Organogermanium Compounds Theory Experiment and Applications Vladimir Ya Lee Full Chapter PDF

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 69

Organogermanium Compounds.

Theory, Experiment, and Applications


Vladimir Ya. Lee
Visit to download the full and correct content document:
https://ebookmass.com/product/organogermanium-compounds-theory-experiment-an
d-applications-vladimir-ya-lee/
More products digital (pdf, epub, mobi) instant
download maybe you interests ...

Organogermanium Compounds: Theory: Experiment, and


Applications, 2 Volumes Lee V.Y. (Ed.)

https://ebookmass.com/product/organogermanium-compounds-theory-
experiment-and-applications-2-volumes-lee-v-y-ed/

The Samantha Granger Experiment: FUSED Kari Lee


Townsend Et El

https://ebookmass.com/product/the-samantha-granger-experiment-
fused-kari-lee-townsend-et-el/

Rheology: Concepts, Methods, and Applications 3rd


Edition Alexander Ya. Malkin

https://ebookmass.com/product/rheology-concepts-methods-and-
applications-3rd-edition-alexander-ya-malkin/

Pincer Compounds: Chemistry and Applications 1st


Edition David Morales-Morales

https://ebookmass.com/product/pincer-compounds-chemistry-and-
applications-1st-edition-david-morales-morales/
Coal Bed Methane: Theory and Applications Pramod Thakur

https://ebookmass.com/product/coal-bed-methane-theory-and-
applications-pramod-thakur/

Elasticity: theory, applications, and numerics Fourth


Edition Sadd

https://ebookmass.com/product/elasticity-theory-applications-and-
numerics-fourth-edition-sadd/

Branching Space-Times : Theory and Applications Nuel


Belnap

https://ebookmass.com/product/branching-space-times-theory-and-
applications-nuel-belnap/

Sound and Recording: Applications and Theory 7th


Edition, (Ebook PDF)

https://ebookmass.com/product/sound-and-recording-applications-
and-theory-7th-edition-ebook-pdf/

Mechanical Design: Theory and Applications T. H. C.


Childs

https://ebookmass.com/product/mechanical-design-theory-and-
applications-t-h-c-childs/
Organogermanium Compounds
Organogermanium Compounds

Theory, Experiment, and Applications

Edited by Vladimir Ya. Lee

Volume 1
This edition first published 2023
© 2023 John Wiley & Sons, Inc.

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

The right of Vladimir Ya. Lee to be identified as the author of editorial material in this work has been asserted in accordance with law.

Registered Office
John Wiley & Sons, Inc., 111 River Street, Hoboken, NJ 07030, USA

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

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

Trademarks: Wiley and the Wiley logo are trademarks or registered trademarks of John Wiley & Sons, Inc. and/or its affiliates in the United
States and other countries and may not be used without written permission. All other trademarks are the property of their respective owners.
John Wiley & Sons, Inc. is not associated with any product or vendor mentioned in this book.

Limit of Liability/Disclaimer of Warranty


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

A catalogue record for this book is available from the Library of Congress

Hardback ISBN: 9781119613435; Set ISBN: 9781394177561 (Volume 1); ePub ISBN: 9781119613527;
ePDF ISBN: 9781119613473; oBook ISBN: 9781119613466

Cover Images: © ALFRED PASIEKA/Getty Images; © Intothelight Photography/Shutterstock


Cover design by Wiley

Set in 9.5/12.5pt STIXTwoText by Integra Software Services Pvt. Ltd, Pondicherry, India
v

Full Table of Contents

Volume 1:

Preface ix
List of Contributors xiii

1 Computational and Theoretical Aspects of Structure and Bonding in Doubly Bonded Organogermanium
Compounds 1
Miriam Karni and Yitzhak Apeloig

2 Organogermanium Compounds of the Main Group Elements 103


Kirill V. Zaitsev

3 Transition Metal Complexes of Germanium 195


Kohtaro Osakada

4 Germanium Cages and Clusters 225


Tanja Kunz and Andreas Schnepf

5 Arylgermanium Hydrides, ArnGeH4-n (n = 1–3) - Synthesis, Characterization, Reactivity 277


Ana Torvisco and Frank Uhlig

6 Germylium Ions and Germylium Ion-like Species 299


Thomas Müller

7 Germanium-Containing Radicals 339


Alexander Hinz and Frank Breher

8 Germanium-Centered Anions 361


Christoph Marschner

9 Germylenes 387
Norio Nakata

10 Multiple Bonds to Germanium 435


Vladimir Ya. Lee
vi Full Table of Contents

Volume 2:

Preface vii
List of Contributors xi

11 Germaaromatic Compounds 477


Yoshiyuki Mizuhata and Norihiro Tokitoh

12 Germanium-centered Ion Radicals 507


Mikhail P. Egorov, Viatcheslav V. Jouikov, Elena N. Nikolaevskaya, and Mikhail A. Syroeshkin

13 Donor-acceptor Stabilization of Species with Low-coordinate Germanium 561


Sakya S. Sen and Herbert W. Roesky

14 Synthesis of the Penta- and Hexacoordinate Germanium(IV) Complexes 597


Naokazu Kano

15 Dynamic Stereochemistry of Penta- and Hexacoordinate Germanium(IV) Complexes 629


Vadim V. Negrebetsky and Alexander A. Korlyukov

16 X-ray Crystallography of Organogermanium Compounds 667


Catherine Hemmert and Heinz Gornitzka

17 Organogermanium Photochemistry 745


William J. Leigh

18 Oligo- and Polygermanes 787


Charles S. Weinert

19 Bioorganic and Medicinal Organogermanium Chemistry 839


Takashi Nakamura, Yasuhiro Shimada, and Katsuyuki Sato

Index 867
vii

Contents

Preface ix
List of Contributors xiii

1 Computational and Theoretical Aspects of Structure and Bonding in Doubly Bonded Organogermanium
Compounds 1
Miriam Karni and Yitzhak Apeloig

2 Organogermanium Compounds of the Main Group Elements 103


Kirill V. Zaitsev

3 Transition Metal Complexes of Germanium 195


Kohtaro Osakada

4 Germanium Cages and Clusters 225


Tanja Kunz and Andreas Schnepf

5 Arylgermanium Hydrides, ArnGeH4-n (n = 1–3) - Synthesis, Characterization, Reactivity 277


Ana Torvisco and Frank Uhlig

6 Germylium Ions and Germylium Ion-like Species 299


Thomas Müller

7 Germanium-Containing Radicals 339


Alexander Hinz and Frank Breher

8 Germanium-Centered Anions 361


Christoph Marschner

9 Germylenes 387
Norio Nakata

10 Multiple Bonds to Germanium 435


Vladimir Ya. Lee
ix

Preface

Germanium is one of the few chemical elements in the Periodic Table, for which the theoret-
ical prediction of its very existence has preceded its actual experimental discovery. This predic-
tion was made by the Russian chemist Dmitri Mendeleev based on the general trends of valence
and atomic weights within his Periodic Table of the chemical elements (1869) [D. Mendelejeff
“Ueber die Beziehungen der Eigenschaften zu den Atomgewichten der Elemente”, Z. Chem.
1869, 12, 405–406]. In an updated version of this Periodic Table (1871-1872) [D. Mendelejeff
“Die Periodische Gesetzmässigkeit der Chemischen Elemente”, Ann. Chem. Pharm. 1872,
Suppl. 8, 133–229; D. Mendelejeff “Zur Frage über das System der Elemente”, Ber. Dtsch. Chem.
Ges. 1871, 4, 348–352], Mendeleev proposed that there was a missing element in the carbon
family with the atomic weight 72 that should be placed in the fourth row, just below silicon and
just above tin within the carbon group. He named this non-existing (at that time) element as
“eka-silicium”. Following this seminal Mendeleev prediction, German chemist Clemens
Winkler finally succeeded in 1886 in the isolation of “eka-silicium” from the mineral argyrodite (Ag8GeS6) and named this
new element as germanium (Ge) [C. Winkler “Germanium, Ge, Ein Neues, Nichtmetallisches Element”, Ber. Dtsch. Chem.
Ges. 1886, 19, 210–211; C. Winkler “Mittheilungen über das Germanium”, J. Prakt. Chem. 1886, 34, 177–229]. Winkler also
pioneered the preparation of the first organic derivative of germanium, namely, tetraethylgermane Et4Ge, in 1887 [C.
Winkler “Mittheilungen über das Germanium”, J. Prakt. Chem. 1887, 36, 177–209]. Since then and up the present date, the
chemistry of organogermanium compounds (that is, compounds featuring Ge–C bonds) has experienced an explosive
growth, especially after the recognition of the key role of metallic germanium in semiconductor electronics in the mid-
twentieth century, followed by the extensive use of germanium and its organic derivatives in optical fibers, polymerization
catalysts, microchip manufacturing, and biomedical applications. Given the undoubted importance of organogermanium
compounds, it comes as no surprise that the field of organogermanium chemistry is continuously growing, thus requiring
regular reviewing and updates on its latest advances. Among the most important previously published books on organo-
germanium chemistry, one should first of all mention excellent monograph by Satgé and coworkers [J. Satgé, M. Lesbre, P.
Mazerolles, “The Organic Compounds of Germanium, Wiley, 1971] and two comprehensive volumes of the Patai’s series of
books [(a) The Chemistry of Organic Germanium, Tin, and Lead Compounds (Eds. S. Patai, Z. Rappoport), Wiley, 1995; (b)
The Chemistry of Organic Germanium, Tin, and Lead Compounds, Volume 2 (Ed. Z. Rappoport), Volume 2, Parts 1–2, Wiley,
2002]. Patai’s latest book was published 20 years ago, and since then, critical progress has been made in organogermanium
chemistry with the majority of milestone developments achieved since 2000. That is why we have attempted in this book
to survey, analyze and summarize the current state of affairs in the field of organogermanium chemistry, focusing on the
latest (published mostly after 2000) groundbreaking advances with comprehensive and up-to-date literature coverage up
to the end of 2021.
Our “Organogermanium Compounds” book is organized into the three major parts, Theory, Experiment, and
Applications, made up of a total of 19 chapters, each written by leading experts in their respective fields comprehensively
covering the relevant literature, first published within the last two decades.
1) The first part, Theory (one chapter) includes a contribution “Computational Aspects of Structure and Bonding in
Doubly Bonded Organogermanium Compounds” from Miri Karni and Yitzhak Apeloig (Technion – Israel Institute of
Technology, Haifa, Israel). In their chapter, the authors present state-of-the-art computational approaches to the flourish-
ing field of doubly bonded organogermanium compounds and thoroughly discuss their structural and bonding aspects
depending on the nature of the double bond and its substituents.
x Preface

2) The second part, Experiment (16 chapters), deals with the most fundamental experimental advances that were
achieved in synthetic and physico-chemical studies, and is accordingly divided into two sections, Synthesis of
Organogermanium Compounds and Physico-Chemical Studies of Organogermanium Compounds. The first section, Synthesis
of Organogermanium Compounds, is further subdivided into three subsections (in accord with the coordination number of
the central germanium): 1) Organogermanium Compounds of Tetracoordinate Germanium, 2) Organogermanium
Compounds of Low-Coordinate Germanium, 3) Organogermanium Compounds of Hypercoordinate Germanium.
In the first subsection, Organogermanium Compounds of Tetracoordinate Germanium, there are four contributions. Kirill
V. Zaitsev (Moscow State University, Moscow, Russia) in his chapter “Organogermanium Compounds of the Main Group
Elements” provides a detailed overview of synthetic and structural aspects of organogermanium compounds containing
Ge–E bonds (where E is an element of groups 13–17). The following chapter “Transition Metal Complexes of Germanium”
by Kohtaro Osakada (Tokyo Institute of Technology, Tokyo, Japan) comprehensively covers chemistry of transition metal
complexes featuring metal–germanium single bonds. In their chapter “Germanium Cages and Clusters”, Tanja Kunz and
Andreas Schnepf (University of Tübingen, Tübingen, Germany) discuss syntheses, structural characterization, and reac-
tivity of germanium clusters of the polyhedral-, metalloid-, and Zintl-type. In the final chapter of this subsection,
“Arylgermanium Hydrides RnGeH4–n (n = 1–3) – Synthesis, Characterization, Reactivity”, Ana Torvisco and Frank Uhlig
(Graz University of Technology, Graz, Austria) describe a variety of aryl-substituted organogermanium hydrides focusing
on their synthesis, structural aspects (including 73Ge NMR properties), and selected applications.
The most abundant second subsection (eight chapters), Organogermanium Compounds of Low-Coordinate Germanium,
focuses on the emerging field concerned with the stable germanium analogues of pivotal organic reactive intermediates,
such as germanium-centered cations, free radicals, anions, ion-radicals, germylenes, multiply-bonded organogermanium
compounds, germaaromatics, and donor-acceptor complexes of low-coordinate germanium. In the first chapter in this sub-
section, “Germylium Ions and Germylium Ion-Like Species”, Thomas Müller (University of Oldenburg, Oldenburg,
Germany) summarizes the chemistry of germylium ions [R3Ge]+ and ion-like species [R3Ge(Do)]+ (Do = electron donor)
with particular emphasis on their synthesis, characterization, and specific reactivity. The chapter “Germanium-Containing
Radicals” by Alexander Hinz and Frank Breher (Karlsruhe Institute of Technology, Karlsruhe, Germany) discusses general
methods for generation and isolation of Ge-centered radicals (including those with redox-non-innocent ligands), polyradi-
cals, as well as their reactivity and synthetic applications. The following chapter “Germanium-Centered Anions”, written
by Christoph Marschner (Graz University of Technology, Graz, Austria), focuses on the general methods for the prepara-
tion and synthetic utilization of a variety of germyl anions, as well as germyl dianions, and compounds with a negative
charge on the sp2 Ge atom. Recent advances in the field of the stable germylenes (including N-heterocyclic germylenes)
and their transition metal complexes, with focus on their preparation, characterization and specific reactivity, are pre-
sented in the chapter “Germylenes” by Norio Nakata (Saitama University, Saitama, Japan). The story of the low-coordinate
organogermanium compounds is further continued by Vladimir Ya. Lee (University of Tsukuba, Tsukuba, Japan) in his
chapter “Multiple Bonds to Germanium,” which summarizes the latest developments in the field of doubly and triply
bonded organogermanium derivatives containing both homonuclear and heteronuclear multiple bonds. Another challeng-
ing topic of contemporary organogermanium chemistry, namely, germaaromatic compounds (both neutral and charged)
including non-classical aromatic systems, is overviewed by Yoshiyuki Mizuhata and Norihiro Tokitoh (Kyoto University,
Kyoto, Japan) in their chapter “Germaaromatic Compounds”. The following chapter “Germanium-Centered Ion Radicals”,
written by Mikhail P. Egorov, Viatcheslav V. Jouikov, Elena N. Nikolaevskaya, and Mikhail A. Syroeshkin (Institute of
Organic Chemistry, Moscow, Russia and University of Rennes, Rennes, France), covers recent progress in the field of
anion-radicals and cation-radicals discussing general methods for their generation and identification, both experimental
and computational. The Organogermanium Compounds of Low-Coordinate Germanium subsection is completed by a con-
tribution “Donor-Acceptor Stabilization of Species with Low-Coordinate Germanium” by Sakuya S. Sen and Herbert W.
Roesky (National Chemical Laboratory, Pune, India and University of Göttingen, Göttingen, Germany), which highlights
the progress in the field of compounds with low-coordinate germanium stabilized by donor-acceptor interactions.
The Synthesis of Organogermanium Compounds section ends with the subsection Organogermanium Compounds of
Hypercoordinate Germanium containing two contributions. In the first one, “Synthesis of the Penta- and Hexacoordinate
Germanium(IV) Complexes”, Naokazu Kano (Gakushuin University, Tokyo, Japan) classifies the title compounds bearing
bi-, tri-, and tetradentate ligands (as well as carbene ligands), and discusses their synthesis and reactivity. The second con-
tribution “Dynamic Stereochemistry of Penta- and Hexacoordinate Germanium(IV) Complexes”, written by Vadim V.
Negrebetsky and Alexander A. Korlyukov (Russian National Research Medical University, Moscow, Russia and Institute
of Organoelement Compounds, Moscow, Russia), summarizes and analyzes the data on stereochemical processes involving
organic complexes of hypercoordinate germanium.
Preface xi

The Experiment part is completed by the Physico-Chemical Studies of Organogermanium Compounds section which
comprises two contributions each dealing with state-of-the-art instrumental techniques for assessing the structures of
organogermanium compounds. In the first one, “X-Ray Crystallography of Organogermanium Compounds”, Catherine
Hemmert and Heinz Gornitzka (University of Toulouse, Toulouse, France) comprehensively overview the vast literature
on the structurally characterized organogermanium compounds discussing their particular X-ray crystallographic fea-
tures. The following chapter “Organogermanium Photochemistry”, by William J. Leigh (McMaster University, Hamilton,
Canada) deals with developments in the photochemistry of organogermanium compounds, particularly emphasizing sys-
tems involving low-coordinate organogermanium compounds as either photoproducts or photoreactants.
3) The last, third, part of the book, Applications (two chapters), reports on the synthetic approaches towards materials
based on organogermanium compounds and their practical use. In the first contribution, “Oligo- and Polygermanes”,
Charles S. Weinert (Oklahoma State University, Stillwater, USA) reviewed and updated the recent advances in the field of
the oligo- and polygermanes, discussing general synthetic methods for their preparation and also their peculiar physical
properties (luminescence, electrochemistry). In the very final chapter of the Applications part (and the whole book)
“Bioorganic and Medicinal Organogermanium Chemistry”, Takashi Nakamura, Yasuhiro Shimada, and Katsuyuki Sato
(Asai Germanium Research Institute, Hakodate, Japan) give an account on the production and use of bioactive organoger-
manium compounds (such as Ge-132) in medicinal chemistry.
As an editor of the volume, I highly appreciate the works of all above-mentioned authors whose excellent contributions
will hopefully turn this book into a guide to contemporary organogermanium chemistry and a useful reference source, to
further inspire those already working in the field and to attract newcomers to join the fascinating world of organogerma-
nium compounds. Although the book is primarily and foremost addressed to specialists in the field of organogermanium
and organometallic chemistry, it could also be of interest to both Main Group and transition metal chemistry experts, as
well as to computational and material science chemists, and to the rest of the chemical community, including undergrad-
uate, graduate, and post-graduate students of the advanced levels.

Vladimir Ya. Lee (Editor)


University of Tsukuba, Tsukuba, Japan
April 2022
xiii

List of Contributors

Yitzhak Apeloig Miriam Karni


Schulich Faculty of Chemistry Schulich Faculty of Chemistry
Technion-Israel Institute of Technology Technion-Israel Institute of Technology
Haifa, Israel Haifa, Israel

Frank Breher Tanja Kunz


Institute of Inorganic Chemistry Institut für Anorganische Chemie
Karlsruhe Institute of Technology (KIT) Universität Tübingen
Karlsruhe, Germany Tübingen, Germany

Mikhail P. Egorov Alexander A. Korlyukov


N. D. Zelinsky Institute of Organic Chemistry Pirogov Russian National Research Medical University
Moscow, Russia Moscow, Russia

Vladimir Ya. Lee


Heinz Gornitzka
Department of Chemistry
Laboratoire de Chimie de Coordination du CNRS
University of Tsukuba
Université Toulouse
Tsukuba, Japan
Toulouse, France

William J. Leigh
Catherine Hemmert Department of Chemistry & Chemical Biology
Laboratoire de Chimie de Coordination du CNRS McMaster University
Université Toulouse Hamilton, Ontario, Canada
Toulouse, France
Christoph Marschner
Alexander Hinz Institut für Anorganische Chemie
Institute of Inorganic Chemistry Technische Universität Graz
Karlsruhe Institute of Technology (KIT) Graz, Austria
Karlsruhe, Germany
Yoshiyuki Mizuhata
Viatcheslav V. Jouikov Institute for Chemical Research
N. D. Zelinsky Institute of Organic Chemistry Kyoto University
Moscow, Russia Kyoto, Japan

Naokazu Kano Thomas Müller


Department of Chemistry Institute of Chemistry
Gakushuin University Carl von Ossietzky University Oldenburg
Tokyo, Japan Oldenburg, Germany
xiv   List of Contributors

Takashi Nakamura Sakya S. Sen


Asai Germanium Research Institute Co., Ltd. Institute of Inorganic Chemistry
Hokkaido, Japan Georg-August-University Göttingen
Göttingen, Germany
Norio Nakata
Department of Chemistry Yasuhiro Shimada
Saitama University Asai Germanium Research Institute Co., Ltd.
Saitama, Japan Hokkaido, Japan

Mikhail A. Syroeshkina
Vadim V. Negrebetsky
N. D. Zelinsky Institute of Organic Chemistry
Pirogov Russian National Research Medical University
Moscow, Russia
Moscow, Russia
Norihiro Tokitoh
Elena N. Nikolaevskaya Institute for Chemical Research
N. D. Zelinsky Institute of Organic Chemistry Kyoto University
Moscow, Russia Kyoto, Japan

Kohtaro Osakada Ana Torvisco


Chemical Resources Laboratory Institute for Inorganic Chemistry
Tokyo Institute of Technology Graz University of Technology
Yokohama, Japan Graz, Austria

Herbert W. Roesky Frank Uhlig


Institute of Inorganic Chemistry Institute for Inorganic Chemistry
Georg-August-University Göttingen Graz University of Technology
Göttingen, Germany Graz, Austria

Charles S. Weinert
Katsuyuki Sato
Department of Chemistry
Asai Germanium Research Institute Co., Ltd.
Oklahoma State University
Hokkaido, Japan
Stillwater, Oklahoma

Andreas Schnepf Kirill V. Zaitsev


Institut für Anorganische Chemie Department of Chemistry
Universität Tübingen Moscow State University
Tübingen, Germany Moscow, Russia
1

Computational and Theoretical Aspects of Structure and Bonding in Doubly


Bonded Organogermanium Compounds

Miriam Karni and Yitzhak Apeloig


Schulich Faculty of Chemistry, Technion - Israel Institute of Technology, Haifa, Israel

In memory of Professor Robert (Bob) West, a pioneer in main group chemistry and group 14 compounds in particular, and
a unique human being.

List of Abbreviations

1-Ad 1- adamantyl
AIM atoms in molecules
Ar aryl
B3LYP Becke, 3-parameter, Lee–Yang–Parr functional
Bbt 2,6-[CH(SiMe3)2]2-4-[C(SiMe3)3]-C6H2
BDE bond dissociation energy
CCSD(T) coupled cluster, singles, doubles (triples)
CDA charge delocalization analysis
CGMT Carter-Goddard-Malrieu-Trinquier model
CV cyclic voltammetry
Dip (or Dipp) 2,6-iPr2-C6H3
DMAP 4-Dimethylaminopyridine
Dmp 2,6-dimethylphenyl
DSSE divalent state stabilization energy
EDA energy decomposition analysis
Eind 1,1,3,3,5,5,7,7-octaethyl-s-hydrindacen-4-yl
EMind 1,1,7,7-tetraethyl-3,3,5,5-tetramethyl-s-hydrindacen-4-yl
EPR electron paramagnetic resonance

Etind
Et Et
Et Et

HOMO highest occupied molecular orbital


LDA/NL local density approximation with non-local corrections
LDF London Dispersion Forces
LUMO lowest unoccupied molecular orbital
Mes 2,4,6-Me3-C6H2
Mes* 2,4,6-tBu3-C6H2
NBO natural bond orbital
NHC N-Heterocyclic carbene
NICS nucleus independent chemical shift
NMR nuclear magnetic resonance
NPA natural population analysis
NRT natural resonance theory
PES potential energy surface

Organogermanium Compounds: Theory, Experiment, and Applications, Volume 1, First Edition. Edited by Vladimir Ya. Lee.
© 2023 John Wiley & Sons, Inc. Published 2023 by John Wiley & Sons, Inc.
2 1 Computational and Theoretical Aspects of Structure and Bonding in Doubly Bonded Organogermanium Compounds

RRKM Rice-Ramsberger-Kassel-Marcus theory


SDD Sttutgart-Dresden Double zeta Electron Core Potential (ECP)
SOJT second-order Jahn Teller
Tbb 2,6-[CH(SiMe3)2]2-4-tBu-C6H2
Tip 2,4,6-triisopropylphenyl
VB valence bond
WBI Wiberg bond index
Xylyl (CH3)2C6H3

1.1 Introduction

Unsaturated hydrocarbons having C=C double bonds; heteroelement- substituted doubly bonded carbon compounds, e.g.,
ketones; alkynes with C≡C triple bonds; hydrocarbons with an extended number of double bonds, such as dienes and
allenes; all are of fundamental importance to chemistry and the chemistry of life. In contrast stable heavier group 14 ana-
logues, with Si, Ge, Sn and Pb atoms, were believed to be nonexistent until several decades ago (see below the historical
overview). The last four decades have witnessed an enormous progress in the synthesis and in the theoretical under-
standing of the properties of these groups of compounds. The experimental progress was accompanied, and in many cases
even preceded, by theoretical and computational studies. These advances were enabled by two important developments:
(a) the substantial advances in quantum mechanical computational methodology [1–10] that now enables reliable, accu-
rate and efficient calculations of the properties of relatively large molecules possessing heavy elements. These develop-
ments make quantum chemical calculations a reliable tool for the quantitative analysis of chemical phenomena and a
reliable guide to experiment; and (b) the dramatic developments in computer technology that have accelerated calculations
dramatically [11, 12].
The synthesis, molecular structure determination, physical and chemical properties, and the development of bonding
models of multiply bonded compounds of heavier group 14 elements, Si, Ge, Sn and Pb, have been reviewed extensively in
the past three decades [13–26]. The numerous experimental and theoretical studies carried out in this field have revealed
fundamental differences in the structures, physical properties and chemical behavior between carbon multiply bonded
compounds and analogous group 14 heavier congener compounds. The heavier group 14 compounds were considered
“unusual”. In contrast, Kutzelnigg suggested that the heavier main group element compounds have “regular” behavior,
while carbon compounds are the exception [27]. This revolutionary view is now widely accepted by main group researchers
(not only for group 14 elements). The trends in the physical properties of group 14 elements, E, are largely responsible for
the unprecedented structures and chemical behavior of the heavier group 14 compounds compared to those of the analo-
gous carbon compounds. Selected properties of group 14 elements are given in Table 1.1. A comparison of the important
physical properties of group 14 elements was also provided by Basch and Hoz [28]. The effects of the trends in the prop-
erties of the E atoms shown in Table 1.1, on the structure and nature of bonding of group 14 multiply bonded compounds,
are discussed in section 1.4.1.
The unusual structures and bonding motifs of multiply bonded heavier group 14 molecules attracted the attention of
many theoreticians, who studied their molecular structures, nature of bonding, kinetic and thermodynamic stability,
potential energy surfaces for a variety of their reactions, etc. Many of these theoretical studies were carried out before stable
compounds of these classes were synthesized, and were therefore truly predictive, and in some cases directed experiment.
Two comprehensive reviews on the theoretical studies of the chemistry of organic Si, Ge, Sn and Pb compounds were pub-
lished: one by Apeloig, Karni and Schleyer in 2001 [44] and one in 2002 by Frenking and coworkers [45].
In this chapter we review the theoretical studies of doubly bonded organogermanium compounds, mainly those
published in the years 2000–2020 but occasionally we added important studies published in 2021. We have highlighted
the insights that these theoretical studies provide on molecular structures, nature of bonding, stability vs. isomeric forms,
and the physical properties, of these intriguing compounds. We do not include in this review their reactions and reaction
mechanisms. Where available, we relate and compare the computational results to reported experimental studies. In our
view, it is important to examine organogermanium compounds in comparison to their closest neighbors in the Periodic
Table, Si and Sn. Thus, when available, we discuss comparisons with analogous organosilicon compounds (the element
most similar to germanium) and other group 14 elements, mainly Sn. For pre-year 2000 theoretical studies, the readers
are referred to the above-mentioned reviews [44, 45] and to specific papers cited in this chapter.
1.2 Computational Methods 3

Table 1.1 Selected physical properties of group 14 elements.

E C Si Ge Sn Pb References

Electronegativity
Allred-Rochow 2.50 1.74 2.02 1.72 1.55 [29,30]
Pauling 2.55 1.90 2.01 1.96 2.33 [31]
Allen
2.54 1.92 1.99 1.82 – [32]
Atomic and ionic radii (pm) 77 118 121 140 175a [33–36]
Neutral 16 40–42 73 93 118–120
2+ 53 69–71 78–84
4+
Valence orbital energy (eV) −19.39 −14.84 −15.52 −13.88 −15.41 [37]
s −11.07 −7.57 −7.29 −6.71 −6.48
p 8.32 7.27 8.23 7.17 8.93
Energy difference
Δr = (rp − rs) (pm)b −0.2 20.3 24.9 28.5 35.8 [37]
Ionization energy (eV) 16.60 13.64 14.43 13.49 16.04 [38–40]
nsc 11.26 8.15 7.90 7.39 7.53
npd
Electron affinity (eV) 1.26 1.76 1.81 1.68 1.91 [41]
Hybridization of E in EH4e sp3.17 sp2.08 sp2.05 sp1.79 sp1.75 [42]
f
Hybridization of E in H2E=EH2 (C2h)
σ(E-E) sp2 sp1.79 sp2.07 sp2.22 sp2.52 [43]
“π(E=E)” p sp18.0 sp7.40 sp5.78 sp3.82
a
Metallic distance;

For the transition: ns2np2(3P) → ns1np2(4P);


b
Difference of orbital radii between the valence p and s orbitals;

For the transition: ns2np2(3P) → ns2np1(2P);


c
d
e
Calculated using NBO analysis;
f
Calculated using NBO analysis at B3LYP/6-31++G(d,p).

The theoretical literature of organogermanium compounds is vast. When we collected the literature for this review, we
were amazed how vast is this field, and because of space limitations we had to limit the scope of this review, which even so
includes more than 540 references! In this chapter, we concentrate on doubly bonded organogermanium compounds,
which in their bonding and other fundamental properties are significantly different from analogous carbon compounds.
We discuss homonuclear and heteronuclear organogermanium compounds, R2Ge=ER2, E=C, Si, Ge, and extended doubly
bonded organogermanium compounds such as germadienes, germaallenes, and germaaromatic derivatives, i.e., germ-
abenezene analogues. For lack of space, we do not discuss doubly bonded compounds between group 14 elements and
groups 13, 15 and 16 elements or triply bonded compounds.
For completeness, we present a brief historic overview of the field, since the isolation of the first stable doubly bonded
heavier group 14 element compounds in 1976.

1.2 Computational Methods

For the benefit of readers unfamiliar with the theoretical methods and terms, we bring a brief overview of the theoretical
methods that are currently frequently used in theoretical calculations of organogermanium compounds. Details of the
specific computational methods or of the basis sets mentioned in this review can be found in the cited papers.
The rapid development of computational packages for electronic structure calculations in the last three decades has been
recently reviewed [10] (for a comprehensive list of available software packages see Ref. [46]). These computational soft-
ware packages apply state-of-the-art theoretical methods for electronic structure modelling of molecules in the gas phase,
4 1 Computational and Theoretical Aspects of Structure and Bonding in Doubly Bonded Organogermanium Compounds

solution, and the solid-state. Currently, these software packages apply mainly two theoretical approaches, ab initio wave-
function based molecular orbital (MO) methods [1–5], and electron density based Density Functional Theory (DFT)
methods [6–8]. These software packages enable highly reliable calculations of molecular structures, potential energy sur-
faces, electron density and population analysis, vibrational frequencies, UV-Vis, NMR, EPR, IR and Raman spectroscopic
properties, electronic circular dichroism (ECD), excitation energies, and other properties [1–9, 47].
Below we briefly summarize the theoretical approaches most commonly used for the calculation of the above-mentioned
molecular properties. A detailed description of the field is given in the monographs cited above. Specific definitions and
description of the computational methods used in the studies reviewed in this chapter are provided in the cited papers and
references therein.

1.2.1 Ab initio Methods


Ab initio MO methods are based on Hartree-Fock (HF) self-consistent field (SCF) calculations, in which the total multi-
electron wave function Ψ is composed from single electron spin orbital wave functions ϕl (MO) via a single anti-symme-
trized determinant (the Slater determinant). The single electron wave functions are expressed by a sum of basis functions
(basis sets). A very large choice of basis sets exists; they differ in size and in their adaptation for a specific calculated
property [48, 49]. In addition to choosing the appropriate computational method, a correct choice of a basis set deter-
mines the computational accuracy and efficiency. However, even with the largest basis set and reaching the HF limit,
the HF method recovers only 95–99% of the total energy. The missing fraction is relatively small in absolute terms, but
most chemistry happens in this small remaining fraction. The missing fraction is corrected by including electron corre-
lation energy.
The most frequently used methods for including electron correlation are by perturbation theory, i.e., Møller-Plesset [50]
in the nth order, (MPn) [51], configuration interaction (CI) [52], and coupled-cluster (CC) [53] methods [1–5, 46]. Today,
coupled-cluster methods provide the most accurate results among the practical ab initio electronic structure theories and
often serve as a benchmark for evaluating, in the absence of experimental results, the reliability of new computational
methods, e.g., density functionals. Ab initio methods can be improved systematically by increasing the size of the basis sets
and the amount of correlation energy that is included. It is assumed that with a complete basis set and full CI the exact
solution to the Schrödinger equation can be obtained. However, the major drawback of these methods is the large com-
puter resources required, and they are therefore applicable only to moderately-sized molecules.

1.2.2 Density Functional Theory ( DFT) Methods


DFT methods [6–8], which include electron correlation indirectly, have become the most popular electronic structure
methods in computational chemistry over the last 30 years. Recent advances in the development of density functionals
(DF) allow one to reliably describe numerous properties of large systems at an affordable computational expense. A major
difficuly in developping reliable DFs is that in contrast to the rigor of ab intio methods, DFT methods do not allow systematic
improvement of the theoretical method. The strategy for improving DFs is by adding new ingredients and parametrizing
them by fitting to a set of properties, e.g., atomization energy, thermochemistry, barrier heights, etc. For example, addition
of a small percentage of exact HF exchange (i.e., hybrid DFs) improves significantly the performance of local DFs as shown
in Figure 1.1. For example, the performance of the PBE (Perdew–Burke-Ernzerhof) functional is improved by a factor of
two in PBE0, which mixes the PBE exchange energy with 25% exact Hartree–Fock exchange energy [8, 54]. Figure 1.2
shows the importance of including London dispersion corrections (D2 [55] and D3(BJ) [56]) that capture noncovalent
interactions more effectively. Based on an assessment of 200 DFs tested on 84 data sets including thermodynamic data,
isomerization energies, barrier heights, etc., it was concluded that addition of dispersion corrections should be applied with
caution [8]. Recommendations for effective use of DFs that are suited for specific applications are given in Refs. [8] and
[57], and the reader is advised to consult these references before applying a specific method to his own research or when
evaluating the research of others. As can be seen in Figures 1.1 and 1.2, calculated errors can be very large, e.g., with BLYP,
B3LYP, and BP86.
The common notations of the theoretical methods used in the literature are the following: each computational level (ab
initio and DFT) is defined by the type of method and the basis sets which are used. By convention, the computational level
is designated as follows: [method 1/basis set 1]//[method 2/basis set 2]. Method 2 and basis set 2 are those used for geom-
etry optimizations while method 1 and basis set 1 are those used for obtaining the final energy or other properties that are
calculated.
1.2 Computational Methods 5

30
NCD ID BH
25
RMSD (kcal/mol)

20

15

10

0
BLYP B3LYP BP86 B3P86 PBE PBE0 revPBE revPBE0 TPSS TPSSh revTPSS revTPSSh

Figure 1.1 Stacked root-mean-square deviations (RMSD, in kcal/mol) for a set of six popular local/hybrid pairs (e.g., BLYP/B3LYP,
BP86/B3P86, etc.) derived from 155 isomerization energies (ID), 206 barrier heights (BH), and 91 noncovalent (NCD) interactions.
Reproduced with permission from Ref. [8]. Copyright (2017) Taylor and Francis, Ltd.

6
NCED IE
5
RMSD (kcal/mol)

0
BLYP -D2 -D3(BJ) B3LYP -D2 -D3(BJ) TPSS -D2 -D3(BJ) TPSSh -D2 -D3(BJ)

Figure 1.2 Stacked root-mean-square deviations (RMSD, in kcal/mol) for a set of four DFs and their dispersion corrected (D2 and
D3(BJ)) counterparts. Based on 1744 non-covalent dimer interactions (NCED) and on 755 isomerization energies (IE). Reproduced with
permission from Ref. [8]. Copyright (2017) Taylor and Francis, Ltd.

1.2.3 Methods for Analysis of Electronic Structure


Important information and deeper insight into the chemical and physical properties of the studied molecules can be gained
by analyzing their wave function, the electron distribution, hybridization at various atoms, the nature of the chemical bond,
etc. Several analysis methods are mentioned throughout this chapter. The most popular is the Natural Bond Orbital method
(NBO) [58–63] that calculates the Natural Population Analysis charges (NPA). The NPA charges are believed to be more
accurate [64, 65] than the Mulliken atomic charges [66, 67], which are still often used. Several other methods for calculating
atomic charges are available and a comparison of their accuracy for specific purposes was discussed, e.g., in References [64]
and [65]. NBO analysis, based on localized orbitals, provides a Lewis structure of the molecule, bond orders, hybridization
at the various atoms, and suggests possible resonance structures that contribute to the molecular electronic structure and
their relative weights (NRT)[68], thus providing important information for analyzing the structures and bonding of the
studied molecules. Other frequently used computational procedures are: (a) Atoms in Molecules (AIM), developed by Bader
[69], which is based on a topological analysis of the electron density in atomic basins, and calculates also atomic AIM
charges; and (b) Energy Decomposition Analysis and Natural Orbitals for Chemical Valence (EDA-NOCV), which proved to
be a powerful tool for bonding analysis. For a detailed description of the above-mentioned methods and their performance,
the reader is referred to the recent review by Frenking, Schwerdtfeger, and coworkers [70].

1.2.4 How Important are Relativistic Effects for Ge?


As the nuclei become heavier the strong attraction of the electrons by the heavier nuclear charge causes the electrons to
move faster and behave relativistically [71, 72], i.e., their relative mass increases and the effective Bohr average radius for
6 1 Computational and Theoretical Aspects of Structure and Bonding in Doubly Bonded Organogermanium Compounds

(a) (b)
–12.0

nonrelativistic energy

Bond length relative to NR (a.u)


–14.0 0.02

0
s orbital energy (eV)

–0.02
relativistic energy
–16.0
–0.04

–0.06

–0.08
–18.0
CH4 SiH4 GeH4 SnH4 PbH4

–20.0
C Si Ge Sn Pb

Figure 1.3 (a) Stabilization of the ns valence orbital due to relativistic effect. Reprinted from Ref. [73] by permission of Gordon and
Breach science publishers, 1999; (b) The contraction of the E–H bond length of EH4 (E=C, Si, Ge, Sn and Pb) calculated using Fock-
Dirac (rectangle) and Fock-Dirac-Breit (*) relativistic approaches relative to non-relativistic calculations. Reprinted with permission
from Ref. [74]. Copyright (1992) Springer Verlag.

the inner-electrons contracts. It was found that the relativistic contraction of the 1s orbital of Ge, Sn, and Pb, is 3, 8, and
20%, respectively [71]. Due to the orthogonality of the higher ns orbitals to the 1s orbital, the relativistic effect on higher ns
orbitals is similar [71].
As shown in Figure 1.3a, the relativistic effect stabilizes the ns orbitals of Ge–Pb atoms. However, the effect for Ge is
relatively small [73]. The relativistic contraction of the E–H bond in EH4 is 0.01% for CH4, 0.05% for SiH4, 0.37% for GeH4
1.16% for SnH4, and 3.9% for PbH4, leading to nearly identical E–H bond lengths in SnH4 (1.734 Å) and PbH4 (1.737 Å).
These data show relatively small relativistic effects for Ge, and consequently most calculations of Ge compounds do not
include relativistic effects.

1.3 Historical Overview

Until the early 1970s, compounds with heavier group 14 elements, E, having double and triple bonds, were unknown (i.e.,
homonuclear or heteronuclear compounds possessing E=E’ double bonds and E≡E’ triple bonds (E, E’=Si, Ge, Sn, and Pb;
E’=group 13–16 elements). Many experimental failures [75–78] to synthesize such compounds and some theoretical studies
[79, 80] led to formulation of the classical “double bond rule” stating that “elements having a principal quantum number
greater than 2 should not be able to form (p–p) π bonds with themselves or with other elements” [81]. Therefore, this group
of compounds were believed not to exist. However, the pursuit for such compounds continued, leading in the late 1970s to
early 1980s to detection of transient intermediates having E=E’ bonds of heavier group 14 elements [22, 75–78, 82, 83]. The
first important breakthrough occurred in 1976, interestingly in germanium and tin chemistry, with the synthesis, isolation,
and characterization of alkyl substituted digermene and distannene, [(Me3Si)2CH]2E=E[CH(SiMe3)2]2, (E=Ge (1) and
E=Sn, (2)) by Lappert and coworkers [84–87]. The X-ray structure of digermene 1 shows a Ge–Ge bond length of 2.347 Å
[86, 87], significantly shorter than the single Ge–Ge bond length of 2.463 Å in (GePh2)6 [87]. The Ge atoms in 1 are pyramidal
featuring a trans-bent structure with sum of angles around Ge, Σ∠(Ge) = 348.5°, and a trans-bent angle of 32°. The X-ray
structure of 2 [84, 85] revealed a Sn–Sn bond distance of 2.764 Å, similar to typical Sn–Sn single bond distances of 2.78–
2.82 Å [44, 87], and pyramidal Sn atoms, Σ∠(Sn) = 342°, and a trans-bent angle of 41.6°. The twist angle around the
1.3 Historical Overview 7

double bond in both 1 and 2 is 0°. Based on these geometries, the authors suggested that 1 and 2 result from a double
donor-acceptor interaction of two singlet 1A1 ER2 monomers, in which the filled lone-pair orbital on one E(II) atom
donates electrons into the vacant p orbital of the second E(II) monomer (Scheme 1.1, for E=Sn). However, 1 and 2
were stable only in the solid state, while in solution they dissociate to their corresponding monomers [87]. (tBu2MeS
i)2Sn=Sn(SiMetBu2)2 synthesized in 2004 by Sekiguchi et al. [88] was the first, of only few isolated distannenes, that
are stable also in solution [15, 89].

Sn Sn Sn Sn

C2h
Lone-pair donation

Scheme 1.1 Mode of interactions of two stannylenes producing a trans-bent distannene.

In 1981, in two milestone papers, Brook and West and their coworkers reported independently the syntheses, isolation,
and characterization, including by X-ray crystallography, of the first stable silene, (Me3Si)2Si=C(OSiMe3)1-Ad (3) [90–92]
and disilene, Mes2Si=SiMes2 (4), [93–96], respectively. Silene 3 has a planar skeleton with a Si=C bond length of 1.764 Å,
(r(Si-C) = 1.869 Å in H3CSiH3 [97]). In disilene 4, r(Si=Si) = 2.143 Å (r(Si-Si) = 2.327 Å in H3SiSiH3 [97]), and the Si centers
are slightly pyramidal with Σ∠(Si) = 358° and a trans-bent angle of 12°.
Since these pioneering landmark achievements, this field has witnessed an enormous progress. The key point to success
is the use of large bulky substituents that protect the highly reactive E=E’ double bonds from further reactions. Stable com-
pounds with a variety of E=E and E=E’ bonds (E, E’=Si, Ge, Sn, Pb; E’=group 13–16 elements), as well as compounds
having an extended number of conjugated double bonds, i.e., heavier allenes and butadiene analogues, and heavier benzene
analogues, were synthesized, isolated, and characterized and their chemistry has been reviewed extensively [16, 18, 24,
98–100].
Heavy alkyne analogues RE≡ER (E=Si, Ge, Sn) have also been isolated, but they are much less abundant than heavy
doubly bonded compounds [17, 18, 24]. Unlike the linear alkynes, all known heavy alkynes have a highly trans-bent
geometry [17, 18, 24]. The first digermyne ArGe≡GeAr (5) and distannyne ArSn≡SnAr (6) [Ar = 2,6-(2,6-iPr2-C6H3)2-
C6H3], were reported by Power and coworkers in 2002 [101]. Several digermynes and distannynes with other substitu-
ents on the aryl groups were reported by these authors in 2010 [102]. In 2006, Tokitoh et al. reported the synthesis of
BbtGe≡GeBbt (7) (Bbt = 2,6-[CH(SiMe3)2]-4-[C(SiMe3)3]-C6H2) [103]. The synthesis and characterization by X-ray crys-
tallography of the first stable disilyne RSi≡SiR, R=[SiiPr{CH(SiMe3)2}2] (8) was reported by Sekiguchi’s group in 2004
[104, 105]. Wiberg reported earlier the detection of a relatively stable disilyne, RSi≡SiR, R=SiMe(Sit-Bu3)2 (9), but its
crystal structure could not be obtained [106]. A second known stable disilyne with R = Bbt (10) was reported in 2008 by
Tokitoh’s group [107, 108]. In 1999, Schwarz and Apeloig reported the identification of two silynes HC≡SiX (X=F, Cl) in
neutralization–reionization mass-spectrometric experiments, and demonstrated their microsecond existence under
high-vacuum conditions [109]. More recently, Kato and Baceiredo reported a base stabilized silyne that was stable up to
−30°C. Its X-ray structure revealed a short Si–C bond length of 1.667 Å, indicating the possible existence of a Si–C triple
bond [110, 111]. Germa- [112] and stanna [113]-acetylenes were identified only as intermediates in a low-temperature
matrix or by their trapping reactions [113]. To date, efforts to synthesize isolable E≡C (E=Si, Ge, Sn) triply bonded mol-
ecules were not successful.
It should be emphasized that all known stable multiply bonded heavier alkene and alkyne analogues are substituted with
very bulky substituents that stabilize them kinetically, preventing their dimerization and significantly slowing their other
reactions. Theoretical studies in the last decade pointed to the importance of attractive noncovalent London Dispersion
Forces (LDF) [114] between the bulky substituents, in stabilizing these weakly bonded multiply bonded molecules against
dissociation to their tetrylene monomers [115–118].
In Table 1.2 we provide a list of characterized doubly bonded germanium compounds and some of their important
geometric parameters (both homo- and heteronuclear) that were synthesized since 2010 and not covered in the reviews of
Power [18] and Lee [24]. Table 1.2 also includes, where available, the results of theoretical calculations of these com-
pounds’ structures.
Table 1.2 Doubly bonded organogermanium compounds, R2Ge=E’R’2, reported in 2010–2020.

Twist angle UV-Vis


Compound Ge=E’ (Å) Σ∠(Ge); Σ∠(E’) (°) around Ge=E’ absorption (nm) Reference Publication year

a) Digermenes
11 (tBuMe2Si)2Ge=Ge(SiMe2tBu)2a 2.270 0.3b 7.5 362, 421 [119,120] 2010
12 (tBu2MeSi)2Ge=Ge(SiMetBu2)2a 2.346 358.8 52.8 618 [121a,b] 2011, 2014
R2Ge=GeR2 (calculated)c 2.278 334.0 0.5 3.11d [121b]
a) R = H3Si 2.284 335.1 13.5 2.75d
b) R = Me3Si
2.347 358.7 55.3 2.21d
c) R = tBu2MeSi
13 R R Cl R
Si Si Xyl
N C Ge Ge C N
334.5
Xyl Si Si 2.294 No twist 507 [122] 2015
37.7b
R Cl R R
R = Tip = 2,4,6-iPr3C6H2,
Xyl = 2.6-Me2C6H3
14 R2Ge2=Ge1RSiPh3 346.53(Ge1)e
R = Tip = 2,4,6-iPr3C6H2 2.328 345.24 (Ge2)f 13.6 426 [123] 2019
15 R2Ge=GeRSiMe2Cl
R = Tip = 2,4,6-iPr3C6H2 –g – – 435 [123] 2019
16 R2Ge=GeRSiMePhCl
R = Tip = 2,4,6-iPr3C6H2 – – – NA [123] 2019
17 R2Ge=GeRSiMe2Ph
R = Tip = 2,4,6-iPr3C6H2 – – – 424 [124] 2018
18 R2Ge=GeRSiMe3
R = Tip = 2,4,6-iPr3C6H2 – – – 424 [124] 2018
19 R2Ge2=Ge1RLi · 2dme
7.1 (Ge1)b
R = Tip = 2,4,6-iPr3C6H2,
dme = 1,2-dimethoxyethane 2.284 12.8 (Ge2)b 19.9 435 [124] 2018
20 K22+[(boryl)Ge=Ge(boryl)] 2–
Boryl = (HCDippN)2B,
Dipp = 2,6-iPr2C6H3h 2.392 – – NA [125] 2016
Twist angle UV-Vis
Compound Ge=E’ (Å) Σ∠(Ge); Σ∠(E’) (°) around Ge=E’ absorption (nm) Reference Publication year

21 R2Ge2=Ge1R(C(R’) = O)
476 (Ge=Ge)
R = Tip = 2,4,6- iPr3C6H2; R’ = (a) tBu, (b) – – – [123] 2019
and 386 (C=O)i
2-methylbutan-2-yl and (c) 1-adamantyl

22 (E)-(L)HGe=GeH(L)
322.5
L = N(SiiPr3)R 2.486 No twist 460 [126] 2013
54.1b
R = 2,6-[CHPh2]-4-iPrC6H2
23 (E)-(L)HGe=GeH(L) 2.535 344.8 No twist NA [127] 2019
Dip
N
39.1b
SiMe3
L= B
N
N
Dip
Dip=Diisopropylphenyl
toluene/18-crown-6
24 (E)-(Eind)XGe=GeX(Eind) (a) 2.415 (a) 337.1, 43.29b No twist (a) 406 [128] 2018
j
X= (a) Br and (b) Cl (b) 2.412 (b) 335.9, 44.34b (b) 390

Et Et Et Et

Et Et
Et Et
Eind

25 ArBrGe=GeBrAr a) 2.509 a) 44.6b a) No twist b) 449 [129,130] 2005,


a) Ar = Bbt = 2,6-[CH(SiMe3)2]2-4- b) 2.406 b) 336.3 2016
[C(SiMe3)3]-C6H2
b) Ar = Tbb = 4-tBu-2,6-[CH(SiMe3)2]2-C6H2

26 Bbt r(Ge=Ge) = 2.413 – 494 [131] 2015


Bbt
C(Bbt) = 133.6
∠C(Bbt)-Ge-Ge-
r(C-C) = 1.532
Ge Ge 39.5b

Bbt = 2,6-[CH(SiMe3)2]-4-[C(SiMe3)3]-C6H2

(Continued)
Table 1.2 (Continued)

Twist angle UV-Vis


Compound Ge=E’ (Å) Σ∠(Ge); Σ∠(E’) (°) around Ge=E’ absorption (nm) Reference Publication year

27 Bbt Bbt 2.414 – NA [132] 2018


Ge–C(Bbt) = 134.1
∠C(Bbt)–Ge–
Ge Ge
36.1/43.0b

Ph

28 Tbb Tbb r(Ge=Ge) = 2.416 351.3 – 369 [130] 2016


Ge Ge r(C=C) = 1.362

Ph Ph

Tbb=4-tBu-2,6-[CH(SiMe3)2]2-C6H2

29 EMind 2.430 Ge1/Ge3 360 – 458, 510 [133] 2018


1 Ge2 334
Ge
Ge4 328
EMind 4 Ge2 EMind
Ge

Ge3

EMind
Et Et Et
Et

Me
Me Me Me
EMind

b) Compounds with Ge=E’ bonds


i) Ge=C bonds

30 SiMe3 2.039k 286.6k,l 88.3l,m NA [134,135] 2019


NiPr2
Me

Ge C

Me
NiPr2
SiMe3
Twist angle UV-Vis
Compound Ge=E’ (Å) Σ∠(Ge); Σ∠(E’) (°) around Ge=E’ absorption (nm) Reference Publication year

31 Se 1.921 (Ge-C) 339.9, 338.7 536 [136] 2018


Tbb Tbb
∠SeGeCC = 11.6
Ge Ge 1.375 (C-C)

Ph Ph
32 1.998n 310.5 (Ge) 18.5 NA [137,138] 2017, 2020
359.8 (C)

R O
33 b) 2.007; b) 304.6 (Ge), – a) 442; b) 463; [138,139] 2015, 2020
K
c) 2.036 359.9 (C); c) 422
Ge
Me2Si SiMe2 c) 310.7 (Ge), 359.9 (C)
Me2Si SiMe2
Ge
Me3Si SiMe3
Toluene/18-crown-6
o p
a) R = Mes; b) R = o-Tol ; c) R = 1-Ad

34 a) 2.055 a) 314 (Ge) – a) 447 [140] 2020


r(C-O) = 1.252 359.6 (C) b) 420
b) 2.047 b) 314.5 (Ge1)
r(C-O) = 1.248 298.2 (Ge2)
359.6 ( C)

Toluene/18-crown-6
a) R=Mes; b) R=1-Ad
35 R O 1.835 351.7 (Ge), – 368 [138,139] 2015, 2020
SiMe3 (r(C = O) = 1.388) 360.0 (C)
Ge
Me2Si SiMe2

Me2Si SiMe2
Ge
Me3Si SiMe3 R = Mes

(Continued)
Table 1.2 (Continued)

Twist angle UV-Vis


Compound Ge=E’ (Å) Σ∠(Ge); Σ∠(E’) (°) around Ge=E’ absorption (nm) Reference Publication year

ii) Ge=Si bonds


36 (tBuMe2Si)2Ge=Si(SiMe2tBu)2 2.221 0.6b 7.5 359, 413 [119] 2010
Tip Tip
37 r(Ge=Si) = 2.246 – planar 495 [141] 2021
Si r(Ge-K) = 3.407 4-membered ring
N C Ge K •18-crown-6

Xyl Si
Tip

iii) Ge=E’ (E’=group 13, 15 and 16


elements)
38 Ph2 1.886 X=Cl: 359.6 (Ge), – NA [142] 2020
P 360.0 (B)
B X
Ge
X= Cl, Br Ar*=2,6-Tip2-C6H3 (Tip=2,4,6-iPr3C6H2)

39 Dipp Dipp r(Ge1-P1) = 2.266, – – NA [143] 2016


N N r(Ge1-P2) = 2.272
P1
Ge1 Ge2
N P2
N
Dipp Dipp

Dipp=2,6-iPrC6H3

40 (tBu2MeSi)2Ge=As-Mes* 2.273 360.0 planar 390q, 450r [144] 2018


Mes* = 2,4,6-tBu3-C6H2
41 Eind2Ge=O 1.647 359.8 planar 297 (IR (Ge=O) [145,146] 2012
916 cm−1)
Twist angle UV-Vis
Compound Ge=E’ (Å) Σ∠(Ge); Σ∠(E’) (°) around Ge=E’ absorption (nm) Reference Publication year

42 1.672s 333.9s – NA [147] 2009

N N
R R
Ar
N Ge O
N
Ar
R = Me, iPr; Ar = 2,6-iPr2C6H3

43 1.646 343.6 – NA [148] 2011


N

N
Ar
N Ge
O
N
Ar
Ar = 2,6-iPr2C6H3

Ar
44 2.077 338.5 – NA [149] 2004
N
S
Ge
N OH
Ar
Ar = 2,6-iPr2C6H3

45 Ar S 2.095 – – NA [150] 2017


N Ge
PCy2
N
Ar
Ar = 2,6-iPr2C6H3; Cy = C6H11

46 Ar S1 S3 2.068 – – NA [150] 2017


(Ge-S2) = 2.234
N Ge PCy2
S2
N
Ar
Ar = 2,6-iPr2C6H3

(Continued)
Table 1.2 (Continued)

Twist angle UV-Vis


Compound Ge=E’ (Å) Σ∠(Ge); Σ∠(E’) (°) around Ge=E’ absorption (nm) Reference Publication year

47 Ar Se 2.221 – – NA [150] 2017


N Ge
PCy2
N
Ar
Ar = 2,6-iPr2C6H3

48 Ar Se 2.163 – – NA [150] 2017


N Ge
P(SiMe3)2
N
Ar
Ar = 2,6-iPr2C6H3

c) Extended doubly bonded compounds

49 2.268, 2.290t, u – – 451 [151] 2017

BbtGe GeBbt
Si
Bbt = 2,6-[CH(SiMe3)2]-4-[C(SiMe3)3]-C6H2

d) Germabenzenes

50 Tbb Tbb 2.312 345.3 – 383 [152] 2015


Ge Ge

Tbb = 4-tBu-2,6-[CH(SiMe3)2]2-C6H2
Twist angle UV-Vis
Compound Ge=E’ (Å) Σ∠(Ge); Σ∠(E’) (°) around Ge=E’ absorption (nm) Reference Publication year

51 R1 R1 r(Ge-C1) = 1.890v 357.3v – 442v [153] 2019


C1 C2 r(Ge-C3) = 1.866v
Tbb Ge Ge Tbb

C3 C4
R2 R3
a) R1 = R2 = R3 = Ph
b) R1 = R2 = R3 = Et
c) R1 = Ph, R2 = Me, R3 = nPr
Tbb = 4-tBu-2,6-[CH(SiMe3)2]2-C6H2

a
A preliminary X-ray diffraction – poor refinement;
b
trans-bent angle;
c
At B3LYP/6–31G(d);
d
HOMO-LUMO gap in eV;
e
Trans-bent angle = 23.7°;
f
Trans-bent angle = 21.3°;
g
Unavailable data; X-ray crystal structure could not be obtained. Characterized as digermenes by their NMR spectra;
h
For previously synthesized digermene dianions, see Ref. [46] cited in Ref. [125] [Pu, L. et al. J. Am. Chem. Soc., 1998, 120, 12682–12683]. For a review, see: Wang, Y. and Robinson, G. H. Chem.
Commun., 2009, 5201–5213;
i
For R’ = 1-adamantyl;
j
For earlier papers on dihalodigermenes, see references [5] and [6] cited in Ref. [128];
k
Best described as a germanium/carbon ylide, with the negative charge located at the germanium atom;
l
Indicates the presence of an active lone pair at the germanium atom, and excludes the possibility of a Ge–C (of the three-membered carbocycle) π bonding;
m
The three-membered ring is oriented almost perpendicular to the germole ring;
n
r(K-Ge) = 3.946 Å, r(K-O) = 2.782 Å, r(C-O) = 1.241 Å;
o
r(K-Ge) = 3.855 Å, r(K-O) = 2.730 Å, r(C-O) = 1.236 Å;
p
r(K-Ge) = 3.613 Å, r(K-O) = 2.740 Å, r(C-O) = 1.231 Å;
q
HOMO [π(As = Ge)] – LUMO [π*(As = Ge)];
r
HOMO-1 [n(As), lone pair] – LUMO [π*(As = Ge)];
s
For R = Me;
t
Best described as R2Ge:→Si(0)←:GeR2;
u
v
∠GeSiGe = 80.1°;
For compound 51a.
16 1 Computational and Theoretical Aspects of Structure and Bonding in Doubly Bonded Organogermanium Compounds

1.4 Doubly Bonded Compounds

The structures of RR’E=ERR’ (E=Si, Ge, Sn, Pb) compounds were studied and analyzed extensively by both theory and
experiment over the last four decades. These studies were reviewed in many comprehensive reviews [15, 16, 18, 24, 25, 44,
45, 70, 154, 155]. The main structural and bonding aspects of these compounds, which were previously reviewed, are
briefly summarized below. New insights gained more recently regarding the nature of bonding in these doubly bonded
compounds have been added.
The most surprising structural feature of the heavier group 14 alkene analogues, R2E=ER2, which attracted great
attention, is probably their nonplanar structures (Table 1.1) [15, 16, 18, 24, 25, 44, 45, 154, 155], in sharp contrast to the
planar structure of alkenes R2C=CR2. This contrast indicates that simplistic analogy between carbon and its heavier ana-
logues is wrong, and that the heavier group 14 elements have entirely different bonding characteristics. This unexpected
experimental observation was predicted by theory before many of these compounds were synthesized [156, 157], and in
many cases after their synthesis, the experimental bending angles were nicely reproduced by theory [44, 45, 70, 158, 159].
The successful computational predictions in this field that preceded experiments were among the early victories of compu-
tational chemistry, which in the 1970s and 1980s still met considerable scepticism by experimentalists.

1.4.1 Bonding Models


As mentioned above, for many decades the heavier group 14 elements were believed to be reluctant to form π bonds
(classical “double bond rule” [79, 80]). After the first multiply bonded compounds were isolated, new questions of great
theoretical relevance arose. Why, in contrast to carbon, do its heavier analogues form compounds with very different
geometries than the analogous carbon compounds? For example, why are doubly bonded compounds non-planar and
have trans-bent structures? Why are triply bonded compounds bent and not linear? What is the nature of these distorted
π bonds? Are they “real” double bonds? These questions raised a burst of theoretical research accompanied by vibrant
debate in the last four decades [18, 25, 44, 45, 70, 160–164]. Below we review the different views of theory on these
intriguing fundamental questions.
A major fundamental difference between elements of the first (C to F) and higher rows of the Periodic Table is their
most probable valence orbital radii. For carbon, the most probable radius of the 2s orbital is relatively expanded due to
Pauli repulsion from the 1s electrons, while the 2p orbital that is not shielded by an inner p shell is relatively contracted.
This results in very similar radii of the 2s and 2p orbitals. Moving down group 14, the atomic np orbitals are much more
expanded as a consequence of intraatomic Pauli repulsion from orthogonal inner shell (n-1)p orbitals. Consequently, the
radii of the ns and np orbitals (n > 2) diverge significantly
[27]. Δr = rp–rs exhibits a zigzag increase from Si to Pb,
caused by d-block contraction (Ge) and relativistic effects
(Sn, Pb) (Table 1.1 and Figure 1.4). The similar radii of the
s and p valence orbitals of carbon (and other first row ele-
ments) facilitates hybridization. In contrast, for the
heavier main group elements, hybridization is much less
favorable due to the dissimilar radii of their valence ns
and np orbitals. Kutzelnigg introduced the term “hybrid-
ization defects” for this phenomena [27]. The increasing
“hybridization defects” down group 14 have a crucial
effect on the structures and bonding of heavier group 14
compounds [70, 165]. The increase in the trans-bending
angle, and in the planarization energy of the trans-bent
H2E=EH2 (C2h) (Table 1.3) moving down group 14, are
manifestations of increasing “hybridization defects” [27,
42, 115]. Similarly, the much higher inversion barriers of
e.g., PH3 (35.1 (calculated) [166]) vs. NH3 (5.8−6.0 kcal/
mol [166–168] (calculated)) and SiH3- (26 kcal/mol
(MP2/6-31+G(d)) [169] vs. CH3- (2.5 kcal/mol (CCSD(T)/
Figure 1.4 Calculated ns and np orbital radii of maximal
electron density of group 14 elements [27, 37, 115]. Reproduced aug-cc-PVQZ) [169]), was also rationalized by the smaller
with permission from reference [115]. Copyright (2020) American tendency for hybridization of the ns and np orbitals in the
Chemical Society. heavier elements [115, 166]. The decrease in the
1.4 Doubly Bonded Compounds 17

dissociation energy of H2E=EH2 (C2h) to their EH2 fragments (see below, Table 1.3, section 1.4.2) was also rationalized by
the increasing destabilization of heavy alkene analogues, due to the lack of effective iso-valent hybridization and the
enhanced stability of the EH2 fragments [42]. The non bonding lone-pair orbital of EH2 has mainly s orbital character,
thus, lowering its energy, while the E–H bonds are essentially made of non-hybridized p orbitals strengthening the E–H
bonds and stabilizing the divalent hydrides (“inert pair effect”) [42, 165]. Electronegative substituents enhance “hybrid-
ization defects” in heavier p-block elements, thus destabilizing the olefin-like molecules, while their effect on the low-
oxidative states (e.g., ER2) where hybridization is less important, is marginal [42, 165].
The Pauli repulsion, caused by core electrons in heavier group 14 elements, has an important effect on the E–E π and
σ bonds strengths. DFT calculations by Ziegler and Jacobson have shown that the increase in core-core Pauli repulsion
between the E atoms forming the E=E bond (i.e., interactions of the valence orbitals of one E atom with the inner shell
orbitals of the other E atom forming the bond, termed interatomic Pauli repulsion [170, 171] or inner shell repulsion
[79]), weakens the E–E σ bond and increases the interatomic E–E distance, which in turn also lowers the overlap of the
p orbitals weakening the π bond [170, 171]. It is argued that the linear p overlap in the σ bonds is more effective than the
sideway π overlap, making the π bonds more sensitive to changes in the bond distance. In contrast, a study of the varia-
tion of orbital overlap as a function of internuclear distance found that π bonds (not supported by an underlying σ bond)
can be shorter than σ bonds and are not inherently weak [172]. Thus, it is the underlying σ bond that actually determines
the E=E bond strength. Analysis of the nature of bonding in trans-bent HE≡EH (E=Si, Ge, Sn) emphasizes the impor-
tance of the E–E σ bond in stabilizing the trans-bent structures. VB calculations [173, 174] show that the σ-frame prefers
trans-bending and the π bonding opposes this distortion. In HC≡CH, the destabilization of the π bond upon bending
overrides the propensity of the σ-frame to distort, while in the heavier congener molecules, the stabilization of the
σ-frames by trans-bending overcomes the destabilization of the total π bonding and thus results in trans-bent molecules
[173, 174]. The increasing core-core repulsion down the group leads to a significant decrease in the stability of a E=E
double bond as well as of a E–E single bond; thus for E2H4, E=Sn and Pb, the hydrogen dibridged isomer, HE(μ‒H)2EH
becomes more stable than the trans-bent doubly bonded isomer H2E=EH2 and the mono-bridged isomer H2E(μ‒H)EH
[70, 162]. For E2R4 (R=electronegative substituent) “hybridization defects” are enhanced significantly and the dibridged
isomer becomes the global minimum for E=Ge, R=F, Cl, Br. For E=Ge, R=F the dibridged isomer is more stable than
F2Ge=GeF2 by 18.3 kcal/mol (at BP86/def2-TZVpp) [70]. At MP2(FCI)/cc-pwCVTZ(-PP) R2Ge=GeR2, R=F, Cl, Br, I,
were located as transition states [175]. At a variety of other DFT levels, F2Ge=GeF2 dissociates to two GeF2 fragments
upon geometry optimization [176]. See section 1.4.2.2.2.1 for a detailed discussion on the effect of substituents on the
structures and stability of various Ge2R4 isomers. The above discussion focuses on the effect of the inherent difference
in the radial extension of the valence ns and np orbitals on structure and bonding. Other bonding models are discussed
below.
An alternative rationale for the preferred trans-bent geometries of heavy group 14 analogues of alkenes and acetylenes
(R2E=ER2 and RE≡ER, respectively) is arising from second order Jahn-Teller (SOJT) effects [43, 87, 158–161, 171, 177].
The SOJT effects are caused by mixing of occupied π (HOMO) and empty σ* (LUMO+1) molecular orbitals which acquire
the same symmetry (bu) upon trans-bending (Figures 1.5 and 1.6). The mixing of σ and π* orbitals is less important [171].
The extent of orbital interactions and the resulting stabilization of the trans-bent structure is inversely proportional to the
energy difference between the interacting orbitals.

Figure 1.5 A schematic description of the π-σ* mixing (SOJT) upon trans-bending of R2E = ER2.
18 1 Computational and Theoretical Aspects of Structure and Bonding in Doubly Bonded Organogermanium Compounds

Figure 1.6 Orbital symmetries and orbital energies of D2h (left) and C2h (right) H2Ge=GeH2. Reprinted with permission from Ref. [159].
Copyright (2020), Royal Society of Chemistry.

The π-σ* energy separation in planar D2h H2E=EH2 decreases upon moving to the heavier elements in group 14; ΔE (eV,
at ωB97X-D/def2-TZVPP) = 14.54 (E=C), 9.78 (E=Si), 9.05 (E=Ge), 7.44 (E=Sn) [159]. As a result, the orbital interaction

lization energies associated with the π → σ* mixing, which increase in the order (in kcal/mol, at B3LYP/def2-TZVPP//
energies and electron delocalization increase going down group 14. This is manifested also in the NBO second order stabi-

CCSD(T)/def2-TZVPP): 3.64 (Si), 8.83 (Ge), 24.41 (Sn) [178], 47.0 (Pb) [179]. Simultaneously, the trans-bending angles
widen [16, 18]. The barriers for planarization (or linearization of heavy acetylene analogues) increase down group 14 [43,
70, 159, 171], thus, For H2E=EH2 they are (in kcal/mol): 0.0 (E=C), 0.9 (E=Si), 2.7 (E=Ge), 10.1 (E=Sn) [159], 23.2 (E=Pb)
[171] (see also Table 1.3), and correlate linearly with the π-σ* energy separation for heavy E-elements (Figure 1.7) [159].

increase in the π(E=E) → σ*(E–E) electron delocalization [178].


The E=E Wiberg bond indices (WBIs) decrease from 2.05 (E=C) to 1.90 (E=Si), 1.79 (E=Ge), 1.54 (E=Sn), reflecting the

Figure 1.7 Planarization (inversion) barriers of H2E=EH2 vs. the π-σ* energy gap (ΔE, eV) for ethylene (D2h, ΔE=14.54) and its heavier
congeners (E=Si (ΔE=9.78), Ge (ΔE=9.05), and Sn (ΔE=7.44)) in C2h symmetry. Reproduced with permission from Ref. [159]. Copyright
(2020) Royal Society of Chemistry.
1.4 Doubly Bonded Compounds 19

Another model for understanding the structures of R4E2 compounds is based on the ground state multiplicity of the R2E
fragments and their singlet-triplet energy gap (ΔEST). Carbenes, R2C:, R = alkyl, aryl, are ground state triplets (3B1), while
X2C:, X = electronegative substituents, e.g., halogen, NR2, OR, are ground state singlets [180, 181]. Heavier R2E: are all
ground state singlets (1A1) [182] (Figure 1.8a, Tables 1.3, 1.5 and 1.6). The reasons for this reversal in multiplicity were
analysed [183]. The association of two triplet ER2 (3B1) molecules, forms a planar alkene-like structure (Figure 1.8b). In
contrast, for two singlet ER2 (1A1) molecules to form a planar double bond, they have first to be excited to the triplet state,
paying the cost of the singlet-triplet energy gap (ΔEST); consequently the E=E bond is weakened [184]. Instead, singlet ER2
(1A1) fragments prefer to interact via two donor-acceptor interactions, in which the lone-pair orbital of one ER2 fragment
donates electrons into the empty p orbital of the other ER2 fragment, forming a trans-bent doubly bonded compound
(Figure 1.8c) [70, 160–162]. Based on the trans-bent X-ray structure of the first isolated distannene [(Me3Si)2HC]2
Sn=Sn[CH(SiMe3)2]2, (2), Lappert et al. suggested that 2 results from a donor-acceptor interaction of two singlet SnR2
monomers (Scheme 1.1) [84–87].

Figure 1.8 Bonding models of the interaction of two triplet (3B1) and singlet (1A1) R2E fragments. (a) triplet and singlet R2E;
(b) Interaction of two R2E in the triplet state, forming a planar R2E=ER2; (c) Lone-pair donor-acceptor interaction of two R2E in the
singlet state, forming a trans-bent R2E=ER2; (d) E–X bond-pair donation, forming a dibridged structure; (e) Lone-pair donation
from X forming a dibridged structure.
20 1 Computational and Theoretical Aspects of Structure and Bonding in Doubly Bonded Organogermanium Compounds

Descending group 14, the lone-pair orbital of R2E acquires more s-character and is stabilized (“inert pair effect”) and is
less prone to donate electrons to the partner R2E fragment, reducing the tendency to form an E=E bond. At the same time
the E–R bonds become better donors, donating charge into the empty p orbital of the second R2E fragment, thus favoring
dibridged isomers (Figure 1.8d) [70]. Substituents that enhance the “inert pair effect”, e.g., F, prefer dibridged structures
over the doubly bonded isomers for the heavier element compounds [70]. Another explanation for the preference of a
dibridged structure for E2R4 when R are electronegative substituents bearing lone pairs (e.g., F), is the mutual donation of
the substituents’ lone pair into the empty p orbital of the tetrylene (Figure 1.8e) [185, 186].
In a series of important papers, by Carter and Goddard [184] and by Malrieu and Trinquier [160–163, 177], they sug-
gested simple criteria that predict the occurrence of trans-bent E=E double bonds (B), the degree of their trans-bending, as
well as the occurrence of other structural isomers, e.g., a dibridged isomer (C) (Scheme 1.2). These criteria, known as the
“CGMT” criteria, compare ΔEST of the R2E fragments that compose E2R4 with the overall σ+π bond energy (Eσ+π) of
planar R2E=ER2. According to the CGMT approach, for homonuclear R4E2 molecules, a trans-bent doubly bonded struc-
ture (B) is predicted to be more stable than the planar structure (A) or the dibridged (C), when ΔEST of ER2 is in the range
given in Eq. (1.1):

¼ Eσ+π ≤ ∆EST ≤ ½ Eσ+π  (1.1)

When ΔEST is smaller than ¼ Eσ+π, a planar doubly bonded species is predicted, while when it is larger than half the
double bond energy a dibridged isomer C is predicted. For E heavier than carbon, the double bond energy (Eσ+π, ΔEdiss)
decreases and ΔEST increases, leading to trans-bent structures with increasing bending angles as E is heavier, or to dibridged
structures (Table 1.3) [70].

1.4.2 Homonuclear Ge=Ge Compounds


1.4.2.1 The Parent H2Ge=GeH2 and Its Isomers: Structures and Potential Energy Surface
As emphasized above, unlike H2C=CH2 which has a planar D2h skeleton (A), calculations show that H2E=EH2
(E=Si-Pb) possess a trans-bent C2h structure B (Table 1.3). The planar structure is a transition state between two iso-
meric trans-bent structures B [44, 45, 70, 158, 159]. Upon moving from carbon to heavier group 14 elements, the
trans-bending angle and the barrier for planarization increases from 31.4° and 0.8 kcal/mol for E=Si to 47.9° and
4.2 kcal/mol for E=Ge, reaching a bending angle θ of 53.2° and a barrier for planarization of ca. 27 kcal/mol for E=Pb
[90] (Table 1.3). A SCF calculated PES of the planarization energy vs. the trans-bending angle (Figure 1.9) shows the
significant increase of the barrier for planarization on moving down group 14 [87]. Simultaneously, the dissociation
energy of H2E=EH2 to two EH2 (1A1) fragments is reduced, i.e., ΔEdiss is (at 0 K, in kcal/mol, at BP86/def2-TZVPP
[70]): 178.8 (E=C, dissociation to two triplet CH2 fragments); 67.8 (E=Si); 51.3 (E=Ge); 28.9 (E=Sn, at LDA/TZ+p)
[171]; and 10 (E=Pb, at LDA/TZ+p) [171], indicating that actually H2Pb=PbH2 dissociates at room temperature
(Table 1.3).
The calculated Ge=Ge bond length in H2Ge=GeH2 is 2.275–2.346 Å (depending on the computational level,
Figure 1.10, Table 1.3), shorter than that of a Ge–Ge single bond length of 2.40–2.46 Å (exp. r(Ge-Ge) = 2.403 Å [187])
in H3Ge-GeH3 [188]. The calculated trans-bending angle is ca. 42–47° [43–45, 70, 188, 189]. The trans-bending angle
in H2Ge=GeH2 is significantly larger than in H2Si=SiH2 (e.g., θ = 31.4°). The experimental Ge=Ge bond lengths and
the trans-bending angles in isolated digermenes, R2Ge=GeR2 (R are bulky alkyl and aryl ligands), span over a large
range of 2.2–2.49 Å [15, 18, 24] (Figure 1.11a) and of ca. 15–45° [18] (Σ∠Ge = 327–358°), respectively [15, 18, 24]. The
bond distances in 60% of the isolated digermenes (Figure 1.11a) are in the range of those calculated for H2Ge=GeH2
(Figure 1.10).
As discussed above, several additional structural isomers, C–F (Scheme 1.2), were located by calculations as minima on
the E2H4 (E=C–Pb) potential energy surface (PES). The calculated relative energies of isomers B, C and F and the struc-
tural parameters of isomer B are presented in Table 1.3. The geometrical parameters of isomers B-F for E=Ge, R=H are
presented in Figure 1.10. The most stable isomer for E=Si and Ge is the doubly bonded trans-bent isomer B, but for E=Sn
and Pb, the most stable isomer is the trans-dibridged isomer C (Table 1.3) [44, 45, 70, 160–162]. The hydrogen-singly
bridged isomer E is a minimum on the PES for E=Ge [189], Sn, and Pb [162, 163]. For Sn2H4 and Pb2H4 the calculated sta-
bility order of the isomers is: C > E > F > B, i.e., the doubly bonded isomer B is the least stable isomer [163].
The PES of Ge2H4 is relatively flat and spans over a range of ca. 14 kcal/mol compared to the PES of Si2H4 that spans over
a range of 22 kcal/mol (Table 1.3) [70, 188, 189, 194].
H2Ge=GeH2 and HGe-GeH3 are nearly degenerate, with H2Ge=GeH2 being slightly favored (at most by ca. 3 kcal/mol,
depending on the level of calculation, Tables 1.3 and 1.4) [70, 188, 189, 194]. The calculated barrier for the isomerization of
1.4 Doubly Bonded Compounds 21

Table 1.3 Calculated geometric parameters and relative energies of isomers B, C and F (ΔE) on the PES of E2H4 (E=Si, Ge, Sn, Pb), the
dissociation energy of B to EH2 fragments (ΔEdiss) and the singlet-triplet energy gaps (ΔEST) of the corresponding fragments.a,b

H2E=EH2 (B) HE(μ-H)2EH (C) HEEH3(F, 1A’) EH2

E r(E=E) θ ΔEc ΔEdissd ΔEe ΔEdissf ΔE ΔESTg


h h i j
C 1.322 180 (0.0) 178.7 164.7 27.3 72.3 ;79.1 −15.8;
−9.03k;
−(8.99l)
Si 2.178 31.4 0.0 (0.8) 67.8 [100.0] 18.7 31.2h 7.9m 16.5; 19.6n;
(21.0o)
Ge 2.300 47.9 0.0 (4.2) 51.3 [93.7] 6.8 26.9h 3.4p 23.3
q h j
Sn 2.175 50.6 0.0 (11.5) 28.9 [74.4] −9.4; 40.5; 33.2 −2.1 27.3
−9.1j
Pb 2.922 53.2 0.0 (27.3) 10.0q [60.4] −21.8; 38.5; 28.7h −16.4 j 35.5
−23.9 j
a
Bond lengths in Å, bond angles in degrees. Energies in kcal/mol;
b
From Ref. [70], at BP86/def2-TZVPP, unless stated otherwise. For other calculated values see Refs. [43–45, 162, 171, 178];

Dissociation energies: H2E=EH2 (C2h) → 2H2E: (1A1). In brackets, ΔEdiss of H2E=EH2 (planar, D2h) → 2H2E: (3B1);
c
In parenthesis, the relative energy of the planar doubly bonded isomer A;
d
e
Energy difference between isomers C and B;
f
Dissociation of C to 2H2E: (1A1);
g
Experimental values in parenthesis;
h
From Ref. [162], at CI//HF/DZ;
i
From Ref. [190];
j
From Ref. [162] at CI/DZP;
k
At CASSCF-SOCI with the TZ3P(2f,2d)+2diff basis set [191];
l
See references cited in Ref. [191];
m
At the G1 level, from Ref. [190];
n
At MRSOCI [192];
o
From Ref. [193];
p
At CCSD(T)/aug-cc-pVTZ//B3LYP/6-311G(d,p) [189], see Table 1.6 for values at other computational levels;
q
At LDA/TZ+P with nonlocal corrections [171].

Scheme 1.2 Possible isomers on the R4E2 potential energy surface.

and barriers were calculated for the isomerization of H2Ge=SiH2 → HGeSiH3. The isomerization energy of H2Ge=SiH2 →
H2Ge=GeH2 to HGeGeH3 at several CCSD levels of theory is 12.9–14.0 kcal/mol (Table 1.4) [195]. Similar reaction energies

HSiGeH3 is significantly higher than the isomerization of H2Ge=SiH2 → HGeSiH3. In part, this results from the different
bond strength of the Si–H and Ge–H bonds in H2Ge=SiH2 and in the corresponding products HGeSiH3 or HSiGeH3. The
22 1 Computational and Theoretical Aspects of Structure and Bonding in Doubly Bonded Organogermanium Compounds

Figure 1.9 Variation of energy (kJ/mol) with


the trans-bending angle of H2E=EH2 calculated
at the SCF level. The curves for C2H4 and Si2H4
40 C2H Si H were calculated for non-adiabatic bending (i.e.,
4 2 4
with constant values for E–E, E–H bond
distances and H–E–H bond angles). The curves

–1
∆ESCF/KJ MOL
for Ge2H4 and Sn2H4 were calculated for
30
adiabatic bending (i.e., with re-optimized values
for the E–E bond distances and H–E–H angles).
Adapted with permission from Ref. [87].
20 Copyright (1986) Royal Chemical Society.

10

–60 –40 –20 20 40 60


Ge H θ/˚
2 4

–10

–20

–30 Sn H
2 4

(a) 1.545 107.8 (b) 109.1


2.275 108.6
2.305
2.278 107.5
2.310 107.5
H 106.4
115.6 2.311 H
2.346
B Ge Ge 1.537
Ge2H4 1Ag 1.540
H
116.7 1.548
H
116.0 1.556
115.4 1.569
115.1
114.5
1.596
104.9
1.764 91.8 1.589 Neutral (C2h 1Ag)
105.6
75.1
87.7 1.771 74.4 1.534
C D 1.539 113.3
1.545 114.1
Ge2H4 1Ag Ge2H4 1A1
H 1.553 113.8 2.510
1.536 1.557 114.6 2.500
1.543 114.3 2.528
1.548 2.521
1.557 2.554
1.561 Ge Ge
113.8 1.543 1.583
H 110.5 89.7
1.929 1.710 1.593
2.528 107.4 H 88.6
110.5 1.600
128.2 89.0
1.595 86.7 1.546 106.8 110.8 1.612
107.0 106.8 110.9 88.1 H
89.0 110.9 1.617
2.412 1.548 1.598 106.1 111.2 88.7
106.7 106.1
E F
Ge2H4 1A
Ge2H4 1A' Neutral (Cs 1A')

Figure 1.10 (a) B3LYP/6-311G(d,p) optimized structures of isomers located on the G2H4 PES. Adapted with permission from Ref.
[189]. Copyright (2006) Elsevier B.V., and (b) Optimized structures of trans-bent H2Ge=GeH2 and of germylgermylene, at a variety of
DFT levels. Bond lengths in Å and bond angles in degrees. Adapted with permission from Ref. [188]. Copyright (2002) Wiley Periodicals.
1.4 Doubly Bonded Compounds 23

(a) (b)
Distribution of Ge=Ge bond lengths in digermenes Distribution of Si=Si bond lengths in disilenes

2.26–2.28 Å >2.28 Å
>2.50 Å 2.20 –2.29 Å 2.23–2.25 Å 3% 1%
4% 14% 8%

2.40–2.49 Å 2.14–2.16 Å
23% 2.20–2.25 Å 34%
22%
2.30 –2.39 Å 2.17–2.19 Å
59% 32%

Figure 1.11 Distribution of (a) Ge=Ge bond lengths and (b) Si=Si bond lengths in experimentally isolated digermenes and disilenes.
Reprinted with permission from Ref. [25]. Copyright (2013) Elsevier.

H2E=E’H2 → HEE’H3 isomerization (E and E’=Ge, Si and C).a


Table 1.4 Calculated reaction energies (ΔE) and reaction barriers (ΔE#) for the

Reaction ΔE ΔE#

H2C=CH2 → HCCH3 74.9b 75.0b


H2Si=SiH2 → HSiSiH3 7.9c 17.3d
H2Ge=GeH2 → HGeGeH3 2.3 to −2.3 12.9–14.0
H2Ge=SiH2 → HGeSiH3 −3.2 to −7.0; −3.6e, −4.8f, −6.3 g 14.4 to 15.7; 11.8e
H2Ge=SiH2 → HSiGeH3 14.1f; 9.4g –
a
In kcal/mol, at various CISD and CCSD levels. From Ref. [195] unless stated otherwise.
For more data, see the citations in previous reviews [44, 45];
b
At MP4/6-311G(d,p) [197];
c
At the G1 level [198];
d
MP3/6-31G(d) [199];
e
At CCSD(T)/6-311(d,p)//B3LYP/6-311G(d,p), from Ref. [200];
f
At CCSD(T)/cc-pVTZ (H, Si) and CCSD(T)/cc-pVTZ-pp (Ge), from Ref. [201];
g
At CISD+Q, from Ref. [196].

Si–H and Ge–H bond strength in H2Ge=SiH2 are 68.5 kcal/mol vs. 57.4 kcal/mol, respectively; in HGeSiH3 the Si–H bond
strength is 90.6 kcal/mol, and the Ge–H bond strength in HSiGeH3 is 84.0 kcal/mol [196]. The contrast with the extremely
high barriers (75 kcal/mol) for the H2C=CH2 to HCCH3 isomerization demonstrates again that analogy between C and its
heavier analogues is misleading.
For Sn2R4, of the four isomers B, C, D and F, a trans-dibridged tin derivative Ar’Sn(μ-H)2SnAr’ (Ar’ = (H3C6-2,6-(C6H2-
2,4,6-iPr3)2) [202] and a dihydrido stannylene derivative Ar”SnSnH2Ar” (Ar” = C6H-2,6-(C6H2-2,4,6-iPr3)2-3,5-iPr2) [203]
were isolated and characterized by X-ray crystallography. Calculations for Ar2H2Sn2 show that stannylene isomer F
becomes more favorable than the dibridged isomer C as the size of the Ar ligand increases [204].
The elusive H 2Ge=GeH2 was isolated and characterized spectroscopically only as a donor-acceptor complex, via

LB1(or LB2)→H2Ge=GeH2→LA, (52, 53), LB1=N-heterocyclic carbene, LB2=nucleophilic N-heterocyclic olefin,


coordination with a Lewis base (LB)/Lewis acid (LA) combination, which stabilize the reactive digermene, i.e.,

digermene, LB2→Cl 2Ge=GeCl2→LA (54) [205] (Figure 1.12), and mixed E=E’ molecules, LB1→H 2Si=EH2→LA,
LA=W(CO)5 (Scheme 1.3 and Figure 1.12) [205]. A similar strategy was used to synthesize and isolate tetrachloro-

E=Ge, Sn [206].
24 1 Computational and Theoretical Aspects of Structure and Bonding in Doubly Bonded Organogermanium Compounds

Scheme 1.3 (a) Possible resonance structures of H2Ge=GeH2 and (b) Stable digermene Lewis acid/Lewis base complex.

Figure 1.12 X-ray molecular structures of LB1→H2Ge=GeH2→W(CO)5 (52) and LB2→Cl2Ge=GeCl2→W(CO)5 (54). Reproduced with
permission from Ref. [205]. Copyright (2013) American Chemical Society.

The Ge–Ge bond length in 52 is 2.421 Å which is in the range of a Ge–Ge single bond (see above). Calculations at
M06-2X/cc-pVTZ of model systems of 52 and 53 where the bulky 2,6-iPr2-C6H3 substituent at N was replaced by a Me
group, designated 52Me and 53Me, predict a polar Ge–Ge linkage with 60% of the charge density residing on the W atom
bonded to Ge, due to electron withdrawing by the carbonyl groups. The Wiberg bond index (WBI) of the Ge–Ge bond is
0.87 and 0.91 in 52Me and 53Me, respectively. Comparison with the Ge–Ge WBI of H3Ge-GeH3 of 0.97 and that of H2Ge=GeH2
of 1.79 supports a single covalent bond in 52 and 53, and the bonding mode shown in Scheme 1.3(b). Similar to 52 and 53,
also the parent H2Si-GeH2 and H2Si-SnH2 were isolated as complexes with W(CO)5 and NHC ligands [206]. WBIs calcu-
lated at B3LYP/cc-pVDZ-pp are 0.88 for the Si–Ge bond and 0.79 for the Si–Sn bond, supporting the bonding shown in
Scheme 1.3(b) also for these compounds [206].
Digermene, H2Ge=GeH2 (X1Ag), and germylene HGeGeH3 (X1A’) and their fully deuterated isotopomers were observed by
IR spectroscopy in GeH4 and GeD4 matrices at 12 K, upon irradiation of the matrices with energetic electrons. The digermene
and D4-digermene were identified by their absorptions at 845 cm−1 (ν11), and at 1476 cm−1 (ν5), attributed by calculations to
a GeH2 scissor vibration, and to the GeH2 asymmetric stretching vibration, respectively [189]. Calculations predict also two
stronger bands at 2125 cm−1 (ν5, GeH2 asymmetric stretching) and at 2093 cm−1 (ν10, GeH2 symmetric stretching). The slightly
less stable HGe-GeH3 (F) and its deuterated derivative were also identified by their absorption at 787 cm−1 and 558 cm−1(ν5,
GeH3/GeD3 umbrella), respectively. The vibrational frequencies and IR intensities of H4Ge2 isomers B-F were calculated at
1.4 Doubly Bonded Compounds 25

B3LYP/6-311G(d,p) aiding in identifying the experimentally observed species [189]. Tetra-methyl digermene was detected in
nitrogen and argon matrices at 5 K and were identified by Raman and IR spectroscopy [207, 208].
H-migration from the sterically hindered dihydrodigermene Ar’HGe=GeHAr’ (Ar’ = C6H3-2,6-(2,6-iPr2C6H3)2), 55
occurs (Eq. 1.2) in the presence of PMe3, producing a base-stabilized germylene, Ar’Ge-GeH2Ar’, 56 [209]. The experi-
mental X-ray structure of 55 exhibits a trans-bending angle of 20.5° and a Ge–Ge bond length of 2.372 Å [209], longer than
the calculated values for H2Ge=GeH2 of 2.275–2.346 Å [188] (see also Table 1.3 and Figure 1.10), and it is in the upper limit
of the measured bond lengths in isolated Ge=Ge compounds, of 2.20–2.45 Å (Figure 1.11) [15, 18, 24, 25]. The Ge–Ge bond
length in germylene 56 is 2.530 Å, similar to the calculated Ge–Ge bond distance in HGe–GeH3 (Figure 1.10), and the C–
Ge–Ge bond angle is 101.6° being larger than calculated for HGeGeH3 of 89° (Figure 1.10) [209]. Other examples of stable
germylgermylenes are limited, e.g., (2,6-Mes2H3C6)GeGetBu3, an analogous germyl stannylene was isolated as well [210].

PMe3
Ar' Ar' H
toluene H
H (1.2)
Ge Ge + PMe3 Ge Ge
H
Ar' Ar'
55 56

1.4.2.2 Substituent Effects


1.4.2.2.1 Correlation between Structures, Bond Dissociation Energies and ΔEST
In agreement with the CGMT approach, Karni and Apeloig [211] found a linear correlation between ΔEST of HRSi (R=Li, BeH,
BH2, SiH3, F, OH, NH2) and the trans-bending angles at the Si atoms in HRSi=SiH2 and its dissociation energy. Electronegative
and π-donor substituents, F, OH, and NH2, increase ΔEST of the silylene fragments, and consequently, increase the trans-bending
angle and reduce dramatically the dissociation energy of disilene to its fragments [211] and the energy difference between the disi-
lenes and their dibridged isomers C (Scheme 1.2) [185]. At MP2/6-31G(d), the dibridged structure HSi(μ-H)2SiH (C) is less stable
than H2Si=SiH2 (B) by 25.8 kcal/mol (18.7 kcal/mol at BP86/def2-TZVPP, Table 1.3), while the fluorine substituted dibridged
isomers FSi(μ-H)2SiF and HSi(μ-F)2SiH are by only 6.1 and 9.5 kcal/mol, respectively, less stable than HFSi=SiHF [185]. Trinquier
predicted computationally that F2Si=SiF2 isomerizes without a barrier to the dibridged isomer FSi(μ-F)2SiF [161].
Chen and coworkers [212] studied computationally the structures and bond dissociation energies of R2Ge=GeH2,
R2Ge=SiH2 and H2Ge=SiR2 (R=H, CH3, NH2, OH, F, Cl) and also found a good linear correlation between ΔEST of GeR2 or
SiR2, the trans-bending angles and the E=E’ bond dissociation energies. The data in Table 1.5 predicts that for electronegative
π-donating substituents, for which ΔEST of GeR2 is significantly larger than that of GeH2, the energy required to planarize the
trans-bent R2Ge=GeH2 increases significantly. Simultaneously, the Ge–Ge bond length elongates from 2.307 Å (R=H) to
2.483 Å (R=NH2), being longer than the single bond length in H3Ge–GeH3 of 2.40–2.46 Å [188] (exp. r(Ge–Ge) = 2.403 Å
[187]). The trans-bending angle at the Ge atom bearing the substituents (θR) is smaller than in H2Ge=GeH2, but for electro-
negative π-donating substituents, θH is very large, 78–94°. The long Ge=Ge bonds that are in the range of a Ge–Ge single bond
length, the large trans-bending angle θH of nearly 90°, and the significantly smaller BDE for R2Ge=GeH2, R=NH2, OH, F and

better described as R2Ge → GeH2 adducts, where the σ-lone pair of GeR2 donates electrons into the empty p orbital on GeH2,
Cl, relative to R=H, indicate a weak Ge–Ge bond. In our view, these geometric parameters indicate that these compounds are

(57, E=Ge) while the back-donation from GeH2 is small, in contrast to a classical donor-acceptor complex (Figure 1.8c),
leading to a trans-bent double bond. Similar geometry patterns are observed in a variety of NHC-stabilized divalent group 14
adducts [213–215], e.g., NHC→SiCl2 [215, 216], NHC→GeMes2 [216, 217], and NHC→GeI2 [218]. In these adducts the NHC
carbene is nearly perpendicular to the plane of the ER2 unit and the C–E bond is longer than a C–E single bond. In NHC→
GeI2, r(C–Ge) = 2.102 Å [218] vs. 1.969 Å [219] in H3C–GeH3 and 1.784 Å in H2Ge=CH2 [220]. NBO calculations [213] of the

resides mainly (91%) on the Si atom. The calculated extent of σ-donation in NHC→SiH2 is 0.557 electrons and the SiH2→NHC
NHC→SiH2 adduct suggests a Lewis structure having a C–Si polarized σ bond and a localized p(π) lone pair orbital which

back-donation is only 0.170 electrons (calculated using charge delocalization analysis (CDA) [221]). These are all indications
of a donor adduct rather than a doubly bonded compound [213].

E E

57
26 1 Computational and Theoretical Aspects of Structure and Bonding in Doubly Bonded Organogermanium Compounds

Table 1.5 Calculated singlet-triplet energy gaps (ΔEST) of R2Ge, relative energies (ΔE) of trans-bent
R2Ge=GeH2 (B) and planar (A) structures, the Ge–Ge bond dissociation energy (BDE) and geometric
parameters of R2Ge=GeH2, R=H, CH3, NH2, OH, F, Cl.a

R H CH3 NH2 OH F Cl

ΔEST 27.5 31.5 59.2 72.3 83.5 63.6


ΔEb −4.83 −6.1 −24.5 −22.6 −26.0 −20.2
c
BDE −52.2 −50.1 −35.9 −29.1 −23.4 −33.5
r(Ge-Ge) (planar, A) 2.224 2.230 2.234 2.232 2.225 2.228
r(Ge-Ge) (trans-bent, B) 2.307 2.337 2.483 2.475 2.467 2.440
θRd 43.45 30.14 12.79 17.61 28.1786.03 32.31
θHe 43.72 59.55 94.64 92.94 78.09
a
Calculated at B3LYP/6-311G(d). Energies in kcal/mol, bond length in Å, angles in degrees. From Ref. [212];
b
Negative values indicate that B is more stable than A. A is not a minimum on the PES;
c
BDE is the difference between the energy of R2Ge=GeH2 and the sum of the energies of GeH2 and GeR2 in
their singlet state. The values given in Table 1.5 were derived from the reported correlation of BDE vs. ΔEST: i.e.,
BDE = 0.515 ΔEST −66.37 (r2 = 0.956) [212]. The individual BDEs calculated directly were not provided;
d
The trans-bending angle at the Ge bonded to R;
e
The trans-bending angle at the Ge bonded to H.

1.4.2.2.2 Substituent Effects on Structures and Isomeric Forms of Ge2R2R’2 Compounds


1.4.2.2.2.1 Electronegative Substituents
1.4.2.2.2.1.1 Halide Substituents
Electronegative substituents significantly stabilize the singlet state of ER2, increasing ΔEST (e.g., Table 1.6, E=Ge). The
large increase in ΔEST from 16.5, 23.3, and 27.3 kcal/mol for EH2 (E=Si, Ge, Sn, respectively), to 69.7, 81.5, and 80.2 kcal/
mol, for EF2 (E=Si, Ge, Sn, respectively), leads to a significant stabilization of the dibridged isomer C [70, 176, 222] and it
is the most stable E2X4 isomer. The trans-bent isomer B of F2Ge=GeF2 is not a minimum on the potential energy surface
[175, 176, 222]. The dibridged Ge2F4 isomer, C, was identified and characterized in matrix by IR and Raman spectroscopy
[164].
Hargittai and coworkers studied computationally the PES of E2R4 (E=Si, Ge, Sn, Pb; R=F, Cl, Br, I) using the
MP2(FCI)/cc-pwCVTZ(-PP) level of theory [175]. The trans-dibridged isomer, C, is the global minimum for all com-
pounds, and it is only slightly more stable than the cis-dibridged isomer. The germylgermylene divalent isomer F is
higher in energy by ca. 30 kcal/mol for R=F, but it becomes relatively more stable for R=Cl, Br, and I (Table 1.6).
Structural and NBO analysis of the bonding in R3Ge–GeR for R=Br, Cl, and I, reveals that these are better described as
mono-bridged isomers E. For example, in R3Ge–GeR, R=Br, Cl, and I (see atom numbering in Table 1.6), r(Ge3–R4) of
2.443, 2.596, and 2.788 Å, respectively, having Wiberg bond indices (WBIs) of 0.43, 0.46, and 0.52, respectively, are
nearly identical to r(Ge1–R4) of 2.389. 2.525, and 2.711 Å and WBIs of 0.47, 0.52, and 0.58, respectively. Meanwhile,
Ge1–R6 and Ge3–R5 bonds (in F), are significantly shorter, e.g., for R=Br, they are 2.284 and 2.323 Å, respectively, and
for both the WBIs are 0.84 [175].
ΔEST for GeHR (R=F, Cl, Br, I) is significantly smaller than for GeR2 (Table 1.6), and according to the CGMT model the
relative stability order of the isomers on the PES of (GeHR)2 is expected to be different than that of the (GeR2)2 PES. The
relative energies calculated for the isomers of E2Me2Br2 shown in Figure 1.13 (the isomeric notations in this paragraph are
as in Figure 1.13) show that for E=Si, (E)-dibromodimetallene (A) is the most stable isomer. The corresponding silylsi-
lylene (E) has almost the same energy [186]. The bulkier substituted Bbt(Br)Si=Si(Br)Bbt (Bbt = 2,6-[CH(SiMe3)2]-4-
[C(SiMe3)3]-C6H2)) is more stable than its isomer E, by 14 kcal/mol [186]. Similarily, dihydrodisilene Bbt(H)Si=Si(H)Bbt
is more stable than the corresponding silylsilylene E by 13.2 kcal/mol, and the barrier for the isomerization of Bbt(H)
Si=Si(H)Bbt to Bbt(H2)Si-SiBbt is 19.6 kcal/mol (at B3PW91/6-311+G(2df)[Si]:6-31G(d)). Bbt(H2)Si-SiBbt was suggested as
an intermediate in the slow decomposition of Bbt(H)Si=Si(H)Bbt at 80° [186]. The dibridged isomers (C, D) are signifi-
cantly less stable for E=Si.
For E2Me2Br2, E=Ge, germylgermylene E is the most stable isomer. The doubly bonded isomer A is 5 kcal/mol higher in
energy. The dibridged isomers C and D are only slightly less stable than the doubly bonded isomer A. For E=Sn, both
dibridged isomers are significantly more stable than the doubly bonded isomer A (Figure 1.13) [186] as also predicted for
Table 1.6 Relative energies (kcal/mol) of isomers B-F on the Ge2R4 (R=H, F, Cl, Br, I) PES.

Isomer B C D E F

R θ R R R5 R5
R Ge Ge Ge1 Ge3
Ge Ge R
R R R4 R2
Ge Ge Ge1 Ge3
R R R R6
R R2
R R R6 R4 GeR2 ΔEST GeHR ΔEST

H 0.0 11.5a, 6.9b, 9.0c, 6.8d 13.6a, 8.6b, 11.6c 4.8a, 1.9b 3.4a, −0.24b, 2.4c; 2.8e 23.3f, 26.7g −
h i g i g
F - 0.0 1.8 , 1.7 - 26.6 , 29.4 85.0 46.2g
i i g,j
Cl - 0.0 1.4 - 10.4 64.1 42.0g,j
Br - 0.0 1.9i - 8.1i 57.2g,j 40.3g,j
i i k,l
I - 0.0 2.7 - 4.6 47.7 37.5k,l
a
At CCSD(T)/aug-cc-pVTZ//B3LYP/6-311G(d,p), from Ref. [189];
b
At B3LYP/6-311G(d,p), from Ref. [189];
c
At CI/DZP from Ref. [162];
d
At BP86/def2-TZVPP from Ref. [70];
e
At BP86/DZP++ from Ref. [188];
f
From Ref. [70], at BP86/def2-TZVPP;
g
At B3LYP/DZP++, from Ref. [223];
h
At BP86/DZP++, from Ref. [176];
i
At MP2(FCI)/cc-pwCVTZ(-PP) from Ref. [175];
j
For values at other computational levels see Refs. [44, 45, 224];
k
At B3LYP/6-311g(d,p), from Ref. [223].
28 1 Computational and Theoretical Aspects of Structure and Bonding in Doubly Bonded Organogermanium Compounds

Figure 1.13 Calculated relative energies (kcal/mol) of E2Me2Br2 isomers at B3PW91/6-311G(3d), [TZ(2d) for Sn, and 6-31G(d) for C
and H]. Energies of E-1,2-dibromo-1,2-dimethyldimetallene are set to zero). Negative relative energies indicate more stable species.
Reproduced with permission from Ref. [186] Copyright (2013) Chemical Society of Japan.

Sn2H2 (Table 1.3) [162]. In fact, all known donor-free dimeric hydridostannylenes, (RHSn)2, adopt the trans-dihydrogen-
bridged structure RSn(μ-H)2SnR (e.g., R = 2,6-Dip2-C6H3, Dip = 2,6-iPr2-C6H3) [203, 225] or stannystannylene structure,
RSnSnH2R (R = 2,6-Tip2-C6H-3,5-iPr2 (Tip = 2,4,6-iPr3-C6H2)) [203, 225], and not the alkene-like structure [226]. It is
interesting to note that for E2Me2Br2, the Ge compounds differ from those of their closest neigbours, Si and Sn, both qual-
itatively (in their most stable isomer) and quantitatively (the PES is flatter for Ge than for Si or Sn (excluding isomer E)).
The calculated dissociation energies (ΔEdiss) of Me(Br)E=E(Br)Me to two E(Br)Me fragments decrease as E is heavier.
ΔEdiss (in kcal/mol, at B3PW91/6-311G(3d), [TZ(2d) for tin, and 6-31G(d) for C and H]) are: 37.5 (E=Si) > 23.1 (E=Ge) > 14.6
(E=Sn)[186]. These calculated dissociation trends are in agreement with experimental results that in solution dibromodisi-
lene is stable, dibromodigermene is in equilibrium with its bromogermylene fragments [129], while for E=Sn only the
Br-bridged isomer is observed [186].
Bbt(Br)Ge=Ge(Br)Bbt (25, Table 1.2) was isolated as stable orange crystals at −78°C. [129, 186] Its Ge=Ge bond length is
2.509 Å, longer than other known Ge=Ge bond length of 2.21–2.44 Å [18], and even longer than the Ge–Ge single bond of
2.463 Å in (GePh2)6 [87]. The trans-bending angle at Ge is 44.6°, more pyramidal than in other digermenes which are in the
range of ca. 10–40° [18], but no twist of the double bond was observed [129]. The calculated (B3LYP/6-31+G(d)) r(Ge=Ge)
and trans-bending angle in H(Br)Ge=Ge(Br)H are 2.378 Å and 53.2°, respectively, and 2.414 Å and 48.1°, respectively, in the
more bulky substituted Mes(Br)Ge=Ge(Br)Mes [129]. These calculations indicate that the Ge=Ge bond in 25 is elongated due
to the steric repulsion between the bulkier Bbt substituents. However, the trans-bending depends on an intrinsic property of
the GeRBr germylene, e.g., its ΔEST (see above). In toluene solution, 25 is in equilibrium with its germylene fragments [129].
The synthesis and X-ray crystallography characterization of additional 1,2-dihalodigermenes, (E)–Ar(Cl)
Ge=Ge(Cl)Ar, Ar = 2,6-Mes2-C6H3 [227, 228] (Mes = 2,4,6-Me3-C6H2)) (58); Ar = 2,6-Tip2-C6H3 (Tip = 2,4,6-iPr3-
C6H2) (59) [227], were reported by Power’s group. Very recently, Matsuo’s group reported the synthesis of (E)–
Eind(R)Ge=Ge(R)Eind, R=Cl (24b) and Br (24a) (for the definition of Eind, see Table 1.2) [128]. The Ge=Ge bond
length in 58, 59, 24b, and 24a are 2.443, 2.363, 2.412, and 2.414 Å, respectively, and the trans-bending angles are
39.0, 36.8, 44.3, and 43.3°, respectively. Twisting about the double bond was not observed [128]. Based on UV-Vis
spectroscopy, it was concluded that in solution 24a dissociates to its germylene fragments, and 25 is in equilibrium
with the GeBrBbt fragments. The different behavior of 24a and 25 was explained by dispersion effects. The computed
dissociation energy (ΔG) of the Ge=Ge bond in 24a (at B3PW91-D3/6-311G(3d,f) (Ge, Br) and 6-311G(d) (H, C, Si),
including dispersion) is 7.8 kcal/mol, compared to 15.3 kcal/mol in 25. Compound 24a can be regenerated from the
germylenes by removal of the solvent [128].
ΔEST of Ge(SiH3)X, X=F, Cl, Br, are (at B3LYP/DZP++): 31.1, 28.6, and 27.7 kcal/mol, respectively, significantly smaller
than those of Ge(CH3)X, X=F, Cl, Br, of 48.0, 43.5, and 41.7 kcal/mol, respectively (at B3LYP/DZP++), as expected for SiH3
vs. CH3 substituents [223]. Based on the CGMT model, we predict that their doubly bonded dimers e.g., X(SiR3)Ge=Ge(SiR3)X,
1.4 Doubly Bonded Compounds 29

X=F, Cl, Br, may be less pyramidal at the Ge atoms and with shorter and stronger Ge=Ge bonds, compared to the large
trans-bending angles and long Ge=Ge bonds found in X(Ar)Ge=Ge(Ar)X (e.g., compounds 24a, 24b, 25, (Table 1.2) and 58,
59). Computational studies are required to assess this prediction.

1.4.2.2.2.1.2 Amino substituents


Calculations and experimental evidence have shown that diamino-tetrylenes (E(NR2)2) do not dimerize to form tetraamino-
E=E molecules, in line with their large ΔEST values, e.g., ΔEST = 79.3 kcal/mol for Si(NH2)2 [229]. Calculations predict that
(H2N)2Si=Si(NH2)2 is not a minimum on the PES, and that the dibridged H2NSi(μ-H2N)2SiNH2 should be observed [229,
230]. Following these calculations Kira and coworkers in an elegant experiment have confirmed the existence of the
dibridged iPr2NSi(μ-iPr2N)2SiNiPr2 [231].
A similar behavior is found for amino substituted germanium and tin compounds. Reaction of ArECl (E=Ge, Sn, Pb) in
liquid ammonia produced trans-dibridged ArE(μ-NH2)2EAr (E=Ge, Ar = Ar′ = 2,6-iPr2-C6H3 or Ar = Ar* = 2,6-(2,4,6-iPr3-
C6H2)2-C6H3; E=Sn, Ar=Ar*; E=Pb, Ar=Ar*) that were characterized spectroscopically [232].
E(N(SiMe3)2)2, E=Ge, Sn, are monomeric in the solid state and do not dimerize in solution [87]. Calculations at B3PW91
reveal that the Ge–Ge bond length in (Ge(N(SiMe3)2)2 is 5.841 Å, more than twice the van der Waals radius of a Ge atom

longer than a Ge–Ge single bond [117]. The calculated dissociation (ΔH) energy of ((Me3Si)2N)2Ge=Ge(N(SiMe3)2)2 → 2
(2.11 Å [233]) and it is contracted to 3.798 Å at B3PW91-D3 when attractive dispersion forces are included, still significantly

Ge(N(SiMe3)2)2 is −7.0 kcal/mol (i.e., exothermic) at B3PW91, but increases to 3.6 kcal/mol when dispersion forces are
included. However, ΔG for dissociation (at 25°C and 1 atm.) is −12.2 kcal/mol (i.e., exergonic), indicating that monomeric
species are favored under these conditions [117].
Additional experimental support for the conclusion that diaminosilylenes avoid dimerization to tetraamino-disilenes
was reported by West, Apeloig, and coworkers [234], who showed that the stable N-heterocyclic silylene 60, R=tBu, does
not dimerize to form disilene 61, E=E’=Si (Scheme 1.4). Instead, silylene 62, E=E’=Si, substituted by one amino group and
a silyl group is formed as an intermediate, and it dimerizes to Z-diaminodisilyldisilene, 63, E=E’=Si, R=tBu, which is sta-
bilized by two silyl groups. Calculations at B3LYP/6-31G(d) found that 63, E=E’=Si, R=Me is a minimum on the PES while
the tetraaminodisilene 61, E=E’=Si, R=Me, is not. 63, E=E’=Si, R=tBu, exhibits a very long Si=Si bond of 2.289 Å (2.285 Å
calc. for R=Me), an extremely large trans-bending angle of ca. 33° and a Si–Si–N twist angle of 25°. The formation of 63 and
not of 61 is consistent with the much larger ΔEST of (Me2N)2Si of 68.7 kcal/mol relative to that of H3Si(Me2N)Si of 45.2 kcal/
mol (B3LYP/6-311G(d,p) [234].

Scheme 1.4 Dimerization of N-Heterocyclic tetrylenes


30 1 Computational and Theoretical Aspects of Structure and Bonding in Doubly Bonded Organogermanium Compounds

Similarly, a reaction of silylene 60, E=Si, R=tBu, with germylene 60’, E’=Ge, R=tBu, did not produce the tetraamino-
substituted germasilene 61, E=Si, E’=Ge, but instead yielded digermene 63, E=Ge, E’=Si (Scheme 1.4) substituted by two
amino groups and two stabilizing silyl groups [235, 236]. In 63, E=Ge, E’=Si, R=tBu, r(Ge-Ge) is 2.45 Å [235, 236], longer
than in alkyl and aryl substituted digermenes. The trans-bending angle is ca. 41° and the twist angle is 22°. Reaction of
silylene 60 with Sn[N(SiMe3)2]2 yields, as a final product, the hydridodisilylstannane, 64 [235]. Calculations to study this
interesting reaction are welcomed.
(E)-(R2N)HGe=GeH(NR2) (22 [126], 23 [127], Table 1.2) were synthesized by hydrogenation of amido-digemynes 65 and
67, respectively (Scheme 1.5a and 1.5b respectively), and were characterized by X-ray crystalography [126, 127]. A UV-Vis
spectroscopic experiment led to the conclusion that digermene 22 is in equilibrium with germylene 66, which was trapped
by DMAP (4-dimethylaminopyridine) and the adduct was characterized by X-ray crystallography [126]. 68 activates H2
even at −10°C yielding amindogermylamidogermylene 69, both in the solid state and in solution. The structure of 69 was
determined by X-ray crystallography (Scheme 1.5c) [237]. Based on variable temperature NMR experiments at 3–70°C in
THF-d8 solution, it was concluded that in solution germylgermylene 69 is in equilibrium with digermene 70 [237]. The
reaction mechanisms of addition of H2 to R2NE-ENR2 (E=Si, Ge, Sn) were studied computationally by Frenking and
coworkers (for detailed PESs the reader is referred to the cited manuscripts) [225, 238]. The hydrogenation of 68 was cal-
culated (at BP86/def2-TZVPP) to be exergonic, forming either 69 (ΔG = −7.5 kcal/mol, −9.4 kcal/mol including dispersion
effects, and −8.9 kcal/mol when solvent effect is included) or 70 (ΔG = −5.6 kcal/mol, −6.6 kcal/mol including dispersion
effects, and –5.7 kcal/mol, when solvent effect is included). The rate-determining free energy barrier for hydrogenation is
20.4 kcal/mol, 18.4 kcal/mol including dispersion effects, and 19.0 kcal/mol with solvent effect [225]. 69 is more stable than
70 by only 2–3 kcal/mol which is in line with the observed equilibrium between these isomers in solution. r(Ge=Ge) is
2.544 Å (70, calculated) [225], 2.486 Å (22, exp., and 2.510 Å, calc.) [126], and 2.535 Å (23, exp.) [127]. These bonds are at
the longest edge of known Ge=Ge bonds, and are close to that of r(Ge-Ge) in 69 of 2.550 Å [237]. These digermenes are also
highly trans-bent, i.e., by 50° (calculated for 70) [237] (54.1° measured for 22) [126], and by a considerably smaller angle of
39.1° in 23 (exp.) [127]. Computational studies are required for analyzing the nature of the bonding in these molecules.

Scheme 1.5 Synthesis of (E )-(R2N)HGe=GeH(NR2) by hydrogenation of amido-digermylenes


1.4 Doubly Bonded Compounds 31

1.4.2.2.2.2 Electropositive Substituents


Electropositive substituents reduce ΔEST of ER2, lower the trans-bending angle of R2E=ER2, and decrease its bond dissoci-
ation energy to the ER2 fragments.

1.4.2.2.2.2.1 Silyl Substituents


The calculated (at UB3LYP/6‑31+G(d,p)//UB3LYP/6‑31+G(d,p) + ZPE) ΔEST for (H3Si)2Si is 8.4 kcal/mol, compared to
20 kcal/mol for SiH2, and it is smaller for bulkier silyl substituents, e.g., ΔEST((Me3Si)2Si) = 2.8 [239]. Thus, H3Si(H)
Si=SiH2 [211] and (H3Si)2Si=Si(SiH3)2 [240] are both calculated to be planar. In tetrasilyldisilene, r(Si=Si) = 2.148 Å.
Isolated disilenes, with bulky silyl substituents, e.g., R3Si(R’3Si)Si=Si(SiR’3)SiR3 (R3Si=SiMe2tBu, R’3Si=SiMeiPr2) (71)
[241] and (R3Si)2Si=Si(SiR3)2 (R3Si=SiMe2tBu) (72) [242] are both planar with r(Si=Si) = 2.196 Å and 2.202 Å, respectively,
and the twist angles about the double bonds are 0° and 8.9°, respectively. Increasing the bulk of the substituents to SiMetBu2
(73) retains the planar geometry, but the double bond is twisted by 54.5°, elongating r(Si=Si) to 2.259 Å [243]. The calcu-
lated structure of 73 (at UBP86-D3/TZVP + D3, including dispersion corrections) [244] is in very good agreement with its
X-ray structure. The significant twist of the double bond enables a facile thermal rotation around the double bond, over-
coming a barrier of only 7.5 kcal/mol to access a triplet biradical that was observed by EPR spectroscopy, and its identity
was supported by calculations [244]. This is the first spectroscopic observation of a triplet diradical resulting from thermal
rotation around an E=E double bond, including C=C bonds.
Similarly to the effect of silyl substitution on silylenes, the calculated singlet-triplet energy gap (ΔEST) of Ge(SiH3)2 is
13.8 kcal/mol (at B3LYP/DZP++) [223], significantly smaller than for GeH2 and Ge(CH3)2 for which ΔEST are 26.7 [223]
(27.5 [212]) and 31.0 [223] (31.5 [212]) kcal/mol, respectively, but larger than for Si(SiH3)2 at 8.4 kcal/mol. The smaller
ΔEST of Ge(SiH3)2 causes a significant decrease in the calculated trans-bending angle θ and in r(Ge=Ge); for example, in
R2Ge=GeR2, R=SiH3, θ = 27.9° and r(Ge=Ge) = 2.292 Å, compared to θ = 41.14° and r(Ge=Ge) = 2.340 Å for R=CH3, and
θ = 43.9° and r(Ge=Ge) = 2.309 Å for R=H (at B3LYP/6-311G(d) [245]; calculated values at other computational levels are
given in Table 1.3 and in Figure 1.10). The silyl effect on geometry is exhibited also in isolated digermenes, e.g., in (iPr2M
eSi)2Ge=Ge(SiMeiPr2)2, (74), (iPr3Si)2Ge=Ge(SiiPr3)2 (75), and (tBuMe2Si)2Ge=Ge(SiMe2tBu)2 (11) in which the trans-
bending angles are: θ = 5.9, 16.4, and 7.5°, and r(Ge=Ge) = 2.268, 2.298, and 2.270 Å, respectively (the twist angle in these
three compounds is 0° [119, 120]), compared to the much larger θ of 32° and longer r(Ge=Ge) of 2.347 Å in [(Me3Si)2CH]2
Ge=Ge[CH(SiMe3)2]2 (1) [86, 87], or in aryl substituted digermenes that have trans-bending angles in the range of 20–40°
[16, 18].
The most significant change in geometry is observed for (tBu2MeSi)2Ge=Ge(SiMetBu2)2, (12) [121b]; i.e., θ = 0°,
r(Ge=Ge) significantly elongated to 2.346 Å, and the twist angle is very large, 52.8°, similarly to that of disilene 73. The
extreme twist reduces the 4pπ–4pπ overlap, destabilizes the HOMO, and decreases the HOMO–LUMO gap (Figure 1.14),
as manifested in a redshift of the UV-Vis absorption to 618 nm [121b] compared to that of 11 (421 nm) and to other iso-
lated digermenes (412–475 nm). The partial breaking of the Ge–Ge π bond in 12 may indicate a contribution of a biradical
character to the bond and a facile rotation to obtain a triplet biradical, as observed in the analogous disilene 73, but a
triplet state rotated digermene analogue has not been yet reported. 12 produces stable cation- and anion-radicals by cyclic
voltammetry [121b].

Figure 1.14 Optimized structure and HOMOs (B3LYP/6-31G(d)) of a set of silyl- substituted digermenes. Reprinted with permission
from Ref. [121b]. Copyright (2014) John Wiley & Sons Inc.
Another random document with
no related content on Scribd:
500 6 5 0
First account total 500 6 5 0
—— ——————
Grand total 1000 £12 10 0
==== ============
Audited and found correct,
(Signed) Thomas W.
Cathcart.
And what do you think she did? “Well, ma tear,” she said, “I can’t let
you go away without something left, in case you met a poor beggar
in the street. You must take back one of those little packets to go on
with, as a present from me;” and she picked up one and placed it in
Alison’s hand, and Alison took it gladly.
And that was the beginning of a new Threepenny Trust, for Mr.
Cathcart also contributed a little heap, and Mr. Muirhead
henceforward made a point of saving every threepenny bit that he
received in change (and I believe that sometimes he asked specially
for them when he went to his bank) and bringing them home for
Alison’s fund; and Uncle Mordaunt must have done the same, for the
last time he came to dinner he said to Alison, “I wish you’d get rid of
this rubbish for me,” and handed her seventeen of the little coins.
So you see that there is every chance of Mr. James Thomson’s kind
scheme going on for a long time yet; but, in so far as his own
thousand threepenny bits are concerned, the story is done.
RODERICK’S PROS.
RODERICK’S PROS.
Once upon a time there was a little boy of ten, who bowled out C. B.
Fry. This little boy’s name was Roderick Bulstrode (or Bulstrode is
the name that we will give him here), and he lived in St. John’s
Wood, in one of the houses whose gardens join Lord’s. His father
played for the M.C.C. a good deal, and practised in the nets almost
every day, to the bowling of various professionals, or pros., as they
are called for short, but chiefly to that of Tom Stick; and in the
summer Roderick was more often at Lord’s than not.
How it came about that Roderick bowled C. B. Fry was this way.
Middlesex were playing Sussex, and Mr. Fry went to the nets early to
practise, and Roderick’s father bowled to him and let Roderick have
the ball now and then. And whether it was that Mr. Fry was not
thinking, or was looking another way, or was simply very good-
natured, I don’t know, but one of Roderick’s sneaks got under his bat
and hit the stumps. (They were not sneaks, you must understand,
because he wanted to bowl sneaks, but because he was not big
enough to bowl any other way for 22 yards. He was only ten.)
Roderick thus did that day what no one else could do, for Mr. Fry
went in and made 143 not out, in spite of all the efforts of Albert Trott
and Tarrant and J. T. Hearne.
Roderick’s bedroom walls had been covered with portraits of
cricketers for years, but after he bowled out C. B. Fry he took away a
lot of them and made an open space with the last picture postcard of
Mr. Fry right in the middle of it, and underneath, on the mantelpiece,
he put the ball he had bowled him with, which his father gave him,
under a glass shade. And other little St. John’s Wood boys, friends
of Roderick’s from the Abbey Road, and Hamilton Terrace, and
Loudoun Road, and that very attractive red-brick village with a green
of its own just off the Avenue Road, used to come and see it, and
stand in front of it and hold their breath, rather like little girls looking
at a new baby.
Roderick also had a “Cricketers’ Birthday Book,” so that when he
came down to breakfast he used to say, “Tyldesley’s thirty-five to-
day,” “Hutchings is twenty-four,” and so on. And he knew the initials
of every first-class amateur and the Christian name of every pro.
That was not Roderick’s only cricketing triumph. It is true that he had
never succeeded in bowling out any other really swell batsman, but
he had shaken hands with Sammy Woods and J. R. Mason, and one
day Lord Hawke took him by both shoulders and lifted him to one
side, saying: “Now then, Tommy, out of the way.” But these were only
chance acquaintances. His real cricketing friend was Tom Stick, the
ground bowler.
Tom Stick came from Devonshire, which is a county without a first-
class eleven that plays the M.C.C. in August, and he lived in a little
street off Lisson Grove, where he kept a bird-fancier’s shop. For
most professional cricketers, you know, are something else as well,
or they would not be able to live in the winter. Many of them make
cricket-bats, many keep inns, many are gardeners. I know one who
is a picture-framer, and another an organist, while George Hirst, who
is the greatest of them all, makes toffee. Well, Tom Stick was a bird-
fancier, with a partner named Dick Crawley, who used to mind the
shop when Tom had to be at Lord’s bowling to gentlemen, Roderick’s
father among them, or playing against Haileybury or Rugby or
wherever he was sent to do all the hard work and go in last.
Roderick’s father was very fond of Tom and was quite happy to know
that Roderick was with him, so that Roderick not only used to join
Tom at Lord’s, but also at the shop off Lisson Grove, where he often
helped in cleaning out the cages and feeding the birds and teaching
the bullfinches to whistle, and was very good friends also with certain
puppies and rabbits. His own dog, a fox-terrier named “Sinhji,” had
come from Tom.
Tom used to bowl to Roderick in the mornings before the gentlemen
arrived for their practice, and he taught him to hold his bat straight
and not slope it, and to keep his feet still and not draw them away
when the ball was coming (which are the two most important things
in batting), and it was he who stopped Roderick from carrying an
autograph-book about and worrying the cricketers for their
signatures. In fact, Tom was a kind of nurse to Roderick, and they
were so much together that, whereas Tom was known to Roderick’s
small friends as “Roddy’s Pro,” Roderick was known to Tom’s friends
as “Sticky’s Shadow.”
Now it happened that last summer Roderick’s father had been
making a great many runs for the M.C.C. in one of their tours.
(Roderick did not see him, for he had to stay at home and do his
lessons; but his father sent him a telegram after each innings.) Mr.
Bulstrode (as we are calling him) batted so well, indeed, that when
he returned to London he was asked to play for Middlesex against
Yorkshire on the following Monday, to take the place of one of the
regular eleven who was ill; and you may be sure he said yes, for,
although he was now thirty-two, this was the first time he had ever
been asked to play for his county.
Roderick, you may be equally sure, was also pleased; and when his
father suddenly said to him, “Would you like to come with me?” his
excitement was almost too great to bear.
“And Tom too?” he asked, after a minute or so.
“Yes, Tom’s going,” said his father. “He’s going to field if anyone is
hurt or has to leave early. But if he’s not wanted he will look after
you.”
“Hurray!” said Roderick. “I know what I shall do. I shall score every
run and keep the bowling analysis too.”
The train left St. Pancras on the Sunday afternoon, and that in itself
was an excitement, for Roderick had never travelled on Sunday
before; but before that had come the rapture of packing his bag,
which on this occasion was not an ordinary one, but an old cricket-
bag of his father’s, which he begged for, in which were not only his
sponge and collars and other necessary things, but his flannels and
his bat and pads.
This bag he insisted upon carrying himself all along the platform,
and, as several of the Middlesex team were also on their way to the
train at the same moment, the presence of so small a cricketer in
their midst made a great sensation among the porters.
“My word!” said one, “Yorkshire will have to look out this time.”
“Who’s the giant,” asked another, “walking just behind Albert Trott? I
shouldn’t like to be in when he bowled his fastest.”
But Roderick was unconscious of any laughter. He was the proudest
boy in London, although his arm, it is true, was beginning to ache
horribly. But when, as he was climbing into the carriage, the guard
lifted him up and called him “Prince Run-get-simply,” he joined in the
fun.
THE PRESENCE OF SO SMALL A CRICKETER MADE A
GREAT SENSATION AMONG THE PORTERS.
It was a deliriously happy journey, for all the cricketers were very
nice to him, and Mr. Warner talked about Australia, and Mr.
Bosanquet showed him how he held the ball to make it break from
the leg when the batsman thought it was going to break from the off,
and at Nottingham Mr. Douglas bought him a bun and a banana.
They got to Sheffield just before eight, and Roderick went to bed
very soon after, in a little bed in his father’s room in the hotel.
The first thing Roderick did the next morning was to buy a scoring-
book and a pencil, and then he and his father explored Sheffield a
little before it was time to go to the ground at Bramall Lane and get
some practice.
The people clustered all round and in front of the nets and watched
the batsmen, and now and then they were nearly killed, as always
happens before a match. They pointed out the cricketers to each
other.
“There’s Warner,” they said. “That’s Bosanquet—the tall one.”
“Where’s Trott? Why, there, bowling at Warner. Good old Alberto!”
and so on.
“Who’s the man in the end net?” Roderick heard some one ask.
“I don’t know. One of Middlesex’s many new men, I suppose,” said
the other.
“But he can hit a bit, can’t he?” the first man said, as Roderick’s
father stepped out to a ball and banged it half-way across the
ground.
Roderick was very proud, and he felt that the time had come to make
his father known. “That’s Bulstrode,” he said.
“Oh, that’s Bulstrode, is it?” said the second man. “I’ve heard of him.
He makes lots of runs on the M.C.C. tours. But I guess Georgy’ll get
him.”
“Who is Georgy?” asked Roderick.
“Georgy—why, where do you come from? Fancy being in Sheffield
and asking who Georgy is. Georgy is Georgy Hirst, of course.”
Roderick walked back to the pavilion with his father very proudly.
“You’ll have to be very careful how you play Hirst,” he said.
“I shall,” said his father; “but why?”
“Because the men were saying he’s going to get you.” Mr. Bulstrode
laughed; but he thought it very likely too.
I’m not going to tell you all about the match, for it lasted three days,
and was very much like other matches. Roderick had a corner seat
in the pavilion, where he could see everything, and for the first day
he scored every run and kept the analysis right through. This
included his father’s innings, which lasted, alas! far too short a time,
for, after making four good hits to the boundary, he was caught close
in at what was called silly mid-on off—what bowler do you think?—
George Hirst.
But the next day Roderick gave up work, because he wanted to see
more of Tom, and Tom made room for him in the professionals’ box
while Yorkshire were in, and he saw all the wonderful men—quite
close too—Tunnicliffe and Denton and Hirst—and even talked with
them. Hirst sat right in front of the box, with his brown sunburned
arms on the ledge, and his square, jolly, sunburned face on his arms,
and said funny things about the play in broad Yorkshire; and now
and then he would say something to Roderick. And then suddenly
down went a wicket, and Hirst got up to go in.
“Give me a wish for luck,” he said to Roddy.
“I wish my father may catch you out,” said Roddy; “but not until,” he
added, “you have made a lot of runs.”
“If he does,” said Hirst, “I’ll give thee some practice to-morrow
morning.”
Poor Roddy, this was almost too much. It is bad enough to watch
your favourites batting at any time, for every ball may be the last; but
it is terrible when you equally want two people to bring something off
—for Roddy wanted Hirst (whom he now adored) to make a good
innings, and, at the same time, he wanted his father to catch Hirst
out.
Hirst was not out when it was time for lunch, and so Roderick was
able to tell his father all about it.
“What’s this, Hirst?” said Mr. Bulstrode, when the teams were being
photographed. “Give me a chance, and let me see if I can hold it.”
Hirst laughed, and when he laughs it is like a sunset in fine weather.
“I have a spy round to see where thee’re standing every over,” he
said, “and that’s where I’ll never knock it.”
“But what about my boy’s practice?” Mr. Bulstrode replied.
“Ah, we’ll see about that,” said the Yorkshireman.
But, as a matter of fact, Roderick got his practice according to the
bargain, for, as it happened, it was Mr. Bulstrode who caught Hirst, at
third man.
I need hardly tell you that Roderick dreamed that night. His sleep
was full of Hirsts, all jolly and all hitting catches which his father
buttered. But in the morning, when he knew how true his luck was,
he was almost too happy. Hirst was as good as his word, and they
practised in the nets together for nearly half an hour, and Roderick
nearly bowled him twice.
In Middlesex’s next innings Roderick’s father made thirty-five, all of
which Roderick scored with the greatest care; but the match could
not be finished owing to a very heavy shower, and so this innings did
not matter very much one way or the other, except that it made Mr.
Bulstrode’s place safe for another match.
Of that match I am not going to tell; but I have perhaps said enough
to show you how exceedingly delightful it must be to have a father
who plays for his county.
THE MONKEY’S REVENGE
THE MONKEY’S REVENGE
Once upon a time there was a little girl named Clara Amabel Platts.
She lived in Kensington, near the Gardens, and every day when it
was fine she walked with Miss Hobbs round the Round Pond. Miss
Hobbs was her governess. When it was wet she read a book, or as
much of a book as she could, being still rather weak in the matter of
long words. When she did not read she made wool-work articles for
her aunts, and now and then something for her mother’s birthday
present or Christmas present, which was supposed to be a secret,
but which her mother, however hard she tried not to look, always
knew all about. But this did not prevent her mother, who was a very
nice lady, from being extraordinarily surprised when the present was
given to her. (That word “extraordinarily,” by the way, is one of the
words which Clara would have had to pass over if she were reading
this story to herself; but you, of course, are cleverer.)
It was generally admitted by Mrs. Platts, and also by Miss Hobbs and
Kate Woodley the nurse, that Clara was a very good girl; but she had
one fault which troubled them all, and that was too much readiness
in saying what came into her mind. Mrs. Platts tried to check her by
making her count five before she made any comment on what was
happening, so that she could be sure that she really ought to say it;
and Kate Woodley used often to click her tongue when Clara was
rattling on; but Miss Hobbs had another and more serious remedy.
She used to tell Clara to ask herself three questions before she
made any of her quick little remarks. These were the questions: (1)
“Is it kind?” (2) “Is it true?” (3) “Is it necessary?” If the answer to all
three was “Yes,” then Clara might say what she wanted to; otherwise
not. The result was that when Clara and Miss Hobbs walked round
the Round Pond Clara had very little to say; because, you know, if it
comes to that, hardly anything is necessary.
Well, on December 20, 1907, the postman brought Mrs. Platts a
letter from Clara’s aunt, Miss Amabel Patterson of Chislehurst, after
whom she had been named, and it was that letter which makes this
story. It began by saying that Miss Patterson would very much like
Clara to have a nice Christmas present, and it went on to say that if
she had been very good lately, and continued good up to the time of
buying the present, it was to cost seven-and-six, but if she had not
been very good it was only to cost a shilling. This shows you the kind
of aunt Miss Patterson was. For myself, I don’t think that at
Christmas-time a matter of good or bad behaviour ought to be
remembered at all. And I think that everything then ought to cost
seven-and-six. But Miss Patterson had her own way of doing things;
and it did not really matter about the shilling at all, because, as it was
agreed that Clara had been very good for a long time, Mrs. Platts
(who did not admire Miss Patterson’s methods any more than we do)
naturally decided that unless anything still were to happen (which is
very unlikely with six-and-sixpence at stake) the present should cost
seven-and-six, just as if nothing about a shilling had ever been said.
Unless anything were to happen. Ah! Everything in this story
depends on that.
Clara was as good as gold all the morning, and she and Miss Hobbs
marched round the Round Pond like soldiers, Miss Hobbs talking all
the time and Clara as dumb as a fish. At dinner also she behaved
beautifully, although the pudding was not at all what she liked; and
then it was time for her mother to take her out to buy the present. So,
still good, Clara ran upstairs to be dressed.
As I dare say you know, there are in Kensington High Street a great
many large shops, and the largest of these, which is called Biter’s,
has a very nice way every December of filling one of its windows
(which for the rest of the year is full of dull things, such as tables,
and rolls of carpets, and coal scuttles) with such seasonable and
desirable articles as boats for the Round Pond, and dolls of all sorts
and sizes, and steam engines with quite a lot of rails and signals,
and clockwork animals, and guns. And when you go inside you can’t
help hearing the gramaphone.
It was into this shop that Mrs. Platts and Clara went, wondering
whether they would buy just one thing that cost seven-and-six all at
once, or a lot of smaller things that came to seven-and-six
altogether; which is one of the pleasantest problems to ponder over
that this life holds. Well, everything was going splendidly, and Clara,
after many changings of her mind, had just decided on a beautiful
wax doll with cheeks like tulips and real black hair, when she
chanced to look up and saw a funny little old gentleman come in at
the door; and all in a flash she forgot her good resolutions and
everything that was depending on them, and seizing her mother’s
arm, and giving no thought at all to Miss Hobbs’s three questions, or
to Kate Woodley’s clicking tongue, or to counting five, she cried in a
loud quick whisper, “Oh, mother, do look at that queer little man! Isn’t
he just like a monkey?”
“DO LOOK AT THAT QUEER LITTLE MAN!”
Now there were two dreadful things about this speech. One was that
it was made before Aunt Amabel’s present had been bought, and
therefore Mrs. Platts was only entitled to spend a shilling, and the
other was that the little old gentleman quite clearly heard it, for his
face flushed and he looked exceedingly uncomfortable. Indeed, it
was an uncomfortable time for every one, for Mrs. Platts was very
unhappy to think that her little girl not only should have lost the nice
doll, but also have been so rude; the little old gentleman was
confused and nervous; the girl who was waiting on them was
distressed when she knew what Clara’s unlucky speech had cost
her; and Clara herself was in a passion of tears. After some time, in
which Mrs. Platts and the girl did their best to soothe her, Clara
consented to receive a shilling box of chalks as her present, and was
led back still sobbing. Never was there such a sad ending to an
exciting expedition.
Miss Hobbs luckily had gone home; but Kate Woodley made things
worse by being very sorry and clicking away like a Bee clock, and
Clara hardly knew how to get through the rest of the day.
Clara’s bedtime came always at a quarter to eight, and between her
supper, which was at half-past six, and that hour she used to come
downstairs and play with her father and mother. On this evening she
was very quiet and miserable, although Mrs. Platts and Mr. Platts did
all they could to cheer her; and she even committed one of the most
extraordinary actions of her life, for she said, when it was still only
half-past seven, that she should like to go to bed.
And she would have gone had not at that very moment a
tremendous knock sounded at the front door—so tremendous that, in
spite of her unhappiness, Clara had, of course, to wait and see what
it was.
And what do you think it was? It was a box addressed to Mrs. Platts,
and it came from Biter’s, the very shop where the tragedy had
occurred.
“But I haven’t ordered anything,” said Mrs. Platts.
“Never mind,” said Mr. Platts, who had a practical mind. “Open it.”
So the box was opened, and inside was a note, and this is what it
said:
“Dear Madam,
“I am so distressed to think that I am the cause of
your little girl losing her present, that I feel there is nothing
I can do but give her one myself. For if I had not been so
foolish—at my age too!—as to go to Biter’s this afternoon,
without any real purpose but to look round, she would
never have got into trouble. Biter’s is for children, not for
old men with queer faces. And so I beg leave to send her
this doll, which I hope is the right one, and with it a few
clothes and necessaries, and I am sure that she will not
forget how it was that she very nearly lost it altogether.
“Believe me, yours penitently,
“The Little-old-gentleman-who-really-is-(as-his-
looking-glass-has-too-often-told-him)-like-a-
monkey.”
To Clara this letter, when Mrs. Platts read it to her, seemed like
something in a dream, but when the box was unpacked it was found
to contain, truly enough, not only the identical doll which she had
wanted, with cheeks like tulips and real black hair, but also frocks for
it, and night-dresses and petticoats, and a card of tortoiseshell toilet
requisites, and three hats, and a diabolo set, and a tiny doll’s parasol
for Kensington Gardens on sunny days.
Poor Clara didn’t know what to do, and so she simply sat down with
the doll in her arms and cried again; but this was a totally different
kind of crying from that which had gone before. And when Kate
Woodley came to take her to bed she cried too.
And the funny thing is that, though the little old gentleman’s present
looks much more like a reward for being naughty than a punishment,
Clara has hardly ever since said a quick unkind thing that she could
be sorry for, and Miss Hobbs’s three questions are never wanted at
all, and Kate Woodley has entirely given up clicking.
THE NOTICE-BOARD
THE NOTICE-BOARD
Once upon a time there was a family called Morgan—Mr. Morgan
the father, Mrs. Morgan the mother, Christopher Morgan, aged
twelve, Claire Morgan, aged nine, Betty Morgan, aged seven, a fox-
terrier, a cat, a bullfinch, a nurse, a cook, a parlourmaid, a
housemaid, and a boy named William. William hardly counts,
because he came only for a few hours every day, and then lived
almost wholly in the basement, and when he did appear above-stairs
it was always in the company of a coal-scuttle. That was the family;
and at the time this story begins it had just removed from
Bloomsbury to Bayswater.
While the actual moving was going on Christopher Morgan, Claire
Morgan, and Betty Morgan, with the dog and the bullfinch, had gone
to Sandgate to stay with their grandmother, who, with extraordinary
good sense, lived in a house with a garden that ran actually to the
beach, so that, although in stormy weather the lawn was covered
with pebbles, in fine summer weather you could run from your
bedroom into the sea in nothing but a bath-towel or a dressing-gown,
or one of those bath-towels which are dressing-gowns. Christopher
used to do this, and Claire would have joined him but that the doctor
forbade it on account of what he called her defective circulation—two
long words which mean cold feet.
When, however, the moving was all done and the new house quite
ready, the three children and the dog and the bullfinch returned to
London, and getting by great good luck a taxicab at Charing Cross,
were whirled to No. 23, Westerham Gardens almost in a minute, at a
cost of two-and-eightpence, with fourpence supplement for the
luggage. Christopher sat on the front seat, watching the meter all the
time, and calling out whenever it had swallowed another twopence.
The first eightpence, as you have probably also noticed, goes slowly,
but after that the twopences disappear just like sweets.
It is, as you know, a very exciting thing to move to a new house.
Everything seems so much better than in the last, especially the
cupboards and the wall-papers. In place of the old bell-pulls you find
electric bells, and there is a speaking-tube between the dining-room
and the kitchen, and the coal-cellar is much larger, and the bath-
room has a better arrangement of taps, and you can get hot water on
the stairs. But, of course, the electric light is the most exciting thing
of all, and it was so at Westerham Gardens, because in Bloomsbury
there had been gas. But Mr. Morgan was exceedingly serious about
it, and delivered a lecture on the importance—the vital importance—
of always turning off the switch as you leave the room, unless, of
course, there is some one in it.
Christopher and Claire and Betty were riotously happy in their new
home for some few days, especially as they were so near
Kensington Gardens, only a very little way, in fact, from the gate
where the Dogs’ Cemetery is.
And then suddenly they began to miss something. What it was they
had no idea; but they knew that in some mysterious way, nice as the
new house was, in one respect it was not so nice as the old one.
Something was lacking.
It was quite by chance that they discovered what it was; for, being
sent one morning to Whiteley’s, on their return they entered
Westerham Gardens by a new way, and there on a board fixed to the
railings of the corner house they read the terrible words:

ORGANS AND STREET CRIES

PROHIBITED.

Then they all knew in an instant what it was that had vaguely been
troubling them in their new house. It was a house without music—a
house that stood in a neighbourhood where there were no bands, no
organs, and no costermongers.

You might also like