41): 5497.
78 Gawande, M.B., Bonifácio, V.D.B., Luque, R. et al. (2014). Solvent-free and
catalysts-free chemistry: a benign pathway to sustainability. ChemSusChem
7 (1): 24.
79 Bhuyan, D., Sarmah, M.M., Dommaraju, Y., and Prajapati, D. (2014).
Microwave-promoted efficient synthesis of spiroindenotetrahydropyridine
derivatives via a catalyst- and solvent-free pseudo one-pot five-component
tandem Knoevenagel/aza-Diels–Alder reaction. Tetrahedron Lett. 55 (37): 5133.
80 Banerjee, B. (2017). Recent developments on ultrasound assisted catalyst-free
organic synthesis. Ultrason. Sonochem. 35: 1.
81 Azarifar, D. and Sheikh, D. (2012). Ultrasound-promoted catalyst-free
synthesis of 2,2′ -(1,4-phenylene)bis[1-acetyl-1,2-dihydro-4H-3,1-benzoxazin-
4-one] derivatives. Helv. Chim. Acta 95 (7): 1217.
82 Liu, S., You, K., Song, J. et al. (2018). Supported TiO2 /MCM-41 as an efficient
and eco-friendly catalyst for highly selective preparation of cyclohexanone
oxime from solvent-free liquid phase oxidation of cyclohexylamine with
molecular oxygen. Appl. Catal., A 568: 76.
83 Tabrizian, E., Amoozadeh, A., and Shamsi, T. (2016). A novel class of
heterogeneous catalysts based on toluene diisocyanate: the first amine-
functionalized nano-titanium dioxide as a mild and highly recyclable solid
nanocatalyst for the Biginelli reaction. React. Kinet. Mech. Catal. 119 (1): 245.
28 1 Introduction: Scope of the Book

84 Maheswari, C.S., Shanmugapriya, C., Revathy, K., and Lalitha, A. (2017).


SnO2 nanoparticles as an efficient heterogeneous catalyst for the synthesis
of 2H-indazolo[2,1-b]phthalazine-triones. J. Nanostructure Chem. 7 (3): 283.
85 Borhade, A.V., Tope, D.R., and Gite, S.G. (2017). Synthesis, characterization and
catalytic application of silica supported tin oxide nanoparticles for synthesis of
2,4,5-tri and 1,2,4,5-tetrasubstituted imidazoles under solvent-free conditions.
Arabian J. Chem. 10: S559.
86 Kumar, A., Rout, L., Achary, L.S.K. et al. (2021). Solvent free synthesis of
chalcones over graphene oxide-supported MnO2 catalysts synthesized via
combustion route. Mater. Chem. Phys. 259: 124019.
87 Moulavi, M.H., Kale, B.B., Bankar, D. et al. (2019). Green synthetic method-
ology: an evaluative study for impact of surface basicity of MnO2 doped MgO
nanocomposites in Wittig reaction. J. Solid State Chem. 269: 167.
88 Bahrami, K., Mehdi, K.M., and Farrokhi, A. (2009). Highly efficient solvent-free
synthesis of dihydropyrimidinones catalyzed by zinc oxide. Synth. Commun.
39 (10): 1801.
89 Tamuly, C., Saikia, I., Hazarika, M. et al. (2015). Bio-derived ZnO nanoflower:
a highly efficient catalyst for the synthesis of chalcone derivatives. RSC Adv.
5 (12): 8604.
90 Tanna, J.A., Chaudhary, R.G., Gandhare, N.V., and Juneja, H.D. (2016).
Alumina nanoparticles: a new and reusable catalyst for synthesis of dihy-
dropyrimidinones derivatives. Adv. Mater. Lett. 7 (11): 933.
91 Shaterian, H.R., Fahimi, N., and Azizi, K. (2014). New applications of
phosphoric acid supported on alumina (H3 PO4 –Al2 O3 ) as a reusable
heterogeneous catalyst for preparation of 2,3-dihydroquinazoline-4(1H)-ones,
2H-indazolo[2,1-b]phthalazinetriones, and benzo[4,5]imidazo[1,2-a]pyrimidines.
Res. Chem. Intermed. 40 (5): 1879.
92 Habibi, D., Kaamyabi, S., and Hazarkhani, H. (2015). Fe3 O4 nanoparticles
as an efficient and reusable catalyst for the solvent-free synthesis of
9,9-dimethyl-9,10-dihydro-8H-benzo-[a]xanthen-11(12H)-ones. Chin. J. Catal.
36 (3): 362.
93 Abbasi, S., Saberi, D., and Heydari, A. (2017). Copper oxide supported on
magnetic nanoparticles (CuO@γ-Fe2 O3 ): an efficient and magnetically
separable nanocatalyst for addition of amines to carbodiimides towards
synthesis of substituted guanidines: CuO@γ-Fe2 O3 -catalyzed hydroamination of
carbodiimides. Appl. Organomet. Chem. 31 (9): e3695.
94 Rostamnia, S., Lamei, K., Mohammadquli, M. et al. (2012, 2012).
Nanomagnetically modified sulfuric acid (γ-Fe2 O3 @SiO2 -OSO3 H): an efficient,
fast, and reusable green catalyst for the Ugi-like Groebke–Blackburn–Bienaymé
three-component reaction under solvent-free conditions. Tetrahedron Lett.
53 (39): 5257.
95 Rostamizadeh, S., Nojavan, M., Aryan, R. et al. (2013). Amino acid-based ionic
liquid immobilized on α-Fe2 O3 -MCM-41: an efficient magnetic nanocatalyst and
recyclable reaction media for the synthesis of quinazolin-4(3H)-one derivatives.
J. Mol. Catal. A: Chem. 374: 102.
References 29

96 Ahmad, I.S., Dhar, R., Hisaindee, S., and Hasan, K. (2021). An environmentally
benign solid acid nanocatalyst for the Green synthesis of carboxylic acid ester.
ChemistrySelect 6 (36): 9645.
97 Carvalho, B.F., da Silva, M.J., Paula de Carvalho Teixeira, A. et al. (2020). Fuel
274: 117799.
98 Zali, A., Ghani, K., Shokrolahi, A., and Keshavarz, M.H. (2008). Carbon-based
solid acid as an efficient and reusable catalyst for cross-aldol condensation of
ketones with aromatic aldehydes under solvent-free conditions. Chin. J. Catal.
29 (7): 602.
99 Tavakoli-Hoseini, N. and Davoodnia, A. (2011). Carbon-based solid acid as
an efficient and reusable catalyst for one-pot synthesis of tetrasubstituted
imidazoles under solvent-free conditions. Chin. J. Chem. 29 (1): 203.
100 Hosseini-Sarvari, M., Sharghi, H., and Etemad, S. (2007). Solvent-free
Knoevenagel condensations over TiO2 . Chin. J. Chem. 25 (10): 1563.
101 Ahmed, A.I., El-Hakam, S.A., Khder, A.S., and Abo El-Yazeed, W.S. (2013).
Nanostructure sulfated tin oxide as an efficient catalyst for the preparation
of 7-hydroxy-4-methyl coumarin by Pechmann condensation reaction. J. Mol.
Catal. A: Chem. 366: 99.
102 Tang, K.Y., Chen, J.X., Legaspi, E.D.R. et al. (2021). Gold-decorated TiO2
nanofibrous hybrid for improved solar-driven photocatalytic pollutant
degradation. Chemosphere 265 (129114): 129114.
103 Hasanzadeh, B.S., Dekamin, M.G., and Yaghoubi, A. (2018). Selective and
highly efficient synthesis of xanthenedione or tetraketone derivatives catalyzed
by ZnO nanorod-decorated graphene oxide. New J. Chem. 42 (17): 14246.
104 Pratap, S.R., Shyamsundar, M., and Shamshuddin, S.Z.M. (2018). Mesoporous
ZrO2 –Al2 O3 (ZA) mixed metal oxide as an efficient and reusable catalyst for
the liquid phase O-methoxymethylation reaction under solvent free conditions.
J. Porous Mater. 25 (5): 1265.
105 Gawande, M.B., Monga, Y., Zboril, R., and Sharma, R.K. (2015). Silica-decorated
magnetic nanocomposites for catalytic applications. Coord. Chem. Rev. 288: 118.
106 Sun, B., Zhou, G., and Zhang, H. (2016). Synthesis, functionalization, and
applications of morphology-controllable silica-based nanostructures: a review.
Prog. Solid State Chem. 44 (1): 1.
107 Asefa, T. and Tao, Z. (2012). Mesoporous silica and organosilica materials –
review of their synthesis and organic functionalization. Can. J. Chem. 90 (12):
1015.
108 Das, S., Goswami, A., Hesari, M. et al. (2014). Reductive deprotection of
monolayer protected nanoclusters: an efficient route to supported ultrasmall
au nanocatalysts for selective oxidation. Small 10 (8): 1473.
109 Konwar, L.J., Boro, J., and Deka, D. (2014). Review on latest developments in
biodiesel production using carbon-based catalysts. Renewable Sustainable Energy
Rev. 29: 546.
110 Qareaghaj, O.H., Ghaffarzadeh, M., and Azizi, N. (2021). A rapid and
quantitative synthesis of xanthene derivatives using sulfonated graphitic carbon
nitride under ball-milling. J. Heterocycl. Chem. 58 (10): 2009.
30 1 Introduction: Scope of the Book

111 Ghodsinia, S.S.E., Akhlaghinia, B., and Jahanshahi, R. (2021). Co3 O4


nanoparticles embedded in triple-shelled graphitic carbon nitride
(Co3 O4 /TSCN): a new sustainable and high-performance hierarchical catalyst
for the Pd/Cu-free Sonogashira–Hagihara cross-coupling reaction in solvent-free
conditions. Res. Chem. Intermed. 47 (8): 3217.
112 Wang, K., Xue, B., Wang, J.L. et al. (2021). Efficient and selective oxidation
of cyclohexane to cyclohexanone over flake hexagonal boron nitride/titanium
dioxide hybrid photocatalysts. Mol. Catal. 505: 111530.
113 Chand, H., Choudhary, P., Kumar, A. et al. (2021). Atmospheric pressure
conversion of carbon dioxide to cyclic carbonates using a metal-free Lewis
acid-base bifunctional heterogeneous catalyst. J. CO2 Util. 51: 101646.
114 Lan, D.H., Wang, H.T., Chen, L. et al. (2016). Phosphorous-modified bulk
graphitic carbon nitride: facile preparation and application as an acid-base
bifunctional and efficient catalyst for CO2 cycloaddition with epoxides. Carbon
100: 81.
115 Dadhania, H., Raval, D., and Dadhania, A. (2021). A highly efficient and
solvent-free approach for the synthesis of quinolines and fused polycyclic
quinolines catalyzed by magnetite nanoparticle-supported acidic ionic liquid.
Polycyclic Aromat. Compd. 41: 440.
116 Wang, S., Zhang, Q., Cui, C. et al. (2021). Ionic liquids-SBA-15 hybrid catalysts
for highly efficient and solvent-free synthesis of diphenyl carbonate. Green
Energy Environ. http://dx.doi.org/10.1016/j.gee.2021.02.010.
117 Dobras, G., Kasperczyk, K., Jurczyk, S., and Orlińska, B. (2020).
N-Hydroxyphthalimide supported on silica coated with ionic liquids containing
CoCl2 (SCILLs) as new catalytic system for solvent-free ethylbenzene oxidation.
Catalysts 10 (2): 252.
118 Yue, C., Mao, A., Wei, Y., and Lü, M. (2008). Knoevenagel condensation
reaction catalyzed by task-specific ionic liquid under solvent-free conditions.
Catal. Commun. 9 (7): 1571.
119 Whang, H.S., Lim, J., Choi, M.S. et al. (2019). Heterogeneous catalysts for
catalytic CO2 conversion into value-added chemicals. BMC Chem. Eng. 1 (1):
http://dx.doi.org/10.1186/s42480-019-0007-7.
120 Filiciotto, L. and Luque, R. (2018). Nanocatalysis for green chemistry. In: Green
Chemistry and Chemical Engineering, Encyclopedia of Sustainability Science
and Technology Series (ed. B. Han and T. Wu), 1. New York, NY: Springer
New York.
121 Navalon, S., Dhakshinamoorthy, A., Alvaro, M. et al. (2017). Active sites on
graphene-based materials as metal-free catalysts. Chem. Soc. Rev. 46: 4501.
122 Pentsak, E.O., Gordeev, E.G., and Ananikov, V.P. (2020). Carbocatalysis: from
acetylene trimerization to modern organic synthesis. A review. Dokl. Phys
Chem. 493: 95.
32 2 Strategies for the Preparation of Nanocatalysts and Supports Under Solvent-Free Conditions

Ar NH3 Ar
Ar NH3
Sn S Zr Ball
Zn Cu(OAc)2
Cu3N
Urea
Cu Calcination at 300 °C
O
CZTS
NH2 NH3
H2N

N-Source Cu(OAc)2
CZTS

H2 O2

Mixture Pulverization Chemical reaction Compound H2O

(a) Mechanochemistry (b) Thermal treatment


methods Solvent-free methods
synthesis NPs and
supports
Power supply Bulk Ni
O2 Ni

Bulk Ni
ICP vapors
N2
Pd(hfac)2
Nucleation
Thermal plasma and growth Formalin
reinforced
: Pd(hfac)2 : -hfac

: Formalin : Polar H
NiO NiO N2
nanocubes nanocubes : Pd :C

(c) Plasma-assisted (d) Deposition


methods methods

Figure 2.1 Synthesis strategies for nanoparticles and supports by solvent-free approach,
(a) mechanochemistry method, Source: Reproduced with permission from Park et al. [15].
Royal Society of Chemistry. (b) Thermal treatment method. Source: Reproduced with
permission from Panda et al. [16]. American Chemical Society. (c) Plasma-assisted method.
Source: Reproduced with permission from Hou et al. [17]. American Chemical Society. (d)
Deposition method. Source: Reproduced with permission from Lu et al. [18]. American
Chemical Society.

are dissolved in the particular solvents or surfactants and reduced to the formation
of desired NP products. Another one is physical methods, wherein NPs are achieved
by evaporation, direct heating, condensation, laser ablation, sputter deposition, and
many other strategies that require solvents. However, solvent-free synthetic proce-
dures possess energy and environmental benefits. The solvent-free NPs synthesis
methods obey the green chemistry principles, such as the prepared NPs used as cat-
alysts, with high atom economy, minimize the by-products of the reactions, and are
environmentally benign [14].
This chapter focuses on recent solvent-free methods for synthesizing various NPs
and supports an update concerning this solvent-free synthesis area’s new progress.
The solvent-free NPs and supports can be achieved by various synthetic strategies,
including mechanochemistry, thermal treatments, plasma-assisted, and deposition
methods (Figure 2.1).

2.2 Mechanochemistry
Mechanochemistry is an interesting method for synthesizing new materials,
NMs, and other composite materials. The development of solvent-free synthesis
2.2 Mechanochemistry 33

via mechanochemistry is imperative for many applications, including mineral


processing, coal industry, pharmacy, waste treatment, agriculture, building indus-
try, and several others. It is important to note that the mechanochemistry process
is employed for several chemical conversions using mechanical energy traditions,
such as grinding, ball milling, and sliding to form new materials [19, 20]. After a
long history, the International Union of Pure and Applied Chemistry (IUPAC) coin
the term mechanochemistry, which involved mechanically supported reactions,
achieved via absorption of mechanical energy. In the nineteenth century, a separate
branch of mechanochemistry in chemistry was introduced, accredited by Matthew
Carey Lea [21, 22]. Theophrastus, a student of Aristotle, mentioned in his book -On
Stones, first mechanochemical reaction performed using a copper mortar and pestle
for grinding of cinnabar with the presence of vinegar yields mercury [20].

2.2.1 Ball Milling


In the eighth decade of the last century, the ball milling method was proposed
worldwide. The mechanical alloying and milling methods have been applied for a
wide range of powder materials and their alloys. In detail, the ball milling method
is not new and dates back to more than 150 years [23]. The ball milling process
uses the various types of milling reactors with metal balls to mix the reaction
mixture/raw material to achieve desired materials. In this part, we have included
recent mechanochemical ball milling processes for a variety of NP syntheses.
Park et al. have synthesized the Cu2 ZnSnS4 (CZTS) NPs by mechanochemi-
cal approach via planetary ball milling (PM) [15]. The corresponding materials
prepared by using a mixture of copper, zinc, tin, and sulfur (molar ratio of
Cu:Zn:Sn:S = 1.7 : 1.2 : 1.0 : 4.0), which further placed in a stainless still (SS) jar
filled along with argon (Ar) gas and zirconia balls (Figure 2.2a). The obtained
milled powder material was examined by transmission electron microscopy (TEM),
high-resolution transmission electron microscopy (HRTEM), and selected area
electron diffraction (SAED) pattern. The TEM analysis confirmed the size of NMs
(15 nm) (Figure 2.2b) and the HRTEM image and SAED pattern confirmed that the
CZTS NPs are crystalline in nature and lattice fringes through interplanar distances
of 0.31, 0.27, and 0.16 nm to the (112), (200), and (312) planes of kesterite CZTS,
correspondingly (Figure 2.2c,d).
The solvent-free mechanochemical synthesis provides the moderately large-scale
production of CZTS NPs (∼20 g).
The ball milling process has been employed in different types of mechanochemi-
cal processes for the preparation of NPs. For example, Malca et al. synthesized the
ultrasmall Bi2 S3 NPs by the solvent-free mechanically activated milling processes,
which included mixer ball milling (MM), PM, and also manual grindings (MGs)
[24]. The MM process was used for the mixing of Bi(NO3 )3 ⋅5H2 O and L-cystine
with the presence of oleylamine (OA) or sodium 6-amino-hexanoate (AHA) as
the capping agent (Figure 2.3a). The resultant pale-yellow product was achieved
after 90 minutes to obtain final Bi2 S3 NPs, after 12 hours aging period at ambient
34 2 Strategies for the Preparation of Nanocatalysts and Supports Under Solvent-Free Conditions

(a)
Zr Ball
Sn S
Zn
Cu

CZTS

CZTS

Mixture Pulverization Chemical reaction Compound

(b) (c)
(d)
d(112) = 0.31 nm
d(200) = 0.27 nm
d(312) = 0.16 nm

2 nm

(d)

100 nm 5 1/nm

Figure 2.2 (a) Schematic representation of CZTS NPs synthesis, (b) TEM image, (c) HRTEM
image, (d) SAED pattern of CZTS NPs. Source: Reproduced with permission from Park et al.
[15]/Royal Society of Chemistry.

temperature. As-prepared Bi2 S3 NP size depends on the stoichiometry ratio of OA


and bismuth to obtain 5 : 1 (5OA@Bi2 S3 -90-MM) and 10 : 1 (10OA@Bi2 S3 -90-MM).
TEM images characterized the 5OA@Bi2 S3 -90-MM and 10OA@Bi2 S3 -90-MM
exhibited very tiny Bi2 S3 NPs diameter sizes 2.09 ± 0.29 and 2.08 ± 0.32 nm, cor-
respondingly. Again 5OA@Bi2 S3 -MG NPs were obtained via MG step, where the
diameter 1.96 ± 0.24 nm (Figure 2.3b–d) is confirmed. The PM process synthesized
5OA@Bi2 S3 and 5AHA@Bi2 S3 NPs were observed a uniform size 1.88 ± 0.27 and
3.0 ± 0.41 nm correspondingly. In this study, the AHA capping agent and planetary
milling reactors are suitable for the scale-up of the process of NPs.
The mesoporous alumina-supported platinum NPs have been prepared by using
the mechanochemical ball milling method [25]. The noble metal precursor used
as acetylacetone platinum (Pt(acac)2 ) is mixed with block copolymers (P123) and
Al(OCHCH3 CH3 )3 in the corundum reactor. The ball milling process formed
the gelatinous yellowish solid compound, which can be calcinated at 400 ∘ C for
2.2 Mechanochemistry 35

(a)
C8H17
NH2
3 OA

Bi2S3

Bi(NO3)3∙ 5H2O
or
+
OH
Toluene
HS OH
AHA
NH2 Mechanical Manual Aging
Grinding Bi2S3 Water
NaO
NH2
O

(b1) (c1) (d1)

(b2) 2.09 ± 0.29 nm (c2) 2.08 ± 0.32 nm (d2) 1.96 ± 0.24 nm


30 30 30
Frequencies (%)

Frequencies (%)

Frequencies (%)

20 20 20

10 10 10

0 0 0
1.0 2.0 3.0 1.0 2.0 3.0 1.0 2.0 3.0
Size distribution (nm) Size distribution (nm) Size distribution (nm)

Figure 2.3 (a) Schematic representation of the synthesis of Bi2 S3 NPs, TEM image (scale
20 nm), and size distribution graph of, (b1 and b2) 5OA@Bi2 S3 -90-MM, (c1 and c2)
10OA@Bi2 S3 -90-MM, (d1 and d2) 5OA@Bi2 S3 -90-MG. Source: Reproduced with permission
from Malca et al. [24]/American Chemical Society.

four hours to achieve the Pt/m-Al2 O3 -400 (Figure 2.4a). The same method was used
for the preparation of Pd/m-Al2 O3 -400 by using palladium metal precursors.
The scanning transmission electron microscope at high angle annular dark-field
(HAADF-STEM) was conducted to examine the pore structure and morphology
of Pt/m-Al2 O3 -400 NPs. It shows the Pt NPs average size ∼4.8 nm dispersed on
Meso-Al2 O3 support (Figure 2.4b,c).
Notably, as-synthesized nanocatalyst has a high surface area (up to 465 m2 g−1 ),
uniform pore spreading with a pore size ∼3.9 nm, and good distribution of noble
metal species. The catalytic activity of Pt/m-Al2 O3 -400 and Pd/m-Al2 O3 -400 materi-
als was further used to hydrogenate nitrobenzene to aniline. This mechanochemical
synthetic strategy would deliver a potential approach for synthesizing mesoporous
alumina-loaded noble metal or transition metal catalyst.
36 2 Strategies for the Preparation of Nanocatalysts and Supports Under Solvent-Free Conditions

(a)
Pt nanoparticles

400 °C

in air
0.5 h

Pt/m-Al2O3-400

P123 Al(OCHCH3CH3)3 Pt(acac)2 Ball milling


(c) 40
(b) 4.8 nm
35
30
Frequency (%)

Pt NPs 25
20
15
10
5
0
m

4. 4.8 m

–8 nm
m

5. 5.5 m
m

nm
7– n

1– n

2– n
4– n

8– n
5– n

9– n
2. 2.7

4. .1

6. 6.2

7.6 7.6
3. 3.4

6. 6.9

.3
4
2–

50 nm
2.

Figure 2.4 Synthesis and characterization of Pt/m-Al2 O3 -400. (a) Schematic


representation of the catalyst, (b) HAADF-STEM image, (c) the particle size distribution of
Pt NPs. Source: Reproduced with permission from Nie et al. [25]/Elsevier.

The graphitic carbon nitride (g-C3 N4 ) is having good electronic and chemical
properties and thermal stability. The boron NPs-carbon nitride is applied in the
sustainable process using sunlight as a catalyst for photocatalytic conversion of
alcohols to ketones. Fernández-García and coworkers reported the preparation of
B/g-C3 N4 using the ball milling process via a mechanochemical approach [26].
The g-C3 N4 was achieved by the calcination of melamine at 400 ∘ C. Further, the
as-prepared g-C3 N4 product mixes with various boron sources by using the Retsch
PM100 planetary ball mill reactor, along with reaction chamber and SS balls.
Following this synthetic strategy, a series of boron-based carbon nitrides, such
as 1 B/g-C3 N4 , 2.5 B/g-C3 N4 , 5 B/g-C3 N4 , and 10 B/g-C3 N4 , have been prepared
(Figure 2.5a). The TEM/ HRTEM analysis of 2.5 B/g-C3 N4 and 10 B/g-C3 N4
composites confirmed the boron’s presence in this composite material (isolated
boron particle range is 5–20 nm) (Figure 2.5b–d).
The as-prepared distinct composites catalysts were delivered high activity
and selectivity for the catalytic photo-oxidation of 2-propanol to acetone. The
2.2 Mechanochemistry 37

(a)

B:

N N N N
N N N N N N N N N N N
N N N N N N N N N
N N N N N N
N N N N Mechanochemical N N N N N N
N N N N N N treatment

(b) (c)

200 nm
20 nm

(d) (e)

200 nm 20 nm

Figure 2.5 (a) Schematic representation of the synthesis of B/g-C3 N4 , (b) TEM image of 2.5
B/g-C3 N4 , (c) HRTEM image of 2.5 B/g-C3 N4 , (d) TEM image of 10 B/g-C3 N4 , (e) HRTEM
image of 10 B/g-C3 N4 samples (arrow shows the boron particles). Source: Reproduced with
permission from Caudillo-Flores et al. [26]/Royal Society of Chemistry/CC BY-NC 3.0.

as-prepared 2.5 B/g-C3 N4 display the best activity for the alcohol oxidation to the
respective ketone.
Recently, Moores and coworkers have prepared the ultrasmall nickel phosphide
NPs (NiP-NPs) by the mechanochemical ball milling process using a zirconia jar
[27]. The synthesis of NiP-NPs was achieved by the general approach of milling
the nickel chloride, sodium phosphide, and various capping ligands. This substance
38 2 Strategies for the Preparation of Nanocatalysts and Supports Under Solvent-Free Conditions

(a) (b)

20 nm 10 nm

(c) 35% n = 318 (d)


30%
Entry Sample name Ligand Ligand Particle size
25% equivalents (nm)
Frequency

1 NixPy-2-C15 Pentadecylamine 2 3.5 ± 1.2


20% 2 NixPy-5-C15 Pentadecylamine 5 4.8 ± 2.1a
15% 3 NixPy-2-C16 Hexadecyl amine 2 3.5 ± 0.8

10% 4 NixPy-5-C16 Hexadecyl amine 5 5.4 ± 1.3


5 NixPy-2-C17 Heptadecyl amine 2 5.0 ± 2.1
5%
6 NixPy-5-C17 Heptadecyl amine 5 2.8 ± 0.6
0%
0 1 2 3 4 5 6 7 NixPy-10-OAm Oleyl amine 10 3.0 ± 0.8

Particle diameter (nm)

Figure 2.6 (a and b) TEM images, (c) size distribution graph of NiP NPs (Nix Py -5-C17),
(d) effect of alkyl amine ligand equivalency on particle sizes. a The mixture of free and
combined particles was noted by bright-field TEM image. Source: Reproduced with
permission from Fiss et al. [27]/American Chemical Society.

was milled in a zirconia jar with a zirconia ball at 30 Hz for 30 minutes. Then the
obtained product was aged at room temperature for 18 hours, to get the final product
of NiP-NPs after washing and drying.
For the mechanochemical synthesis of NiP-NPs, the heptadecyl amine, a
C17-substituted amine, has been used as a capping ligand to form desired prod-
ucts. The diameters of NiP-NPs were confirmed by TEM, which was found
to be 5.0 ± 2.1 nm and 2.8 ± 0.6 nm (for 2 and 5 mol equivalent), respectively
(Figure 2.6a–c). The table (Figure 2.6d) described the various types of NiP-NPs and
the size of NPs with the various ligand equivalents. Notably, the mechanochem-
ical synthesis technique of NiP-NPs production can be scaled up to 2.5 gm. The
prepared ultrasmall NiP-NPs were used for the photocatalytic hydrogen evolution
reaction (HER).

2.2.2 Mortar and Pestle


Another simple mechanochemical method is the traditional grinding method,
which can be done using a simple mortar and pestle [19]. In this grinding process,
the chemical changes start within the crush solid raw materials. The first grinding
2.2 Mechanochemistry 39

(a) (b) (c)

50 nm 50 nm

(d)

NaOH

+
Ground Crystallization
Ni, Mg, Al nitrate Na2CO3

Ni-Mg-Al LDH

Figure 2.7 (a) The sample of α-Ni(OH)2 , (b and c) TEM images of α-Ni(OH)2 . Source:
Reproduced with permission from Cui et al. [28]/Elsevier. (d) Synthesis route of the
Ni-Mg-Al catalyst. Source: Reproduced with permission from Du et al. [29]/Elsevier.

was used to form foodstuffs; then, it treated other categories of ingredients, like
minerals, paints, and medicines. Some of the solvent-free NPs synthesis using
mortar and pestle has been discussed in this part.
The large-scale production of α-Ni(OH)2 nanosheets has been reported using
the mortar and pestle via the mechanochemical process [28]. The nickel nitrate
hexahydrate (Ni(NO3 )2 ⋅6H2 O) and morpholine (C4 H9 NO) have been used as
starting materials for nanosheet preparation. In the first step, the metal precursor
(Ni(NO3 )2 ⋅6H2 O) was ground by using the mortar and pestle, and then C4 H9 NO
was added to this powder material to obtain the final product α-Ni(OH)2 nanosheets
(Figure 2.7a).
It is noted that the TEM image shows the thin and smooth Ni(OH)2 nanosheets
with the size of more than 100 nm, along with 4–10 nm nanoflakes also observed
(Figure 2.7b,c).
The α-Ni(OH)2 nanosheets have been used for the supercapacitors due to their
specific capacitance, high stability, and outstanding rate capacity. Hou and cowork-
ers synthesized the Ni–Mg–Al layered double hydroxide (LDH) via a solvent-free
approach using a mortar and pestle grinding [29]. For the synthesis of Ni–Mg–Al
LDH, the mixture of hydrated metal nitrate of Ni(NO3 )2 ⋅6H2 O, Mg(NO3 )2 ⋅6H2 O, and
Al(NO3 )3 ⋅9H2 O is used. All precursors were mixed and formed the mixture powder,
then in this mixture, sodium hydroxide (NaOH) and Na2 CO3 have been added. After
the all-mixture grinding, the obtained product was placed in an autoclave at 120 ∘ C
for one day. The residue was washed by deionized (DI) water and dried in a vacuum
oven the next day. Furthermore, the obtained Ni3.6 Mg2.4 Al2 (OH)16 CO3 (solvent-free)
product was calcinated at 700 ∘ C to achieve the Ni3.6 Mg2.4 Al2 O9 under solvent-free
40 2 Strategies for the Preparation of Nanocatalysts and Supports Under Solvent-Free Conditions

conditions. After that the corresponding product reduced at 850 ∘ C in hydrogen (H2 )
environments, to achieved the final catalyst (Ni3.6 /Mg2.4 Al2 O5.4 ) (Figure 2.7d).
The solvent-free precursor Ni3.6 Mg2.4 Al2 (OH)16 CO3 shows a high crystallinity
and TEM analysis shows the uniform, hexagonal plates with diameters of 25–30 nm.
The prepared Ni3.6 Mg2.4 Al2 (OH)16 CO3 by the solvent-free method is sharper and
more robust than the Ni3.6 Mg2.4 Al2 (OH)16 CO3 (prepared by the coprecipita-
tion (CP) method). It is noted that the Ni3.6 Mg2.4 Al2 (OH)16 CO3 (solvent-free)
is more crystalline than the Ni3.6 Mg2.4 Al2 (OH)16 CO3 (CP). The as-synthesized
Ni3.6 Mg2.4 Al2 (OH)16 CO3 (solvent-free) is highly reactive and stable for the reaction
of hydrogenolysis of D-sorbitol below mild conditions. The high activity might
be accredited to its uniform dispersal throughout the reaction procedure and
small-sized particles, and Ni clusters provide more available solid basic places and
hydrogenation reactive places for starting materials.
Over the last few decades, the bimetallic nanoparticles (BMNPs) have been inves-
tigated for various applications, such as homogeneous and heterogeneous catalysis,
nanomedicine, imaging, and several other research areas. The core–shell NPs are the
class of NMs that attract attention due to their unique properties and a wide range
of applications in catalysis, biology, materials chemistry, and sensors [8].
Choi et al. synthesized the core–shell and alloy Ag–Cu BMNPs using the
mechanochemical mortar and pestle grinding process (Figure 2.8) [30]. In the first
part, the synthesis of bimetallic Ag–Cu NPs, the mixture of copper(II) formate
tetrahydrate, and silver nitrate are dissolved in a mixture of hexylamine/xylene
(1 : 5 v/v). Later, the formed mixture was mixed with sodium sulfate powder by
using mortar and pestle. Furthermore, the formed powder product was calcinated
using a tube furnace at 300 ∘ C in a nitrogen (N2 ) environment for one hour and was
washed by an aqueous hydrazine solution to obtain the final Ag@Cu core–shell
NPs. The same procedure is used for forming the AgCu homogeneous alloy NPs;
here the higher temperature is essential for the calcination process (calcination
done at 400 ∘ C) (Figure 2.8a).
The TEM image observation of Ag@Cu core–shell shows the flower-like mor-
phologies; the core–shell NPs average diameter is 37.8 ± 8.3 nm with an average core
size is 21.4 ± 5.9 nm and the shell thickness is 8.2 ± 2.1 nm (Figure 2.8b). TEM image
of AgCu alloy NPs average diameter of 27.7 ± 5.4 nm displays the particles in spher-
ical morphology (Figure 2.8c).
The morphology of the core–shell versus alloy structures can be tuned by only
regulating the annealing temperature. The mechanochemical synthetic approach
using mortar and pestle is simply too scalable for large-scale NPs production. The
produced BMNPs might be useful in various fields, such as electronics, biomedical
research, and catalysis.
In a similar line, Lee and coworkers synthesized the Co@NC core–shell nanos-
tructures from metal-organic frameworks via a mechanochemical process by using
a mortar and pestle [31]. In this synthesis, the three solid precursors, such as
Zn(ac)2 , pyrazole (C3 H4 N2 ), Co(NO3 )2 , are mixed in a mortar and grind around
25 minutes (pink-colored observed after product formation) to obtain the dried
powder of synchronized self-assembled metal organic framework (MOF) structures
2.2 Mechanochemistry 41

(a) Cu formate (b)


AgNO3
Hexylamine
Xylene

Sodium sulfate
50 nm
Mix

(c)

300 °C 400 °C
Bake
Heat treatment
under N2
50 nm

Wash (d)
Hydrazine
aqueous solution

Ag@Cu core–shell AgCu alloy 100 nm

Figure 2.8 (a) Schematic synthesis representation of AgCu bimetallic NPs, (b) TEM image
of Ag@cu core–shell NPs, (c) TEM image of AgCu alloy, (d) scanning electron microscope
(SEM) image of pure Cu NPs. Source: Reproduced with permission from Choi et al.
[30]/Royal Society of Chemistry.

Znx Co1−x (C3 H4 N2 ). After the pyrolyzed product is formed at 900 ∘ C, pyrazoles
decompose into N-doped carbon, and carbo-reduction of the Co2+ ions occurs
instantaneously. These outcomes achieved pure Co@NC-MOF-900 electrocatalyst
(Figure 2.9).
The as-prepared electrocatalyst shows that the average particle size is 17 nm (with
a range of 10–20 nm). The Co@NC-MOF-2-900 electrocatalyst has been used for the
oxygen reduction reaction (ORR).
In this section, we have explained the different examples of NPs, containing the
metal trihalide NPs. Jana et al. synthesized the trihalide NPs (APbBr3 , A = FA,
MA, and Cs) using mortar and pestle via the mechanochemical process [32].
In this NPs synthesis, in the first step, MABr and FABr (FA = Formamidinium
and MA = Methylammonium) have been prepared. The preparation of MAPdBr3
(orange-colored solid product) can be obtained by mixing and grinding MABr
(CH3 NH3 Br), PdBr2 , and n-octyl ammonium at room temperature for 10 minutes.
Then the obtained solid was centrifuged with acetone and dried in a vacuum oven.
The same procedure was used to produce FAPbBr3 and CsPbBr3 with FABr and
CsBr correspondingly instead of MABr.
42 2 Strategies for the Preparation of Nanocatalysts and Supports Under Solvent-Free Conditions

Physical Left at room


grinding temperature

Zn (ac)2 + Pyrazole + Co(NO3)2 [Zn(py)(Co)] [Zn(py)(Co)]


Scratch out
the powder

Pyrolysis at Scratch out


900 °C the powder

Under Ar
2h

Co@NC-MOF-2-900 °C [Zn(py)(Co)] [Zn(py)(Co)]

Figure 2.9 Synthesis route of Co@NC-MOF-900 electrocatalyst. Source: Reproduced with


permission from Peera et al. [31]/John Wiley & Sons, Inc.

The as-prepared CsPbBr3 , MAPbBr3 , and FAPbBr3 materials show the square
shape of NPs in addition to rectangular, spherical, and parallelogram shapes NPs,
correspondingly. The TEM analysis shows the 6–8 nm average diameter of MAPbBr3
NPs, and the effective Bohr diameters of MAPbBr3 and CsPbBr3 are 4 and 7 nm,
correspondingly.
Mazonde et al. synthesized the Co-based zeolite catalyst using the mechanochem-
ical process [33]. The Co-based zeolite synthesis was achieved by mixing the
boehmite (Al2 O3 ), Co/SiO2 , NaOH, and ethylenediamine (EDA). This mixture
was ground (using mortar and pestle), and then the mixture was heated in the
autoclave for 48 hours at 180 ∘ C. The obtained material was calcinated at 550 ∘ C to
get final zeolite catalyst (Co@NaZSM-5) (Figure 2.10). This produced catalyst from
SiO2 with the pore size of 10 and 50 nm, mentioned as the Co@NaZSM-5(10) and
Co@NaZSM-5(50).

Co/SiO2 NaOH
EDA

AIO(OH)
HO
Al
O

Co@NaZSM-5

Grinding

Figure 2.10 Schematic representation of the synthesis of Co-based zeolite. Source:


Reproduced with permission from Mazonde et al. [33]/Royal Society of Chemistry.
2.2 Mechanochemistry 43

(a)

Mortar–pestle milling

Carbon

Washing

Drying

PdAg alloy NPs PdAg/C

(c)
(b) 30
Frequency (%)

20
2 nm

10

0
50 nm 2 3 4 5 6
Particle size (nm)

Figure 2.11 (a) Schematic representation of the PdAg alloy synthesis, (b) TEM image
(insets image is the HRTEM images), and (c) particle size distributions of Pd5 Ag5 /C. Source:
Reproduced with permission from Xu et al. [34]/Elsevier.

This synthesis method offers a new protocol for emerging zeolites and can prove
crucial for the additional advanced applications of metal@zeolite nanocatalyst.
In another work, Xu et al. synthesized the PdAg alloy NPs by using the
mechanochemical approach [34]. In a typical synthesis of PdAg alloy, PdCl2 , and
AgNO3 are mixed by homogeneous grinding with carbon for 30-minutes process.
Then in this mixture, metal/NaOH ( with ratio 1 : 4 mol) was added and again
ground the mixture for 30 minutes, followed by the addition of sodium borohydride
(NaBH4 ) ground this complete mixture for more than 20 minutes. Further, the
product was washed by a mixture of solvent acetone and DI water and achieved the
final (PdAg/C) product (Figure 2.11a). The same procedure was used to synthesize
the PdAg/C series with the various molar ratios of Pd and Ag metal precursors.
It is noted that when the Ag content increases in the PdAg/C NPs, the disper-
sal of NPs is progressively better (Figure 2.11b). Notably, the Pd5 Ag5 /C NPs showed
homogeneous distribution on the carbon support, and HRTEM analysis shows the
44 2 Strategies for the Preparation of Nanocatalysts and Supports Under Solvent-Free Conditions

lattice fringe of 0.231 nm (Figure 2.11b insect). The TEM image and size distribution
display that the NPs of Pd/C, Pd9 Ag1 /C, Pd5 Ag5 /C (Figure 2.11b,c), Pd1 Ag9 /C, and
Ag/C had average sizes of 4.8, 5.3, 4.9, 4.5, and 4.7 nm, correspondingly. The same
procedure has been employed to prepare Pd5 Ag5 /C NPs without the use of NaOH;
notably, a bit larger size of 9.74 nm has been observed. The Pd9 Ag1 /C has shown
excellent activity for electrocatalytic reactions, like the electrocatalytic HER.

2.3 Thermal Treatment


This part of the chapter discusses the selected and important solvent-free synthe-
sis methods of NPs and supports by using the thermal treatment. Various types of
NPs synthesis methods have been highlighted in this thermal treatment method,
including simple thermal treatment, calcination, pyrolyzing, annealing treatment,
and thermal deposition methods.

2.3.1 Simple Thermal Treatment


The solvent-free simple thermal treatment is useful for a metal precursor to metal
NPs, metal oxides, etc., also the synthesis of several structures such as the disks,
wires, platelets, and rods. The products prepared by the solvent-free thermal synthe-
sis method are monodisperse NPs that are due primarily to minimal inter-particle
collisions [35].
Wang et al. synthesized the sulfur and nitrogen-co-doped carbon nanoparticles
(SNCNPs) using a one-pot solid-phase thermal treatment method [36]. The SNCNPs
preparation was achieved by the thermal treatment on glutathione in SS autoclave
at 260 ∘ C for one hour. The formed product was centrifuged with water to obtain
SNCNPs (Figure 2.12a). The TEM image shows that the SNCNPs were spherical,
with an average diameter of 2.93 nm (Figure 2.12b,c); the HRTEM image displays
the lattice fringe of 0.203 nm (insets in Figure 2.12b).
The as-prepared SNCNPs have a bright blue fluorescence located at 442 nm and
a relatively high quantum yield of 39.9%, which can serve as a fluorescent probe
for detecting Hg2+ in real samples like the lake water samples with high sensitivity
and selectivity. This existing synthetic plan affords a simple and solvent-free way
of using modest chemicals and can be employed to prepare various carbon NMs in
bioimaging, energy renovation and storing, catalysis, and biosensors.
Shahsavani et al. synthesized the AgBr NPs by using the solvent-free thermal
treatment [37]. In the first step Ag(Brcatsc)(NO3 ) precursor has been prepared
using the AgNO3 and thiosemicarbazone ligand Brcatsc. Further, this prepared
Ag(Brcatsc)(NO3 ) was heated in an electrical furnace up to 600 ∘ C for three hours.
The obtained AgBr NPs (average size under 20 nm confirmed by TEM) were washed
by ethanol for removing the material impurities.
The CuO/TiO2 nanocomposites were synthesized by using the annealing thermal
treatment reported by Li et al. [38]. In a typical preparation of CuO/TiO2 nanocom-
posites, in the first step, TiO2 material was added to the solution of Cu(NO3 )2 ⋅3H2 O
2.3 Thermal Treatment 45

(a) HS
NH2 NH2
H2N
O O COOH HS
HS NH2
H 260 °C NH2
H 2N
OH H O
NH H COOH
1h HS HS
N NH2 NH2
HS OH H2N
COOH
O HS
NH2
Glutathione

(b) (c) 40

30
Frequency (%)

2 nm 20

10

20 nm
0
2.0 2.5 3.0 3.5 4.0
Size (nm)

Figure 2.12 (a) Schematic representation of the synthesis of SNCNPs, (b) TEM image
(insets, HRTEM image), (c) size distribution of SNCNPs. Source: Reproduced with permission
from Wnag et al. [36]/Elsevier.

for the annealing thermal treatment and stirred for 24 hours at room temperature.
Further, the obtained material dried under a vacuum oven and calcined at 400 ∘ C to
get the CuO/TiO2 nanocomposite. The same procedure has been employed for the
preparation of CuO/SiO2 nanocomposites.
The TEM image observation of 20 wt% CuO/TiO2 nanocomposite displays that
NPs range of 50–60 nm and the HRTEM image was displayed that the surface
of TiO2 NPs size is about 40–50 nm enclosed with small NPs of CuO of 8.0 nm
(Figure 2.13a,b). The CuO/TiO2 nanocomposites were used for the photocatalytic
Sonogashira coupling reaction under visible light irradiation. The CuO/TiO2
nanocomposites study offered a green and maintainable approach for the synthesis
of substituted diacetylenes and the excessive probability of photocatalysis in light
encouraged organic syntheses.
Chen and coworkers synthesized the hierarchically structured Cu-based electro-
catalysts with nanowires (NWs) array by the thermal treatment [39]. In the first step
the Cu(OH)2 NWs array on the copper foam (Cu(OH)2 /CF) has been prepared by the
anodization of copper foam (CF). The prepared Cu(OH)2 /CF was kept in a furnace
oven with an air atmosphere at 150 ∘ C for three hours and 200 ∘ C for three hours,
achieving the black-colored CuO/CF electrode. Also, the Cu3 P/CF NWs electrode
was prepared by using the thermal treatment. The NaH2 PO2 and CuO/CF were
placed in the furnace at 300 ∘ C for 60 minutes after the material cooling formed the
Cu3 P/CF electrode (Figure 2.14).
46 2 Strategies for the Preparation of Nanocatalysts and Supports Under Solvent-Free Conditions

(a) (b)
d = 0.233 nm
CuO (2 0 0)

d = 0.347 nm
TiO2 (1 0 1)

50 nm 5 nm

Figure 2.13 (a) TEM image, (b) HRTEM image of 20 w% CuO/TiO2 nanocomposite. Source:
Reproduced with permission from Li et al. [38]/Royal Society of Chemistry.

Anodic Annealing
treatment
R
OE

CF CuO/CF CuO NWs


ER
O

Phosphorization
Potentiostat

HER

Overall water splitting Cu3P/CF Cu3P NWs

Figure 2.14 Schematic representation of the synthesis of Cu-based NWs. Source:


Reproduced with permission from Lu et al. [39]/American Chemical Society.

The prepared CuO/CF NWs are used as effective electrocatalysts for the OER, and
Cu3 P NWs are used as electrocatalysts for the HER. The as-synthesized Cu-based
NM improves the Cu-based NMs list; it also delivers a new plan to ease the synthetic
technique of electrolyzers for water splitting reaction.
Wang et al. synthesized the Cu, N–Co doped carbon nanodisks (N–C NDs)
using the thermal treatment [40]. In the synthesis of N–C NDs catalyst, the
Zn-TPyP-MOFs precursors were prepared in the first step by using the acetic
solution of 5,10,15,20-tetra pyridyl porphine (H2 TPyP), which was stirred with
copper acetate monohydrate (CuAc2 ⋅H2 O) and sodium dodecyl sulfate (SDS), then
formed mixture ultrasonicated. The achieved Cu-TPyP-MOFs product was kept in a
2.3 Thermal Treatment 47

porcelain boat and placed in the tube furnace with an Ar environment at 800 ∘ C, and
then black-colored Cu@Cu-N-C product was obtained. It was the resulting material,
treated with FeCl3 and 0.5 M H2 SO4 solution at 80 ∘ C for 12 hours and washed
by DI water for removing the free bonded Cu NPs or nanoclusters by obtained
Cu@Cu–N–C catalyst. After that, the formed sample product was reannealed at
800 ∘ C for one hour with an Ar atmosphere. The last step received the Cu–N–C NDs
through the bio mimic stomata-like interconnected hierarchical porous topology
that is denoted as Cu–N–C–ICHP (interconnected hierarchical porous). The TEM
image of Cu@Cu–N–C catalyst was observed the black spots of Cu NPs with a
size range of 30–100 nm. The 2D Cu–N–C–ICHP have the high-performance and
methanol accepting electrocatalyst for oxygen reduction reaction (ORR), for 2D
electrocatalysts have the bio mimic pore structure engineering and encourage their
applications in fuel cells, artificial photosynthesis, or also other clean energies.
Panda et al. synthesized the Cu3 N nanostructure by using the thermal treatment
method [16]. The copper acetate and urea were kept on two separate porcelain boats
in a quartz tube and treated in Ar atmosphere at 300 ∘ C for two hours. The obtained
black powder was washed and dried to get the Cu3 N nanostructure (Figure 2.15).
The morphological observation by scanning electron microscope (SEM) image
revealed the cubical Cu3 N NPs with variable sizes of 60–100 nm. The Cu3 N NPs
deposited on nickel foam (NF) and show outstanding activity and constancy when
used as the cathode and anode for HER and OER.
The synthesis of Cu-MOF-74 derived Cu–Cu2 O–C nanocomposites was achieved
by using a thermal treatment method [41]. After the preparation of Cu-MOF-74
in the first step, it is placed on the ceramic boat and kept in a furnace oven with
an N2 environment at a rising temperature up to 1000 ∘ C for one hour. The formed
product was ultrasonicated with NaOH solution at 40 ∘ C for 45 minutes. After
the ultrasonication, the dried product was placed in a ceramic boat and heated at

Ar NH3 Ar
Ar NH3

Cu3N
Urea Cu(OAc)2

Calcination at 300 °C
O

H2N NH2 NH3

N-Source Cu(OAc)2

H2 O2

H2O

Figure 2.15 Schematic representation of the synthesis of Cu3 N cube nanostructure.


Source: Reproduced with permission from Panda et al. [16]. American Chemical Society.
48 2 Strategies for the Preparation of Nanocatalysts and Supports Under Solvent-Free Conditions

1000 ∘ C in the furnace with an N2 environment to achieve the final Cu–Cu2 O–C
nanocomposites.
The SEM image of Cu–Cu2 O–C nanocomposites observed the spherical particles
with a diameter of around 200 nm, and the size distribution observed the Cu NPs
size is 1.7 nm. The Cu–Cu2 O–C nanocomposite efficiently catalyzed the pheny-
lacetylene coupling to form 1,4-diphenyl butadiyne underneath the visible-light
irradiation using CO2 as the oxidant. This work not only provides an alternative
yet very promising strategy for the light-initiated Glaser coupling reaction, but also
highlights MOFs’ imperative role as precursors to synthesize carbon-supported
metal/metal oxide nanocomposites for heterogeneous cooperative catalysis.
Zhu and coworkers synthesized the Au–Cu alloy NPs by using the thermal treat-
ment [42]. The Au2.6 Cu0.4 @ZrO2 NPs synthesized using the ZrO2 powder added in
a mixture of HAuCl4 and Cu(NO3 )2 aqueous solutions underneath magnetic stir-
ring at ambient temperature. In this mixture, the aqueous solution of lysine was
added with vigorous stirring, at that time pH was 8–9. NaBH4 was added slowly;
later, this solution mixture was aged for 24 hours. Then achieved dried powder was
thermally treated with an H2 and Ar environment at 450 ∘ C for a half-hour to obtain
the Au2.6 Cu0.4 @ZrO2 NPs catalyst. The Au–Cu NPs average size is around 5 nm (by
TEM) and X-ray photoelectron spectroscopy (XPS) analysis confirmed the Au con-
tent more than the Cu, the Au2.6 Cu0.4 alloy NPs. The as-prepared alloy NPs have an
excellent photocatalytic activity for the nitroaromatics reduction to respective ani-
line, the bimetallic character of NPs tends to the selective reduction.
Xiang et al. synthesized the α-Fe2 O3 (hematite) NPs by using thermal treatment
[43]. In the typical synthesis of α-Fe2 O3 NPs, the first step ferrous sulfate waste
(received from industry) and pyrite (FeS2 ) was heated at 180 ∘ C for four hours, then
crushed the mixture product with the 12 : 1 ratio of ferrous sulfate waste and pyrite
to form 200 mesh powder products. Furthermore, the mixed materials were kept in
the tube furnace with an N2 environment and heated at 550 ∘ C at a heating rate of
10 ∘ C min−1 for 30 minutes to achieve the α-Fe2 O3 NPs.
The hematite NPs was shown size around 30.8 nm before the photo-Fenton reac-
tion and 38.7 nm after the photo-Fenton reaction. Also, SEM image observed spher-
ical morphology of α-Fe2 O3 NPs and size around 50 nm. The as-prepared hematite
NPs observed an admirable catalytic role for the decolorization of methyl orange by
the photo-Fenton reaction.
Takanabe and coworkers synthesized the quaternary metal sulfide NPs by metal
oxide and thiourea via the thermal treatment method [44]. The synthesis of qua-
ternary metal sulfide NPs was achieved in a semi closed crucible with an open-air,
by using the metal oxides such as copper oxide, gallium oxide, and indium oxide
with stoichiometric millimolar ratio (1 : 2 : 3) and mixed with several quantities
of thiourea (thiourea/Cu ratios of 8, 32, and 210). The crushed mixture material
was placed in a semi closed crucible and kept in a muffle furnace at 300 ∘ C for
12 hours. The resultant product was washed, and dried, to achieve the quaternary
(CuGa2 In3 S8 = CGIS) metal sulfide NPs. The specific amount of Na2 S used for the
additional quaternary mixture preparation with the thiourea.
2.3 Thermal Treatment 49

(a)

Glucose Mixing Calcining

CoCl2 Thiourea
Co9S8 nanopartcles
Precursors Intermediates entrapped, N, S-codoped
mesoporous carbon

(b) (c)

200 nm

Figure 2.16 (a) Schematic representation of NPs synthesis, (b) SEM image, (c) TEM image
of G2.0 T1.0 Co0.3 -900. Source: Reproduced with permission from Yu et al. [45]/Elsevier.

Also, the CGIS was prepared with the ratio of Cu:S = 1 : 32 at the temperature
of 200 and 550 ∘ C in the presence of Na2 S, denoted as CGIS-200S and CGIS-550S
correspondingly. The 13 C and 15 N cross-polarization magic angle spinning dynamic
nuclear polarization (CP-MAS DNP) for thiourea was heated up to 550 ∘ C for
four hours, and the outcomes confirmed the development of a g-C3 N4 structure.
TEM image observed the CGIS-300S NPs ranging in sizes from 5 to 10 nm. The
as-prepared CGIS sample was investigated for photocatalytic applications.
Yu et al. synthesized the Co9 S8 NPs N, S-co-doped mesoporous carbon using the
calcinated carbonization process [45]. The mixture of glucose, thiourea, and CoCl2
was prepared mechanically. The formed intermediate materials were carbonized
at various temperatures for two hours. The achieved products were mentioned as
Gx Ty Coz -Temp, whereas x, y, and z stand for the masses of glucose, thiourea, and
CoCl2 , correspondingly, and temp stands for the calcination reaction temperature
(Figure 2.16a).
The SEM image observed the bulky mesoporous morphology of the
G2.0 T1.0 Co0.3 -900 sample, insides of high Co, N, and S fillings (Figure 2.16b).
The TEM image observed (Figure 2.16c) the Co9 S8 NPs uniformly distributed in
50 2 Strategies for the Preparation of Nanocatalysts and Supports Under Solvent-Free Conditions

the carbonaceous medium, with NPs size range 20–50 nm. The new mesoporous
G2.0 T1.0 Co0.3 -900 catalyst displays excellent electrocatalytic activities for HER.
Davar and coworkers synthesized the cobalt oxide (Co3 O4 ) by using the thermal
treatment method [46]. In the synthesis of Co3 O4 NPs, the first step Co(salophen)
sample has been prepared. Then Co(salophen) was heated up to 500 ∘ C with air envi-
ronment, with a heating rate of 30 ∘ C min−1 for five hours to obtain the Co3 O4 NPs
(diameter = 30 nm). The optical property of Co3 O4 NPs changes to a short wave-
length, and the blue shift phenomenon may be attributed to the quantum effect.
Jagirdar and coworker synthesized the Pd(0), PdS, and Pd@PdO core–shell NPs
by solventless thermolysis of a Pd–thiolate cluster powder [35]. In the typical syn-
thesis of NPs, the Pd–thiolate powder was heated at 298 ∘ C with air environment for
three hours, to achieve Pd(0) NPs. For the synthesis of Pd@PdO core–shell NPs, Pd
thiolate powder is heated at a higher temperature such as 325 ∘ C, and then 425 ∘ C
with an air environment. The PdS NPs synthesized by the Pd thiolate powder was
heated at 430 ∘ C with an Ar environment for three hours. The TEM image of Pd(0)
NPs observed the oval-shaped NPs with sizes ranging from 13 to 22 nm. TEM image
of Pd@PdO core–shell and PdS NPs shows the polygonal-shaped particle with a size
ranging from 20 to 100 nm and spherical shape with a size of 9.27 ± 0.9 nm corre-
spondingly. The Pd–thiolate complex powder is a useful precursor for synthesizing
several Pd nanophase materials like Pd(0), PdS, and Pd@PdO core–shell NPs by sol-
ventless thermolysis methods.

2.3.2 Thermal Decomposition


The thermal decomposition occurs at the metallic ion and force of reaction along
with ligands in the coordination compounds effects on the pressure and tempera-
ture to form monodispersed NPs [47]. The thermal decomposition method is used to
prepare the monodispersed NPs, which are stable and have small particle sizes. The
thermal decomposition method has some benefits, like modest process cost-effective
and easy to get the high pure compound product for the capable route of industrial
base applications [48].
In the year 2013, Mohadesi et al. synthesized the HgO nanostructure using the
simple thermal decomposition method [49]. The first step involves the as-prepared
Hg(OAc)2 nanostructures kept in a silicon boat, placed in a tube furnace, heated
with an air environment at 350 ∘ C for two hours to obtain the HgO nanostructure.
Subsequently, the formed product was cooled to room temperature certainly, and the
precipitations form achieved the HgO NPs. The crystallite of HgO NPs considered
diameter size as 85 nm by the Scherrer equation calculation.
Goudarzi et al. synthesized the ZnO NPs by the ball milling-thermal deposition
process [50]. The ZnO NPs were prepared from the Zn(NO3 )⋅6H2 O, and cochineal
mixed by a ball milling reactor using the zirconia grinding media. Furthermore, the
obtained powder mixture was calcinated at 600 ∘ C for three hours, to achieve the
spherical shape ZnO NPs with a size range of 15–25 nm.
The as-prepared ZnO NPs have been used for the photocatalytic decolorization
of methylene orange. The preparation strategy of ZnO NPs includes numerous
2.3 Thermal Treatment 51

advantages, such as it is a controllable, solvent-free, template-free, and economical


synthesis method. A similar kind of ZnO NPs synthesis was reported by Ranjbar et al.
using Zn(OAc)2 powder via the solvent-free simple thermal deposition method [51].
In the year 2018, Togashi et al. synthesized the monodisperse Cu NPs by the ther-
mal decomposition method [52]. In the first step, the prepared OA-Cu(ox) complex
was mixed with OA by glass centrifugation. That obtained mixture was kept on a
hotplate stirrer at 140 ∘ C for 20 minutes for regular mixing, then this mixture was
stirred at a higher temperature around 200–260 ∘ C for one hour, to get final cata-
lyst of Cu NPs. The TEM image observed the as-synthesized Cu NPs size changes
with increasing the heating temperature; the spherical Cu NPs morphology with a
size ranging from 10 to 30 nm at 220 ∘ C, and the Cu NPs size reduces to 13.8 nm at
260 ∘ C. The thermal decomposition of alkylamine–oxalate complexes is established
as an innovative method for the synthesis of monodisperse metal NPs.
Similarly, Togashi et al. synthesized the Fe3 O4 NPs by using the thermal decompo-
sition method [53]. The oleylamine-coordinated iron oxalate complex (OA-Fe(ox))
precursor has been prepared first and then heated up to 300 ∘ C without any addi-
tions achieving the spherical shape (confirmed by TEM) Fe3 O4 NPs (diameter of
10.2 ± 1.2 nm) product.

2.3.3 Microwave Heating Energy


The microwave irradiation technique is the best energy source for material
preparation in various areas, especially for the medicinal chemistry and drug
discovery area, because the microwave technique reduces the reaction time [54, 55].
Microwave chemistry is becoming a vital tool for the fast-paced, time-delicate arena.
The solvent-free microwave heating is a much more fast and thus higher output
alternative to the efficient and scalable heating method [56].
For example, the synthesis of metal NPs on carbon nanotube and graphene
was achieved by using the solvent-free microwave reaction. Lin et al. synthesized
the metal and metal oxide NPs by using solvent-free microwave heating [56].
In this synthesis, the multiwalled carbon nanotube (MWCNT) powder and silver
acetate powder with a specific ratio are mixed using mortar and pestle. Then
this mixer powder is kept in a microwave reactor quartz vial. In this synthesis,
the CEM discovered a microwave reactor (300 W) used at 300 ∘ C for 60 minutes.
The obtained product was dried overnight with air or N2 environment to achieve the
Ag-MWCNT material. Also, some other metal or metal oxide (Co, Ni, Au, CoO,
MnO, etc.) decorated MWCNT, via the same synthesis method with different
metal salts with similar quantities. Notably, the Ag NPs diameter ranging from 20
to 100 nm was observed by SEM analysis. Also, other prepared NPs showed the
various size distributions; Ag (6.4 ± 2.5 nm), Pd (8.7 ± 2.3 nm), Pt (3.1 ± 0.9 nm),
Au (9.7 ± 5.4 nm), Ni (5.6 ± 2.3 nm), Co/CoO (65 ± 41 nm; most in the clusters
form), Fe3 O4 (11 ± 3.7 nm), TiO2 (2.4 ± 0.8 nm), and MnO (36 ± 15 nm), these NPs
were prepared at 300 ∘ C, with 50 W power level. The as-synthesized NPs are used
in various fields, such as electrical, composite, energy, catalysis, and biological
applications.
52 2 Strategies for the Preparation of Nanocatalysts and Supports Under Solvent-Free Conditions

Ni et al. synthesized the Ru NPs-supported carbon nano tube (CNT) by using the
microwave-assisted synthesis method [57]. In this synthesis, the CNT and Ru3 (CO)12
were mixed using a grinding process to form uniform mixing, then formed mix-
ture placed in the quartz-tube reactor with Ar environment at room temperature
for two hours. After that, this reactor was kept in the domestic microwave oven at
2.45 GHz with an 800-W power level. The obtained Ru/CNT product cooled at room
temperature with Ar gas. The achieved Ru NPs distributed on the support of car-
bon nanotubes are spherical with the size of 2–4 nm. The as-synthesized Ru/CNT
catalyst is used for the cinnamaldehyde hydrogenation reaction.

2.4 Plasma-Assisted Methods


The nanocatalysts and supports have been a core part of boosting chemical reac-
tions in the field of industries and academics [58]. Various synthesis strategies for
nanocatalysts and supports can be found in nanotechnology, but strategies have dif-
ficulty to maintaining their active properties and need a solvent to prepare catalysts
[59]. The plasma-assisted method has resolved the critical limitations of the previ-
ously reported conventional synthesis strategies [60]. Nowadays, the plasma method
has been regarded as a benign and environment-friendly method, with no need for
additional solvent, stabilizer, and surfactant for the nanocatalysts preparation. As
we know, matter has three standard states like solids, liquid, and gas; after that, the
fourth one is plasma, and the fifth one is Bose–Einstein condensate (BEC) states that
existed in the universe [61]. In 1928, Irving Langmuir described the term Plasma,
which can be generated from the collision between gases constituent by increas-
ing their sufficient energy and later denoted as an ionized gas of the “fourth state
of matter.” It promotes the dissociation and ionizes in the form of atoms, molecules,
radicals, electrons dissociate and ionize in atoms, molecules, radicals, electrons, neu-
tral particles, ions species (positive and negative), and electrically neutral [60–62].
Even the plasma state has been shown as the electrically active or conductive mate-
rial. The figure (Figure 2.17) mentioned that the first step of plasma is converting the
feed materials into atoms or molecules at the gas phase to make nanocatalysts and
supports [63]. Before the vaporizing temperature must be above the feeding materi-
als boiling point and residence time depends on feeding materials particle size. The

Feed Primary
particles particles

Vapor

Nanoparticles/
supports

15 000 K 500 K

Figure 2.17 Process of particles formation in a plasma state. Source: Adapted from Zhao
et al. [63].
2.4 Plasma-Assisted Methods 53

plasma process is estimated that 8 milliseconds (ms) is required for 25 μm particles,


and 50 ms is required for 100 μm particles. The particles will be formed through
nucleation of the vapor with particle size, morphology, and crystalline form con-
trolled through a fast-quenching gas process [63].
The excellent source and properties of plasma employed for the preparation of
nanocatalyst and supported with good dispersion of tiny particles size on supporter,
and its catalytic activity, selectivity, and stability are significantly improved. Due to
the versatile properties of the plasma, it is used for many applications like materials
science, agriculture, and microelectronics industries, and even in many emerging
areas such as medical fields [64–67]. Plasma is divided into two parts, based on the
electron’s internal energy, i.e. thermal Plasma and cold Plasma [62]. This section
discusses the plasma types concerning the solvent-free plasma-assisted methods for
employing the nanocatalyst and supports synthesis with maintaining their selectiv-
ity, activity, stability, and surface.

2.4.1 Thermal Plasma Method


Nowadays, thermal plasma is employed in different fields of material chemistry,
polymers, and biomedicals. Notably, it is included in mineral processing, waste treat-
ment, and gasification, the synthesis of NPs and supports, and the preparation of
SACs [58, 68]. In the thermal plasma synthesis method, the plasma’s prominent
role is to provide a concentrated heat source. Because of (pressure > 10 kPa), the
gas and the electron temperatures are comparable and are of many thousands of
degrees. Starting materials are typically atomized at these temperatures, and the
powder synthesis occurs during condensation of the plasma when the gases cool
[62]. The thermal plasma method plays a central role in preparing various nanocat-
alysts and supports in solvent-free conditions with environmental sustainability.
In this section, the bimetallic oxide alloy, PtPd oxide NPs were homogenously
coated over the surfaces of the exfoliated two-dimensional (2D) MXene Ti3 C2 Tx
nanosheets [PtOa (a = 0, 1, 2) PdOb (b = 0,1) @Ti3 C2 Tx ], which is efficiently used for
the electrocatalytic water-splitting reaction (Figure 2.18a) [69]. He and coworkers
have synthesized nanosheets via plasma irradiation and provided the power of
200 W at periods (1, 3, and 5 minutes) with the constant pressure at 0.1 Pa. During
the synthesis of PtOa PdOb NPs@Ti3 C2 Tx nanosheets, three steps were performed:
etching, sonication, and the last one is solvent-free plasma irradiation. After the
characterization of HRTEM, the size distribution graph indicates that the alloy
PtOa PdOb NPs@Ti3 C2 Tx NPs average size is 2.2, 2.5, and 6 nm at 1, 3, and 5 minutes,
respectively. The PtOa PdOb NPs@Ti3 C2 Tx nanosheets obtained at 3 minutes are
uniformly more dispersed than the preparation time of 1 and 5 minutes. These
results concluded that the plasma treatment crystal growth depended on time [69].
The TEM (Figure 2.18b) and HRTEM images (Figure 2.18c) confirmed the strong
chemical coupling between the noble metal NPs [Pt(0), Pt(II), Pt(IV), Pd(0), and
Pd(IV)] and the Ti3 C2 Tx nanosheets at the interface. The as-prepared nanocompos-
ite electrode was shown the superior catalytic activity for OER and HER in alkaline
solution [69].
54 2 Strategies for the Preparation of Nanocatalysts and Supports Under Solvent-Free Conditions

(a) Ti3C2
HF DMSO

Etching Sonication
AI
Ti3AIC2 Max Multilayers Ti3C2Tx Few-layer Ti3C2Tx

PdCl2,
H2PtCl6∙xH2O

Overall Irradiation
water-splitting under plasma

Cathode H2 Anode O2 PtOaPdObNPs@Ti3C2Tx Pd2+/PtCl62–/Ti3C2Tx

Mxene Pd2+ PtCl62– Pd Pt PdO PtOa

(b) (c)

50 nm 1 2 3 4 5 nm
Particle size (nm)

Figure 2.18 (a) Schematic diagram of the preparation of PtOa PdOb NPs@Ti3 C2 Tx
nanosheets, (b) TEM, and (c) HR-TEM images of PtOa PdOb NPs@Ti3 C2 Tx nanosheets. Source:
Reproduced with permission from Cui et al. [69]/American Chemical Society.

Yuan and his team have reported the scalable preparation of ultrafine NiO
nanocubes with high dispersion and exposed crystal facets reinforced by the
thermal plasma method (Figure 2.19a) [17].
The NiO nanocubes were synthesized by using the irregular bulk metal Ni
(30–50 μm) as a precursor material transported in the plasma by a carrier gas
(O2 ) which further grew into the NiO nanocubes at low temperature. The SEM
and TEM images revealed that the morphology of the NiO nanocubes with a size
of 50–100 nm is dominant in the as-synthesized products (Figure 2.19b,c) [17].
Here, the significantly activated ultrafine Ni particles may catalyze the irreversible
Li-ion’s decomposition in the solid electrolyte interlayer and increase capacity. The
as-synthesized NiO nanocubes exhibited versatile electrochemical performance:
high capacity for long cycles (>1000 mAh g−1 ; 600 cycles), improved Coulombic
efficiency (>99%), and superior rate properties. This work can help to design
atomic-scale catalysts and also development for advanced materials with high
performance.
2.4 Plasma-Assisted Methods 55

(a) (b)

Power supply
O2 Ni Bulk Ni

Bulk Ni
ICP vapors 100 nm

Nucleation (c)
Thermal plasma and growth
reinforced 50 nm

NiO
NiO
nanocubes 50 nm
nanocubes

Figure 2.19 (a) The schematic diagram of the synthesis process of NiO nanocubes.
(b) SEM, and (c) TEM images of NiO nanocubes. Source: Reproduced with permission from
Hou et al. [17]/American Chemical Society.

(a) To positive terminal


of power supply (b)

Double walled
water cooled dome
and chamber

Plasma torch

Plasma plume 100 nm


Inlet and outlet
of water
Graphite (c)
crucible
Nucleation and assembly
growth of View port
nanoparticles

Movable anode Wilson seal


To negative terminal
Water inlet of power supply
and outlet
for anode 5 1/nm

Figure 2.20 (a) Schematic representation of DC transferred arc thermal plasma reactor
(DC-TATPR) for the synthesis of Ni NPs, (b) TEM images, and (c) SAED pattern for Ni NPs.
Source: Reproduced with permission from Ghodke et al. [70]/Elsevier.

The nickel NPs were synthesized by a thermal plasma route, reported by Mathe
and coworkers (Figure 2.20a).[70] This method obtained the highly crystalline
Ni NPs with a maximum number of NPs size of around 30–50 nm. The observed
TEM image reveals that Ni NPs possess a spherical shape (Figure 2.20b). The
SAED pattern of Ni NPs has shown their polycrystalline nature (Figure 2.20c). The
as-synthesized Ni NPs were thoroughly characterized and employed as catalysts for
hydrogen production using NaBH4 at temperature 370 K [70].
56 2 Strategies for the Preparation of Nanocatalysts and Supports Under Solvent-Free Conditions

Nanocrystalline Ni NPs catalyst was reused for hydrogen production up to five


cycles at optimized reaction parameters. The as-synthesized Ni NPs may be helpful
for large-scale hydrogen production, especially in the industries.

2.4.2 Cold Thermal Plasma Method


Cold plasma is a fast, facile, environmentally friendly, and energy-efficient
method for synthesizing nanocatalyst and supports [58, 71]. In the cold plasma
a low-pressure (p < 200 Pa) low-temperature plasma (cold plasma) the electron
temperature is much higher than the gas or ion temperature, which is close to room
temperature because of the low collisional coupling between electrons and heavy
particles at reduced pressures [60, 62]. The cold plasma method prepares various
nanocatalysts and supports them in solvent-free conditions, avoids the aggregation
of metal NPs, and has high catalytic performance with environmental sustainability.
Al-Masoodi et al. have synthesized Ag NPs from Ag wire of 3 mm length and
0.5 mm diameter as a precursor via plasma-assisted hot-filament evaporation strate-
gies [72]. During the preparation, Ag NPs had played the crucial role of compo-
nents like home-built reaction chamber, filament heating power supply, RF match-
ing impedance device, RF power generator, substrate heating power supply, and H2
gas cylinder in the plasma-assisted hot-filament evaporation system (Figure 2.21a).
The Ag NPs deposited on glass substrate were investigated by TEM images, and the
size distribution graph illustrated the Ag NPs exhibited as a circular-like shape and
size in the range 6–29 nm, respectively (Figure 2.21b). In the HRTEM image shown,
two crystal structures were observed Ag NPs with a lattice spacing of about 0.22
and 0.26 nm corresponding to the crystalline cubic Ag (111) and silver oxide (Ag2 O)
(111) (Figure 2.21c). Then Fast Fourier Transform (FFT) images were further con-
firmed crystalline structures of the Ag and Ag2 O (Figure 2.21d,e) correspondingly
[72]. From this work, we can modify glass materials by these Ag NPs on organic
light-emitting diodes performance.
Anthony and coworker have prepared Au NPs via the plasma-based gas phase
method [73]. During the synthesis of Au NPs, provided DC power to Pt–Au was get-
ting gold surface evaporation the Au atom entered in gas phase and then condensed
to form Au NPs and exhibited a reactor image (Figure 2.22a) [73]. It is noted that
for the synthesis of Au NPs, the flow rate and other parameters affect the yield of
Au NPs product, which is mentioned in the table (Figure 2.22a). The TEM image
was elucidated in smaller Au NPs uniformly distributed over the grid at the end of
the reactor (Figure 2.22b), and single Au NPs with the crystallographic plane were
identified by FFT shown in Figure 2.22c. The average Au NPs diameter is of ∼4 nm,
which can be suitable for industrial applications [73]. From this work, the strategy
may be explored for the high production of Ag NPs with suitable diameters using
the plasma method.
Bhattacharyya and coworkers synthesized the Pd NPs-decorated Hydrogen
plasma-treated TiO2 under plasma treatment [74]. The PdCl2 and HCl solution was
stirred at 50 ∘ C for 3 h. The as-prepared TiO2 films were kept in the solution for
2−3 min. The films were thoroughly washed with DI water and dried under nitrogen
2.4 Plasma-Assisted Methods 57

(a)

Silver wire RF electrode


Tungsten filament H2

Substrates Shutter

View port Thermocouple

Heater electrode

25
250

To vacuum system

(b) (c) (d)

(e)

100 nm 5 nm

Figure 2.21 (a) Schematic setup Ag NPs preparation via plasma-assisted hot-filament
evaporation system, (b) TEM, (c) HRTEM images of typical Ag NPs deposited on the glass
substrate, (d and e) the Fast Fourier Transform (FFT) images of the respective crystals in (c).
Inset (b) represents the graph of diameter distribution. Source: Reproduced with permission
from Hamood Al-Masoodi et al. [72]/Elsevier.

(a)
DC power (b)

10 cm
Pressure: 100 nm
1–8 Torr
D = 5–55 cm

Flow- DC
Time Distance Pressure
(c)
Sample rate power AuNPs
(min) (D, cm) (Torr)
(sccm) (W)
1 100 10–13.6 1-5 5.1 No 1.4
2 600 9.9–14.1 3-5 20.3 Yes 2.25
Stainless 10.2–
steel filter 3 900 5 50.8 Yes 5.1
13.9

4 1200 11.3–15 5 55.8 Yes 7.4

Vacuum pump

Figure 2.22 (a) Schematic representation of hot-wire Au NPs synthesis, including inset
table of reaction parameters of Ag NPs, (b) TEM image of Au NPs, and (c) FFT of the Au NPs.
Source: Reproduced with permission from Izadi et al. [73]/John Wiley & Sons, Inc.
58 2 Strategies for the Preparation of Nanocatalysts and Supports Under Solvent-Free Conditions

(a) Hydrothermal Synthesis


Δ = 150 °C, 4 : 30 h

FTO
FTO

Hydrogen cold plasma Pd2+ deposition

Pd NPs decorated TiO2

reduced in H2 plasma
FTO
FTO

= Pd2+
= Electron
= Pd metal
FTO

(b) (c)
0.48 μm
2.02 μm

Figure 2.23 (a) Schematic process for Pd NPs-decorated Hydrogen plasma-treated TiO2 ,
(b) FE-SEM image of TiO2 , and (c) FE-SEM image of Pd@TiO2 thin films. Source: Reproduced
with permission from Adak et al. [74]/American Chemical Society.

airflow after deposition. To modify an existing RF magnetron system, an ultrahigh


purified H2 was used to obtain H plasma. The Pd doped TiO2 film was placed on
the reactor cathode to achieve Pd@TiO2 photocatalyst (Figure 2.23a) [74]. The
as-synthesized Pd@TiO2 photocatalyst showed photoelectrochemical behavior as
well as the ability to oxidize water in water splitting devices. The high-magnification
field-emission scanning electron microscopy (FE-SEM) analysis shows the mod-
erately thick square top nanorod-like structure of TiO2 (Figure 2.23b). The top
surface of TiO2 nanorods was decorated with Pd NPs via cold H plasma treatment
(Figure 2.23c).
The H plasma treatment was found to be beneficial for enhancing the photore-
sponse due to the complete reduction of Pd2+ to Pd metal NPs. This H plasma-treated
strategy that will help the photocatalyst synthesis could be used in sensors, solar
cells, fuel cells, biomedical devices, and some other uses in the nanotechnology area.
2.5 Deposition Method 59

(a) high pressure (b) 45 m Torr


Carbon Fe(acac)3
xerogel crystallized
freeomagnetic
+
Fe3N nanoparticle
on carbon

magnet
(c) 6 m Torr

Ar:MH3 plasma at low pressure


various pressure amorphous
non-magnetic

Figure 2.24 (a) Schematic representation of the synthesis of Fe3 N NPs on a xerogel carbon
(CXG ) substrate, HAADF-STEM images of Fe3 N/CXG treated under (b) 45 mTorr, and (c) 6 mTorr.
Source: Reproduced with permission from Haye et al. [75]/American Chemical Society.

Haye et al. have designed Fe3 N-type NPs evenly deposited over high-surface-area
porous carbon support via the plasma synthesis process [75]. Fe NPs were formed
from the degradation of iron acetylacetonate precursor mixed with a carbon xerogel
treated in an RF inductively coupled plasma reactor at 13.56 MHz. Then the mix-
ture was treated at four diverse pressures, namely 6, 12, 25, and 45 mTorr (0.8, 1.6,
3.3, and 6 Pa) (Figure 2.24a). During plasma treatment, the deviation in pressure
directly affects the residence time of reactive species, as well as the crystalline state of
NPs. The Fe3 N NPs phase was controlled by pressure, where low-pressure treatment
resulted in amorphous surfaces and high-pressure treatment resulted in crystallized
surfaces. STEM image confirmed that homogeneously dispersed NPs designed con-
sistently on porous carbon surfaces (Figure 2.24b,c). The results concluded that the
synthesized magnetic material is an efficient, cheap, nontoxic filler and environmen-
tally friendly for such hybrid membranes.

2.5 Deposition Method


In recent years, nanocatalysts and supports significantly impacted chemical pro-
cesses, especially in industrial and academic fields [76, 77]. There are a variety of syn-
thesis strategies reported for the nanocatalyst and support. However, those strategies
have crucial problems for synthesizing nanocatalysts and supports in solvent-free
conditions while maintaining their high surface, selectivity, activity, thermal
stability, active porous site, and morphology [78, 79]. In this section, we will be
discussing recently reported advanced atomic layer and chemical vapor deposition
(CVD) strategies for nanocatalyst and supports by solvent-free and benign protocol.

2.5.1 Atomic Layer Deposition (ALD) Method


The atomic layer deposition (ALD) is an excellent technique for synthesizing thin
film, photoactive materials, metal oxide, various supports, and nanocatalyst [80–82].
60 2 Strategies for the Preparation of Nanocatalysts and Supports Under Solvent-Free Conditions

The ALD has a significant parameter for controlling surface morphology with an
active site to synthesize different types of nanocatalysts and supports [83–86]. Now,
this method is called greener due to catalysts synthesized by solvent-free approach.
This section highlighted the solvent-free and environmentally friendly synthesis of
nanocatalyst and supports by the ALD method.
Detavernier and coworkers have synthesized BMNPs of noble and non-noble
metal of Pt–In NPs via the ALD route [87]. This bimetallic alloy NPs syn-
thesis included three steps, (i) ALD deposition of MOx layer using a metal
(In or Ga) precursor and O2 plasma, (ii) ALD deposition of a Pt layer using
trimethyl(methyl cyclopentadienyl)platinum(IV) [(MeCp)PtMe3 ] and O3 , and
(iii) temperature-programmed reduction (TPR) in H2 to convert the as-deposited
Pt/MOx bilayer into Ma Ptb BMNPs (Figure 2.25a) [87]. The Ma Ptb BMNPs catalyst
was shown superior activity for the dehydrogenation of propane to propene reaction.
The SEM images and size distributions showed that the Pt–In NPs were dependent
on the atomic ratio (Pt+In) total thickness (dTotal) of the three samples. Two trends
are observed: (i) the average particle size reduces linearly with reduced dTotal
(Figure 2.25b–e) and, (ii) At Pt/(Pt+In) atomic ratios of 75%, larger BMNPs are
obtained for the most Pt-rich phase, InPt3 , than for the other phases of Pt/(Pt+In)
atomic ratios of 60% and 40% [87]. The presented work can help create openings

(a) Substrate
Metal precursor (b) 80

60
Counts

ALD
40
O2 plasma 20
Step 1
MOx 0
0 12 20 30 40 50
150 nm Particle size (nm)
MeCpPtMe3
(c) 90
ALD Ozone
Counts

Step 2 60

30
Pt
0
0 12 20 30 40 50
150 nm Particle size (nm)
Hydrogen
(d) 200
ΔT SEM
Step 3 150
Counts

Ma Ptb 100
50
0
0 12 20 30 40 50
150 nm Particle size (nm)
200
(e)
150
Counts

100

–H2 50
0
Porous SiO2 0 12 20 30 40 50
150 nm Particle size (nm)

Figure 2.25 (a) Schematic representation of ALD-based synthesis of BMNPs on silica


supports, (b–e) SEM images and particle size distributions graph measured after
temperature-programmed reduction (TPR) up to 700 ∘ C of Pt/In2 O3 bilayers with Pt/(Pt+In)
atomic ratios of c. 75%. Source: Reproduced with permission from Ramachandran et al.
[87]/American Chemical Society.
2.5 Deposition Method 61

(a)

Pd(hfac)2 N2

Formalin

: Pd(hfac)2 : –hfac

: Formalin : Polar H
N2
: Pd :C

(d) 50
(b) (c) 3.4 ± 0.4 nm

Frequency (%)
40
30
20
10
20 nm 5 nm 0
2 3 4 5 6
Particle diameter (nm)

Figure 2.26 (a) Schematic representation of Pd/CM-TiO2 -H catalyst synthesis, (b–d) TEM
images and size distribution graph of Pd/CM-TiO2 -H. Source: Reproduced with permission
from Lu et al. [18]/American Chemical Society.

for fields searching bimetallic materials, including non-noble metals, for example,
Pt–Fe or Pt–Co NPs, for magnetic applications.
Chen and coworkers successfully loaded Pd NPs and deposition of TiO2 on
Al2 O3 ceramic membranes (CMs) obtained Pd/CM–TiO2 –H via the ALD route
(Figure 2.26a) [18]. The membrane surface was modified by deposition and
subsequently calcination with TiO2 by ALD to obtain the Al2 O3 CMs pores. This
Pd/CM-TiO2 -H catalyst was shown superior catalytic activity toward reduce the
p-nitrophenol to p-aminophenol because of the surface porous properties of Al2 O3
CMs. TEM image of Pd/CM-TiO2 -H catalysts observed that Pd NPs were success-
fully distributed on the porous surface of CM-TiO2 -H by ALD. The Pd(111) plane in
the lattice of 0.22 nm and an average particle diameter of 3.5 nm for metallic Pd NPs
was observed (Figure 2.26b–d). The preliminary work can be encouraged for future
study that will improve catalytic performance, specifically for catalytic stability.
Ren et al. have fabricated the Cu/ZnO-50 catalyst by an ALD method
[88]. For the synthesis of Cu/ZnO-50 (50 is the ALD cycle) the precursors
of copper(II)-hexafluoroacetylacetonate [Cu(hfac)2 ] and diethylzinc [DEZ,
(C2 H5 )2 Zn] were used. The precursors were distinctly introduced to the reactor
chamber with N2 gas, the ALD cycles prepared the series of ZnO/Cu, such as
ZnO NWs, ZnO/Cu-5, ZnO/Cu-10, ZnO/Cu-20, ZnO/Cu-50, and ZnO/Cu-100
62 2 Strategies for the Preparation of Nanocatalysts and Supports Under Solvent-Free Conditions

(a)
ALD
ZnO seeds

Silicon
Hydrothermal
ZnO ZnO NWs

Cu

ALD Cu

(b) (c)

Cu
EDS

5 nm

ZnO NW
47 nm

50 nm 5 nm

Figure 2.27 (a) Schematic representation of the preparation process of ZnO NWs
decorated with Cu NPs via thermal ALD, (b) TEM image of ZnO/Cu-50, and (c) HRTEM
images of ZnO/Cu-50. Source: Reproduced with permission from Ren et al. [88]/American
Chemical Society.

(Figure 2.27a) [88]. The TEM image of the ZnO/Cu-50 observed ZnO NWs diameter
size is ∼47 nm (Figure 2.27b). The corresponding SAED pattern revealed strong
reflection spots of the (002) and (100) reflections of ZnO and a diffraction ring
equivalent to the polycrystalline structure of Cu (Figure 2.27c). The enhancement
photoluminescence (PL) of ZnO NWs with a decreasing number of ALD cycles
decreased the size of the Cu NPs. This work is studied to design Cu NPs with highly
controllable size and explain that the surface passivation and electron transmission
between Cu and ZnO can effectively manipulate the PL intensities of ZnO NWs.
Chen and coworkers have designed CeOx deposited on Pt NPs to obtain nanofence
Pt/CeOx catalyst via the facet-selective ALD method [89]. The Pt/CeOx catalyst was
achieved by four steps: Pt precursor attached to active surface, ligand exchanged
for Ce precursor’s chemisorption, O3 combustion reaction, and surface reactivated
after 700 ∘ C calcination. The HRTEM images revealed 200 cycles, the CeOx (100)
2.5 Deposition Method 63

deposited over Pt (100) surfaces and achieved the core–shell-like structure. However,
at 50 and 100 cycles, CeOx (100) was rarely seen deposited on Pt NPs and shown in
an atomic model of HRTEM images. The neighboring planes’ distance in Pt NPs
is 0.23 nm for (111) and 0.19 nm for (100) facets, CeOx is also crystalline [89]. The
study reveals that forming a metal oxide nanofence structure to encapsulate precious
metal NPs effectively enhances the catalytic activity and thermal stability.

2.5.2 Chemical Vapor Deposition (CVD) Method


CVD is widely used for materials processes in the research area and various
industries [90]. The CVD is a powerful technology for producing a high-quality
polymeric thin film, 2D materials, metal oxide, nanocatalyst, carbon-based sup-
porter, and producing single-metal or bi-metal catalysts without solvent [91, 92].
CVD is a fundamental process in which gases precursor adsorbs on the particular
active surface (supports) in the gas phase. In this section, we discuss the advanced
synthesis of nanocatalysts and supports by the CVD method under solvent-free
conditions.
In this research work, the CVD-assisted synthesis is adapted for the deposition
of highly dispersed metal NPs and specifically for multimetallic systems. Kim
and coworkers have fabricated the deposition of Pt-Co alloys NPs on the surface
of carbon black via a low-temperature CVD technique [93]. The Pt-Co/C(600),
Pt-Co/C(500) and Pt-Co/C(400) BMNPs catalysts were obtained from MeCpPtMe3 ,
and di-carbonyl cyclopentadienyl cobalt(I) (CpCo(CO)2 ) as a precursors. Then
metal was deposited on the surface of carbon supports, such as propane (C3 H8 ),
methane (CH4 ), cyclopropane (C3 H4 ), and cyclopentadiene (HCp) in the gas
phase [93]. As prepared BMNPs catalysts were analyzed by thermogravimetric
analysis (TGA), inductively coupled plasma-mass spectrometry (ICP-MS), and
X-ray fluorescence spectroscopy (XRF). In HRTEM images and SAED pattern,
the as-deposited Pt and Co NPs were revealed highly dispersed over the carbon
surface. Among as-prepared catalysts, Pt-Co/C(500) catalysts demonstrates the best
performance electrocatalytic OER to maintain high mass activity, specific activity,
and long-term stability. This method can be helpful for the synthesis of BMNPs by
deposition monodispersed with composition-controllable and applied it to various
catalyst synthesis processes [93].
Similarly, the CVD-assisted synthesis method is adapted for encapsulating
metal and nonmetal alloys NPs in graphene shells. Kim and coworkers have
synthesized ultra-stable graphene-encapsulated Fe@G, Co@G, and Ni@G
NPs by the CVD method [94]. These catalysts were made from organic metal
precursors (ferrocene (Fe(C5 H5 )2 ) for Fe@G, CpCo(CO)2 for Co@G, and
bis(methylcyclopentadienyl)nickel(II) ((MeCp)2 Ni) for Ni@G) and ethanol as
a starting materials for the metal core and graphene shell, correspondingly. The
aforementioned procedure followed for designing of cobalt-promoted catalysts, Co
in situ post-synthetic treatments such as oxidation, phosphidation, and sulfidation
with nonmetal (O, P, and S) to obtained Co3 O4 @G, CoP@G, and CoS2 @G catalysts.
Among various catalysts, Co3 O4 @G catalyst exhibit excellent activity for lithium-ion
64 2 Strategies for the Preparation of Nanocatalysts and Supports Under Solvent-Free Conditions

(a) 2-step floating catalyst CVD for high purity spin-capable CNT forest Spin-capable CNT forest

Catalyst solution mist Long


CNT (b)
Catalyst particle

SiO2

Si Carbon source
Very low metallic impurity
(c)100
80

Weight (%)
Inactive particle 60
Mist deposition time
40
Short
STEP-1: Catalyst particle formation 20
0
STEP-2: CNT growth 200 300 400 500 600 700 800
Temperature (°C)

Figure 2.28 (a) Synthesis of CNT forest via CVD route, (b) SEM image and the dry-spin
capable CNT forest, and (c) TGA profile of the CNT sample with mist deposition for
3.5 minutes. Inset shows the TEM image of the CNT. Source: Reproduced with permission
from Kinoshita et al. [95]/Elsevier.

battery anode application. The HRTEM analysis confirmed that Co3 O4 @G, CoP@G,
and CoS2 @G were successfully synthesized with ultrasmall size with a narrow
diameter and atomically distributed the core particles over graphene shell. The
as-fabricated Co3 O4 @G, CoP@G, and CoS2 @G catalysts can be used for reliable
high-performance catalysis and energy storage/conversion.
Inoue and co-workers have prepared CNT forest growth with low impurity levels
by the two-step floating CVD method. The CNT preparation was used Fe solution
mist over silica (SiO2 ) and ferrocene as a carbon source (Figure 2.28a) [95]. The
ferrocene was decomposed on the substrate at 700 ∘ C, iron NPs were formed by sur-
face diffusion and agglomeration on the substrate. After catalyst formation, the mist
supply was stopped, and then the chamber was purged with Ar for 10 seconds to
exhaust the residual mist to obtained CNT and low metal loaded with high purity
[95]. According to the TGA profile and SEM image of the CNT, growth occurred for
15 minutes and mist deposition occurred for 3.5 minutes (Figure 2.28b, c). In terms
of weight loss, 99.2 weight percent was estimated, which corresponds to a level of
impurities of less than 1%.
The simple two-step floating catalysts-CVD process can contribute to applied
CNT-based applications, such as large-scale structural materials with dry spinning
phenomena and electrodes for supercapacitors and lithium-ion batteries.

2.6 Conclusion and Future Perspective

This chapter has highlighted recent advances of various solvent-free methods


for NPs and supports synthesis. The NPs synthesis has been achieved by using
environment-friendly, solvent-free methods, such as the mechanochemical
approach with the ball milling and mortar and pestle grinding, thermal treatment
method, plasma-assisted method, and deposition method. The solvent-free synthetic
routes are the benign, cost-effective, and safest method for the environment. The
References 65

solvent-free synthetic technique has found the best way for various kinds of material
synthesis. Overall, this chapter has provided great generality and scope toward
synthesizing material without using solvents. Therefore, we have faith that the
NMs area is well-organized, environment-friendly, and sustainable processes via
the solvent-free synthesis of novel and highly active nanocatalysts and supports.
We hope that this chapter would help the researcher to better understand the
potential values of these solvent-free synthetic techniques, which may lead to
designing new and more effective processes. The mortar and pestle grinding with
a mechanochemical technique is very well developed by the solvent-free approach
for various kinds of material synthesis for NPs, metal oxides, and many different
materials. Furthermore, the solvent-free synthesis area is also exploring for the
other materials. For example, newly derived SACs, sub-NMs, solid-supported mixed
metals catalyst, and many more types of materials syntheses can be used for a
variety of catalytic and environmental technologies.

Acknowledgments

The author is very much thankful to the Council for Scientific and Industrial
Research (CSIR) and Union Grant Commission (UGC), New Delhi, to award the
junior research fellowship and also grateful to the Institute of Chemical Technology
(ICT), Mumbai-Marathwada Campus, Jalna, Maharashtra, India.

References

1 Varma, R.S. (2014). Appl. Sci. 4: 493–497.


2 Centi, G. and Perathoner, S. (2003). Catal. Today 77: 287–297.
3 Gawande, M.B., Goswami, A., Felpin, F.-X. et al. (2016). Chem. Rev. 116:
3722–3811.
4 Rathi, A.K., Zboril, R., Varma, R.S., and Gawande, M.B. (2016). Ferrites and
Ferrates: Chemistry and Applications in Sustainable Energy and Environmental
Remediation. American Chemical Society https://doi.org/10.1021/bk-2016-1238
.ch002.
5 de Oliveira, P.F.M., Torresi, R.M., Emmerling, F., and Camargo, P.H.C. (2020). J.
Mater. Chem. A 8 (32): 16114–16141.
6 Ahmed, M. (2020). Polymer Science and Nanotechnology. Elsevier https://doi.org/
10.1016/B978-0-12-816806-6.00016-9.
7 Gawande, M.B., Velhinho, A., Nogueira, I.D. et al. (2012). RSC Adv. 2:
6144–6149.
8 Gawande, M.B., Goswami, A., Asefa, T. et al. (2015). Chem. Soc. Rev. 44:
7540–7590.
9 Dhand, C., Dwivedi, N., Loh, X.J. et al. (2015). RSC Adv. 5: 105003–105037.
10 Baig Rb, N. and Varma, R.S. (2013). An Introduction to Green Chem. Methods,
18–38. Future Science Ltd https://doi.org/10.4155/ebo.13.4.
>
H H
Neat N N
R1 NCS + R2 NH2 R1 R2
Ball milling, rt S
3.2.1.20 3.2.1.21
3.2.1.22
1 49 examples
R = alkyl
R2 = alkyl, aryl >99%

O NC CN

CN H CN
Neat
2 +
CN OH Grinding, rt
O NH2
R R
3.2.1.23 3.2.1.24 3.2.1.25
11 examples
89–98%

R = H, 3-OMe, 4-OMe, 5-OMe, 3-OEt, 5-NO2, 5-Cl, 3,5-di-I, 3-OMe-5-NO2, 3,5-di-NO2, 3-Br-5-Cl
O
R1 NH2 Br R1 N
Ar/Het Neat Ar/Het
N + N
2 Grinding, 25–30 °C, 3–5 min R2
R

3.2.1.26 3.2.1.27 3.2.1.28


30 examples
96–99%
R1 = H, Me
R2 = 4-Me, 4-OMe, 3-OMe, 4-Br, 3-Br, 4-Cl, 3,4-di-Cl, 4-F, 4-NO2, 4-CN, 2-naphthyl, 4-acetylbiphenyl,
3-(bromoacetyl)coumarin

O O Ar
S
Neat NH
+ Ar CHO +
O H2N NH2 Grinding, rt, 50–90 min N S
H
3.2.1.17 3.2.1.29 3.2.1.7
3.2.1.30
4 examples
Ar = 4-Cl-C6H4, 4-Br-C6H4, 4-NO2-C6H4, 2-NO2-C6H4
79–85%
O O
O O
3.2.1.32 Quartz sand O O
NO2 or R1 R2
+ Ar NO2
or Grinding, rt NO2
Ar Ar
O O 3.2.1.33
3.2.1.34 3.2.1.35
R1 R2
17 examples
3.2.1.31
58–99%
dr = 55 : 45 to 99 : 1
R1 = H, aryl
R2 = Ph, OEt, Me

O
3
R2 R
+ R1 R3 H
Neat Ph3P C 3.2.1.39
R1 CH2X + PPh3 Ph3P CH2R1 X
K2CO3 H R2 R1
3.2.1.36 3.2.1.37 Ball milling 3.2.1.38 3.2.1.40
rt, 3–20 h 8 examples
1
R = H, Ph, C(O)Ph, C(O)(OEt) 70–99%
X = Br, Cl E/Z ratio = 1.6 : 1 to 3.5 : 1
R2 = H, Me

R3 = Br,
O O Selectfluor O O O O
(2.0 equiv)
R1 R 2
R1 R 2 and/or R1 R2
Ball milling, rt F F
0.5–2 h F
3.2.1.41
3.2.1.42 3.2.1.43
1 R2
R = = alkyl, aryl 10 examples
68–98%
Cl
N Conditions Mono-fluorinated:di-fluorinated
N 2BF4
Neat 1.2 : 1 to 17 : 1
F
Selectfluor MeCN 5 : 1 to 50 : 1
Na2CO3 1 : 8 to 1 : 50

NHSO2R3 R2

FeCl3 (3 equiv)
+
Ball milling, rt, 2 h
R1 R2
3.2.1.45 R1
3.2.1.44 3.2.1.46
R1 = H, 4-Me, 3,4-di-Me, 4-Cl 8 examples
R2 = H, 4-Cl, 2-Cl, 4-Br 15–41%
R3 = Ts, Ms
X
Br O

R1 12 examples
OH
76–99%
X R1

3.2.1.50 NBS, ball milling, rt, 1 h 3.2.1.53


R2
O OH DBDMH, ball milling, rt, 1 h
O
+
O 11 examples
Br
R3 69–99%
R2
3.2.1.51 3.2.1.52

R3
1
R = 4-Me, 4-CF3, 3,5-di-CF3, 4-OMe, 4-F, 4-Cl, 4-COCH3 3.2.1.54
R2 = H, 2,4-di-Me, 2,4,6-tri-Me, 2-Br, 2-F, 4-NO2
R3 = H, 4-Me, 2,4,6-tri-Me, 4-iPr, 4-F, 4-Cl, 4-Br
X = O, H2

O Br
O
N
N Br O
N
O Br
NBS DBDMH
OH OH
O H 2N R
NH2 Neat N R HCl (aq)
+ R H
Ball milling, 3–4 h THF, 22 °C H2N R
NH2 N R
3.2.1.55 12 h
OH OH 3.2.1.57
20 examples
3.2.1.54 3.2.1.56 65–99%
99% ee 99% ee

R = alkyl, aryl
NH2 O N
Neat
+ R
X R Ball milling, 60 min N
NH2
H
3.2.1.58 3.2.1.59
3.2.1.60
7 examples
R = H, alkyl, aryl
80–97%
X = H, OH
R2
O
1 1 N
R O O R
O
O Neat
O
N + + N
O NH2 MW, 110 °C, 10 min N
R2 R3 O N R3
H
3.2.2.1 3.2.2.2 3.2.2.3
3.2.2.4
R1 = H, F, Cl, Br, NO2 20 examples
R2 = H, Me 83–95%
R3 = H, -(CH3)2
R

O H
O N
CHO
Neat
+ + NH4OAc
MW, 700 W, 110 °C, 7 min O
O R 3.2.2.7 O

3.2.2.5 3.2.2.6

R
R = H, 2-Br, 2-Cl, 3-Br, 2-Me, 3-Me, 3-NO2, 4-Br, 4-F, 4-Cl, 4-NO2,
4-Me, 4-OMe, 2,4-di-Cl 3.2.2.8
14 examples
76–85%

O
O
PPh3
Ar CHO + R N3 + HS OH R
N S
Neat
Ar
3.2.2.9 3.2.2.10 3.2.2.11 80 W, 130–150 °C, 10–15 min
3.2.2.12
9 examples
O O 92–96%
R= , , O , O ,

Proposed mechanism
R
N
PPh3 3.2.2.9 HC 3.2.2.11
R N3 R N PPh3 3.2.2.12
Straudinger aza-Wittig Ar –H2O
3.2.2.10 3.2.2.13
3.2.2.14
Ph
O NH
H
Ph
N CONHBn PhNCO Ph * N CONHBn

CN (CH2)3NHBoc MW, 80 °C, 2 h CN (CH2)3NHBoc

(S)-3.2.2.15 (R,S)-3.2.2.16
98%
(epimer ratio = 1 : 1)

Epimerization process

Ph Ph
Ph
O NH O NH
O NH
(S) N (S) CONHBn MW (S) CONHBn
Ph (S) N (S) CONHBn Ph * N
H Δ Ph
CN (CH2)3NHBoc C CN (CH2)3NHBoc
(CH2)3NHBoc
(S)-3.2.2.16 NH (R,S)-3.2.2.16
3.2.2.17
O O R1
CN CN
Neat
R1 CHO + + R2 R2
CN O MW, 80 °C, 7 min O NH2
R2 R2
3.2.2.18
3.2.1.23 3.2.2.19 3.2.2.20
17 examples
R1 = alkyl, aryl, heteroaryl, cinnamyl 72–88%
R2 = H, Me

OH

NH2 X2 NH2
OH
N N
O X1 H O
N H O X2 N
3.2.2.24 OH 3.2.2.22
R1 n n
N R1 150 W, 180 °C, 2 min X1 N R1 2
H R2
H R2 O 150 W, 180 °C, 2 min H R
3.2.2.21 3.2.2.23
3.2.2.25
R1 = R2 = cycloalkyl, aryl 12 examples
5 examples
92–95% n = 1,2 90–98%
1 2
>99% de X = X = H, Cl, F, Me, OMe

Proposed mechanism

OH O
R2 OH
NH2 O N
O
OH Meyers’ reaction R1 N
N + R1 n O
H R2 O followed by
2
Pictet–Spengler N R1 R
3.2.2.22 3.2.2.21 N H
H
(n = 1)
3.2.2.26 3.2.2.27

3.2.2.25
>
R1 R1 R1
O
O Neat N
+ H N NH2 + R2 CHO N
2 +
MW, 130 °C
NH2
3.2.2.29 4 min N R2 N R2
H
3.2.2.28 3.2.1.7
3.2.2.30 3.2.2.31
R1 = aryl
Major Minor
R2 = alkyl, aryl 70–91%
(major/minor = 80 : 20 to >99 : trace)
Proposed mechanism
R1 R1 R1
O
NH NH
N
+ H R2 –H O
NH2 R2
2 N N R2
3.2.2.29 H
3.2.2.32 3.2.2.30
3.2.2.33
(major)
Pathway A
Aromatization
R1
R1
O + Urea + R2 CHO
N
NH2 3.2.1.7 3.2.2.29
N R2
3.2.2.28
3.2.2.31
Pathway B (minor)
R1 R1
NH Aromatization
O N
H R2
NH2 –H2O N R2
H
3.2.2.34 3.2.2.30
3.2.2.28
(major)
R
CN Neat n CN
O + + NH4OAc NH2
R n CN MW, 150 °C, 30 min n
3.2.2.7 R N
3.2.2.35 3.2.1.23 CN
NH2
3.2.2.36
n = 0, 1, 2
R = H, Me, Et 6 examples
41–87%

R2
B(OH)2 R1
CHO Neat N
R2 R 2
R2
R1 + N + OH
H MW, 120 °C, 2 h
OH
3.2.2.37 3.2.2.38 3.2.2.39
3.2.2.40
R1 = H, 2-Me, 3-Me, 4-Me, 3-OMe, 3,4,5-tri-OMe, 3-NO2 15 examples
(a) R2 = H, allyl, Bn, -CH2CH2OCH2CH2-, -CH2CH2CH2CH2CH2- 72–96%

O
O CHO N
Neat O
B(OH)2 + +
O OH
N OH MW, 120 °C, 2 h
H
(b) 3.2.2.37a 3.2.2.38a 3.2.2.39
3.2.2.40a
95%
OH
OMe

NH2 O
R3 MeO Neat
OH
+
MW, 145 °C, 5 min HN O
N R2 HO
R1 R3

3.2.2.41 3.2.2.42 N R2
R1
R1 = R2 = R3 = H, Me, OMe
3.2.2.43
7 examples
85–91%

Ar
Ar N
Ar
N N
Ar Neat
N +
O N 4 Å MS, MW, 80 °C
H N
10–40 min
3.2.2.45
3.2.2.44 3.2.2.46
19 examples
Me 15–94%
O N
, , , , N , N
N = N N N N
H H H H H
H H

Proposed mechanism
Ar
Ar
N
N Ar
N
N N
N
3.2.2.44 + 3.2.2.45 3.2.2.46
HN Ar H
N
N
H
3.2.2.47 3.2.2.48
O O O

O Neat N
+ O R2
N O N MW, 120 °C, 15 min N
R1 H R2 H R1 O

3.2.2.49 3.2.2.50 3.2.2.51


1 21 examples
R = H, 6-F, 6-Cl, 7-Cl, 6-Br, 6-OMe, 6-OCF3, 7-Me
73–96%
R2 = H, 5-F, 5-Cl, 5-Me, 5,7-di-Me

Proposed mechanism
R2
O 2
H R O
N
O + O N 3.2.2.51
–CO2 O –H2O
N O NH2 O
R1 H O
R1
3.2.2.49 3.2.2.50
3.2.2.52

Gram scale reaction


O O
O
O Neat N
+ O
N MW, 120 °C, 15 min N
N O H
H O

3.2.2.49a 3.2.2.50a 3.2.2.51a


(30.67 mmol) (30.67 mmol) (6.77 g, 89%)
O O
R2
CHO Neat N
H2N R2 (CH2O)n
O MW, 130 °C, 20 min O N
R1 3.2.2.54 3.2.2.55 R1
R2
3.2.2.53 3.2.2.56
R1 = R2 = alkyl, aryl 25 examples
27–99%

Proposed mechanism

O OH O O
CHO (CH2O)n R2
CHO CHO
MW 3.2.2.55 N
O R2 R2 H N R2 CH2
O N OH N 2 OH N
R1 H2N R2 R1 H R1 H 1
3.2.2.54 R R2
3.2.2.53 3.2.2.54
3.2.2.57 3.2.2.58 3.2.2.59

3.2.2.56
N O
Neat N
NH2 Br
+ Ar Ar
MW, 65 –C, 15–20 min N
R
R
3.2.2.60 3.2.2.61
3.2.2.62
10 examples
R = H, 4-Me, 5-Me, 5-Cl
84–90%

CHO R
O
Neat
2 Ar CH3 + NH4OAc
MW, 120 °C, 30 min
R 3.2.2.64 3.2.2.7 Ar N Ar
3.2.2.63
3.2.2.65
R = 2-OH, 3-OH, 4-OH 15 examples
71–86%
Ar2
Ar2 N
Ar1 HN N Ph
N
O Neat
N O
NH O 230 °C, 15 min
Ph
Ar1
3.2.3.1
3.2.3.2
7 examples
70–78%
R R
O OH

N O
N O NO2
NO2 R O O NO2
O
N 3.2.3.7 H 3.2.3.5
N O NH + S NH
O NH
110 °C, 45 min 110 °C, 30 min
3.2.3.3 3.2.3.4
3.2.3.6
3.2.3.8 15 examples
14 examples 79–91%
80–92%

R= H, 4-Cl, 4-Br, 4-F, 4-Me, 4-Et, 4-NO2, 4-OMe, 3-Cl, 3-Br, 3-NO2, 2-Cl, 3,4-di-OMe, 3,4,5-tri-OMe, pyrid-3-yl

Proposed mechanism

O Knoevenagel Michael NO2


condensation NO2 addition
H NH
O O –H2O O O S
S NH
3.2.3.5 or 3.2.3.7 H
3.2.3.9
3.2.3.4 3.2.3.10

Intramolecular H
O-cyclization NO2
NO2
NH –MeSH
O O NH
S

3.2.3.11 3.2.3.6 or 3.2.3.8

O R1
R1
Neat R3
+
R3 100 °C, 6–10 h N
N R2
R2 OH
3.2.3.12 3.2.3.13
3.2.3.14
R1 = H, F, Cl, Br, alkoxy
26 examples
R2 = alkyl, aryl 55–89%
R3 = alkyl, aryl
O
R1 R1
HOOC Neat N
NH +
O N
NH2 100 or 150 °C, 20–60 min
R2 R3 R2
3.2.3.15 R3
3.2.3.17
3.2.3.16 18 examples
>97%

R1 = H, Ph, Me, i-Pr, t-Bu, -CH2CH2OH, 2-Cl-C6H4, 3-Cl-C6H4, 4-F-C6H4, 4-NO2-C6H4, 4-Me-C6H4
R2 = H, 3,4-di-OMe
R3 = H, Me, Ph, 4-Me-C6H4, 4-Cl-C6H4

>

O OTMS
TMSCN
R H R CN
rt or 100 °C H
3.2.3.18
3.2.3.19
15 examples
(a) 95–99%

O CN
Neat H CN
R H +
CN 140 °C R CN
3.2.3.18 3.2.1.23
3.2.3.20
14 examples
(b) R = alkyl, aryl 93–99%
X R1
R1 X Neat
P H R2 R2 P
+
R1 H 80 °C, 5–13 h R1
H
3.2.3.21 3.2.3.22
3.2.3.23
11 examples
X = S, Se
60–94%
R1 = aryl
R2 = aryl, -COOMe, -SiMe3, -CH2OH

X
Ph
H P
Ph
C C
H
3.2.3.24
O

OH H2N OH HO
R NH2 OH R
RCHO
3.2.3.27 3.2.3.26
N OO RCHO N O 100 °C, 15–45 min N
N O O
3.2.3.26
3.2.3.29 100 °C, 10–30 min 3.2.3.25
3.2.3.28
6 examples 11 examples
88–94% R = alkyl, aryl 81–94%

O
O R1 P R2
R2 Neat
R1 P 2 + R2
R 120 °C, 48 h
3.2.3.30 3.2.3.31
3.2.3.32
R1 = R2 = alkyl, aryl 17 examples
46–98%
R2
R2 H
Neat N
R1 + O + R3 N C R1 Ar
N NH2 Ar 160 °C N
3.2.3.35
3.2.3.33 3.2.3.34 NH
R3
R1 = H, 2-Me, 3-NO2 3.2.3.36
R2 = H, OCH2Ph 13 examples
R3 = alkyl 80–98%

Proposed mechanism
R3
N+
O 3.2.3.35
C
Ar R3
H
R1 R1 R1 N+
R2 N H N

N NH2 –H2O N Ar N Ar
R2 R2 H
3.2.3.33 H+
3.2.3.38
3.2.3.37

R3 R3
R1 NH
R 1 N
N H
Ar N
Ar
N
N
R2
R2
3.2.3.36 3.2.3.39
R1 R2 O R1 R2
Neat N
N +
H O 60 °C, 3 h
O
3.2.3.40 3.2.3.41
3.2.3.42
R1 = H, alkyl 20 examples
R2 = alkyl, aryl 20–100%

O
NH2 HN
OH O OH
Neat
+
O 60 °C, 3 h

3.2.3.43 3.2.3.41 3.2.3.44


100%
X X
X
O R2 R3
R1 Neat
N + R3
H R2 80 °C, 40–120 min
HN NH
3.2.3.45 3.2.3.46 R1 R1

X = H, NO2 3.2.3.47
R1 = H, Me 25 examples
80–94%
R2 = alkyl, phenyl
R3 = H, alkyl, phenyl

O Ar Ar
O O N O O
N NH2 S O
S S
Neat N H2O2 N
ArCHO + + N N N
85 °C H N
O H
3.2.3.48
3.2.3.49 3.2.3.50
3.2.3.51 3.2.3.52
10 examples
80–89%
O R2 NH2 R2
HO OEt N
3.2.3.55
R1 CF3 or
H CF3 90–120 °C CF3
R1
3.2.3.53 3.2.3.54 3.2.3.56
15 examples
75–99%
R1 = H, Ph (100% E isomer)
H
R2 = H, alkyl, N N N
Ph , ,

OMe

O N 6

H + NH2
3 6
3 H

3.2.3.57 3.2.3.58 3.2.3.59


96%
E/Z ratio = 3 : 1
Ar
CHO HO
NHAr N
R1 = Me, Et, Pr, Pent, Ph
N O Neat O
+ N R2 = tBu, Bn
N 65 °C, 2 h N
R2O2C Ar = Ph, 4-ClC6H4, 4-FC6H4,
R1 CO2R2
4-MeOC6H4
R1
3.2.3.60 3.2.3.61 3.2.3.63
12 examples
less 10–60%
A1,3 strain Ar
H
H N O
H N
N
CO2R2
R1
3.2.3.62
Cl S
Neat
+ S + H2 N Ar Ar
NO2 130 °C, 24 h N
R R
3.2.3.65
3.2.3.64 3.2.3.66
28 examples
R = H, 3-Cl, 4-Me, 5-Me, 5-OMe, 5-Cl, 5-Br 52–80%

Proposed mechanism

S H2S 3.2.3.65 NH3 H2 S S, NH3


CH2
Ar NH2 Ar NH Ar N Ar Ar

3.2.3.65 3.2.3.67 3.2.3.68 3.2.3.69


Cl

NO2
R
3.2.3.64
Cl H2O S Cl
S
Cyclization
3.2.3.66 Ar Ar
Reduction N N
R R O O
O

3.2.3.71 3.2.3.70

O O
CHO NH2
O O
Neat R
Cl + N
100 °C, 30 min
R
3.2.3.72 3.2.3.73 3.2.3.74
7 examples
R = H, 4-Me, 4-OMe, 4-F, 4-Cl, 4-Br, 3-Me, 3-Cl 95–98%
O O
O
OH OH
OH Neat
+ + ArNH2 NHAr
OH O
75 °C, 10–60 min
O O 3.2.3.77 OO
3.2.3.75 3.2.3.76

3.2.3.78
10 examples
O 81–87%
HO
Proposed mechanism
O O
O O
N 3.2.3.75 NHAr Tautomerization
3.2.3.76 + 3.2.3.77 O 3.2.3.78
–H2O Ar
O H
OO
3.2.3.79

3.2.3.80
H2N S R2 S R2
N N
N
3.2.4.3
HN R1
)))), 60 °C, 1 h N N
N N

CHO 3.2.4.4
TMSN3 8 examples
1
R NC + 84–98%
N Cl
S R3
3.2.4.1 3.2.4.2 S
H2N R3
N N
N
R1 = c-Hex, t-Bu, Bn, 4-MeOPh 3.2.4.5
R2 = H, CN HN
)))), 60 °C, 1 h N N R1
R3 = H, F
N N

3.2.4.6
8 examples
79–94%
H CHO
N HCOOH
1 2 N
R R )))), rt, 2–240 min R1 R2
3.2.4.7 3.2.4.8
21 examples
R1 = H, alkyl, aryl 87–97%
R2 = aryl

H (Boc)2O Boc
N
R1 R2 N
)))), rt, 2–6 min R1 R2
3.2.4.9 3.2.4.10
1 2
R = R = H, aryl, alkyl 11 examples
quantitative yields

NH2 NHBoc
(Boc)2O
3 OH OH
R R3
)))), rt, 2–7 min
3.2.4.11
3.2.4.12
R3 = H, alkyl, aryl, -COOMe 13 examples
quantitative yields

HN(SiMe3)2
R OH R OTMS
)))), rt
3.2.4.13 3.2.4.14
16 examples
R = alkyl, aryl 60–98%
CH(OEt)3 Ar Ar
Ar NH2 N N
)))), rt, 50–120 min H
3.2.4.15 3.2.4.16
17 examples
64–98%

R1

NH2 CHO
O O
Neat O
+ + P HN P
O O O
)))), rt, 20–45 s
R1 R 2

3.2.4.17 3.2.4.18 3.2.4.19


R2
1 3.2.4.20
R = 4-Cl, 4-F, 4-CH3, 4-CF3
R2 = 4-Cl, 4-N(CH3)2, piperonyl, naphthyl 17 examples
80–99%
CHO O
H H OR2
H NH2 N N P
N Neat R1 S OR2
R1 S + + P(OR2)3
O )))), rt, 2–3 h O O
O

3.2.4.21 3.2.4.22 3.2.4.23


3.2.4.24
R1 = alkyl, aryl 17 examples
R2 = Me, Et 65–95%

O R1 R2
O N
Ar
R1 R2 Ar
Fe Neat
N
+ Fe
H )))), rt

3.2.4.25 3.2.4.26
3.2.4.27
R1 = R2 = alkyl 18 examples
83–98%
R2 R2
R1 R1

O
HOOC
N O N O
CHO
R1 NH2 Me
Ac2O
+
R2 COOH )))), 55–60 °C Me
CHO
N O N O
3.2.4.28 3.2.4.29 COOH
O
R1 = H, OMe, OH, Br, Cl 1
R R1
R2 = H, OMe, COOH
R2 R2

3.2.4.30 3.2.4.31
6 examples
70–85%

CHO
R2 R1 R1 R2 R2 R1 R1 R2
CHO O O
3.2.4.32
O O
HOOC N N COOH HOOC N N COOH
O O
Me Me

3.2.4.33 3.2.4.34
Not formed
O R1 O R1
Cl P Cl R1 N P N
Neat
+ HN R2 R2
R2 )))), rt, 5–15 min

3.2.4.35 3.2.4.36 3.2.4.37


18 examples
R1 = R2 = alkyl, aryl 89–98%
122 4 Metal Oxides as Catalysts/Supports in Solvent-Free Organic Reactions

Catalysis

Clean
energy Sensor
production

Nano Applications of
electronic Solar cells
devices metal oxides

Biomedical Gas storage

Environmental
remediation

Figure 4.1 Applications of metal oxides.

different spectroscopic, microscopic, and crystallographic techniques to determine


their composition and structure, an essential prerequisite for understanding MO’s
properties [6]. The thermal, morphological, optical, and structural properties of
nanomaterials are mainly characterized by thermal conductivity, high-resolution
transmission electron microscopy (HR-TEM), photoluminescence spectroscopy,
helium ion microscopy, scanning electron microscopy (SEM), UV-visible, and X-ray
diffraction (XRD) [3, 7].
Metal oxides, either as a catalyst or as support, have been employed as hetero-
geneous catalysts in several organic reactions (Figure 4.2) [8]. MOs involve simple
oxides such as alumina, titania, tin, zinc, zirconia, manganese, cerium, iron, and
porous and mesoporous metal oxides. The transition metal-based metal oxides are
often employed in several reactions such as isomerization, dehydration, dehydro-
genation, condensation reactions, and oxidation [9]. It is because of their selective
actions, ease of preparation and regeneration, and low production cost. Recent
advancements in metal oxide research have led to the development of metal oxide
catalysts with different shapes, compositions, and sizes, which help attain high
selectivity in organic synthesis. Metal oxides of a few nanometers have shown very
high activity because of their high-reaction surface area per weight of catalyst, but
they also pose the risk of loss of activity because of agglomeration and thermal
sintering. The use of various mesoporous and microporous supports such as carbon,
alumina, and silica has helped to overcome these nano metal oxides’ drawbacks.
Apart from that, this support plays an essential role in achieving the high activity of
a catalyst because of synergistic activation of the substrate. Hence, several supported
4.2 Different Metal Oxides as a Catalyst/Support in Solvent-Free Reactions 123

5%
6%
Zeolites
Oxides, Complex oxides 9%
Ion-exchange resins
41%
Phosphates
9%
Clays
Other

30%

Figure 4.2 Heterogeneous catalyst utilization in industrial applications.


Source: Reproduced from ref. [8]/Royal Society of Chemistry.

metal oxide catalysts have been utilized in organic synthesis. In recent times, metal
oxides have been used as an active site of supported catalysts and find application as
catalyst support because of the control and advanced synthesis method. The metal
oxides possess an excellent surface area and can also be tailormade. Due to their
inherent acidic, basic, and redox properties, they present a significant choice for
a variety of organic reactions. The mesoporous metal oxides are notable examples
that find application in the synthesizing of several useful intermediates, adsorption,
purification, and so forth.
The utilization of metal oxides as a catalyst or support in solvent-free organic
synthesis is of great interest. Several metal oxides have been used as catalysts and
supports in industrially significant synthesis, including titanium oxides (TiO2 ),
manganese oxides (MnO2 ), tin oxides (SnO2 ), zinc oxide (ZnO), iron oxide (Fe2 O3 or
Fe3 O4 ), and aluminum oxide (Al2 O3 ). The current chapter gives information about
the utilization of different metal oxides in solvent-free organic synthesis. It also
discusses the role of metal oxide and a solvent-free approach to achieve better results.

4.2 Different Metal Oxides as a Catalyst/Support


in Solvent-Free Reactions
4.2.1 Titanium Dioxide-Based Catalysts
Titanium dioxide (TiO2 ) has a tetragonal unit cell with a rutile phase in its elemental
form, as discovered by German chemist M.H. Klaproth, is considered among the top
20 industrially important inorganic materials [4]. TiO2 has three different phases,
that is, rutile, anatase, and brookite, and has different crystal structures [10].
124 4 Metal Oxides as Catalysts/Supports in Solvent-Free Organic Reactions

Anatase and rutile phases have a tetragonal crystal structure, whereas brookite
possesses an orthorhombic morphology [11, 12]. The mixture of rutile and anatase
phases is more effective than each phase separately. TiO2 is easy to synthesize,
inexpensive, safe to use, cheaper, with high chemical stability, commercially
accessible, biocompatible, and nontoxic, making it a widely used MO in several
applications [13]. TiO2 is used in solar cells for the production of electric energy and
hydrogen, as a corrosion-protective coating, in electric devices, in bone implants,
as a gas sensor, as an optical coating, in heterogeneous catalysis, in ceramics, as a
spacer material in the spin-valve system, and as a photocatalyst [13, 14]. TiO2 -based
materials can be prepared by different methods, such as the hydrothermal method,
the co-precipitation method, the solvothermal method, and the sol–gel method.
TiO2 and other metal oxides have been used as a catalyst for solvent-free reac-
tions. A rutile phase of TiO2 NPs has been prepared and utilized for the solvent-free
synthesis of quinoxaline derivatives from the isatin-derivative condensation with
o-phenylenediamine [15]. The TiO2 catalyst has shown higher activity than various
homogeneous and heterogeneous catalysts under optimum reaction conditions.
Compared with different solvents, solvent-free reaction conditions have helped to
achieve the highest product yield in a shorter reaction time (Figure 4.3) [15]. The
nanocrystalline TiO2 was synthesized via a green combustion method containing
citric acid as a fuel and titanium nitrate as a precursor under microwave heating
[16]. The prepared catalyst has shown excellent activity in the Xanthene synthesis
under solvent-free conditions. Different solvents’ effects were also compared
with the solvent-free condition, and the solvent-free condition has achieved high
product yield with a short reaction time. The TiO2 (30%) supported on MCM-41
(TiO2 /MCM-41) has been used in the solvent-free liquid-phase oxidation of cyclo-
hexylamine to cyclohexanone oxime [17]. The catalyst has shown promising activity
and can be easily recovered and reused for several cycles. The presence of a titanium
hydroxyl group on the catalyst surface acted as an active site for the reaction [17].
The metal oxide supported on zeolites with very minimum metal oxides (<10%)
has shown promising activity in several organic syntheses. TiO2 supported on

Figure 4.3 Effect of solvent


95 on isatin condensation
reaction with
o-phenylenediamine.
90 Source: Copyright © 2019
Elsevier B.V. All rights
reserved, reproduced with
% Yield

85
permission [15].

80

75

70
Ethanol Acetonitrile Xylene Tolune Solvent-free
4.2 Different Metal Oxides as a Catalyst/Support in Solvent-Free Reactions 125

the nanoclinoptilolite (TiO2 /NCP) catalyst has demonstrated promising activity


in the Gewald reaction to synthesize 2-aminothiophenes under the solvent-free
condition from three components [18]. The Gewald reaction mainly involves
different organic solvents like ethanol, 1,4-dioxane, and methanol, making the
process environmentally hazardous, and hence the current protocol has shown a
benign route with no separation of solvents. The current protocol has shown a good
yield of products comparable to reported literature with a short reaction duration.
TiO2 , as support, has also found a significant application in organic synthesis.
The catalyst’s active site can be introduced either by simple integration or by
functionalizing the required acid or basic site on the surface of TiO2 . Rahmani et al.
have synthesized sulfonic acid-functionalized TiO2 using chlorosulfonic acid as a
sulfating agent [19]. The prepared catalyst was completely characterized by XRD,
SEM, FT-IR, and TGA analysis to find the structure, morphology, and thermal
stability. The catalyst has shown efficient activity in the solvent-free synthesis of
different organic derivatives such as chalcone, pyrimidines, and benzothiazoles.
The nickel(II) Schiff-based complex supported on TiO2 NPs was synthesized and
showed high activity in the solvent-free synthesis of sulfoxide via oxidation of
sulfide using hydrogen peroxide oxidant at room temperature [20]. The prepared
catalyst has demonstrated excellent activity and reusability in the reaction. A
similar study has shown the application of ZnCl2 -loaded TiO2 nanomaterial as an
efficient catalyst for one-pot synthesis of propargylamine from aromatic aldehydes,
phenylacetylene, and amines under solvent-free conditions [21]. The catalyst
was prepared by a two-step method, firstly the synthesis of nanocrystalline TiO2
via the sol–gel method, followed by the impregnation of ZnCl2 . The solvent-free
reaction has shown a high product yield compared with different organic solvents
because of the reactant’s facile contact with the catalyst. Bimetallic gold–palladium
(Au–Pd) supported on different phases of TiO2 (Au–Pd/TiO2 ) as a catalyst has
been used for the solvent-free oxidation of benzyl alcohol [22]. Au–Pd supported
on the rutile TiO2 catalyst has shown high conversion of benzyl alcohol compared
with Au–Pd loaded on anatase and brookite TiO2 . The selectivity of toluene was
high, while the benzaldehyde selectivity was low for the rutile TiO2 -supported
catalyst. The TiO2 support offers the strongest interaction between the metals and
carrier and enables facile activation of molecular oxygen and enhances the catalyst
activity [22]. A novel sulfonic acid-functionalized TiO2 nanomaterial has been
prepared and used for the one-pot solvent-free synthesis of chalcone derivatives
via the Claisen–Schmidt reaction [23]. The catalyst was prepared by simple reflux
method using a mixture of nitric acid-activated TiO2 , triethylenetetramine, and
3-mercaptochloroporpyl followed by filtration (Figure 4.4). The prepared catalyst
has shown enhanced activity in the synthesis of several derivatives of chalcones,
giving a yield of between 85% and 97% in a short time of reaction. The catalyst has
also shown good reusability, with a 10% decrease in activity after five cycles. In
another study, the mesoporous vanadium ion-doped TiO2 was used for the one-pot
four-component reaction for the synthesis of polyhydroquinoline derivatives via the
Hantzsch reaction under solvent-free conditions at 80 ∘ C [24]. The prepared catalyst
and solventless protocol have shown better activity in comparison to the reported
126 4 Metal Oxides as Catalysts/Supports in Solvent-Free Organic Reactions

OH H3CO Toluene
TiO2 OH + H3CO Si Reflux, 24 h
Cl
OH H3CO

O Xylene
TiO2 O Si H2N N N NH2
Cl Reflux, str.24 h H H
O
3-TiO2PCl TETA

(i) CISO3H/CHCl3/
O str.6 h
TiO2 O Si
N N N NH2 (ii) Washed with
O H H H
EtOH
3-TETANPTiO2

O
TiO2 O Si
N N N NH
O
SO3H SO3H SO3H SO3H

TiO2-BPTETSA

Figure 4.4 Synthesis of sulfonic acid-functionalized TiO2 catalyst. Source: Copyright


© 2017 Taiwan Institute of Chemical Engineers. Published by Elsevier B.V. All rights
reserved, reproduced with permission [23].

literature. The surface-functionalized metal oxide particles have found extensive


research applications in catalysis and biotechnology. Amine-functionalized TiO2 , a
novel organic–inorganic hybrid basic catalyst, has been prepared and used for the
solvent-free synthesis of dihydropyrimidinone (DHPM) derivatives via the Biginelli
reaction at 100 ∘ C [25]. The different protic and aprotic solvents have been used and
found to have a lower yield of the product under solvent-free conditions.
Mesoporous TiO2 has unique properties, such as high surface area and surface
functional group distribution. This contributes to a synergistic effect in the catalytic
activities and the desired product selectivity. Mesoporous TiO2 has been used to sup-
port various metals such as Pd, Ru, Cu, and Cr, and the prepared catalyst has been
utilized for the solvent-free conversion of levulinic acid to γ-valerolactone (GVL)
[26]. The mesoporous nature of TiO2 and the solvent-free condition have helped to
achieve the high selectivity of GVL. The NiO-modified mesoporous TiO2 (NiO/TiO2 )
catalyst was synthesized by a simple sol–gel method used for the efficient adsorp-
tion of CO2 and the conversion of epoxides [27]. The prepared catalyst has shown
promising activity in the solvent-free synthesis of propylene carbonate (100% yield)
from the cycloaddition of propylene oxide and CO2 in the presence of tetrabutylam-
monium iodide as a co-catalyst. The prepared catalyst has shown good reusability
without significant loss in the original activity.
As a mixed oxide, TiO2 can either form a core–shell structure or be prepared in the
proper molar ratio, depending on the applications. The TiO2 -based mixed oxide has
been used to functionalize several acidity groups to generate a high-acidic catalyst.
Recently, Fe3 O4 @TiO2 magnetic nanoparticles were prepared and functionalized
Table 4.1 TiO2 -based catalyst/catalyst support for solvent-free reactions.

Yield of
Sr No. Catalyst Reaction T (∘ C) products (%) References

1. TiO2 Synthesis of 3,4-dihydropyrimidin- 70 82–98 [30]


2(1H)-ones
2. TiO2 Knoevenagel condensations 80 52–95 [31]
3. Ru/meso-TiO2 Hydrogenation of LA to GVL 150 99.4 [26]
4. Fe3 O4 @TiO2 Synthesis of quinazolinone derivative 80 92–97 [28]
5. NiO/TiO2 Synthesis of cyclic carbonate derivatives 100 86.9–100 [27]
6. Rutile-phase nano-TiO2 Synthesis of quinoxaline derivatives via 100 88–95 [15]
condensation reaction
7. ZnCl2 –TiO2 Synthesis of propargylamine derivatives 100 88–97 [21]
8. Au–Pd/TiO2 Oxidation of benzyl alcohol to 120 65.1 [22]
benzaldehyde
9. 30%TiO2 /MCM-41 Oxidation of cyclohexylamine 100 69.4 [17]
10. TiO2 -BPTETSA Claisen–Schmidt reaction 120 80–97 [23]
11. TiO2 /NCP Synthesis of substituted 100 60–90 [18]
2-aminothiophenes
12. TiO2 –CNTs nanocomposite Synthesis of pyrimidinones 80 90–98 [32]
13. ZrO2 –TiO2 Synthesis of coumarin through rt 97 [33]
Pechmann reaction
14. Nanocrystalline TiO2 Synthesis of 80 82–95 [16]
1,8-dioxooctahydroxanthenes
15. Nanocrystalline TiO2 Synthesis of 80 85–94 [16]
1,8-dioxodecahydroacridines
16. TiO2 NPs Ring-opening of epoxides with aromatic rt 95 [34]
amines
17. Pt/TiO2 Selective hydrosilylation of 1,3-diynes 70 55−98 [35]
18. TiO2 –SO4 2− Synthesis of chalcones MW heating 60–99.2 [36]
19. Phosphate-impregnated titania Aza–Michael reactions 50 80–95 [37]
20. Amine-functionalized TiO2 Biginelli reaction 100 82–96 [25]
21. Sulfonic acid-functionalized TiO2 (n-TSA) Synthesis of pyrimidinone derivatives 70 83–95 [19]
22. SO3 H-functionalized TiO2 (n-TSA) Synthesis of benzothiazole 70 79–93 [19]
23. Magnetic core–shell TiO2 nanoparticles Domino Knoevenagel–Michael 100 93–96 [38]
cyclocondensation reaction
24. Vanadium on TiO2 Synthesis of polyhydroquinoline 80 74–90 [24]
derivatives
25. Ni(II) Schiff base complex–TiO2 Oxidation of sulfides to sulfoxides rt 85–93 [20]
26. Titanate nanotubes Isomerization of α-pinene to camphene 120 76.7 [29]
27. Pd-DTP/TNT Cascade synthesis of n-pentyl 140 69 [14]
tetrahydrofurfuryl ether
130 4 Metal Oxides as Catalysts/Supports in Solvent-Free Organic Reactions

SnO2 nanoparticles have been used for the solvent-free synthesis of phtha-
lazine derivatives from the condensation reaction of three components (such
as dimedone, aromatic aldehydes, and phthalhydrazide) as a catalyst [39]. The
catalyst has shown high activity and reusability at ambient conditions, while the
solvent-free reaction condition has shown better conversion and yield because of
the absence of solvent molecule interruption between the catalyst and reactant
molecules. The SnO2 nanoparticles were prepared using the thermal decom-
position method for the solvent-free Knoevenagel condensation reaction [40].
The various synthesis parameters such as temperature, time of calcination, the
molar concentration of NaNO3 , and SnCl4 were found to affect the size of SnO2 .
The combination of SnO2 with the solvent-free method has shown an excellent
yield of condensate product for diverse aldehydes and active methylene groups.
The mesoporous structure’s presence helps to overcome the diffusion resistance,
and the high surface area of the catalyst results in ease of access to the active
site and enhances the catalytic activity [41]. Mesoporous tin oxide (Meso-SnO2 ),
having a high surface area ranging from 35 to 160 m2 g−1 and pore size from 0.08
to 0.12 cm3 g−1 , is synthesized using both template-free and template-assisted
methods and calcined at a different temperature to generate different acid
natures (Lewis and Brønsted) and the acidity strength in a sample [42]. The
prepared catalyst has shown promising activity in the solvent-free ketalization and
acetalization reactions of glycerol with acetone and benzaldehyde, respectively
(Figure 4.5) [42]. The template-assisted Meso-SnO2 catalyst has shown a good
product yield compared with various solid catalysts because of its nature and high
acidity.

OH
OH Acid-catalyzed O
reactions
O O O O
+ +
H H OH
Sn
HO OH
O O Sn
Sn
O H
O O
Mesoporous O H
Acetalization Ketalization
tin oxide
Sn Sn
O
O
O O
+ H O Sn O O O O
O
OH H Sn +
HO OH HO
Oxidation reaction OH

Epoxidation

+ H2O2
O

Figure 4.5 Mesoporous SnO2 as a catalyst for different reactions. Source: Copyright
© 2017 Elsevier B.V. All rights reserved, reproduced with permission [42].
4.2 Different Metal Oxides as a Catalyst/Support in Solvent-Free Reactions 131

Tin oxide is also used to support various metal particles and oxides and can
be functionalized with various functional groups via the postsynthesis method
to generate a new and strong acidic site in SnO2 . At 8 bar hydrogen pressure,
platinum-loaded SnO2 (Pt/SnO2 ) has been used in the solvent-free reductive
dimerization of various carboxylic acids to esters [43]. The detailed characterization
reveals that the SnO2 -supported catalyst is better because of the presence of Sn4+
Lewis acid sites and has shown a better activity compared with other supported
catalysts with very high reusability. Sulfated tin oxides (STOs) were prepared via
thermal decomposition of stannous sulfate with varying sulfur content (1–8%)
[44]. The STO was used in the solvent-free liquid-phase synthesis of isosorbide via
sorbitol dehydration (Figure 4.6) [44]. The prepared catalyst has shown high activity
in comparison to reported literature in a short reaction time. Various reaction and
catalyst synthesis parameters were optimized, and catalyst STO-400 (calcined at
400 ∘ C) have provided 65% isosorbide selectivity at 100% conversion of sorbitol in
two hours at the reaction temperature of 180 ∘ C. The acidity of sulfated metal oxide
is generally controlled by the amount of sulfur present in the catalyst [14]. The var-
ious sulfate precursors, such as sulfuric acid, chlorosulfonic acid, and ammonium
sulfate, have been used to get a highly active catalyst [45]. The hydroxylated SnO2
has been prepared and wet impregnated by sulfuric acid solution followed by calci-
nation at different temperatures, resulting in the STO with 5–30 wt% sulfur content
[46]. The prepared catalyst has been used in the solvent-free synthesis of coumarin
derivatives via the Pechmann condensation reaction of ethyl acetoacetate and
resorcinol. The catalyst has shown an excellent yield of product, which increases
with an increase in the sulfur content. A similar type of catalyst, that is, sulfated
SnO2, was used for the solvent-free Friedel craft-acylation of toluene with acetic
postanhydride and was found to have high activity compared with other catalysts
[47]. Further, to enhance the sulfated SnO2 catalyst activity, microwave heating has
been used for the solvent-free synthesis of naphthopyranopyrimidine derivatives
[48]. The utilization of microwave under solvent-free conditions helped to get mild,
rapid preparation with a high yield of different derivatives (72–96%) in a short
reaction time. A phosphate-modified mesoporous tin oxide (m-SnO2 ) as a catalyst
has been used for the solvent-free synthesis of hydroquinone diacetate [49]. The
reaction involved hydroquinone and acetic anhydride having a dual application,
that is, reactant and solvent. The catalyst and solvent-free protocol have helped
to get a high yield of products (93.1%). Overall, SnO2 as a catalyst and support
for functionalization of different groups has shown a significant catalytic activity

OH
HO OH H OH
HO OH
O
OH OH OH +
OH O OH
O –H2O O
HO –H2O OH H
HO
OH OH
1,4-Sorbitan 1,5-Sorbitan Isosorbide
Sorbitol

Figure 4.6 Dehydration of sorbitol to isosorbide. Source: Copyright © 2013 Elsevier B.V. All
rights reserved, reprinted with permission [44].
132 4 Metal Oxides as Catalysts/Supports in Solvent-Free Organic Reactions

Table 4.2 SnO2 -based catalyst/catalyst support for solvent-free reactions.

Temperature Yield of
Sr No. Catalyst Reaction (∘ C) products (%) References

1. SnO2 Nps Synthesis of phthalazine 80 84–96 [39]


derivatives
2. Mesoporous SnO2 Glycerol acetalization 100 48 [42]
3. Mesoporous SnO2 Glycerol ketalization 60 90.1 [42]
4. Tin oxide Knoevenagel rt 80 [40]
nanoparticles condensation
(NP-SnO2 )
5. SnO2 /SiO2 Imidazole-derivative 80 87–95 [50]
synthesis
6. 5 wt% Pt/SnO2 Reductive dimerization of 100 73–95 [43]
various carboxylic acids
7. Sulfated tin oxides Dehydration of sorbitol 180 65 [44]
(STO)
8. Sulfated tin oxide Friedel–Crafts acylation of 70 98 [47]
toluene
9. Sulfated tin oxide Acylation of a variety of rt 88–90 [51]
(SO4 2− /SnO2 ) alcohols, phenols, and
amines
10. Nanostructure Pechmann condensation 120 95.8 [46]
sulfated tin oxide
11. (H3 PO4 /m-SnO2 ) Synthesis of hydroquinone 120 93.2 [49]
diacetate

in the synthesis of different fine chemicals and intermediates under solvent-free


conditions (Table 4.2).

4.2.3 Manganese Oxide-Based Catalysts


Manganese oxides (MnOx) possess variable oxidation states of Mn ions and also
contain lattice oxygen vacancies, which makes them interesting supports for a
variety of redox reactions. Moreover, MnOx is suitable for scale-up because of its
low toxicity, high stability, relatively low cost of MnOx materials, and abundant
commercial availability at large scale [52]. The most common forms of manganese
oxides are MnO, MnO2 , Mn2 O3 , and Mn3 O4 , which have the manganese oxidation
states of +II, +IV, and +III and have been extensively used in catalysis, alkaline
and rechargeable batteries, sensors, wastewater treatment, and supercapacitors
[53, 54]. The redox properties of manganese oxides have attracted lots of attention
in the oxidation and hydrogenation reactions. Manganese oxide, or MnO2 -based
nanomaterials, has been used extensively in various catalytic reactions. On the
basis of the different ways of linkage of the basic octahedral unit [MnO6 ], MnO2
possesses various polymorphic forms such as 3D structure (λ-MnO2 ), chain-like
4.2 Different Metal Oxides as a Catalyst/Support in Solvent-Free Reactions 133

tunnel structure (𝛼-, 𝛽-, and 𝛾-MnO2 ), and layered or sheet structure (𝛿-MnO2 )
[55, 56]. The activity of these materials is a function of their morphology and size
and can be tuned by the synthesis method’s help or by employing suitable thermal
treatments. Manganese oxide-based catalysts have been used in several solvent-free
reactions, as shown in Table 4.3.
A manganese oxide-based catalyst has been used for the solvent-free aerobic
oxidation of benzyl alcohol under microwave heating [70]. Different manganese
oxides such as MnO2 , Mn2 O3 , Mn3 O4 , and MnO have been tested at 80 ∘ C, and
MnO2 was found to have a high activity because of its high surface area. The
MnO2 catalyst has shown high selectivity to benzaldehyde and is stable after
three reuses. Similarly, amorphous MnO2 has been used for solvent-free aerobic

Table 4.3 MnO2 -based catalyst/catalyst support for solvent-free reactions.

Yield of
Sr No. Catalyst Reaction T (∘ C) products (%) References

1. Amorphous MnO2 Aerobic oxygenation of 130 87–99 [57]


primary amines to amides
2. Mn3 O4 Oxidation of ethylbenzene 130 41 [58]
to acetophenone
3. MnO2 /GO Claisen–Schmidt 110 86–98 [59]
condensation
4. MnO2 @Zeolite-Y Synthesis of 4-arylidene- 100 85–97 [60]
Nanoporous isoxazolidinones
5. MnO2 -doped MgO Synthesis of E-ethyl rt 100 [61]
cinnamate by Wittig
reaction
6. MnO2 –MWCNT Biginelli reaction under 30–100 W 92–97 [62]
nanocomposites microwave radiation
7. 1.0 mol% Mo–MnO2 Allylic oxidation of 60 32.4 [63]
cyclohexene
8. Phosphate Synthesis of xanthene 100 85–97 [64]
anchored MnO2 laser dyes
9. Au/MnO2 Aerobic oxidation of 120 40.1 [65]
benzyl alcohol
10. Pd supported on Aerobic oxidation of 120 67 [66]
MnOx benzyl alcohol
11. Co3 O4 –MnO2 /ZnO Synthesis of some 100 75–96 [67]
bis(indolyl)methane
derivatives
12. Co3 O4 –MnO2 /ZnO Synthesis of 2-aryl-1- 80 97 [68]
arylmethyl-1H-
1,3-benzimidazoles
13. Co3 O4 –MnO2 /ZnO Synthesis of 3,4- 80 98 [69]
dihydropyrimidin-
2(1H)-ones-thiones
4.2 Different Metal Oxides as a Catalyst/Support in Solvent-Free Reactions 135

Mo in various quantities has shown an influence on the redox properties and Lewis
acid sites of Mo-supported catalysts. The catalyst has shown a promising result with
high selectivity. Under solvent-free conditions, gold (Au) nanoparticles supported
on manganese oxide nanorods (β-MnO2 ) have been synthesized and used in the
aerobic oxidation of benzyl alcohol [65]. The prepared catalyst has shown better
activity than the commercial MnO2 because of the well-defined nanorod structure,
which helps to get a synergistic effect to get the enhanced activity of the prepared
catalyst. In a similar kind of reaction, palladium (Pd) supported on different man-
ganese oxides (MnO, Mn3 O4 , Mn2 O3 , and MnO2 ) were prepared and studied in the
solvent-free oxidation reaction [66]. Because of the synergistic effect between metal
and support, the activity of manganese oxide-based catalysts was better than that
of CeO2 - and Fe3 O4 -supported catalysts. The catalyst has shown good reusability
for four runs. In another study, phosphate-anchored MnO2 (P-MnO2 ) catalysts
have been studied in the solvent-free synthesis of xanthene and benzoxanthene
derivatives [64]. The prepared catalyst has shown a high activity toward different
derivatives giving a yield between 85% and 98% at optimized conditions. The
reaction protocol using P-MnO2 catalyst has shown a high yield of product in a
short reaction time and presents an easy workup because of solvent-free conditions.
Another method of increasing the activity of an MnO2 -based catalyst is the syn-
thesis of mixed oxides with different mole ratios. Cobalt-doped manganese oxide
supported on zinc oxide (Co3 O4 -MnO2 /ZnO) catalyst has a cubic phase of Co3 O4
and Mn2 O3 and has been used in several solvent-free reactions. Ahmadian et al. have
used the prepared catalyst in the solvent-free synthesis of benzimidazole deriva-
tives [68]. The catalyst activity was measured under various solvents and solvent-free
conditions, where the prepared catalyst shows a high activity and yield toward the
product under solvent-free conditions. Similarly, the supported mixed oxide cata-
lyst has also shown high activity and yield in the solvent-free synthesis of thione
derivatives [69] and bis(indolyl) methane derivatives [67]. The overall process has
been green, with ease of product separation because of solvent-free conditions, and
required a short reaction time compared with reported literature.
The manganese oxide and supported catalyst have shown promising results
toward the solvent-free synthesis of several products via different reactions. man-
ganese oxide as a catalyst is mainly used in the oxidation reaction because of its
different oxidation states, which helps to achieve high oxidation activity.

4.2.4 Zinc Oxide-Based Catalysts


Zinc oxide (ZnO) is a semiconductor oxide with a high breakdown field strength,
high electron mobility, and a melting point of around 2248 K [4]. ZnO is cheap,
nontoxic, stable at high temperatures, and resistant to different chemicals, making
it one of the most extensively used metal oxides for many applications such as
solar cells, gas sensors, antibacterial materials, photoactivity, flame-retardancy,
catalysis, and semiconductor manufacturing [71, 72]. ZnO possesses three different
morphologies, such as cubic (zinc blende), rocksalt, and wurtzite, among which
wurtzite is the most stable structure [4, 73]. Like other metal oxides, ZnO can
136 4 Metal Oxides as Catalysts/Supports in Solvent-Free Organic Reactions

also be prepared by various methods like sol–gel, hydrothermal, co-precipitation,


hydrolysis, peptization, and so forth.
In the last decade, ZnO NPs have been used in several crucial organic trans-
formations as catalysts under solvent-free conditions, such as 2-aminothiophene
synthesis via the Gewald reaction [74], DHPM synthesis [75], 2-acylpyrroles and
their derivative synthesis [76], flavon synthesis [77], and many more [78]. Tang et al.
have recently demonstrated ZnO as an efficient catalyst for the room temperature
synthesis of chloroesters under solvent-free conditions via the reaction of various
ethers and acyl chloride [78]. The ZnO catalyst has shown a moderate-to-good
yield (73–94%) of a range of products for different reacting ethers such as tri-,
tetra-, penta-, and hexacyclic ethers. The catalyst has shown better reusability
after three cycles and can be potentially used for several acylative ether cleavages.
Recently, ZnO nanorods have been synthesized via a simple reflux method and
have shown a promising result in the solvent-free synthesis of thiophene derivatives
via a multicomponent reaction of isoquinoline, isothiocyanates, alkylbromides,
and activated acetylenic compounds at room temperature [79]. The catalyst has
shown great potential under solvent-free conditions, giving excellent product yield
under mild reaction conditions and ease of separation and operational procedure.
Similarly, ZnO nanorods have also shown promising results in synthesizing
pyrimido-isoquinolines and pyrimidoquinoline derivatives under solvent-free
synthesis [80]. The catalyst with solvent-free conditions has shown a synergistic
effect in the product yield under mile reaction conditions and easy operational
procedure. In another study, Rasal et al. prepared ZnO nanoflowers (ZnO NFs) via a
simple wet chemical method (Figure 4.7) and used them for ultrasonication-assisted
solvent-free synthesis of α-aminophosphonates [81]. The catalyst has shown better
activity than various reported literature and given better activity under solvent-free

OH–

pH = 11 Refluxing at 100 °C

Zn2+
ZnO nuclei
Zn(OH)42–

Aging at 150 °C

ZnO crystals
ZnO nanoflowers

Figure 4.7 ZnO nanoflower synthesis mechanism. Source: Copyright © 2018 Taylor &
Francis. All rights reserved, reproduced with permission [81].
4.2 Different Metal Oxides as a Catalyst/Support in Solvent-Free Reactions 137

conditions, giving excellent α-aminophosphonate-derivative yield. A similar kind of


ZnO NFs has been prepared and found to be a very active catalyst for synthesizing
chalcones under solvent-free conditions and microwave heating [82].
ZnO Nps have also been supported on a variety of supports in order to
increase the activity and prevent the agglomeration of NPs. Naz et al. have
reported ZnO dispersed in an N-doped carbon matrix (ZnO/NCs) as a promising
catalyst for the solvent-free polymerization of ε-caprolactone to cyclic poly-
caprolactone [83]. The catalyst was prepared by controlled thermal treatment of
Zn-DABCO (1,4-diazabicyclo[2.2.2]-octane) at different temperature conditions.
The basic and acidic properties of the ZnO/NCs catalyst have helped to synthesize
high-molecular-weight polycaprolactone. The catalyst has shown a good yield
with high reusability with no loss in activity for seven runs. In another study,
magnesium oxide-supported nano-ZnO (ZnO/MgO) has shown promising results
in synthesizing various benzimidazole derivatives under solvent-free conditions
[84]. The solvent-free method, the use of cheap and reusable catalysts, and the
excellent yield of benzimidazole derivatives were the current protocol’s advantages.
Under solvent-free conditions, ZnO-supported metal oxide, mixed metal oxides,
and NPs as catalysts have been explored in organic synthesis. A ZnO-supported
CuO/Al2 O3 catalyst has been prepared by the co-precipitation method and demon-
strated as a promising catalyst for solvent-free synthesis of propargylamines in high
yield [85]. The prepared catalyst activity was compared under solvent-free and dif-
ferent solvents such as toluene, water, ethanol, and acetonitrile and showed bet-
ter results under solvent-free conditions because of the absence of solvation effects
caused by different solvents. Further, the current protocol has shown the highest
activity compared with published literature at mile reaction conditions. Banaker
et al. have prepared a ZnO decorated with graphene oxide (GO/ZnO) nanocom-
posite as a catalyst having both Bronsted and Lewis acid sites [86]. The prepared
catalyst has shown promising results in the selective synthesis of xanthenedione
derivatives via solvent-free multicomponent reaction, giving an excellent yield of dif-
ferent derivatives under mild reaction conditions with low catalyst loading. The easy
separation because of solvent-free conditions and the better stability and reusability
of the catalyst were the significant advantages of the current protocol. Palladium sup-
ported on magnetically separable ZnO (Pd/Fe3 O4 /ZnO) has been an efficient catalyst
for the solvent-free synthesis of 2-oxazoline benzoxazoles from aromatic nitriles [87].
The prepared catalyst under the solvent-free protocol has significant features such
as short reaction time, simple operation, high yields of products, and negligible loss
of activity after reusability.
Pure ZnO as a catalyst in various shapes or supported on various materials and
as a catalyst support has demonstrated excellent performance in various reactions
while being inexpensive, nontoxic, and highly reusable (Table 4.4).

4.2.5 Aluminum Oxide-Based Catalysts


Aluminum(III) oxide, also called alumina (Al2 O3 ), is an inexpensive and the most
widely used metal oxide in catalytic and adsorption applications because of its
138 4 Metal Oxides as Catalysts/Supports in Solvent-Free Organic Reactions

Table 4.4 ZnO-based catalyst/catalyst support for solvent-free reactions.

Yield of
Sr No. Catalyst Reaction T (∘ C) products (%) References

1. ZnO Nps Gewald reaction 100 77 [74]


2. ZnO Dihydropyrimidinone 80 92–97 [75]
synthesis
3. ZnO Flavon synthesis MW 86–88 [77]
heating
4. ZnO Nps Synthesis of chloroesters rt 79–95 [78]
5. ZnO nanorods Synthesis of thiophene rt 75–93 [79]
derivatives
6. ZnO nanorods Pyrimido-isoquinoline- rt 85–93 [80]
derivative synthesis
7. ZnO nanorods Pyrimidoquinoline- rt 83–87 [80]
derivative synthesis
8. ZnO nanoflowers Synthesis of US 84–99 [81]
α-aminophosphonates assisted rt
9. Bioderived ZnO NFs Chalcone synthesis 100 88–98 [82]
10. ZnO/NCs Polymerization of 160 99 [83]
ε-caprolactone
11. ZnO/MgO Synthesis of 60 75–90 [84]
benzimidazole derivatives
12. ZnO-supported Synthesis of 80 90 [85]
CuO/Al2 O3 propargylamines
13. GO/ZnO Synthesis of 100 89–99 [86]
xanthenediones
14. Pd/Fe3 O4 /ZnO Synthesis of 2-oxazolines 120 67–96 [87]
15. Pd/Fe3 O4 /ZnO Synthesis of benzoxazoles 120 67–95 [87]

insolubility in water and organic solvents [88, 89]. Al2 O3 is generally produced by
thermal decomposition of alumina oxides and hydroxides and exists in different
polymorphic phases such as α-Al2 O3 , β-Al2 O3 , γ-Al2 O3 , δ-Al2 O3 , θ-Al2 O3 , η-Al2 O3 ,
κ-Al2 O3 , χ-Al2 O3 , and ρ-Al2 O3 [88–90]. The different nanomaterials based on
alumina have found application in dental composites, drug delivery carriers,
ceramics, catalysis, absorbent, abrasive, surgical implants, and electronics [88, 90].
Among the various alumina forms, γ-Al2 O3 is the most widely used catalyst and
catalyst support because of its high surface area, acid and base characteristics, and
better thermal and chemical stabilities [90]. Several organic transformations have
been reported using Al2 O3 -based nanomaterials (Table 4.5).
Because of its acidic and basic properties, Al2 O3 has shown a promising result in
various organic transformations as a catalyst. Yadav et al. have used the basic alu-
mina catalyst to synthesize N-arylamines via the solvent-free reaction of activated
aryl halides and secondary amines under microwave irradiation [92]. The reported
Table 4.5 Al2 O3 -based catalyst/catalyst support for solvent-free reactions.

Yield of
Sr No. Catalyst Reaction T (∘ C) products (%) References

1. Basic Al2 O3 Aminolysis of triglyceride rt 98 [91]


2. Basic Al2 O3 Synthesis of N-arylamines MW heating 73–90 [92]
3. γ-Al2 O3 Synthesis of benzothiazoles 120 85–92 [93]
4. γ-Al2 O3 Synthesis of tetrasubstituted imidazoles Heating 85–94 [94]
5. Al2 O3 Biginelli reaction rt 70–91 [95]
6. γ-Al2 O3 /SbCl5 Synthesis of 2,3-dihydroperimidines rt 50–95 [96]
7. Al2 O3 /MeSO3 H Synthesis of quinoline derivatives rt 80–91 [97]
8. Al2 O3 /KF Synthesis of vinyl sulfides 60 52–95 [98]
9. Basic Al2 O3 /PCl5 Synthesis of nitriles from aldehydes 120 84–94 [99]
10. B2 O3 /Al2 O3 Enamination of β-dicarbonyl compounds with amines rt 73–95 [100]
11. Au/Al2 O3 Oxidation of benzyl alcohol to benzoic acid 100 59 [101]
12. Cu@Mg/γ-Al2 O3 Hydrogenation of furfural to furfuryl alcohol 170 93 [102]
13. Mesoporous ZrO2 –Al2 O3 O-methoxymethylation reaction 40 97–99 [103]
14. TiCl2 /nano-γ-Al2 O3 Synthesis of 1,4-dihydropyridine derivatives 90 62–95 [104]
15. Sulfonic acid-functionalized Synthesis of per-O-acetylation carbohydrate derivatives 50 70–98 [105]
nano-γ-Al2 O3
(Continued)
Table 4.5 (Continued)

Yield of
Sr No. Catalyst Reaction T (∘ C) products (%) References

16. Mo/γ-Al2 O3 Biginelli reaction 100 82 [106]


17. H3 PO4 –Al2 O3 Synthesis of benzo[4, 5] imidazo[1,2-a]pyrimidine 100 76–94 [107]
derivatives
18. H3 PO4 –Al2 O3 Synthesis of 2,3-dihydroquinazolin-4(1H)-one 100 80–95 [107]
derivatives
19. H3 PO4 –Al2 O3 Synthesis of 2H-indazolo[2,1-b]phthalazine-trione 100 76–93 [107]
derivatives
20. K2 CO3 /Al2 O3 Synthesis of α-aminophosphonates 140 94 [108]
21. SbCl3 –Al2 O3 Synthesis of benzo[b]1,4-diazepines 60 83–94 [109]
22. KF–Al2 O3 Synthesis of coumarin derivatives MW 64–80 [110]
23. KF–Al2 O3 Synthesis of Benzocoumarin derivatives MW 60–70 [110]
24. CuO/ZnO/Al2 O3 Synthesis of propargylamines 80 86–92 [85]
25. Ni/Al2 O3 Hydrogenation of levulinic acid to GVL 200 92 [111]
26. Pt–Sn/γ-Al2 O3 β-alkylation of secondary alcohols 155 80–96 [112]
4.2 Different Metal Oxides as a Catalyst/Support in Solvent-Free Reactions 143

properties, a simple method of synthesis, high surface area, and volume ratio
[116]. Iron oxide NPs can be synthesized by a wide range of synthesis methods
such as sol–gel, hydrothermal, microemulsion, and sonochemical. The potential
field of iron oxide nanoparticle applications is in biomedical applications such as
drug delivery; diagnostic magnetic resonance imaging (MRI); thermal therapy;
durable pigments (colored concretes, coatings, and paints); and catalysis. Fe3 O4 and
Fe2 O3 as catalysts and catalyst supports have been used for several reactions such as
oxidation, hydrogenation, coupling reactions, and multicomponent reactions. Apart
from that, separation from reaction media because of its magnetic nature offers
the advantage of the utilization of these MO for supporting different functional
groups and metal oxides to make a mixed metal oxide-based material (either having
the core–shell structure or simple mixed oxide) to get a highly active catalyst is
also extensively studied. The application of Fe3 O4 - and Fe2 O3 -based catalysts in
solvent-free organic synthesis has been given in Table 4.6.

4.2.6.1 Fe3 O4 -Based Catalyst/Support


Magnetite (Fe3 O4 ) possesses ferromagnetism and is a highly stable form of iron
oxide and hence is used to synthesize catalysts to utilize its magnetic property. Bare
iron oxides possess redox and acid properties and have been used for solvent-free
reactions. A Fe3 O4 -based nanocatalyst has shown a promising result in isatin ketal-
ization under solvent-free conditions and sonic irradiation [117]. Fe3 O4 synthesized
via chemical methods has shown promising results, giving good-to-excellent
product yield under solvent-free and ambient reaction conditions. Using Fe3 O4
NPs with an average particle diameter of 45 nm, they were used for the one-pot
synthesis of 1,8-dioxo-decahydroacridine derivatives via multicomponent reactions
(MCRs) of aldehyde, amine, and dimedone under solvent-free conditions [118].
The prepared catalyst has shown excellent activity under a solvent-free condition
in a short reaction time compared with several solvents. The catalyst was quickly
recovered using an external magnetic field and showed a negligible loss in activity
up to six cycles. In another MCR, ultrasound-assisted synthesis of 2-naphthols
under solvent-free conditions was carried out using Fe3 O4 as a catalyst, which has
shown excellent activity, giving more than 90% yield of product at 80 ∘ C and a short
reaction duration [119]. In another study, Fe3 O4 NPs have shown promising results
in the solvent-free synthesis of vinyl thioethers from terminal alkynes and thiols
via a hydrothiolation reaction [120]. Under solvent-free conditions, the catalyst
has given good-to-excellent yields of different vinyl thioether derivatives in a short
reaction time and at ambient reaction conditions with negligible loss of activity
for five cycles. In addition to these, bare magnetite NPs have also shown excellent
activity in synthesizing various products, including as α-aminophosphonates
via multicomponent coupling [121], calix[4]resorcinarene derivatives [122], and
benzoxanthene derivatives [123] via condensation reaction with high reusability
and ease of separation via the magnetic field.
Fe3 O4 NPs can be easily grafted onto support or can be easily functionalized to
induce magnetic properties into nanomaterials. A novel indium nanoparticle sup-
ported on magnetic carbon nanotube (Fe3 O4 -CNT-IN) was synthesized and used for
Table 4.6 Iron oxide-based catalyst/catalyst support for solvent-free reactions.

Yield of
Sr No. Catalyst Reaction T (∘ C) products (%) References

1. Fe3 O4 Satine ketalization 75 71–97 [117]


2. Fe3 O4 Synthesis of 1,8-dioxo-decahydroacridine derivatives 120 70–95 [118]
3. Fe3 O4 Ultrasound-assisted synthesis of 2-naphthol derivatives 80 90–98 [119]
4. Fe3 O4 Hydrothiolation reaction of alkynes and thiols 40 55–88 [120]
5. Fe3 O4 Synthesis of α-aminophosphonate derivatives 50 76–94 [121]
6. Fe3 O4 Synthesis of calix[4]resorcinarene derivatives 120 78–95 [122]
7. Fe3 O4 Synthesis of benzoxanthene derivatives 90–110 80–95 [123]
8. Fe3 O4 -CNT-IN Synthesis of isochromeno[4,3-c]pyrazole-5(1H)-one 100 85–94 [124]
derivatives
9. Fe3 O4 -CNT-PTh Synthesis of octahydroquinazolinone derivatives 80 83–95 [125]
10. Fe3 O4 -CNT-PTh Synthesis of dihydropyrimidinone derivatives via Biginelli 80 84–92 [125]
reaction
11. KI@Fe3 O4 Synthesis of β-nitroalcohol via Henry reaction rt 65–99 [126]
12. LAIL@Fe3 O4 Synthesis of polyhydroquinoline derivatives 80 79–97 [127]
13. LAIL@Fe3 O4 Synthesis of propargylamine derivatives 80 39–85 [127]
14. Fe3 O4 @firboin-SO3 H Synthesis of tetrazole derivatives 100 76–92 [128]
15. S-ZrO2 @Fe3 O4 particle (SFZ) Synthesis of ethyl levulinate 120 96 [129]
16. Fe3 O4 @SiO2 –SO3 H Synthesis of pyrazole-fused isocoumarin derivatives 100 78–85 [130]
17. Fe3 O4 @SiO2 –SO3 H Condensation of aldehydes with acetic anhydride rt 85–98 [131]
18. Fe3 O4 @SiO2 –SO3 H Synthesis of xanthene derivatives 130 90–98 [132]
19. Fe3 O4 @SiO2 –TiCl3 Synthesis of 4H-pyrimido[2,1-b]benzothiazole derivatives 100 82–90 [133]
20. Fe3 O4 @SiO2 @imidazol-bisFc Synthesis of dihydropyrano[2,3-c]coumarin derivatives 100 80–94 [134]
[HCO3 ]
21. Fe3 O4 @SiO2 –MoO3 H Synthesis of 1,8-dioxo-decahydroacridine derivatives 100 86–94 [135]
22. Fe3 O4 @SiO2 –ZrCl2 Synthesis of tetrahydrobenzimidazo[2, 100 88–95 [136]
1-b]quinazolin-1(2H)-one
23. Fe3 O4 @SiO2 –ZrCl2 Synthesis of 2H-indazolo[2,1-b]phthalazine-trione 110 88–95 [136]
derivatives
24. Fe2 O3 Synthesis of spiro[4H-pyran-oxindoles]derivatives 90 76–89 [137]
25. γ-Fe2 O3 Synthesis of 2-phenylquinazoline derivatives 85 82–96 [138]
26. γ-Fe2 O3 Friedel–Crafts benzoylation of arenes rt 73–92 [139]
27. Ag–Al2 O3 @Fe2 O3 Acylation of substituted benzyl alcohol 80 94–98 [140]
28. γ-Fe2 O3 @KIT-1 Synthesis of 1-(aryl(piperidin-1-yl)methyl)naphthalen-2-ol 80 70–90 [141]
derivatives
29. CuO@γ-Fe2 O3 Synthesis of substituted guanidines via hydroamination of 80 60–97 [142]
carbodiimides
30. γ-Fe2 O3 @HAp-SO3 H Synthesis of pyrido[2,3-d]pyrimidine derivatives 60 88–95 147
31. γ-Fe2 O3 @HAp-SO3 H N-formylation of amines rt 90–97 [143]
32. γ-Fe2 O3 @HAp-TUD Synthesis of pyranopyridine derivatives 100 82–98 [144]
33. γ-Fe2 O3 -HAp-(CH2 )3 -NHSO3 H Synthesis of quinoline derivatives via Friedlander reaction rt 93–97 [145]
34. GO/α-Fe2 O3 /CuL Synthesis of 2-amino 4H-benzo[h]choromene derivatives 70 89–96 [146]
35. CoCuFe2 O4 NPs Synthesis of pyranopyrazol derivatives rt 89–92 [147]
36. ZnFe2 O4 NPs Claisen–Schmidt condensation for chalcone synthesis 80 94–97 [148]
37. NiFe2 O4 @SiO2 –FHS NPs Synthesis of 1,8-dioxodecahydroacridines 80 82–94 [149]
146 4 Metal Oxides as Catalysts/Supports in Solvent-Free Organic Reactions

the synthesis of isochromeno[4,3-c]pyrazole-5(1H)-one derivative using ninhydrin


and aryl hydrazones under solvent-free conditions [124]. The prepared catalyst has
shown a promising result and is used for large numbers of substrates and is found
to give a high product yield in a short reaction time. Furthermore, because of the
catalyst’s magnetic nature, the separation of catalyst and reaction workup was very
simple, and catalysts have shown no distinguished loss in activity after being reused
for five consecutive cycles. Similarly, polythiophene immobilized on magnetic
carbon nanotube (Fe3 O4 -CNT-PTh) has been used in the solvent-free synthesis of
octahydroquinazolinone and DHPM derivatives via the Biginelli reaction, where
the solvent-free condition and prepared catalyst have given good-to-excellent yields
of different products [125]. Functionalization of bare Fe3 O4 particles directly is also
an attractive method to prepare the highly active catalyst. Potassium iodide loaded
on Fe3 O4 (KI@Fe3 O4 ) catalyst was prepared and used for solvent-free synthesis of
β-nitroalcohol via Henry reaction (Figure 4.8) [126]. The catalyst under solvent-free
conditions gave an excellent yield of different nitroalcohols at room temperature in a
short reaction time and showed enhanced activity compared with different catalysts
and solvents. Similar to above, ionic liquid-functionalized Fe3 O4 NPs have shown
promising activity in different MCRs. Nguyen et al. have prepared the Lewis acidic,
that is, 3-(3-(trimethoxysilyl)propyl)-1H-imidazol-3-ium chlorozincate(II) ionic
liquid supported on Fe3 O4 (LAIL@MNP) and used for the solvent-free synthesis
of benzoxanthenes, pyrroles, propargylamine, and polyhydroquinoline heterocycle
derivatives [127]. The catalyst has shown excellent activity under ultrasound with
a short reaction time, easy separation of catalyst NPs, easy workup because of
solvent-free conditions, and high reusability.
The utilization of Fe3 O4 particles as a core in the catalyst serves a dual purpose,
one is the ease of separation because of its magnetic property, and second, it helps

1 : 2 molar ratio
Fe2+ Fe3+ NH3.H2O
BLACK ppt.
Fe3O4 NP

Filtered,
Distilled water Stirred at 65 °C Aged for washed with
60 min water, and dried

O
R1 NO2
R H
Water
Room temp, evaporated Fe3O4 NPs KI solution
solvent-free
Fe3O4 NP
OH Dried and
NO2 calcined at 450 °C
R
KI Stirred for 30 min
R1
K2O

Figure 4.8 KI-loaded Fe3 O4 catalyst for the synthesis of β-nitroalcohol via Henry reaction.
Source: Copyright © 2019, Springer Science Business Media, LLC, part of Springer Nature.
All rights reserved, reproduced with permission [126].

OH O O R1 O
R1CHO+ NH4OAc
Silica

(MeO)3Si SH
Silica

OH O Si SH
O O EtO OEt
Toluene, Reflux, 24 h
OH O 2 OEt N
H
Ionic
(1) 30% H2O2
(2) 0.5 M H2SO4 liquid 60 °C
(3) Distilled H2O layer

Ionic
Ionic
Silica-SO3

Silica-SO3

liquid
Silica-SO3

Silica-SO3

Ionic liquid
Silica-SO3
Silica-SO3

Silica-SO3

liquid

Ionic O
O liquid
Silica
Silica

O Si SO3H O Si SO3H

O Ionic [BMIM][PF6] O
liquid
Ionic
liquid
Ion
ic
liquid
Support material
181

Carbon-Based Materials as Catalysts/Supports


in Solvent-Free Organic Reactions
Nagy L. Torad 1,2,3 , Ahmed A.M. El-Amir 4 , Fathy M. Hassan 2 ,
Amira A. El-Maddah 4 , Wael A. Amer 2,5 , and Mohamad M. Ayad 2,6
1 National Institute for Materials Science (NIMS), International Center for Materials Nanoarchitectonics

(WPI-MANA), 1-1 Namiki, Tsukuba, Ibaraki 305-0044, Japan


2
Tanta University, Chemistry Department, Faculty of Science, Tanta 31527, Egypt
3
The University of Queensland, Australian Institute for Bioengineering and Nanotechnology (AIBN),
School of Chemical Engineering, Brisbane, QLD 4072, Australia
4
Refractory & Ceramic Materials Division (RCMD), Central Metallurgical R&D Institute (CMRDI),
P.O. Box 87 Helwan, Cairo 11421, Egypt
5
Department of Chemistry, College of Science, University of Bahrain, Sakhir 32038, Bahrain
6 Egypt-Japan University of Science and Technology (E-JUST), Institute of Basic and Applied Sciences,

New Borg El-Arab City, Alexandria 21934, Egypt

6.1 Introduction
Historically, scientists have made enormous progress in realizing economical
synthetic alternatives to advance new processes that are not only efficient, selective,
and high yielding but also environmentally friendly protocols to synthesize a wide
variety of essential and vital compounds from naturally occurring materials by
greener chemical transformations to produce the required intermediates/chemicals
that can be assembled in a truly practical fashion [1]. Catalysis is a chief pillar of
green chemistry that allows target molecules to be prepared in a shorter time and
with greater energy efficiency, enabling the sustainability of synthesis processes
and achieving dual goals of environmental protection and economic benefits
[2, 3]. It is generally assumed that organic reactions are carried out in a solvent
medium. The rationale behind this hypothesis is that the solvent forms a homo-
geneous solution with the reactants, facilitating their stirring and allowing them
to interact effectively. In many cases, the solvent is an inevitable component of
the reaction where it associates with the reaction partners via the solvation of the
reactants, products, or transition states, affecting the yield and speed of reaction
considerably [4]. It is worth mentioning that catalyst efficiency is frequently highly
sensitive to the nature of the solvent, and the largest contributors to the magnitude
of the product are organic solvents. Although many organic solvents, e.g. ethers,
hydrocarbons, esters, chlorinated hydrocarbons, alcohols, sulphoxides, ammonia,
carbon disulfide, and amide derivatives, are ecologically harmful and require
expensive remediation, alternative measures are the “greenness” of a reaction and
Solvent-Free Methods in Nanocatalysis: From Catalyst Design to Applications, First Edition.
Edited by Rafael Luque, Manoj B. Gawande, Esmail Doustkhah, and Anandarup Goswami.
© 2023 WILEY-VCH GmbH. Published 2023 by WILEY-VCH GmbH.
182 6 Carbon-Based Materials as Catalysts/Supports in Solvent-Free Organic Reactions

the volume productivity [5]. For instance, chlorinated hydrocarbons, widely used
on large scales for organic reactions, have posed a severe threat to the environment
due to their high volatility and toxicity.
As global environmental legislation becomes stricter, environmental concerns
should be addressed at the early stages of product development. An alternative
strategy to reduce the impact of organic solvents on the environment is to conduct
organic reactions under solvent-free conditions. Among the benefits of solvent-
free processes are cost-effectiveness, reduced reaction times, decreased energy
consumption, enhanced volume/product ratio, and a significant reduction in reac-
tor size [6, 7]. The design of solvent-free catalytic reactions has gained tremendous
attention in the area of green synthetic chemistry in recent years. In this context,
various approaches for solvent-free organic reactions have been successfully
developed [8, 9].
Solvent-free or solid-state organic reactions usually are executed using the reac-
tants solely or by combining them in zeolites, clays, alumina, silica, or other cata-
lysts to decrease the by-products while achieving a high degree of stereoselectivity
and/or higher reaction rate [10, 11]. Typically, reactants are intensely mixed and then
carefully heated until reaction completion. These reactions follow the basic princi-
ples of green chemistry as well as being cheaper to operate [12]. More importantly,
solvent-free reactions are easy to handle and work up; this would be significantly
important in industry for mass production. Besides, the solid-state reaction gives a
unique opportunity for the product molecules to form crystals with high orientation
and a remarkable degree of stereoselectivity. The potential economic incentive and
the environmental benefits have created significant demand for solvent-free organic
reactions in the area of green synthesis.
This chapter overviews the recent developments in the field of carbon-based
materials as recyclable catalysts/supports in solvent-free organic reactions and their
outstanding performances in the controlled oxidation of alcohols and hydrocar-
bons in fine-chemical production with a high activity using atmospheric oxygen.
The authors wish to provide a fundamental background on the utilizations of
carbon-based materials as catalysts/support materials that are later described
based on their types of functionalities and porosity. These carbon materials include
activated carbons (ACs), carbon-based solid acids (CBSA), porous carbons, carbon
nanotubes (CNTs), graphene (including graphene oxide [GO], reduced graphene
oxide [rGO]), and the merits of atom-level control over composition, as well as
well-defined pore architectures.

6.2 Solvent-Free Catalysis Using Carbon-Based


Materials
Since most catalysts are noxious compounds and are associated with heavy-metal
waste, there is still a tremendous demand for environmentally friendly catalysts
under mild conditions in which solvent use is reduced, and work-up procedures
are simplified [13]. Carbon materials as a catalyst have been developed since over
6.2 Solvent-Free Catalysis Using Carbon-Based Materials 183

100 years ago, but it has not been mainstream materials due to the low activity.
Hitherto, a growing interest in the production of catalysts from sustainable mate-
rials, such as carbon-based nanomaterials with large surface area, could replace the
transition metal-based catalytic systems that make these materials promising as het-
erogeneous catalysts. Besides their low environmental impact and cost-effectiveness
[14, 15], and easy recovery and separation from the reaction products via simple fil-
tration, carbon-based nanomaterials as recyclable catalysts/supports possess many
extra advantages such as high surface area, controllable pore volume. Further, their
chemical surface nature is tunable by a chemical treatment that can be altered for
further modifying their catalytic activity [16].

6.2.1 Activated Carbons (ACs)


ACs, with their unique surface area (ca. 950–2000 m2 g−1 ), not only can be used
as catalyst supports but also as catalysts themselves in fine-chemical syntheses
[17, 18], and suitable alternative to the liquid phase systems [19, 20], owing to
their tunable acidic and basic characteristics. ACs have successfully catalyzed
various types of organic transformations such as synthesis of chalcones 6.2.1.1 [20],
synthesis of α,β-unsaturated nitriles 6.2.1.2 and 6.2.1.3 [21], and N-alkylation of
imidazole 6.2.1.4 [22, 23] under solventless conditions (Figure 6.1).

6.2.1.1 Acetylation Reactions


The formation of esters via acetylation (protection of hydroxyl group) of alcohols,
phenols, and carbohydrates is an essential reaction in organic and pharmaceutical

Figure 6.1 Schematic illustration of alkaline ACs catalyzed various types of organic
transformations under-solvent free conditions.
184 6 Carbon-Based Materials as Catalysts/Supports in Solvent-Free Organic Reactions

synthesis, cosmetics, and the food industry. Typically, acetylation reactions have
been achieved by an excess of acid anhydride or acid halide in the presence of tertiary
amine catalysts, including triethylamine and pyridine [24, 25]. Furthermore, pyri-
dine derivative, 4-dialkylaminopyridine [26], amine bases [27], metal triflates [28],
metal perchlorates [29], salen cobalt complex [30], iodine [31], bicarbonates and
carbonates [32], modified zirconia [33], tributylphosphine [34], p-toluenesulfonic
acid [35], Lewis acids (ZnCl2 [36] and cobalt chloride [37]), and scandium tri-
fluoromethanesulfonate [38], as well as task-specific ionic liquids [39] have been
successfully investigated as catalysts for acetylation of various kinds of hydroxylated
substrates. However, most of these acetylation approaches suffer from several limi-
tations, such as long-time and high-energy consumption, harsh reaction conditions,
and complicated work-up procedures. For example, pyridine and triethylamine
have unpleasant odors and difficult exclusions. 4-dialkylaminopyridine has compli-
cated, and costly preparation procedures, messy work-up, and tributylphosphine is
expensive, irritant, and highly flammable.
Environmentally friendly AC catalysts have very good catalytic activities for the
acetylation of aliphatic alcohols, benzylic alcohols, and aromatic alcohols such as
phenols, β-naphthol, and monosaccharides via clean, cost-effective, and efficient
solvent-free protocols [40]. The catalytic activity of several kinds of AC materials
has been investigated in the solvent-free acetylation reaction of alcohols, phenols,
and monosaccharides with acetic anhydride (AA) at 60 ∘ C. Three commercial basic
ACs, including carbon M, carbon gel X, and carbon N, in addition to three acidic ACs
(M-S, X-S, and N-S) that have been obtained via the sulfuric acid treatment of the
commercial carbons, have been studied for their catalytic activities. A distinguished
catalytic activity has been observed for the highest acidic X-S AC of large meso-
pore volume among all tested ACs. Primary, secondary, allylic, benzylic, and glycol
alcohols are acetylated in three hours or less in the optimum conditions. In con-
trast, complete acetylation of β-naphthol and phenol takes place in 8 and 24 hours,
respectively. Further, in the case of per-o-acetylation of monosaccharides, D-fructose
is hyperacetylated within one hour. In contrast, acetylation of D-xylose and D-glucose
yields pentaacetates in seven and eight hours, respectively. It is also noteworthy
that the acetylation of benzyl alcohol cannot be completed in the absence of an X-S
catalyst even after 24 hours under identical conditions, confirming the higher per-
formance of the X-S catalyst.

6.2.1.2 Oxidation of Cyclohexane


Systematic oxidation of cyclohexane into a mixture of cyclohexanol and cyclohex-
anone (so-called ketone- alcohol [KA] oil) is a pivotal process that is commonly
applied as solvent lacquers in the synthesis of polymers (e.g. Nylon-6,6 and Nylon-6),
varnishes, as well as shellacs, and as stabilizer and homogenizer in detergent and
soap industries [41]. In the early stage, Co(II)-based homogeneous catalyst has been
applied for the production of KA oil from cyclohexane [42, 43]. However, this pro-
cess shows various limitations, such as the difficult recovery of the catalyst due to
its easiness to be charged with salts, low conversion yield, and a huge amount of
alkali (110 kg NaOH are normally wasted for every single produced kilogram KA oil).
6.2 Solvent-Free Catalysis Using Carbon-Based Materials 185

Figure 6.2 (a) Cyclohexane to KA oil conversion in the presence of AC material,


(b) Proposed mechanism for the oxidation of cyclohexane with quinonic group grafted-AC.
Source: Sadiq et al. [46]/Hindawi/CC BY.

A wide variety of catalysts have been successfully applied [44, 45]; however, these
catalytic systems have shown a loss of sensitivity, poor catalytic activity, toxicity,
inconvenience for mass production, low selectivity, and high cost.
Recently, Sadiq and coworkers have reported AC materials via a two-step
chemothermal reaction involving the carbonization of the peanut shell in an inert
atmosphere at 380 ∘ C followed by activation with H2 O2 as an efficient catalyst
for solvent-free synthesis oxidation of cyclohexane 6.2.1.2.1 into KA oil 6.2.1.2.2
and 6.2.1.2.3 in the presence of molecular oxygen at a moderate temperature
(Figure 6.2) [46]. In addition to the ease of synthesis and recovery, nontoxicity,
and cost-effectiveness, the used AC catalyst has good recyclability for several
consecutive runs without a significant decay in activity. Thus, AC is a durable
heterogenous catalyst with a long lifespan for the solvent-free oxidative transfor-
mation of cyclohexane to KA oil with quinonic group grafted-AC, and its plausible
mechanism can be investigated according to Figure 6.2.

6.2.2 Carbon-Based Solid Acid (CBSA) Catalysts


6.2.2.1 Cross-Aldol Condensation of Ketones with Aromatic Aldehydes
Synthesis of CBSA catalyst can be achieved via several approaches such as through
a heating treatment of aromatic naphthalene in sulfuric acid under two reaction
steps: (i) the aromatic compound is sulfonated at 473–573 K; (ii) the obtained
186 6 Carbon-Based Materials as Catalysts/Supports in Solvent-Free Organic Reactions

sulfonated compound is then semi-carbonized to give rise to the desired catalyst


[47, 48]. Since the main approach for the synthesis of CBSA is sulfonation, a
variety of precursors have already been developed for the sulfonated carbons,
depending on the type of need for a catalyst [49]. The CBSA catalysts have attracted
tremendous attention in the area of solid-acid catalysts for catalyzing various
organic reactions due to their simple preparation procedure, recyclability without
any change in their activity or yield, high acid–base stability, sustainable catalytic
activity, and corrosion resistance [50]. In addition, CBSAs have been employed
for many critical reactions such as cross-aldol condensation of cycloalkanones
with aldehydes and ketones, multi-component condensation reaction, Hantzsch
condensation reaction, Biginelli reaction, nitration of aromatic compounds, and
reductive amination of aldehydes and ketones [50]. Among them, cross-aldol
condensation of cycloalkanones with aldehydes/ketones is a crucial reaction for
the synthesis of α,α′ -bis-substituted benzylidene cycloalkanones, a feedstock for
the synthesis of bioactive pyrimidine derivatives [50]. Various catalytic systems
including ZrCl4 [51], bis(p-methoxyphenyl)telluroxide (BMPTO) [52], KF–Al2 O3
[53], microwave-irradiated alumina [54], anhydrous RuCl3 , TiCl3 (SO3 CF3 ), and
perfluorosulfonated ionomers (e.g. nafion) [55, 56] have been also used as solid acid
catalysts for the cross-aldol condensation reaction of cycloalkanones. However,
those compounds have significantly lower acid activity, high toxicity, and high
cost of preparation compared to sulfonated carbons [57]. In addition, several aldol
condensation reactions catalyzed by such catalysts may involve a reverse reaction
or self-condensation of the starting materials [58, 59].
Utilizing CBSAs as catalysts in the cross-aldol condensation reaction of cycloalka-
nones is an effective pathway to overcoming the above-mentioned catalytic flaws
since CBSAs have promising catalytic features that have been mentioned in the pre-
vious paragraph [58]. A CBSA catalyst with dense SO3 H groups (chemical composi-
tion of CH0.4 O0.32 S0.13 ) has been successfully used in the cross-aldol condensation of
cyclohexanone derivatives with aromatic aldehydes under solvent-free conditions at
70 ∘ C [60]. This reaction resulted in products with a high yield in the range of 85–95%
within 65–120 minutes with no self-condensation of the ketones. The condensation
reactions have not proceeded in the absence of the catalyst, confirming the practical
role of the catalyst. Moreover, increasing the amount of CBSA catalyst (from 0.25 to
0.75 g mmol−1 of cyclohexanone) has drastically shortened the reaction time with an
excellent reusability without any loss in the reaction yield.
Multi-component reactions are the further condensation process that is catalyzed
by CBSA catalysts and can deliver a fast, reliable, and efficient synthesis of complex
organic molecules without isolation of any intermediate [61]. It is a recently emerged
and efficient tool in modern organic synthesis that typically involves three or more
reactants to form one ideal product that possesses the essential motifs of all the start-
ing substances. In addition, multi-component reactions have a major contribution to
the environment since they reduce energy consumption, waste production, and the
number of synthetic steps. The produced compounds from the multi-component
condensation reaction, such as 1,4-dihydropyridines (1,4-DHPs), are of major
importance. DHPs are common elements of various bioactive compounds as
6.2 Solvent-Free Catalysis Using Carbon-Based Materials 187

bronchodilator, vasodilator, anticancer, anti-atherosclerotic, and antidiabetic


agents [62]. Also, many derivatives of DHP are considered prospective drug candi-
dates for curing congestive heart failure. In addition, 1,8-dioxodecahydroacridines
and their derivatives are important polyfunctionalized derivatives of 1,4-DHP that
have gained less attention than other derivatives.
A typical synthesis method of DHP compounds is the reaction of aldoximes with
dimedone under microwave irradiation [63]. However, such procedures suffer from
long reaction times, low yields, harsh reaction conditions, and the use of toxic
organic solvents. Therefore, applying a multicomponent condensation of dimedone
6.2.2.1.1, aromatic aldehydes, and ammonium acetate or amines in the presence
of a suitable catalyst can produce DHP 6.2.2.1.2 compounds in high yields and
eco-friendly procedures depending on the type of catalyst. Among various catalysts
[64–66], CBSA catalyst, synthesized via heat treatment of naphthalene in the
presence of concentrated H2 SO4 at 250 ∘ C under a flow of N2 , has been efficiently
used through a one-pot synthesis of broad derivatives of DHPs 6.2.2.1.2 through
three-component condensation of dimedone 6.2.2.1.1, aromatic aldehydes, and
ammonium acetate or aromatic amines at 100 ∘ C in an oil bath for 15–40 minutes,
as represented in Figure 6.3a [67]. The catalyst mainly affects the efficiency of
the reaction under solvent-free conditions where no products are obtained in the
absence of the catalyst, and the yield linearly increases with increasing the catalyst
amount. More interestingly, the catalyst is recoverable and reusable several times
without significant loss in its activity. Unlike conventional processes, hot ethanol
extraction is used to restore the catalyst’s activity. Compared to other reported
approaches, the eco-friendly synthesis process of DHP using a three-component
condensation process in the presence of the green and heterogeneous CBSA catalyst
shows shorter reaction times, higher yields (over 90% obtained after 20 minutes),
and a simpler procedure [67].

(a)

(b)

Figure 6.3 (a) Synthesis of 1,8-dioxodecahydroacridines catalyzed by CBSA catalyst.


(b) CBSA-catalyzed-based synthesis of polyhydroquinoline derivatives.
188 6 Carbon-Based Materials as Catalysts/Supports in Solvent-Free Organic Reactions

Recently, polyhydroquinoline compounds have attracted an increasing area of


interest due to their outstanding pharmacological and therapeutic figure-of-merits
such as antitumor, hepatoprotective, vasodilator, bronchodilator, geroprotective,
and anti-atherosclerotic activity [68]. These compounds are typically synthesized
via cyclocondensation of β-ketoester, aldehyde, and ammonia substances, through
the Hantszch condensation method [69]. However, this approach suffers from
many drawbacks such as unsatisfactory product yield, long reaction time, harsh
refluxing conditions, and high consumption of toxic solvents. To overcome these
major shortcomings, the application of an efficient, CBSA-catalyzed Hantszch
reaction as a green heterogeneous catalyst has been used for the synthesis of
polyhydroquinoline 6.2.2.1.3 compounds under solventless conditions providing
high-product yields and exhibiting short-reaction times and easy work-up pro-
cedures, as well as high recyclability and catalytic efficiency of the catalyst [70].
The use of the CBSA catalyst under solvent-free conditions renders the Hantszch
reaction superior for the synthesis of polyhydroquinoline 6.2.2.1.3 compounds. A
variety of polyhydroquinoline 6.2.2.1.3 derivatives have been obtained in high yields
(up to 95%) and short times (18–35 minutes) using the Hantszch condensation
reaction in the presence of CBSA catalyst under solvent-free conditions at 90 ∘ C, as
represented in Figure 6.3b [70].

6.2.2.2 Substituted Imidazoles


The CBSA catalyst is highly efficient and eco-friendly for the synthesis of 1,2,4,5-
tetrasubstituted imidazoles via a one-pot four-component condensation reaction
of benzil, primary amines, aromatic aldehydes, and ammonium acetate under
solvent-free conditions. These substituted imidazole compounds have remarkable
pharmaceutical and therapeutic activities, include herbicides, fungicides, and
plant growth regulators [71]. Most of the reported protocols for the synthesis of
tetrasubstituted imidazole derivatives such as multi-component condensation
of 1,2-diketones, aromatic aldehydes, ammonium acetate, and primary amines
[72, 73], condensation of 1,2-diketones with primary amines and aryl nitriles
under microwave irradiation [74], hetero-cope rearrangement [75], and reaction of
N-(2-oxo) amides with ammonium trifluroracetate [76] are suffering from several
disadvantages such as poor product yields, harsh reaction conditions, long reaction
time, and utilization of toxic and expensive acid catalysts. Application of CBSA
catalyst provides excellent yields (86–93%), mild reaction conditions, short reaction
times (1–2 hours), easy work-up procedures in addition to excluding the volatile
toxic organic solvents during the synthesis of 1,2,4,5-tetrasubstituted imidazoles via
one-pot four-component condensation reaction at 130 ∘ C [77].

6.2.2.3 Amidoalkyl Naphthols


A CBSA-based catalyst also has a high catalytic performance toward the synthesis
of amidoalkyl naphthols via a one-pot three-component reaction of β-naphthol
6.2.2.3.1, aryl aldehydes 6.2.2.3.2, and acetamide 6.2.2.3.3 under solvent-free con-
ditions, as shown in Figure 6.4 [78]. Amidoalkyl naphthols are commonly prepared
through the three-component condensation reaction of amides, aldehydes, and
6.2 Solvent-Free Catalysis Using Carbon-Based Materials 189

Figure 6.4 CBSA-catalyzed synthesis of amidoalkyl naphthols. Source: Davoodnia et al.


[78]/with permission from Elsevier.

β-naphthols in the presence of numerous catalysts [79]. The majority of these


synthesis protocols suffer from various disadvantages such as utilizing a toxic and
expensive catalyst, the use of hazardous volatile solvents, long reaction time, and
unsatisfactory product yields. The new synthetic protocol, catalyzed by CBSA, is
beneficial to synthesize a wide range of amidoalkyl naphthol 6.2.2.3.4 products via
a one-pot three-component reaction under solvent-free conditions at 130 ∘ C [78].
Utilization of CBSA catalyst provided high-selectivity, high atom economy, and
clean-reaction profiles with high-yield and short-reaction time. The catalyst can be
easily recovered/reused several times without a considerable loss in the activity,
which makes this synthesis process more cost-effective, efficient, safer, and cleaner.
In addition, the organic reactions are carried out very efficiently without produc-
ing any side products. Furthermore, aromatic aldehydes with electron-withdrawing
groups have produced higher product yields and reacted faster as compared to those
containing electron-withdrawing groups [78].
The CBSA-based material acts as an outstanding and highly efficient solid catalyst
for nitration of aromatic rings in combination with sodium nitrate under solvent-free
conditions at room temperature (Figure 6.5a) [80]. Nitration of aromatic compounds
is a pivotal industrial approach that is principally responsible for the preparation
of nitroaromatic compounds that are indispensable in the synthesis of enormous
industrial compounds, such as explosives, pharmaceuticals, dyes, and plastics [81].
Constantly, the commercial nitration approach is a noxious process, where it usually
190 6 Carbon-Based Materials as Catalysts/Supports in Solvent-Free Organic Reactions

(a)

(b)

(c)

Figure 6.5 Schematic illustration for (a) the nitration of aromatic rings, (b) the in situ
generation of nitronium ion, and (c) the production of diphenyl disulfide from thiophenol
catalyzed by CBSA catalyst.

involves the application of excessive amounts of concentrated nitric and sulfuric


acid, generating enormous amounts of waste acids and NOx fumes [82]. Moreover,
the separation of product from acid via the commercial processes is a quite difficult
and energy-consuming step that requires basic work-up procedures, which in turn
leads to excessive basic waste streams and enhances the costs. The disadvantages of
the commercial processes have motivated scholars to develop solid-acid-catalyzed
nitration reactions for the synthesis of nitro-aromatic compounds. A wide range of
nitro-aromatic compounds have been successfully synthesized using CBSA/NaNO3
system with high yields up to 98% and short times (2–60 minutes) [80]. Accordingly,
the nitration reactions have been efficiently carried out at room temperature, and
the proposed mechanism for the in situ generation of nitronium ion and the pro-
duction of diphenyl disulfide from thiophenol is represented in Figure 6.5b,c [80].
Such a nitration process, catalyzed by CBSA, possesses many advantages over the
commercial processes such as sustainability, high-catalytic efficiency, ease of separa-
tion process, cost-effective and simple synthesis of the reagents, as well as excellent
yields of the products, which might be attractive for the large-scale production of
nitro-aromatic compounds.

6.2.2.4 Reductive Amination of Aldehydes and Ketones


Reductive amination of aldehydes and ketones by sodium borohydride has been suc-
cessfully achieved by utilizing CBSA as a catalyst. Commonly, sodium borohydride
has been employed as a reducing agent with different Brønsted acid catalysts for
the direct and indirect reductive amination of aldehydes and ketones into alkylated
amines [83, 84]. However, most of the used Brønsted acids are highly toxic, corro-
sive, and difficult to recover from reaction products. The CBSA catalyst has achieved
6.2 Solvent-Free Catalysis Using Carbon-Based Materials 191

(a)

(b)

(c)

Figure 6.6 (a) and (b) represent the direct and indirect reductive amination of ketones,
respectively. (c) The direct solvent-free reductive amination reaction of benzaldehyde with
aniline in the presence of CBSA and NaBH4 .

an outstanding catalytic activity over toxic Brønsted acids for the direct and indirect
reductive amination of aldehydes or ketones by NaBH4 under solvent-free conditions
at room temperature [85].
Under the optimized conditions, direct reductive amination (DAR) of aromatic
and aliphatic ketones, and aldehydes through amines by sodium borohydride
(NaBH4 ) and 70% wet CBSA catalyst has been carried out, and these reactions have
been found to be highly efficient in producing high-to-excellent product yields
(Figures 6.6a) [86]. The chemo-selectivity of DAR reactions has been studied in
the optimal conditions on functional ketones 6.2.2.4.1 and aldehydes containing
other reducible groups (i.e. cyano, ester, carboxylic acid, nitro, and amide groups).
These reactions have given rise to the corresponding alkylated amines 6.2.2.4.2 at
good-to-excellent yields without reduction of the reducible functional group [86].
However, DAR reactions suffer from producing a small amount of the corre-
sponding alcohol as a side product due to the reduction of the carbonyl group of
aldehyde or ketone. To overcome this flaw, it is proposed to conduct the indirect
reductive amination reactions with NaBH4 in the presence of 70% w/w wet CBSA
catalyst, which forms sequential reactions including the imination and further
reduction of the imine 6.2.2.4.3, resulting in excellent yields of amines 6.2.2.4.2 in
few minutes under solvent-free conditions at room temperature (Figure 6.6b) [86].
For example, DAR of benzaldehyde and aniline with one molar equivalent of
NaBH4 in the presence of dry CBSA under solvent-free condition can be taken
192 6 Carbon-Based Materials as Catalysts/Supports in Solvent-Free Organic Reactions

place with 90% conversion after 25 minutes at room temperature. However, when
the reaction is carried out in the presence of wet CBSA (prepared by simply
mixing a small amount of water with CBSA), the rate of reduction can be greatly
accelerated and the reaction is completed in eight minutes (Figure 6.6c). When
benzaldehdye 6.2.2.4.4 is thoroughly mixed with aniline 6.2.2.4.5 for 15 minutes
under solvent-free conditions at room temperature and the resulting mixture
has been successively stirred with a mixture of NaBH4 and CBSA, the reaction
gives N-benzylaniline 6.2.2.4.6 (94%) and benzyl alcohol 6.2.2.4.7 (6%). At the
early stage of mixing, the imine product 6.2.2.4.8 with 95% yield is produced,
which is rapidly reduced with NaBH4 in the presence of CBSA to N-benzylaniline
6.2.2.4.6.

6.2.2.5 Xanthenes and Dibenzoxanthenes


Synthesis of xanthenes and dibenzoxanthenes that cover various industrial products
and intermediates such as visualization materials, dyes, agricultural bactericide,
photodynamic therapy, antiviral and anti-inflammatory drugs have been widely
catalyzed by CBSA [87]. Such a method can be considered as a solution for
many issues that appear in other synthetic approaches of xanthenes [88]. This is
because the other methods suffer from many limitations, including significant
production of wastes, long reaction times (up to five days), low yields, utilization
of toxic and expensive reagents/solvents, high corrosiveness, nonrecoverable, and
difficult-to-dispose acid catalysts (H2 SO4 , CH3 COOH, H3 PO4 , and HCl). The utiliza-
tion of green and promising alternate homogenous CBSA-based catalyst has been
used for the condensation of β-naphthol 6.2.2.3.1 with aromatic aldehydes toward
the synthesis of Aryl-14H-dibenzo[a,j]xanthenes 6.2.2.5.2 under solvent-free
conditions at mild reaction conditions and easy work-up procedures, as shown in
Figure 6.7 [89].
Various substituted xanthene derivatives have been obtained via the condensation
of different aromatic aldehydes with β-naphthol 6.2.2.3.1 under solventless condi-
tions at 85 ∘ C. In the absence of the CBSA catalyst, about 10% of xanthene derivatives
have been produced after a prolonged reaction time. After the addition of the CBSA
catalyst, the reaction yield increased up to 94% within 30–45 minutes. Thus, utiliza-
tion of non-corrosive and green CBSA for the synthesis of xanthene derivatives via
a solventless condensation process has led to large-scale production of xanthenes
6.2.2.5.2 through a facile, cost-effectiveness, and environmentally-benign procedure
at mild-reaction conditions (Figure 6.7) [89].

6.2.2.6 Dihydropyrimidinone Compounds (Biginelli Reaction)


CBSA catalysts can also be synthesized through direct carbonization [90] of coal
pretreated with concentrated H2 SO4 under a flow of N2 gas [91]. The so-called
“coaled carbon-based solid acid (CCBSA)” catalyst have been successfully applied
for one-pot synthesis of 3,4-dihydropyrimidin-2(1H)-ones via the Biginelli reaction
of β-ketoesters 6.2.2.6.1, aldehydes 6.2.2.6.2 with urea/thiourea 6.2.2.6.3 under
solvent-free conditions, as those compounds have attracted considerable atten-
tion in the pharmaceutical industry because of their outstanding biological and
6.2 Solvent-Free Catalysis Using Carbon-Based Materials 193

Figure 6.7 Preparation of 14-aryl-14-H-dibenzo[a,j]xanthenes and suggested


transformation mechanism.

pharmaceutical activities. However, the main process used for the synthesis of
those compounds using acid-catalyzed Biginelli reaction may deal with several
limitations [92]. The introduction of CCBSA catalyst in Biginelli reaction gives
rise to synthesize a wide range of 3,4-dihydropyrimidinone 6.2.2.6.4 compounds
at high yield (up to 92.8%) in short reaction times under solventless conditions
(Figure 6.8) [91]. In addition, the products could be easily separated from the
reaction mixture, and the CCBSA catalyst can be successfully recovered and reused
for four consecutive runs without any decay in its activity, which makes Biginelli
reaction beneficial and efficient, especially with simple and clean procedures, and
the easy work-up under solventless conditions [91].

Figure 6.8 Synthesis of dihydropyrimidin-2(1H)-ones using CCBSA.


194 6 Carbon-Based Materials as Catalysts/Supports in Solvent-Free Organic Reactions

6.2.2.7 Acylation, Acetalization, Thioacetalization of Aldehydes


The green Pistachio peel (Pis-SO3 H) can be a source for the synthesis of a highly
sulfonated CBSA catalyst through simultaneous sulfonation, dehydration, and
carbonization of green Pistachio peel as biomass. Very recently, a novel CBSA
derived from green Pistachio peel has been applied as an efficient catalyst for
the chemoselective acetalization, thioacetalization, acylation of aldehydes, and
synthesis of biscoumarins under solvent-free conditions [93]. Acetals are considered
as protecting groups for the carbonyl compounds and act as intermediates and
products in the pharmaceutical, phytopharmaceutical, fragrance, and lacquer
industries. The acetalization, thioacetalization, and acylation of carbonyl group
have been achieved using a variety of catalysts [94–96]. However, the development
of a novel catalyst for such kind of solvent-free reactions is a vital endeavor. Due
to the easy preparation in large-scale from waste biomass, the green Pis-SO3 H
catalyst is also interesting [93]. This excellent performance of Pis-SO3 H catalyst has
resulted in the acetalization/acylation of benzaldehyde 6.2.2.7.1 or acetophenone
6.2.2.7.2 to 1,3-dioxolane 6.2.2.7.3, and 1,3-dithiolanes 6.2.2.7.4 with a yield up to
100%, and in the synthesis of 1,1-diacetates 6.2.2.7.5 or 1,1-diesters 6.2.2.7.6 under
solvent-free conditions at room temperature (Figure 6.9) [93]. In addition, the

Figure 6.9 Chemoselectivity of green Pis-SO3 H catalyst in acetalization, acylation, and


thioacetalization of aldehydes under solvent-free reaction at room temperature.
6.2 Solvent-Free Catalysis Using Carbon-Based Materials 195

sulfonated CBSA catalyst has been applied as a heterogeneous acid catalyst in the
synthesis of 3,3′ -arylmethylene-bis(4-hydroxycoumarins) derivatives at 80 ∘ C with
a yield that reaches 95% [93].

6.2.3 Carbon Nanotubes (CNTs)


6.2.3.1 Esterification of Alcohols
Selective oxidation of alcohols is a challenging reaction that is applied for the
direct synthesis of esters which have been widely used in the field of cosmetics,
fine chemicals, and polymer industries [97, 98]. For such a reaction, different
noble-metal-based homogeneous catalysts have been applied [99–101]; however,
these catalysts industrially seem impractical due to the use of nonrecoverable
homogeneous catalysts, costly noble metals, and the overwhelming work-up pro-
cesses. Hence, replacing these catalysts with inexpensive, reusable, and non-noble
metal-based heterogeneous catalysts is an important step for the selective con-
version of alcohols to esters [97, 102]. The development of heteroatom-doped
carbon-based supports has been progressed impressively due to their semicon-
ductor nature and the ability to tune the electronic density by the amount of
heteroatom doping [103]. For instance, nitrogen-doped carbons with electron-rich
surfaces have recently acted as promoters that catalyze specific reactions and also
as active supporting materials [104, 105]. In addition, surface functional groups
can vary the CNT supports [106]. CNT supports have contributed as a catalyst
support material due to their thermal stability, mechanical properties, and electrical
conductivity [107, 108]. Since the activation of aliphatic C—H bonds in aromatic
hydrocarbons could be proceeded on nitrogen-doped carbon, the co-doped carbon
might be able to offer noticeable activity. A catalytic system of cobalt-enriched
N-, O-, and S-tridoped CNTs (Co@NOSC) has been used as an efficient multi-
functional sustainable catalyst for the base-free selective oxidation of alcohol to
esters [109]. In their report, a scalable and sustainable method has been reported
to synthesize Co@NOSC catalyst by direct pyrolysis of a mixture of cobalt salt,
urea, and a renewable raw biomaterial (i.e. carrageenan), as a source of carbon and
oxygen.
A selective one-step aerobic oxidative esterification of benzyl alcohol 6.2.3.1.1 to
methyl benzoate ester 6.2.3.1.3 has been achieved by the catalysis of Co@NOSC
under base-free conditions, giving a catalytic oxidation conversion of 97% at 60 ∘ C in
24 hours (Figure 6.10). The recyclability of the Co@NOSC catalyst has been studied
by its treatment with hydrogen at 400 ∘ C for one hour, achieving 85.5% conversion
and 96.5% selectivity after six recycle times [109]. Co@NOSC catalyst has also been
tested for aerobic esterification of various substituted benzyl alcohols, bibenzyl
alcohols, S-heterocyclic alcohols, N-heterocyclic alcohols, aliphatic alcohol, and
benzyl alcohol with long-chain aliphatic alcohols under optimum conditions [109].
Electron paramagnetic resonance (EPR) study of the oxidation of benzyl alcohol
6.2.3.1.1 in the presence of the Co@NOSC catalyst with spin-trapping experiments
using α-(4-pyridyl-1-oxide)-N-tert-butylnitrone (POBN) as the spin-trapping agent
further confirms the formation of • CH2 OH radicals and H• radicals, wherein the
196 6 Carbon-Based Materials as Catalysts/Supports in Solvent-Free Organic Reactions

Figure 6.10 A possible reaction mechanism for the oxidative esterification of benzyl
alcohols in the presence of the Co@NOSC catalyst. Source: Nandan et al. [109]/with
permission from Royal Society of Chemistry.

solvent plays an active role in a nonconventional manner. This measurement is


initially conducted by dissolving POBN in MeOH in the presence of the substrate
benzyl alcohol 6.2.3.1.1, followed by worming the solution at 60 ∘ C for 30 minutes.
The metallic cobalt core serves as a source of electrons, whereas the CNTs shell
provides active sites for efficiently binding of the substrate to the catalyst surface
and facilitate electron transfer reactions.
On the basis of the EPR results (Figure 6.10), a possible mechanism for the
oxidative esterification of benzyl alcohol 6.2.3.1.1 by the Co@NOSC catalyst has
been proposed and may proceed as follows: (i) at the early stage, molecular triplet
oxygen is presumed to form quickly and is converted into the superoxide anion
radical (O2 •− ) after accepting an electron from the metallic cobalt; (ii) then the
reactive O2 •− abstracts a hydrogen radical (• H) from the surface-adsorbed molecules
of methanol and transforms into a hydroperoxide radical (• HO2 ), whereas methanol
forms a radical intermediate (• CH2 OH); (iii) in the next step, POBN effectively
traps these • CH2 OH radicals, forming the POBN–CH2 OH radical adduct. The • HO2
radical further abstracts an • H radical from • CH2 OH to form hydrogen peroxide and
an aldehyde (CH2 =O); (iv) then, the formed H2 O2 is dissociated again into • OH
in the presence of the catalyst; (v) subsequently, these • OH radicals attack benzyl
alcohol 6.2.3.1.1 to form benzaldehyde 6.2.3.1.2; (vi) finally, benzaldehyde 6.2.3.1.2
reacts with the excess amount of methanol to form a reactive hemiketal intermedi-
ate, which eventually transforms into methyl benzoate 6.2.3.1.3 via abstraction of
6.2 Solvent-Free Catalysis Using Carbon-Based Materials 197

hydrogen by the • OH radical, because of the presence of more heteroatoms on the


surface of the Co@NOSC catalyst.
The presence of more hydrophilic heteroatoms on the surface of the catalyst may
increase the conversion and selectivity of –NO2 -substituted benzyl alcohol up to
90% comparing to previously reported noble-metal-based homogeneous catalysts
[110]. Lactonization of diols has also been performed via Co@NOSC catalyst. Such
a reaction is important due to the potential of lactones in polymer industries and
biological applications [111]. A good conversion (66–68%) has been observed during
lactonization of diols with a good selectivity (91–92%) under ambient O2 at a reac-
tion temperature of 100 ∘ C. The high base-free catalytic activity of the catalyst can
be attributed to the synergistic cooperation between the inner metallic core of the
cobalt nanoparticles (NPs) and the electron-rich outer part due to the N-, O-, and
S-doped CNTs. After several cycles, the catalyst still possesses very stable and robust
activity without any significant loss of its activity. This base-free approach coupled
with non-hazardous oxidants is very promising for selective oxidation of different
organic transformations in high yield.

6.2.3.2 Benzyl Alcohol Oxidation


Pd supported catalysts have superior catalytic performance in alcohol oxidation such
as hydroxyapatite (Pd/HAP) [112], Pd-supported zeolite [113], and Pd-supported
graphene [114, 115], where the catalytic activity of small Pd NPs is promoted in the
aerobic oxidation of benzyl alcohol [116]. The metal dispersions and metal-support
interactions of catalysts can be fine-tuned by controlling the surface chemistry
of the catalyst’s support. For instance, the basic functional group localized on
the surface of support material can enhance the activation of benzyl alcohol in
aerobic dehydrogenation, accelerate the desorption of intermediate products, and
decrease the extent of side reactions [117]. The aerobic oxidation of benzyl alcohol is
achieved by CNT-supported metals such as Pt alloyed with Fe catalyst [118] and Pd
NPs supported on MnO-doped CNT [119]. In an interesting study, the Pd NPs have
been supported on organosilane-functionalized CNT, where their catalytic activity
has been demonstrated by the conversion of benzyl alcohol 6.2.3.1.1 by aerobic
oxidation reaction under solvent-free conditions (Figure 6.11) [120]. The grafting
method of CNT to generate the functional groups onto the surface of CNT is the
utilization of organosilanes, i.e. (3-aminopropyl)triethoxysilane (APS), hexamethyl-
disilazane (HMDS), and [3-(2-aminoethylamino)propyl]trimethoxysilane (ATMS).
Pd NPs have been eventually supported on the functionalized CNT using the
adsorption-reduction method, and the catalyst has been obtained after carbother-
mal treatment at 400 ∘ C under a flow of H2 . The aerobic oxidation of benzyl alcohol
6.2.3.1.1 determines the catalytic activity of the prepared catalysts with ambient
oxygen (20 ml min−1 ) under stirring of 1200 rpm at 160 ∘ C. It has been observed that
the pH measurements are differentiated according to the surface-functionalized
CNT-supported catalyst. For instance, the pretreated CNT surface with nitric
acid is mildly acidic, where the pH of APS and ATMS functionalized CNT is
increased from 6.2 to 9.1 and 9.8, respectively, indicating that the basicity of
CNT surfaces can be enhanced with amino-containing functional groups. However,
198 6 Carbon-Based Materials as Catalysts/Supports in Solvent-Free Organic Reactions

Figure 6.11 The preparation process and catalytic evaluation of surface-functionalized


CNT-supported Pd catalysts. Source: Yan et al. [120]/with permission from Elsevier.

HMDS-functionalized CNT shows a neutral surface. The highest conversion (27.8%)


of benzyl alcohol 6.2.3.1.1 to benzaldehyde 6.2.3.1.2 and trace amounts of toluene
and benzoic acid with turnover frequency (TOF) equals 288 755 hours−1 has been
observed for Pd/2.4 APS-CNT catalyst [120].
This high catalytic activity can be attributed to the small mean size and the high
dispersion of Pd NPs after surface functionalization of CNT with APS organosilane
modifiers, where Pd/APS-CNT possesses the smallest Pd NPs size and the narrowest
size distribution. The APS functional groups on CNT increase CNT electron density
due to enhancing the electron density of Pd NPs. The lowest catalytic activity has
been observed for Pd/HMDS-CNT with only 12.8% conversion of benzyl alcohol and
TOF of 125 089 hours−1 . Furthermore, increasing the amount of APS functionalized
groups from 0.3 to 2.4 wt% linearly increases with the conversion of benzyl alco-
hol 6.2.3.1.1 up to 27.8%, while further increase of APS to 3.6 wt% leads to a huge
decrease in catalytic conversion (7.9%).
A wet impregnation method followed by a metal ion adsorption-reduction process
for the preparation of Pd NP catalysts supported on rare-earth oxide functionalized
6.2 Solvent-Free Catalysis Using Carbon-Based Materials 199

CNT surfaces Pd/LnOx -CNT (Ln = La, Ce, Sm, Gd, Er, Yb) has been proposed,
as well [121]. The rare-earth oxide functionalized CNT supported Pd catalysts show
higher catalytic activity in the aerobic solvent-free oxidation of benzyl alcohol
6.2.3.1.1 compared to the nonfunctionalized Pd/CNT catalysts due to the fine-tuned
properties of active catalytic sites such as the particle size of Pd, valence status,
size distribution, electron density, surface acidity and basicity, and metal-support
interactions. The functionalized catalyst by an appropriate amount of Sm2 O3
(1Pd/10Sm-CNT) achieves a significant high TOF of 318 760 hours−1 and selectivity
of 95.7%, whereas 1Pd/5Sm-CNT produces the highest conversion value of 29.6%
for aerobic oxidation of benzyl alcohol.

6.2.3.3 Phenol Derivatives Antioxidants


Phenol derivatives are widely used for the production of herbicides, pharmaceuti-
cals, dyes, agrochemicals, and antioxidants [122]. 6,6′ (Arylmethylene)bis(2-(tert-
butyl)4-methylphenol( are antioxidant compounds that prevent oxidative damage
caused by reactive oxygen species [123]. These compounds are interacting and
intercepting with reactive radicals by neutralizing the reactive oxygen species. The
preparation of antioxidants is conducted by using homogeneous and heterogeneous
catalysts [124]; however these catalysts have some limitations, such as poor yield,
harsh conditions, very long reaction time, and using an excessive amount of cata-
lyst, toxic organic solvents, and non-recyclable catalysts. Therefore, the synthesis
of bisphenolic antioxidants with direct, efficient, and environmentally friendly
methods becomes a major concern nowadays. One of the attractive approaches to
synthesize bisphenolic antioxidants is the multi-component reaction that allows
the direct formation of the desired products with an increased structural complexity
from a collection of simple and small starting materials [125]. Involving catalysts in
these reactions, particularly metals-supported catalysts, can promote the catalytic
activity and, hence, enhance the resulted product. The main essential features of
suitable support are surface area, mechanical properties, chemical composition,
and stability. In this regard, carbon can act as a promising support material due
to its valency, which allows the formation of many allotropes. A wide variety of
carbon nanostructures have been used as supports, such as graphene, diamond, and
fullerene. Some of these carbon catalysts have many disadvantages, including the
low specific surface area, instability at high temperatures, non-reusability in organic
reactions at high temperatures, low thermal and mechanical properties, as well as
the low active sites on the catalyst. In contrast, CNTs have attracted tremendous
attention for their superior properties, which have been widely applied in potential
applications, such as electronic devices, medicine [126], optical devices [127] and
catalytic applications [128]. CNTs are very effective supports as solid acid catalysts
due to their thermal and mechanical stability and their high specific surface
area. Synthesis of trisphenol and calix[4]arene derivatives have been successfully
achieved using sulfonated multi-walled carbon nanotubes (SO3 H@MWCNTs) as
a catalyst where different calix[4]arenes have been synthesized with a product
yield of 65–82%. The preparation method of (SO3 H@MWCNTs) is described in
Figure 6.12 [129].
200 6 Carbon-Based Materials as Catalysts/Supports in Solvent-Free Organic Reactions

(b)

(a)

(c)

Figure 6.12 Preparation of the heterogeneous catalytic system. Source: Fareghi-Alamdari


et al. [129]/with permission from Royal Society of Chemistry.

Moreover, MWCNTs-SO3 H has been used as a solid catalyst for the synthesis of
6,6′ -(arylmethylene)bis(2-(tert-butyl)4-methylphenol) antioxidants under solvent-
free conditions [130]. The MWCNTs-SO3 H catalyst is synthesized through two
steps as follows: (i) the conversion of MWCNTs to MWCNTs-COOH with HCl and
HNO3 , introducing a high content of oxygen-based functional groups on the surface
of MWCNTs [131, 132], (ii) the surface functionalization with sulfonated groups
through the reaction of MWCNTs-COOH with H2 SO4 to obtain MWCNTs-SO3 H.
Using the prepared catalyst, 6,6′ -(arylmethylene)bis(2-(tert-butyl)4-methylphenol)
antioxidants have been synthesized by the reaction of p-cresol, methyl tert-butyl
ether (MTBE), and 4-nitrobenzaldehyde under solvent-free conditions at 100 ∘ C in
four hours with a high conversion yield of 95%. By increasing the temperature from
70 to 110 ∘ C, the possibility for the formation of 2,6-di-tert-butyl-p-cresol (2,6-DTBC)
as a by-product is dramatically increased, and thus prevents the formation of the
required antioxidants. The heterogeneous MWCNTs-SO3 H catalysts have been
reused for eight consecutive runs without significant loss of the catalytic activity,
which clarifies their practical application over the homogeneous catalyst.

6.2.3.4 Acrylonitrile Derivatives


Very recently, 2-(1H-tetrazole-5-yl)acrylonitrile has been green-synthesized under
solvent-free conditions using a novel magnetic CNT-functionalized chlorosulfonic
acid decorated to the spherical NPs (Fe3 O4 -CNT-SO3 H) [133]. By applying, a
multi-component domino Knoevenagel condensation/1,3-dipolar cycloaddition
reaction between aromatic aldehydes, malononitrile, and sodium azide, the
2-(1H-tetrazole-5-yl)acrylonitrile can be successfully obtained under solvent-free
conditions. The Fe3 O4 -CNT-SO3 H nanotube catalyst achieves a conversion yield up
to 90% at 80 ∘ C after 2.5 hours and can be magnetically separated and recycled up to
five runs without an obvious loss in the catalytic efficacy.

6.2.4 Graphene Oxide (GO)


During the development processes of green heterogeneous carbon-based cata-
lysts, graphene, with its excellent optical, thermal, and mechanical properties,
6.2 Solvent-Free Catalysis Using Carbon-Based Materials 201

has been discovered and used as a promising catalyst for organic transformations
[134]. Graphene-based materials are potential candidates for many applications
due to the exceptional properties of graphene, such as thermal conductivity
(∼5000 W m−1 K−1 ), ultrahigh electron mobility (2 00 000 cm2 V−1 s−1 ), high planar
surface (2630 m2 g−1 ), outstanding electronic properties, and superlative mechani-
cal strength (Young’s modulus, ∼1100 GPa) [135, 136]. The high surface to volume
ratio and the single-atom thickness allow graphene to emerge as incomparable
2D supports and catalysts for different catalytic reactions [137]. Furthermore, the
catalytically active site can be highly loaded due to the ultrahigh thermal conduc-
tivity and the high chemical stability of graphene. Graphene can be activated either
by doping the substrate underneath epitaxial graphene [138] or by introducing
defects [139, 140]. GO is an oxidized form of pristine graphene that has various
oxygen-containing functional groups on its surfaces, where these groups offer supe-
rior hydrophilic character and analogous chemical reactivity [139]. The hydrophilic
GO can be further reduced to produce hydrophobic rGO of similar properties to
GO, including mechanical, conductive, and optoelectronic [141]. rGO is attractive
for a wide range of applications since it consists of a heterogeneous structure
involving graphene-like basal planes that have structural defects and filled with
areas containing oxidized chemical groups [141]. As a catalyst, graphene and its
derivatives can act as support materials for metal NPs such as Pt, Ag, and Au, which
significantly improve their catalytic activity. Here, the unsupported NPs have poor
catalytic performance due to the agglomeration of NPs, which causes a reduction
in the active surface area. Graphene as a catalyst support can allow stabilization of
catalyst particles, resulting in enhanced catalytic performance [13]. For instance,
supporting of CuO NPs on GO/rGO has been widely used for various organic
transformations, such as CO oxidation [142], synthesis of coumarin-based triazoles
[143], synthesis of imidazo[1,2-a]pyridines [144], nitroaromatics hydrogenations
[145], decomposition reaction of dye molecules [146], and the synthesis of ynones,
1,3-diynes and 1,5-benzodiazepines [147].

6.2.4.1 Alkylaminophenols Derivatives


A very simple, convenient, and green method for the synthesis of a CuO NPs/rGO
nanocomposite catalyst has been synthesized using GO support, Cu(OAc)2 mono-
hydrate, and hydrazine hydrate as a reducing agent, where the reduction of GO,
protection, and functionalization occurred in a one-synthetic step (Figure 6.13)
[148]. The catalyst has been then used to catalyze the selective synthesis of
alkylaminophenols 6.2.4.1.2 derivatives via the Petasis-Borono-Mannich (PBM)
reaction of boronic acids, salicylaldehydes 6.2.4.1.1, and amines under microwave
irradiation [148]. Compared to conventional methods, the catalytic activity of the
catalyst to catalyze the synthesis of alkylaminophenols 6.2.4.1.2 under microwave
irradiation has been enhanced by 12-fold. Using microwave irradiation is more
favorable for several advantages, including cost-effective, environmentally benign,
simple work-up procedures, short reaction times, and high yields. Furthermore,
the CuO NPs/rGO catalyst enhances the rates of several of these transformations
and exhibits high recyclability using a simple filtration process, implying its high
202 6 Carbon-Based Materials as Catalysts/Supports in Solvent-Free Organic Reactions

Figure 6.13 Proposed reaction mechanism for the CuO NPs/rGO composite catalyzed PBM
reaction. Source: Dandia et al. [148]. Copyright 2018, the Royal Society of Chemistry,
permitted under CC-BY License.

robustness and stability. It is proposed that the CuO NPs/rGO catalyst serves as
a Lewis acid for the reaction of salicylaldehydes 6.2.4.1.1 and amines to form the
corresponding imine. In the subsequent step, the catalyst facilitates the coordina-
tion between the oxygen anion of the imine and the boron atom of the boronic
acid (H3 BO3 ) and forms an intermediate of tetracoordinate borate. After that, the
aryl moiety of the H3 BO3 migrates from the boron to the iminium carbon and
forms the stable intermediate, which upon hydrolysis affords the corresponding
alkylaminophenol 6.2.4.1.2 in a high yield (93%) by the loss of a H3 BO3 molecule.

6.2.4.2 N-Arylation Reactions


The N-arylation reaction plays a significant role in organic synthesis since the
resultant products, arylamines, and N-aryl heterocyclic compounds are ubiquitous
compounds in pharmaceuticals, crop-protection chemicals, and material science
[149]. The copper-catalyzed N-arylation of amides, i.e. the Goldberg reaction, is an
efficient method for the construction of products relevant to both industry and
academic settings [150]. Core–shell Cu@Cu2 O NPs on rGO (Cu@Cu2 O NPs-rGO)
has been developed as a catalyst for the N-arylation reaction of N-heterocycles [151].
6.2 Solvent-Free Catalysis Using Carbon-Based Materials 203

The CuO/rGO nanocomposites catalyst has attracted significant interest in the


synthesis of hybrid molecules containing versatile heterocyclic moiety, flavanone
for numerous biological activities such as antitumor, anticancer, antioxidant, and
anti-inflammatory [152, 153].
Very recently, graphene has been applied for various organic reactions under
solvent-free conditions [154]. An eco-friendly low-cost aminically modified
GO-based zinc metal catalyst has been designed and proficiently catalyzed CO2 fix-
ation through N-formylation and carbamate formations reactions of both aromatic
and aliphatic amines with high yield in the presence of polymethylhydrosiloxane
(PMHS) as a reducing agent under mild temperature [154]. Under the same CO2
pressure, the catalyst achieves good product selectivity of carbamates synthesized
from aniline or its derivatives and alkyl/aryl bromide with solvent-free conditions
at room temperature. The catalyst is efficiently reusable for six consecutive cycles.

6.2.4.3 Oxidation of Benzylic Alcohols


The catalytic role of noble metals NPs in the oxidation of alcohols has a key role in
chemical industries and academic laboratories due to their impressive catalytic per-
formances compared to non-noble metals [155]. The high catalytic activity of metal
NPs catalysts can be attributed to the high potential of Pd species to be in narrow size
distribution and high dispersion throughout the graphene sheets. For the oxidation
of alcohol, the size of the metal NPs mainly controls the catalytic behavior [156].
For instance, the catalysis of benzyl alcohol by acidic and basic sites is suggested to
be conducted with a dehydrogenation process to form ketones or aldehydes [157].
This dehydrogenation of benzyl alcohol can be achieved by using Au NPs supported
on hydrotalcite, which affords the best catalytic activity due to the strong acidity and
basicity of the surface sites [158]. Hence, the surface property is important to con-
trol the reaction conversion and selectivity toward the required product. In addition,
the conversion of benzyl alcohol to benzaldehyde is higher with basic surface [159].
Therefore, the best catalytic activity can be indicated with highly dispersed metal
NPs that have narrow size distribution [158].
A facile method combines lyophilization, vacuum, and reduction treatments for
the synthesis of Pd/rGO aerogels with large surface areas and enriched oxygen-
containing functional groups [160]. The oxygen-containing functional groups
facilitate the active phases for tuning the interaction between palladium and
the rGO support, and thus, the Pd NPs can be homogeneously dispersed in
nanoscale on the Pd/rGO aerogels. Compared with Pd/AC as a reference, the
Pd/rGO aerogel catalyst with large surface areas contains a higher proportion
of fine-tuned Pd species than the Pd/AC and thus, exhibits improved catalytic
performance in the selective oxidation of benzyl alcohol as the main product
under solvent-free conditions at 140 ∘ C and 10 bar O2 , affording a conversion of
about 80% after three hours reaction time [160]. In contrast, the Pd/AC catalyst
showed a low conversion yield of approximately 40% under the same condition
with relatively poor selectivity. In other words, the Pd/rGO aerogels exhibit better
activity and selectivity compared to Pd/AC catalysts. In this method, toluene
and benzyl benzoate that are respectively derived from hydrogenolysis of benzyl
204 6 Carbon-Based Materials as Catalysts/Supports in Solvent-Free Organic Reactions

alcohol and esterification between benzyl alcohol and benzaldehyde are detected as
by-products.
There has been considerable interest in the utilization of Pd-anchored graphene as
an ideal catalyst for green oxidation of aromatic alcohols using molecular oxygen as
an oxidant without solvent or additives [161]. First, the GO is prepared by oxidization
of graphite, followed by ultrasonication to exfoliate the obtained graphite oxide into
GO which is then calcined in flowing He at 300 ∘ C for four hours. The calcined GO
has been used as carbon support for anchoring palladium species. For comparison,
CNTs and AC supports have also been used to anchor palladium species. The pre-
pared Pd anchored-GO, -CNT, and -AC have been examined for their catalytic activ-
ities under solvent-free aerobic oxidation of aromatic alcohols with ambient oxygen
at 110 ∘ C. As it is well known, oxygen-based functional groups on the surface of
GO are thermally more stable than those on CNT and AC, and thus explain the
high weight loss of Pd/CNT (7.4%) and Pd/AC (9.3%), compared to Pd/GO of lower
weight loss (2.5%) by means of thermogravimetric analysis (TGA) [161]. The Pd/GO
catalyst has shown higher conversion for benzyl alcohol that reached up to 90%
of benzaldehyde as the main product under solvent-free aerobic oxidation and has
increased with increasing the reaction time up to eight hours. However, some small
quantities of benzoic acid and benzyl benzoate have been detected as by-products.
Without Pd NPs, the oxidation of benzyl alcohol with GO, CNT, and AC supports is
insignificant (<1%). Even commercial Pd/AC with high palladium loading of 5.92%
has shown a lower activity than Pd/GO catalyst with very low palladium content
(0.68%). The outstanding catalytic activity of Pd/GO (TOF = 30 137 mol h−1 molPd )
in the aerobic oxidation of aromatic alcohols is attributed to the promoted adsorp-
tion sites of oxygen on Pd/GC and the spill-over of oxygen from palladium sites to
the adjacent bridge sites of graphene [161]. In addition, the adsorption of benzyl
alcohol on Pd/GO catalyst is promoted by π-electron interaction between benzene
π-conjugation system and graphene.

6.2.4.4 Aldol and Konevenagel Condensation Reaction


GO has been successfully applied as a catalyst for both aldol and Knoevenagel
condensation reactions [162]. It has shown a potential catalytic activity due to
the presence of various surface functional groups (i.e. −OH, −COOH, −CO−,
O=C−O−). In addition, graphene-based catalysts show good biocompatibility, high
surface area [163], and high adsorption capacity [164]. The GO-based organocatalyst
prepared via Hummer’s method has shown high catalytic activity in aldol and Kon-
evenagel condensation reaction under solvent-free conditions at room temperature,
which is a profitable process for green chemistry according [162]. The catalytic
activity of GO catalyst in aldol condensation of acetophenone and benzaldehyde
shows a high conversion yield up to 93% under solvent-free conditions at room
temperature in four hours, while the reaction does not proceed in the absence of
the catalyst. The catalytic activity of GO for Knoevenagel condensation reaction
of benzaldehyde with malononitrile to produce benzylidene malononitrile as the
main product is also significant. The catalyst shows a significantly high conversion
of 97% under the same conditions.
6.2 Solvent-Free Catalysis Using Carbon-Based Materials
. 205

6.2.4.5 Oxidation of Cyclohexene


Metal (Au, Pt, and Ag)-graphene complex nanostructures as catalysts have been
successfully used by Huang and coworkers for cyclohexene oxidation in ambient
air with high activities [165]. They have synthesized a core-shell structure in which
the metal NPs act as core and graphene as the shell. This core-shell structure is
considered challenging since the planar nature of graphene can be an obstacle.
The chemical oxidation method is followed to obtain graphene. Hollow graphene
nanospheres and a high density of metal NPs supported on graphene are synthe-
sized by adjusting the molar ratios of graphene to metal salts (Figure 6.14a–c, where
R = 1 mg(graphene) per 100 mM(metal salt) ) [165]. The catalytic conversion of cyclohexene
by Au-graphene core-shell NPs has been reported as 55.5% after 24 hours at 60 ∘ C
with ambient air as an oxidant (Figure 6.14a). However, the conversions of cyclohex-
ene 6.2.4.5.1 to 1,2-cyclohexanediol 6.2.4.5.2 with Pt-graphene and Ag-graphene
core-shell nanocatalysts are 46.10% and 38.05%, respectively, under the same
conditions.
On the other hand, a high density of Au NPs supported on graphene as a catalyst
has achieved 52.75% of cyclohexene 6.2.4.5.1 conversion, whereas hollow graphene
nanospheres have exhibited a catalytic conversion of only 15.4%. Based on these
results, the higher catalytic activity of Au-graphene core–shell NPs over the Au
NPs-supported on graphene is most likely due to the synergetic confinement effect
in the core–shell nanostructures.
Furthermore, functionalized rGO has been prepared by using p-arsanilic acid
and dithiooxamide [166]. Ruthenium (Ru) NPs of <6 nm have been immobilized
onto rGO by the hemi-micelle method to prepare a nanocatalyst (rGO-AO-TO/Ru).
In this process, the hydrophilic groups on the surface of graphene dispersed

Figure 6.14 Proposed synthesis scheme for different metal-graphene complex


nanocatalysts by adjusting the molar ratios of graphene to metal salts. (a) Core-shell
nanostructures, (b) hollow graphene nanospheres, and (c) a high density of metal NPs
supported on graphene (M and n represents metal ion and valence, respectively;
R = 1 mg(graphene) per 100 mM(metal salt) ). Source: Huang et al. [165]. Copyright 2012, the Royal
Society of Chemistry.
206 6 Carbon-Based Materials as Catalysts/Supports in Solvent-Free Organic Reactions

(a)

(b)

Figure 6.15 (a) A schematic illustration of the preparation of rGO-AO-TO/Ru nanocatalyst


and (b) its catalytic oxidation of cyclohexene with molecular oxygen under solvent-free
conditions.

in n-hexane act as reverse micelles, which is considered as a micro reactor to


immobilize Ru3+ ions (Figure 6.15a) [166]. The synthesized nanocatalyst exhibits
high catalytic oxidation of cyclohexene 6.2.4.5.1 to cyclohexenol 6.2.4.5.2 and
cyclohexenone 6.2.4.5.3 with a conversion yield up to 91% and selectivity (90%)
under solvent-free conditions with molecular oxygen (Figure 6.15b). During
oxidation process, molecular oxygen is converted to adsorbed very high energy O2 •−
because of the electron transfer from the Ru sites on the surface of rGO-AO-TO/Ru
6.2 Solvent-Free Catalysis Using Carbon-Based Materials 207

nanocatalyst to oxygen molecules. The oxidation reaction of cyclohexene 6.2.4.5.1 is


dependent on the reaction temperature (optimum temperature is 80 ∘ C), therefore
cyclohexene 6.2.4.5.1 converts into the gas phase in the higher temperatures (boiling
point is 83 ∘ C). Subsequently, in the gas phase, the effective collision of cyclohexene
6.2.4.5.1 is decreased with the active sites of the catalyst, thus the conversion rate of
the reaction is diminished, and the required activation energy to run the oxidation
reaction cannot be supplied at the low temperatures. Furthermore, increasing the
pressure higher than 6 bar has a reverse effect on the oxidation of cyclohexene
6.2.4.5.1, which may be ascribed to the following reasons: (i) at high pressure, the
surface of the catalyst is totally saturated by oxygen and there is no empty space
support to be entered; and (ii) Ru NPs may be oxidized and their catalytic activities
are dramatically dropped. The recyclability of Ru-based catalyst is high even
after five consecutive runs without loss of the reaction efficiency under optimum
conditions since the catalyst can be easily separated from the reaction mixture by
centrifugation.

6.2.4.6 Oxidation of Hydrazide and Pyrazole Derivatives


Metal-free carbon-based acid catalysts have gained much interest to support many
synthetically useful organic transformations. Phenyl sulfonated rGO (rGOPhSO3 H)
exhibits an encouraging efficiency, cost-effectiveness, and environmentally friendly
feature. The solid acid rGOPhSO3 H nanocatalyst has been prepared with the
introduction of aromatic sulfonic acid radicals onto the surface of rGO by frac-
tional removal of oxygenated functions under ultrasound irradiation (Figure 6.16)
[167]. This catalyst has been used to catalyze the cyclization involving oxidation
of various hydrazides 6.2.4.6.1 for direct synthesis of 5-substituted-3H-[1,3,4]-
oxadiazol-2-ones 6.2.4.6.2 in outstanding yields (86–99%) from hydrazides with
carbon dioxide (1.0 MPa) under solvent-free sonochemical conditions (Figure 6.17)
[167]. The rGOPhSO3 H is effective after recycling experiments since the catalyst
can be separated by very simple filtration procedures.
Developing water-tolerant and dispersible catalysts has gained significant interest
in improving several catalytic processes for the synthesis of organic compounds in
water. A new type of highly dispersible phosphoric acid-functionalized GO-based
nanocatalyst with hydrophilic acidic and basic sites as well as hydrophobic

(a) (b)

Figure 6.16 Model structures of graphite oxide (GO) (a) and sulfonated reduced graphene
oxide (rGOPhSO3 H) (b).
208 6 Carbon-Based Materials as Catalysts/Supports in Solvent-Free Organic Reactions

Figure 6.17 Proposed reaction mechanism of 5-substituted-3H-[1,3,4]-oxadiazol-2-ones


catalyzed by rGOPhSO3 H catalyst under solvent-free sonochemical conditions.

carbon sheet sites has been synthesized via sonication of polyphosphoric acid and
phosphoric acid, then the addition of dispersed GO in deionized water, followed by
heating at 95 ∘ C for several hours [168]. The functionalization method is flexible,
allowing the control of the acid content. The catalytic activity of the prepared
catalyst has been evaluated toward a one-pot synthesis of a four-component
reaction of hydrazines, malononitrile, ethyl acetoacetate, and aldehydes to syn-
thesize pyrano[2,3-c]pyrazole derivatives with an excellent yield of 80–90% within
15 minutes in an aqueous medium. The water-tolerant phosphonated GO solid
acid nanocatalyst has exhibited high stability and excellent recyclability since the
catalyst can be simply retrieved even after six successive cycles of reactions without
loss of the catalytic activity.

6.2.5 Porous Carbon Materials


Carbon materials have shown potential as catalysts in the selective oxidation
reaction of aromatic alcohols and their corresponding aldehydes, due to their low
cost, high specific area, stability, and porous textural properties [169]. Mesoporous
carbon with its narrow pore size distribution can act as catalyst support that
improves the catalytic activity of substrates and products. Particularly, ordered
mesoporous carbons (OMCs) have been extensively studied in a wide variety of
6.2 Solvent-Free Catalysis Using Carbon-Based Materials 209

applications in the field of electrochemistry, adsorption, and separation, compared


to ordinary mesoporous carbon due to the uniform pore size distribution and the
regular arrangement of the channel in OMCs that maximize the interactions with
analytes and accelerate the mass transport [170, 171]. This improvement in catalytic
activity arises from the narrow pore size distribution that enables providing a
large reaction space and shortens heat and mass transfer path for substrate and
products [172].

6.2.5.1 Oxidation of Alcohol and Hydrocarbons


Oxidation of natural materials has been widely adopted in preparing various impor-
tant industrial molecules [173]. Particularly, the selective oxidation of alcohols and
hydrocarbons is a primary reaction for the synthesis of organic compounds and
industrial chemistry [174]. Compared to homogeneous catalysts that are used in
aerobic oxidation [175], the heterogenous catalysts are more suitable for large-scale
production, readily used in flow reactors, and easy to recycle without loss of their
activity. Introduction of nitrogen atoms into ordered mesoporous carbons (NOMCs)
can further enhance the adsorption properties; this concept has been used in the
fabrication of the effective NOMCs with a 2D hexagonal mesoporous structure via
a one-pot hydrothermal route using aqueous self-assembly of amphiphilic triblock
copolymer Pluronic F127 as a template [172]. Melamine resin is added to modify
phenolic resin, where the –NH– group is connected to carbon atoms to form an
orderly mesoporous structure and function as an effective support to anchor Pd
NPs [172]. The developed Pd/NOMC catalyst is employed for the aerobic selective
oxidation of benzyl alcohol under solvent-free conditions with excellent catalytic
performance, affording TOF up to 8698 h−1 . The catalyst shows good catalytic sta-
bility for five consecutive cycles of the reaction representing its potential application.
Novel nanoporous nitrogen-doped carbon materials show a unique behavior
as heterogeneous catalysts [176]. Their outstanding performance is associated
with their natural features such as thermal and chemical stability, large surface
area, high nanoporosity, active functionalities, and hydrophobic surface proper-
ties. Nanoporous nitrogen-doped carbons show an outstanding performance in
oxidation reactions for their large surface area, high nanoporosity, and active func-
tionalities. nitrogen-doped carbon material derived from glucose as basic support
to construct Pd-based nitrogen-doped carbon (Pd-based C-GluA-550) catalysts
with different loadings (0.5–4 wt%) by ultrasonic-assisted deposition method has
been successfully prepared [176]. The two-step hydrothermal carbonization (HTC)
process of D-glucose has been realized at 200 ∘ C for eight hours with N-containing
additives, followed by carbonization of the HTC carbon materials at 550 or 1000 ∘ C
under the flow of N2 atmosphere. The successful solvent-free oxidation of indane,
hydrocarbons, and alcohols by Pd-based nitrogen-doped carbon catalysts with
atmosphere air or molecular oxygen is an important step toward the fabrication of
medicinal and biological chemicals. Solvent-free oxidation of indane in ambient air
(1 atm) with nitrogen-doped carbon (C-GluA-550) support has shown noncatalytic
oxidation activity. Using 4% Pd-based C-GluA-550, the oxidation conversion of
indane has reached 5.7% with 1-indanone as the main product at 80 ∘ C; this is most
210 6 Carbon-Based Materials as Catalysts/Supports in Solvent-Free Organic Reactions

likely because Pd-based nitrogen-doped carbon catalyst promotes the air oxidation
of sp3 C—H bond. However, the Pd-based heterogeneous catalyst without nitrogen
doping has exhibited lower indane conversion by one-third to 3.7% under the
same reaction conditions, compared with the nitrogen-doped one. With decreasing
the amount of Pd-based catalyst from 4 to 0.5 wt%, the oxidation reaction activity
typically increases four-fold affording TOF of 113–452 h−1 when the temperature
is increased to 120 ∘ C, respectively, with high selectivity for 1-indanone [176]. It
has demonstrated that the use of various N- or O-type ligands into the indane
oxidations to test the effect of different organic ligands on Pd NPs. The introduction
of acetylacetone and 8-hydroxyquinoline cannot significantly change the catalytic
activity, while N-type ligands remarkably improve the catalytic oxidations of
Pd-based nitrogen-doped carbon catalyst to some degree and the TOF is enhanced
from 452–863 h−1 by adding a small amount of 2-acetylpyridine, suggesting that a
suitable ligand can accelerate the catalytic process as the organic ligands may be
able to destabilize the indane-PdII intermediate and increase the reduction potential
PdII /Pd0 , thereby promoting the redox process [177].
An extended study has explored the potential use of Pd-based nitrogen-doped car-
bon catalyst in solvent-free oxidation of some other hydrocarbons, such as tetralin
to the corresponding α-tetralol, α-tetralone as the main product with 477 h−1 under
the optimized condition, which is a key step in the commercial production insecti-
cide and α-naphthol [178]. The catalyst is also applied in the selective oxidation of
ethylbenzene into acetophenone, affording a TOF of 477 h−1 , however, some mass
loss can be detected, probably because of the volatility of ethylbenzene by air bub-
bling. The novel Pd-based nitrogen-doped carbon is suitable for the selective oxi-
dation of alcohols by air with good activity, such as 1-phenylethanol, 2-octanol, and
benzyl alcohol to give acetophenone, 2-octanone, and benzaldehyde and benzyl ben-
zoate (giving a TOF of 14 802 h−1 ) as main products, respectively [176]. The high
catalytic performance of the catalyst can be attributed to the nanostructure proper-
ties of nanoporous nitrogen-containing spheres of approximately 20 nm in diameter
and Pd metal support-substrate junctions, which not only activates the electronic
properties of the metal NPs but also enhances the adsorption of substrates, giving
rise to increasing the selective oxidation. This catalytic oxidation method based on
Pd-based nitrogen-doped carbon catalyst is active, scalable, sustainable, recyclable,
and allows the in situ use of produced air in the oxidation reactions of many trans-
formations. In addition, the catalyst works in a heterogeneous manner, which make
it an ideal environmentally benign chemical process for potential use in industrial
applications.

6.2.5.2 Coupling of Amines


Oxidative coupling of amine is a crucial reaction that produces substituted imines
and their derivatives, which are paramount compounds in pharmaceuticals [179],
various fine chemicals [180], and molecular motors [181]. Synthesis of imines
through the oxidative coupling of amines prevails over the other conventional
methods that require the use of acid catalysts and unstable aldehydes [182].
Although the coupling reaction has been developed significantly, few obstacles can
6.2 Solvent-Free Catalysis Using Carbon-Based Materials 211

act as key challenges, as the rapid dehydrogenation of amines intermediates into


nitriles’ by-products which can reduce selectivity [183]. In addition, the purification
and separation of unstable imines are very difficult when using an organic solvent,
which also generates environmental issues. Furthermore, the metal catalysts,
which are mainly involved in these reactions, such as Pd [184], Au [185], Ru [186],
Cu [187], V [188], and Fe [189], are of high cost and contaminate the final products,
especially in pharmaceutical aims. Hence, developing a heterogeneous metal-free
catalyst with high selectivity and low cost for the oxidative coupling of amines is of
significant concern.
A few carbon-based catalysts [190], such as GO [191, 192], carbon nitride (g-C3 N4 )
[193], and boron- and nitrogen-codoped holey graphene (BNG) [194], have been
applied for the oxidative coupling of amines. However, these carbon-based cata-
lysts are limited either for their hard and costly preparation procedures or the huge
consumption of the volatile precursor [192, 194]. In contrast, mesoporous carbon
materials have presented an outstanding catalytic performance for oxidative cou-
pling of amines to imines. The high catalytic activity of these materials is related to
the physical and chemical tunability of the surface functionality [195, 196], the high
pore volume and large surface area [197], and the adaptability of an electron upon
doping with other elements (such as N, B, P, S) [198]. For instance, the enriched S
and P co-doped graphitic porous carbon is highly active and selective for the aerobic
oxidation of benzylic alcohols [199].
Mesoporous carbon catalysts have been synthesized via the carbothermal treat-
ment of macrocyclic compounds such as phthalocyanine (Pc) and porphyrin at
different temperatures (600, 700, 800, or 900 ∘ C) for the self-coupling of primary
benzylic amines [200]. Colloidal silica or ordered mesoporous silica SBA-15
has been employed as the hard template to fulfill the mesoporous structures.
The prepared mesoporous carbon catalysts have been used for the oxidative
self-coupling of benzylamine 6.2.5.2.1 under solventless reaction with ambient
oxygen at 100 ∘ C for five hours [200]. The catalyst has shown a major role in the
conversion reaction yield of benzylimine 6.2.5.2.2. For example, the self-coupling
of benzylamine 6.2.5.2.1 at a 100 mmol scale can smoothly proceed with only
0.47 wt% mesoporous carbon under solventless conditions, and the catalyst can
be reused several times without the need for additional reactivation treatment,
which highlighted the practical application of these materials. For instance, no
products have been produced in the absence of catalyst and using commercial
phthalocyanine has achieved only 20% conversion to an imine. The mesoporous
carbon catalysts carbonized at 800 ∘ C exhibit the highest catalytic efficiency with
a 90% conversion yield at 100 ∘ C after 5.5 hours, whereas catalysts carbonized at
600, 700, and 900 ∘ C have a dramatic decrease in the conversion yields. Defect sites
generated from the pyrolysis procedure, as well as surface area and pore volume,
are proposed as the crucial factors for the carbon catalysts to achieve the high
activity.
At the initial state of the reaction, both benzylamine 6.2.5.2.1 and molecular
oxygen are activated on the defective sites of the carbon catalyst and subsequently
are transformed into benzylimine 6.2.5.2.2 and H2 O2 intermediate, which in turn
212 6 Carbon-Based Materials as Catalysts/Supports in Solvent-Free Organic Reactions

Figure 6.18 An illustration of proposed reaction pathways for mesoporous carbon catalyst
(Pc-Ludox-8)-catalyzed oxidative coupling of primary amines to imines. Source: Chen et al.
[200]/American Chemical Society.

reacts immediately with another benzylamine 6.2.5.2.1 molecule to synthesize


benzylimine 6.2.5.2.2 as shown in Figure 6.18. The proposed transformation
of benzylamine to benzylimine may follow two pathways: (i) benzylamine may
hydrolyze to produce the corresponding aldehyde, and subsequently, condensed
with another molecule of benzylamine 6.2.5.2.1 to form benzylimine 6.2.5.2.2
(path A); (ii) addition of another benzylamine 6.2.5.2.1 molecule successively
releases NH3 to produce benzylimine 6.2.5.2.2 (path B). No detectable signal
of aldehyde has been found by GC-MS, implying that path A is unfavorable for
the Pc-Ludox-8-catalyzed solventless reaction of benzylamine. For confirming
this hypothesis, N-benzylidenemethyl amine has been added as the analogy of
benzylimine 6.2.5.2.2 to 1 equivalent of benzylamine 6.2.5.2.1 in the presence of
Pc-Ludox-8 catalyst (Figure 6.18). About 61% of N-benzylidenemethyl amine has
been transformed into benzylimine 6.2.5.2.2, and the residual amount has been
successfully recovered from the reaction mixtures. Based on this observation, it is
strongly supported that benzylimine 6.2.5.2.2 is intermediate, and the hydrolytic
process to produce the aldehyde (path A) is unfavorable, demonstrating that the
oxidative coupling of benzylamines proceeds through the path B.
6.3 Summary and Future Perspectives 213

Replacing silica colloid (Ludox HS-40) with mesoporous silica SBA-15 as a tem-
plate leads to a pronounced decrease in the yield to 75%. Further, the non-use of
the template hinders the conversion yield. A moderate yield is observed upon using
tetraphenylporphyrin (TPP) as the precursor. In addition to the role of the catalyst,
the O2 oxidant also shows an essential behavior for the reaction, where only 4% of
imine yield is obtained after the replacement of O2 with N2 . The catalysts show
potential performance in large-scale production and can be reused several times
without remarkable loss of catalytic efficiency. The unprecedented behavior of the
catalysts is attributed to their high total pore volume, large surface area, as well as
homogeneous distribution of defect sites.

6.3 Summary and Future Perspectives


Green chemistry has been widely used for the production of chemical products,
where it can achieve sustainability at the molecular level without involving
hazardous substances for industrial sectors. The utilization of green chemistry in
catalysis has achieved major economic and environmental benefits since the tar-
geted molecules can be prepared within short time and great energy efficiency. The
catalysis of organic reactions under solvent-free conditions is a largely green and
sustainable method compared to solvent-based reactions that have a potential risk of
causing serious hazards to release to the environment by the accumulation of organic
solvents. In addition, the solventless reactions are also controlled by the type of the
involved catalyst since its activity, surface area, porosity, porevolume, and func-
tionality play a key role in the yield of the products. Carbon-based catalysts possess
several advantageous properties that prevail their presence in solventless organic
reactions. Apart from the low environmental impact and the low costs, carbon-based
catalysts have a large surface area that makes them interesting for heterogeneous
catalysis. Moreover, porous carbons with their large pore volume, high nanoporos-
ity, and surface functionalities can be readily tuned to improve their catalytic
performances. Their use may save energy and may diminish solvent use in many
reactions.
The research on carbon nanomaterials with target functionalities can offer a plat-
form for nanoarchitectonics in the field of catalysis science and technology for their
sustainable use toward solventless organic reactions for which other toxic materi-
als catalysts are not appropriate or do not provide satisfactory performances. Apart
from these, although doping of heteroatom into carbon nanostructures generally
promotes catalytic performance, fundamental understandings of the effect of this
modification are still required. The continuously increasing demand and develop-
ment of cutting-edge environmentally friendly, energy-efficient, and cost-effective
synthetic approaches for catalysis technology, “ab initio” synthesis of optimized,
green carbon catalysts for on-demand applications in various organic transforma-
tions with the aid of advanced computational and analytical techniques is crucial
and further research must be intensively considered in future.
214 6 Carbon-Based Materials as Catalysts/Supports in Solvent-Free Organic Reactions

References

1 Sheldon, R.A. and Kochi, J.K. (1981). Metal-Catalyzed Oxidations of Organic


Compounds (Elsevier ed.). New York: Academic Press.
2 Anastas, P.T., Kirchhoff, M.M., and Williamson, T.C. (2001). Catalysis as a
foundational pillar of green chemistry. Appl. Catal., A 221 (1, 2): 3–13.
3 Anastas, P.T. and Warner, J.C. (2000). Green Chemistry: Theory and Practice.
Oxford: Oxford University Press.
4 Singh, M.S. and Chowdhury, S. (2012). Recent developments in solvent-free
multicomponent reactions: a perfect synergy for eco-compatible organic
synthesis. RSC Adv. 2 (11): 4547–4592.
5 Sheldon, R.A., Arends, I.W.C.E., and Hanefeld, U. (2007). Green Chemistry and
Catalysis. Wiley-VCH.
6 Gawande, M.B., Bonifácio, V.D.B., Luque, R. et al. (2013). Benign by design:
catalyst-free in-water, on-water green chemical methodologies in organic
synthesis. Chem. Soc. Rev. 42 (12): 5522–5551.
7 Doustkhah, E., Rostamnia, S., Zeynizadeh, B. et al. (2018). Efficient H2
generation using thiourea-based periodic mesoporous organosilica with Pd
nanoparticles. Chem. Lett. 47 (9): 1243–1245.
8 Zebardasti, A., Dekamin, M.G., Doustkhah, E. et al. (2020). Carbamate-
isocyanurate-bridged periodic mesoporous organosilica for van der Waals CO2
capture. Inorg. Chem. 59 (16): 11223–11227.
9 Doustkhah, E., Mohtasham, H., Farajzadeh, M. et al. (2020). Organosiloxane
tunability in mesoporous organosilica and punctuated Pd nanoparticles growth;
theory and experiment. Microporous Mesoporous Mater. 293: 109832.
10 Sharma, R.K., Sharma, S., Dutta, S. et al. (2015). Silica-nanosphere-based
organic–inorganic hybrid nanomaterials: synthesis, functionalization and
applications in catalysis. Green Chem. 17 (6): 3207–3230.
11 Tanaka, K. and Toda, F. (2000). Solvent-free organic synthesis. Chem. Rev.
100 (3): 1025–1074.
12 Li, C.J. (2005). Organic reactions in aqueous media with a focus on carbon–
carbon bond formations: a decade update. Chem. Rev. 105 (8): 3095–3166.
13 Sachdeva, H. (2020). Recent advances in the catalytic applications of GO/rGO
for green organic synthesis. Green Process. Synth. 9 (1): 515–537.
14 Assadi, M.H.N. and Sahajwalla, V. (2014). Recycling end-of-life polycarbonate
in steelmaking: ab initio study of carbon dissolution in molten iron. Ind. Eng.
Chem. Res. 53 (10): 3861–3864.
15 Assadi, M.H.N. and Sahajwalla, V. (2014). Polymers’ surface interactions with
molten iron: a theoretical study. Chem. Phys. 443 107–111.
16 Yang, Y., Chiang, K., and Burke, N. (2011). Porous carbon-supported catalysts
for energy and environmental applications: a short review. Catal. Today 178 (1):
197–205.
17 Rodriguez-Reinoso, F. (1998). The role of carbon materials in heterogeneous
catalysis. Carbon 36 (3): 159–175.
References 215

18 Calvino-Casilda, V., López-Peinado, A., Durán-Valle, C. et al. (2010). Last


decade of research on activated carbons as catalytic support in chemical
processes. Catal. Rev. 52 (3): 325–380.
19 Perozo-Rondon, E., Calvino-Casilda, V., Martin-Aranda, R. et al. (2006).
Catalysis by basic carbons: preparation of dihydropyridines. Appl. Surf. Sci.
252 (17): 6080–6083.
20 Durán-Valle, C.J., Fonseca, I.M., Calvino-Casilda, V. et al. (2005). Sonocatalysis
and alkaline-doped carbons: an efficient method for the synthesis of chalcones
in heterogeneous media. Catal. Today 107 500–506.
21 Rubio-Gómez, J., Mart𝚤n-Aranda, R., Rojas-Cervantes, M. et al. (1999).
Ultrasound enhanced reactions involving activated carbons as catalysts:
synthesis of α,β-unsaturated nitriles. Carbon 37 (2): 213–219.
22 Calvino-Casilda, V., Martín-Aranda, R., and López-Peinado, A. (2009).
Microwave assisted green synthesis of long-chain 1-alkylimidazoles and
medium-chain 1-alkyl-2-methylimidazoles with antiviral properties catalyzed
by basic carbons. Catal. Lett. 129 (3-4): 281–286.
23 Durán-Valle, C., Madrigal-Martínez, M., Martínez-Gallego, M. et al. (2012).
Activated carbon as a catalyst for the synthesis of N-alkylimidazoles and
imidazolium ionic liquids. Catal. Today 187 (1): 108–114.
24 Sartori, G., Ballini, R., Bigi, F. et al. (2004). Protection (and deprotection) of
functional groups in organic synthesis by heterogeneous catalysis. Chem. Rev.
104 (1): 199–250.
25 Otera, J. and Nishikido, J. (2010). Esterification: Methods, Reactions, and
Applications. Weinheim: Wiley-VCH.
26 Scriven, E.F. (1983). 4-Dialkylaminopyridines: super acylation and alkylation
catalysts. Chem. Soc. Rev. 12 (2): 129–161.
27 Sano, T., Ohashi, K., and Oriyama, T. (1999). Remarkably fast acylation of
alcohols with benzoyl chloride promoted by TMEDA. Synthesis 1999 (7):
1141–1144.
28 Karimi, B. and Maleki, J. (2003). Lithium trifluoromethanesulfonate (LiOTf)
as a recyclable catalyst for highly efficient acetylation of alcohols and
diacetylation of aldehydes under mild and neutral reaction conditions. J. Organ.
Chem. 68 (12): 4951–4954.
29 Miyashita, M., Shiina, I., Miyoshi, S. et al. (1993). A new and efficient
esterification reaction via mixed anhydrides by the promotion of a catalytic
amount of Lewis acid. Bull. Chem. Soc. Jpn. 66 (5): 1516–1527.
30 Rajabi, F. (2009). A heterogeneous cobalt (II) Salen complex as an efficient
and reusable catalyst for acetylation of alcohols and phenols. Tetrahedron Lett.
50 (4): 395–397.
31 Phukan, P. (2004). Iodine as an extremely powerful catalyst for the acetylation
of alcohols under solvent-free conditions. Tetrahedron Lett. 45 (24): 4785–4787.
32 Lugemwa, F.N., Shaikh, K., and Hochstedt, E. (2013). Facile and efficient
acetylation of primary alcohols and phenols with acetic anhydride catalyzed by
dried sodium bicarbonate. Catalysts 3 (4): 954–965.
216 6 Carbon-Based Materials as Catalysts/Supports in Solvent-Free Organic Reactions

33 Osiglio, L., Sathicq, Á.G., Romanelli, G.P. et al. (2012). Borated zirconia
modified with ammonium metatungstate as catalyst in alcohol acetylation.
J. Mol. Catal. A: Chem. 359 97–103.
34 Vedejs, E. and Diver, S. (1993). Tributylphosphine: a remarkable acylation
catalyst. J. Am. Chem. Soc. 115 (8): 3358–3359.
35 Cope, A.C. and Herrick, E.C. (1963). Organic Synthesis Collective
(ed. N. Rabjohn). New York: Wiley.
36 Baker, R., Bordwell, F., Hauser, C. et al. (1955). tert-Butyl acetate. Org. Synth.
3 141.
37 Iqbal, J. and Srivastava, R.R. (1992). Cobalt(II) chloride catalyzed acylation of
alcohols with acetic anhydride: scope and mechanism. J. Organ. Chem. 57 (7):
2001–2007.
38 Ishihara, K., Kubota, M., Kurihara, H. et al. (1996). Scandium trifluoromethane-
sulfonate as an extremely active Lewis acid catalyst in acylation of alcohols
with acid anhydrides and mixed anhydrides. J. Organ. Chem. 61 (14):
4560–4567.
39 Lopez, I., Bravo, J.L., Caraballo, M. et al. (2011). Task-oriented use of ionic
liquids: efficient acetylation of alcohols and phenols. Tetrahedron Lett. 52 (26):
3339–3341.
40 López-Coca, I.M., Izquierdo, S., Silvero, G. et al. (2019). Sustainable carbon-
based materials as heterogeneous catalysts in solvent-free acetylation reactions.
Proceedings 9 (1): 40.
41 Zhou, W.J., Wischert, R., Xue, K. et al. (2014). Highly selective liquid-phase
oxidation of cyclohexane to KA oil over Ti-MWW catalyst: evidence of
formation of oxyl radicals. ACS Catal. 4 (1): 53–62.
42 Schuchardt, U., Cardoso, D., Sercheli, R. et al. (2001). Cyclohexane oxidation
continues to be a challenge. Appl. Catal., A 211 (1): 1–17.
43 Lü, G., Zhao, R., Qian, G. et al. (2004). A highly efficient catalyst Au/MCM-41
for selective oxidation cyclohexane using oxygen. Catal. Lett. 97 (3): 115–118.
44 Kishore, M.J.L. and Kumar, A. (2007). Heteronuclear macrocyclic iron-copper
complex catalyst covalently bonded to modified alumina catalyst for oxidation
of cyclohexane. Ind. Eng. Chem. Res. 46 (14): 4787–4798.
45 Liu, Y., Tsunoyama, H., Akita, T. et al. (2011). Aerobic oxidation of cyclohexane
catalyzed by size-controlled Au clusters on hydroxyapatite: size effect in the
sub-2 nm regime. ACS Catal. 1 (1): 2–6.
46 Sadiq, M., Khan, M., Numan, M. et al. (2017). Tuning of activated carbon for
solvent-free oxidation of cyclohexane. J. Chem. 2017: 5732761 (1-8).
47 Laohapornchaiphan, J., Smith, C.B., and Smith, S.M. (2017). One-step
preparation of carbon-based solid acid catalyst from water hyacinth leaves
for esterification of oleic acid and dehydration of xylose. Chem. Asian J. 12 (24):
3178–3186.
48 Doustkhah, E., Lin, J., Rostamnia, S. et al. (2019). Development of sulfonic-
acid-functionalized mesoporous materials: synthesis and catalytic applications.
Chem. Eur. J. 25 (7): 1614–1635.
References 217

49 Konwar, L.J., Mäki-Arvela, P., and Mikkola, J.P. (2019). SO3 H-Containing
functional carbon materials: Synthesis, structure, and acid catalysis. Chem.
Rev. 119 (22): 11576–11630.
50 Deli, J., Lóránd, T., Szabó, D. et al. (1984). Potentially bioactive pyrimidine
derivatives. 1. 2-Amino-4-aryl-8-arylidene-3,4,5,6,7,8-hexahydroquinazoline.
Pharmazie 39 (8): 539–540.
51 Khodaei, M.M., Bahrami, K., and Khedri, M. (2007). ZrCl4 as an efficient
catalyst for crossed-aldol condensation of cyclic ketones with aromatic
aldehydes in refluxing ethanol. J. Chin. Chem. Soc. 54 (3): 807–810.
52 Zheng, M., Wang, L., Shao, J. et al. (1997). A facile synthesis of α,α’-
bis(substituted benzylidene) cycloalkanones catalyzed by bis(p-ethoxyphenyl)
telluroxide (bmpto) under microwave irradiation. Synth. Commun. 27 (2):
351–354.
53 Yadav, J., Reddy, B.S., Nagaraju, A. et al. (2002). Microwave assisted synthesis
of α,α′ -bis(benzylidene) ketones in dry media. Synth. Commun. 32 (6): 893–896.
54 Esmaeili, A.A., Tabas, M.S., Nasseri, M.A. et al. (2005). Solvent-free crossed
aldol condensation of cyclic ketones with aromatic aldehydes assisted by
microwave irradiation. Chem. Monthly 136 (4): 571–576.
55 Olah, G.A., Iyer, P.S., and Prakash, G.S. (2003). Perfluorinated resinsulfonic
acid (Nafion-H®) catalysis in synthesis. In: Across Conventional Lines, 594–612.
World Scientific.
56 Inagaki, S., Guan, S., Ohsuna, T. et al. (2002). An ordered mesoporous
organosilica hybrid material with a crystal-like wall structure. Nature
416 (6878): 304–307.
57 Okuhara, T. (2002). Water-tolerant solid acid catalysts. Chem. Rev. 102 (10):
3641–3666.
58 Hathaway, B.A. (1987). An aldol condensation experiment using a number of
aldehydes and ketones. J. Chem. Educ. 64 (4): 367.
59 Nakano, T., Irifune, S., Umano, S. et al. (1987). Cross-condensation reactions
of cycloalkanones with aldehydes and primary alcohols under the influence of
zirconocene complexes. J. Organ. Chem. 52 (11): 2239–2244.
60 Abbas, Z., Ghani, K., Shokrolahi, A. et al. (2008). Carbon-based solid acid
as an efficient and reusable catalyst for cross-aldol condensation of ketones
with aromatic aldehydes under solvent-free conditions. Chin. J. Catal. 29 (7):
602–606.
61 Dömling, A. and Ugi, I. (2000). Multicomponent reactions with isocyanides.
Angew. Chem. Int. Ed. 39 (18): 3168–3210.
62 Gordeev, M.F., Patel, D.V., and Gordon, E.M. (1996). Approaches to combina-
torial synthesis of heterocycles: a solid-phase synthesis of 1,4-dihydropyridines.
J. Organ. Chem. 61 (3): 924–928.
63 Tu, S., Miao, C., Gao, Y. et al. (2004). A Novel Cascade reaction of aryl
aldoxime with dimedone under microwave irradiation: the synthesis of
N-hydroxylacridine. Synlett 2004 (02): 0255–0258.
64 Das, B., Thirupathi, P., Mahender, I. et al. (2006). Amberlyst-15: an effi-
cient reusable heterogeneous catalyst for the synthesis of 1,8-dioxo-
218 6 Carbon-Based Materials as Catalysts/Supports in Solvent-Free Organic Reactions

octahydroxanthenes and 1,8-dioxo-decahydroacridines. J. Mol. Catal. A: Chem.


247 (1,2): 233–239.
65 Chandrasekhar, S., Rao, Y.S., Sreelakshmi, L. et al. (2008).
Tris(pentafluorophenyl)borane-catalyzed three-component reaction for the
synthesis of 1,8-dioxodecahydroacridines under solvent-free conditions.
Synthesis 2008 (11): 1737–1740.
66 Shen, W., Wang, L.M., Tian, H. et al. (2009). Brønsted acidic imidazolium
salts containing perfluoroalkyl tails catalyzed one-pot synthesis of 1,8-dioxo-
decahydroacridines in water. J. Fluorine Chem. 130 (6): 522–527.
67 Davoodnia, A., Khojastehnezhad, A., and Tavakoli-Hoseini, N. (2011). Carbon-
based solid acid as an efficient and reusable catalyst for the synthesis of
1,8-dioxodecahydroacridines under solvent-free conditions. Bull. Korean Chem.
Soc. 32 (7): 2243–2248.
68 Niknam, K., Panahi, F., Saberi, D. et al. (2010). Silica-bonded S-sulfonic acid
as recyclable catalyst for the synthesis of 1,8-dioxo-decahydroacridines and
1,8-dioxo-octahydroxanthenes. J. Heterocycl. Chem. 47 (2): 292–300.
69 Loev, B. and Snader, K.M. (1965). The Hantzsch reaction. I. Oxidative
dealkylation of certain dihydropyridines. J. Organ. Chem. 30 (6): 1914–1916.
70 Davoodnia, A. and Khojastehnezhad, A. (2012). Carbon-based solid acid
catalyzed synthesis of polyhydroquinoline derivatives via Hantzsch reaction
under solvent-free conditions. J. Chil. Chem. Soc. 57 (4): 1385–1387.
71 Heeres, J., Backx, L., Mostmans, J. et al. (1979). Antimycotic imidazoles. Part 4.
Synthesis and antifungal activity of ketoconazole, a new potent orally active
broad-spectrum antifungal agent. J. Med. Chem. 22 (8): 1003–1005.
72 Kidwai, M., Mothsra, P., Bansal, V. et al. (2007). One-pot synthesis of highly
substituted imidazoles using molecular iodine: a versatile catalyst. J. Mol. Catal.
A: Chem. 265 (1, 2, 182): 177.
73 Sharma, S.D., Hazarika, P., and Konwar, D. (2008). An efficient and one-pot
synthesis of 2,4,5-trisubstituted and 1,2,4,5-tetrasubstituted imidazoles catalyzed
by InCl3 ⋅3H2 O. Tetrahedron Lett. 49 (14): 2216–2220.
74 Balalaie, S., Hashemi, M.M., and Akhbari, M. (2003). A novel one-pot synthesis
of tetrasubstituted imidazoles under solvent-free conditions and microwave
irradiation. Tetrahedron Lett. 44 (8): 1709–1711.
75 Lantos, I., Zhang, W.Y., Shui, X. et al. (1993). Synthesis of imidazoles via
hetero-Cope rearrangements. J. Organ. Chem. 58 (25): 7092–7095.
76 Claiborne, C.F., Liverton, N.J., and Nguyen, K.T. (1998). An efficient synthesis
of tetrasubstituted imidazoles from N-(2-oxo)-amides. Tetrahedron Lett. 39 (49):
8939–8942.
77 Tavakoli-Hoseini, N. and Davoodnia, A. (2011). Carbon-based solid acid as
an efficient and reusable catalyst for one-pot synthesis of tetrasubstituted
imidazoles under solvent-free conditions. Chin. J. Chem. 29 203–206.
78 Davoodnia, A., Mahjoobin, R., and Tavakoli-Hoseini, N. (2014). A facile,
green, one-pot synthesis of amidoalkyl naphthols under solvent-free conditions
catalyzed by a carbon-based solid acid. Chin. J. Catal. 35 (4): 490–495.
References 219

79 Nguyen, V.T., Nguyen, H.T., and Tran, P.H. (2021). One-pot three-component
synthesis of 1-amidoalkyl naphthols and polyhydroquinolines using a deep
eutectic solvent: a green method and mechanistic insight. New J. Chem. 45 (4):
2053–2059.
80 Shokrolahi, A., Zali, A., and Keshavarz, M.H. (2007). Wet carbon-based solid
acid/NaNO3 as a mild and efficient reagent for nitration of aromatic compound
under solvent free conditions. Chin. Chem. Lett. 18 (9): 1064–1066.
81 Yan, G. and Yang, M. (2013). Recent advances in the synthesis of aromatic nitro
compounds. Org. Biomol. Chem. 11 (16): 2554–2566.
82 Olah, G.A., Malhotra, R., and Narang, S.C. (2003). Nitration: methods and
mechanisms. In: Across Conventional Lines, 975–979. World Scientific.
83 Boros, E.E., Thompson, J.B., Katamreddy, S.R. et al. (2009). Facile reductive
amination of aldehydes with electron-deficient anilines by acyloxyborohy-
drides in TFA: application to a diazaindoline scale-up. J. Organ. Chem. 74 (9):
3587–3590.
84 Devi, C.L., Olusegun, O.S., Kumar, C.N.S.S.P. et al. (2009). Novel combination
of sodium borohydride and reusable polyaniline salt catalyst for rapid and
efficient reductive amination of carbonyl compounds. Catal. Lett. 132 (3-4): 480.
85 Hara, M., Yoshida, T., Takagaki, A. et al. (2004). A carbon material as a strong
protonic acid. Angew. Chem. 116 (22): 3015–3018.
86 Shokrolahi, A., Zali, A., and Keshavarz, M.H. (2011). Reductive amination of
aldehydes and ketones by NaBH4 using carbon-based solid acid (CBSA) as
catalyst. Green Chem. Lett. Rev. 4 (3): 195–203.
87 Ghulam, S., Aamer, S., and Pervaiz Ali, C. (2018). A review on the recent
trends in synthetic strategies and applications of xanthene dyes. Mini Rev
Organ. Chem. 15 (3): 166–197.
88 Bhattacharya, A.K. and Rana, K.C. (2007). Microwave-assisted synthesis of
14-aryl-14H-dibenzo[aj]xanthenes catalysed by methanesulfonic acid under
solvent-free conditions. Mendeleev Commun. 4 (17): 247–248.
89 Mirkhani, V., Moghadam, M., Tangestaninejad, S. et al. (2009). Highly efficient
synthesis of 14-aryl-14H-dibenzo[a,j]xanthenes catalyzed by carbon-based solid
acid under solvent-free conditions. Synth. Commun. 39 (24): 4328–4340.
90 Ramin, L., Assadi, M.H.N., and Sahajwalla, V. (2014). High-density
polyethylene degradation into low molecular weight gases at 1823 K: an
atomistic simulation. J. Anal. Appl. Pyrolysis 110 318–321.
91 Liu, W.Y., Liu, N., Zheng, R.Y. et al. (2013). Coaled carbon-based
solid acid: a new and efficient catalyst for click synthesis of 3,4-
dihydropyrimidin-2(1H)-ones under solvent-free conditions. Adv. Mater. Res.
634–638 504–507.
92 Chitra, S. and Pandiarajan, K. (2009). Calcium fluoride: an efficient and
reusable catalyst for the synthesis of 3,4-dihydropyrimidin-2(1H)-ones and
their corresponding 2(1H)thione: an improved high yielding protocol for the
Biginelli reaction. Tetrahedron Lett. 50 (19): 2222–2224.
93 Ghorbani, F., Pourmousavi, S.A., and Kiyani, H. (2019). Novel carbon-based
solid acid from green pistachio peel as an efficient catalyst for the
220 6 Carbon-Based Materials as Catalysts/Supports in Solvent-Free Organic Reactions

chemoselective acylation, acetalization and thioacetalization of aldehydes,


synthesis of biscoumarins and antimicrobial evaluation. Curr. Organocatal.
7 (1): 55–80.
94 Fahid, F. and Pourmousavi, S.A. (2014). Sulfonated polyanthracene-catalyzed
highly efficient and chemoselective thioacetalization of carbonyl compounds
and transthioacetalization of acetals and acylals. J. Sulfur Chem. 36 (1): 16–29.
95 Zhang, F., Liu, H., Zhang, Q.J. et al. (2010). Silica phosphoric acid: an
efficient and recyclable catalyst for the solvent-free synthesis of acylals and
their deprotection in MeOH. Synth. Commun. 40 (21): 3240–3250.
96 Hoseinabadi, Z., Pourmousavi, S.A., and Zamani, M. (2016). Synthesis of
sulfonated carbon-based solid acid as a novel and efficient nanocatalyst for
the preparation of highly functionalized piperidines and acylals: a DFT study.
Res. Chem. Intermed. 42 (6): 6105–6124.
97 Su, H., Zhang, K.X., Zhang, B. et al. (2017). Activating cobalt nanoparticles
via the Mott-Schottky effect in nitrogen-rich carbon shells for base-free aerobic
oxidation of alcohols to esters. J. Am. Chem. Soc. 139 (2): 811–818.
98 Saberi, F., Rodríguez-Padrón, D., Doustkhah, E. et al. (2019).
Mechanochemically modified aluminosilicates for efficient oxidation of vanillyl
alcohol. Catal. Commun. 118 65–69.
99 Zhang, J., Leitus, G., Ben-David, Y. et al. (2005). Facile conversion of alcohols
into esters and dihydrogen catalyzed by new ruthenium complexes. J. Am.
Chem. Soc. 127 (31): 10840–10841.
100 Bai, X.F., Ye, F., Zheng, L.S. et al. (2012). Hydrosilane and bismuth-accelerated
palladium catalyzed aerobic oxidative esterification of benzylic alcohols with
air. Chem. Commun. 48 (68): 8592–8594.
101 Gowrisankar, S., Neumann, H., and Beller, M. (2011). General and selective
palladium-catalyzed oxidative esterification of alcohols. Angew. Chem. Int. Ed.
50 (22): 5139–5143.
102 Jagadeesh, R.V., Junge, H., Pohl, M.M. et al. (2013). Selective oxidation of
alcohols to esters using heterogeneous Co3 O4 –N@C catalysts under mild
conditions. J. Am. Chem. Soc. 135 (29): 10776–10782.
103 Zhu, C., Li, H., Fu, S. et al. (2016). Highly efficient nonprecious metal catalysts
towards oxygen reduction reaction based on three-dimensional porous carbon
nanostructures. Chem. Soc. Rev. 45 (3): 517–531.
104 Zhao, Y., Zhang, J., Li, K. et al. (2016). Electrospun cobalt embedded porous
nitrogen doped carbon nanofibers as an efficient catalyst for water splitting.
J. Mater. Chem. A 4 (33): 12818–12824.
105 Sadeghi Rad, T., Ansarian, Z., Khataee, A. et al. (2021). N-doped graphitic
carbon as a nanoporous MOF-derived nanoarchitecture for the efficient
sonocatalytic degradation process. Sep. Purif. Technol. 256: 117811.
106 Du, C.Y., Zhao, T.S., and Liang, Z.X. (2008). Sulfonation of carbon-nanotube
supported platinum catalysts for polymer electrolyte fuel cells. J. Power Sources
176 (1): 9–15.
107 Serp, P. and Castillejos, E. (2010). Catalysis in carbon nanotubes.
ChemCatChem 2 (1): 41–47.
References 221

108 Aalinejad, M., Pesyan, N.N., and Doustkhah, E. (2020). Diaza crown-type
macromocycle (kryptofix 22) functionalised carbon nanotube for efficient Ni2+
loading; a unique catalyst for cross-coupling reactions. Mol. Catal. 494: 111117.
109 Nandan, D., Zoppellaro, G., Medřík, I. et al. (2018). Cobalt-entrenched N-, O-,
and S-tridoped carbons as efficient multifunctional sustainable catalysts for
base-free selective oxidative esterification of alcohols. Green Chem. 20 (15):
3542–3556.
110 Liu, C., Wang, J., Meng, L. et al. (2011). Palladium-catalyzed aerobic oxidative
direct esterification of alcohols. Angew. Chem. Int. Ed. 50 (22): 5144–5148.
111 Albertsson, A.C. and Varma, I.K. (2003). Recent developments in ring opening
polymerization of lactones for biomedical applications. Biomacromolecules 4 (6):
1466–1486.
112 Mori, K., Hara, T., Mizugaki, T. et al. (2004). Hydroxyapatite-supported
palladium nanoclusters: a highly active heterogeneous catalyst for selective
oxidation of alcohols by use of molecular oxygen. J. Am. Chem. Soc. 126 (34):
10657–10666.
113 Li, F., Zhang, Q., and Wang, Y. (2008). Size dependence in solvent-free aerobic
oxidation of alcohols catalyzed by zeolite-supported palladium nanoparticles.
Appl. Catal., A 334 (1): 217–226.
114 Rostamnia, S., Doustkhah, E., Karimi, Z. et al. (2015). Surfactant-exfoliated
highly dispersive Pd-supported graphene oxide nanocomposite as a catalyst for
aerobic aqueous oxidations of alcohols. ChemCatChem 7 (11): 1678–1683.
115 Khan, M., Shaik, M.R., Adil, S.F. et al. (2020). Facile synthesis of Pd@graphene
nanocomposites with enhanced catalytic activity towards Suzuki coupling
reaction. Sci. Rep. 10 (1): 11728.
116 Chen, Y., Guo, Z., Chen, T. et al. (2010). Surface-functionalized TUD-1
mesoporous molecular sieve supported palladium for solvent-free aerobic
oxidation of benzyl alcohol. J. Catal. 275 (1): 11–24.
117 Brönnimann, C., Mallat, T., and Baiker, A. (1995). Localized basification of
catalytic surfaces enhances the selective oxidation of L-sorbose over supported
Pt catalysts modified with tertiary amines. J. Chem. Soc., Chem. Commun.
13 1377–1378.
118 Zhou, C., Chen, Y., Guo, Z. et al. (2011). Promoted aerobic oxidation of benzyl
alcohol on CNT supported platinum by iron oxide. Chem. Commun. 47 (26):
7473–7475.
119 Tan, H.T., Chen, Y., Zhou, C. et al. (2012). Palladium nanoparticles supported
on manganese oxide–CNT composites for solvent-free aerobic oxidation of
alcohols: Tuning the properties of Pd active sites using MnOx. Appl. Catal., B
119, 120 166–174.
120 Yan, Y., Chen, Y., Jia, X. et al. (2014). Palladium nanoparticles supported on
organosilane-functionalized carbon nanotube for solvent-free aerobic oxidation
of benzyl alcohol. Appl. Catal., B 156-157 385–397.
121 Yan, Y., Jia, X., and Yang, Y. (2016). Palladium nanoparticles supported on CNT
functionalized by rare-earth oxides for solvent-free aerobic oxidation of benzyl
alcohol. Catal. Today 259 292–302.
222 6 Carbon-Based Materials as Catalysts/Supports in Solvent-Free Organic Reactions

122 Zhang, M., Tang, J., Mujumdar, A.S. et al. (2006). Trends in microwave-related
drying of fruits and vegetables. Trends Food Sci. Technol. 17 (10): 524–534.
123 Consoli, G.M.L., Galante, E., Daquino, C. et al. (2006). Hydroxycinnamic acid
clustered by a calixarene platform: radical scavenging and antioxidant activity.
Tetrahedron Lett. 47 (37): 6611–6614.
124 Shimizu, K., Kontani, S., Yamada, S. et al. (2010). Design of active centers for
bisphenol-A synthesis by organic–inorganic dual modification of heteropolyacid.
Appl. Catal., A 380 (1): 33–39.
125 D’Souza, D.M. and Müller, T.J.J. (2007). Multi-component syntheses of
heterocycles by transition-metal catalysis. Chem. Soc. Rev. 36 (7): 1095–1108.
126 Zhang, W., Zhang, Z., and Zhang, Y. (2011). The application of carbon
nanotubes in target drug delivery systems for cancer therapies. Nanoscale
Res. Lett. 6 (1): 555.
127 Derycke, V., Auvray, S., Borghetti, J. et al. (2009). Carbon nanotube chemistry
and assembly for electronic devices. C.R. Phys. 10 (4): 330–347.
128 Serp, P., Corrias, M., and Kalck, P. (2003). Carbon nanotubes and nanofibers in
catalysis. Appl. Catal., A 253 (2): 337–358.
129 Fareghi-Alamdari, R., Golestanzadeh, M., and Zekri, N. (2016). Solvent-free
synthesis of trisphenols as starting precursors for the synthesis of calix[4]arenes
using sulfonated multi-walled carbon nanotubes. New J. Chem. 40 (4):
3400–3412.
130 Reza, F.A., Mohsen, G., Farima, A. et al. (2013). Regiospecific, one-pot, and
pseudo-five-component synthesis of 6,6′ -(arylmethylene)bis(2-(tert-butyl)
4-methylphenol) antioxidants using highly sulfonated multi-walled carbon
nanotubes under solvent-free conditions. Can. J. Chem. 91 (10): 982–991.
131 Osorio, A.G., Silveira, I.C.L., Bueno, V.L. et al. (2008).
H2 SO4 /HNO3 /HCl-Functionalization and its effect on dispersion of carbon
nanotubes in aqueous media. Appl. Surf. Sci. 255 (5, Part 1): 2485–2489.
132 Peng, H., Alemany, L.B., Margrave, J.L. et al. (2003). Sidewall carboxylic acid
functionalization of single-walled carbon nanotubes. J. Am. Chem. Soc. 125 (49):
15174–15182.
133 Akbarzadeh, P., Koukabi, N., and Hosseini, M.M. (2020). Magnetic carbon
nanotube as a highly stable and retrievable support for the heterogenization
of sulfonic acid and its application in the synthesis of 2-(1H-tetrazole-5-yl)
acrylonitrile derivatives. J. Heterocyclic Chem. 57 (6): 2455–2465.
134 Banerjee, A.N. (2018). Graphene and its derivatives as biomedical materials:
future prospects and challenges. Interface Focus 8 (3): 20170056.
135 Wang, J., Gong, C., Wen, S. et al. (2018). Proton exchange membrane based
on chitosan and solvent-free carbon nanotube fluids for fuel cells applications.
Carbohydr. Polym. 186 200–207.
136 Rostamnia, S., Zeynizadeh, B., Doustkhah, E. et al. (2015). Exfoliated Pd
decorated graphene oxide nanosheets (PdNP–GO/P123): Non-toxic, ligandless
and recyclable in greener Hiyama cross-coupling reaction. J. Colloid Interface
Sci. 451 46–52.
References 223

137 Yam, K.M., Guo, N., Jiang, Z. et al. (2020). Graphene-based heterogeneous
catalysis: role of graphene. Catalysts 10 (1): 53.
138 Kong, X.K., Chen, C.L., and Chen, Q.W. (2014). Doped graphene for metal-free
catalysis. Chem. Soc. Rev. 43 (8): 2841–2857.
139 Georgakilas, V., Otyepka, M., Bourlinos, A.B. et al. (2012). Functionalization of
graphene: covalent and non-covalent approaches, derivatives and applications.
Chem. Rev. 112 (11): 6156–6214.
140 Jia, Y., Zhang, L., Gao, G. et al. (2017). A heterostructure coupling of exfo-
liated Ni–Fe hydroxide nanosheet and defective graphene as a bifunctional
electrocatalyst for overall water splitting. Adv. Mater. 29 (17): 1700017.
141 Tarcan, R., Todor-Boer, O., Petrovai, I. et al. (2020). Reduced graphene oxide
today. J. Mater. Chem. C 8 (4): 1198–1224.
142 Wang, Y., Wen, Z., Zhang, H. et al. (2016). CuO nanorods-decorated reduced
graphene oxide nanocatalysts for catalytic oxidation of CO. Catalysts 6 (12):
214.
143 Jain, Y., Kumari, M., Laddha, H. et al. (2019). Ultrasound promoted
fabrication of CuO-graphene oxide nanocomposite for facile synthesis of
fluorescent coumarin based 1, 4-disubsituted 1, 2, 3-triazoles in aqueous media.
ChemistrySelect 4 (23): 7015–7026.
144 Hussain, N., Gogoi, P., Das, M.R. et al. (2017). Development of novel efficient
2D nanocomposite catalyst towards the three-component coupling reaction for
the synthesis of imidazo [1, 2-a] pyridines. Appl. Catal., A 542 368–379.
145 Parmee, R.J., Collins, C.M., Milne, W.I. et al. (2015). X-ray generation using
carbon nanotubes. Nano Convergence 2 (1): 1–27.
146 Choi, J., Oh, H., Han, S.W. et al. (2017). Preparation and characterization
of graphene oxide supported Cu, Cu2 O, and CuO nanocomposites and their
high photocatalytic activity for organic dye molecule. Curr. Appl Phys. 17 (2):
137–145.
147 Sarkar, R., Gupta, A., Jamatia, R. et al. (2020). Reduced graphene oxide
supported copper oxide nanocomposites: An efficient heterogeneous
and reusable catalyst for the synthesis of ynones, 1, 3-diynes and 1,5-
benzodiazepines in one-pot under sustainable reaction conditions. Appl.
Organomet. Chem. 34 (7): e5646.
148 Dandia, A., Bansal, S., Sharma, R. et al. (2018). Microwave-assisted
nanocatalysis: A CuO NPs/rGO composite as an efficient and recyclable catalyst
for the Petasis-borono–Mannich reaction. RSC Adv. 8 (53): 30280–30288.
149 Halder, P., Roy, T., and Das, P. (2021). Recent developments in selective
N-arylation of azoles. Chem. Commun. 57 5235–5249.
150 Strieter, E.R., Bhayana, B., and Buchwald, S.L. (2009). Mechanistic studies on
the copper-catalyzed N-arylation of amides. J. Am. Chem. Soc. 131 (1): 78–88.
151 Movahed, S.K., Dabiri, M., and Bazgir, A. (2014). A one-step method for
preparation of Cu@ Cu2 O nanoparticles on reduced graphene oxide and their
catalytic activities in N-arylation of N-heterocycles. Appl. Catal., A 481 79–88.
152 Gupta, A., Jamatia, R., Patil, R.A. et al. (2018). Copper oxide/reduced graphene
oxide nanocomposite-catalyzed synthesis of flavanones and flavanones with
224 6 Carbon-Based Materials as Catalysts/Supports in Solvent-Free Organic Reactions

triazole hybrid molecules in one pot: a green and sustainable approach. ACS
Omega 3 (7): 7288–7299.
153 Frattaruolo, L., Carullo, G., Brindisi, M. et al. (2019). Antioxidant and anti-
inflammatory activities of flavanones from Glycyrrhiza glabra L.(licorice) leaf
phytocomplexes: Identification of licoflavanone as a modulator of NF-kB/MAPK
pathway. Antioxidants 8 (6): 186.
154 Khatun, R., Biswas, S., Islam, S. et al. (2019). Modified Graphene oxide based
zinc composite: an efficient catalyst for N-formylation and carbamate formation
reactions through CO2 fixation. ChemCatChem 11 (4): 1303–1312.
155 Hutchings, G.J. (2008). Nanocrystalline gold and gold palladium alloy catalysts
for chemical synthesis. Chem. Commun. (10): 1148–1164.
156 Zhang, Q., Deng, W., and Wang, Y. (2011). Effect of size of catalytically active
phases in the dehydrogenation of alcohols and the challenging selective
oxidation of hydrocarbons. Chem. Commun. 47 (33): 9275–9292.
157 Gervasini, A. and Auroux, A. (1991). Acidity and basicity of metal oxide
surfaces II. Determination by catalytic decomposition of isopropanol. J. Catal.
131 (1): 190–198.
158 Fang, W., Chen, J., Zhang, Q. et al. (2011). Hydrotalcite-supported gold catalyst
for the oxidant-free dehydrogenation of benzyl alcohol: Studies on support and
gold size effects. Chem. Eur. J. 17 (4): 1247–1256.
159 Wang, B., Lin, M., Ang, T.P. et al. (2012). Liquid phase aerobic oxidation of
benzyl alcohol over Pd and Rh catalysts on N-doped mesoporous carbon:
Effect of the surface acido-basicity. Catal. Commun. 25 96–101.
160 Zhao, Z., Zhang, J., Wang, Y. et al. (2020). Thermally stable Pd/reduced
graphene oxide aerogel catalysts for solvent-free oxidation of benzyl alcohol.
Chem. Phys. Lett. 746: 137306.
161 Wu, G., Wang, X., Guan, N. et al. (2013). Palladium on graphene as effi-
cient catalyst for solvent-free aerobic oxidation of aromatic alcohols: Role of
graphene support. Appl. Catal., B 136 177–185.
162 Islam, S.M., Roy, A.S., Dey, R.C. et al. (2014). Graphene based material as a
base catalyst for solvent free Aldol condensation and Knoevenagel reaction at
room temperature. J. Mol. Catal. A: Chem. 394 66–73.
163 Allen, M.J., Tung, V.C., and Kaner, R.B. (2010). Honeycomb carbon: a review of
graphene. Chem. Rev. 110 (1): 132–145.
164 Rostamnia, S., Doustkhah, E., Golchin-Hosseini, H. et al. (2016). Efficient
tandem aqueous room temperature oxidative amidations catalysed by supported
Pd nanoparticles on graphene oxide. Catal. Sci. Technol. 6 (12): 4124–4133.
165 Huang, H., Zhang, H., Ma, Z. et al. (2012). Tunable synthesis of metal-graphene
complex nanostructures and their catalytic ability for solvent-free cyclohexene
oxidation in air. Nanoscale 4 (16): 4964–4967.
166 Nejad, M.S., Behzadi, S., and Sheibani, H. (2019). Fabrication of ultra-small
ruthenium nanoparticles on porous modified reduced graphene oxide and its
application in solvent-free oxidation of cyclohexene with molecular oxygen.
Appl. Organomet. Chem. 33 (4): e4804.
References 225

167 Brahmayya, M., Dai, S.A., and Suen, S.Y. (2017). Sulfonated reduced
graphene oxide catalyzed cyclization of hydrazides and carbon dioxide to 1,
3, 4-oxadiazoles under sonication. Sci. Rep. 7 (1): 4675.
168 Zakeri, M., Abouzari-lotf, E., Miyake, M. et al. (2019). Phosphoric acid
functionalized graphene oxide: A highly dispersible carbon-based nanocata-
lyst for the green synthesis of bio-active pyrazoles. Arabian J. Chem. 12 (2):
188–197.
169 Şenocak, A., Khataee, A., Demirbas, E. et al. (2020). Ultrasensitive detection
of rutin antioxidant through a magnetic micro-mesoporous graphitized carbon
wrapped Co nanoarchitecture. Sens. Actuators, B 312: 127939.
170 Zhu, G., Zhang, F., Li, X. et al. (2019). Engineering the distribution of carbon in
silicon oxide nanospheres at the atomic level for highly stable anodes. Angew.
Chem. Int. Ed. 58 (20): 6669–6673.
171 Xu, H., Xu, H., Chen, Z. et al. (2019). Bimetallic PdCu nanocrystals
immobilized by nitrogen-containing ordered mesoporous carbon for electro-
catalytic denitrification. ACS Appl. Mater. Interfaces 11 (4): 3861–3868.
172 Song, H., Liu, Z., Gai, H. et al. (2019). Nitrogen-dopped ordered mesoporous
carbon anchored Pd nanoparticles for solvent free selective oxidation of benzyl
alcohol to benzaldehyde by using O2 . Front. Chem. 7 458.
173 Wang, Y., Zhang, J., Wang, X. et al. (2010). Boron-and fluorine-containing
mesoporous carbon nitride polymers: metal-free catalysts for cyclohexane
oxidation. Angew. Chem. Int. Ed. 49 (19): 3356–3359.
174 bin Saiman, M.I., Brett, G.L., Tiruvalam, R. et al. (2012). Involvement of
surface-bound radicals in the oxidation of toluene using supported Au-Pd
nanoparticles. Angew. Chem. Int. Ed. 51 (24): 5981–5985.
175 Brink, G.J., Arends, I.W., and Sheldon, R.A. (2000). Green, catalytic oxidation of
alcohols in water. Science 287 (5458): 1636–1639.
176 Zhang, P., Gong, Y., Li, H. et al. (2013). Solvent-free aerobic oxidation of
hydrocarbons and alcohols with Pd@N-doped carbon from glucose.
Nat. Commun. 4 (1): 1593.
177 Campbell, A.N. and Stahl, S.S. (2012). Overcoming the “oxidant problem”:
Strategies to use O2 as the oxidant in organometallic C–H oxidation reactions
catalyzed by Pd (and Cu). Acc. Chem. Res. 45 (6): 851–863.
178 Llabrés i Xamena, F.X., Casanova, O., Tailleur, R.G. et al. (2008). Metal organic
frameworks (MOFs) as catalysts: A combination of Cu2+ and Co2+ MOFs as an
efficient catalyst for tetralin oxidation. J. Catal. 255 (2): 220–227.
179 Kobayashi, S., Mori, Y., Fossey, J.S. et al. (2011). Catalytic enantioselective
formation of C−C bonds by addition to imines and hydrazones: A ten-year
update. Chem. Rev. 111 (4): 2626–2704.
180 He, R., Jin, X., Chen, H. et al. (2014). Mn-Catalyzed three-component reactions
of imines/nitriles, grignard reagents, and tetrahydrofuran: An expedient access
to 1,5-amino/keto alcohols. J. Am. Chem. Soc. 136 (18): 6558–6561.
181 Greb, L. and Lehn, J.M. (2014). Light-driven molecular motors: Imines
as four-step or two-step unidirectional rotors. J. Am. Chem. Soc. 136 (38):
13114–13117.
226 6 Carbon-Based Materials as Catalysts/Supports in Solvent-Free Organic Reactions

182 Largeron, M. and Fleury, M.B. (2013). Bioinspired oxidation catalysts. Science
339 (6115): 43–44.
183 Tayade, K.N. and Mishra, M. (2014). Catalytic activity of MCM-41 and Al
grafted MCM-41 for oxidative self and cross coupling of amines. J. Mol. Catal.
A: Chem. 382 114–125.
184 Furukawa, S., Suga, A., and Komatsu, T. (2014). Highly efficient aerobic
oxidation of various amines using Pd3Pb intermetallic compounds as catalysts.
Chem. Commun. 50 (25): 3277–3280.
185 Grirrane, A., Corma, A., and Garcia, H. (2009). Highly active and selective
gold catalysts for the aerobic oxidative condensation of benzylamines to imines
and one-pot, two-step synthesis of secondary benzylamines. J. Catal. 264 (2):
138–144.
186 Ho, H.A., Manna, K., and Sadow, A.D. (2012). Acceptorless photocatalytic
dehydrogenation for alcohol decarbonylation and imine synthesis. Angew.
Chem. Int. Ed. 51 (34): 8607–8610.
187 Largeron, M. and Fleury, M.B. (2012). A biologically inspired CuI/topaquinone-
like Co-catalytic system for the highly atom-economical aerobic oxidation of
primary amines to imines. Angew. Chem. Int. Ed. 51 (22): 5409–5412.
188 Kodama, S., Yoshida, J., Nomoto, A. et al. (2010). Direct conversion of
benzylamines to imines via atmospheric oxidation in the presence of
VO(Hhpic)2 catalyst. Tetrahedron Lett. 51 (18): 2450–2452.
189 Dhakshinamoorthy, A., Alvaro, M., and Garcia, H. (2010). Aerobic oxidation of
benzyl amines to benzyl imines catalyzed by metal–organic framework solids.
ChemCatChem 2 (11): 1438–1443.
190 Wu, S., Yu, L., Wen, G. et al. (2021). Recent progress of carbon-based metal-free
materials in thermal-driven catalysis. J. Energy Chem. 58 318–335.
191 Su, C., Acik, M., Takai, K. et al. (2012). Probing the catalytic activity of porous
graphene oxide and the origin of this behaviour. Nat. Commun. 3 (1): 1298.
192 Huang, H., Huang, J., Liu, Y.M. et al. (2012). Graphite oxide as an efficient and
durable metal-free catalyst for aerobic oxidative coupling of amines to imines.
Green Chem. 14 (4): 930–934.
193 Su, F., Mathew, S.C., Möhlmann, L. et al. (2011). Aerobic oxidative coupling of
amines by carbon nitride photocatalysis with visible light. Angew. Chem. Int.
Ed. 50 (3): 657–660.
194 Li, X.H. and Antonietti, M. (2013). Polycondensation of boron- and nitrogen-
codoped holey graphene monoliths from molecules: carbocatalysts for selective
oxidation. Angew. Chem. Int. Ed. 52 (17): 4572–4576.
195 Liang, C., Li, Z., and Dai, S. (2008). Mesoporous carbon materials: synthesis
and modification. Angew. Chem. Int. Ed. 47 (20): 3696–3717.
196 Zhang, J., Liu, X., Blume, R. et al. (2008). Surface-modified carbon nanotubes
catalyze oxidative dehydrogenation of n-butane. Science 322 (5898): 73–77.
197 Titirici, M.M., White, R.J., Brun, N. et al. (2015). Sustainable carbon materials.
Chem. Soc. Rev. 44 (1): 250–290.
References 227

198 Paraknowitsch, J.P. and Thomas, A. (2013). Doping carbons beyond nitrogen:
an overview of advanced heteroatom doped carbons with boron, sulphur and
phosphorus for energy applications. Energy Environ. Sci. 6 (10): 2839–2855.
199 Patel, M.A., Luo, F., Savaram, K. et al. (2017). P and S dual-doped graphitic
porous carbon for aerobic oxidation reactions: enhanced catalytic activity and
catalytic sites. Carbon 114 383–392.
200 Chen, B., Wang, L., Dai, W. et al. (2015). Metal-free and solvent-free oxida-
tive coupling of amines to imines with mesoporous carbon from macrocyclic
compounds. ACS Catal. 5 (5): 2788–2794.
229

Nitride-Based Nanostructures for Solvent-Free Catalysis


Fatemeh Majidi Arlan 1 , Ahmed A.M. El-Amir 2 , Amira A. El-Maddah 2 ,
Nagy L. Torad 3,4,5 , and Esmail Doustkhah 4,6
1
Iranian Academic Center for Education, Culture and Research, Research Department of Chemistry,
PO Box 57155-397, Urmia, Iran
2
Refractory & Ceramic Materials Division (RCMD), Central Metallurgical R&D Institute (CMRDI), P.O. Box 87
Helwan, 11421, Cairo, Egypt
3
Tanta University, Chemistry Department, Faculty of Science, 31527 Tanta, Egypt
4
International Center for Materials Nanoarchitectonics (WPI-MANA), National Institute for Materials Science
(NIMS), 1-1 Namiki, 305-0044 Tsukuba, Ibaraki, Japan
5
The University of Queensland, School of Chemical Engineering and Australian Institute for Bioengineering
and Nanotechnology (AIBN), 4072 Brisbane, QLD, Australia
6
Koç University, Tüpra¸s Energy Center (KUTEM), 34450 Istanbul, Turkey

7.1 Carbon Nitride

7.1.1 Introduction
In 1989, Liu and Cohen proposed an empirical model to define the bulk modulus of
covalent solids based on β-Si3 N4 structure; then, based on their studies, they claimed
the possible existence of a hypothetic material β-C3 N4 with potential excellent
mechanical properties [1, 2]. Then, during the past decade, efforts were devoted to
investigating carbon nitride (CN) materials and their polymorphs in a wide variety
of applications and developing the synthesis methods [3, 4], the construction of
heterojunctions [5, 6], and element doping [7, 8]. These have been utilized to
promote the light absorption and photocatalytic ability of carbon nitride through
providing a large specific surface area and proper band structure [9]. Among CN
derivatives, graphitic CN is popular because of its two-dimensionality, which can
trigger a large surface area after delamination. Several top-down methods have
emerged to delaminate the bulk graphitic CN, which relies on mechanical methods
[10, 11]. Top-down strategies are, most of the time, necessary with bottom-up
processes, as most bottom-up procedures produce the stacked form of the bulk
graphitic CN [12]. For the bottom-up approach, there are several precursors such
as melamine, cyanuric acid, and triazine [13]. Some common synthetic methods of
CNs are shown in Table 7.1.
CN is considered an environmentally nonmetal-friendly polymeric material
with superior semiconducting characteristics; high stability, practical catalytic
Solvent-Free Methods in Nanocatalysis: From Catalyst Design to Applications, First Edition.
Edited by Rafael Luque, Manoj B. Gawande, Esmail Doustkhah, and Anandarup Goswami.
© 2023 WILEY-VCH GmbH. Published 2023 by WILEY-VCH GmbH.
230 7 Nitride-Based Nanostructures for Solvent-Free Catalysis

Table 7.1 List of some common methods for synthesis of modification of CNs.

Synthesis method Precursors Product References

Direct synthesis (C3 N3 )(NH2 )3 + C3 N3 Cl3 C6 N9 H3 ⋅HCl [14]


Hexagonal graphitic
structure
Excellent crystallinity
Direct synthesis NH4 SCN + NH3 flow Low-defected CN [15]
nanocrystallites with
decreased sp2 N defects
bonding
Direct synthesis via Ball-milled β-C3 N4 nanorods [16]
mechanosynthesis graphite + NH3
Direct synthesis via Ball-milled Graphitic CN [17]
mechanosynthesis graphite + NH3 flow
Flux-assisted conversion C3 N4 H4 + NiMnCo α-C3 N4 + β-C3 N4 [18]
(catalyst)
Solvothermal synthesis (C3 N3 ) Nitrogen-rich graphitic [19]
(NH2 )3 + C3 N3 Cl3 + Ni CN hollow spheres
Synthesis under C3 N3 Cl3 + Li3 N in CN hollow spheres [20]
moderate conditions presence of porous
substrate

activities under visible light, unique physiochemical properties, nontoxic structure,


and facile synthesis procedure via cheap and abundant precursors are notable
beneficial features of these materials [1, 2]. Carbon nitrides can possess various
crystalline forms and structures, including graphitic CN (g-C3 N4 ), cubic-C3 N4 ,
defect zinc blende-C3 N4 , β-C3 N4 , and α-C3 N4 . Among those mentioned, g-C3 N4
with a bandgap of 2.7 eV is an important class of nonmetal-conjugated layered
material, analogous to graphene, with weak van der Waals forces between the
layers, leading to stacking of the layers [3, 4]. Despite the lack of hydrogen in the
chemical formula of g-C3 N4 , hydrogen appears on the edges and also in defective
areas, and as a result, its amount is not zero. There are two common types of g-C3 N4
varying in the chemical structure; one is condensed with a triazine structure and
heptazines condense the other one through the bridging tertiary amines in both
types [5]. It is worth noting that each type has its own various monomers, dimers,
and trimers. For example, the fully condensed polyheptazine is called tri-s-triazine,
and its dimer and monomer are called melon and melem, respectively. In the
case of polytriazine-based g-C3 N4 , the monomer is melamine, and the dimeric
form is melam, as depicted in Figure 7.1. Interestingly, this is not all the story,
and there is a considerable diversity in the chemical structures and formulas of
the g-C3 N4 ; diverse monomers can be co-condensed together to create a hybrid
form. A good example of this would be the co-condensation of polytriazine with
7.1 Carbon Nitride 231

(a) (b) (c)

Carbon
Nitrogen
Hydrogen

(d) (e) (f)

Figure 7.1 3D structures of carbon nitrides’ motifs: (a) melamine, (b) melam, (c) melem,
(d) melon, (e) fully condensed triazine-based C3 N4 structure (TGCN), and (f) fully condensed
polyheptazine (tri-s-triazine) C3 N4 structure. Source: Miller et al. [5]/Royal Society of
Chemistry/CC BY.

polyheptazine, which has produced a highly sensitive sensor toward formic acid
and other carboxylic acids [5, 6].
Diverse reactions such as CO2 activation [7–11], photocatalysis [12–20], and
synthesis of fine chemicals [21–24] can take the advantage of g-C3 N4 as a semi-
conductive catalyst. The underlying reason for the catalytic activity of g-C3 N4
is its similarity to metal oxide semiconductors from the viewpoint of bandgap
structure and the surface-rich Lewis basicity [6]. As g-C3 N4 has a metal-free compo-
sition, it has gained popularity over heavy metal-containing catalysts [6]. Therefore,
g-C3 N4 -based catalysts act as green and separable heterogeneous catalysts compared
with homogeneous organocatalysts such as ionic liquids or organic compounds [6].
The other merit of g-C3 N4 -based catalysts is their independent pathway to UV irra-
diation. As the bandgap of g-C3 N4 falls in the visible light range, the electron–hole
photogeneration can occur simply by solar light, which causes g-C3 N4 to be a much
greener photocatalyst [25]. Further improvements in the photocatalytic activity of
g-C3 N4 can be achieved through doping or supporting diverse heteroatoms such
as nonmetals like N and S or metals like Ag, Pd, or Pt [26, 27]. Utilizing g-C3 N4 as
a catalyst support because of added catalytic activity is of great interest in a wide
range of organic transformations such as hydroxylation of benzene [28], oxidation
of alkane [29], N-arylation of heteroaromatic compounds [30], reduction of nitrile
[31], oxidative amidation of aromatic aldehyde [32], Suzuki–Miyamura coupling
[33], oxidative of amine to imine [34], hydrogenation of quinoline [35], conversion
of CO2 into cyclic carbonate [36], and hydroxylation of aryl iodide [37].
232 7 Nitride-Based Nanostructures for Solvent-Free Catalysis

7.1.2 Synthesis of Carbon Nitride


Li et al. [38] have synthesized well-structured g-C3 N4 and mesoporous graphitic CN
(mpg-C3 N4 ) via a simple method. A g-C3 N4 nanosheet has been prepared through
a conventional thermal condensation reaction of cyanamide and mpg-C3 N4 via the
same method, with the introduction of diverse amounts of Ludox HS-40 solution
as a silica template. The gained nanoporous g-C3 N4 has been treated with NH4 HF2
after being calcined at 550 ∘ C to remove the silica template.
The structure of carbon nitride can also be modified and tuned by various
methods; a well-regulated and high-performance α/β-C3 N4 nanowire has been
recently fabricated [39]. They used an innovative hot melt reduction approach for
applying the structural control during the synthesis. For this control, they used
polyvinylchloride, ammonium chloride, and ferric oxide as cost-effective and easily
accessible precursors. Typically, α/β-C3 N4 has been synthesized in an autoclave
under sealed conditions by heating of the mixture of starting materials. The synthe-
sized nanowires exhibit good crystallinity with an average diameter of 10 nm and a
length range of 1–4 μm. It has been indicated that the surface structure of α/β-C3 N4
nanowires would facilitate further modification of the surface and functionalization
because of their high surface area and also rough surface with abundant exposed
atoms/prominences. Furthermore, the obtained band gap for α/β-C3 N4 nanowires
was found to be 4.38 eV, which had obvious ultraviolet luminescence properties at
the maximum emission peak of about 340 nm.
Crystalline CN phases have been synthesized using a facile solvent-free chemical
reaction method by Liu and coworkers [40] by mixing NH4 Cl with C3 N3 Cl3 in an
autoclave under a nitrogen atmosphere. The dominant phase in the synthesized
product is polycrystalline graphitic carbon nitride with a piece-like morphology
along with a probable α-C3 N4 phase. Graphite or onion-like lamellar-structured
carbon nitrides have been synthesized for the first time by Guo et al. [41] through the
autoclave method under a nitrogen atmosphere by heating a mixture constituting
C3 N3 Cl3 , NaNH2 , K, or NaN3 . g-C3 N4 with ultra-thin porous layers of carbon nitrides
has been synthesized by Devthade et al. [42], through facile and straightforward
routes. g-C3 N4 has been prepared by the conventional thermal condensation reac-
tion of melamine. However, mesoporous g-C3 N4 can be synthesized from melamine
as raw material in the presence of HCl and tetraethyl orthosilicate as a silica tem-
plate. The mixture has been heated to afford the desired carbon nitride; then, the
silica template is removed through NH4 HF2 treatment to give mesoporous g-C3 N4 .

7.1.3 Modification of Carbon Nitrides


Various methods widely report modification of g-C3 N4 . The modifications can
occur during the synthesis of g-C3 N4 or after its synthesis. When the modification
occurs through co-condensation of the g-C3 N4 ’s monomers, a drastic change occurs
in the intrinsic structure and properties of the modified g-C3 N4 . For example,
a solvent-free thermal condensation process for synthesizing polyimide-based
g-C3 N4 is reported from easily accessible melem and diverse dianhydrides [43].
7.1 Carbon Nitride 233

Figure 7.2 Synthesis of polyimides: M-FDA, M-NTDA, and M-ODPA from melem and
dianhydrides (FDA, NTDA, and ODPA).

This method first obtains melem from melamine in a ceramic crucible under a
semiclosed atmosphere at 200 ∘ C, followed by heating at 425 ∘ C. They utilize melem
to co-condense it with polyimide monomer, as shown in Figure 7.2. By copolymer-
ization of melem with three different dianhydrides, such as 1,4,5,8-naphthalene
tetra-carboxylic dianhydride (NTDA), 4,4′ -oxydiphthalic anhydride (ODPA), and
2,2′ -bis(3,4-dicarboxyphenyl)hexafluoropropanedianhydride (FDA), three other
structures have been obtained [44]. The solvent-free synthesis method of this mod-
ified g-C3 N4 includes the solvent-free grinding of the melem and dianhydride and
further heating at 350 ∘ C under an N2 atmosphere. The resulting polyimide-based
g-C3 N4 , named M-FDA (white powder, 87%), M-ODPA (light brown powder, 89%),
and M-NTDA (dark brown powder, 84% in yield), where M represents the melem.
The g-C3 N4 can be modified by loading catalytically active metals into it. For
instance, Au anchored on g-C3 N4 nanosheets has been reported by impregnating a
dilute chloroauric acid aqueous solution into g-C3 N4 nanosheets at room temper-
ature under dark conditions, which is subsequently reduced by a H2 /Ar (5%) gas
mixture in a tube furnace [45]. Then, the atomically dispersed gold catalyst (labeled
as Au1 /g-C3 N4 ) is produced with metal loading at 110 ∘ C. The gold nanoparticles
on g-C3 N4 are around 3–5 nm [45]. Modification of g-C3 N4 can also be achieved by
supramolecules, which is another trending approach to trigger the intermolecular
interaction potential of g-C3 N4 . In this regard, metallophthalocyanine is used to
support g-C3 N4 to deliver a hybrid [Mpc/g-C3 N4 (M = Co, Cu)] via direct annealing
of a mixture of dicyandiamide and Mpc under a flowing nitrogen atmosphere [46].
In this method, cobalt phthalocyanine or copper phthalocyanine have been used to
obtain the modified g-C3 N4 . It is worth noting that in the modified C3 N4 , the Mpc
moiety can act as a mild Lewis acid site whereas the g-C3 N4 species can function
as a solid base and behave as cocatalyst because of the appearance of amine groups
at the edges of the graphitic sheets. Availability, air and moisture tolerance, and
recyclability are noticeable features of synthesized catalysts [46].
234 7 Nitride-Based Nanostructures for Solvent-Free Catalysis

Doping is another promising technique to modify the performance of g-C3 N4 .


The dopant can occur in the main 2D structure of g-C3 N4 either in carbon’s site or
in nitrogen’s site. Boron doping is the most popular method to modify and tune the
catalytic activity of the g-C3 N4 because of the trivalent capability of the B. Thus,
it adopts an identical 2D orientation when placed in the g-C3 N4 [6]. Increased
surface acidity and surface basicity are beneficial aspects of doping B atoms into
the CN framework. This doping can also be achieved under solvent-free conditions.
Co-doping CN has also been a proper method for catalytic activity boosting. The
synthesis method is usually used in the solvent-free process at high temperatures
(i.e. 500–600 ∘ C). Cu is one example that is codoped with boron in CN [47]. To pro-
duce codoped Cu-boron-doped carbon nitride (BCN) by this method, appropriate
amounts of CpmimBF4 , urea, and Cu(NO3 )2 ⋅3H2 O have been annealed at 550 ∘ C.
Karkeabadi et al. established the synthesis of graphite-like carbon nitride
and modified g-C3 N4 /Cu2 O through conventional thermal polycondensation of
urea [48]. Initially, the g-C3 N4 can be prepared through the previously reported
thermal polycondensation process [49], in which urea is heated at 525 ∘ C for
two hours under a semiclosed atmosphere in a crucible covered with a lid. The
obtained yellow g-C3 N4 is then added to a solution of CuCl2 ⋅2H2 O at pH = 13.
The immediate addition of an appropriate amount of NH2 OH⋅HCl solution gave
the desired g-C3 N4 /Cu2 O composite [50]. A simple preparation of g-C3 N4 /Cu2 O
has been reported by immobilizing CuO2 on the surface of g-C3 N4 through reduc-
tion of CuCl2 ⋅2H2 O under alkaline conditions in the presence of hydroxylamine
hydrochloride.
Priya et al. synthesized heterostructure g-C3 N4 /CuO nanocomposites through
a mechanochemical process [51]. g-C3 N4 can be prepared by directly heating
melamine as an easily accessible precursor under atmospheric air conditions.
Typically, melamine was heated at 500 ∘ C for two hours and dried at room temper-
ature. The CuO NPs have been prepared through the previously reported method.
Then, diverse amounts of synthesized g-C3 N4 have been mixed with CuO NPs
under solvent-free mechanochemical conditions till the appearance of a navy-blue
color. The obtained product is heated for two hours at 550 ∘ C and cooled to room
temperature to give the g-C3 N4 /CuO.

7.1.4 Solvent-Free Catalysis with Carbon Nitrides


Sharma and Sasson [25] have investigated the synthesis of Ru-doped g-C3 N4 through
conventional methods with some modifications. Then, the efficiency of synthesized
nanosheets as photocatalysts has been evaluated via an environmentally friendly
solvent-free H-transfer reaction of the carbonyl group in the presence of alcohol as
an H source under visible light irradiation. It has been found that the holes and elec-
trons are able to oxidize the absorbed alcohol while reducing the absorbed aldehyde
on the surface of CN. The reduction of carbonyl groups by alcohol as H contributor
is depicted in Figure 7.3. From a catalytical point of view, it should be mentioned
that CN has gained π-conjugated system because of the presence of nitrogen
atoms and can act as a semiconductor. Also, doped ruthenium atoms might form
7.1 Carbon Nitride 235

Figure 7.3 Solvent-free photocatalytic H-transfer reaction.

a heterojunction between the g-C3 N4 π-electrons, which promotes photocatalytic


activity of Ru-g-C3 N4 . Solvent-free and mild reaction conditions, recyclable catalyst,
and high yields are some of the notable features of the mentioned methodology [25].
Sulfonated CN (Sg-CN) has been used as an organosulfonated heterogeneous acid
catalyst in the manufacture of biodiesel [52–54], where its preparation process con-
sists of modest calcination of urea [34] and a subsequent sulfonation step. Qareaghaj
et al. [55] have utilized environmentally benign for the synthesis of xanthene deriva-
tives, applying Sg-CN as a high performance and green catalyst under neat condi-
tions. The reaction has been carried out under neat conditions via mechanochemical
ball-milling of diverse aldehydes, β-naphthol, and 1,3-indandion with catalytic load-
ing of Sg-CN to yield the xanthene derivatives via the Knoevenagel and Michael
reactions (Figure 7.4). Workup simplicity, green conditions, and high yields are ben-
eficial aspects of the developed method.
The hydrosilylation of alkynes is a powerful process for producing vinylsilane
compounds, which are synthetically versatile organosilicon reagents in organic
chemistry. Such a reaction was achieved under solvent-free conditions [45], using
the prepared catalysts Au1 /g-C3 N4 and AuNPs/g-C3 N4 where phenylacetylene

Figure 7.4 One-pot synthesis of xanthene derivatives through solvent-free reaction using
Sg-CN as catalyst.
236 7 Nitride-Based Nanostructures for Solvent-Free Catalysis

Figure 7.5 Hydrosilylation of alkynes affording three possible isomeric vinylsilane


compounds. Source: Adapted from Sun et al. [40].

R = cyclohexane, CH2CI, CH3, C6H5

Figure 7.6 Mpc/g-C3 N4 -promoted chemical fixation of CO2 to cyclic carbonate.

and triethylsilane have been used as substrates for the hydrosilylation reaction,
as shown in Figure 7.5. It has been indicated that upon using Au1 /g-C3 N4 , the
conversion of phenylacetylene has reached 78% after 12 hours of the reaction. In
addition, the conversion has been increased by increasing the reaction temper-
ature; however, the selectivity has decreased. On the other hand, by using Au
NPs/g-C3 N4 as a catalyst, the conversion has reached only 6%, which has been
explained by the active sites that are mainly derived from the Au–N coordination on
Au nanoparticles.
Zhang et al. [46] synthesized Mpc-CN hybrids, which are utilized as bifunctional
nucleophile–electrophile catalyst for the chemical fixation of CO2 to cyclic carbon-
ates (Figure 7.6). The reaction has been carried out in an autoclave under solvent-free
conditions between diverse epoxides and CO2 with the role-playing of Mpc/g-C3 N4
as a highly efficient catalyst. The cycloaddition reaction between epichlorohydrin
and CO2 gave the 3-chloro-1,2-propylenecarbonate in high yields of 97.6% with full
conversion of ECH. The high efficiency of the catalyst is rooted in both CO2 and
ECH activation simultaneously; CO2 activation is done through nucleophilic attack
because of the g-C3 N4 moiety, which is rich in primary and secondary amine groups
at the edges of graphitic sheets, and ECH activation is done via electrophilic attack
because of the presence of Mpc species, which act as Lewis acid.
Zhu et al. [6] have developed diverse graphitic CN-based materials and have
assessed the efficiency of prepared catalysts via a green and sustainable solvent-free
cycloaddition reaction of CO2 and styrene oxide (Figure 7.7). BCN, supported on

Figure 7.7 Cycloaddition of CO2 and styrene oxide to produce styrene carbonate.
7.1 Carbon Nitride 237

Figure 7.8 g-C3 N4 -NaOH-utilized CO2 cycloaddition reaction to epoxide.

mesoporous silica SBA-15, exhibited excellent catalytic activity with 95% conversion
and selectivity for cycloaddition reaction of CO2 and styrene oxide to afford the
desired styrene carbonate under neat conditions. The high catalytic efficiency of
BCN/SBA-15 is based on acid–base duality induced via B doping, which activates
both CO2 and epoxide.
Xu et al. [56] have synthesized a series of g-C3 N4 materials and then have assessed
the efficiency of diverse as-synthesized catalysts via green solvent-free cycloaddi-
tion of CO2 to diverse epoxides. The reaction has been carried out in an autoclave
among diverse epoxides and CO2 with the introduction of ZnI2 and catalytic loading
of g-C3 N4 at 140 ∘ C (Figure 7.8). The synthesized g-C3 N4 -NaOH and g-C3 N4 -KOH
exhibited high and stable catalytic activity with 83–92% conversion under optimized
reaction conditions, and the catalysts are reusable for at least five times with no loss
of efficiency. It is worth noting that the authors have reported a much easier method
for the synthesis of CNs as promising catalysts in CO2 cycloaddition reactions.
From both academic and industrial viewpoints, the Sonogashira–Hagihara
cross-coupling reaction is considered as one of the most useful reactions in organic
transformations, as it provides a strong tool for C(sp)–C(sp2 ) construction via
the reaction of aryl halides with terminal acetylenes. In this regard, Ghodsinia
et al. [57] have developed a green and facile pathway for the construction of
cross-coupled products through the solvent-free reaction of diverse aryl halides
and terminal acetylenes in the presence of Co3 O4 /TSCN as catalyst at 70 ∘ C
(Figure 7.9). Co3 O4 /TSCN has been synthesized through the hard-templating
method and impregnation process and exhibited excellent catalytic activity in
the Sonogashiro–Hagihara reaction. The results revealed that aryl halides with
both electron-withdrawing and electron-rich groups are well-tolerated; however,
aryl halides with electron-withdrawing groups are coupled more quickly than

Figure 7.9 The Sonogashira–Hagihara cross-coupling reaction in the presence of the


Co3 O4 /TSCN (III).
238 7 Nitride-Based Nanostructures for Solvent-Free Catalysis

electron-rich aryl halides and resulted in higher yields. Indeed, the cross-coupling
reaction of aliphatic terminal alkynes has also been investigated, and the results
revealed that aliphatic alkynes could couple with aryl halides in the presence of
Co3 O4 /TSCN, and the desired products can be achieved in moderate yields.
Cu2 O-modified g-C3 N4 has been utilized as a superior catalyst for coupling reac-
tion between aryl aldehydes, alkynes, and secondary amines [48]. The reaction is
carried out at 110 ∘ C under neat conditions with a catalytic loading of g-C3 N4 /Cu2 O
(Figure 7.10). The corresponding propargylamines were achieved via C–H activa-
tion of terminal alkyne by a catalyst that reacted with an in situ produced iminium
ion from the coupling reaction of amine and aldehyde. Reusable heterogeneous cat-
alysts, high yields, and short reaction times are notable merits of the developed
methodology.
Priya et al. [51] developed a one-pot three-component synthesis of pyrimidoin-
dazole derivatives under microwave conditions in the presence of g-C3 N4 /CuO
as an efficient CN-based catalyst. The reaction of diverse aldehydes, alkynes, and
amines is carried out under neat microwave conditions at 100 W at 70 ∘ C with
little catalyst loading to afford the corresponding pyrimidoindazole derivatives
(Figure 7.11). After optimizing the reaction conditions, diverse aldehydes, including
electron-withdrawing and electron-releasing groups, have been investigated. The
results revealed that aldehydes with electron-releasing groups reacted much more
quickly and gave higher yields.

Figure 7.10 g-C3 N4 /Cu2 O-utilized synthesis of propargylamines.

Figure 7.11 One-pot three-component synthesis of pyrimidoindazoles applying


g-C3 N4 /CuO under MW irradiation.
7.1 Carbon Nitride 239

Figure 7.12 p-CNN photocatalytic synthesis of DHPMs under solvent-free conditions.

Li et al. [58] established porous ultrathin carbon nitride nanosheets (p-CNNs) via
a facile and straightforward method. Then, they assessed the photocatalytic activity
of synthesized p-CNNs in a one-pot, solvent-free synthesis of dihydropyrimidine
derivatives. The condensation reaction of 1,3-dicarbonyl compounds, aldehydes,
and urea or thiourea with catalytic loading of p-CNNs under visible light irradiation
at 90 ∘ C gave the corresponding dihydropyrimidines in high yields (Figure 7.12).
Eco-friendliness, high yields, neat condition, and simple workup procedure are
worth-noting advantages of the mentioned method.
In another attempt, 12-phenyl-9,9-dimethyl-8,9,10,12-tetrahydrobenzo[a] l.l
xanthen-11-one and 5-phenyl-1(4-methoxyphenyl)-3[(4-methoxyphenyl)-amino]-
1H-pyrrol-2(5H)-one were synthesized. The reaction of 2-naphthol, benzaldehyde,
and trimethadione, at 40 ∘ C under visible light irradiation in the presence of 1.1 wt%
p-CNNs yielded corresponding 12-phenyl-9,9-dimethyl-8,9,10,12-tetrahydrobenzo
[a]xanthen-11-one in 97% yield (Figure 7.13), and the reaction of the prepara-
tion of p-methoxyaniline with benzaldehyde, and ethyl pyruvate with 2.5 wt%
p-CNNs at room temperature under visible light afforded the corresponding
5-phenyl-1(4-methoxyphenyl)-3[(4-methoxyphenyl)-amino]-1H-pyrrol-2(5H)-one
in 93% yield (Figure 7.14). The developed method exhibits various economic and
environmental merits, such as excellent yields and eliminating the use of solvents
and metal-based photocatalysts.
Wang et al. [59] reported on the synthesis of hydrolyzed carbon nitride (HCN).
They evaluated this material in a CO2 cycloaddition reaction under solvent-free and
autoclave conditions (Figure 7.15). This reaction is carried out at 110 ∘ C and under
2 MPa of CO2 for six hours. Here, the cycloaddition of CO2 has been conducted with

Figure 7.13 p-CNN photocatalytic reaction for the synthesis of 12-phenyl-9,9-dimethyl-


8,9,10,12-tetrahydrobenzo[a]xanthen-11-one under solvent-free conditions.
240 7 Nitride-Based Nanostructures for Solvent-Free Catalysis

Figure 7.14 Photocatalytic p-CNNs for the synthesis of 5-phenyl-1(4-methoxyphenyl)-


3[(4-methoxyphenyl)-amino]-1H-pyrrol-2(5H)-one under solvent-free conditions.

Figure 7.15 Cycloaddition reactions of various epoxides with CO2 catalyzed by


HCN-FD/Bu4 NI.

and without a cocatalyst. When there is no cocatalyst in the reaction, a trace amount
of the product is yielded. On the other hand, the addition of Bu4 NI as a cocatalyst to
the freeze–drying hydrolyzed carbon nitride (HCN-FD) catalyst has resulted in the
highest yield (76%) of cycloaddition of CO2 with propylene oxide (PO).
A novel and heterogeneous carbon nitride-based organocatalyst has been devel-
oped [60]. The catalytic potential of prepared 2D materials has been explored via
the green and neat preparation of substituted quinoxalines through a condensa-
tion reaction under solvent-free conditions. The reaction’s starting compounds
include the 1,2-diketones and aryl 1,2-diamines at 100 ∘ C, yielding quinoxalines
in high yields within a short reaction time (Figure 7.16). Simplicity in catalyst
preparation, obtaining high yields in the reaction product in a short reaction
time, reusability, and thermal stability of the catalyst are advantages of this
solvent-free method.
β-Hydroxy nitriles are considered essential building blocks in industrial organic
transformations; hence, exploring a cost-effective and scalable synthetic approach
for the synthesis of β-hydroxyl nitriles is of great importance. In this regard,
Saadat et al. [61] established an innovative solvent-free method for the syn-
thesis of β-hydroxy nitriles through epoxide ring-opening in the presence of
Fe3 O4 @g-C3 N4 -K as a green and biocompatible catalyst. The rapid and regioselec-
tive ring-opening of diverse epoxides with TMSCN under solvent-free conditions
at 70 ∘ C gave the substituted hydroxyl nitriles high yields (Figure 7.17). Excellent
scalability, short reaction time, and workup simplicity are the significant highlights
of the mentioned process.
Selective oxidation of saturated hydrocarbons via molecular oxygen is of
significant importance in catalysis; hence, the development of high-performance
catalysts has always been in great demand. Regarding that, Zhan and
7.1 Carbon Nitride 241

Figure 7.16 Synthesis of quinoxaline derivatives applying CN-Pr-VB1 as catalyst.

Figure 7.17 Ring-opening of various epoxides promoted by magnetic support catalyst.

Figure 7.18 Selective oxidation of cyclohexane to cyclohexanol via Co-g-C3 N4 .

coworkers [62] established a series of codoped g-C3 N4 , and the potential of


the synthesized catalyst is explored in the selective oxidation of cyclohexane
(Figure 7.18). t-Butyl hydroperoxide (TBHP) as an initiator is added to the cyclohex-
ane in an autoclave to produce cyclohexanol via molecular oxygen and a catalyst.
An excessive amount of PPh3 is added to ultimately convert the produced cyclohexyl
hydroperoxide into cyclohexanol. The catalyst is recoverable for at least five runs
without any significant loss in activity. The results revealed that the Co content of
the catalyst plays a vital role in the oxidation process, and the highest conversion
of 11.2% is obtained by applying a catalyst containing 11.9% Co content. Co-g-C3 N4
242 7 Nitride-Based Nanostructures for Solvent-Free Catalysis

Figure 7.19 Ring-opening reaction of diverse epoxides with arylamines in the presence of
Fe3 O4 -g-C3 N4 .

exhibited good catalytic activity in the selective oxidation process of C–H (sp3 )
to alcohol, and the results can be extended to produce alcohols and ketones via
molecular oxygen.
Novel magnetic g-C3 N4 has been developed by Ahooie et al. [63], and the catalytic
efficiency of the prepared nanocomposites is explored through the aminolysis pro-
cess under the solvent-free condition at 60 ∘ C. Ring-opening of diverse aryl and alkyl
epoxides by aromatic amines can be carried out in the presence of Fe3 O4 -g-C3 N4 as
a heterogeneous catalyst (Figure 7.19). The substituted 𝛽-aminoalcohols achieved
good-to-excellent yields within a short reaction time. The proposed mechanism for
activating the epoxide ring might be associated with the synergistic contribution of
Fe3 O4 and g-C3 N4 sheets, based on the hydrogen bonds between N–H and epoxide.
The notable features of the developed methodology are a green and eco-friendly cat-
alyst, economically viable reagents, high conversions within a short reaction time,
and a simple workup procedure.
Xu et al. [64] synthesized mesoporous CN (mp-C3 N4 ) grafted with an
n-bromobutane sheet. The catalytic performance has been examined by the
cycloaddition of CO2 with propylene oxide to produce corresponding propylene
carbonate under solvent-free conditions at 140 ∘ C. The reaction resulted in the
desired product in a high yield (Figure 7.20). The authors also explored the reaction
using diverse mp-C3 N4 grafted with other alkyl halides instead of n-bromobutane,
exhibiting high catalytic performance. Amine species at the edges of graphitic
sheets are catalytically active sites that enhance CO2 adsorption while activating
propylene oxide molecules.
Sulfonated mesoporous C3 N4 (omp-g-C3 N4 /SO3 H) has been synthesized [65] via
a facile and straightforward method using SBA-15 as a hard template. Then, the
catalytic potential of prepared omp-g-C3 N4 /SO3 H was evaluated via multicompo-
nent reaction of diverse aldehydes, 2-naphthol, and urea. The reaction is carried
out under solvent-free conditions at 100 ∘ C with small catalyst loading, and corre-
sponding 1,2-dihydro-1-arylnaphtho[1,2-e][1,3]oxazin-3-one derivatives have been
achieved in high yields (Figure 7.21). The merits of the reported strategy are the

Figure 7.20 n-butBr/mp-C3 N4 utilized synthesis of propylene carbonate.


7.1 Carbon Nitride 243

Figure 7.21 ompg-C3 N4 /SO3 H-catalyzed reaction for the synthesis of


1,2-dihydro-1-arylnaphtho[1,2-e][1,3]oxazin-3-ones.

Figure 7.22 Synthesis of 2,3-dihydroquinazolin-4(1H)-one derivatives catalyzed by


ompg-C3 N4 /SO3 H.

easy workup procedure, short reaction time, increased conversion, environmentally


benign, and recoverable nonmetal catalyst.
In a similar study reports the well-structured sulfonated mesoporous g-C3 N4
and then the potential catalytic activity has been explored through the con-
densation reaction of diverse carbonyl compounds with anthranilamide. The
condensation reaction of various aldehydes or ketones with 2-aminobenzamide
under the neat condition at 120 ∘ C in the presence of omp-g-C3 N4 /SO3 H as a
heterogeneous catalyst gained the substituted 2,3-dihydroquiazoline-ones in high
yields (Figure 7.22). After the optimization reaction, the scope of the reaction
can be extended by using diverse substituted aromatic aldehydes and ketones
with either electron-donating or electron-withdrawing groups. The corresponding
products can be obtained in good-to-excellent yields. The developed method
highlights the broad substrate scope, mild and green reaction conditions, and short
reaction time.
Lan et al. [67] have synthesized a series of phosphorous graphitic carbon nitrides
(P-g-C3 N4 ) via simple and facile thermolysis of melamine and hexachlorocyclot-
riphosphazene (HCCP). They examined the catalytic efficiency of the P-g-C3 N4 in
the solvent-free synthesis of cyclic carbonate. The cycloaddition of CO2 to diverse
epoxides under solvent-free conditions at 100 ∘ C via P-g-C3 N4 as a catalyst is effi-
cient (Figure 7.23). The P-g-C3 N4 composites with different P content and different

Figure 7.23 Phosphorous g-C3 N4 catalyzed the synthesis of cyclic carbonates.


244 7 Nitride-Based Nanostructures for Solvent-Free Catalysis

Figure 7.24 Synthesis of dihydropyrimidine derivatives through g-C3 N4 @SO3 Ch-catalyzed


process.

cocatalysts under the examination exhibit different catalytic activities. Phosphorous


content plays as an active acid site that is attributed to the ring-opening process
of epoxide molecules. An increase in P content improved the reaction yield; using
Bu4 NBr as a cocatalyst significantly increased the yield of the process.
Azizi and Edrisi [68] have reported a novel and green choline sulfate ionic
liquid-supported g-C3 N4 (g-C3 N4 @SO3 Ch) via a facile method. Then, the catalyst
was evaluated in a solvent-free Biginelli reaction to synthesize N-containing
heterocyclic scaffolds. Under solvent-free conditions at 80 ∘ C, the reaction includes
urea, 𝛽-ketoesteres, and aldehydes, with the catalysis of g-C3 N4 @SO3 Ch yielding
the corresponding heterocycles in high yields (Figure 7.24). Incorporating ionic
liquid into CN provides an attractive strategy to enhance the surface area, charge
separation efficiency, and overall catalytic activity. It is worth mentioning that the
developed catalyst has shown excellent catalytic performance at room temperature
for the construction of various biologically active heterocyclic compounds with
convenient and excellent reusability without any loss in mass and activity.
Li et al. [38] have synthesized g-C3 N4 and mpg-C3 N4 as catalysts in the oxidation of
toluene to benzaldehyde under solvent-free conditions in an autoclave under stirring
and pressurized with 10 bar O2 . The conversion obtained by using 100 g of mpg-C3 N4
is 26% after 16 hours of the reaction at 185 ∘ C. The nanostructure tuning of g-C3 N4
has impacted a high selectivity by turning the homogeneous oxidation into hetero-
geneous oxidation and capturing all free ⋅O2 − radicals to effectively suppress the
overoxidation of aldehydes. The multicomponent Biginelli reaction has been stud-
ied by Devthade et al. [42] to evaluate the photocatalytic activity of mgp-C3 N4 . The
reaction is tested with a model reaction of benzaldehyde, acetylacetone, and urea
at different temperatures, times, and doses of catalyst, as given in Figure 7.25. The
investigation into Biginelli condensation of active methylene compounds, aryl alde-
hydes, and urea has been catalyzed by mpg-C3 N4 under visible light irradiation at

Figure 7.25 Photocatalyzed multicomponent Biginelli reaction. Source: Devthade et al.


[42]/John Wiley & Sons.
7.2 Boron Nitride 245

ambient temperature. The highest yield obtained upon using 20 mg of mpg-C3 N4


catalyst is 84% under the light conditions with 1 mmol of methyl substrate and at
70 ∘ C for 120 minutes. This method has offered an excellent yield of pure products
within a short reaction time and the stability and recyclability of the heterogeneously
dispersed mpg-C3 N4 catalyst with consistent catalytic activity under visible light
irradiation.

7.2 Boron Nitride

7.2.1 Introduction
The structure of boron nitride (BN) involves both the bulk form and the nanoform.
The bulk BN form has two allotropes of hexagonal boron nitride (h-BN) and
cubic boron nitride (c-BN), with orbital hybridizations of sp2 and sp3 , respectively,
that are the most stable. The fabrication of a nanomorphology from h-BN can be
achieved from the bulk form thanks to the weak interlayer forces of the layered
h-BN. Figure 7.26 exhibits the crystal structure of hexagonal-layered BN.
BN nanomaterials have excellent chemical stability against different chemical
reagents, which has increased their demand for many applications that involve
corrosion-resistant films [70, 71], high-temperature manufacturing [72], and
oxidation resistance layers [73]. The type and strength of present σ bonds and
the lack of surface states and dangling bonds can explain the negligible chemical
reactivity of BN nanomaterials. Boron nitride compounds are insoluble in common
weak and strong acids such as HCl, HBr, H2 SO4 , HNO3 , acetic acid (CH3 COOH),
and hydrofluoric acid (HF). However, highly oxidizing acids such as HNO3 have
been used in the surface/edge oxidation of BN through chemical reactions as boron

a
c

Figure 7.26 Crystal structure of hexagonal-layered BN adapted from [69] projection


vector. The cell parameters are obtained from computed cell parameters; Materials Project
websites (mp-7991).
246 7 Nitride-Based Nanostructures for Solvent-Free Catalysis

sites on the surface act as Lewis acids. Therefore, electron pairs in central atoms
can be taken by Lewis acid–base interactions [74].
Furthermore, chemical etching of edges and chemical substitution of h-BN mate-
rial can occur by alkali and alkaline earth fluorides at elevated temperatures and
pressure [75]. The high chemical inertness of h-BN has been shown to make them
reliable for corrosion resistance films. On the edge of mechanical performance,
many factors have been affecting the mechanical performance of boron nitride
nanosheets (BNNS), such as temperature, defect levels, and chirality [76–78]. By
elevating the temperature from 0 to 2000 K, the young’s modulus decreases from
0.96 to 0.68 TPa, and it becomes more extreme at zigzag edges than at armchair
ones [79]. This can result from enhancing atomic vibrations and boosting the rate
of viable plastic deformation at higher temperatures. However, local anisotropic
rims are the main reason for the different mechanical performances of edge
configurations, where there can be various existing forces (tensile or compressive)
with different distributions [80, 81].
BNNSs with a hexagonal structure are a structural and isoelectronic analog of
graphene, where they possess high thermal conductivity and mechanical strength
[82–88]. They are differentiated from graphene by their high chemical inertness, ulti-
mate electrical insulation (bandgap ∼6 eV), high oxidation resistance, and large sur-
face area [82, 88, 89]. Due to the unique characteristics of BNNSs, they can be utilized
in various technological applications such as catalyst supports [90, 91], polymeric
nanocomposites [92–94], electronic devices [95, 96], and lithium batteries [97].

7.2.2 Synthesis and Modification of Boron Nitride


BNNSs have recently gained theoretical and experimental interest with many
potential synthesis methods. There are two methods for preparing BNNSs: chemical
vapor deposition (CVD), growing from various B and N precursors [98–101]. The
second method is the mechanical cleavage or solvent exfoliation of bulk h-BN
[102–106]. A synthetic procedure developed by Lei et al. [107] has achieved a
low-cost, facile, solvent-free, and high-yield synthesis of BNNSs, where a mechani-
cal solid-state exfoliation has been used. The synthesis process involves the milling
of BN powders and NH4 Cl powders under a nitrogen atmosphere. The milled
powder then undergoes calcination under nitrogen/hydrogen flow. The obtained
BNNSs from this solvent-free procedure have exhibited high surface area and
excellent recycling stability, in addition to outstanding adsorption capacities for
proteins. Hence, there is high potential for the as-synthesized BNNSs to be utilized
as adsorbents for water treatment and purification.
Sun et al. [108] discovered that BNNSs can be obtained by exfoliating bulk h-BN
powder in thionyl chloride without using any dispersion agents. The reported
procedure involves the addition of h-BN powder to SOCl2 in order to get large
pellets of BNNSs yielding 20%. The as-synthesized BNNSs have been then used to
prepare the Pd/BNNS catalyst through the deposition–precipitation method, which
has been achieved by dispersing BNNSs in isopropanol through mild sonication.
It is worth mentioning that the authors have used the same procedure to obtain
Pd/h-BN. The activity and recyclability of the as-synthesized Pd/BNNS catalyst in
7.3 Molybdenum Nitride 247

the hydrogenation of nitroaromatics outperformed the commercial Pd/C and many


reported catalysts. So far, no reports about the solvent-free catalysis of BNs, to our
knowledge, is not reported.

7.3 Molybdenum Nitride

7.3.1 Introduction
Molybdenum nitrides (MoNs) also belong to the remarkable class of metal nitrides
with unique physicochemical properties. MoNs generally form several crystal struc-
tures, such as hexagonal and triclinic structures (Figure 7.27) [109, 110].

7.3.2 Synthesis of Molybdenum Nitride


Generally, synthetic procedures used in MoN preparation are physical and chem-
ical. Physical methods include laser ablation, arc discharge and evaporation,
plasmon-enhanced chemical vapor deposition (PECVD), and pulsed laser deposi-
tion [111]. Despite some drawbacks such as side product formation and shapeless
and polydisperse nanomaterials, the most common and popular chemical syn-
thesis of MoN is direct nitridation of elements [69], nitridation of metal oxides
[112, 113], ammonolysis of metal chlorides [114], solvothermal synthesis [115],

b b b

c a [001] Projection vector c a [100] Projection vector a [111] Projection vector


c
(a)

b b
b

c a [001] Projection vector c [111] Projection vector c a [100] Projection vector


(b) a

Figure 7.27 (a) Crystal structure of hexagonal MoN from different projection vectors. The
unit-cell parameters are based on a computed cell reported on the Materials Project
website (mp-1078389). (b) Crystal structure of triclinic molybdenum nitride from different
projection vectors. The unit-cell parameters are based on a computed cell reported on the
Materials Project website (mp-1080195).
248 7 Nitride-Based Nanostructures for Solvent-Free Catalysis

Table 7.2 Comparing PECVD versus thermal methods for MoN synthesis.

Method Type Reagents Condition References

Plasma-enhanced δ1 -MoN MoCl5 + nitrogen plasma 500–650 ∘ C [121]


chemical vapor
deposition
Plasma-enhanced δ2 -MoN MoCl5 + nitrogen 730 ∘ C [121]
chemical vapor
deposition
Thermal δ3 -MoN Mo metal + ammonia 700 ∘ C [122]

and supercritical processes [116]. It should be mentioned that the nature of the
precursor is of great importance in determining the final structure of MoN. The
use of urea as a nitrogen source for the synthesis of bulk MoN has been known
since 1994 [117]. However, further treatment and modified reaction conditions
achieved MoN nanoparticles in 2003 [118]. However, metal–urea precursors are
still popular for use because of their potential for easy separation and purification
[119]. MoNs possess extreme mechanical strength, thermal stability, durability, and
optical and electronic properties, and such versatility and potentiality have made
them attractive in catalysis [120]. The synthetic method of three diverse MoN is
depicted in Table 7.2.

7.3.3 Solvent-Free Catalytic Application of Molybdenum Nitride


Li et al. [123] used a modified MoN-based nanocatalyst through hybridization of
PMo12 O40 3− (PMo)-paired ionic, pyridine-4-carboxylic acid (PC), and melamine
(Mel), followed by calcination in N2 (Figure 7.28). The synthesis process involves
the dissolving of PC in distilled water, and then the aqueous solution PMo is added
under stirring for 12 hours at room temperature. The obtained product is PC-PMo
OH

NH2
CO
CO

H+

OH

nm

N N
N
N

4
H

0.2
+

H2N N NH2
H+ N

HOOC
O

H+ OH Mo2N
N
NH2 CN
Hybridization Calcination
H+ N
N N
H+ H2N N NH2
N 800 °C, N2
OH HO O

O
PC-PMo-Mel-800
PC-PMo PC-PMo-Mel

Figure 7.28 Schematic illustration for the preparation of PC-PMo-Mel-800 through two
steps: hybridization and calcination. Source: Reproduced with permission from ref.
[123]/John Wiley & Sons.
7.4 Aluminum Nitride 249

Figure 7.29 Oxidative coupling of benzylamine promoted by modified MoN.

solid powder mixed with melamine in distilled water and stirred for 15 minutes
at 80 ∘ C to produce PC-PMo-Mel. The obtained PC-PMo-Mel solid product is
pyrolyzed in a tube furnace at 400, 600, and 800 ∘ C under an N2 atmosphere for two
hours, where the obtained sample at 800 ∘ C (PC-PMo-Mel-800) is more efficient
than the others.
Melamine has shown an outstanding role as a carburization and nitridation
reagent that can convert CN and form MoN by releasing N elements to react with
Mo species in PMo. The obtained PC-PMo-Mel-800 has exhibited stability and
high activity for the oxidative coupling of amines to imines (Figure 7.29) under
solvent-free and atmospheric conditions. In addition, the catalyst can be efficiently
reused at least seven times.

7.4 Aluminum Nitride


7.4.1 Introduction
Aluminum nitride (AlN) with a wide bandgap has become the critical component
of semiconductors in piezoelectric and electronic applications [124]. High thermal
conductivity [125], low thermal expansion coefficient [126], and a large bandgap
[127], chemical and physical stability at relatively high temperatures, high hardness
[128], and high resistance to molten metals, wear and corrosion [126] have charac-
terized AlN and assigned it myriad applications ranging from sensors, thin-film res-
onators, metal oxide semiconductors, to microelectronic and optoelectronic devices
[129, 130]. AlN can be found in several crystal structures with different properties:
hexagonal (wurtzite-AlN) and two cubic phases (rock salt and zinc-blende), which
is a metastable phase at ambient conditions, and involves difficult synthetic process
compared with hexagonal structure [131]. It is worth noting that greater thermal
conductivity, electrical resistivity, and hardness are provided by rock salt cubic AlN
because of the higher symmetry in its crystal structure [132].

7.4.2 Synthesis of Aluminum Nitride


The conventional synthetic approaches of AlN powder include the carbothermal
reduction–nitridation of alumina [133], direct nitridation of Al powder [134], non-
transferred arc plasma method [135], chemical routes [136], microwave-assisted
urea route [137], pulsed laser ablation [138], and transferred type arc plasma [139].
Moreover, a different methodology on the basis of solid-state metathesis has been
developed for the synthesis of nitrides using solid nitrogen-containing organic
compounds (SNCOCs). In this route, the metal oxide is first mixed with a SNCOC,
250 7 Nitride-Based Nanostructures for Solvent-Free Catalysis

Table 7.3 List of some common methods for synthesis of AlN.

Method Reagents References

Carbothermal reduction Al2 O3 + N2 + C/CH4 [141]


CVD AlCl3 + NH3 + N2 [142]
Sol–gel C9 H21 O3 Al + dextrose [143]
Mechanochemical Al powder + Melamine [144]
Solvent thermal synthesis AlCl3 + NaN3 [145]
Transferred Arc-plasma Al powder + N2 + NH3 [139]

such as a cyanamide, dicyanamide, or melamine; then, the mixture is heated in


sealed quartz tubes to form the corresponding nitride [140]. Table 7.3 summarizes
diverse methods of AlN synthetic processes.

7.4.2.1 Solvent-Free Synthesis


The major synthetic method of AlN is via the mechanochemical process with the
milling of aluminum or aluminum oxide under a nitrogen or ammonia atmosphere
[146–148]. Among solid compounds, melamine and urea have been utilized as the
common nitrogen sources to synthesize aluminum nitride [149, 150]. Melamine
with 66% nitrogen content is considered a safe and suitable nitrogen source for syn-
thesizing a wide range of metal nitrides from their oxides by low-temperature treat-
ment methods [151]. Rounaghi et al. [144] have synthesized hexagonal AlN by the
decomposition of melamine through a solid-state reaction of melamine and alu-
minum powders using high-energy ball-milling. The results disclosed that Al causes
melamine deammoniation at the first stages of milling, and further milling leads to
the s-triazine ring degradation.

7.4.3 Solvent-Free Application of Aluminum Nitride


AlN-based solvent-free catalysis has been proposed by Tekale et al. [152] to prepare
a set of active α-aminophosphonates through the Kabachnik–Fields reaction
(Figure 7.30). They achieved the reaction’s accomplishment by adding the catalyst
to aldehyde, amine, and dialkyl phosphite and heating at 50 ∘ C for one hour
under solvent-free conditions. They obtained a yield of up to 94%, higher than
using solvent under the same conditions. The applied protocol has shown a

Figure 7.30 AlN-catalyzed solvent-free synthesis of α-aminophosphonates. Source: Tekale


et al. [152]/Springer Nature.
References 251

convenient, clean, and high-yielding one-pot, three-component route for synthe-


sizing α-aminophosphonates. The procedure has been achieved in a short reaction
time with a high product yield and a simple technique. In addition, the products
have exhibited promising activities such as anti-inflammatory, antioxidant, and
cytotoxic.

7.5 Conclusion

Based on what is outlined in this chapter, nitrides can be efficiently utilized


as catalysts for a broad set of solvent-free reactions and result in high yields of
products. The main reason nitrides can be used as catalysts is because nitrides
possess nonneutral features (i.e. acidic and basic). More importantly, they can be
assumed as heterogeneous catalysts that can be subsequently recovered and reused
for the reactions. Therefore, this group of catalysts with some unknown aspects in
catalysis and green features can be excellent candidates for the catalysis of organic
reactions, whether in solvent-free or solvent-based reactions.

References

1 Wang, Y., Wang, X., and Antonietti, M. (2012). Polymeric graphitic carbon
nitride as a heterogeneous organocatalyst: from photochemistry to multipurpose
catalysis to sustainable chemistry. Angew. Chem. Int. Ed. 51 (1): 68–89.
2 Thomas, A., Fischer, A., Goettmann, F. et al. (2008). Graphitic carbon nitride
materials: variation of structure and morphology and their use as metal-free
catalysts. J. Mater. Chem. 18 (41): 4893–4908.
3 Ong, W.-J., Tan, L.-L., Ng, Y.H. et al. (2016). Graphitic carbon nitride
(g-C3 N4 )-based photocatalysts for artificial photosynthesis and environmen-
tal remediation: are we a step closer to achieving sustainability? Chem. Rev.
116 (12): 7159–7329.
4 Zhang, Y., Zhao, H., Hu, Z. et al. (2015). Protic salts of high nitrogen content as
versatile precursors for graphitic carbon nitride: anion effect on the structure,
properties, and photocatalytic activity. Chem. Tech. 80 (7): 1139–1147.
5 Miller, T.S., Jorge, A.B., Suter, T.M. et al. (2017). Carbon nitrides: synthesis and
characterization of a new class of functional materials. Phys. Chem. Chem. Phys.
19 (24): 15613–15638.
6 Zhu, J., Diao, T., Wang, W. et al. (2017). Boron doped graphitic carbon nitride
with acid-base duality for cycloaddition of carbon dioxide to epoxide under
solvent-free condition. Appl. Catal., B 219: 92–100.
7 Lin, J., Pan, Z., and Wang, X. (2014). Photochemical reduction of CO2 by
graphitic carbon nitride polymers. ACS Sustainable Chem. Eng. 2 (3): 353–358.
8 Huang, Z., Li, F., Chen, B. et al. (2013). Well-dispersed g-C3 N4 nanophases
in mesoporous silica channels and their catalytic activity for carbon dioxide
activation and conversion. Appl. Catal., B 136, 137: 269–277.
252 7 Nitride-Based Nanostructures for Solvent-Free Catalysis

9 Ma, D., Li, X., Wang, X. et al. (2021). Research development on graphitic car-
bon nitride and enhanced catalytic activity on ammonium perchlorate. RSC
Adv. 11 (10): 5729–5740.
10 Xu, J., Wu, F., Jiang, Q. et al. (2015). Metal halides supported on mesoporous
carbon nitride as efficient heterogeneous catalysts for the cycloaddition of CO2 .
J. Mol. Catal. A: Chem. 403: 77–83.
11 Goettmann, F., Thomas, A., and Antonietti, M. (2007). Metal-free activation of
CO2 by mesoporous graphitic carbon nitride. Angew. Chem. Int. Ed. 46 (15):
2717–2720.
12 Zheng, D., Cao, X.-N., and Wang, X. (2016). Precise formation of a hollow
carbon nitride structure with a janus surface to promote water splitting by
photoredox catalysis. Angew. Chem. Int. Ed. 55 (38): 11512–11516.
13 Zhang, G., Lan, Z.-A., and Wang, X. (2016). Conjugated polymers: catalysts for
photocatalytic hydrogen evolution. Angew. Chem. Int. Ed. 55 (51): 15712–15727.
14 Ou, H., Lin, L., Zheng, Y. et al. (2017). Tri-s-triazine-based crystalline carbon
nitride nanosheets for an improved hydrogen evolution. Adv. Mater. 29 (22):
1700008.
15 Cui, Y., Zhang, G., Lin, Z. et al. (2016). Condensed and low-defected graphitic
carbon nitride with enhanced photocatalytic hydrogen evolution under visible
light irradiation. Appl. Catal., B 181: 413–419.
16 Lan, Z.A., Zhang, G., and Wang, X. (2016). A facile synthesis of Br-modified
g-C3 N4 semiconductors for photoredox water splitting. Appl. Catal., B 192:
116–125.
17 Qin, J., Wang, S., Ren, H. et al. (2015). Photocatalytic reduction of CO2 by
graphitic carbon nitride polymers derived from urea and barbituric acid. Appl.
Catal., B 179: 1–8.
18 Wang, X., Blechert, S., and Antonietti, M. (2012). Polymeric graphitic carbon
nitride for heterogeneous photocatalysis. ACS Catal. 2 (8): 1596–1606.
19 Zheng, Y., Lin, L., Ye, X. et al. (2014). Helical graphitic carbon nitrides
with photocatalytic and optical activities. Angew. Chem. Int. Ed. 53 (44):
11926–11930.
20 Yang, C., Wang, B., Zhang, L. et al. (2017). Synthesis of layered carbonitrides
from biotic molecules for photoredox transformations. Angew. Chem. Int. Ed.
56 (23): 6627–6631.
21 Goettmann, F., Fischer, A., Antonietti, M. et al. (2006). Metal-free catalysis of
sustainable Friedel–Crafts reactions: direct activation of benzene by carbon
nitrides to avoid the use of metal chlorides and halogenated compounds. Chem.
Commun. 4530–4532. https://doi.org/10.1039/b608532f.
22 Goettmann, F., Fischer, A., Antonietti, M. et al. (2007). Mesoporous graphitic
carbon nitride as a versatile, metal-free catalyst for the cyclisation of functional
nitriles and alkynes. New J. Chem. 31 (8): 1455–1460.
23 Zhang, P., Gong, Y., Li, H. et al. (2013). Selective oxidation of benzene to
phenol by FeCl3 /mpg-C3 N4 hybrids. RSC Adv. 3 (15): 5121–5126.
24 Xu, J., Long, K.-Z., Wang, Y. et al. (2015). Fast and facile preparation of
metal-doped g-C3 N4 composites for catalytic synthesis of dimethyl carbonate.
Appl. Catal., A 496: 1–8.
References 253

25 Sharma, P. and Sasson, Y. (2017). A photoactive catalyst Ru-g-C3 N4 for hydro-


gen transfer reaction of aldehydes and ketones. Green Chem. 19 (3): 844–852.
26 Ge, L., Han, C., Liu, J. et al. (2011). Enhanced visible light photocatalytic
activity of novel polymeric g-C3 N4 loaded with Ag nanoparticles. Appl. Catal.,
A 409, 410: 215–222.
27 Maeda, K., Wang, X., Nishihara, Y. et al. (2009). Photocatalytic activities of
graphitic carbon nitride powder for water reduction and oxidation under visible
light. J. Phys. Chem. C 113 (12): 4940–4947.
28 Verma, S., Baig, R.B.N., Nadagouda, M.N. et al. (2017). Hydroxylation of
benzene via C–H activation using bimetallic CuAg@g-C3 N4 . ACS Sustainable
Chem. Eng. 5 (5): 3637–3640.
29 Wang, Y., Yao, J., Li, H. et al. (2011). Highly selective hydrogenation of phenol
and derivatives over a Pd@carbon nitride catalyst in aqueous media. J. Am.
Chem. Soc. 133 (8): 2362–2365.
30 Nandi, D., Siwal, S., and Mallick, K. (2017). Carbon nitride supported cop-
per nanoparticle composite: a heterogeneous catalyst for the N-arylation of
hetero-aromatic compounds. New J. Chem. 41 (8): 3082–3088.
31 Li, Y., Yutong, G., Xuan, X. et al. (2012). A practical and benign synthesis of
amines through Pd@mpg-C3 N4 catalyzed reduction of nitriles. Catal. Commun.
28: 9–12.
32 Wang, L., Yu, M., Wu, C. et al. (2016). Synthesis of Ag/g-C3 N4 composite as
highly efficient visible-light photocatalyst for oxidative amidation of aromatic
aldehydes. Adv. Synth. Catal. 358 (16): 2631–2641.
33 Sun, J., Fu, Y., He, G. et al. (2015). Green Suzuki–Miyaura coupling reaction
catalyzed by palladium nanoparticles supported on graphitic carbon nitride.
Appl. Catal., B 165: 661–667.
34 Chen, B., Wang, L., Dai, W. et al. (2015). Metal-free and solvent-free oxidative
coupling of amines to imines with mesoporous carbon from macrocyclic com-
pounds. ACS Catal. 5 (5): 2788–2794.
35 Wei, Z., Chen, Y., Wang, J. et al. (2016). Cobalt encapsulated in N-doped
graphene layers: an efficient and stable catalyst for hydrogenation of quinoline
compounds. ACS Catal. 6 (9): 5816–5822.
36 Su, Q., Yao, X., Cheng, W. et al. (2017). Boron-doped melamine-derived car-
bon nitrides tailored by ionic liquids for catalytic conversion of CO2 into cyclic
carbonates. Green Chem. 19 (13): 2957–2965.
37 Ding, G., Han, H., Jiang, T. et al. (2014). Heterogeneous copper-catalyzed
hydroxylation of aryl iodides under air conditions. Chem. Commun. 50 (65):
9072–9075.
38 Li, X.H., Wang, X., and Antonietti, M. (2012). Solvent-free and metal-free
oxidation of toluene using O2 and g-C3 N4 with nanopores: nanostructure boosts
the catalytic selectivity. ACS Catal. 2: 2082–2086.
39 Wang, J., Zhang, L., Long, F. et al. (2016). Solvent-free catalytic synthesis and
optical properties of super-hard phase ultrafine carbon nitride nanowires with
abundant surface active sites. RSC Adv. 6 (28): 23272–23278.
40 Sun, G., Liu, Z., Hu, Q. et al. (2008). Solvent-free synthesis of crystalline carbon
nitride compounds. J. Alloys Compd. 455: 303–307.
254 7 Nitride-Based Nanostructures for Solvent-Free Catalysis

41 Guo, Q., Yang, Q., Yi, C. et al. (2005). Synthesis of carbon nitrides with
graphite-like or onion-like lamellar structures via a solvent-free route at low
temperatures. Carbon 43 (7): 1386–1391.
42 Devthade, V., Kamble, G., Ghugal, S.G. et al. (2018). Visible light-driven
Biginelli reaction over mesoporous g-C3 N4 Lewis-base catalyst. ChemistrySelect
3 (14): 4009–4014.
43 Habib, S., Serwar, M., Rana, U.A. et al. (2021). A (solvent-free) approach to
metal-free photocatalysts for methylene blue degradation. Iran. Polym. J. 30:
1029–1039.
44 Chu, S., Wang, C., Feng, J. et al. (2014). Melem: a metal-free unit for photocat-
alytic hydrogen evolution. Int. J. Hydrogen Energy 39: 13519–13526.
45 Feng, X., Guo, J., Wang, S. et al. (2021). Atomically dispersed gold anchored on
carbon nitride nanosheets as effective catalyst for regioselective hydrosilylation
of alkynes. J. Mater. Chem. A 9 (33): 17885–17892.
46 Zhang, T., Wang, X., Huang, X. et al. (2016). Bifunctional catalyst of a
metallophthalocyanine-carbon nitride hybrid for chemical fixation of CO2 to
cyclic carbonate. RSC Adv. 6 (4): 2810–2818.
47 Han, H., Ding, G., Wu, T. et al. (2015). Cu and boron doped carbon nitride
for highly selective oxidation of toluene to benzaldehyde. Molecules 20:
12686–12697.
48 Karkeabadi, M., Nemati, F., Elhampour, A. et al. (2019). Cu2 O modified g-C3 N4
as an effective catalyst for the synthesis of propargylamines: experimental,
quantum mechanical mechanistic and kinetic study. React. Kinet. Mech. Catal.
126: 265–282.
49 Gao, D., Xu, Q., Zhang, J. et al. (2014). Defect-related ferromagnetism in
ultrathin metal-free g-C3 N4 nanosheets. Nanoscale 6 (5): 2577–2581.
50 Elhampour, A. and Nemati, F. (2016). Nano-Fe3 O4 @TiO2 /Cu2 O core–shell
composite: a convenient magnetic separable catalyst for A3 and KA2 coupling.
J. Chin. Chem. Soc. 63: 653–659.
51 Priya, D.D., Khan, M.M.R., and Roopan, S.M. (2020). Fabricating a g-C3 N4 /CuO
heterostructure with improved catalytic activity on the multicomponent synthe-
sis of pyrimidoindazoles. J. Nanostruct. Chem. 10: 289–308.
52 Choudhary, P., Bahuguna, A., Kumar, A. et al. (2020). Oxidized graphitic
carbon nitride as a sustainable metal-free catalyst for hydrogen transfer
reactions under mild conditions. Green Chem. 2 (15): 5084–5095.
53 Baig, R.B.N., Verma, S., Nadagouda, M.N. et al. (2016). Room temperature
synthesis of biodiesel using sulfonated graphitic carbon nitride. Sci. Rep. 6:
39387.
54 Bahuguna, A., Kumar, A., Chhabra, T. et al. (2018). Potassium-functionalized
graphitic carbon nitride supported on reduced graphene oxide as a sustain-
able catalyst for Knoevenagel condensation. ACS Appl. Nano Mater. 1 (12):
6711–6723.
55 Qareaghaj, O.H., Ghaffarzadeh, M., and Azizi, N. (2021). A rapid and quantita-
tive synthesis of xanthene derivatives using sulfonated graphitic carbon nitride
under ball-milling. J. Heterocycl. Chem. 58 (10): 2009–2017.
References 255

56 Xu, J., Shang, J.K., Jiang, Q. et al. (2016). Facile alkali-assisted synthesis
of g-C3 N4 materials and their high-performance catalytic application in
solvent-free cycloaddition of CO2 to epoxides. RSC Adv. 6 (60): 55382–55392.
57 Ghodsinia, S.S.E., Akhlaghinia, B., and Jahanshahi, R. (2021). Co3 O4
nanoparticles embedded in triple-shelled graphitic carbon nitride
(Co3 O4 /TSCN): a new sustainable and high-performance hierarchical catalyst
for the Pd/Cu-free Sonogashira–Hagihara cross-coupling reaction in solvent-free
conditions. Res. Chem. Intermed. 47: 3217–3244.
58 Li, Y., Ma, J., Liu, Z. et al. (2021). Fabrication of porous ultrathin carbon
nitride nanosheet catalysts with enhanced photocatalytic activity for N- and
O-heterocyclic compound synthesis. New J. Chem. 45 (1): 365–372.
59 Wang, X., Shi, L., Chen, Y. et al. (2021). Facile synthesis of carboxyl- and
hydroxyl-functional carbon nitride catalyst for efficient CO2 cycloaddition. Mol.
Catal. 515: 111924.
60 Rashidizadeh, A., Ghafuri, H., Zand, H.R.E. et al. (2019). Graphitic carbon
nitride nanosheets covalently functionalized with biocompatible vitamin
B1: synthesis, characterization, and its superior performance for synthesis of
quinoxalines. ACS Omega 4 (7): 12544–12554.
61 Saadat, M., Qomi, M., and Azizi, N. (2020). Greener and regioselective ring
opening of epoxides with TMSCN using potassium salts of magnetic carbon
nitride. Monatsh. Chem. 151: 1597–1602.
62 Fu, Y., Zhan, W., Guo, Y. et al. (2017). Highly efficient cobalt-doped carbon
nitride polymers for solvent-free selective oxidation of cyclohexane. Green
Energy Environ. 2 (2): 142–150.
63 Ahooie, T.S., Azizi, N., Hashemi, M.M. et al. (2018). Magnetic g-C3 N4
nanocomposite-catalyzed environmentally benign aminolysis of epoxide. Res.
Chem. Intermed. 44: 1425–1436.
64 Xu, J., Wu, F., Jiang, Q. et al. (2015). Mesoporous carbon nitride grafted with
n-bromobutane: a high-performance heterogeneous catalyst for the solvent-free
cycloaddition of CO2 to propylene carbonate. Catal. Sci. Technol. 5 (1): 447–454.
65 Goodarzi, N., Rashidizadeh, A., and Ghafuri, H. (2020). ompg-C3 N4 /SO3 H
organocatalyst-mediated green synthesis of 1,2-dihydro-1-arylnaphtho[1,2-e][1,3]
oxazin-3-ones under solvent-free and mild conditions: a fast and facile one-pot
three-component approach. Monatsh. Chem. 151: 791–798.
66 Ghafuri, H., Goodarzi, N., Rashidizadeh, A. et al. (2019). ompg-C3 N4 /SO3 H:
an efficient and recyclable organocatalyst for the facile synthesis of
2,3-dihydroquinazolin-4(1H)-ones. Res. Chem. Intermed. 45: 5027–5043.
67 Lan, D.-H., Wang, H.-T., Chen, L. et al. (2016). Phosphorous-modified bulk
graphitic carbon nitride: facile preparation and application as an acid-base
bifunctional and efficient catalyst for CO2 cycloaddition with epoxides. Carbon
100: 81–89.
68 Azizi, N. and Edrisi, M. (2020). Preparation of choline sulfate ionic liquid
supported on porous graphitic carbon nitride nanosheets by simple surface
modification for enhanced catalytic properties. J. Mol. Liq. 300: 112263.
256 7 Nitride-Based Nanostructures for Solvent-Free Catalysis

69 Chen, X.Z., Dye, J.L., Eick, H.A. et al. (1997). Synthesis of transition-metal
nitrides from nanoscale metal particles prepared by homogeneous reduction of
metal halides with an alkalide. Chem. Mater. 9 (5): 1172–1176.
70 Mahvash, F., Eissa, S., Bordjiba, T. et al. (2017). Corrosion resistance of mono-
layer hexagonal boron nitride on copper. Sci. Rep. 7: 42139.
71 Ren, S., Cui, M., Pu, J. et al. (2017). Multilayer regulation of atomic boron
nitride films to improve oxidation and corrosion resistance of Cu. ACS Appl.
Mater. Interfaces 9 (32): 27152–27165.
72 Depeursinge, A., Racoceanu, D., Iavindrasana, J. et al. (2010). Fusing visual and
clinical information for lung tissue classification in high-resolution computed
tomography. Artif. Intell. Med. 50 (1): 13–21.
73 Liu, Z., Gong, Y., Zhou, W. et al. (2013). Ultrathin high-temperature
oxidation-resistant coatings of hexagonal boron nitride. Nat. Commun. 4: 2541.
74 Li, H., Yu, C., Chen, R. et al. (2012). Novel ionic liquid-type Gemini surfac-
tants: synthesis, surface property and antimicrobial activity. Colloids Surf., A
395: 116–124.
75 Angizi, S., Alem, S.A.A., Azar, M.H. et al. (2022). A comprehensive review on
planar boron nitride nanomaterials: from 2D nanosheets towards 0D quantum
dots. Prog. Mater Sci. 124: 100884.
76 Li, C., Bando, Y., Zhi, C. et al. (2009). Thickness-dependent bending modulus
of hexagonal boron nitride nanosheets. Nanotechnology 20: 385707.
77 Liang, Y., Qin, H., Huang, J. et al. (2019). Mechanical properties of boron
nitride sheet with randomly distributed vacancy defects. Nanotechnol. Rev. 8 (1):
210–217.
78 Le, M.Q. (2014). Size effects in mechanical properties of boron nitride nanorib-
bons. J. Mech. Sci. Technol. 28: 4173–4178.
79 Außendorf, G., Jüttemann, F., Muktavat, K. et al. (2006). Total polarization of
the 185 nm emission line of mercury excited by electron impact. J. Phys. B: At.
Mol. Opt. Phys. 39 (21): 4435–4442.
80 Ding, N., Chen, X., and Wu, C.M. (2016). Mechanical properties and failure
behaviors of the interface of hybrid graphene/hexagonal boron nitride sheets.
Sci. Rep. 6: 31499.
81 Reddy, C.D., Ramasubramaniam, A., Shenoy, V.B. et al. (2009). Edge elastic
properties of defect-free single-layer graphene sheets. Appl. Phys. Lett. 94 (10):
101904.
82 Lin, Y. and Connell, J.W. (2012). Advances in 2D boron nitride nanostructures:
nanosheets, nanoribbons, nanomeshes, and hybrids with graphene. Nanoscale
4 (22): 6908–6939.
83 Meziani, M.J., Song, W.-L., Wang, P. et al. (2015). Boron nitride nanomaterials
for thermal management applications. ChemPhysChem 16 (7): 1339–1346.
84 Lu, F., Wang, F., Cao, L. et al. (2012). Hexagonal boron nitride nanomaterials:
advances towards bioapplications. Nanosci. Nanotechnol. Lett. 4: 949–961.
85 Lee, C., Wei, X., Kysar, J.W. et al. (2008). Measurement of the elastic properties
and intrinsic strength of monolayer graphene. Science 321 (5887): 385–388.
References 257

86 Balandin, A.A., Ghosh, S., Bao, W. et al. (2008). Superior thermal conductivity
of single-layer graphene. Nano Lett. 8 (3): 902–907.
87 Li, L.H. and Chen, Y. (2016). Atomically thin boron nitride: unique properties
and applications. Adv. Funct. Mater. 26 (16): 2594–2608.
88 Pakdel, A., Bando, Y., and Golberg, D. (2014). Nano boron nitride flatland.
Chem. Soc. Rev. 43 (3): 934–959.
89 Li, L.H., Cervenka, J., Watanabe, K. et al. (2014). Strong oxidation resistance of
atomically thin boron nitride nanosheets. ACS Nano 8 (2): 1457–1462.
90 Huang, C., Chen, C., Ye, X. et al. (2013). Stable colloidal boron nitride
nanosheet dispersion and its potential application in catalysis. J. Mater. Chem.
A 1 (39): 12192–12197.
91 Wang, L., Ni, S.Q., Guo, C. et al. (2013). One pot synthesis of ultrathin boron
nitride nanosheet-supported nanoscale zerovalent iron for rapid debromination
of polybrominated diphenyl ethers. J. Mater. Chem. A 1 (21): 6379–6387.
92 Khan, U., May, P., O’Neill, A. et al. (2013). Polymer reinforcement using
liquid-exfoliated boron nitride nanosheets. Nanoscale 5 (2): 581–587.
93 Yi, M., Shen, Z., Zhang, W. et al. (2013). Hydrodynamics-assisted scalable
production of boron nitride nanosheets and their application in improving
oxygen-atom erosion resistance of polymeric composites. Nanoscale 5 (21):
10660–10667.
94 Song, W.-L., Wang, P., Cao, L. et al. (2012). Polymer/boron nitride nano-
composite materials for superior thermal transport performance. Angew. Chem.
Int. Ed. 51 (26): 6498–6501.
95 Sajjad, M., Morell, G., and Feng, P. (2013). Advance in novel boron nitride
nanosheets to nanoelectronic device applications. ACS Appl. Mater. Interfaces
5 (11): 5051–5056.
96 Dean, C.R., Young, A., Meric, I. et al. (2010). Boron nitride substrates for
high-quality graphene electronics. Nat. Nanotechnol. 5: 722–726.
97 Lei, W., Qin, S., Liu, D. et al. (2013). Large scale boron carbon nitride
nanosheets with enhanced lithium storage capabilities. Chem. Commun. 49 (4):
352–354.
98 Chatterjee, S., Luo, Z., Acerce, M. et al. (2011). Chemical vapor deposition
of boron nitride nanosheets on metallic substrates via decaborane/ammonia
reactions. Chem. Mater. 23 (20): 4414–4416.
99 Song, L., Ci, L., Lu, H. et al. (2010). Large scale growth and characterization of
atomic hexagonal boron nitride layers. Nano Lett. 10 (8): 3209–3215.
100 Shi, Y., Hamsen, C., Jia, X. et al. (2010). Synthesis of few-layer hexagonal boron
nitride thin film by chemical vapor deposition. Nano Lett. 10 (10): 4134–4139.
101 Nag, A., Raidongia, K., Hembram, K.P.S.S. et al. (2010). Graphene analogues of
BN: novel synthesis and properties. ACS Nano 4 (3): 1539–1544.
102 Bhimanapati, G.R., Kozuch, D., and Robinson, J.A. (2014). Large-scale synthesis
and functionalization of hexagonal boron nitride nanosheets. Nanoscale 6 (20):
11671–11675.
103 Lin, Y., Williams, T.V., Xu, T.-B. et al. (2011). Aqueous dispersions of
few-layered and monolayered hexagonal boron nitride nanosheets from
258 7 Nitride-Based Nanostructures for Solvent-Free Catalysis

sonication-assisted hydrolysis: critical role of water. J. Phys. Chem. C 115 (6):


2679–2685.
104 Coleman, J.N., Lotya, M., O, Neill, A. et al. (2011). Two-dimensional nanosheets
produced by liquid exfoliation of layered materials. Science 331 (6017): 568–571.
105 Chen, X., Dobson, J.F., and Raston, C.L. (2012). Vortex fluidic exfoliation of
graphite and boron nitride. Chem. Commun. 48 (31): 3703–3705.
106 Marsh, K.L., Souliman, M., and Kaner, R.B. (2015). Co-solvent exfoliation and
suspension of hexagonal boron nitride. Chem. Commun. 51 (1): 187–190.
107 Lei, W., Liu, D., and Chen, Y. (2015). Highly crumpled boron nitride
nanosheets as adsorbents: scalable solvent-less production. Adv. Mater.
Interfaces 2 (3): 1400529.
108 Sun, W., Ming, Y., Fu, Q. et al. (2016). High-yield production of boron nitride
nanosheets and its uses as a catalyst support for hydrogenation of nitroaromat-
ics. ACS Appl. Mater. Interfaces 8 (15): 9881–9888.
109 Marchand, R., Tessier, F., and DiSalvo, F.J. (1999). New routes to transition
metal nitrides: and characterization of new phases. J. Mater. Chem. 9 (1):
297–304.
110 Tessier, F. and Marchand, R. (1997). An original way to prepare nitride-type
compounds from sulfide precursors. J. Alloys Compd. 262, 263: 410–415.
111 Deno, H., Kamemoto, T., Nemoto, S. et al. (2008). Formation of TiN–Ir particle
films using pulsed-laser deposition and their electrolytic properties in producing
hypochlorous acid. Appl. Surf. Sci. 254 (9): 2776–2782.
112 Lee, Y., Terashima, H., Shimodaira, Y. et al. (2007). Zinc germanium oxynitride
as a photocatalyst for overall water splitting under visible light. J. Phys. Chem.
C 111 (2): 1042–1048.
113 Buha, J., Djerdj, I., Antonietti, M. et al. (2007). Thermal transformation of
metal oxide nanoparticles into nanocrystalline metal nitrides using cyanamide
and urea as nitrogen source. Chem. Mater. 19 (14): 3499–3505.
114 Xiang, D., Liu, Y., Gao, S. et al. (2008). Evolution of phase and microstructure
during carbothermal reduction–nitridation synthesis of Ti(C,N). Mater. Charact.
59 (3): 241–244.
115 Wang, J., Grocholl, L., Gillan, E.G. et al. (2002). Facile azidothermal metathesis
route to gallium nitride nanoparticles. Nano Lett. 2 (8): 899–902.
116 Desmoulins-Krawiec, S., Aymonier, C., Loppinet-Serani, A. et al. (2004). Synthe-
sis of nanostructured materials in supercritical ammonia: nitrides, metals and
oxides. J. Mater. Chem. 14 (2): 228–232.
117 Podsiadło, S. (1995). Stages of the synthesis of indium nitride with the use of
urea. Thermochim. Acta 256 (2): 375–380.
118 Qiu, Y. and Gao, L. (2004). Metal-urea complex-a precursor to metal nitrides.
J. Am. Ceram. Soc. 87 (3): 352–357.
119 Sardar, K., Dan, M., Schwenzer, B. et al. (2005). A simple single-source precur-
sor route to the nanostructures of AlN, GaN and InN. J. Mater. Chem. 15 (22):
2175–2177.
120 Ponce, F.A. and Bour, D.P. (1997). Nitride-based semiconductors for blue and
green light-emitting devices. Nature 386 (6623): 351–359.
References 259

121 Ganin, A.Y., Kienle, L., and Vajenine, G.V. (2006). Synthesis and charac-
terisation of hexagonal molybdenum nitrides. J. Solid State Chem. 179 (8):
2339–2348.
122 Schönberg, N.I.L.S. (1954). Contributions to the knowledge of the
molybdenum-nitrogen and the tungsten-nitrogen systems. Acta Chem. Scand.
8 (2): 204–207.
123 Li, Y., Xiao, K., Li, J. et al. (2018). Molybdenum nitride nanocatalyst derived
from melamine and polyoxometalate-based hybrid for oxidative coupling of
amines to imines with air. ChemCatChem 10 (19): 4317–4323.
124 Iqbal, A. and Mohd-Yasin, F. (2018). Reactive sputtering of aluminum nitride
(002) thin films for piezoelectric applications: a review. Sensors 18 (6): 1797.
125 Oh, S.-M. and Park, D.-W. (1998). Preparation of AlN fine powder by thermal
plasma processing. Thin Solid Films 316 (1, 2): 189–194.
126 Sung, M.-C., Kuo, Y.-M., Hsieh, L.-T. et al. (2017). Two-stage plasma nitridation
approach for rapidly synthesizing aluminum nitride powders. J. Mater. Res. 32:
1279–1286.
127 Yaddanapudi, K. (2018). First-principles study of structural phase transforma-
tion and dynamical stability of cubic AlN semiconductors. AIP Adv. 8: 125006.
128 Kim, K. (2005). Plasma synthesis and characterization of nanocrystalline alu-
minum nitride particles by aluminum plasma jet discharge. J. Cryst. Growth 283
(3, 4): 540–546.
129 Kumari, N., Singh, A.K., and Barhai, P.K. (2014). Study of properties of AlN
thin films deposited by reactive magnetron sputtering. Int. J. Thin Films Sci.
Techol. 3: 34–39.
130 Dutheil, P., Orlianges, J.-C., Crunteanu, A. et al. (2015). AlN, ZnO thin films
and AlN/ZnO or ZnO/AlN multilayer structures deposited by PLD for surface
acoustic wave applications: deposited AlN, ZnO thin films and AlN/ZnO or
ZnO/AlN multilayer structures. Phys. Status Solidi A 212 (4): 817–825.
131 Shahien, M., Yamada, M., Yasui, T. et al. (2010). Cubic aluminum nitride coat-
ing through atmospheric reactive plasma nitriding. J. Therm. Spray Technol. 19:
635–641.
132 Kudyakova, V.S., Shishkin, R.A., Elagin, A.A. et al. (2017). Aluminium nitride
cubic modifications synthesis methods and its features. Rev. J. Eur. Ceram. Soc.
37 (4): 1143–1156.
133 Jackson, T.B., Virkar, A.V., More, K.L. et al. (1997). High-thermal-conductivity
aluminum nitride ceramics: the effect of thermodynamic, kinetic, and
microstructural factors. J. Am. Ceram. Soc. 80 (6): 1421–1435.
134 Da Cruz, A.C. and Munz, R.J. (1999). Review on the vapour-phase synthesis
of aluminum nitride powder using thermal plasmas. Kona Powder Part. J. 17:
85–94.
135 Kim, T.-H., Choi, S., and Park, D.-W. (2013). Effects of NH3 flow rate on
the thermal plasma synthesis of AlN nanoparticles. J. Korean Phys. Soc. 63:
1864–1870.
136 Choi, S., Im, H., and Kim, J. (2017). Synthesis of AlN particles by chemical
route for thermal interface material. Adv. Mater. Lett. 8 (9): 939–943.
260 7 Nitride-Based Nanostructures for Solvent-Free Catalysis

137 Ahn, J.B., Kim, D.S., Kim, Y.K. et al. (2016). Synthesis of AlN particles by
microwave-assisted urea route. Appl. Mech. Mater. 851: 191–195.
138 Grigoriu, C., Hirai, M., Nishiura, K. et al. (2000). Synthesis of nanosized
aluminum nitride powders by pulsed laser ablation. J. Am. Ceram. Soc. 83
(10): 2631–2633.
139 Iwata, M., Adachi, K., Furukawa, S. et al. (2004). Synthesis of purified AlN
nano powder by transferred type arc plasma. J. Phys. D: Appl. Phys. 37:
1041–1047.
140 Rounaghi, S., Eshghi, H., Scudino, S. et al. (2016). Mechanochemical route to
the synthesis of nanostructured aluminium nitride. Sci. Rep. 6: 33375.
141 Gálvez, M.E., Frei, A., Meier, F. et al. (2009). Production of AlN by carbother-
mal and methanothermal reduction of Al2 O3 in a N2 flow using concentrated
thermal radiation. Ind. Eng. Chem. Res. 48: 528–533.
142 Wu, N.C., Tsai, M.S., Wang, M.C. et al. (2000). Morphology and formation
mechanism of aluminum nitride nanocrystals synthesized by chemical vapor
deposition. J. Cryst. Growth 208: 189–196.
143 Chaundhuri, M.G. (2016). Feasibility study of synthesis of nanostructured
aluminum nitride through sol–gel route. Int. J. Eng. Res. Appl. 6: 20–22.
144 Rounaghi, S.A., Rashid, A.R.K., Eshghi, H. et al. (2012). Formation of nanocrys-
talline h-AlN during mechanochemical decomposition of melamine in the pres-
ence of metallic aluminum. J. Solid State Chem. 190: 8–11.
145 Li, L., Hao, X., Yu, N. et al. (2003). Low-temperature solvent thermal synthesis
of cubic AlN. J. Cryst. Growth 258: 268–271.
146 McCormick, P.G. and Froes, F.H. (1998). The fundamentals of mechanochemi-
cal processing. JOM 50 (11): 61–65.
147 Suryanarayana, C. (2001). Mechanical alloying and milling. Prog. Mater Sci.
46 (1, 2): 1–184.
148 Li, P., Xi, S., and Zhou, J. (2009). Phase transformation and gas-solid reaction
of Al2 O3 during high-energy ball milling in N2 atmosphere. Ceram. Int. 35 (1):
247–251.
149 Zhang, W., Li, Z., and Zhang, D. (2010). Synthesizing AlN powder by
mechanochemical reaction between aluminum and melamine. J. Mater. Res.
25 (3): 464–470.
150 Cintas, J., Montes, J.M., Cuevas, F.G. et al. (2008). Heat-resistant bulk nano-
structured P/M aluminium. J. Alloys Compd. 458 (1, 2): 282–285.
151 Zhao, H., Lei, M., Chen, X. et al. (2006). A facile route to metal nitrides
through melamine and metal oxides. J. Mater. Chem. 16 (45): 4407–4412.
152 Tekale, S.U., Kauthale, S.S., Shaikh, R.U. et al. (2014). Aluminium
nitride catalyzed solvent-free synthesis of some novel biologically active
α-aminophosphonates. J. Iran. Chem. Soc. 11: 717–724.
261

Supported Ionic Liquids for Solvent-Free Catalysis


Vahid Khakyzadeh 1 and Sahra Sheikhaleslami 2
1
Department of Chemistry, K.N. Toosi University of Technology, 15875-4416, Tehran, Iran
2
K.N. Toosi University of Technology, Faculty of Science, Department of Chemistry, PO Box 15875-4416,
Tehran, Iran

8.1 Introduction

Catalysts are the mainstay of laboratory and industrial processes, accelerating


chemical reactions. The utilization of catalysts in organic synthesis provides
extra attractive advantages, empowering chemists to maximize the total atom
economy and minimize the energy consumption of selected processes under milder
reaction conditions. According to the number of aggregate physical states present
in the reaction system, there are two main conventional disciplines in catalysis
research, including homogeneous and heterogeneous catalysis. Either of these
parallel disciplines has advantageous or disadvantageous practical aspects; thus,
enhancing their applicability is still considered one of the stimulating subjects
of modern chemistry and has led to the emergence and growth of new areas in
catalysis. In homogeneous catalysis, the interaction between catalyst and reactants,
selectivity, and efficiency is extremely high, while separating the expensive catalysts
from products and recycling them can be troublesome. Also, many homogenous
catalysts cannot resist prolonged exposure to relatively high temperatures during
the reactions and separation processes, including distillation [1]. On the other
hand, separating heterogeneous catalysts from the reaction mixture is straight-
forward. At the same time, they are often associated with low selectivity and
activity because of the restricted number of available and active sites on the
catalyst. Nevertheless, the beneficial aspects of heterogeneous catalysts outweigh
their drawbacks. As the quick recoverability of expensive catalysts is paramount
for large-scale industrial processes, they still play a dominant role in these
processes [2].
Therefore, the ultimate goal in catalysis is to create a catalyst that exhibits the
selectivity of a homogeneous one with the stability and recyclability of a hetero-
geneous one. Among various approaches implemented for the heterogenization

Solvent-Free Methods in Nanocatalysis: From Catalyst Design to Applications, First Edition.


Edited by Rafael Luque, Manoj B. Gawande, Esmail Doustkhah, and Anandarup Goswami.
© 2023 WILEY-VCH GmbH. Published 2023 by WILEY-VCH GmbH.
262 8 Supported Ionic Liquids for Solvent-Free Catalysis

of homogeneous catalysts, the immobilization of homogeneous catalysts on


solid-support materials has attracted a great deal of attention. Preparing the immo-
bilized catalyst is an elaborate process that facilitates catalyst separation. It renders
the reusability of the catalyst while maintaining its high reactivity and selectivity
for a longer time. The drawback of this strategy is that the immobilization via
covalent binding of catalyst and support requires the costly functionalization of
both the catalyst and the surface, which is a synthetically complex process and
may, in turn, change the inherent attributes of the catalyst. The immobilization
via other alternative noncovalent interactions may also cause some other problems
such as leaching [3]. Another executed approach, namely two-liquid phase catalysis
(biphasic homogenous catalysis), involves dispersing the catalyst molecularly in a
liquid phase (catalyst phase) while the reactants and products are in a second fluid
(bulk fluid). Depending on different factors, such as phase behavior of the catalyst as
well as diffusion constants and relative solubilities of the reactants, intermediates,
and products, the reaction can occur in the catalyst phase, bulk fluid, or at the
phase boundary. The downside of this classic two-phase system is the continuous
aid of mechanical energy that is inevitable to sustain the large interphase area of
two liquid phases [4].
While researchers have been dealing with problems of the methods mentioned
above, other scientists were endeavoring to puzzle out the dilemmas related to
the hazardous and volatile organic solvents, representing an apparent handicap
for the progression of sustainable chemical processes. Along this path, searching
for alternative, easily recoverable solvents with lower environmental impacts,
scientists have realized the unique properties of ILs and their predominant appli-
cability in a wide array of scientific areas. Ionic liquids (ILs) are known by various
names, such as molten salts, ionic fluids, designer solvents, and neoteric solvents.
These compounds are organic–ionic hybrids with melting points below 100 ∘ C
that are composed of an organic cation with low symmetry and an inorganic
anion. Although ILs initially emerged as an eco-friendly alternative to volatile
organic solvents, discovering new ILs with unique properties and applications
in various fields, including material science, electrochemistry, spectroscopy, and
catalysis, led to their ever-growing popularity [5]. The ILs are suitable green
reaction media for many organic catalytic reactions, presenting a general IL effect,
including control of product distribution, increased reaction rate, and facile product
separation.
Moreover, ILs can be versatile and mobile carriers of charged or polar functional
units such as transition-metal complexes, organocatalysts, and ligands via covalent
bonds. Consequently, with time, IL methods have been further expanded for
designing and functionalizing the third generation of ILs, namely task-specific
ionic liquids (TSILs) in catalysis, employing them as the supporters of substrates.
Nonetheless, from a practical standpoint, homogeneously catalyzed reactions via
biphasic liquid–liquid systems always require large amounts of expensive ILs,
which affect the economic viability of the systems. Additionally, because of the high
viscosity and low-diffusion coefficient of ILs, they are often tedious to manipulate,
and the catalytic applicability of these compounds is limited [6].
8.2 Supported Ionic Liquids 263

Solid-supported ionic liquid catalysis


High catalyst activity, selectivity and stability
Facile product separation
Mild reaction conditions
Homogeneous catalysis Reusability of catalyst Ionic liquid catalysis
High reaction rates
High activity and selectivity High activity, selectivity, and stability
Mild reaction conditions Facile product separation
High reaction rates High reaction rates
Complicate, waste-generating, and
Difficult separation expensive preparation
High waste production Use of large amounts of Ionic liquid
Mass transfer and toxicity problems
Heterogeneous catalysis
Two-liquid phase catalysis
Readily separated and recycled High activity and selectivity
Long service life Mild reaction conditions
High reaction rates
Low activity and selectivity
High-temperature reactions Costly preparation Difficult separation
High activity
Relatively low reaction rates and selectivity Catalyst leaching Continuous use of mechanical energy

Solid-supported catalysis

Figure 8.1 Advantages and disadvantages of different catalysis strategies.

During the past two decades, ILs have received wide attention and become avail-
able as innovative fluids offering a wide mixability gap with additional substances
or solvents and low volatility that made them ideal candidates and perfectly suited
fluids for the catalyst phase to implement supported catalysts and overcome the
drawbacks of other catalysis strategies (Figure 8.1). Apart from the nonvolatile
nature of ILs, the principal benefit of these compounds lies in their significant
structural variability, which is an excellent factor in designing favored supported
ionic liquid (SIL)-phase catalysts for the desired reaction. As each segment of
SIL catalytic systems can be selected independently, the concept allows precise
assembling of the required catalyst from predefined building blocks. Our goal in
this chapter of the book is to provide an utter introduction to the SILs and survey the
outstanding applications of these substances as catalysts in solvent-free reactions.

8.2 Supported Ionic Liquids

In the past few years, there has been a surge of interest in the immobilization of ILs
onto solid supports, the so-called supported ILs (SILs), to overcome the drawbacks
of heterogeneous, homogeneous, and unsupported IL catalytic systems. Although
the resulting SILs are solid, the active species can be dissolved in the IL phase and
perform like a homogeneous catalyst. The SIL catalysts have the benefits of IL
media as well as solid-support materials and offer multiple promising advantages:
(i) concomitant use of small amounts of ILs; (ii) high selectivity and activity due to
a monotonic distribution of catalytically active species within IL on the surface of
the supports, and; (iii) easy separation from the reaction mixture for further reuse.
To immobilize a thin layer of ILs onto the solid support, the strong interactions
between ILs and solid supports have to prevail over the high surface tension of
ILs that decreases the stability of ILs on the support surface. The ILs can be fixed
264 8 Supported Ionic Liquids for Solvent-Free Catalysis

+ – + – + –
– + – + – + + + + + + Covalent bonding
+ – + – + – – – – – –
+ Ionic liquid cation
Solid support Solid support – Ionic liquid anion
(a) (b)

Figure 8.2 Different types of IL-support interaction. (a) Physisorption of the IL on solid
support. (b) Chemisorption of the IL on solid support.

onto solid supports through two very different systems, including physisorption
and chemisorption (Figure 8.2). In physisorption, only weak interactions such as
hydrogen bonding and van der Waals forces bind IL to the surface of the support.
Typically, the fixation of physisorbed ILs can be improved via a physical coating of
polar oxidic supports with a thin layer of IL. In such a case, three primary forces,
including polar interactions, OH bridges, and van der Waals forces, would bind the
IL to the support surface.
Furthermore, the high viscosity of ILs would further stabilize the liquid film.
Issues arise when the catalyst is stationary for long periods, and gravity leads to a
gradual downward shift and redistribution of the IL film or when the solubility of
IL in the solvent conduces to leaching and washing the catalyst out of the IL phase.
The chemisorption method has been employed to boost the interaction between
support and IL and circumvent the aforementioned issues. A molecular layer of IL
can be covalently bound to the support, providing stronger interactions [7]. Apart
from the type of interaction between IL and solid support, IL stabilization and
distribution are dependent on the chemical, physical, and structural properties of
supports.
Given that applications of SILs benefit from the properties of both ILs and solid
supports, it is essential to fully understand their origin and unique characteristics
to develop such applications further. This knowledge opens the door for designing
ILs and solid supports that can impart further improvements in the performance
of SILs. To identify the structure–performance relationship of SILs, the character-
istics of each part of these catalysts, including ionic and solid-support segments,
should be investigated.

8.3 Building Blocks of SILs


The building-block system of SILs is composed of two main constituents, including
the IL and support segments. However, according to the demands of catalysis, a vari-
ety of components such as molecularly dispersed particles and dissolved additives
could be incorporated into the system, affording several SIL catalysts. Depending
upon the choice of SILs, the catalysis can be done by task-specified IL, dispersed
molecular catalysts in the ILs, or catalytically active solid support. Therefore, the
precise choice of ionic and solid-support segments is the first and foremost factor in
assembling a SIL building-block system.
8.3 Building Blocks of SILs 265

Common cations Common anions

+
N + N N R N + N+
R R1 R R1 R R1 Br – Cl– BF4– PF6–
lmidazolium Pyridinium Pyrrolidinium Piperidinium Bromide Chloride Tetrafluoroborate Hexafluorophosphate
R + R1
N
R3 R3
R + R1 + O O
+
R2 N N R5 S N P –
F3C S N S CF3
R R2 R R2
R3 R4 R2 R1 R1 CF3SO3– O O
Guanidinium Sulfonium Ammonium Phosphonium Trifluomethylsulfate Bis(trifluoromethylsulfonyl)imide

Figure 8.3 Common cations and anions used to prepare ILs. R, R1 , R2 , R3 , R4 , R5 = mainly
aryl, alkyl, and hydroxyl groups.

8.3.1 Ionic Segment


The ionic segment of SILs plays a decisive part in the applicability of these
compounds. Numerous possible combinations of cations and anions that meet
the definition of ILs enable the design of ILs with a diverse suite of behaviors
and a multitude of properties. Some of the common cations and anions used for
preparing various ILs are displayed in Figure 8.3. A suitable combination of cation
and anion can tune the polarity, miscibility, hydrophilicity/hydrophobicity, and
other physicochemical properties of ionic liquids. According to the composition of
SILs and the required reaction conditions, the IL segment can play four different
roles in the catalytic building-block system of SILs: (i) the functionalized IL can act
as a catalyst itself; (ii) it can be a favored solvent for other molecular catalysts such
as transition metal catalysts, organocatalysts, and nanoparticles (NPs) dispersed
in it; (iii) it can perform in a twofold manner via acting as a co-catalyst as well as
providing a physical solvent effect for catalytically active supports; and (iv) it can
serve as a support surface modifier [8].

8.3.2 Solid-Support Segment


According to the chosen SIL technique and the role of support, the solid support
can be either an inert or catalytically active segment of the SIL. In most cases, solid
supports are porous materials with a high internal surface area to stabilize a larger
interphase area. The porosity and pore size of solid support significantly impact the
fixation of the IL layer. According to the International Union of Pure and Applied
Chemistry (IUPAC) classification, porous materials have been classified into three
groups by their pore sizes: macroporous, mesoporous, and microporous solids.
In the nonporous and macroporous supports, a moderately thick film of the IL
and its proximity to the outer surface of the support particles form. Therefore, the
IL film becomes susceptible to mechanical stress, especially in the case of a liquid
bulk fluid, where droplets of the IL would be easily removed with the bulk fluid.
Hence, nonporous and macroporous supports are not appropriate for fixed-bed
applications. In the case of microporous supports, wherein the small pores are
entirely filled up with IL, because of the significant pore diffusion limitations,
266 8 Supported Ionic Liquids for Solvent-Free Catalysis

1 10 100
Micropores Mesopores Macropores
<2 nm <50 nm >50 nm

Crystalline Pseudocrystalline Amorphous (also: partly crystalline)


Classification
Layered Framework Corpuscular Glassy

Layered double
Silica-based hydroxides Zeolites MCM-41, MCM-48, Silica gel Vycor Pore Glass (VPG),
(LDHs) clays SBA-15, SBA-16 Controlled Pore Glass (CPG)

Carbon-based Graphite, Buckyballs, Activated carbons,


graphene carbon nanotubes charcoals

Figure 8.4 Classification of common support materials according to chemical and


structural features.

the substrate and product molecules are unable to pass the ions of the IL. Hence,
those parts of the catalyst phase adjacent to the support’s outer surface would be
the only available parts. Among various porous materials, mesoporous solids have
been considered ideal solid supports to immobilize ILs, as the film of IL within the
mesopores can be stable under any mechanical stress [9].
Apart from the porosity of solid supports, these compounds’ structural and
chemical features also influence the stabilization of IL films. Generally, support
materials are categorized into three types based on their structure: amorphous,
crystalline, and pseudocrystalline (Figure 8.4). Crystalline solids possess a reg-
ular geometry, whereas the constituent particles of amorphous solids have an
irregular arrangement. Also, the long-range structural order of these materials
differs while their short-range orders often appear similar. The pseudocrystalline
solids are mesoporous materials with a controlled porosity range of 1–10 nm, in
which the pore-forming materials are arranged irregularly. At the same time,
the packing of the pores is quite regular. The crystalline materials themselves
are subdivided into the “layered” and “framework” types, and the amorphous
materials are subdivided into “corpuscular” or “glassy” according to their
character.
On the other hand, from a chemistry perspective of these materials, techni-
cal supports can be categorized into two groups, including carbon-based and
oxide-based supports [10]. Each of these groups consists of different types of com-
pounds. For instance, the oxide-based supports include silica-based, alumina-based,
silica–alumina materials, and so on. We emphasize that the nature of the surface
chemistry of these materials, which is defined by the surface groups found at the
interface, is another important factor in assembling catalytic systems based on
SILs [11].
Numerous supports have already been utilized to immobilize a thin film of IL,
such as precipitated silica, silica gel, ordered mesoporous silica, porous glass, zeo-
lite, carbon nanotube (CNT), buckyballs, polymer, magnetic nanoparticles (MNPs),
metal–organic frameworks (MOFs) [12, 13]. Given that a considerable number of
support materials have been utilized in SILs, the description of each support mate-
rial could be outside the scope of this chapter. However, it would be beneficial to
8.3 Building Blocks of SILs 267

study the most attractive and prevalent support materials developed for these cat-
alytic systems up until now.

8.3.2.1 Silica Gels


Silica gel is an amorphous silicon dioxide constituting microsized irregular aggre-
gates of alternating oxygen and silicon atoms. Most of the silica gel supports utilized
in supported ionic liquid phase (SILP) catalysis are commercially available silica gels
with a surface area of 300–500 m2 g−1 . However, various kinds of silica gels possess-
ing the desired pore size, high surface area (300–800 m2 g−1 ), and optical properties
(transparency and whiteness) can be attained by altering the temperature, pH, the
applied acid, and operating drying conditions (aerogel: supercritical drying; xerogel:
slow drying; cryogel: freeze–drying). Silica gels are mainly produced in two ways
as follows: (i) the neutralization of sodium silicate solutions utilizing hydrochloric
acid or sulfuric acid, wherein the resulting silica hydrogel is then aged, washed, and
dried; and (ii) the sol–gel synthesis of silica gels based on the gelation of silicon alkox-
ides, such as tetramethoxysilane and tetraethoxysilane, in the presence of water and
ethanol. The surface of the silica gel support contains vicinal, geminal, and isolated
silanol groups (Si–OH) and siloxane groups (O–Si–O) that provide the chemical fix-
ation of IL onto the support (Figure 8.5) [14].
In recent years, silica gels have been widely used as inorganic supports for the
immobilization of ILs to develop a recyclable and efficient catalytic system in indus-
trial processes. This support material has been successfully used in SILP catalysis
and solid catalysts with ionic liquid layers (SCILLs) [15, 16].

8.3.2.2 Ordered Mesoporous Silicas


Since the discovery of highly ordered mesoporous silica materials having
well-defined pore sizes of about 1–30 nm, their use as supports, in general,
and for stabilizing ILs, in particular, has been growing significantly. This increasing
popularity is owing to their high specific pore volumes, high surface areas, and
the ability to modify their internal surfaces by various functional ILs. To date,
several mesoporous material-based SIL catalytic systems have been developed
by combining different ILs in the presence or absence of homogeneous catalysts.
The most frequently used mesoporous silica materials include Mobil Composition
of Matter (MCM)-41, MCM-48, Santa Barbara Amorphous (SBA)-15, and SBA-16.

Siloxane Vicinal silanol


bridge Geminal silanol
Isolated H H H
silanol O H O
O O H
H O
Si Si Si O
O Si O Si
O O
O O Si
Si O O O O O O
O O
O O

Figure 8.5 Schematic silica gel surface.


268 8 Supported Ionic Liquids for Solvent-Free Catalysis

– + Si(OR3)

Postsynthetic functionalization
(grafting)

MCM-41, SBA-15

+ –
+ –

+ –
– + Ionic liquid segment
Si Si Si
OH OH OH OH OH O O OR O O O RO O OR
R Organic functional group

SiO2 SiO2

Figure 8.6 Grafting silylated ILs onto ordered mesoporous silicas.

The advantage of these attractive support materials is that their pore surface
resembles crystalline silica, while the outside of the crystallites can be defined as
amorphous. According to this structural feature, the ILs having moieties with a
chance of multiple H-bonding interactions can be adsorbed selectively into the
mesopores via multiple H-bonds [17]. Among various methods for preparing
mesoporous materials-based SILs, grafting or immobilization is the most standard
method in which the functionalized (silylated) ILs covalently anchor to the internal
surface of the mesopores (Figure 8.6).
Additionally, the immobilization of IL functionalities can also be achieved by
using bis-silylated IL-bridging precursors, leading to periodic mesoporous organosil-
ica (PMO)-SIL material. It has been proven that mesoporous material-based SIL
exhibited better performances in an overwhelming number of industrially
important catalytic reactions. They are indeed much better in terms of efficiency
and reusability than the amorphous silica-IL catalytic systems [18].

8.3.2.3 Carbon Nanotubes (CNTs)


CNTs can be visualized as graphite sheets rolled into hollow cylinders with
nanometer-scale diameters. In general, CNTs are either single-walled carbon
nanotubes (SWCNTs) or multi-walled carbon nanotubes (MWCNTs). The unique
properties of CNTs, including high thermal and electrical conductivities, high
mechanical strength, high surface area, and distinct optical characteristics, make
them ideal for a wide range of biomedical, electrical, and other industrial appli-
cations [19]. One major shortcoming of CNTs is their tough processability and
dispersibility related to the nonpolar and inert character of their surface. As a result,
CNTs often require further functionalization and modification, attaching organic
or inorganic moieties to the tubular structure of them through one of four different
approaches as follows: (i) noncovalent functionalization; (ii) covalent functional-
ization; (iii) endohedral filling; (iv) external decoration with inorganic materials.
Recently, these mesoporous materials have been widely used to fabricate CNT-SILs.
In this regard, the oxidative treatment of the CNTs would be implemented in order
to grow carbonyl, carboxylic, and hydroxyl groups at the defect sites of CNTs that
could serve as a starting point for further covalent functionalization reactions,
including IL immobilization [20]. The use of CNTs offers significant advantages
for SILP catalytic systems, including high catalytic activities stemming from their
8.4 SIL Catalytic Systems 269

mesoporous nature that evades mass transfer limitations; high activity/selectivity


arising from the possibility of reaction in the inner cavity of CNTs (confinement
effect); and well-defined and tunable support structure [21].

8.3.2.4 Silica-Coated Magnetic Nanoparticles (SMNPs)


The immobilization of ILs onto the surface of nanoparticles (NPs) leads to novel
properties in the resulting hybrid while retaining the beneficial features of both
moieties. In recent years, there has been a survey of interest in developing magnetic
NP-supported ILs (MNP-ILs) as the catalyst can be easily separated from the
reaction mixture via an external magnetic field. MNPs have a number of commend-
able properties, including high thermostability, controllable size and morphology,
high surface area-to-volume ratio, chemical inertness, ease of availability, and
excellent magnetic separability, which prevents unnecessary loss of catalysts and
eliminates the need of time- and energy-consuming separation procedures. Iron
oxide NPs have gained maximum attention among all magnetic supports because
of their outstanding features, including economic viability, easy availability,
high-saturation magnetization value, and uncomplicated functionalization [22].
Although the employment of MNPs possesses enormous beneficial aspects, the
use of naked MNPs may not be very functional for large-scale industrial processes
because of their tendency to aggregate, intrinsic structural changes under oxidative
atmospheric conditions, and cytotoxicity damage. Accordingly, for the efficient
and safe catalytic applicability of MNPs, core–shell NPs are preferred, wherein
MNPs are encapsulated via an appropriate passive coat that would help prevent
the agglomeration of MNPs along with improving their chemical stability. Among
several passive coats utilized to coat MNPs, silica has emerged as the most widely
employed coating agent. The silica-coat efficiently screens the strong interactions
between the MNPs, prevents chemical and oxidative degradation of the magnetic
core, improves the dispersibility of MNPs in aqueous systems, and provides
surface silanol groups that can form strong covalent bonds for further surface
functionalization [23].

8.4 SIL Catalytic Systems


For SIL catalytic systems, there are two alternatives with respect to the phase behav-
ior of SILs, including covalently attaching a monolayer of IL and fixing a multilayer
of IL onto the solid support (Figure 8.7). In the case of immobilizing a monolayer
of IL via covalent anchoring onto the support surface, the IL becomes part of the
support and loses its specific bulk-phase properties such as conductivity, solvation
strength, and viscosity. Here, the monolayer of IL acts as a surface modifier, transfer-
ring certain functionality to the support surface. In the second case, wherein mul-
tilayers of IL are stabilized on a support, the IL retains its bulk-phase properties. In
this method, various functionalities can be incorporated into the system by dissolv-
ing molecular catalysts, such as metal salts, transition metal complexes, acids, and
NPs [24].
286 8 Supported Ionic Liquids for Solvent-Free Catalysis

fused polycyclic quinoline via Friedländer condensation with comparable reaction


benefits and catalytic activity (Figure 8.29b) [75].

8.9.2 Solvent-Free Knoevenagel Condensation


As Knoevenagel condensation has great significance in synthetic organic chem-
istry, the catalytic activity of SILs has been investigated for this reaction as well.
In this regard, Liu et al. synthesized a SIL via supporting MTESP-ImCl and
1-[(triethoxysilyl)propyl]pyridinium chloride (TESPPyCl) on the pore surface of
SBA-15 mesoporous materials for solvent-free Knoevenagel condensation, namely,
aldehydes with ethyl cyanoacetate/malononitrile [76].
In another study, Wang et al. fabricated mesoporous polyionic liquid (MPIL)
solid-base catalyst via radical copolymerization of the amine-functionalized
vinylimidazolium bromide hydrobromide [AVIM]Br⋅HBr IL with divinylbenzene
(DVB) followed by removal of HBr and ion exchange of bromide anions with
hydroxyls (Figure 8.30). The MPIL-derived solid-base catalyst served as a superior
and recyclable catalyst for Knoevenagel condensation of benzaldehyde with ethyl
cyanoacetate under solvent-free conditions. The high activity of the catalyst was
attributed to the large surface area and high content of functionalized IL precursor
on the hydrophobic polymer matrix, which enables possible dual-base synergistic
cooperation between the Lewis base of the tethered amine and the Brønsted base of
the hydroxyl anion [77].
A tertiary amine-containing basic IL catalyst developed by Zhang et al. could also
efficiently catalyze the solvent-free Knoevenagel condensation of benzaldehyde
derivatives and ethyl cyanoacetate or malononitrile [59].

8.9.3 Esterification
Synthesis of diester derivatives has constantly been a matter of great importance
because of their broad utility in the production of agrochemicals, plasticizers, and
fine chemicals. Therefore, a variety of catalytic systems have been designed to

Figure 8.30 The MPIL-derived solid-base catalyst used for Knoevenagel condensation
reaction.
8.10 Solvent-Free CO2 Conversion Reactions 287

Figure 8.31 Solvent-free diesterification of phthalic anhydride, maleic anhydride, and


succinic acid catalyzed by SDAIL@magnetic nanosilica.

promote their synthesis. Apart from the economic viability and recoverability of
catalysts, the selectivity of the synthesis has always been challenging. Accordingly,
in order to fabricate a reusable metal-free catalyst and improve the selectivity of
synthesis, a highly acidic magnetic nanosilica-SIL catalyst (Supported diacidic
ionic liquid [SDAIL]@magnetic nanosilica) was developed for the solvent-free
diesterification of maleic anhydride, phthalic anhydride, and succinic acid using
various alcohols. The Fe3 O4 NPs were supported on nanosilica, and the and
2,2-bis((3-methylimidazolidin-1-yl)methyl)propane-1,3-diol bromide ILs were
covalently grafted onto the support using chloride functionalities of magnetic
nanosilica (Figure 8.31). Here, multiple factors, such as the large surface area of
the NPs, the high acidity of IL, and the high hydrophilicity of the catalyst, led to the
extraordinary efficiency of the catalyst [78].
In another study, Jiang et al. developed an SMNP-SIL to immobilize Candida
rugosa lipase with high loading capacity. The functional SMNP-SIL was fabricated
by covalent bonding of IL–silane having different cation chain lengths and anions
(Cl− , BF4 − , and PF6 − ) onto the surface of silica-coated Fe3 O4 NPs followed by
immobilization of lipase onto the MNP via ionic physisorption. The enzyme
activity and stability of the immobilized lipase were investigated in catalyzing the
esterification reaction between oleic acid and butanol. The IL was not only used for
the stabilization of MNP to immobilize lipase but also served as a medium for lipase
catalysis. The catalytic activity of immobilized lipase was 1.07–1.18 times higher
than that of free lipase, and it could preserve 60% of its initial activity after eight
repeated batch reactions even at high reaction temperatures. In contrast, free lipase
lost its activity after six cycles [79].

8.10 Solvent-Free CO2 Conversion Reactions


The conversion of greenhouse carbon dioxide (CO2 ) gas into value-added organic
compounds is a topic of interest for researchers in academia and society. Along these
288 8 Supported Ionic Liquids for Solvent-Free Catalysis

Br
N
N
N
N
N
N
N N
Br K2S2O8/H2O Br
or Br
AIBN/EtOH
N
N Br
Br
N N
N Br AIBN: Azobisisobutyronitrile
N
bV-Imi N N bV-Imi-NT
Br

O
R TON
O bV-Imi-NT O
R CH2Cl 1058
O
40 bar, 150 °C, 3 h CH3 279
R Ph 103
CH2OH 1184

Figure 8.32 Schematic representation of bV-Imi-NT used for solvent-free synthesis of


cyclic carbonates.

lines, the production of five-membered organic cyclic carbonates via CO2 fixation
into epoxides has become a safe alternative to the reaction of appropriate alcohols
with reagents such as phosgene or its derivatives and has attracted significant
interest because of its wide applicability in organic synthesis. For this purpose,
many effective catalytic systems have been developed, among which SIL catalysts
are highly efficacious systems affording the reaction with high selectivity under
mild reaction conditions, even without using any toxic organic solvents [80]. In this
regard, imidazolium-functionalized single-walled carbon nanotubes (Im-SWCNTs)
have been fabricated via covalently anchoring two different vinyl-substituted
imidazolium-based polymers to the π-skeleton of nanotubes through a one-pot
procedure involving self-assembly and radical-initiated polymerization of the
imidazolium moieties (Figure 8.32). The retrievable Im-SWCNT materials exhibited
superior catalytic activity for the reaction of CO2 and various epoxides to synthesize
cyclic carbonates, as highlighted by the excellent turnover numbers (TONs) higher
than similar catalysts reported in the literature. Moreover, numerous active sites
cause the use of a meager amount of catalyst [81].
Following this path, bisvinylimidazolium-functionalized pristine carbon
nanohorns (bVImi-CNHs) have also been developed and used for the same pur-
pose, exhibiting excellent TON and productivity. Among these Im-CNHs, two
catalysts containing bromide or iodide counterion and an octyl organic linker
between the two imidazolium units displayed no activity loss after three cycles,
and two other catalysts bearing a p-xylyl linker and bromide anion with different
CNHs/bVImiX ratio exhibited an unprecedented increase in catalytic activity after
reusing [82].
8.10 Solvent-Free CO2 Conversion Reactions 289

Figure 8.33 The MIL-101(Cr)-TSIL catalyst used for solvent-free cycloaddition of CO2 with
epoxides.

Recently, another efficient heterogeneous catalyst has been prepared for the
solvent-free cycloaddition of CO2 with epoxides in which the MOF Matérial Institut
Lavoisier (MIL)-101(Cr) is functionalized by the TSIL imidazolium-based ionic
liquid-containing carboxylic acid (Figure 8.33). The MIL-101(Cr)-TSIL catalyst
demonstrated easy recyclability, good thermal stability, and excellent catalytic activ-
ity due to the electrostatic interactions of imidazolium cation and hydrogen bonding
of the carboxylic acid group, which activate CO2 and epoxide, respectively [83].
Also, Yu et al. created an imidazolium bromide-functionalized Mn(III)-porphyrin
MOF, namely (ImBr−MOF-545(Mn)) through postsynthetic metalation and
subsequent ImBr IL functionalization to efficiently catalyze the one-pot cascade
epoxidation–CO2 cycloaddition reaction of olefins into cyclic carbonates under
solvent-free conditions. The presented catalyst displayed superior catalytic activity
under mild conditions and good reusability for up to five runs [84]. To date, several
other solid-SIL catalytic systems have been utilized for the cycloaddition of CO2
with epoxides to promote the synthesis of cyclic carbonates under solvent-free
conditions [85–87].
The successful synthesis of oxazolidinones has also been conducted via the
solvent-free coupling of CO2 to aziridines using an efficient and recyclable
polymer-supported diol-functionalized IL catalyst, namely (PS-DFILXs). The
developed catalyst provided excellent chemo- and regioselectivity towards
5-aryl-2-oxazolidinones under mild conditions. The hydroxyl groups of PS-DFILX
proved to play a crucial role in the higher efficiency of the catalyst, and the catalyst
could be reused four times without a significant loss in catalytic activity and
selectivity [88]. Following this path, a cost-effective fibrous nanosilica-supported
nano-Ni@Pd-containing IL (KCC-1/IL/Ni@Pd) has been fabricated and is being
used to efficiently catalyze the solvent-free synthesis of 2-oxazolidinones via the
cyclization of propargylic amines with CO2 in short reaction times. Thanks to the
presence of IL robust anchors on the high surface area of KCC-1, the high loading
capacities of NPs can disperse well without aggregation and leaching [89].
290 8 Supported Ionic Liquids for Solvent-Free Catalysis

8.11 Solvent-Free Oxidation Reactions

Strategies toward synthesizing carbonyl compounds, including aldehyde and


ketones, have always been among the most fertile arenas in academic research,
owing to the importance of these intermediates in the fine chemical and pharma-
ceutical industries. Among various synthetic routes, selective oxidation of alcohols
with aqueous hydrogen peroxide (H2 O2 ) has great potential in terms of green
chemistry. To date, many efforts have been made to develop an efficient catalytic
system for this transformation [90].
Toward this end, Tan et al. designed a hydrophilic aminofunctionalized IL
(NH2 -IL) in order to covalently attach it to the SBA-15 material and then introduced
the Keggin-type H3 PW12 O40 (HPW) into the channel of the NH2 -IL-functionalized
SBA-15 with the aid of the IL moiety (Figure 8.34). The as-prepared recyclable
catalyst efficiently catalyzed the solvent-free oxidation of benzyl alcohols with H2 O2
to synthesize benzaldehyde. The hydrophilic NH2 -IL moiety had an acceleration
effect on this heterogeneously catalytic process via enhancing the accessibility of
H2 O2 , which gives higher H2 O2 efficiency in the oxidation [91].
Also, Zheng et al. prepared a graphene oxide-supported heteropolyacid-based
IL heterogeneous catalyst (PW@IL-GO) via a facile aminoprotonation and
anion-exchange reaction for the selective oxidation of alcohols with an H2 O2 oxi-
dant. The as-prepared PW@IL-GO exhibited good amphipathic nature, accelerating
mass transfer efficiency in a two-phase system and offered improved dispersity,
accessibility, and stability of active sites. It displayed high activity and selectivity
for carbonyl compound production under organic solvent-free conditions, as
well as convenient recovery, straightforward preparation, and steady reuse of
the catalyst [92]. Recently, Dobras et al. developed a SCILL catalytic system via
immobilizing the N-hydroxyphthalimide on silica gel (SiOCONHPI) with the
aid of ester bonds followed by coating it with various CoCl2 -containing ILs. The
developed catalyst could efficiently catalyze solvent-free ethylbenzene oxidation
with oxygen under mild conditions, and the highest productivity was obtained
while 1-butyl-3-methylimidazolium octylsulfate [bmim][OcOSO3 ] IL was used for
the system [93].
In another study, a highly selective solvent-free hydroxylation of aromatic
compounds to phenols was performed using H2 O2 oxidant and a series of silica-

Figure 8.34 The SBA-15-supported aminofunctionalized IL used for solvent-free oxidation


of benzyl alcohols.
8.12 Miscellaneous Solvent-Free Organic Reactions 291

Figure 8.35 Solvent-free hydroxylation of aromatic compounds using silica-supported


imidazolium-based ILs.

supported imidazolium-based ILs containing different anions. The CuCl3 -IL-SiO2


was proven to be the most efficient catalyst in the titled reaction (Figure 8.35) [94].

8.12 Miscellaneous Solvent-Free Organic Reactions


In addition to the reactions as mentioned above, SMNP-SIL catalysts have been
utilized in a series of other significant organic transformations, including diazo-
tization, N-formylation/N-arylation, C—C bond-forming reactions, etherification,
protection reactions, cyanosilylation, and the Strecker reaction. Diazotization is
an essential chemical reaction used for the preparation of azo compounds. In this
regard, Issad supported Brønsted acidic N-propyl-2-pyrrolidinium hydrogen sulfate
IL onto SMNP and utilized it as an efficient and green nanocatalyst for the one-pot
diazotization–halogenation reaction of different aromatic amines under solventless
conditions at room temperature (Figure 8.36). The diazonium salts were stable
at room temperature and reacted rapidly with sodium iodide to give aryl iodides
in good-to-excellent yields. The advantages of the developed protocol include
rapid access to products; low pollution; simple workup; and easy separation of the
catalyst [95].
Mozumdar and coworkers immobilized MTESP-Im/Cl IL onto silica-coated Fe3 O4
NPs and employed it for the solvent-free N-formylation of amines with formic acid at
room temperature. They obtained the desired compounds in excellent yields within
a short reaction time, and the catalyst was reusable for five cycles with almost similar
catalytic activity (Figure 8.37) [96].

Figure 8.36 The solvent-free diazotization–halogenation reaction of aromatic amines


using N-propyl-2-pyrrolidinium hydrogen sulfate IL.
292 8 Supported Ionic Liquids for Solvent-Free Catalysis

Figure 8.37 Solvent-free N-formylation of amines using SMNP-supported


imidazolium-based IL.

SiO2
O N O
N
Fe3O4 O Si
NH2 N O
EtO OAc
O IL-OAc@FSMNP
R O R
O 120 °C, 12 h 14 examples
80–100%
R = H, Cl, Br, Me, OMe, NO2, COCH3 Selectivity = 84–100%

FSMNP: Iron Silica Magnetic Nano Particles

Figure 8.38 Solvent-free N-aryl oxazolidin-2-ones using SMNP-supported acetate-based


butylimidazolium IL.

In another study, the same SMNP support was used to immobilize an acetate-based
butylimidazolium IL to employ as a highly efficient nanocatalyst for a straightfor-
ward one-pot synthesis of bioactive N-aryl oxazolidin-2-ones under ligand-, metal-,
and solvent-free conditions (Figure 8.38). The imidazolium cation and acetate anion
of the nanocatalyst offered excellent assemblies of H-bond donors and acceptors,
which cooperatively activated the substrates, delivering excellent product yields with
high conversion and selectivity percentage. Furthermore, the catalyst’s reusability
for up to eight consecutive cycles suggests the possibility of scaling-up in a variety
of industrial and pharmaceutical applications [97].
Among various C—C bond-formation reactions, the nonmetal-catalyzed asym-
metric Michael addition reaction has recently drawn much attention. To date,
multiple methods have been developed in this regard. Accordingly, a silica gel-
supported pyrrolidine-based chiral catalyst was readily prepared from (S)-Boc-
L-proline and found to be a highly effective organocatalyst for the solvent-free
Michael addition of ketones to β-nitrostyrene, generating the desired products in
excellent yields (up to 94%) with high diastereoselectivities and excellent enan-
tioselectivities (up to >99 : 1 dr and up to >99% ee) (Figure 8.39). In addition, the
catalyst was reused five times without a significant loss in catalytic activity and
stereoselectivity [98].
The Friedel−Crafts (FC) benzylation of arenes is one of the most important
synthetic reactions to produce privileged diarylmethane structural units. Its
conventional methods, in which strong Lewis acids, Brønsted acids catalyze the
reaction, or heterogeneous catalysts, suffer from large amounts of undesirable,
8.12 Miscellaneous Solvent-Free Organic Reactions 293

Figure 8.39 Solvent-free Michael addition of ketones to β-nitrostyrene using supported


pyrrolidine-based chiral IL.

environmentally harmful by-products along with poor catalyst reusability and


selectivity. Therefore, it is highly desirable to design efficient and reusable solid
acid catalysts for the titled transformation. Accordingly, Li et al. prepared a
heteropolyanion-based sulfonated IL-functionalized mesoporous copolymer (P(VB-
VMS)PW) and investigated its catalytic performance in solvent-free Friedel–
Crafts benzylation of single-ring aromatic compounds with benzyl alcohol. The
catalyst displayed superior yield and selectivity for the FC reaction due to
the enhanced acidity of its SO3 H-functionalities as a result of the existence of
HPA-anions as well as its suitable mesoporous copolymeric structure [99].
Bivona et al. have covalently attached multilayered thiazolium salts onto the sur-
face of SBA-15 via radical reactions between bis-vinylthiazolium dibromide salts and
mercaptopropyl-modified SBA-15 mesoporous silica and then used it as a catalyst
for the synthesis of ethers (Figure 8.40). The as-prepared hybrid materials efficiently
catalyzed the solvent-free etherification reaction of 1-phenylethanol using different
gas phases (oxygen, air, nitrogen, and argon) in seven consecutive cycles [100].
Developing efficient strategies toward selective protection and deprotection of
different functionalities is very beneficial for multistep syntheses to manipulate
other functional groups during the synthesis. In this regard, Pourjavadi et al. have
developed a green functionalized poly(ionic liquid)-coated MNP (Fe3 O4 @PIL)
acidic catalyst and protocol to protect aldehydes in solvent-free and mild conditions.
The catalyst was fabricated by a simple and effective distillation–precipitation
polymerization procedure in the presence of MNPs and exhibited excellent activity
for the protection of aldehydes, good reactivity for the deprotection of acylals, good
thermal stability, simple recovery by an external magnet, and recyclability without
significant loss in the catalytic activity [101].

Figure 8.40 Solvent-free synthesis of ethers using SBA-supported multilayered


thiazolium-based IL.
294 8 Supported Ionic Liquids for Solvent-Free Catalysis

F3C CF3
O O
O N F3C O Co O CF3
SiO2 N
O Si O O
EtO
CF3 CF3 O
R NH2 R 11 example
N O 79–90%
Dimethyl Carbonate (DMC)/60 °C, N2, Solvent-free H
R = Alkyl, Aryl

Figure 8.41 Solvent-free synthesis of carbamates using nanostarch-supported


cobalt-containing IL.

Karimian and Tajik have developed an efficient and straightforward method for
the N-Boc protection of amines with (Boc)2 O using silica-supported propyl(N-
methyl) imidazolium chloride IL([Sipmim]Cl) catalyst. The catalyst exhibited
excellent catalytic activity and reusability for the targeted protection under
solvent-free and mild reaction conditions [102].
Continuing with the use of recyclable SILs in organic transformations, a
nanostarch-supported imidazolium-based IL containing cobalt chelate anion
has been prepared and used to catalyze the one-pot synthesis of carbamates
from dimethyl carbonate and amines, affording excellent product yields under
solvent-free and mild reaction conditions (Figure 8.41) [103].
Nasresfahani et al. have prepared an efficient IL-functionalized mesoporous sil-
ica NP, namely ([pmim]FeCl4 /MSNs) nanocatalyst via covalent immobilization of
imidazolium-based ILs on the surface of mesoporous silica using the postgrafting
method. They then used it to catalyze the solvent-free reaction of triethyl orthofor-
mate with arylamines, resulting in short reaction times and high isolated yields of
N,N ′ -diaryl-substituted formamidines (Figure 8.42) [104].
In 2014, Martín et al. designed organocatalysts based on polymeric SILLP to
efficiently catalyze the solvent-free cyanosilylation of various carbonyl compounds
utilizing trimethylsilyl cyanide (Figure 8.43). The implementation of SILLP systems
provides tuning and optimization of the catalytic efficiency by changing parameters
such as the SILLP loading, the type of polymeric backbone, the nature of the
counterion, and the substitution pattern of the imidazolium units that offer a stable
organocatalytic supported system for batch cyanosilylation reaction and, more
specifically, for continuous flow cyanosilylation reaction with high productivity
and proper catalyst recyclability. The presented strategy conformed to the concept

Figure 8.42 Solvent-free synthesis of N,N′ -diaryl-substituted formamidines using


[pmim]FeCl4 /MSNs as a catalyst.
8.12 Miscellaneous Solvent-Free Organic Reactions 295

TMSCN Cl
N N CH3
PS-DVB
Pump4
Catalytic platform 2

Cl
Ru O
Solvent-free
O Pump3
TMSO CN
Ru-4 TMSCN
+
PhNH2
TfO OTf
Sc
N N SO3 OTf
Pump1 Pump6 PS-DVB
Catalytic platform 3

IL + Ru-4

O Solvent-free
Pump2 Catalytic platform 1 Pump5

OH CN
PhHN

Figure 8.43 Solvent-free continuous flow cyanosilylation of carbonyl compounds using


polymeric SILLP system.

O
ATR-FTIR
TMSN
R1 R2 Cl
N
R1 = Ar, Alkyl PS-DVB N CH3
R2 = H, Alkyl
rt, 24 h, Solvent-free
TMSO CN
Pump TMSN: Trimethylsilyl azide R1
ATR: Attenuated total reflectance R2
FTIR: Fourier transform infrared
7 examples
TMS: Trimethylsilyl group
Yields: 99->99%

Figure 8.44 The task-specific SILLP system containing imidazolium-sulfonic acid used for
three different catalytic platforms in a single continuous flow process.

of green chemistry: solvent-free, no waste regarding unconverted reactants and


side-products, 100% atom economy, no necessity for separation, and excellent
catalytic activity, along with continuous monitoring [105].
Recently, Perís et al. combined three different IL-based catalytic platforms in
a single continuous flow process to synthesize both the protected cyanohydrin
and the α-amino nitrile from the allylic alcohol (Figure 8.44) alternatively. In
the first catalytic platform, Ru-complexes were employed as catalysts for the
isomerization of different allylic alcohols into ketones under mild conditions using
the combination of [bmim][NTf2 ] ILs with supercritical CO2 (scCO2 ), wherein the
IL phase is utilized for the homogeneous immobilization of the Ru-catalyst, and
the scCO2 phase is intended to promote the delivery of reactants to the catalytic
sites of IL and to facilitate the separation of products. In catalytic platform 2, the
produced ketone from platform one transforms into its cyanohydrin trimethylsi-
lyl ether by reacting with trimethylsilyl cyanide (TMSCN) through a fixed-bed
reactor containing an efficient organocatalytic system based on polymeric SILLP.
In catalytic platform 3, the produced ketone from platform one is used for the
Strecker reaction catalyzed by task-specific SILLP-containing imidazolium-sulfonic
296 8 Supported Ionic Liquids for Solvent-Free Catalysis

acid. The cooperative effect between the scandium sites and the IL-like fragments
that leads to improving catalytic efficiency highlights the importance of these
fragments. Also, the substrates can be activated through H-bonding with both the
OTf− anion and the imidazolium cation. The last two systems were applicable
under solvent-free conditions, and no purification steps were required between
platforms [106].

8.13 Conclusion

Among various developed catalytic systems, the immobilized ionic liquids (ILs) on
solid supports, the so-called SILs, have proven promising catalysts for a wide array
of organic transformations. The immobilization of ILs can combine the beneficial
properties of ILs with those of solid supports, driving novel features while maintain-
ing the desired properties of both moieties. The stabilization of the catalysts via SILs
can actively contribute to eradicating the concerns about heterogeneous, homoge-
nous, and free IL catalytic systems, including catalyst leaching, recoverability, and
reusability, activity, and selectivity of catalyst, as well as using a large amount of
ILs, which has critical importance from economic, environmental, and scientifical
points of view.

References

1 Cornils, B., Herrmann, W.A., Beller, M., and Paciello, R. (ed.) (2017). Applied
Homogeneous Catalysis with Organometallic Compounds: A Comprehensive
Handbook in Four Volumes. Wiley-VCH.
2 Li, C. and Liu, Y. (ed.) (2014). Bridging Heterogeneous and Homogeneous
Catalysis: Concepts, Strategies and Applications. Wiley-VCH.
3 Benaglia, M. and Puglisi, A. (ed.) (2019). Catalyst Immobilization: Methods and
Applications. Wiley-VCH.
4 Cole-Hamilton, D.J. (2003). Homogeneous catalysis – new approaches to
catalyst separation, recovery, and recycling. Science 299: 1702–1706. https://
doi.org/10.1126/science.1081881.
5 Eftekhari, A. (ed.) (2017). Ionic Liquid Devices. Royal Society of Chemistry.
6 Lei, Z., Chen, B., Koo, Y.M. et al. (2017). Introduction: ionic liquids. Chem. Rev.
117: 6633–6635. https://doi.org/10.1021/acs.chemrev.7b00246.
7 Fehrmann, R., Riisager, A., and Haumann, M. (ed.) (2014). Supported Ionic
Liquids: Fundamentals and Applications. Wiley-VCH.
8 Scholz, J. and Haumann, M. (2012). Supported ionic liquid thin film
technology. In: Nanomaterials in Catalysis (ed. P. Serp and K. Philippot),
251–280. Wiley-VCH.
9 Müller, T.E. (2014). Supported ionic liquids as part of a building-block system
for tailored catalysts. In: Supported Ionic Liquids: Fundamentals and Applica-
tions (ed. R. Fehrmann, A. Riisager and M. Haumann), 209–232. Wiley-VCH.
References 297

10 Kaur, P. and Chopra, H.K. (2020). Exploring the potential of supported ionic
liquids as building block systems in catalysis. ChemistrySelect 5: 12057–12086.
https://doi.org/10.1002/slct.202002826.
11 Somorjai, G.A. and Li, Y. (2010). Introduction to Surface Chemistry and
Catalysis. Wiley.
12 Campisciano, V., Giacalone, F., and Gruttadauria, M. (2017). Supported ionic
liquids: a versatile and useful class of materials. Chem. Rec. 17: 918–938.
https://doi.org/10.1002/tcr.201700005.
13 Selvam, T., Machoke, A., and Schwieger, W. (2012). Supported ionic liquids on
non-porous and porous inorganic materials – a topical review. Appl. Catal., A
445: 92–101. https://doi.org/10.1016/j.apcata.2012.08.007.
14 Vansant, E.F., Van Der Voort, P., and Vrancken, K.C. (ed.) (1995).
Characterization and Chemical Modification of the Silica Surface. Elsevier.
15 Brünig, J., Csendes, Z., Weber, S. et al. (2018). Chemoselective supported
ionic-liquid-phase (SILP) aldehyde hydrogenation catalyzed by an Fe(II) PNP
pincer complex. ACS Catal. 8: 1048–1051. https://doi.org/10.1021/acscatal
.7b04149.
16 Jalal, A. and Uzun, A. (2017). An exceptional selectivity for partial
hydrogenation on a supported nickel catalyst coated with [BMIM][BF4].
J. Catal. 350: 86–96. https://doi.org/10.1016/j.jcat.2017.04.002.
17 Hoffmann, F., Cornelius, M., Morell, J. et al. (2006). Silica-based mesoporous
organic–inorganic hybrid materials. Angew. Chem. Int. Ed. 45: 3216–3251.
https://doi.org/10.1002/anie.200503075.
18 Vangeli, O.C., Romanos, G.E., Beltsios, K.G. et al. (2010). Grafting of
imidazolium based ionic liquid on the pore surface of nanoporous materials –
study of physicochemical and thermodynamic properties. J. Phys. Chem. B 114:
6480–6491. https://doi.org/10.1021/jp912205y.
19 Balasubramanian, K. and Burghard, M. (2005). Chemically functionalized
carbon nanotubes. Small 1: 180–192. https://doi.org/10.1002/smll.200400118.
20 Mallakpour, S. and Soltanian, S. (2016). Surface functionalization of carbon
nanotubes: fabrication and applications. RSC Adv. 6: 109916–109935. https://doi
.org/10.1039/C6RA24522F.
21 Rodríguez-Pérez, L., Teuma, E., Falqui, A. et al. (2008). Supported ionic liquid
phase catalysis on functionalized carbon nanotubes. Chem. Commun.
4201–4203: https://doi.org/10.1039/B804969F.
22 Gupta, R., Yadav, M., Gaur, R. et al. (2020). Magnetically supported ionic
liquids: a sustainable catalytic route for organic transformations. Mater. Horiz.
7: 3097–3130. https://doi.org/10.1039/D0MH01088J.
23 Gawande, M.B., Monga, Y., Zboril, R. et al. (2015). Silica-decorated magnetic
nanocomposites for catalytic applications. Coord. Chem. Rev. 288: 118–143.
https://doi.org/10.1016/j.ccr.2015.01.001.
24 Fehrmann, R., Haumann, M., and Riisager, A. (2014). Introduction.
In: Supported Ionic Liquids: Fundamentals and Applications (ed. R. Fehrmann,
A. Riisager and M. Haumann), 1–10. Wiley-VCH.
298 8 Supported Ionic Liquids for Solvent-Free Catalysis

25 Li, H., Bhadury, P.S., Song, B. et al. (2012). Immobilized functional ionic
liquids: efficient, green, and reusable catalysts. RSC Adv. 2: 12525–12551.
26 Giacalone, F. and Gruttadauria, M. (2016). Covalently supported ionic liquid
phases: an advanced class of recyclable catalytic systems. ChemCatChem
8: 664–684.
27 Bartlewicz, O., Da˛bek, I., Szymańska, A. et al. (2020). Heterogeneous catalysis
with the participation of ionic liquids. Catalysts 10: 1227.
28 Singh, S.K. and Savoy, A.W. (2020). Ionic liquids synthesis and applications: an
overview. J. Mol. Liq. 297: 112038.
29 Migowski, P., Luska, K.L., and Leitner, W. (2016). Nanoparticles on supported
ionic liquid phases – opportunities for application in catalysis. In: Nanocatalysis
in Ionic Liquids (ed. M.H. Prechtl), 249–273. Wiley-VCH.
30 Kernchen, U., Etzold, B., Korth, W. et al. (2007). Solid catalyst with ionic
liquid layer (SCILL) – a new concept to improve selectivity illustrated by
hydrogenation of cyclooctadiene. Chem. Eng. Technol. 30: 985–994. https://
doi.org/10.1002/ceat.200700050.
31 Xin, B. and Hao, J. (2014). Imidazolium-based ionic liquids grafted on solid
surfaces. Chem. Soc. Rev. 43: 7171–7187. https://doi.org/10.1039/C4CS00172A.
32 Zhang, W.i. and Cue, B.W. (ed.) (2018). Green Techniques for Organic Synthesis
and Medicinal Chemistry. Wiley.
33 Dhakshinamoorthy, A., Asiri, A.M., Alvaro, M. et al. (2018). Metal organic
frameworks as catalysts in solvent-free or ionic liquid assisted conditions. Green
Chem. 20: 86–107. https://doi.org/10.1039/C7GC02260C.
34 Andrade, M.A. and Martins, L.M. (2020). Supported palladium nanocatalysts:
recent findings in hydrogenation reactions. Processes 8: 1172.
35 Huang, J., Jiang, T., Gao, H. et al. (2004). Pd nanoparticles immobilized on
molecular sieves by ionic liquids: heterogeneous catalysts for solvent-free
hydrogenation. Angew. Chem. Int. Ed. 43: 1397–1399. https://doi.org/10.1002/
anie.200352682.
36 Tao, R., Miao, S., Liu, Z. et al. (2009). Pd nanoparticles immobilized on
sepiolite by ionic liquids: efficient catalysts for hydrogenation of alkenes and
Heck reactions. Green Chem. 11: 96–101. https://doi.org/10.1039/B811587G.
37 Chen, W., Zhang, Y., Zhu, L. et al. (2007). A concept of supported amino acid
ionic liquids and their application in metal scavenging and heterogeneous
catalysis. J. Am. Chem. Soc. 129: 13879–13886. https://doi.org/10.1021/
ja073633n.
38 Jagtap, S. (2017). Heck reaction – state of the art. Catalysts 7: 267. https://doi
.org/10.3390/catal7090267.
39 Ma, X., Zhou, Y., Zhang, J. et al. (2008). Solvent-free Heck reaction catalyzed by
a recyclable Pd catalyst supported on SBA-15 via an ionic liquid. Green Chem.
10: 59–66. https://doi.org/10.1039/B712627A.
40 Liu, G., Hou, M., and Song, J. (2010). Immobilization of Pd nanoparticles with
functional ionic liquid grafted onto cross-linked polymer for solvent-free Heck
reaction. Green Chem. 12: 65–69. https://doi.org/10.1039/B913182E.
References 299

41 Zhang, J., Zhao, G.F., Popović, Z. et al. (2010). Pd-porphyrin functionalized


ionic liquid-modified mesoporous SBA-15: an efficient and recyclable catalyst
for solvent-free Heck reaction. Mater. Res. Bull. 45: 1648–1653. https://doi.org/
10.1016/j.materresbull.2010.07.006.
42 Firouzabadi, H., Iranpoor, N., Ghaderi, A. et al. (2014). Design and synthe-
sis of a new phosphinite-functionalized clay composite for the stabilization
of palladium nanoparticles. Application as a recoverable catalyst for C–C
bond formation reactions. RSC Adv. 4: 27674–27682. https://doi.org/10.1039/
C4RA03645J.
43 Insuasty, D., Castillo, J., Becerra, D. et al. (2020). Synthesis of biologically active
molecules through multicomponent reactions. Molecules 25: 505. https://doi
.org/10.3390/molecules25030505.
44 Kumar, D., Sharma, P., Singh, H. et al. (2017). The value of pyrans as
anticancer scaffolds in medicinal chemistry. RSC Adv. 7: 36977–36999. https://
doi.org/10.1039/C7RA05441F.
45 Zhang, Q., Su, H., Luo, J. et al. (2012). A magnetic nanoparticle supported
dual acidic ionic liquid: a “quasi-homogeneous” catalyst for the one-pot syn-
thesis of benzoxanthenes. Green Chem. 14: 201–208. https://doi.org/10.1039/
C1GC16031A.
46 Nguyen, H.T., Le, N.P.T., Chau, D.K.N. et al. (2018). New nano-Fe3 O4 -supported
Lewis acidic ionic liquid as a highly effective and recyclable catalyst for the
preparation of benzoxanthenes and pyrroles under solvent-free sonication. RSC
Adv. 8: 35681–35688. https://doi.org/10.1039/C8RA04893B.
47 Zolfigol, M.A., Ayazi-Nasrabadi, R., and Baghery, S. (2016). The first urea-based
ionic liquid-stabilized magnetic nanoparticles: an efficient catalyst for the
synthesis of bis(indolyl)methanes and pyrano[2,3-d]pyrimidinone derivatives.
Appl. Organomet. Chem. 30: 273–281. https://doi.org/10.1002/aoc.3428.
48 Estakhri, E., Nasr-Esfahani, M., Mohammadpoor-Baltork, I. et al. (2017).
Chloroaluminate ionic liquid-modified silica-coated magnetic nanoparti-
cles: efficient and reusable catalyst for selective synthesis of mono- and
bis-dihydropyrano[3,2-b]chromenediones. Appl. Organomet. Chem. 31: e3799.
https://doi.org/10.1002/aoc.3799.
49 Mohammadi Ziarani, G., Mollabagher, H., Gholamzadeh, P. et al. (2018).
Synthesis of the biologically active henna based benzochromene derivatives
using ionic liquid functionalized SBA-15 as a nanoreactor. Iran. J. Catal.
8: 59–67.
50 Zhang, Q., Wei, H., Li, J. et al. (2017). One-pot synthesis of benzopyrans
catalyzed by silica supported dual acidic ionic liquid under solvent-free
conditions. Heterocycl. Commun. 23: 411–414. https://doi.org/10.1515/hc-2017-
0163.
51 Saadati-Moshtaghin, H.R. and Abbasinohoji, F. (2019). LaMnO3 supported
ionic liquid; an efficient catalyst for one-pot three-component synthesis of
tetrahydrobenzo[b]pyran derivatives under solvent-free conditions. Polycyclic
Aromat. Compd. 41: 455–466. https://doi.org/10.1080/10406638.2019.1596135.
300 8 Supported Ionic Liquids for Solvent-Free Catalysis

52 Teimuri-Mofrad, R., Gholamhosseini-Nazari, M., Payami, E. et al.


(2018). Ferrocene-tagged ionic liquid stabilized on silica-coated magnetic
nanoparticles: efficient catalyst for the synthesis of 2-amino-3-cyano-4H-pyran
derivatives under solvent-free conditions. Appl. Organomet. Chem. 32: e3955.
https://doi.org/10.1002/aoc.3955.
53 Rigi, F. and Shaterian, H.R. (2017). Silica-supported ionic liquids prompted
one-pot four-component synthesis of pyrazolopyranopyrimidines, 3-methyl-
4-aryl-4, 5-dihydro-1H-pyrano[2,3-c]pyrazol-6-ones, and 1,6-diamino-2-oxo-
1,2,3,4-tetrahydropyridine-3,5-dicarbonitriles. Polycyclic Aromat. Compd. 37:
314–326. https://doi.org/10.1080/10406638.2015.1112821.
54 Li, J.J. (2021). Hantzsch dihydropyridine synthesis. In: Name Reactions.
Springer https://doi.org/10.1007/978-3-030-50865-4_63.
55 Safari, J. and Zarnegar, Z. (2013). A magnetic nanoparticle supported
Ni2+ -containing ionic liquid as an efficient nanocatalyst for the synthesis
of Hantzsch 1,4-dihydropyridines in a solvent-free dry-system. RSC Adv.
3: 26094–26101. https://doi.org/10.1039/C3RA43601B.
56 Taheri, N., Heidarizadeh, F., and Kiasat, A. (2017). A new magnetically
recoverable catalyst promoting the synthesis of 1,4-dihydropyridine and
polyhydroquinoline derivatives via the Hantzsch condensation under
solvent-free conditions. J. Magn. Magn. Mater. 428: 481–487. https://doi.org/
10.1016/j.jmmm.2016.09.099.
57 Sharma, P. and Gupta, M. (2015). Silica functionalized sulphonic acid coated
with ionic liquid: an efficient and recyclable heterogeneous catalyst for the
one-pot synthesis of 1,4-dihydropyridines under solvent-free conditions. Green
Chem. 17: 1100–1106. https://doi.org/10.1039/C4GC00923A.
58 Zolfigol, M.A. and Yarie, M. (2015). Synthesis and characterization of novel
silica-coated magnetic nanoparticles with tags of ionic liquid. Application in the
synthesis of polyhydroquinolines. RSC Adv. 5: 103617–103624. https://doi.org/10
.1039/C5RA23670C.
59 Zhang, Q., Ma, X.M., Wei, H.X. et al. (2017). Covalently anchored tertiary
amine functionalized ionic liquid on silica coated nano-Fe3 O4 as a novel,
efficient and magnetically recoverable catalyst for the unsymmetrical Hantzsch
reaction and Knoevenagel condensation. RSC Adv. 7: 53861–53870. https://doi
.org/10.1039/C7RA10692K.
60 Asadian, M., Davoodnia, A., Beyramabadi, S.A. et al. (2019). The first
pyrazolium-based ionic liquid containing phosphotungstic acid immobilized
on CuFe2 O4 @SiO2 : a recyclable organic-inorganic nanohybrid catalyst for the
synthesis of polyhydroquinolines. Iran. J. Chem. Chem. Eng. 38: 65–77.
61 Chopda, L.V. and Dave, P.N. (2020). Recent advances in homogeneous and
heterogeneous catalyst in Biginelli reaction from 2015-19: a concise review.
ChemistrySelect 5: 5552–5572.
62 Mashhoori, M.S., Sandaroos, R., and Moghaddam, A.Z. (2020). Polymeric
imidazolium ionic liquid-tagged manganese Schiff base complex: an efficient
catalyst for the Biginelli reaction. Res. Chem. Intermed. 46: 4939–4954. https://
doi.org/10.1007/s11164-020-04230-8.
References 301

63 Yao, B.J., Wu, W.X., Ding, L.G. et al. (2021). Sulfonic acid and ionic liquid
functionalized covalent organic framework for efficient catalysis of the Biginelli
reaction. J. Organomet. Chem. 86: 3024–3032. https://doi.org/10.1021/acs.joc
.0c02423.
64 Zhang, Q., Luo, J., and Wei, Y. (2010). A silica gel supported dual acidic ionic
liquid: an efficient and recyclable heterogeneous catalyst for the one-pot
synthesis of amidoalkyl naphthols. Green Chem. 12: 2246–2254. https://doi
.org/10.1039/C0GC00472C.
65 Kotadia, D.A. and Soni, S.S. (2012). Silica gel supported – SO3 H functionalised
benzimidazolium based ionic liquid as a mild and effective catalyst for rapid
synthesis of 1-amidoalkyl naphthols. J. Mol. Catal. A: Chem. 353: 44–49. https://
doi.org/10.1016/j.molcata.2011.11.003.
66 Zare, A., Merajoddin, M., Moosavi-Zare, A.R. et al. (2016). Design and
characterization of nano-silica-bonded 3-n-propyl-1-sulfonic acid imidazolium
chloride {nano-SB-[PSIM]Cl} as a novel, heterogeneous and reusable catalyst for
the condensation of arylaldehydes with β-naphthol and alkyl carbamates. Res.
Chem. Intermed. 42: 2365–2378. https://doi.org/10.1007/s11164-015-2154-7.
67 Mohammadi, Z.G., Seiedakbari, L., Gholamzadeh, P. et al. (2017). Preparation
and characterization of ionic liquid functionalized SBA-15 and its application in
the synthesis of 2,3-dihydroquinazolinones. Iran. J. Catal. 7: 137–145.
68 Azgomi, N. and Mokhtary, M. (2015). Nano-Fe3 O4 @SiO2 supported ionic
liquid as an efficient catalyst for the synthesis of 1,3-thiazolidin-4-ones under
solvent-free conditions. J. Mol. Catal. A: Chem. 398: 58–64. https://doi.org/10
.1016/j.molcata.2014.11.018.
69 Sadeghzadeh, S.M. (2015). A heteropolyacid-based ionic liquid immobilized
onto Fe3 O4 /SiO2 /salen/Mn as an environmentally friendly catalyst in a
multi-component reaction. RSC Adv. 5: 17319–17324. https://doi.org/10.1039/
C4RA16726K.
70 Naeimi, H. and Aghaseyedkarimi, D. (2016). Ionophore silica-coated magnetite
nanoparticles as a recyclable heterogeneous catalyst for one-pot green synthesis
of 2,4,5-trisubstituted imidazoles. Dalton Trans. 45: 1243–1253. https://doi.org/
10.1039/C5DT03488D.
71 Hamadi, H., Kooti, M., Afshari, M. et al. (2013). Magnetic nanoparticle
supported polyoxometalate: an efficient and reusable catalyst for solvent-free
synthesis of α-aminophosphonates. J. Mol. Catal. A: Chem. 373: 25–29. https://
doi.org/10.1016/j.molcata.2013.02.018.
72 De, S. (2021). Applied Organic Chemistry: Reaction Mechanisms and
Experimental Procedures in Medicinal Chemistry. Wiley-VCH.
73 Yassaghi, G., Davoodnia, A., Allameh, S. et al. (2012). Preparation, character-
ization and first application of aerosil silica supported acidic ionic liquid as
a reusable heterogeneous catalyst for the synthesis of 2,3-dihydroquinazolin-
4(1H)-ones. Bull. Korean Chem. Soc. 33: 2724–2730. http://dx.doi.org/10.5012/
bkcs.2012.33.8.2724.
74 Rezayati, S., Jafroudi, M.T., Nezhad, E.R. et al. (2016). Imidazole-functionalized
magnetic Fe3 O4 nanoparticles: an efficient, green, recyclable catalyst for one-pot
302 8 Supported Ionic Liquids for Solvent-Free Catalysis

Friedländer quinoline synthesis. Res. Chem. Intermed. 42: 5887–5898. https://doi


.org/10.1007/s11164-015-2411-9.
75 Dadhania, H., Raval, D., and Dadhania, A. (2021). A highly efficient and
solvent-free approach for the synthesis of quinolines and fused polycyclic
quinolines catalyzed by magnetite nanoparticle-supported acidic ionic liquid.
Polycyclic Aromat. Compd. 41: 440–453. https://doi.org/10.1080/10406638.2019
.1595057.
76 Liu, Y., Peng, J., Zhai, S. et al. (2006). Synthesis of ionic liquid functionalized
SBA-15 mesoporous materials as heterogeneous catalyst toward Knoevenagel
condensation under solvent-free conditions. Inorg. Chem. 15: 2947–2949.
https://doi.org/10.1002/ejic.200600289.
77 Wang, X., Li, J., Chen, G. et al. (2015). Hydrophobic mesoporous poly(ionic
liquid)s towards highly efficient and contamination-resistant solid-base cata-
lysts. ChemCatChem 7: 993–1003. https://doi.org/10.1002/cctc.201402995.
78 Fareghi-Alamdari, R., Niri, M.N., Hazarkhani, H. et al. (2018). Diacidic
ionic liquid supported on magnetic-silica nanocomposite: a novel, stable,
and reusable catalyst for selective diester production. J. Iran. Chem. Soc. 15:
2615–2629. https://doi.org/10.1007/s13738-018-1450-8.
79 Jiang, Y., Guo, C., Xia, H. et al. (2009). Magnetic nanoparticles supported ionic
liquids for lipase immobilization: enzyme activity in catalyzing esterification.
J. Mol. Catal. B: Enzym. 58: 103–109. https://doi.org/10.1016/j.molcatb.2008.12
.001.
80 Monfared, A., Mohammadi, R., Hosseinian, A. et al. (2019). Cycloaddition of
atmospheric CO2 to epoxides under solvent-free conditions: a straightforward
route to carbonates by green chemistry metrics. RSC Adv. 9: 3884–3899. https://
doi.org/10.1039/C8RA10233C.
81 Buaki-Sogó, M., Vivian, A., Bivona, L.A. et al. (2016). Imidazolium function-
alized carbon nanotubes for the synthesis of cyclic carbonates: reducing the
gap between homogeneous and heterogeneous catalysis. Catal. Sci. Technol.
6: 8418–8427. https://doi.org/10.1039/C6CY01068G.
82 Calabrese, C., Liotta, L.F., Carbonell, E. et al. (2017). Imidazolium-
functionalized carbon nanohorns for the conversion of carbon dioxide:
unprecedented increase of catalytic activity after recycling.
ChemSusChem 10: 1202–1209. https://doi.org/10.1002/cssc.201601427.
83 Bahadori, M., Tangestaninejad, S., Bertmer, M. et al. (2019). Task-specific ionic
liquid functionalized–MIL–101(Cr) as a heterogeneous and efficient catalyst
for the cycloaddition of CO2 with epoxides under solvent free conditions. ACS
Sustainable Chem. Eng. 7: 3962–3973. https://doi.org/10.1021/acssuschemeng
.8b05226.
84 Yu, K., Puthiaraj, P., and Ahn, W.S. (2020). One-pot catalytic transformation of
olefins into cyclic carbonates over an imidazolium bromide-functionalized
Mn(III)-porphyrin metal–organic framework. Appl. Catal., B 273: 119059.
https://doi.org/10.1016/j.apcatb.2020.119059.
85 He, Q., O’Brien, J.W., Kitselman, K.A. et al. (2014). Synthesis of cyclic carbon-
ates from CO2 and epoxides using ionic liquids and related catalysts including
References 303

choline chloride–metal halide mixtures. Catal. Sci. Technol. 4: 1513–1528.


https://doi.org/10.1039/C3CY00998J.
86 Li, Z.J., Sun, J.F., Xu, Q.Q. et al. (2021). Homogeneous and heterogeneous ionic
liquid system: promising “ideal catalysts” for the fixation of CO2 into cyclic
carbonates. ChemCatChem 13: 1848–1866. https://doi.org/10.1002/cctc
.202001572.
87 Appaturi, J.N., Ramalingam, R., Gnanamani, M.K. et al. (2021). Review on
carbon dioxide utilization for cycloaddition of epoxides by ionic liquid-modified
hybrid catalysts: effect of influential parameters and mechanisms insight.
Catalysts 11: 4. https://doi.org/10.3390/catal11010004.
88 Watile, R.A., Bagal, D.B., Deshmukh, K.M. et al. (2011). Polymer supported diol
functionalized ionic liquids: an efficient, heterogeneous and recyclable catalyst
for 5-aryl-2-oxazolidinones synthesis from CO2 and aziridines under mild and
solvent free condition. J. Mol. Catal. A: Chem. 351: 196–203. https://doi.org/10
.1016/j.molcata.2011.10.007.
89 Sadeghzadeh, S.M., Zhiani, R., and Emrani, S. (2018). Ni@Pd nanoparticles
supported on ionic liquid-functionalized KCC-1 as robust and recyclable
nanocatalysts for cycloaddition of propargylic amines and CO2 . Appl.
Organomet. Chem. 32: e3941. https://doi.org/10.1002/aoc.3941.
90 Kopylovich, M.N., Ribeiro, A.P., Alegria, E.C. et al. (2015). Catalytic oxidation
of alcohols: recent advances. Adv. Organomet. Chem. 63: 91–174.
91 Tan, R., Liu, C., Feng, N. et al. (2012). Phosphotungstic acid loaded on
hydrophilic ionic liquid modified SBA-15 for selective oxidation of alcohols
with aqueous H2 O2 . Microporous Mesoporous Mater. 158: 77–87. https://doi.org/
10.1016/j.micromeso.2012.03.023.
92 Zheng, W., Wu, M., Yang, C. et al. (2020). Alcohols selective oxidation
with H2 O2 catalyzed by robust heteropolyanions intercalated in ionic
liquid-functionalized graphene oxide. Mater. Chem. Phys. 256: 123681. https://
doi.org/10.1016/j.matchemphys.2020.123681.
93 Dobras, G., Kasperczyk, K., Jurczyk, S. et al. (2020). N-hydroxyphthalimide
supported on silica coated with ionic liquids containing CoCl2 (SCILLs) as new
catalytic system for solvent-free ethylbenzene oxidation. Catalysts 10: 252.
94 Hu, Y.L. and Fang, D. (2016). Preparation of silica supported ionic liquids
for highly selective hydroxylation of aromatics with hydrogen peroxide under
solvent-free conditions. J. Mex. Chem. Soc. 60: 207–217.
95 Isaad, J. (2014). Acidic ionic liquid supported on silica-coated magnetite
nanoparticles as a green catalyst for one-pot diazotization–halogenation
of the aromatic amines. RSC Adv. 4: 49333–49341. https://doi.org/10.1039/
C4RA05705H.
96 Garkoti, C., Shabir, J., and Mozumdar, S. (2017). An imidazolium based ionic
liquid supported on Fe3 O4 @SiO2 nanoparticles as an efficient heterogeneous
catalyst for N-formylation of amines. New J. Chem. 41: 9291–9298. https://doi
.org/10.1039/C6NJ03985E.
97 Gupta, R., Yadav, M., Gaur, R. et al. (2017). A straightforward one-pot
synthesis of bioactive N-aryl oxazolidin-2-ones via a highly efficient
304 8 Supported Ionic Liquids for Solvent-Free Catalysis

Fe3 O4 @SiO2 -supported acetate-based butylimidazolium ionic liquid


nanocatalyst under metal- and solvent-free conditions. Green Chem.
19: 3801–3812. https://doi.org/10.1016/j.molcata.2014.11.018.
98 Li, P., Wang, L., Zhang, Y. et al. (2008). Silica gel supported pyrrolidine-based
chiral ionic liquid as recyclable organocatalyst for asymmetric Michael addition
to nitrostyrenes. Tetrahedron 64: 7633–7638. https://doi.org/10.1016/j.tet.2008.05
.039.
99 Li, J., Zhou, Y., Mao, D. et al. (2014). Heteropolyanion-based ionic liquid-
functionalized mesoporous copolymer catalyst for Friedel–Crafts benzylation of
arenes with benzyl alcohol. Chem. Eng. J. 254: 54–62. https://doi.org/10.1016/j
.cej.2014.05.124.
100 Bivona, L.A., Quertinmont, F., Beejapur, H.A. et al. (2015). Thiazolium-based
catalysts for the etherification of benzylic alcohols under solvent-free
conditions. Adv. Synth. Catal. 357: 800–810. https://doi.org/10.1002/adsc
.201400733.
101 Pourjavadi, A., Hosseini, S.H., Doulabi, M. et al. (2012). Multi-layer
functionalized poly(ionic liquid) coated magnetic nanoparticles: highly
recoverable and magnetically separable brønsted acid catalyst. ACS Catal.
2: 1259–1266. https://doi.org/10.1021/cs300140j.
102 Karimian, S. and Tajik, H. (2016). N-Boc protection of amines using silica
propyl(N-methyl)imidazolium chloride as an efficient and reusable catalyst.
Iran. J. Sci. Technol. Trans. Sci. 40: 267–274. https://doi.org/10.1007/s40995-016-
0091-y.
103 Kumar, S. and Jain, S.L. (2013). A nanostarch functionalized ionic liquid
containing imidazolium cation and cobalt chelate anion for the synthesis of
carbamates from amines and dimethyl carbonate. Dalton Trans. 42:
15214–15218. https://doi.org/10.1039/C3DT52127C.
104 Nasresfahani, Z., Kassaee, M.Z., and Eidi, E. (2017). Ionic liquid-functionalized
mesoporous silica nanoparticles ([pmim]FeCl4 /MSNs): efficient nanocatalyst for
solvent-free synthesis of N,N ′ -diaryl-substituted formamidines. Appl. Organomet.
Chem. 31: e3800. https://doi.org/10.1002/aoc.3800.
105 Martín, S., Porcar, R., Peris, E. et al. (2014). Supported ionic liquid-like phases
as organocatalysts for the solvent-free cyanosilylation of carbonyl compounds:
from batch to continuous flow process. Green Chem. 16: 1639–1647. https://doi
.org/10.1039/C3GC42238K.
106 Peris, E., Porcar, R., García-Álvarez, J. et al. (2019). Divergent multistep
continuous synthetic transformations of allylic alcohol enabled by catalysts
immobilized in ionic liquid phases. ChemSusChem 12: 1684–1691. https://doi
.org/10.1002/cssc.201900107.
305

Present Status and Future Outlook


Elka Kraleva 1 , Maria L. Saladino 2 , and Izabela S. Pieta 3
1
Leibniz-Institut für Katalyse e.V. (LIKAT Rostock), Albert-Einstein-Straße 29a, 18059 Rostock, Germany
2
University of Palermo, Department of Biological, Chemical and Pharmaceutical Science and Technology
(STEBICEF), Viale delle Scienze, Bld 17, 90128 Palermo, Italy
3
Institute of Physical Chemistry Polish Academy of Sciences, Kasprzaka 44/52, 01-224 Warsaw, Poland

9.1 Summary
The increasing environmental and health concerns have been considered in
many synthetic protocols and chemical reactions based on organic and inorganic
solvents. Among them, (i) organic solvents such as chlorinated hydrocarbons (i.e.
chloromethane), C6 and C6+ hydrocarbons (i.e. cyclohexane, heptane, aromatics,
i.e. toluene), glycols (i.e. ethylene glycol [EG], polyethylene glycol [PEG]), alcohols
(i.e. ethanol, benzyl alcohol), (ii) ionic liquids (composed of organic cations and
inorganic anions), and (iii) inorganic solvents (i.e. alkaline aqueous solution) have
been mostly reported [1–7]. Most of them are widely used in massive amounts
and are harmful and toxic to natural ecosystems and humans. Therefore, the
development of the procedure, environmentally, and economically acceptable, for
solvent-free (SF) material syntheses, mild condition material functionalization,
as well as solvent-free/solvent-less reaction protocols, has received enormous
attention in recent times in the area of green synthesis.
Efforts have been made to design material manufacturing and process pathways
to meet the green chemistry principles – minimizing (i) waste formation, (ii) waste
product effect on the environment and maximizing (i) desired product selectivity
and yield, (ii) ease and convenience of waste disposal, or (iii) recycling of waste
formed in current accessible technologies (Figure 9.1) [1, 5, 8–12].
Based on such rules, the primary objective is to simplify and adapt the classi-
cal protocols and synthesis procedures in a modern way, aiming at reducing con-
sumption of energy and raw materials while minimizing pollution. Among the most
promising ways to achieve these goals, priority is given to solvent-free techniques
because solvents are often harmful to the environment, toxic, carcinogenic, prob-
lematic to storage, and expensive. Besides, usually, complex issues apply while they
need to be removed from reaction/post-reaction environment. These approaches can
be extended to perform catalyst synthesis with minimal and well-designed volumes
Solvent-Free Methods in Nanocatalysis: From Catalyst Design to Applications, First Edition.
Edited by Rafael Luque, Manoj B. Gawande, Esmail Doustkhah, and Anandarup Goswami.
© 2023 WILEY-VCH GmbH. Published 2023 by WILEY-VCH GmbH.
306 9 Present Status and Future Outlook

Prevent wase Renewable feed stocks –


01 formation use if technically and 07
economically practicable

Less hazardous Reduce derivatives –


02 chemical synthesis simplified processes 08
Designing before chemicals Catalysis – feasibility of non
03 desired function and Green feasible reaction, increased 09
minimized toxicity. chemistry selectivity, better yield
Atom economy – principles Design for degradation
04 one step versus multistep products do not persists in 10
process the environment
Safer solvents and Real – time analysis for pollution
05 auxiliaries substances prevention control of 11
minimized hazardous substances formation

Design for energy Inherently safer chemistry


06 efficiency for accident prevention 12

Figure 9.1 Green chemical principles for material synthesis and reaction protocols
summarized from refs. [1, 5, 8].

of strong mineral acids (i.e. HCl, H2 SO4 ) or alkaline solutions (NaOH, KOH) that
can cause corrosion, safety, handling, and pollution problems as waste. On the other
hand, considering greener reaction protocols, this type of strong acid can be replaced
advantageously by solid, recyclable acids, i.e. clays [13–15]. Moreover, to increase
reaction efficiency and maximize atom economies in a given reaction, starting mate-
rials, solvents, and catalysts should be designed properly. Some of solvents must be
avoided at all, i.e. benzene, as it is carcinogenic, or if possible, should be replaced by
water (carrying out reactions in the aqueous phase) [2, 16–19].
Along with the advancements in the development of material nanoengineering,
metal economy, and material surface engineering toward a given functionality,
solvent-free (SF) techniques of catalyst synthesis provide a specific, supportable
route for various catalytic applications [1, 2, 7, 16–18, 20, 21]. Such an application
in the effective design of advanced nanomaterials minimizes conventional, i.e.
hydrothermal, synthesis. It simultaneously can assure notable increases in product
yields with advantageous characteristics, such as reduced waste production during
catalyst preparation and activation protocols providing the desired catalyst shapes
and structures, e.g. a hierarchical pore structure [22–26]. The reported achievements
in this area have shown that while targeting solvent-free material preparation, the
active surface of working catalysts needs a fundamental understanding through
in situ/operando studies and establishment of an intrinsic correlation between
the chemistry and structure of the authentic surface catalyzing a reaction and its
corresponding catalytic performance [3]. Thus, the type of material can be suitably
prepared regarding its application (Figure 9.2).
Many studies are devoted to mechanochemical synthesis of advanced nanomateri-
als for catalytic applications [3, 4, 13, 20, 31]. The development of mechanochemical
approaches to prepare catalysts with novel properties with improved catalytic activ-
ity and selectivity has significantly evolved in recent years, with most reports
available in the field over the past decade [2, 16–19]. Mechanochemistry (MCH)
has emerged as a highly promising and simple methodology and highly competitive
9.1 Summary 307

Impregnation Atomic Facet


doping engineering

Galvanic
MOF pyrolysis replacement
Alloying

Pyrolysis

In
Pe
Co-precipitation
– rfo

is
Interface
tu rm
si

es
engineering
/o a
h pe nc
nt
ra e
Sy

don

Catalysis

C – C coupling
Electrocatalysis

Suzuki – Miyura
Vapor-phase reaction

Asymmetric synthesis Henry reaction


RedOx Applications

Figure 9.2 Preparation of tailor-made and sustainable catalysts for a given functionality.
Source: Adapted from references [27–30].

with traditional synthetic protocols for catalyst/nanomaterial syntheses, which


generally involve multistep processes, heating, and/or addition of expensive and
hazardous reagents [3, 7, 13, 31]. This protocol has been used for the preparation of
a wide range of nanomaterials with catalytic applications, such as supported metal
and metal-oxide nanoparticles (NPs), perovskites, nanocomposites, metal–organic
frameworks (MOFs), and hybrid nanomaterials (e.g. based on proteins and
metal-oxide nanoparticles).
The inherent disadvantages of conventional material synthetic protocols could
be overcome by a simple grinding step under ball-milling conditions. During
grinding (or milling), a mixture/material accumulates an excess of potential
energy that, together with shear and friction forces, can introduce a large vari-
ety of defects/changes in the final material, drastically improving/modifying its
reactivity [13, 20, 32]. Moreover, mechanochemical synthesis of solid materials
(mainly heterogeneous catalysts) can remarkably influence their catalytic prop-
erties for specific applications, e.g. synthesis of oxide nanoparticles as supports.
Mechanochemical synthesis has emerged as an outstanding approach with a high
level of simplicity, reproducibility, and versatility. High-energy ball milling can
induce a variety of transformations in metal oxides such as amorphization, grain
boundary disordering, changes in particle size and surface area, and polymorphic
transformations [13, 19, 31, 33]. During grinding, mechanical energy can also create
308 9 Present Status and Future Outlook

crystal defects such as Schottky or Frenkel defects or crystallographic shear planes.


Sometimes, these kinds of defects can lead to interesting catalytic properties [3].
Generally, during the MCH treatment of solid materials, several non-steady-state
processes are induced; defects and destruction of the crystal lattice, as well as varia-
tion in the form and size of particles, are observed [34]. The energy accumulation
in defects increases the temperature of solid particles and stimulates the formation
of new phases, e.g. formation of layered titanate from anatase/rutile TiO2 [35, 36].
Local chemical reactions occur at contact points of particles in a mixture through
the atomic movement produced by local plastic deformation. This atomic move-
ment, however, is limited to near-surface regions in the oxide systems. It has been
shown that the properties of oxide mixtures prepared by MCH treatment methods
usually depend on many factors, such as experimental setup, component ratio, and
milling time [3]. Fine powders activated by MCH treatment are highly reactive, and
it is known that composite substances can be formed [33]. This synthesis process is
environmentally friendly because it is a nonthermal solid-state reaction. In some
cases, the yielded powders have unique properties that were unobtainable using
conventional thermal methods, but the introduction of impurities can be possible.
There are several examples of mechanochemical preparation of metal-oxide
nanoparticle systems with enhanced catalytic applications compared to conven-
tional systems, especially titania and zirconia (TiO2 –ZrO2 ) as catalysts or catalyst
supports for heterogeneous catalytic reactions [37]. TiO2 –ZrO2 systems are strong
solid acids showing catalytic activity in different acid-catalyzed reactions, such
as isomerization and cracking alkanes, hydration, and polymerization of alkenes
[38]. The nature of the active sites in solid acid catalysts is defined by the pres-
ence of surface protons generating Bronsted acidity or coordinately unsaturated
cationic centers, i.e. Lewis acid sites [35]. The obtained results are very interest-
ing and show some advantages of MCH methods for mixed oxide preparation
[3, 13, 19, 20, 31, 33, 34]. Moreover, perovskite, zeolites, and metal oxides prepared
via solvent-free techniques still present unique properties compared to those
prepared via solvent-based routes [13, 39]. Chitrakar et al. reported the first example
of the solvent-free route for the synthesis of Zn-Al LDH (layered double hydroxides)
by grinding metal salts in the mortar [40], while Thomas prepared layered hydroxy
salts with variable Ni/Zn ratio in a simple one-minute synthesis method by grinding
the mixture of metal salts and NaOH [41].
The nature of the active phase in the prepared catalyst is determined by the
structure of the oxide precursor species and is strictly dependent on the preparation
method [2]. One of the critical factors that affect the catalytic activity is the interac-
tion between the active components and support, which may influence the chemical
nature of the catalyst/support [42]. Furthermore, the metal–support interaction
plays an essential role in the catalytic activity by influencing both dispersion of
active ingredients and their ability to be reduced and sulfided [43]. Kraleva et al.
showed the effect of MCH as the synthesis method, and the nature of heteroatoms
(Fe, Co, Ni) of molybdenum heteropolyoxometalates as the promoters in the
TiO2 -supported catalysts (heteropolycompounds of Anderson type) on catalyst
phase composition and catalytic activity in the thiophene hydrodesulphurization
9.1 Summary 309

reaction (HDS). The study showed that HDS activity is affected by the hetero-cation
presence, and the maximum activity can be reached when optimum interaction
between the active phase and support is realized, leading to the formation of
nanosized particles.
Importantly, applying or coupling microwave irradiation during solvent-free
material synthesis can result in even more benefits, as this route is fast and energy
efficient [44]. This technique can replace conventional ways of, e.g. perovskite
preparation due to some featured properties: (i) short synthesis time, in the scale
of minutes instead of hours; (ii) a green method with limited use of solvents; and
(iii) high energy efficiency [45]. In general, microwave-assisted synthesis is based
on irradiating stoichiometric mixtures of precursors by microwaves for a specific
period of time. [44, 46–48]. Such a synthesis method is very fast compared to
conventional protocols, for example, GaAlO3 is synthesis within one minute instead
of several hours by a solid-state method [49, 50]. Moreover, metal oxides (i.e. La2 O3
and CuO to synthesize La2 CuO4 [51]) or carbonate/nitrate salts can be used as
starting materials. However, the precursor salts that are not microwave susceptors
can be mixed with a microwave susceptor to (i) assist in the initial heating without
reacting with other reactants and (ii) provide a safe decomposition of carbonates to
oxides and then to perovskites upon continuous irradiation [52].
In recent years, extensive work has been conducted on designing novel porous
nanocomposites with different functionalities for catalytic applications, sensing,
drug delivery, and adsorption. Novel dry-milling-assisted strategy for the synthe-
sis of supported transition metal–oxide nanoparticles (NPs, i.e. Fe, Co, Pd) on
mesoporous aluminosilicates (i.e. Al-Santa Barbara Amorphous mesoporous silica
[SBA]-15, with a ratio of Si/Al = 15) [25]. Iron oxide NPs were initially targeted in
many reports because of their excellent catalytic properties in a range of catalytic
processes, including oxidation, acid-catalyzed reactions, and coupling chemistries
[3, 53, 54]. Moreover, they have been shown as an important preparation strategy
during the (i) development of MOFs [55], (ii) functionalization with monosaccha-
rides of the magnetic nanoparticle (MNP) surfaces [56], and (iii) development of
bio-nanocomposites based on proteins and dopamine (DA)-coated metal-oxide
MNPs [4].
Solvent-free procedures are also extremely attractive for the synthesis of NPs and
bimetallic nanocatalysts (bi-NCs), with well-controlled shape and structure [57–59].
They can offer many advantages from the aspects of synthetic chemistry, general
principles, and new strategies on surface structure control of bi-NCs at the atomic
level [60–64]. Notably, green methods to remove capping agents from surfaces of
bi-NCs without affecting their structural parameters also need to be investigated.
Besides, mild reductants need to be developed if the noble-to-non-noble metals
transition is targeted and for further cost reduction of these nanocatalysts. Thus,
the synthetic methods based on organic solvents, especially the solvothermal
method, should be further developed [58]. For such development, operando and
in situ studies of surface reconstruction and segregation in the catalytic process
have to be considered, e.g. ambient-pressure X-ray photoelectron spectroscopy or
environmental transmission electron microscopy. The complex study will result in
310 9 Present Status and Future Outlook

the authentic information of surface morphology, structure, and chemistry changes


under reaction conditions and guide toward tailored material synthesis with high
activity, selectivity, and chemical and thermal stability in return.
Key advantages of mechanochemically assisted preparations include operation
without a need for additives (e.g. solvents, templates, surfactants) and post-synthetic
treatments in many cases. This avoids the generation of solvent and/or gaseous
wastes and simplifies the overall synthetic protocol, ultimately simplifying scale-up.
Besides, this synthetic strategy provides materials with unique properties, often
impossible to achieve by classical methods.
Another aspect of solvent-free techniques is related to many important organic
reactions and transformations [7, 65]. According to the green chemistry principles,
the catalyst design, targeting improved selectivity and maximizing reaction yield,
must be performed, accounting for economic process viability minimizing waste
formation [5, 45, 50]. An important issue is to maximize the atom economy and
avoid toxic by-product formation and unacceptable environmental consequences.
Additional issues are catalyst pretreatment and use of renewable feedstocks. In this
context, one of the main challenges in the field of solvent-free reactions is the new
product/highly demanded product synthesis to replace the traditional processes pro-
viding higher efficiencies, social and environmental acceptance [5].
Numerous organic reactions can be successfully carried out under solvent-free
conditions [1, 5, 7, 11]. In many applications, enzymes have also been proposed to
carry out reactions without solvents. Such reactions are environmentally friendly,
e.g. the synthesis of polyglycerol polyricinoleate through Rhizopus arrhizus lipase
as a biocatalyst [66, 67]. Moreover, multistep and multicomponent reactions, which
include many reagents, have to be modified to improve selectivity toward a single
product. They should be redesigned using green methods and based on alternative
substrates, e.g. diketene application for a new Biginelli-like reaction.
In Table 9.1, examples of the most investigated solvent-assisted and their
corresponding solvent-free reactions are summarized. As presented in Table 9.1,
chemical transformations can be performed in the gaseous, liquid, or solid state, but
the majority of these transformations are carried out in the solution phase. A proper
design catalyst is usually needed to control their selectivity and efficiencies [20].
For instance, among all the known oxidative transformations, the oxidation of
alcohols to ketones and aldehydes has gained a particular interest of the research
community due to its broad range of industrial applications [6]. The scale-up of
the oxidation reactions has been significantly restricted due to use of heavy metals,
limited selectivity for highly functionalized compounds, and the process’s high
temperature. In this regard, new alternatives to conventional oxidation methods
have been developed, requiring the stoichiometric amounts of inorganic oxidants,
which are highly toxic and polluting, aiming to minimize chemical waste in these
catalytic processes. For such purposes, cleaner oxidants (“green oxidants”), such as
molecular oxygen or H2 O2 , can be employed in the process [6].
Bioethanol to valuable chemicals or intermediates transformation is another
important example as this reaction is considered to benefit the sustainable
development of the current chemical-, energy-, and transportation industry and
9.1 Summary 311

Table 9.1 Solvent-assisted and their corresponding solvent-free reactions.

Reactions Classical/solvent-free conditions

Acylation reaction catalyzed by carbon tetrabromide Substrates: phenols, alcohols, and thiols
with acylating agents, e.g. acyl halides
and anhydrides of carboxylic acids.
OH O O
Catalyst. AlCl3 . Disadvantages: low
CBr4(5 mol%)
yields, improved efficiency, and
+ AC2O selectivity needed
solvent-free, RT
Air, 24h 91 % Substrates: phenols, alcohols, and thiols.
Catalyst: carbon tetrabromide (CBr4 ).
High yields, regioselectivity and
chemoselectivity in various organic
transformations
Aldol condensation Substrates: 3-nitrobenzaldehyde and
O O cyclohexanone using an acid/base
O OH
catalyst. Disadvantages: high
Catalyst 1c temperature, low yields, use of organic
R R
+
0°C solvents
X X
Substrates: ketones and aryl aldehydes
Catalyst: glucosamide-derived
prolinamide. High yields, avoidance of
organic solvents, easy extraction
Cyclic ketals synthesis Substrates: carbonyl group and diols.
R1
Catalyst: acidic catalysts, benzene, or
O
O toluene as solvents. Disadvantages:
O R2 harmful to the environment
1 2
R R Substrates: carbonyl groups and diols.
O 2 a-e
O HO OH Catalyst: dry conditions using P-toluene
+ sulfonic acid (PTSA) in toluene under
R4 R4 R3
R3 R5 microwave. High yields, reduced
O R4
O reaction time
1 OH OH
O R4
R5
R1 R2
2a H H 3 a-c
2b H CH3 R3 R4 R5
2c CH3 CH3 3a CH3 H H
2d H C2H5 3b CH3 H CH3
2e H C3H7 3c H CH3 H

Jasminaldehyde synthesis Substrates: 1-heptanal and


O benzaldehyde Catalyst: liquid alkali like
NaOH or KOH. Disadvantages: lack of
O catalyst reusability, hazardous liquid
+
O alkali wastes, difficult disposal of spent
liquid bases
O
Substrates: 1-heptanal and
benzaldehyde in 1 : 5 ratio. Catalyst:
chitosan in nitrogen atmosphere.
Environmentally benign, no hazardous
waste production, short reaction time,
high selectivity
(Continued)
312 9 Present Status and Future Outlook

Table 9.1 (Continued)

Reactions Classical/solvent-free conditions

Heck reaction Substrates: coupling of olefins with aryl


or vinyl halides. Catalyst: palladium.
+ Disadvantages: catalyst recovery and
CO2Me
recycling, expensive, unstable, and toxic
ligands, e.g. phosphine for Pd
Base, SBA-TMG-pd
stabilization and Pd black formation
120–140 °C
Substrates: olefins and aryl halides.
CO2Me Catalyst: SBA-TMG-Pd, Pd supported on
1,1,3,3-tetramethylguanidinium (TMG)
via an ionic liquid. Environmentally
benign, economically profitable,
SBA-TMG-Pd more active and stable
Oxidative coupling of phenols Substrates: phenols in solution.
Catalyst: metal salts such as FeCl3 or
manganese tris(acetylacetonate)
FeCl3*6H2O OH aluminum. Disadvantages: expensive,
solid OH quinines as by-products
OH
Substrates: coupling reactions in the
solid state accelerated by irradiation
with ultrasound. Catalyst: FeCl3 .
Environmentally benign, fast, and more
efficient
Pyrazole chalcones synthesis Substrates: ketones and aryl halides.
Catalyst: alkali metal hydroxide or
sodium ethoxide. Disadvantages:
AR CHO
O expensive, use of harmful organic
solvents and difficult extraction process
N +
N R Substrates: pyrazole aldehydes,
acetophenones, and activated barium
Ba(OH)2 hydroxide mixture grinding. Catalyst:
grinding BaOH. Catalyst: high yields aryl
aldehydes. Environmentally benign,
short reaction time, high yields, mild
O N reaction conditions
N
R
AR

Tishchenko reaction Substrates: aldehydes to their dimeric


O esters. Catalyst: aluminum-based
Catalyst catalysts. Disadvantages: catalyst
2RCHO R O R recovery, more efficient reagents needed
Substrates: olefins and aryl halides.
Catalyst: ball milling and sodium
hydride. Environmentally benign,
economically profitable, successful in a
liquid nitrogen environment
9.1 Summary 313

Table 9.1 (Continued)

Reactions Classical/solvent-free conditions

Transesterifications Substrates: esters, alcohols. Catalyst:


R R strong acids, soluble bases e.g. caustic
soda. Disadvantages: catalyst types, not
R1OH + O OR2 R2OH + safe to handle and store, hazardous
O OR1
reagents in the preparation of catalysts
Substrates: esters, alcohols. Catalyst:
solid, reusable fluorapatite and
hydroxyapatite-supported zinc chloride.
Environmentally benign, reusable
catalyst, sustainable fuels production
Baeyer–Villiger oxidation Substrates: ketones, peracids, i.e.
O peroxyacid O 3-chloroperbenzoic acid (MCPBA),
or peroxide trifluoroperoxyacetic acid (TFPAA), or
R2
R1 R2 R1 O hydrogen peroxide, CHCl3 . Catalyst:
Lewis acid, poly(ionic) liquids,
Ketone Ester
polyoxometalates, transition metals,
O Ga-based catalysts. Disadvantages:
O peroxyacid catalyst types, not safe to handle and
or peroxide O store, hazardous reagents in the
[ ]n [ ]n preparation of catalysts
Substrates: ketones, MCPBA. Catalyst:
Cyclic ketone Lactone
Environmentally benign, solid-state
reaction with high product yield
Reduction of ketones to alcohols Substrates: ketones, HCl. Catalyst:
Ph2CO Ph2CH-OH Zn-based catalysts. Disadvantages:
catalyst types, hazardous reagents in the
NaBH4 preparation of catalysts
trans-PhCH=CHCOPh trans-PhCH=CHCPh
RT Substrates: ketones, NaBH4 . Catalyst:
OH
1:1 none. Environmentally benign, room
PhCH2CH2CHCPh temperature, solid-state reaction with
OH high product yield

Dearomatization – Birch reduction Substrates: arenes, i.e. benzene, liquid


E NH3 , proton donors, tripyrrolidinophos-
10-26°C, H2N(CH2)2NH2 phoric acid (TPPA), diethylene glycol
R R monobutyl ether (DBB), tetrahydro-
t-BuOH, THF, E furan (THF), dimethylurea (DMU),
highly reactive electrophile (E).
Catalyst: alkali metals. Disadvantages:
difficult control, cryogenic conditions,
catalyst types, hazardous and expensive
reagents
Substrates: arenes, C4 H8 O, lithium in
ethylenediamine, BuOH for selectivity
control, highly reactive electrophile (E).
Catalyst: none. Room temperature
reaction, ammonia-free condition, high
product yield
(Continued)
314 9 Present Status and Future Outlook

Table 9.1 (Continued)

Reactions Classical/solvent-free conditions

Hydrogenation of phenols Substrates: phenols, water, protic


H H solvents. Catalyst: Co, Fe, Ni complexes,
OH
Rh-based catalysts. Disadvantages:
OH OH catalyst types, hazardous reagents in the
H2
H + OH preparation of catalysts
OH PtO2, solid HO H
Substrates: phenols, H2 at pressure
36 cis-37 trans-38 1 bar. Catalyst: PtO2 . Environmentally
benign, mild condition, solid-state
reaction with product yield 68%
Halogenation of o-stilbene carboxylic acids Substrates: o-stilbene carboxylic acids,
O OH O OH acyl chlorides or organic acid chlorides,
H R Br NH4 Cl, CHCl3 , SnCl4 . Catalyst: strong
H R
Br2 acids, AlCl3 . Disadvantages: catalyst
H Gas-solid H types, not safe to handle and store,
Br
hazardous reagents
Substrates: o-stilbene carboxylic acids,
Br2 or pyridine HBr Br2 complex.
Catalyst: none. Environmentally
benign, room temperature (RT)
selective reaction to erythro-1,2-
dibromo-1,2-dihydro-o-stilbene
carboxylic acid
C–N forming reaction – Michael addition of amines, Substrates: amines, organic solvents,
thiophenol and methyl acetate to chalcone protic acids. Catalyst: Lewis or Brønsted
rt, acids, transition metal salts.
+ n-Bu2NH H2O Disadvantages: catalyst types, not safe to
O
NnBu2 O handle and store, hazardous reagents,
MeO MeO economic reasons
rt, H2O
PhSH
+ K2CO3 Substrates: amines, thiophenol and
O SPh O methyl acetate, water suspension,
rt, H2O
C19 H42 BrN (cetrimonium bromide),
+ MeCOCH2CO2Me K CO
2 3
n-BuNH2 , K2 CO3 . Catalyst: none.
O Me O
Environmentally benign, economic,
CO2Me product yields 92–98%
O

Alcohol dehydration Substrates: alcohols. Catalyst: Lewis or


rt,HCl gas Brønsted acids, transition metal salts.
PhR1C CH2R2 PhR C 1
CH2R 2

solid
Disadvantages: catalyst types, not safe to
OH handle and store, hazardous reagents in
the preparation of catalysts, reaction
temperature too high, economic factors
Substrates: alcohols, HCl. Catalyst:
Cl3 CCO2 H. Environmentally benign,
solid-state reaction, high yield 99%
(Continued)
9.1 Summary 315

Table 9.1 (Continued)

Reactions Classical/solvent-free conditions

Cycloaddition – Diels–Alder reaction Substrates: dienes, alkene (dienophile).


O Catalyst: Lewis acids, i.e. zinc chloride,
boron trifluoride, tin tetrachloride,
NH2
aluminum chloride. Disadvantages:
reactants stability (dienes), catalyst
+ + types, not safe to handle and store,
hazardous reagents and catalysts
Substrates: diene, alkene, maleic
MW assisted
Urea anhydride, melamine, urea. Catalyst:
zeolites. Solvent free, microwave
assisted or thermal conditions, reusable
NH catalyst, 10–150 ∘ C, high selectivity,
solid-state reaction
NH

Reformatsky reaction Substrates: aldehydes or ketones, α-halo


RCHO + BrCH2COOEt esters, anhydrous solvents, benzene,
ethers i.e. diethyl ether, THF, glyme or
Zn dimethoxymethane. Catalyst: metallic
NH4Cl zinc/zinc dust. Disadvantages: catalyst
RCH(OH)CH2COOEt type, not safe to handle and store,
hazardous reagents
Substrates: aromatic aldehydes, ethyl
bromoacetate, Zn–NH4 Cl. Catalyst:
Zn–NH4 Cl. Environmentally benign,
solid-state reaction, grinding, yields
above 90%
Pinacol coupling Substrates: aromatic aldehydes,
aromatic ketones, THF. Catalyst: Zn,
X CHO ZnCl2 . Disadvantages: catalyst types,
solvents, hazardous reagents in the
Zn-ZnCl2
preparation of catalysts
solid Substrates: aromatic aldehydes,
aromatic ketones. Catalyst: Zn, ZnCl2 .
Environmentally benign, reusable
X CH2OH catalyst, reasonable yields, selective
α-glycols synthesis from aromatic
+ ketones

X CH CH X

OH OH

Zn-ZnCl2
ArCOAr1 ArAr1 C CArAr1C
solid
OH OH

(Continued)
316 9 Present Status and Future Outlook

Table 9.1 (Continued)

Reactions Classical/solvent-free conditions

Benzilic acid rearrangement Substrates: benzil derivatives, alkali


O O 1
metal hydroxide, aqueous organic
KOH –Ar solvents, ethanol. Catalyst: none.
CAr CAr1 Ar COOH Disadvantages: heating, economic
solid
OH factors, hazardous reagents,
moderate yields
Substrates: benzil derivatives, KOH.
Catalyst: none. Solid state reaction at
80 ∘ C with 90% yield
Photocyclization Substrates: acetophenone, aqueous
organic solvents, methanol. Catalyst:
none. Disadvantages: mixture of cis
OH and trans isomers, economic factors
O hν
Substrates: acetophenone. Catalyst:
none. Solid state photochemical
reaction with 100% selectivity
toward trans isomer
Photodecarbonylation Substrates: cyclopentanone
derivative, organic solvents,
alcohols. Catalyst: none.
O
Ph Ph Ph Disadvantages: 3 : 1 mixture of
Ph Ph Ph 1,1-diphenylethylene and
Ph hν 1,1,4,4-quatrophenyl cyclobutene,
Ph economic factors
Substrates: cyclopentanone
derivative. Catalyst: none. Solid-state
reaction under hν irradiation, with
100% selectivity toward
1,1,4,4-quatrophenyl cyclobutene
Carbohydrate chemistry
Glycosyl azides synthesis Substrates: acetylated compounds,
OAc silver or sodium azides, water,
OAc
NaN3 O organic solvents i.e.
O dimethylformamide (DMF), ionic
8h, 500–600 rpm
AcO AcO liquids, basic media, sodium azide
N3 NaN3 . Catalyst: Lewis acid.
X
X=Br, Cl Disadvantages: poor solubility and
stability of azides, solvents,
economic factors
Substrates: acetobromoglucose,
NaN3. Catalyst: none. Solid state
reaction under mechanochemical
conditions (stainless steel [SS] balls)
with high yield 87–99%
(Continued)
9.1 Summary 317

Table 9.1 (Continued)

Reactions Classical/solvent-free conditions

Complex glycopeptides and ureas Substrates: N,N ′ -carbonyldiimidazole


COOH (CDI), dry organic solvents, such as
CHCl3 , THF, DMF, or toluene under a
CDI (CH2)8 N2 atmosphere, carboxylic acid, sugars,
linker. Catalyst:
COOH 1,8-diazabicyclo[5.4.0]undec-7-ene
(DBU) catalysis. Disadvantages: time
Ball mill, 2 h, 500 rpm consuming, multistep process,
–H2O +CO2 economic factors, expensive and
hazardous coupling agents and solvents
O HN
N
N that cause irreversible environmental
N damage
(CH2)8 Substrates: 1,2-dicarboxylic acid,
N
N galactose, p-aminophenyl linker,
O OAc p-nitrophenol, K2 CO3 . Catalyst: Pd/C
AcO
O hydrogenation catalyst, H-Cube.
O NH2
Ball mill AcO
Environmentally benign, reusable
HN OAc catalyst, high yields, solid-state reaction
N 2 h, 500 rpm
under mechanochemical conditions
(ball milling)
AcO OAc O
O
O NH (CH2)4
AcO OAc 2

AcO OAc O
O
O NH
AcO OAc 2

Drug development, pharmaceutical, and medical application


Organoselenium synthesis Substrates: NaSeH or Na2 Se2 , NaBH4 ,
Se NaOEt, alcohols, alkaline solutions,
Cl THF, LiBHEt3 , NH2 NH2 /PhNH2 NH2 in
+ DMF, NaOH, Na in NH3lq . Catalyst: Pt
N group metal complexes. Disadvantages:
time consuming, multistep process,
rt, N2
economic factors, expensive and
Glycerol
hazardous solvents, reaction
temperature
Se Substrates: N,N-dimethylaniline,
haloselenium compound, glycerol.
Catalyst: none. High yield (99%),
N reaction under an inert atmosphere.
Waste product HCl

(Continued)
318 9 Present Status and Future Outlook

Table 9.1 (Continued)

Reactions Classical/solvent-free conditions

Hydrogenation of benzaldehyde Substrates: benzaldehyde, toluene,


O benzene, phenol, acetic, butyric, or
lauric acids, S- or N-containing solvents,
ethanol, dioxane, n-octane, dodecane.
OH Catalyst: Au, Pd, Ru, Ni, Cu, Pt alumina
+ HO OH based. Disadvantages: High pressure
20–40 bar, alcohol yields 40–70%,
Ru-catalyst economic factors, expensive and
Glycerol, base hazardous reagents and solvents
Substrates: benzaldehyde, glycerol, base.
OH Catalyst: Ru-based. Environmentally
benign, reusable catalyst

O
+ HO OH

Carbonyl protection Substrates: molecule with a carbonyl


O group, 1,2- or 1,3-dithiole, polyethylene
S S glycol, dimethyl acetals, perchloric
Glycerol acid in CH2 Cl2 –H2 O, HBF4 . Catalyst:
+ SH 90 °C SiO2 , hydrides, Fe-, Mn-based.
HS Disadvantages: low yields, economic
factors, expensive, and hazardous
solvents
Substrates: cyclohexanone, 1,2-dithiole,
aldehydes and ketones
(thioacetalization), glycerol. Catalyst:
none. Environmentally benign, reusable
solvent
Dipyrromethane synthesis Substrates: pentanone, HCl, pyrrole.
R2 Catalyst: strong organic acids, i.e.
H2O trifluoroacetic acid (TFA).
O
R1
+ catalyst Disadvantages: time consuming,
NH
NH R1 R2 HCl multistep process, purification required,
HN
economic factors, hazardous solvents
and reagents
Substrates: pyrrole, diethyl ketone,
water. Catalyst: HCl. Environmentally
benign, one-step process, little/no
purification, high yields
(Continued)
9.1 Summary 319

Table 9.1 (Continued)

Reactions Classical/solvent-free conditions

Sitagliptin phosphate synthesis Substrates: asymmetric enamine


F hydrogenation for antidiabetic
F compound sitagliptin. Catalyst: Rh.
O O Disadvantages: yield 97%, time
N N
consuming, multistep process, high
N pressure, purification required (trace
N F amount of Rh)
F3C DHP
i-PrNH2 Substrates: (i) direct amination of
prositagliptin ketone to enantiopure
F sitagliptin followed by (ii) phosphate
F salt formation to get sitagliptin
O NH2 phosphate, dehaloperoxidase (DHP).
N Catalyst: H3 PO4 , i-PrNH2 . Two-step
N
N process, high yields (99.95%), economic
N F factors
F3C
H3PO4

– F
H2PO4 +
F
O NH2
N N
N
N F
F3C

Keto-ester to lactam cyclization Substrates: aldehydes, amines, esters,


O organic solvents, i.e. toluene, binapine,
strong base, H2 . Catalyst: Ni/transition
metal-based catalyst. Disadvantages:
high pressure, toxic solvents, catalyst
+ D-alanine
type, economic factors
EtO O
Substrates: 4-oxo ester, lactate
dehydrogenase, D-alanine. Catalyst:
Transaminaze transaminase. Environmentally benign,
NH pH 7, 30 °C low reaction temperature (30 ∘ C), high
yields 92%
O

(Continued)
320 9 Present Status and Future Outlook

Table 9.1 (Continued)

Reactions Classical/solvent-free conditions

Sertraline synthesis (Zoloft) Substrates: ethyl acetate, ether solution


O of the free base, HCl(g) , sertraline,
NH2Me
aqueous methylamine (NH2 Me), TiCl4 ,
NH2Me THF, toluene, hexene. Catalyst: metal
salts. Disadvantages: economic
factors – unwanted trans-isomer and
waste elimination, expensive and
hazardous catalysts and solvents that
Cl
Cl cause irreversible environmental
Cl damage
Cl
Pd/CaCO3/H2
EtOH Substrates: sertraline, aqueous
methylamine (NH2 Me), ethanol.
NH2Me Catalyst: Pd/C, Pd/CaCO3 , H2 .
NH2Me
Environmentally benign, reusable
catalyst, high yields, cis/trans product
EtOH
ratio = 20 : 1
D-mandelic
acid
Cl
Cl
Cl
Cl

Paroxetine synthesis (Paxil, Seroxat) Substrates: coupling reaction of 3- and


F 4-substituted bromocyclohexane
F MeO2C CO2Me derivatives i.e.
MeO2C CO2NH2 4-bromo-N-Boc-piperidine
(Boc = tert-butoxycarbonyl), alkyl
MeO2C CO2Me bormide, THF, PhMgBr, ArMgBr,
CHO tetramethylethylenediamine (TMEDA),
O N O
hexamethylenetetramine (HMTA),
toluene, DMF, NaOH. Catalyst: Co,
Protease Fe-based. Disadvantages: time
consuming, multistep process,
economic factors, expensive, and
F hazardous coupling agents and solvents
that cause irreversible environmental
damage. Yield 51–76%
HO2C CO2Me Substrates: salt of an arecoline
derivative, 4-fluorophenylmagnesium
O N O halide, racemic cis/trans piperidine
ester (after extraction), strong base,
hydride reducing agent, trans carbinol.
–CO2 Catalyst: protease enzyme.
F
F Environmentally benign, short time of
O synthesis, more cost-efficient process,
Demethylation
O O Global reduction enzymes as catalyst, high yields
Etheryfication (doubled to solvent-based synthesis)
HO2C CO2Me
N
H
O N O

(Continued)
9.1 Summary 321

Table 9.1 (Continued)

Reactions Classical/solvent-free conditions

Pregabalin synthesis (Lyrica) Substrates: isovaleraldehyde, diethyl


O O malonate, potassium cyanide,
O O
EtO OEt (S)-mandelic acid, THF, isopropanol;
EtO OEt isopropyl alcohol (IPA), KOH, MeOH,
Rasemization
NC
NC EtOH. Catalyst: Raney nickel.
Disadvantages: time consuming,
multistep process and a long synthesis
for a relatively simple molecule,
Lipase economic factors, expensive chiral
auxiliary, hazardous azide chemistry,
O O low temperature (−78∘ C)
HO OH Substrates: cyano diester, H2 O, KOH,
NC IPA, Ca(OAc)2 . Catalyst:
enzyme – lipase (Lipolase), Ni/H2 .
Environmentally benign, reduced (up to
7 times less) chemical usage, reusable
H2O
catalyst, high yields
O
OH
KOH O
NC
Ni/H2 HO
NH2

Source: Adapted from [1, 5–12, 68–71].

thus all humanity in the long term [63, 72]. Different catalytic systems guide the
ethanol conversion by different routes toward various products [5, 8]. The design of
catalysts for this process should consider the complexity of reaction pathways such
as dehydration, dehydrogenation, aldolization, and Tishchenko reaction, which can
be controlled by the tailored design of the surface properties.
Currently, MCH is not yet ready to fully replace solution-based synthesis meth-
ods. However, it is undoubtedly a powerful and eco-friendly tool that should be
investigated and better exploited in the coming years. Fundamental studies on the
mechanochemical synthesis of nanomaterials, such as metal oxides and composite
materials, are vital. It is believed that novel sorbents, catalysts, and other nanoma-
terials and practical technologies utilizing MCH will emerge in the near future in
strong relation with in situ operando characterization [63, 72].
The solvent role in organic transformation is vast, ranging from (i) providing a
reaction medium for the efficient collision of reactant molecules to the (ii) solva-
tion of reactants, (iii) affecting transition state and (iv) products, and (v) activation
energy reduction [5, 6]. Therefore, the solvent/solvent-type plays a crucial role in
each step of a given chemical reaction. Moreover, the change in the solvent can affect
the reaction rate and/or pathway and, consequently, results in a change in overall
conversion, yield, and targeted product distribution. Even though the solvent plays
an essential role in reaction, it does not become an integral part of the product and
322 9 Present Status and Future Outlook

can be recovered after reaction completion. However, taking the green chemistry
principles into account as well as sustainable development and process economy, it
is accepted that the best solvent is " “no solvent.”

9.2 Future Outlook


Solvent-free techniques represent a clean, economical, efficient, and safe procedure
that can lead to substantial savings in money, time, and products. They can be effi-
ciently coupled to nonclassical methods of activation that include ultrasound and
microwaves [5–7, 40, 50, 67]. Further challenges in MCH, including improvements
related to scalability, quantitativeness, and greater control over the reaction condi-
tions, can also profit in reaction economy as well as in whole process development
sustainably [20].
Solvent-free or synthetic organic protocols applying greener and more efficient
methodologies are crucial to developing more sustainable processes in many indus-
trial sectors. Recent advances in organic chemistry, various substances synthesis,
solid-state chemistry have shown that solvents are not always necessary for perform-
ing chemical reactions effectively [9, 11]. In Table 9.1, it has already been shown
that a wide variety of essential chemical syntheses and molecule transformations
are possible in the absence of any solvent. The reaction can be carried out often in
the solid state, implying shaking, milling, acoustic waves, etc., instead of thermal
energy. Moreover, some of the described reactions (Table 9.1) can proceed spon-
taneously between solid reactants applying mechanochemical protocols, i.e. upon
milling while water and/or organic traces, vapors, and catalyst addition facilitate the
reaction initiation and govern the overall process selectivity and yield. As shown in
Table 9.1, these innovative synthesis procedures and chemistry-based strategies are
essential for green pharmaceutical development to produce and deliver drugs in a
safer, more economical, and healthy/environmentally benign way [12, 69]. Herein,
the open issue is to rethink the technological steps and bring together chemical
innovations to channel complex, waste-generating pharmaceutical transformations
into simple and economically viable processes at production scale and meet
current global challenges (water pollution and extensive environmental utilization,
zero-waste policy, circular economy). Several industries have already proposed
that it is possible to provide a more sustainable chemical pathway to important
therapies (synthesis of Sitagliptin, Zoloft R© , Paxil, or Lyrica synthesis, Table 9.1)
[9, 11, 69]. Green solvents, biocatalysts – enzymes, and recyclable nanocatalysts
offer many opportunities for greener methods to reduce pharmaceutical products’
environmental impact and cost.
Another crucial issue is connected with efficient water removal from the reaction
environment [1, 73–75], e.g. during ether synthesis, water inhibits active sites for
etherification, thus lowering the etherification rate and selectivity. Therefore, find-
ing an alternative pathway or reaction protocols coupled with the green chemistry
method can result in reaction rate enhancement and improvement of ether yields.
Further investigations on tuning catalyst properties, morphology composition, and
References 323

structure into direct biomass-derived ether synthesis processes may reflect a promis-
ing sustainable pathway for renewable ether synthesis. An industrial opportunity
can also decrease the number of transition stages of the overall process, opening
other possibilities for direct ether synthesis from biomass-derived molecules [34] or
relevant chemicals, fuels, or drop-in-fuels in benign environmental processes follow-
ing green chemistry principles.

Acknowledgments
This research was supported by funding from the Foundation for Polish Science
(POWROTY/2016-1/5). I.S.P. extends his sincere appreciation to the NAWA
(The Polish National Agency for Academic Exchange) through Bekker grants
PPN/BEK/2019/1/00348 “C1-C4 alkanes to oxygenated fuel electrochemical
transformation.”

References

1 Marvaniya, H.M., Kaumil, K.N., Modi, N., and Sen, D.J. (2011). Greener reactions
under solvent free conditions. Int. J. Drug Dev. Res. 3 (2): 334–343. ISSN
0975-9344.
2 Gawande, M.B., Goswami, A., Felpin, F.-X. et al. (2016). Cu and Cu-based
nanoparticles: synthesis and applications in catalysis. Chem. Rev. 116: 3722–3811.
3 Xu, C., De, S., Balu, A.M. et al. (2015). Mechanochemical synthesis of advanced
nanomaterials for catalytic applications. Chem. Commun. 51: 6698–6713.
4 Rodríguez-Padrón, D., Balu, A.M., Romero, A.A., and Luque, R. (2017).
New bio-nanocomposites based on iron oxides and polysaccharides applied to
oxidation and alkylation reactions. Beilstein J. Org. Chem. 13: 1982–1993.
5 Baiga, R.B.N. and Varma, R.S. (2013). Solvent-free synthesis. In: An Introduction
to Green Chemistry Methods, Future Science Book Series (ed. R. Luque and J.C.
Colmenares), 18–38. Future Science Ltd.
6 Enache, D.I., Edwards, J.K., Landon, P. et al. (2006). Solvent-free oxidation of
primary alcohols to aldehydes using Au–Pd/TiO2 catalysts. Science 311: 362–365.
7 Loupy, A. (1999). Solvent-free reactions. Top. Curr. Chem. 206: 155–207.
8 Malipedi, H., Das, P., and Karigar, A. (2011). Green technique-solvent free
synthesis and its advantages. Int. J. Ayurveda Res. 2 (4): 1079–1086.
9 Burrows, J., Kamo, S., and Koide, K. (2021). Scalable Birch reduction with
lithium and ethylenediamine in tetrahydrofuran. Science 374: 741–746.
10 Pieta, I.S., Ishaq, M., Wells, R.P.K., and Anderson, J.A. (2010). Quantitative
determination of acid sites on silica–alumina. Appl. Catal., A 390: 127–134.
11 Tanaka, T. and Toda, F. (2000). Solvent-free organic synthesis. Chem. Rev. 100:
1025–1074.
12 Koenig, S. (2013). Scalable Green Chemistry. Case Studies from the
Pharmaceutical Industry. Boca Raton: Taylor & Francis Group, CRC Press.
324 9 Present Status and Future Outlook

13 Intasard, S.G., Imwiset, K., Bureekaew, S., and Ogawa, M. (2018). Mechano-
chemical methods for the preparation of intercalation compounds, from
intercalation to the formation of layered double hydroxides. Dalton Trans. 47:
2896–2916.
14 Cavani, F., Trifiro, F., and Vaccari, A. (1991). Hydrotalcite-type anionic clays:
preparation, properties, and applications. Catal. Today 11: 173–301.
15 Hernández, W.Y., Lauwaert, J., Van Der Voort, P., and Verberckmoes, A. (2017).
Recent advances on the utilization of layered double hydroxides (LDHs) and
related heterogeneous catalysts in a lignocellulosic-feedstock biorefinery scheme.
Green Chem. 19: 5269–5302.
16 Markova, Z., Novak, P., Kaslik, J. et al. (2014). Eco-friendly magnetic iron
oxide-pillared montmorillonite for advanced catalytic degradation of dichloro-
phenol. ACS Sustainable Chem. Eng. 2: 1674.
17 Polshettiwar, V. and Varma, R.S. (2010). Green chemistry by nano-catalysis.
Green Chem. 12: 1540.
18 Virkutyte, J., Baruwati, B., and Varma, R.S. (2010). Visible light induced photo-
bleaching of methylene blue over melamine-doped TiO2 nanocatalyst. Nanoscale
2: 1109.
19 Rahmanivahid, B., Pinilla-de Dios, M., Haghighi, M., and Luque, R. (2019).
Mechanochemical synthesis of CuO/MgAl2 O4 and MgFe2 O4 Spinels for vanillin
production from isoeugenol and vanillyl alcohol. Molecules 14: 2597–2613.
20 Ralphs, K., Hardacre, C., and James, S.L. (2013). Application of heterogeneous
catalysts prepared by mechanochemical synthesis. Chem. Soc. Rev. 42: 7701–7718.
21 Lewalska-Graczyk, A., Pieta, P., Garbarino, G. et al. (2020). Graphitic carbon
nitride–nickel catalyst: from material characterization to efficient ethanol
electrooxidation. ACS Sustainable Chem. Eng. 18: 7244–7255.
22 Mazonde, B., Cheng, S., Zhang, G. et al. (2018). A solvent-free in situ synthesis
of a hierarchical co-based zeolite catalyst and its application to tuning
Fischer–Tropsch product selectivity. Catal. Sci. Technol. 8: 2802–2808.
23 Gottuso, A., Kochritz, A., Saladino, M.L. et al. (2020). Catalytic and photo-
catalytic epoxidation of limonene: using mesoporous silica nanoparticles as
functional support for a Janus-like approach. J. Catal. 391: 202–211.
24 Saladino, M.L., Spinella, A., Caponetti, E., and Minoja, A. (2008). Characteri-
zation of Nd-MCM41 obtained by impregnation. Microporous Mesoporous Mater.
113: 490–498.
25 Kraleva, E., Saladino, M.L., Spinella, A. et al. (2011). H3 PW12 O40 supported on
mesoporous MCM-41 and Al-MCM-41 materials: preparation and characteri-
zation. J. Mater. Sci. 46 (22): 7114–7120.
26 Saladino, M.L., Kraleva, E., Todorova, S. et al. (2011). Synthesis, characterization
and catalytic activity of mesoporous Mn-MCM-41 materials. J. Alloys Compd.
509: 8798–8803.
27 Yuan, N., Majeed, M.H., Bajnoczi, E.G. et al. (2019). In situ XAS study of the
local structure and oxidation state evolution of palladium in a reduced graphene
oxide supported Pd(II) carbene complex during an undirected C–H acetoxylation
reaction. Catal. Sci. Technol. 9: 2025–2031.
References 325

28 Lukashuk, L., Yigit, N., Rameshan, R. et al. (2018). Operando insights into CO
oxidation on cobalt oxide catalysts by NAP-XPS, FTIR, and XRD. ACS Catal. 8
(9): 8630–8641.
29 Li, X., Wang, H.-Y., Yang, H. et al. (2018). In situ/operando characterization
techniques to probe the electrochemical reactions for energy conversion. Small 2:
1700395.
30 Peng, L. and Wei, Z. (2020). Catalyst engineering for electrochemical energy
conversion from water to water: water electrolysis and the hydrogen fuel cell.
Engineering 6: 653–679.
31 Do, J.-L. and Friscic, T. (2017). Mechanochemistry: a force of synthesis. ACS
Cent. Sci. 3 (1): 13–19.
32 Ciriminna, R., Fidalgo, A., Meneguzzo, F. et al. (2019). Vanillin: the case for
greener production driven by sustainability megatrend. ChemistryOpen 6:
658–658.
33 McCormick, P.G., Tsuzuki, T., Robinson, J.S., and Ding, J. (2001). Nanopowders
synthesized by mechanochemical processing. Adv. Mater. 13: 1008–1010.
34 Yan, S., Wang, C., Hu, H. et al. (2020). Mechanochemical preparation of a
H3 PO4 -based solid catalyst for heterogeneous hydrolysis of cellulose. ACS Omega
5 (46): 29971–29977.
35 Kraleva, E., Saladino, M.L., Matassa, R. et al. (2011). Phase formation in mixed
TiO2 –ZrO2 oxides prepared by sol–gel method. J. Struct. Chem. 52 (2): 340–348.
36 Doustkhah, E., Assadi, M.H.N., Komaguchi, K. et al. (2021). In situ blue titania
via band shape engineering for exceptional solar H2 production in rutile TiO2 .
Appl. Catal., B 297: 120380.
37 Kraleva, E., Spojakina, A., Saladino, M.L. et al. (2010). Mechanical treatment of
TiO2 and ZrO2 oxide mixtures. J. Nanosci. Nanotechnol. 10: 8417–8423.
38 Wang, X., Yu, J., Chen, Y. et al. (2006). ZrO2 -modified mesoporous nanocrys-
talline TiO2−x Nx as efficient visible light photocatalysts. Environ. Sci. Technol. 40:
2369–2383.
39 Villegas, J., Giraldo, O., Laubernds, K., and Suib, S.L. (2003). New layered
double hydroxides containing intercalated manganese oxide species: synthesis
and characterization. Inorg. Chem. 42: 5621–5631.
40 Chitrakar, R., Tezuka, S., Sonoda, A. et al. (2007). A solvent-free synthesis of
Zn–Al layered double hydroxides. Chem. Lett. 36: 446–447.
41 Thomas, N. (2012). Mechanochemical synthesis of layered hydroxy salts. Mater.
Res. Bull. 47: 3568–3572.
42 Li, M., Liu, D., Chen, X. et al. (2021). Radical-driven decomposition of graphitic
carbon nitride nanosheets: light exposure matters. Environ. Sci. Technol. 55 (18):
12414–12423.
43 Spojakina, A.A., Kraleva, E.Y., and Jiratova, K. (2010). Heteroatom effect on
hydrodesulphurization activity of TiO2 -supported molybdenum heteropolyox-
ometalates of Anderson type. Kinet. Catal. 51: 385–393.
44 Franco, A., De, S., Balu, A.M. et al. (2018). Integrated mechanochemical/
microwave-assisted approach for the synthesis of biogenic silica-based Catalysts
from Rice husk waste. ACS Sustainable Chem. Eng. 6: 11555–11562.
326 9 Present Status and Future Outlook

45 Baiga, R.B.N. and Varma, R.S. (2012). Alternative energy input: mechano-
chemical, microwave and ultrasound-assisted organic synthesis. Chem. Soc.
Rev. 41: 1559–1584.
46 Caponetti, E., Martino, D.C., Leone, M. et al. (2006). Microwave-assisted
synthesis of anhydrous CdS nanoparticles in water in oil microemulsion.
J. Colloid Interface Sci. 304: 413–418.
47 Rizzuti, A., Leonelli, C., Corradi, A. et al. (2009). Structural characterization
of zirconia nanoparticles prepared by microwave-hydrothermal synthesis.
J. Dispersion Sci. Technol. 30: 1511–1516.
48 Saladino, M.L., Nasillo, G., Martino, D.C., and Caponetti, E. (2010). Synthesis
of Nd:YAG nanopowder using the citrate method with microwave irradiation.
J. Alloys Compd. 491: 737–741.
49 Cook, D.S., Hooper, J.E., Dawson, D.M. et al. (2020). Synthesis and polymor-
phism of mixed aluminum−gallium oxides. Inorg. Chem. 59: 3805–3816.
50 Crawford, D.E. (2017). Solvent-free sonochemistry: sonochemical organic synthe-
sis in the absence of a liquid medium. Beilstein J. Org. Chem. 13: 1850–1856.
51 Abboudi, M., Messali, M., Kadiri, N. et al. (2011). Synthesis of CuO, La2 O3 ,
and La2 CuO4 by the thermal-decomposition of oxalates precursors using a new
method. Synth. React. Inorg. Met. -Org. Chem. 41: 683–688.
52 Rizzuti, A., Leonelli, C., Corradi, A. et al. (2009). Structural characterization
of zirconia nanoparticles prepared by microwave-hydrothermal synthesis.
J. Dispersion Sci. Technol. 30: 487–491.
53 Cao, S., Tao, F., Tang, Y. et al. (2016). Size- and shape-dependent catalytic
performances of oxidation and reduction reactions on nanocatalysts. Chem.
Soc. Rev. 45: 4747–4765.
54 Saberi, F., Rodríguez-Padrón, D., Doustkhah, E. et al. (2019). Mechanochemi-
cally modified aluminosilicates for efficient oxidation of vanillyl alcohol. Catal.
Commun. 118: 65–69.
55 Meng, J., Liu, X., Niu, C. et al. (2020). Advances in metal–organic framework
coatings: versatile synthesis and broad applications. Chem. Soc. Rev. 49:
3142–3186.
56 Rodríguez-Padrón, D., Puente-Santiago, A.R., Balu, A.M. et al. (2017). Solventless
mechanochemical preparation of novel magnetic bioconjugates. Chem. Commun.
53: 7635–7637.
57 Banadaki, A.D. and Kajbafvala, A. (2014). Recent advances in facile synthesis of
bimetallic nanostructures: an overview. J. Nanomaterials 28: 1–29. http://doi.org/
10.1155/2014/985948.
58 Gu, J., Zhang, J.-W., and Tao, F. (2012). Shape control of bimetallic nanocatalysts
through well-designed colloidal chemistry approaches. Chem. Soc. Rev. 41:
8050–8065.
59 Kuna, E., Mrdenovic, D., Jönsson-Niedziółka, M. et al. (2021). Bimetallic
nanocatalysts supported on graphitic carbon nitride for sustainable energy
development: the shape-structure–activity relation. Nanoscale Adv. 3: 1342–1351.
60 Pieta, I.S., Epling, W.S., Kazmierczuk, A. et al. (2018). Waste into fuel-catalyst
and process development for MSW valorisation. Catalysts 8: 113–128.
References 327

61 Pieta, I.S., Kadam, R.G., Pieta, P. et al. (2021). The hallmarks of copper single
atom catalysts in direct alcohol fuel cells and electrochemical CO2 fixation. Adv.
Mater. Int. 8: 2001822.
62 Pieta, I.S., Lewalska-Graczyk, A., Kowalik, P. et al. (2021). CO2 hydrogenation
to methane over Ni-catalysts: the effect of support and Vanadia promoting.
Catalysts 11: 433.
63 Pieta, I.S., Michalik, A., Kraleva, E. et al. (2021). Bio-DEE synthesis and dehy-
drogenation coupling of bio-ethanol to bio-butanol over multicomponent mixed
metal oxide catalysts. Catalysts 11: 660.
64 Pieta, I.S., Rathi, A., Pieta, P. et al. (2019). Electrocatalytic methanol oxidation
over Cu, Ni and bimetallic Cu–Ni nanoparticles supported on graphitic carbon
nitride. Appl. Catal., B 244: 272–283.
65 Borges, M.E. and Díaz, L. (2012). Recent developments on heterogeneous
catalysts for biodiesel production by oil esterification and transesterification
reactions: a review. Renewable Sustainable Energy Rev. 16: 2839–2849.
66 Bolm, C. and Hernandez, J.G. (2018). From synthesis of amino acids and
peptides to enzymatic catalysis: a bottom-up approach in mechanochemistry.
ChemSusChem 11: 1410–1420.
67 Gómez, J.L., Bastida, J., Máximo, M.F. et al. (2011). Solvent-free polyglycerol
polyricinoleate synthesis mediated by lipase from Rhizopus arrhizus. Biochem.
Eng. J. 54: 111–116.
68 Crowe, D., Ward, N., and Wells, A. S. (1999). Process for the preparation of
paroxetine. Smithkline Beecham PLC. International patent WO2001029032A1.
69 Menges, N. (2017). The role of green solvents and catalysts at the future of drug
design and of synthesis. Intechopen https://doi.org/10.5772/intechopen.71018.
70 Kumar, V., Kumar Giri, S., Venugopalan, P., and Kartha, R. (2014). Synthesis
of cross-linked glycopeptides and ureas by a mechanochemical, solvent-free
reaction and determination of their structural properties by TEM and X-ray
crystallography. ChemPlusChem 79: 1–10.
71 Bhuyan, D., Sarm, R., Dommarajua, Y., and Prajapati, D. (2014). Catalyst- and
solvent-free, pot, atom and step economic synthesis of tetrahydroquinazolines by
an aza-Diels–Alder reaction strategy. Green Chem. 16: 1158.
72 Saladino, M.L., Chillura Martino, D., Kraleva, E., and Caponetti, E. (2013). Effect
of the cerium loading on the HMS structure. Its catalytic performance in the
reaction of partial oxidation of bio-ethanol. Catal. Commun. 36: 10–15.
73 Jean, J.M.R., Gallo, M.R., Bueno, J.M.C., and Schuchardt, U. (2014). Catalytic
transformations of ethanol for biorefineries. J. Braz. Chem. Soc. 25: 2229–2243.
74 Karatzos, S., McMillan, J., and Saddler, J. (2019). The Potential and Challenges of
“Drop in” Biofuel. International Energy Agency Bioenergy. IEA Bioenergy. ISBN:
978-1-910154-61-8.
75 Matsakas, L., Gao, Q., Jansson, S. et al. (2017). Green conversion of municipal
solid wastes into fuels and chemicals. Electron. J. Biotechnol. 26: 69–83.
329

Index

a ammonium persulfate 171


acetalization of aldehydes 194 aromatic amines, solvent-free
acetalization glycerol 130 diazotization-halogenation 291
acetic anhydride (AA) 108, 131, 141, aromatic compounds, solvent-free
147, 169, 184 hydroxylation 290
acetohydroxyacid synthase (AHAS) 92 aryl aldehyde 15, 72, 188, 238, 244, 283
acetylation reactions 98, 183–184 aryne formation 166
acid-catalyzed Biginelli reaction 193 atmospheric oxygen 182
acrylonitrile derivatives 200 atomic layer deposition (ALD) 8, 59–63
activated carbons (ACs) 182–185 Au-anchored on g-C3 N4 nanosheets
acylation of aldehydes 194
233
adsorption-reduction method 197
Au1 /g-C3 N4 catalyst 233, 235, 236
aerosil silica-supported acidic IL 285
AuNPs/g-C3 N4 catalyst 235, 236
alcohols
esterification 195–197
b
oxidation 37, 182, 195, 197, 203,
ball milling 6, 10, 17, 33–38, 64, 70, 71,
209–210, 310
73, 75–77, 79, 142
aldol condensation 70, 173, 186, 204
benzaldehyde 14, 28, 72, 75, 81, 82, 94,
alkylaminophenols derivatives 201–202
95, 97, 106, 107, 125, 130, 141, 149,
aluminum nitride (AlN)
166, 167, 169, 170, 173, 191, 194,
properties 249
aluminum oxide (Al2 O3 ) 15, 123, 250 196, 198, 203, 204, 210, 239, 244,
amidoalkyl naphthols 188–190, 282 278, 280, 290
1-amidoalkyl naphthol 282–283 benzoxanthenes 276
amines benzyl alcohols 148, 192, 196, 210, 293
coupling 210–213 aerobic oxidative esterification 195,
N-formylation 291, 292 196
2-amino-3-cyano-4H-pyran 279 oxidation 6, 125, 133, 135, 171, 195,
α-aminophosphonates 197–199, 204, 209, 290
AlN-catalyzed solvent-free synthesis benzylamine, oxidative coupling
250 249
synthesis 106, 136, 250, 284 benzylic alcohols oxidation 203–204
α/β-C3 N4 nanowires, band gap 232 BeW12 O40 ILMLMNPs catalysts 278
Solvent-Free Methods in Nanocatalysis: From Catalyst Design to Applications, First Edition.
Edited by Rafael Luque, Manoj B. Gawande, Esmail Doustkhah, and Anandarup Goswami.
© 2023 WILEY-VCH GmbH. Published 2023 by WILEY-VCH GmbH.
330 Index

bifunctional nucleophile–electrophile catalysis 4, 13, 14, 16, 31, 40, 44, 48, 51,
catalyst 236 121, 126, 127, 132, 135, 138, 163,
Biginelli reaction 13, 15, 72, 126, 169, 171, 181
141, 146, 186, 192–193, catalysts, in organic synthesis 261
280–282 chemical vapor deposition (CVD) 8, 59,
bioactive N-aryl oxazolidin-2-ones 63–64, 246
synthesis 292 chemisorption method 264, 270
choline sulfate ionic liquid-supported
bisphenolic antioxidants 199
g-C3 N4 (g-C3 N4 @SO3 Ch) 244
bisvinylimidazolium functionalized
Claisen–Schmidt condensation reaction
pristine carbon nanohorns
14, 125, 134, 149
(bVImi-CNHs) 288
classical heating 69, 91–103
boron doping in g-C3 N4 234
coaled carbon-based solid acid (CCBSA)
boron nitride (BN)
catalyst 192, 193
nanomaterials 245 cold-plasma methods 8, 56
boron nitride nanosheets (BNNSs) cold thermal plasma method 56–59
chemical vapor deposition 246 condensation reactions 285
high surface area and recycling stability aldol 70, 173, 186, 204
246 Claisen–Schmidt 14, 125, 134, 149
mechanical cleavage/solvent exfoliation Friedländer condensation 285–286
246 Hantszch 188
borono-Mannich reaction 86 isatin 124
Bose–Einstein condensate (BEC) states Knoevenagel 20, 92, 172, 204, 286
52 Co3 O4 /TSCN catalyst 237, 238
bottom-up approaches 4, 229 copolymer-supported Pd nanocatalyst
bromofunctionalization 77, 78 274
Brønsted acid catalysts 190 co-precipitation (CP) method 40, 121,
Brønsted acidic 1-methyl-3- 124, 127, 136, 137
(3-trimethoxysilylpropyl) core-shell structure 31, 126, 143, 147,
205
imidazolium hydrogen sulfate IL
cross-aldol condensation, of ketones with
285
aromatic aldehydes 185–188
Brønsted acidic N-propyl-2-pyrrolidinium
cross-linked polymer-supported
hydrogen sulfate IL 291
amine-functionalized IL 274
crystalline solids 266
c cubic boron nitride (c-BN) 245
carbamates 203, 294
Cu2 O modified g-C3 N4 238
carbon nanotubes (CNTs) 8, 52, 182, cyclic carbonates 17, 236, 243, 288,
195–200, 268–269 289
carbon nitride (CN) cycloaddition
characteristics 229 of CO2 and styrene oxide 236
crystalline 230 of epoxides 240
carbonyl compounds, solvent-free cyclohexane oxidation 17, 184, 185,
continuous flow cyanosilylation 241
295 cyclohexene oxidation 205–207
Index 331

d Fe3 O4 @PIL acidic catalyst 293


DAIL@SiO2 catalysts 278 Fe3 O4 /SiO2 /salen/Mn-supported IL
deionized (DI) water 39, 43, 47, 56, 208 containing tungstophosphoric acid
diazotization 291 283
dibenzoxanthenes and xanthenes 192 ferric hydrogen sulfate (FHS) 149
β-dicarbonyl compounds 141 ferrocene (Fc) 64, 107
1,2-dihydro-1-arylnaphtho[1,2-e][1,3] fibrous nanosilica-supported
oxazin-3-one derivatives 242, 243 nano-Ni@Pd-containing IL
1,4-dihydropyridines (1,4-DHPs) 174, (KCC-1/IL/Ni@Pd) 289
186, 187, 279, 280 field-emission scanning electron
dihydropyrimidine derivatives 239, 244 microscopy (FE-SEM) 58
dihydropyrimidinone (DHPMs) 15, 72, fluorescence proteins (FPs) 92
134, 136 Friedel–Crafts (FC) benzylation, of arenes
dihydropyrimidinone compounds 292, 293
192–193 Friedländer condensation 285–286
3,4-dihydropyrimidin-2(1H)-ones/thiones fused heterocycles 10, 73, 104
281
2,3-dihydroquinazolin-4(1H)-one g
derivatives 142, 243 γ-valerolactone (GVL) 126
2,3-dihydroquinazolinones, solvent-free g-C3 N4 -based catalysts 231
multicomponent synthesis 283 g-C3 N4 /CuO nanocomposites 234
2,3-dihydroquinazolin-4(1H)-ones g-C3 N4 -KOH 237
synthesis 285 g-C3 N4 -NaOH 237
direct reductive amination (DAR) 191 Gewald reaction 125, 136
donor-π-acceptor system (D-π-A) 99 graphene oxide (GO) 182, 200–208
graphene oxide-supported
e heteropolyacid-based IL
electron paramagnetic resonance (EPR) heterogeneous catalyst
195, 196 (PW@IL-GO) 290
electrophilic substitution reaction 167 graphite-like carbon nitride synthesis
epimerization 83 234
epoxides with arylamines, ring-opening graphitic carbon nitride (g-C3 N4 ) 36, 229
reaction of 242 Cu2 O modified 238
esterification 204, 286–287 doping 234
of alcohols 195–197 polytriazine-based 230
of benzyl alcohol 195, 196 melamine 232
ethyl chloroacetate 166 graphitic CN-based materials 236
ethyl levulinate (EL) 147 green chemistry 1–5, 32, 163, 181, 182,
204, 213, 272, 280, 290, 295, 305,
f 310, 322, 323
Fe2 O3 based catalyst/support 148–149 green combustion method 124
Fe3 O4 based catalyst/support 143–148 grinding technique 70, 74–76
Fe3 O4 -g-C3 N4 heterogeneous catalyst Groebke–Blackburn–Bienayme reaction
242 97
Fe3 O4 @g-C3 N4 -K catalyst 240 guanidine-based IL (SH-IL-Pd) 273
332 Index

h imidazolium bromide-functionalized
Hantszch condensation reaction 186, Mn(III)-porphyrin metal-organic
188, 280 framework 289
Hantzsch reaction 125, 279–280 imidazolium functionalized single-walled
Heck reaction, solvent-free 273–275 carbon nanotubes (Im-SWCNT)
helium ion microscopy 122 288
henna-based benzochromenes, imidazopyridines 74, 109
solvent-free synthesis 278 indium nanoparticle 143
heterocyclic compounds synthesis 279 inductively coupled plasma-mass
heterogeneous carbon nitride-based spectrometry (ICP-MS) 63
organocatalyst 240 ionic liquids
non-volatile nature 263
heterogeneous catalysts 4, 14, 122–124,
properties 262
150, 169, 183, 199, 209, 231, 242,
ionic liquids-based materials 18–20
243, 251, 261, 273, 276, 280, 289,
ionic segment, of SILs 265
292
iron-oxide based catalysts 142–149
heteropolyanion-based sulfonated IL
iron oxide nanoparticles 15, 143, 269,
functionalized mesoporous
309
copolymer (P(VB-VMS)PW) 293
isatin condensation reaction 124
hexagonal born nitride (h-BN) 17, 245
hexagonal MoN structure 247
k
high-resolution transmission electron
Kabachnik–Fields reaction 250
microscopy (HR-TEM)
KA oil 184, 185
33, 122
Knoevenagel condensation reaction 12,
homogeneous catalysts 4, 21, 22, 200,
20, 21, 72–74, 92, 94, 95, 130, 172,
209, 261–263, 267
204, 286
immobilization of 262
hydrazide oxidation 207–208 l
hydrocarbons oxidation 20, 209–210 layered double hydroxide (LDH) 39, 308
hydrogen evolution reaction (HER) 38, Lewis acidic chloroaluminate-based
46, 47, 50, 53 pyridinium IL 277
hydrolyzed CN (HCN) 239 liquid assisted grinding (LAG) 76
hydrophilic aminofunctionalized IL
(NH2-IL) 290 m
hydrosilylation, of alkynes 235, 236 magnetically retrievable
β-hydroxy nitriles 240 organic–inorganic nanohybrid
catalyst 280
i magnetic g-C3 N4 242
IL functionalized mesoporous silica NPs magnetic NP-supported ILs (MNP-ILs)
294 269
IL-modified SBA-15 catalyst 274, 275 magnetic urea-based nanocatalyst 276
IL-support interaction types 264 maleic anhydride, solvent-free
imidazoles, substituted 168, 188 diesterification 287
imidazolinones 92 manganese oxide (MnOx ) 14, 132, 135
Index 333

manganese oxides (MnOx )-based catalysts multicomponent reactions (MCRs) 10,


132–135 17, 21, 71, 90, 127, 136, 137, 142,
manual grindings (MG) 33 143, 149, 164, 166, 186, 199, 242,
mechanical grinding 142 272, 275–284, 310
mechanochemistry (MCH) 32, 33, 70,
80, 306, 308, 321, 322 n
ball milling 33–38 nanoarchitectonics 213
mortar and pestle 38–44 nanocatalysts (NCs) 1, 3, 4, 10, 14, 15,
melem 230–233 20, 21, 31–65, 205, 283, 309, 322
melon 230, 231 carbon-based 13, 16, 17
mesoporous carbon catalysts 211, 212 silica-based 16
mesoporous polyionic liquid (MPIL) solvent-free synthesis 6–8, 21, 22
solid-base catalyst 286 nanomaterials (NMs) 1–6, 8, 15, 21, 22,
metal ferrites (M-Fe2 O4 ) 149 31, 121, 122, 127, 132, 138, 143,
metal ion adsorption-reduction process 183, 213, 245, 247, 306, 307, 321
198 nano-silica-supported 3-propyl-1-sulfonic
metallophthalocyanine-CN hybrids 236
acid imidazolium chloride IL 282
metal oxides (MOs) 6, 31, 44, 48, 65,
nanostarch supported imidazolium-based
121–124, 127, 135, 141, 143, 150,
IL 294
168, 247, 307–309, 321
naphthopyranopyrimidine 131
methyl tert-butyl ether (MTBE) 200
N-arylation reaction 202–203
methyl vinyl ketone (MVK) 167
N-Boc protection of amines 104, 105,
Michael addition/reaction 70, 73, 75, 76,
294
92, 96, 166, 167, 292, 293
n-butBr/mp-C3 N4 242
microwave-assisted synthesis method
N–Co doped carbon nanodisks (N-C NDs)
52, 309
46
microwave heating energy 51–52
nitride-based materials 17–18
microwave irradiation 51, 69, 70, 80–91,
nitronium ion 164, 190
138, 149, 163, 166, 168, 187, 188,
201, 309 β-Nitrostyrene 292, 293
MIL-101(Cr)-TSIL catalyst 289 N,N ′ -diaryl-substituted formamidines
mixer ball milling (MM) 33 294
MNP supported acidic IL catalyst 285 non-metal catalyzed asymmetric Michael
MNP-supported Lewis acidic IL 276 addition reaction 292
g-C3 N4 /Cu2 O 234
MoN-based nanocatalyst 248 o
molecular oxygen 125, 185, 204, 206, octahydroquinazolinones 75
209, 211, 240–242, 310 olefins
molybdenum nitrides (MoN) solvent-free hydrogenation 272–273
crystal structures 247 ompg-C3 N4 /SO3 H
PECVD 248 1,2-dihydro-1-arylnaphtho[1,2-e]
mortar and pestle grinding 33, 38–44, [1,3]oxazin-3-ones synthesis 243
64, 65, 70, 74, 75 ordered mesoporous silicas 211, 266–268
chemical fixation of CO2 236 oxazolidinones synthesis 289
solid-base catalyst 286 oxide-based supports 266
334 Index

p polyhydroquinoline compounds 188


Paal-Knorr synthesis, of pyrroles polyhydroquinolines (PHQ) synthesis
276 280
p-CNNs photocatalytic reaction polyimides-based g-C3 N4 , solvent-free
for 12-phenyl-9,9-dimethyl- thermal condensation process
8,9,10,12-tetrahydrobenzo[a] 232
xanthen-11-one 239 polymeric imidazolium IL-functionalized
5-phenyl-1(4-methoxyphenyl)- manganese Schiff base
3[(4-methoxyphenyl)-amino] organometallic catalyst
-1H-pyrrol-2(5H)-one 240 (PIL-SB-Mn(III) 281
p-CNNs photocatalytic synthesis, DHPMs polymeric SILLP based organocatalysts
239 294
PC-PMo-Mel-800 preparation 248, 249 polymer-supported diol-functionalized IL
Pd/BNNS catalyst 246, 247 catalyst (PS-DFILXs) 289
Pd-porphyrin functionalized IL-modified
polymer-supported task-specific ionic
SBA-15 275
liquid 273
Pd-porphyrin functionalized
polysubstituted quinolines 285
imidazolium-based IL 274
polytriazine-based g-C3 N4 230
Pd/rGO aerogels 203
porous carbon materials 208–213
Pd-sepiolite catalyst 273
porous ultrathin CN nanosheets
Pd-soaked supported TSIL 273
(p-CNNs) 239
periodic mesoporous organosilica (PMO)
potassium fluoride (KF) 142
supported IL material 268
propargylamines 137, 238
Petasis-Borono-Mannich (PBM) reaction
pseudo-crystalline solids 266
201
pyran-based hetercycles, synthesis
Petasis reaction 86, 87
phenol derivatives antioxidants 199–200 275–279
12-phenyl-9,9-dimethyl-8,9,10,12- pyrazole oxidation 207–208
tetrahydrobenzo[a]xanthen-11-one pyrimidoindazoles 238
239 pyrroles, Paal-Knorr synthesis 276
phosphorous graphitic carbon nitrides
(P-g-C3 N4 ) 243 q
photo catalytic activity, of Ru-g-C3 N4 quantum effect 50
235 quartz-tube reactor 52
photocatalyzed multicomponent Biginelli quinoline-linked covalent organic
reaction 244 framework 281
photo-Fenton reaction 48 quinoxaline derivatives 124, 241
photoluminescence spectroscopy 122
phthalic anhydride, solvent-free r
diesterification 287 reactive oxygen species (ROS) activation
Pistachio peel 194 101
planetary ball milling (PM) 33 reduced graphene oxide (rGO) 182, 201,
plasma-assisted methods 6, 32, 52–59, 205
64 reductive amination 186, 190, 191
([pmim]FeCl4 /MSNs) nanocatalyst 294 Ru-doped g-C3 N4 234
Index 335

s solid supports, SIL


SBA-15 supported amino-functionalized carbon nanotubes 268–269
IL 290 categorization 266
SBA-supported imidazolium-based IL ordered mesoporous silicas 267–268
278 porosity and pore size 265
SBA-supported multilayered silica coated magnetic nanoparticles
thiazolium-based IL 293 (SMNPs) 269
scanning electron microscopy (SEM) silica gels 267
41, 47–49, 51, 54, 55, 60, 64, solvent-free approach 14, 21, 32, 39, 60,
122, 125 65, 69, 123
SCILL catalytic system 280, 290 solvent-free condensation reactions
[SDAIL]@magnetic nanosilica 287 Friedländer condensation 285–286
silica coated magnetic nanoparticles Knoevenagel condensation 286
169, 269, 275 ethylbenzene oxidation 290
silica gel 20, 163, 168, 266, 267 Heck reaction 273–275
silica gel-supported acidic hydrogenation of olefins 272–273
benzimidazolium-based IL 282 ketalization 130
pyrrolidine-based chiral catalyst 292 solvent-free Mizoroki–Heck reaction of
silica nanoparticles 168–171 bromobenzene 274, 275
silica-supported Brönsted acidic IL solvent-free multicomponent reactions
nanocatalyst 282 1-amidoalkyl naphthol synthesis
silica-supported propyl(N-methyl) 282–283
imidazolium chloride Biginelli reaction 280–282
IL([Sipmim]Cl) catalyst 294 Hantzsch reaction 279–280
SMNP supported acidic N-acyl miscellaneous 283–284
imidazolium chloride IL 280 pyran-based hetercycles 275–279
SMNP-supported acidic pyridinium-based solvent-free oxidation reactions 135,
IL 277 290–291
SMNP supported IL 287 solvent-free photocatalytic H transfer
SMNP supported IL catalysts 291 reaction 235
SMNP-supported imidazolium-based IL solvent-free reactions 31, 69, 103, 121,
292 194, 235, 237, 251, 263, 294, 310,
SMNP supported 311
multi-SO3 H-functionalized silica gel 163–168
Brönsted acidic IL 284 solvent-free synthesis 1–3, 5–8, 15, 20,
SMNP supported urea-based acidic IL 32, 44, 65, 89, 124–127, 130, 131,
276, 277 134–137, 141–143, 146–150, 168,
SMNP supported urea-based IL 277 172, 233, 239, 243, 250, 276–280,
sol–gel methods/techniques 5, 16, 121, 282–285, 288, 289, 293, 294
124–127, 136, 141, 143, 274 sonochemistry 103, 109
solid catalyst with ionic liquid layers Sonogashira coupling reaction
(SCILL) 267, 270, 271, 280, 290 17, 45
solid nitrogen-containing organic Sonogashira–Hagihara cross-coupling
compounds (SNCOCs) 249 reaction 17, 237
solid-supported IL catalytic systems 272 spirooxindole core 81
336 Index

styrene carbonate 236, 237 thermal conductivity 122, 201, 246,


succinic anhydride, solvent-free 249
diesterification 287 thermal decomposition 50–51, 130, 131,
sulfated tin oxides (STO) 131 138
3-sulfobutyl-1-(3-propyltriethoxysilane) thermal plasma method 8, 53–56
imidazolium hydrogen sulfate IL thermogravimetric analysis (TGA) 63,
275 64, 125, 204
sulfonated CN (Sg-CN) 235 thiazolidinones core 81
sulfonated mesoporous C3 N4 1,3-thiazolidin-4-ones synthesis 283
(omp-g-C3 N4 /SO3 H) 242, thiazoloquinolines synthesis
243 283, 284
sulfur and nitrogen-co-doped carbon NPs thioacetalization of aldehydes
(SNCNPs) 44 194–195
supported ionic liquid phase (SILP) tin oxide 13–14, 123, 127, 131
catalysis 267 tin oxide-based catalysts 127, 130–132
supported ionic liquids (SILs) titanate nanotube 127
advantages 263 titanium dioxide (TiO2 )-based catalysts
building-block system 264 123–127
ionic segment 265 titanium oxide (TiO2 ) 12–13
solid support segment 265–269 top-down approaches 4, 6
catalytic system 269 transmission electron microscopy
solvent-free catalysis 272 (TEM) 33–51, 53–57, 61,
62, 64
t triclinic molybdenum nitride structure
task-specific ILs (TSILs) 184, 262, 270, 247
271 tri-s-triazine 230, 231
task-specific SILLP containing 2,4,5-trisubstituted imidazoles synthesis
imidazolium-sulfonic acid 284
295–296 twin-screw extruder (TSE) 10, 71
task-specific SILLP system 295 twin-screw extrusion 71
t-butyl hydroperoxide (TBHP) 134, two-liquid phase catalysis 262
241
tert-butoxycarbonyl (Boc) 104 u
tert-butyl hydroperoxide (TBHP) 134, ultrasound irradiation 10, 69, 103–109,
241 148, 207
tertiary amine-containing basic IL catalyst UV-visible 122
280, 286
tetra butyl ammonium bromide 17 v
tetrabutylammonium iodide 126 volumetric core heating 80
tetrahydrobenzo[b]pyrans 278
1,1,3,3-tetramethylguanidinium lactate w
(TMGL) 273 water splitting reaction 46, 53
tetraphenylporphyrin (TPP) 213 wet impregnation method 141, 198
thermal condensation method 17 Wittig reaction 75, 76, 134, 166
Index 337

x z
xanthenes zeolite structures
dibenzoxanthenes 192 171–175
X-ray diffraction (XRD) 122 zinc oxide (ZnO)-based catalysts
X-ray fluorescence spectroscopy (XRF) 135–137
63 zinc oxide (ZnO) nanoparticles
X-ray photoelectron spectroscopy (XPS) 14
48 zirconia jar 37, 38

You might also like