Nagai (Principal)

Download as pdf or txt
Download as pdf or txt
You are on page 1of 228
At a glance
Powered by AI
Covalent organic frameworks (COFs) are crystalline porous polymers constructed from light organic building blocks via covalent bond linkages. This passage discusses various aspects of COFs including their design, synthesis, structures, properties and applications.

Some common linkages used in COFs include boroxine, boronic ester, spiroborate, imine, hydrazone, azine, squaraine, imide, phenazine, triazine and multihetero linkages.

Common synthesis methods for COFs discussed in the passage include solvothermal synthesis, ionothermal synthesis, microwave synthesis, mechanochemical synthesis and room-temperature synthesis.

Covalent Organic

Frameworks
Covalent Organic
Frameworks

edited by
Atsushi Nagai
Published by
Jenny Stanford Publishing Pte. Ltd.
Level 34, Centennial Tower
3 Temasek Avenue
Singapore 039190

Email: [email protected]
Web: www.jennystanford.com

British Library Cataloguing-in-Publication Data


A catalogue record for this book is available from the British Library.

Covalent Organic Frameworks


Copyright © 2020 by Jenny Stanford Publishing Pte. Ltd.
All rights reserved. This book, or parts thereof, may not be reproduced in any form
or by any means, electronic or mechanical, including photocopying, recording
or any information storage and retrieval system now known or to be invented,
without written permission from the publisher.

For photocopying of material in this volume, please pay a copying fee through
the Copyright Clearance Center, Inc., 222 Rosewood Drive, Danvers, MA 01923,
USA. In this case permission to photocopy is not required from the publisher.

ISBN 978-981-4800-87-7 (Hardcover)
ISBN 978-1-003-00469-1 (eBook)
Contents

Preface ix

1. Design and Synthesis: Covalent Organic Frameworks 1


1.1 Introduction 1
1.2 Design and Synthesis 2
1.2.1 COF Synthesis in the Dynamic Covalent
Chemistry Concept 2
1.2.2 Dynamic Linkages of Building Blocks 4
1.2.3 Topology and Geometry of 2D Porous
Materials Containing COFs 9
1.3 Synthetic Methods of COFs 12
1.3.1 Solvothermal Synthesis 12
1.3.2 Ionothermal Synthesis 13
1.3.3 Microwave Synthesis 13
1.3.4 Mechanochemical Synthesis 13
1.3.5 Room-Temperature Synthesis 14

2. Crystallization and Structural Linkages of COFs 21


2.1 b–O Linkages 21
2.1.1 Boroxine-Linked COFs 21
2.1.2 Boronic Ester (Dioxaborole)-Linked
COFs 24
2.1.3 Spiroborate-Linked COFs 33
2.1.4 Borazine-Linked COFs 36
2.2 Imine Linkages 38
2.3 Hydrazone Linkages 49
2.4 Azine Linkages 52
2.5 Squaraine Linkages 54
2.6 Imide Linkages 54
2.7 Phenazine Linkages 55
2.8 Triazine Linkages 57
2.9 Multihetero Linkages in One COF Skeleton 59
2.10 Perspectives and Challenges 61
vi Contents

3. Gas Adsorption and Storage of COFs 69


3.1 Gas Sorption 69
3.2 Physical and Chemical Adsorption 70
3.3 Brunauer–Emmett–Teller Theory 73
3.4 Hydrogen Gas Storage 74
3.5 Methane Gas Storage 76
3.6 Carbon Dioxide Gas Storage 78
3.7 Membrane Separation of COFs 83
3.7.1 Key Properties of COFs for Membrane
Separation 84
3.7.2 Fabrication of COF-Based Membranes 85
3.7.2.1 Design principles 85
3.7.2.2 Blending 85
3.7.2.3 In situ growth 90
3.7.2.4 Layer-by-layer stacking 92
3.7.2.5 Interfacial polymerization 94
3.7.3 Gas Separation of COF-Based
Membranes 95
3.8 Outlook and Conclusions 99

4. Heterogeneous Catalytic Application of COFs 103


4.1 Heterogeneous Catalysts of COFs for C–C
Bond Coupling Reactions 103
4.1.1 Suzuki–Miyaura Reaction 106
4.1.2 Heck, Sonogashira, and Silane-Based
Cross-Coupling Reactions 112
4.2 Chiral Heterogeneous Catalysts of COFs for
Asymmetric C–C Bond Coupling Reactions 116
4.3 Heterogeneous Bimetallic or Bifunctional
Catalysts of COFs 125
4.4 Heterogeneous Photo- and Electrocatalysts
of COFs 129
4.5 Heterogeneous Catalysts of 3D COFs 132
4.6 Conclusions and Outlook 135

5. Energy Storage Applications of 2D COFs 139


5.1 2D COFs for Optoelectronics and Energy
Storage 139
5.2 Semiconducting and Photoconducting
2D COFs 140
Contents vii

5.3 P-Type Semiconducting 2D COFs 141


5.4 N-Type Semiconducting 2D COFs 153
5.5 Ambipolar Semiconducting 2D COFs 155
5.6 Lithium-Ion Batteries Using 2D COFs as
Electrodes 166
5.6.1 Battery Cathode Application 167
5.6.2 Battery Anode Application 172
5.7 Summary and Perspective 177

6. Biomedical Applications of COFs 183


6.1 Introduction of Biomedical Application 183
6.2 COF Properties of Biomedical Applications 184
6.3 Biomedical COF Applications 187
6.3.1 Drug Delivery 187
6.3.2 Photothermal and Photodynamic
Therapy 195
6.4 Biosensing and Bioimaging 196
6.5 Other Biomedical Applications 198
6.6 Conclusions of Biomedical Applications 200

Index 209
Preface

Rational synthesis of extended arrays of organic matter in bulk, in


solution, in crystals, and in thin films has always been a paramount
goal of chemistry. The classical synthetic tools to obtain long-range
regularity are, however, limited to noncovalent interactions, and not
covalent polymerization reactions, which usually yield structurally
more random products. The most challenging hurdle in the synthesis
of extended yet precisely defined 2D and 3D structures based on
covalent chemistry is the requirement that the reaction linking
individual organic constitutes be reversible, allowing the scaffold to
arrange into a thermodynamic, well-ordered product rather than a
kinetic, amorphous structure. Although the synthesis of crystalline
inorganic materials largely relies on dynamic polymerization
processes, only a few examples that use dynamic polymerizations
for the generation of crystalline organic frameworks are known.
Hence, a combination of porosity and regularity in organic
covalently bonded materials requires not only the design of
molecular building blocks that allow for growth into a nonperturbed,
regular geometry, that is, two or three latent bonding–forming
sites spanning an appropriate angle, but also a condensation
mechanism that progresses under reversible, thermodynamic, self-
optimizing conditions. Ideally, a single precursor molecule fulfills
all these criteria, making even (if any) known from step-growth
polymerization reactions such as polycondensation and polyaddition.
In 2005, crystalline covalent organic frameworks (COFs) as 2D
polymers were for the first time reported by Yaghi et al. 2D COFs
resemble an sp2-carbon-based graphene sheet, but their structures
have a different molecular skeleton formed by orderly linkage of
building blocks to constitute a flat organic sheet. Nanocarbons are
also intriguing molecules and motifs with a discrete size and 2D
conformation but beyond the scope of this book. In the past decade,
COFs have emerged as a new class of highly ordered crystalline
organic porous polymers. They have attracted tremendous research
interest because of their unique structures and potentially wide
x Preface

applications in gas storage and separation, energy storage, catalysis,


and optoelectronic materials development.
This book, Covalent Organic Frameworks, with a selection of
subjects, is a handbook of COF research from design to application,
providing basics of and current trends in organic porous materials
for successful development and application and describing the
concepts of COF design and synthesis; COF crystallization and
structural linkages; theory of gas sorption; and various applications
of COFs, such as heterogeneous catalysts, energy storage (such as
semiconductors and batteries), and biomedical approaches.
Atsushi Nagai
Delft University of Technology (TU Delft), Netherlands
2020
Chapter 1

Design and Synthesis: Covalent Organic


Frameworks

Covalent organic frameworks (COFs) are newly emerged crystalline


porous polymers with well-defined skeletons and nanopores mainly
consisting of lightweight elements (H, B, C, N, and O) linked by
dynamic covalent bonds. Compared with conventional materials,
COFs possess some unique and attractive features, such as a large
surface area, a predesignable pore geometry, excellent crystallinity,
inherent adaptability, and high flexibility. Their large surface area,
tunable porosity, and p conjugation, with unique photoelectronic
properties, will enable COFs to serve as a promising platform for
a wide variety of applications. This chapter traces the evolution
of COFs and highlights the important issues related to synthetic
method and structural design.

1.1 Introduction
Recent decades have seen an explosion of interest in organic
materials with permanent nanometer-scale pores that grow very
quickly due to their specific properties and have broad applications
in gas storage, gas separation, drug delivery, energy conversion,
catalysis, and optoelectronics. Until now, a large number of typical
porous materials have been designed and constructed, such as

Covalent Organic Frameworks


Atsushi Nagai
Copyright © 2020 Jenny Stanford Publishing Pte. Ltd.
ISBN 978-981-4800-87-7 (Hardcover), 978-1-003-00469-1 (eBook)
www.jennystanford.com
2 Design and Synthesis

zeolites, metal-organic frameworks (MOFs), mesoporous silica,


organosilica, microporous polymers, porous carbons, covalent
organic frameworks (COFs), and organic molecular cages. Among
them, the covalent chemistry of organic and inorganic molecules
has been the focus of chemists, leading to many important advances
in science. The building and modification of organic molecules by
covalent bonds to make pharmaceuticals, chemicals, and polymers
have fundamentally changed our way of life. Likewise, covalent
synthesis through the construction of inorganic complexes has
led to useful catalysts capable of high activity and selectivity. The
precision and versatility with which covalent chemistry on such
molecules is practiced have not been translated to either the
building units of extended structures or their modification. This
is the next-generating idea of MOFs [1] and COFs [2–4], which are
two popular classes of porous materials on account of their unique
properties and promising application potential, attracting a lot of
research attention. Constructing novel MOFs and COFs is necessary
to expand the family of porous materials. This book focuses on COFs.
Generally, a 2D COF that is a covalent 2D polymer would resemble
a sp2-carbon-based graphene sheet but its structure would have a
different molecular skeleton, formed by orderly linkage of building
blocks to constitute a fat organic sheet. Two-dimensional COFs are
not single-atom-thick sheets but consist of 2D sheets layered as a
p-p stack. Because the pore size and shape are determined by the
structure of the building blocks, 2D COFs represent a class of porous
materials with tailor-made components and compositions. This
chapter is divided into three sections, which focus on the design and
synthesis of COFs.

1.2 Design and Synthesis

1.2.1 COF Synthesis in the Dynamic Covalent


Chemistry Concept
The crystallization of COFs has generally been dominated by
kinetically controlled reactions that form irreversible covalent
bonds. Higher reversibility, that is, dynamic covalent chemistry
(DCC), must be achieved for efficient error correction [5]. DCC is
Design and Synthesis 3

thermodynamically controlled and thus offers reversible reaction


systems with “error checking” and “proof-reading” characteristics,
leading to the formation of the most thermodynamically stable
structures. Figure 1.1a presents one such reversible assembly of
molecular components, proceeding along kinetic or thermodynamics
pathways. Due to the lower barrier of activation (DG‡), kinetic
intermediates dominate initially due to a fast rate of formation of
the kinetic products, which can be trapped during quenching of the
reaction and form amorphous products with a gain of entropy from
their randomness of arrangement. Meanwhile the thermodynamic
products have the lowest overall Gibb’s free energy (DG°), which
gives the opportunity for the reaction to re-equilibrate toward the
global minimum to form the most stable and probably crystalline
products. In contrast, kinetic products can also be transformed into
thermodynamic products by using certain thermotreatments for
recrystallization. Thermodynamic products have in common COF
chemistry of the amorphous products initially obtained until the
synthetic conditions are optimized by varying the crystallization
time, reaction temperatures, reaction solvents, and catalysts (such
as acids, bases, and metal salts acting as Lewis acids). Simply put,
on using the DCC concept for the construction of COFs, a polymer
skeleton is formed alongside the crystallization process, while the
self-healing feedback reduces the incidence of structural defects and
assists in the formation of an ordered structure. As a result, the final
COF product possesses an ordered crystalline structure with high
thermodynamic stability. Figure 1.1b shows the assembled polymer
formation states of products between transition state and crystalline
products; first the building blocks prepare hyperblanching polymers
or cross-linking polymers. Then pore formation occurs with some
pore errors. Finally, error correction using the DCC concept helps
form a layer with no pore errors and the COFs are precipitated as
crystalline products in the reactive solution.
On the other hand, rational design of building blocks may also
increase the correctness of the assembly of monomer components,
enhancing the crystallinity of the obtained products. It is generally
believed that 2D COFs are easier to form, having less probability
of resulting in networks than the 3D ones, with more structural
diversity.
4 Design and Synthesis

(a) (b)

Figure 1.1 (a) Thermodynamic control in the dynamic covalent chemistry of


COFs and (b) assembly state of COFs between thermodynamic control.

1.2.2 Dynamic Linkages of Building Blocks


One of the most significant characters of COFs is their periodic 2D or
3D polymeric networks linked by dynamic covalent bonds (DCBs),
which are generally formed through reversible organic reactions
as described above. Different from traditional interactions, DCBs
possess robustness of covalent bonds and error-checking capability
of reversible bonds at the same time [6], which endow COFs
with great thermostability, multifarious adaptiveness, as well as
considerable crystallinity. COFs exhibit various performance with
different dynamic linkages. For example, robust linkage will enhance
their chemical stability at the expense of crystallinity since there
is a tradeoff between the two properties [7]. Table 1.1 presents a
list of typical dynamic linkages and their structures, stability, and
crystallinity.
In 2005, the Yaghi research group synthesized two microcrystal-
line 2D frameworks [8], COF-1 and COF-5, by the self-condensation
of 1,4-benzenediboronic acid (BDBA) and co-condensation of BDBA
with hexahydroxyltriphenylene (HHTP). Further, borosilicate [9],
borazine [25], and ionic spiroborate linkages [10] were used for the
formation of COFs. Lewis acid/base interactions of boronate esters
with nitrogen and oxygen nucleophiles [27] were utilized for the
formation of supramolecular frameworks [28].
For imine condensation, the Yaghi research group pioneered both
imine [11, 29] and hydrazine [12] functionalities for the construction
of COFs. Furthermore, azine [17] and imide [18] linkages have been
utilized for the synthesis of COFs. The cyclotrimerization of cyano
Design and Synthesis 5

groups can lead to covalent triazine frameworks (CTFs) under


ionothermal conditions [21]. In most cases, however, amorphous
materials are obtained, especially at higher temperatures [30].
Several strategies have been developed to increase the hydrolytic
stability of self-assembled materials based on C=N bonds. The
Banerjee research group used 1,3,5-triformylphloroglucinol (Tp)
as a trigonal building block for COFs [15]. This resulted in the
irreversible conversion of reversibly formed enol imines into
the more stable b-ketoenamines. Framework TpPa-2 [15] (Pa-2:
2,5-dimethyl-p-phenylendiamine) showed exceptional stability
toward both the acidic 9 N HCl and the basic 9 N NaOH. In another
study, however, Dichtel and coworkers showed that such linkages
can still undergo dynamic exchange with amines and attributed the
stabilizing effect merely to hydrogen bonding instead of irreversible
tautomerization [31]. The stabilization of imine by intermolecular
hydrogen bonding was shown for the porphyrin-containing
COF 2,5-dihydroxyterephthalaldehyde–5,10,15,20-tetrakis(4-
aminophenyl)-21H,23H-porphyrin (DhaTph), whereby Dha was
utilized as a linear linker [32]. Samples of DhaTph remained stable
even after suspension in water for more than 7 days. Quantitative
conversion of imine linkages into the more stable amides resulted in
the improved stability of the COFs [14], which was stabilized by the
irreversible reduction of imine to the more stable amide.
Other types of dynamic covalent reactions have been utilized
to a much lesser extent. Conjugated COFs were obtained by
Michael addition elimination reactions [16] or the condensation
of aldehydes with benzylic nitriles [24], with the latter approach
resulting in an all-carbon framework. The polycondensation of
squaric acid with a tetraaminoporphyrin led to the formation of a
crystalline squaraine framework [19]. Wuest and coworkers utilized
the reversible dimerization of the nitroso group to form azodioxides
in the synthesis of single-crystalline 3D COFs [26].
This section summed up briefly what kinds of dynamic covalent
linkages are utilized for the preparation of COFs. In Chapter 2, the
COFs that build typical bonding structure are presented in detail by
distinguishing DCBs as B–O (boroxine, dioxaborate, and spiroborate
bonds) and C–N (imine, hydroazone, azine, squaraine, imide,
phenazine, imidazole, and triazine bonds) linkages.
6 Design and Synthesis

Table 1.1 Typical dynamic linkages for the construction of COFs

Bonds Linkages Characters Ref.


Boronate ester Formation temp.: 120°C [8]
OH HO O Crystallinity: Excellent
B + B
OH HO O Thermostability: 600°C
Chemical stability: Sensitive
to water, acid, base, alcohols,
and moisture
Boroxine Formation temp.: 120°C [8]
Crystallinity: Excellent
Thermostability: 500°C
OH Chemical stability: Sensitive
B
OH O
B
O to water, acid, base, alcohols,
B
O
B and moisture
B–O
Borosilicate Formation temp.: 120°C [9]
Crystallinity: Excellent
HO
B
OH
Si
OO
Thermostability: 450°C
B O B
HO
Si
OH
+
O O
B Chemical stability: For 1 h
OH O
Si
in air

Spiroborate Formation temp.: 120°C [10]


OH OH Crystallinity: Excellent
+
O
B
O O
B
O Thermostability: 400°C
O O O
(weight loss 7%–12%)
Chemical stability: 2 days
in water and 1 M LiOH;
sensitive to acid
Imine Formation temp.: 120°C [11]
O Crystallinity: Good
+ H2N
H
N
Thermostability: 500°C
Chemical stability: Better
than boron-based COFs
C–N
Hydrazone Formation temp.: 120°C [12]
O O
Crystallinity: G
H
+
H2N NH
N NH
Thermostability: 280°C
O
Chemical stability: Better
than imine-linked COFs
Design and Synthesis 7

Bonds Linkages Characters Ref.


Imide Formation temp.: 200°C [13]
O O Crystallinity: Good
O + H2 N N Thermostability: 530°C
O O Chemical stability: Excellent
Amide Formation temp.: 120°C/r.t. [14]
O Crystallinity: Good
+ H 3N
H
N
Thermostability: 400°C
O
Chemical stability: 1 day in
HN
12 HCl and 1 M NaOH
b-ketoenamine Formation temp.: 120°C [15]
Crystallinity: Good
CHO
Thermostability: 350°C
HO OH
+ H2N
N
Chemical stability: More than
HO OH
OHC
OH
CHO
N N
7 days in boiling water, 9 M
OH HCl, and 9 M NaOH
HN

O O
H H
N N

C–N b-ketoenamine Formation temp.: 130°C [16]


O O
Crystallinity: Moderate
OH + H2N HN
Thermostability: 300°C
Chemical stability: 7 days
in hot water (50°C) and 9 M
HCl; partial hydrolysis in 9 M
NaOH
Azine Formation temp.: 120°C [17]
Crystallinity: Moderate
O

H
+ H2N NH2 N N Thermostability: 300°C
Chemical stability: 1 day in
water, 1 M HCl, 1 M NaOH,
and common organic solvents
Phenazine Formation temp.: 120°C [18]
O H 2N N Crystallinity: Moderate
O
+
H 2N N
Thermostability: More than
1000°C
Chemical stability: 1 day in
water, 1 M HCl, 1 M NaOH,
and common organic solvents
(Continued)
8 Design and Synthesis

Table 1.1 (Continued)


Bonds Linkages Characters Ref.
Squaraine Formation temp.: 85°C [19]
Crystallinity: Moderate
O O
O Thermostability: 300°C
+ H 2N HN NH Chemical stability: 1 day in
HO OH
O water, 1 M HCl, 1 M NaOH,
and common organic solvents
Viologen Formation temp.: 120°C [20]
NO2 Crystallinity: Moderate
NH2 + N
O2N N
Thermostability: 400°C
Chemical stability: 3 days in
boiling water and 6 M HCl
and sensitive to 1 M NaOH
Triazine Formation temp.: 150°C [21]
Crystallinity: Poor
C–N Thermostability: 350°C
Chemical stability: High (no
N
CN N precise description)
N

Melamine Formation temp.: Microwave [22]


Crystallinity: Poor
Thermostability: 400°C
NH2 N N
N Chemical stability: Stable in
H 2N N N NH
N
N N
water and common solvents
NH2 N N N
+
HN N N
H
OHC CHO

N N

C–C Formation temp.: 300°C [23]


(annealing)
X
Crystallinity: Not reported
X : Cl, Br.....
Thermostability: 250°C
Chemical stability: Not
C–C reported
C=C Formation temp.: 120°C [24]
O NC
NC Crystallinity: Moderate
+
H Thermostability: 250°C
Chemical stability: High (no
precise description)
Design and Synthesis 9

Bonds Linkages Characters Ref.


Borazine Formation temp.: 120°C [25]
Crystallinity: Good
Thermostability: 420°C
B–N Chemical stability: No
NH HB
N
BH
reported
BH 2 N N
B
H

Azodioxide Formation temp.: 37°C [26]


Crystallinity: Monocrystalline
O
N Thermostability: Less than
N–N NO N
O 130°C
Chemical stability: Not
reported

1.2.3 Topology and Geometry of 2D Porous Materials


Containing COFs
Porous materials represent an important subject from a scientific
point of view to study phenomena related to surfaces and confined
spaces and have significant practical technological applications
[33–35]. The design of methodologies to precisely control pore
size and distribution to achieve a large surface area is a central
theme in the exploration of porous materials containing COFs.
For example, inorganic porous frameworks, the so-called zeolites,
have recently attracted significant attention for their controllable
porosity and systematically tailored chemical environment within
the pores ranging from the micro- (<2 nm) to the mesoscale
(>2 nm). Microporous zeolites, which are crystalline silicates or
aluminosilicates, are typically synthesized under hydrothermal
conditions from reactive gels in alkaline or acidic media at elevated
temperatures [36–39].
In contrast to zeolites, MOFs with organic components in the
frameworks can be synthesized using the coordination chemistry of
metal ion bonds [40–44]. Thus, the formation of MOFs is driven by
noncovalent coordination bonds [45–47], the diversity of metal ions,
and the availability of different organic structures and topologies.
The pore size can be tuned systematically from micro- or mesometer
scales by using organic ligands of different lengths, while the
10 Design and Synthesis

topology of the framework is retained [48]. MOFs with large surface


areas are very useful for gas storage and separation [49, 50]. MOFs
have also been explored for their ability to act as catalysts for organic
transformation [51] and polymerization [52] to show chirality [53],
conductivity [54], luminescence [55], and magnetism [56] and to
trigger spin transition [57] and nonlinear optical phenomena [58].
MOFs are unique in that they allow for topological design of
porous structures because of the discrete angles and distances
required for orbital overlap between metal ions and organic ligands.
The discovery of MOFs has thus opened up new possibilities for
designable molecular porous materials. However, one drawback
of MOFs is that their structures are based on coordination bonds,
which are much weaker than covalent bonds and make the structure
fragile.
From the molecular design point of view, one of the significant
characters of 2D COFs is the topological construction of the
framework. This is possible because the geometries of building blocks
determine the porosity and pore size. The connections between the
precursors are built from covalent bonds and can be considered as
rather rigid and an approximately linear connection between two
subunits. Therefore, in this section, the topology and geometry of the
assemblies are predominantly defined by both the number and the
spatial distribution of reactive units in the molecular precursors. As
a consequence of the rich chemistry of functionalized (poly)aromatic
hydrocarbons, a great variety of planar precursors are available and
numerous examples with C2, C3, C4, and C6 symmetries have been
used to design and synthesize 2D COFs with various symmetries
(Fig. 1.2). For example, condensation of C3 monomers with C2
monomers leads to 2D COFs with a hexagonal pore structure. On the
other hand, condensation of the C4 and C2 monomers forms 2D COFs
with tetragonal pores. Furthermore, triangular pore structures are
formed by the condensation of C2 with C6 monomers.
To obtain 3D COFs, at least one building has to be extend in
all three spatial directions. As a result of the maximum valency of
four for sp3-hybridized C atoms, simple 3D organic building blocks
are to date exclusively based on tetrahedral tetraphenylmethane
derivatives. The first work on 3D COFs was presented by Yaghi et al., in
2007, when they reported the successful formation of the 3D organic
frameworks COF-102, COF-103, COF-105, and COF-108 for the first
time by self-condensation of tetrakis-(4-dihydroxyborylphenyl)
Design and Synthesis 11

methane and its silane analogue tetrakis[4-dihyroxyborly]phenyl]


silane and co-condensation with HHTP [4]. Overall, 3D COFs are
still very limited in number compared to 2D COFs, presumably
because of the increased difficulties and challenges in the synthesis
of appropriate building blocks, crystallization, and structure
verification.

S
R1 R1 N N O O
R2 R2 (HO)2B B(OH)2
R1 = B(OH)2, CHO, NH2 (HO)2B B(OH)2 (HO)2B N N B(OH)2
R1 = B(OH)2, NH2 O O
= R3 R4
HO OH HO OH

HO OH (HO)2B B(OH)2 (HO)2B B(OH)2


HO OH
R3 R4 etc.
R3 = H, CH3, CH2CH3, C2H4CH3 R4 = H, CH3

R1
HO OH
R1 CHO
= R2 R2

R1 R1 OHC CHO
R2 HO OH
HO OH
R1 R1 HO OH OHHO OH OH
R2 = H, OH etc.
R1 = B(OH)2, CHO, NH2

R
HO OH

N N N
HO OH
= R
N
M1
N
R
HO
N M2 N
OH
N

etc.
HO OH
R
M1 = H2, Zn, Cu, Co M2 = Ni, Zn
R = B(OH)2, NH2

H2 N NH2

H2 N NH2

= H2 N NH2 H 2N NH2

H2 N NH2 etc.

H2 N NH2

Figure 1.2 Different topologies that have been realized in 2D COFs and
selected examples of planar building blocks for the formation of 2D COFs.
12 Design and Synthesis

1.3 Synthetic Methods of COFs


Irrespective of the kind of material, a facile and productive method
can be intriguing and makes it possible for the industrialization of
production. Besides that, some special processes are also needed to
build ideal structures. Given that, the building units and synthetic
routes can be carefully selected on the basis of the design principles
discussed above. Finding suitable synthetic conditions for COF
synthesis is by no means a trivial issue. Since Yaghi and coworkers
exploited the solvothermal (ST) conditions [3], many synthetic
methods have been tried in order to search for a suitable way to satisfy
the needs of extensive applications. Herein, the primary synthetic
methods, such as ST synthesis [4, 11], ionothermal synthesis [21],
microwave synthesis [59, 60], mechanochemical (MC) synthesis [61,
62], and room-temperature (RT) synthesis [63, 64], are discussed.

1.3.1 Solvothermal Synthesis


Most COFs are prepared via the ST synthesis method. A typical ST
condition for the synthesis of COFs is as follows. Monomers and
mixed solvents are placed in a Pyrex tube and degassed via several
freeze-pump-thaw cycles. The tube is then sealed and heated to a
designated temperature for a certain reaction time (takes 2 to 9
days). The precipitate is collected, washed with suitable solvents,
and dried under vacuum to yield the COF as a solid powder. Issues
and challenges such as solubility, the reaction rate, crystal nucleation,
the crystal growth rate, and a self-healing structure are the most
important points to consider when selecting the reaction media and
conditions for the synthesis of COFs.
∑ Solvent combinations and ratios are important factors in
balancing between framework formation and crystallization
when synthesizing highly crystalline COFs.
∑ A suitable temperature is also important to ensure the
reversibility of the utilized linkage reaction. Generally, COFs
have been prepared at temperatures ranging between 80°C
and 120°C, depending on the chemical reactivities of the
building blocks. A closed reaction environment is required to
allow the presence of water molecules, which could trigger
the reverse reaction in the system.
Synthetic Methods of COFs 13

1.3.2 Ionothermal Synthesis


Ionothermal synthesis was utilized to prepare triazine-linked COFs
and was first reported by Thomas and coworkers [21]. Normally,
aromatic nitrile as building blocks (e.g., 1,4-dicyanobenzene;
2,6-dicyanobenzene; and 1,3,5-tri(4-cyanophenyl)benzene)
are dissolved in molten ZnCl2 at 400°C and reacted for 40 h.
In this process, ZnCl2 acts as solvent as well as catalyst so the
cyclotrimerization reaction seems to be partially reversible. When
the system is cooled down to ambient temperature, the salts are
washed out and crystalline CTFs are prepared after purification.
However, harsh reaction conditions limit the building block
availability. Most synthesized CTFs are amorphous materials that
lack long-range molecular order.

1.3.3 Microwave Synthesis


Microwave synthesis has a long history in organic chemistry [65],
but it was not until 2009 that Cooper and coworkers used this
method to prepare COFs [59, 66]. The primal process was similar
to ST synthesis. Microwave heating could be used to obtain 2D
COF-5 and 3D COF-102 in 20 min., which is more than 200 times
faster than the reaction time of 72 h required in the ST synthesis [8].
Moreover, the Brunauer–Emmett–Teller (BET) surface area of COF-
5 (2019 m2/g) obtained via microwave synthesis is slightly higher
than that solvothermally synthesized in a sealed vessel (1590 m2/g).
Compared with ST methods, microwave heating allows the synthetic
process to be completed in a faster and cleaner manner, providing a
new possibility for further applications on larger scales.

1.3.4 Mechanochemical Synthesis


General preparations of COFs are used by dissolving building blocks
in solvents, in which they can be limited by the physicochemical
properties of monomers and the dissolving capacity of the solvents.
The residual organic solvents may also feature toxic products [67].
MC synthesis, a traditional process technology on an industrial scale,
may be an alternative to avoid these drawbacks [68, 69]. In the last
few years, the MC method has been utilized in the preparation of
porous organic polymers, including COFs [70].
14 Design and Synthesis

In 2013, Banerjee and coworkers were first reported to have


built b-ketoenamine-linked COFs by manual grinding in a mortar and
pestle [61]. Interestingly, visual color changes could be observed and
indicated the extent of polycondensation. After 45 min., the color of
mixtures turned to dark red, which stood for complete COF formation
(referred to as MC COFs). In comparison to the same COFs obtained
from ST synthesis (referred to as ST COFs), MC COFs showed similar
chemical stability but their crystallinity was moderate. To reduce
such shortcomings, the liquid-assisted grinding method was applied
to prepare imine, b-ketoenamine, and hydrogen-bonded imine-
linked COFs [71]. Mechanical grinding can also be used to produce
covalent organic nanosheets (CONs) from bulk COFs [72]. The
prepared CONs retained their structural integrity and were stable
in water, acids, and bases before exfoliation. Recently, Banerjee and
coworkers developed a new MC method to continuously obtain
COFs with a twin-screw extruder [62]. These COFs feature highly
crystalline, ultrahigh porosity, and a high surface area. What’s more,
the COFs can be molded into any desired shapes and sculptures
through heating, mimicking the ancient terracotta process. This
approach is widely suitable for imine-linked COFs.
As mentioned above, CTFs were prepared by the ionothermal
method in the beginning [21]. Later, the strategies of microwave
synthesis and RT synthesis were developed with the help of a strong
Brønsted acid catalyst by Cooper and coworkers. But the operating
conditions and product yield were still unsatisfactory. Banerjee and
coworkers expanded MC synthesis to construct CTFs [73]. Instead of
the conventional cyclotrimerization of nitriles, the new approach is
based on Friedel–Crafts alkylation.
In brief, MC synthesis is a user-friendly, ecofriendly, and time-
saving alternative to produce COFs on a massive scale. It provides us
a powerful method to realize the industrial production of COFs.

1.3.5 Room-Temperature Synthesis


RT synthesis was first mentioned in the review of Wang and
coworkers, who first found that imine-based COFs could be facilely
synthesized at RT in an ambient atmosphere [74], though their
research paper was not published until 2017 [63]. By avoiding both
the use of sealed vessels and the difficulty in controlling different
References 15

synthetic parameters, this method makes bulk production of


COF materials possible. The generality of this approach is under
investigation. Further Zamora and coworkers used m-cresol and
DNSO as a solvent instead of the traditional mixture solvent and
prepared RT-COF-1 in minutes [75, 76]. Later, they combined this
methodology with microfluid technology and constructed COFs in
fibrillar microstructures [77]. When Sc(TfO)3 was used to replace
acetic acid, the COFs were synthesized within 10 min. and possessed
a large specific surface area [78].

References

1. (a) Eddaoudi, M., Moler, D. B., Li, H., Chen, B., Reineke, T. M., O’Keefe,
M., Yaghi, O. M. (2001). Acc. Chem. Res., 34, 319–330; (b) Kitagawa,
S., Kitamura, R., Noro, S. (2004). Angew. Chem. Int. Ed., 43, 2334–2375.
2. (a) Ockwig, N. W., O’Keefe, M., Matzger, A. J., Yaghi, O. M. (2005). Science,
310, 1166–1170; (b) Han, S. S., Frukawa, H., Yaghi, O. M. Goddard, W. A.
(2008). J. Am. Chem. Soc., 130, 11580–11580.
3. (a) Hunt, J. R., Doonan, C. J., LeVangie, J. D., Côté, A. P., Yaghi, O. M.
(2008). J. Am. Chem. Soc., 130, 11872–11873; (b) Côté, A. P., EI-Kaderi,
H. M., Furukawa, H., Hunt, J. R., Yaghi, O. M. (2007). J. Am. Chem. Soc.,
129, 12914–12915.
4. EI-Kaderi, H. M., Hunt, J. R., Mendoza-Cortez, A. P., Côté, A. P., Taylor, R.
M., O’Keefe, M., Yaghi, O. M. (2007). Science, 316, 268–272.
5. Rowan, S. J., Cantrill, S. J., Cousins, G. R. L., Sanders, J. K. M., Stoddart, J.
F. (2002). Angew Chem. Int. Ed., 41, 898–952.
6. Jin, Y., Yu, C., Denman, R. J., Zhang, W. (2013). Chem. Soc. Rev., 42, 6634–
6654.
7. Zhu, L., Zhang, Y. B. (2017). Molecules, 22, 1149.
8. Côté, A. P., Benin, A. I., Ockwig, N. W., O’Keefe, M., Matzager, A. J., Yaghi,
O. M. (2005). Science, 310, 1166–1170.
9. Hunt, J. R., Doonan, C. J., Levangie, J. D., Côté, A. P., Yaghi, O. M. (2008). J.
Am. Chem. Soc., 130, 11872–11873.
10. Du, Y. Yang, H., Whiteley, J. M., Wan, S., Jin, Y., Lee, S. H., Zhang, W. (2016).
Angew. Chem. Int. Ed., 55, 1737–1741.
11. Uribe-Romo, F. J., Hunt, J. R., Furukawa, H., Klock, C., O’Keefe, M., Yaghi,
O. M. (2009). J. Am. Chem. Soc., 131, 4570–4571.
12. Uribe-Romo, F. J., Doonan, C. J., Furukawa, H., Oisaki, K., Yaghi, O. M.
(2011). J. Am. Chem. Soc., 133, 11478–11481.
16 Design and Synthesis

13. Fang, Q., Zhuang, Z., Gu, S., Kaspar, R. B., Zheng, J., Wang, J., Qiu, S., Yan,
Y. (2014). Nat. Commun., 5, 4503.
14. Waller, P. J., Lyle, S. J., Osborn Popp, T. M., Dierck, C. S., Reimer, J. A.,
Yaghi, O. M. (2016). J. Am. Chem. Soc., 138, 15519–15522.
15. Kandambath, S., Mallick, A., Likose, B., Mane, M. V., Heine, T., Banerjee,
R. (2012). J. Am. Chem. Soc., 134, 19524–19527.
16. Rao, M. R., Fang, Y., Feyter, S. D., Perepichka, D. F. (2017). J. Am. Chem.
Soc., 139, 2421–2427.
17. Dalapati, S., Jin, S., Gao, J., Xu, Y., Nagai, A., Jiang, D. (2013). J. Am. Chem.
Soc., 135, 173101–17313.
18. Guo, J., Xu, Y., JIn, S., Gao, J., Chen, L., Kaji, T., Honsho, Y., Addicoat, M. A.,
Kim, J., Saeki, A., Ihee, H., Seki, S., Irle, S., Hiramoto, M., Gao, J., Jaing D.
(2013). Nat. Commnun., 4, 2736.
19. Nagai, A., Chen, X., Feng, X., Ding, X., Guo, Z., Jiang, D. (2013). Angew.
Chem. Int. Ed., 52, 3770–3774.
20. Shi, W., Xing, F., Bai, T.-L., Hu, M., Zhao, Y., Li, M-X., Zhu, S. (2015). ACS
Appl. Mater. Interface, 7, 14493–14500.
21. Kuhn, P., Antonietti, M., Thomas, A. (2008). Angew. Chem. Int. Ed., 47,
3450–3453.
22. Zhang, W., Qiu, L. G., Yuan, Y. P., Xie, A. J., Shen, Y. H., Zhu, J. F. (2012). J.
Hazard. Mater., 221, 147–157.
23. Gutzler, R., Walch, H., Eder, G., Kloft, S., Heckl, W. H., Lackinger, M.
(2009). Chem. Commoun., 7, 4456–4458.
24. Jin, E., Asada, M., Xu, Q., Dalapti, S., Addicoat, M. A., Brady, M. A., Xu, H.,
Nakamura, T., Heine, T., Chen, Q., Jiang, D. (2017). Science, 357, 673–
676.
25. Jackson, K. T., Reich, T. E., EI-Kaderi, H. M. (2012). Chem. Commun., 48,
8823–8825.
26. Beaudoin, D., Maris, T., Wuest, J. D. (2013). Nat. Chem., 5, 830–834.
27. Höpfl, H. (1999). J. Organomet. Chem., 581, 129–149.
28. Sheepwash, E., Krampl, V., Scopelliti, R., Sereda, O., Neels, A., Sverin, K.
(2011). Angew. Chem. Int. Ed., 50, 3034–3037.
29. Ding, S.-Y., Gao, J., Wang, Q., Zhang, Y., Song, C.-G., Su, C.-Y., Wang, W.
(2011). J. Am. Chem. Soc., 133, 19816–19822.
30. Sakaushi, K., Antonietti, M. (2015). Acc. Chem. Res., 48, 1591–1600.
31. DeBlase, C. R., Silberstein, K. E., Truong, T.-T., Abruña, H. D., Dichtel, W.
R. (2013). J. Am. Chem. Soc., 135, 16821–16824.
References 17

32. Kandambeth, S., Shinde, D. B., Panda, M. K., Lukose, B., Heine, T.,
Banerjee, R. (2013). Angew. Chem. Int. Ed., 123, 13052–13056.
33. Davis, M. S. (2002). Nature, 417, 813–821.
34. Cundy, C. S., Cox, P. A. (2003). Chem. Rev., 103, 663–701.
35. Wan, Y., Zhao, D. (2007). Chem. Rev., 107, 2812–2860.
36. Sakamoto, Y., Kaneda, M., Terasai, O., Zhan, D. Y., Kim, J. M., Stucky, G. D.,
Shim, H. J., Ryoo, R. (2000). Nature, 408, 449–453.
37. Che, S., Garcia-Bennett, A. E., Yokoi, T., Sakamoto, K., Kunieda, H.,
Terasaki, O., Tatsumi, T. (2003). Nat. Mater., 2, 801–805.
38. Shen, S. D., Garcia-Bennett, A. E., Liu, Z., Lu, Q. Y., Shi, Y. F., Yan, Y., Yu, C.
Z., Liu, W. C., Cai, Y., Terasdaki, O., Zhao, D. Y. (2005). J. Am. Chem. Soc.,
127, 6780–6787.
39. Tan, B., Dozier, A., Lehmler, H. J., Knutson, B. L., Rankin, S. E. (2004).
Langmuir, 20, 6981–6984.
40. Tranchemontague, D. J., Mendoza-Cortés, J. L., O’Keeffe, M., Yaghi, O. M.
(2009). Chem. Soc. Rev., 38, 1257–1283.
41. Uemura, T., Yanai, N., Kitagawa, S. (2009). Chem. Soc. Rev., 38, 1228–
1236.
42. Wang, Z., Cohen, S. M. (2009). Chem. Soc. Rev., 38, 1315–1329.
43. Shimizu, G. K. H., Vaidhyanathan, R., Taylor, J. M. (2009). Chem. Soc.
Rev., 38, 1430–1449.
44. Li, J. R., Kuppler, R. J., Zhou, H. C. (2009). Chem. Soc. Rev., 38, 1477–
1504.
45. Asefa, T., MAcLachlan, M. J., Coombs, N., Ozin, G. A. (1999). Nature, 402,
867–871.
46. Eddaoudi, M., Kim, J., Rosi, N., Vodak, D., Wachter, J., O’Keeffe, M., Yaghi,
O. M. (2002). Science, 295, 469–472.
47. James, S. L. (2003). Chem. Soc. Rev., 32, 276–288.
48. Yaghi, O. M., O’Keefe, M., Ockwig, N. W., Chae, H. K., Eddoudi, M., Kim, J.
(2003). Nature, 423, 705–714.
49. Rowsell, J. L. C., Milward, A. R., Park, K. S., Yaghi, O. M. (2004). J. Am.
Chem. Soc., 126, 5666–5667.
50. Chen, B. L., Liang, C. D., Yang, J., Contreas, D. S., Clancy, Y. L., Lobkovsky,
E. B., Yaghi, O. M. (2006). Angew. Chem. Int. Ed., 45, 1390–1393.
51. Heibaum, M., Glorius, F., Escher, I. (2006). Angew. Chem. Int. Ed., 45,
4732–4762.
18 Design and Synthesis

52. Kitagawa, S., Kitaura, R., Noro, S. (2004). Angew. Chem. Int. Ed., 43,
2334–2375.
53. Kepert, C. J., Prior, T. J., Rosseinsky, M. J. (2000). J. Am. Chem. Soc., 122,
2334–2375.
54. Sadakiyo, M., Yamada, T., Kitagawa, H. (2009). J. Am. Chem. Soc., 131,
9906–9907.
55. Chandler, B. N., Cramb, D. T., Shimizu, G. K. H. (2006). J. Am. Chem. Soc.,
128, 10403–10412.
56. Zheng, M. H., Wang, B., Wang, X. Y., Chen, X. M., Gao, S. (2006). Inorg.
Chem., 128, 10403–10412.
57. Agusti, G., Munoz, M. C., Gaspar, A. B., Real, J. A. (2009). Inorg. Chem., 48,
3371–3381.
58. Zhang, L. J., Yu, J. H., Xu, J. Q., Lu, J., Bie, H. Y., Zhang, X. (2005). Chem.
Commun., 8, 638–642.
59. Campbell. N. L., Clowes, R., Ritchie, L. K., Cooper, A. I. (2009). Chem.
Mater., 21, 204–206.
60. Dogru, M., Sonnauer, A., Gavryushin, A., Knochel, P., Bein, T. A. (2011).
Chem. Commun., 47, 1707–1709.
61. Biswal, B. P., Chandra, S., Kandambeth, S., Lukose, B., Heine, T., Banerjee,
R. (2013). J. Am. Chem. Soc., 135, 5328–5331.
62. Karak, S., Kandambeth, S., Biswal, B. P., Sasmal, H. S., Kumar, S., Pachfule,
P., Banerjee, R. (2017). J. Am. Chem. Soc., 139, 1856–1862.
63. Ding, S. Y., Cui, X. H., Feng, J., Lu, G., Wang, W. (2017). Chem. Commun.,
53, 11956–11959.
64. Marumoto, M., Dasari, R. R., Ji, W., Feriante, C. H., Parker, T. C., Marder, S.
R., Dichtel, W. R. (2017). J. Am. Chem. Soc., 139, 4999–5002.
65. De La Hoz, A., Diaz-Ortiz, A., Moreno, A. (2005). Chem. Soc. Rev., 34,
164–178.
66. Ritchie, L. K., Trewin, A., Reguera-Galan, A., Hasell, T., Cooper, A. I.
(2010). Miroporous Mesoporous Mater., 132, 132–136.
67. Friscic, T. (2012). Chem. Soc. Rev., 41, 3493–3510.
68. Friscic, T., James, S. L., Boldyreva, E. V., Bolm, C., Jones, W., Mack, J.,
Steed, J. W., Suslick, K. S. (2015). Chem. Commun., 51, 6248–6256.
69. James, S. L., Adams, C. L., Bolm, C., Braga, D., Collier, P., Priscic, T.,
Grepioni, F., Harris, K. D., Hyett, G., Jones, W. (2012). Chem. Soc. Rev.,
41, 413–447.
70. Zhang, P., Dai, S. (2017). J. Mater. Chem. A, 5, 16118–16127.
References 19

71. Das, G., Balaji Shinde, D., Kandambeth, S., Biswal, B. P., Banerjee, R.
(2014). Chem. Commnun., 50, 12615–12618.
72. Chandra, S., Kandambeth, S., Biswal, B. P., Lukose, B., Kunjir, S. M.,
Chaudhary, M., Babarao, R., Heine, T., Banerjee, R. (2013). J. Am. Chem.
Soc., 135, 17853–17861.
73. Troschke, E., Grätz, S., Lübken, T., Borchardt, L. (2017). Angew. Chem.
Int. Ed., 56, 14149–14153.
74. Ding, S. Y., Wang, W. (2013). Chem. Soc. Rev., 42, 548–568.
75. De la Pena Ruigomez, A., Rodriguez-San-Miguel, D., Stylianou, K.
C., Cavallini, M., Gentili, D., Liscio, F., Milita, S., Roscioni, O. M., Ruiz-
Gonzalez, M. L., Carbonell, C. (2015). Chemistry (Easton), 21, 10666–
10670.
76. Montoro, C., Rodriguez-San-Miguel, D., Polo, E., Escudero-Cid, R., Ruiz-
Gonzalez, M. L., Navarro, J. A. R., Ocon, P., Zamora, F. (2017). J. Am. Chem.
Soc., 139, 10079–10086.
77. Rodriguez-San-Miguel, D., Abrishamkar, A., Navarro, J. A., Rodriguez-
Trujillo, R., Amabilino, D. B., Mas-Balleste, R., Zamora, F., Puigmarti-
Luis, J. (2016). Chem. Commun., 52, 9212–9215.
78. Matsumoto, M., Dasari, R. R., Ji, W., Feriante, C. H., Parker, T. C., Marder,
S. R., Dichtel, W. R. (2017). J. Am. Chem. Soc., 139, 4999–5002.
Chapter 2

Crystallization and Structural Linkages of


COFs

COFs are prepared by integrating organic molecular building blocks


into predetermined network structures entirely through covalent
bonds. The crystallization problem has been conquered by DCC in
synthesis and reticular chemistry in materials design. This chapter
reviews recent progress in the crystallization and structural linkages
of COFs.

2.1 b–O Linkages


With lightweight compositions and high crystallinity in view, boron-
linked covalent organic framework (COFs) have been intensively
studied, including their crystallization mechanism, design of new
topologies, achievement of high surface areas, and their utilization in
gas storage applications. The successful building of such crystalline
materials gained from adequate reversibility of the b–O linkages,
including boroxine (–B3O3), boronic ester (–BO2C2), and spiroborate
(–O4B–), and containing borazine (–B3N3), for example.

2.1.1 Boroxine-Linked COFs


Boroxines are the cyclotrimeric anhydrides of boronic acids. Their
properties and applications have been recently reviewed [1].

Covalent Organic Frameworks


Atsushi Nagai
Copyright © 2020 Jenny Stanford Publishing Pte. Ltd.
ISBN 978-981-4800-87-7 (Hardcover), 978-1-003-00469-1 (eBook)
www.jennystanford.com
22 Crystallization and Structural Linkages of COFs

By virtue of boron’s vacant orbital, boroxines are isoelectronic


to benzene, but it is generally accepted that they possess little
aromatic character [2]. Several theoretical and experimental studies
have addressed the nature and structure of these derivatives [1]; in
particular, the X-ray crystallographic analysis of triphenylboroxine
has confirmed that it is virtually flat [3]. Boroxines are easily
produced by the simple dehydration of boronic acids, either
thermally through azeotropic removal of water or by exhaustive
drying over sulfuric acid or phosphorous pentoxides [4]. These
compounds can be employed invariably as substrates in many
of the same synthetic transformations known to affect boronic
acids. Interest in the applications of boroxines as end products
has increased in the past decade. Their use has been proposed as
flame retardants [5] and as functional materials [5]. The formation
of boroxine cross-linkages has been employed as a means to
immobilize blue light–emitting oligofluorene diboronic acids [6].
A study examined the thermodynamic parameters of boroxine
formation in water (Eq. 2.1) [7]. Using hydrogen-1 nuclear magnetic
resonance (1H NMR) spectroscopy, the reaction was found to be
reversible at room temperature (RT) and the equilibrium constants,
relatively small ones, were found to be subject to substituent
effects. For example, boroxines with a para-electron-withdrawing
group have smaller equilibrium constants. The observation was
interpreted as an outcome of a back-reaction, that is, boroxine
hydrolysis, that is facilitated by the increased electrophilicity of
boron. Steric effects also come into play, as indicated by a smaller K
value for orthotolyboronic acid compared to that for the paraisomer.
Variable temperature studies have provided useful thermodynamic
information, which was found consistent with a significant entropic
drive for boroxine formation due to the release of three molecules
of water.
R

OH O B
3 R B R B O + 3 H 2O (2.1)
OH O B

R
[boroxine] [H 2O] 3
K=
[boronic acid] 3
b–O Linkages 23

According to the stereochemistry and thermodynamic information


of boroxines, the emergence of COF chemistry was triggered by Yaghi
and coworkers’ seminal work on the self-condensation of aromatic
polyboronic acids (Fig. 2.1) [8]. The successful crystallization of
these materials (COF-1, COF-102, and COF-103) was attributed
to the implementation of a closed reaction system to sustain the
availability of water for maintaining reversible conditions conducive
to crystal growth. The solvents used and their mixture are a handle
to control the diffusion of building blocks into the crystallization
mother liquor. The crystallization was accomplished intentionally
slowly and terminated after sufficient time had elapsed.
B(OH)2 B(OH)2

(HO)2B B(OH)2
(HO)2B B(OH)2 (HO)2B Si B(OH)2

B(OH)2 B(OH)2

O B B O
B O O B
O B B O B O O B
B O O B O B
O B B O
B O B B O O B
B B O O B O O
O O
B O O B B B O BO
O B B O
B
O
B O BO O
B B
B O O B B B Si O
O Si
B OB
B OB
B O O
B O O B B O O B O O B
O B B O O B B O B B B O B O
B O O B B O O B B O B O B O
B O B
B B O
B O
B O O B
B O O B B Si B O
B O O B O O
O B B O B
O B B O B O
B O O B
B O O B

B
B O O
O O
O B B O B B
B B B B O
O O O
O B B O

COF-1 COF-102 COF-103

B
O O
B B
O

O O
B B B B
O O O O
B B

(HO)2B B(OH)2

PDBF
B B
O O O O
B B B B
O O

O
B B
O O
B

PPy COF

Figure 2.1 Construction of classic boroxine-linked COFs (COF-1, COF-102,


COF-103) and pyrene-based boroxine-linked COF (PPy COF).

The selection of solvents and temperature to maximize the error


correction associated with the boroxine ring–forming reactions
24 Crystallization and Structural Linkages of COFs

can optimize the crystallinity of the boroxine-linked COFs. COF-1


with 2D architectures was prepared by the self-condensation of
1,4-phenylenediboronic acid under solvothermal conditions to
produce an extended staggered layered structure with hexagonal
pores (7 Å) and a Brunauer–Emmett–Teller (BET) surface of
711 m2/g. Further, 3D COFs COF-102 and COF-103 were also
synthesized by the self-condensation of tetrahedral (3D-Td) nodes
tetra(4-dihydroxyborylphenyl)methane and its silane analog,
respectively (Fig. 2.1) [9]. These two COFs are the crystal porous
organic frameworks with high surface areas (3472 cm2/g for
COF-102 and 4120 m2/g for COF-103) and a pore size distribution of
11.5 Å for COF-102 and 12.5 Å for COF-103.
After these pioneer reports, Jiang and coworkers demonstrated
the first example of a photoconductive COF, in which sheets composed
of arene building blocks lie one above the other in an eclipsed
arrangement (Fig. 2.1) [10]. COF-1 possesses a pore 7 Å in diameter
with a BET surface area of 711 m2/g. On the other hand, a large p
monomer, for example, pyrene-2,7-diboronic acid, under otherwise
identical conditions yields PPy COF with a large pore size (18.8 Å)
[10]. The large pyrene building blocks prefer a supericomposed p–p
interaction. As a result, PPy COF has an eclipsed stack structure.
Microwave heating can accelerate the reaction times, yielding
high amounts of much cleaner COFs. Cooper and coworkers reported
the synthesis and purification of COF-5 and COF-105 by microwave
heating. The microwave synthetic condition reported by them was
200 times faster than the solvothermal condition reported by Yaghi
et al., without changing the physical properties of COFs [11].

2.1.2 Boronic Ester (Dioxaborole)-Linked COFs


OH HO -2 H2O O
B + B (2.2)
OH HO O

The preparation of boronic esters (dioxaboroles) from catechols is


straightforward by generating water molecules under solvothermal
conditions (Eq. 2.2). The overall process is an equilibrium, and
the forward reaction is fast with preorganized catechols and
particularly favorable when the boronate product is insoluble in the
reaction solvent. Therefore, the cross-condensation of polyboronic
acids with catechols opens the way to stitch different building
blocks into one framework (Fig. 2.2). The maximum utilization of
HO OH (HO)2B B(OH)2

:
HO OH COF-5 (HO)2B B(OH)2

HO OH
HHTP (HO)2B B(OH)2
COF-10

(HO)2B B(OH)2
PPy COF

(HO)2B S
O O T-COF 1 B(OH)2
B B
O O
(HO)2B S
O O O Pore size O T-COF 3 S B(OH)2
BO B O O B
OB 2.7 nm COF-5
S
3.2 nm COF-10 T-COF 4 (HO)2B B(OH)2
B
F-4 S
CO B O B 2.1 nm T-COF1 B O
O O m O O
n F-8 O 3.2 nm T-COF3 O
0.9 CO (HO)2B B(OH)2
nm 2.6 nm T-COF4 HHTP-DPB-COF
1.6
O O 4.7 nm HHTP-DPB COF N HN
B B O 4.6 nm TP-Por-COF O (HO)2B B(OH)2
O O TP-Por-COF
O B 5.3 nm DTP-ANDI-COF B O NH N
5.3 nm DTP-APyrDI-COF
1-S X
B(OH)2 B O O B 1-Se (HO)2B B(OH)2
O O X
1-Te
B(OH)2 X = S, Se, Te
B(OH)2 O O
: B B O O
O O
B(OH)2 (HO)2B B(OH)2 DTP-ANDI-COF (HO)2B N N B(OH)2
(HO)2B
O O
(HO)2B B(OH)2
COF-4
O O
COF-8
DTP-APPuyDI-COF (HO)2B N N B(OH)2

O O

Figure 2.2 Construction of dioxaborole-linked COFs via the cross-condensation of triangular HHTP with ditopic or tritopic boronic acid to
form 2D honeycomb layer structures.
b–O Linkages
25
26

R
O O
B B
O O
R O B R B O
HO OH R O O R

HO OH O R R O
R B O O B
-H2O
Crystallization and Structural Linkages of COFs

+ THF/CH3OH (99 : 1)
reflux for 3 days B O O B
B(OH)2 O R R O
R = H (18 Å)
= CH3 (16 Å)
R O O R
(HO)2B B(OH)2 O B R B O = CH2CH3 (14 Å)
O O = CH2CH2CH3 (11 Å)
B B
O O
R
Me
O O
B B
O O
Me Me
HO OH
OB BO
O O
HO OH
Me Me Me
Me Me
-H2O O O
+ BO OB
THF/CH3OH (99 : 1)
B(OH)2 reflux for 3 days 4 nm

BO OB
O O
Me Me
(HO)2B B(OH)2
Me Me
O O
OB BO

Me
O O
B B
O O
Me

BTP COF
b–O Linkages

Figure 2.3 Construction of dioxaborole-linked COFs via the cross-condensation of ditopic catechol-based linkers with tritopic boronic
acids to form 2D honeycomb layer structures. The pore sizes can be modulated by the decorated functionalities of the ditopic linkers and
27

expanded by elongating both boronic acid and catechol linkers.


28 Crystallization and Structural Linkages of COFs

2,3,6,7,10,11-hexahydroxytriphenylene (HHTP) as a triangular


building block with ditopic, tritopic, and tetratopic boronic acids
leads to 2D and 3D networks (Fig. 2.2) [12–14]. Using a sealed
reaction system under solvothermal conditions, COF-5, COF-6, COF-
8, and COF-10 were successfully produced in the cocktail solvent
system by mixing mesitylene and 1,4-dioxane for 3–5 days at
85°C–100°C [12, 14]. Therefore, the same COF using HHTP as the
colinker can be also prepared by stirring under an inert atmosphere
[15].
By using the tetrahydroxybenzene derivatives and 1,4-diboronic
acid benzene, 2D hexagonal COFs can be prepared by stitching
tritopic boronic acids together. The reaction was carried out in a
mixed solvent of tetrahydrofuran and methanol (MeOH) by stirring
under an insert atmosphere [16, 17].
Boronic ester linkage–built up COFs have high crystallinity but
low hydrolytic and chemical stability due to the reversibility of
the reactions, which leads to their decomposition upon exposure
to water or acid. This disadvantage limits their applications. COF-
anchored alkyl groups in the pores of COFs improve their stability by
decreasing the hydrolysis rate of the connected linkage. Lavigne and
coworkers were the first to incorporate alkyl groups into channel
walls during the synthesis of boronate ester–linked COFs to increase
the stability of COFs, as well as to fine-tune the pore sizes [17].
The presence of alkyl chains in the channels can help control the
pore sizes so they remain in the range of 1.8 nm to 1.1 nm (Fig. 2.3),
which was observed for the series of COFs with no alkyl groups (18
Å), COFs with methyl groups (16 Å), COFs with ethyl groups (14 Å),
and COFs with propyl groups (11 Å). Meanwhile, the surface area
decreased from 1263 to 105 cm2/g and the pore volume decreased
from 0.69 to 0.052 cm3/g as the alkyl chain length increased.
Interestingly, modification of the pore interior with increasingly
larger alkyl groups causes a decline in nitrogen uptake but an
increase in the molar amount of hydrogen adsorbed. Once they
are submerged in aqueous media, the porosity of alkylated COFs
decreases by about 25%, while the nonalkylated COFs are almost
completely hydrolyzed, losing virtually all porosity. In addition, the
degree of crystallinity decreases by about 40% for alkylated COFs
and by 95% for nonalkylated COFs. By using microwave synthesis,
Bein et al. extended the framework into materials having large
(4 nm) openings (Fig. 2.3), featuring some of the largest pores in
crystalline materials at this time [18].
b–O Linkages 29

Octahydroxyphthalocyanine (Pc) as a large p flat building block


can also be utilized for the preparation of COFs. Dichtel’s group was
able to expand the pore size of the ZnPc lattice from 2.7 to 4.4 nm
(Fig. 2.4) [19]. Furthermore, the selective and patterned growth of
a 2D phthalocyanine COF on single-layer graphene was reported by
Dichtel and coworkers. This provides a promising pathway to direct
the growth of COF films [20].

B B
O O O O

N N N N N N
O O O O
B NH HN B B NH HN B
O O O O
N N N N N N

O O O O
B B

B B
O O O O

N N N N N N
O O O O
B NH HN B B NH HN B
O O O O
N N N N N N

O O O O
B B

(HO)2B B(OH)2 Pore sizes

(HO)2B B(OH)2 ZnPc-Py COF 2.7 nm

(HO)2B B(OH)2 ZnPc-DPB COF 3.4 nm

(HO)2B B(OH)2 ZnPc-PPE COF 4.0 nm

O O

(HO)2B N N B(OH)2 ZnPC-NDI COF 4.4 nm

O O

Figure 2.4 Preparation of dioxaborale-linked phthalocyanate COFs via the


cross-condensation of ditopic catechol-based linkers with tritopic boronic acids
to form a 2D honeycomb layer.
30 Crystallization and Structural Linkages of COFs

To avoid the oxidation and insolubility of polycatechols,


protected catechols were proposed as starting building blocks for
the synthesis of COFs, which could be deprotected in situ under
catalysis by a Lewis acid such as trifluoro boron etholate ether
complex (BF3·OEt2) [21]. According to this strategy, triangular HHTP
and square phthalocyanine as well as its metalated derivatives have
been successfully incorporated into boronate ester–linked COFs
(Fig. 2.5). Mechanistic studies present that before a Lewis acid is
added, the boronic acids self-condensate reversibly to form boroxine-
linked products and water. The addition of BF3·OEt2 catalyzes
acetonide hydrolysis of protected catechols, and the resulting
catechol will rapidly condense with boronic acids to form boronate-
linked COFs [22]. Once the free boronic acids are consumed, the
formed products will hydrolyze and release more boronic acids,
which will produce more boronate-linked COFs.

O O O O

+ (HO)2B B(OH)2
N N N
O O O O
(HO)2B B(OH)2 N Ni N + (HO)2B B(OH)2
O O O O
COF-5 : pore size = 2.7 nm N N N

(HO)2B B(OH)2

COF-10 : pore size = 3.2 nm O O


BF3OEt2
BF3OEt2 Mestylene/ClCH2CH2Cl
Mestylene/Dioxane 120 / 3 days
90 / 3 days

B
O O
B B
O O O O

O O
BO OB N N N N
N N
O O O O
OB BO OB
BO B N Ni N B B N Ni N B
O O O O
O O O O N N
N N N N

O O O O O O O O
B B B B
Pc-PBBA COF :
pore size = 2.0 nm

B B B B
O O O O O O O O

N N N N N N
O O O O O O O O
BO B N Ni N B B N Ni N B
OB BO OB O O O O
N N
BO OB N N N N
O O

O O O O
B B
O O
B

Figure 2.5 Structures of dioxaborole-linked COFs prepared from protected


catechol-based linkers with boronic acids for improvement of solubility and
antioxidation.
b–O Linkages 31

To focus on the preparation of highly crystalline COFs, a two-step


microwave synthesis approach has been used for protected boronic
acids as starting materials with the addition of HHTP. The resulting
product proved to be highly crystalline and possessed large pore
openings (Fig. 2.6). Control experiments did not yield COFs via the
open-pot reaction. This work also provided the first high-resolution
transmission electron microscopy image of a COF showing the pores
clearly (Fig. 2.6b,c) [23].

O O
B B
O O
N N
O S O
(a) BO OB

SN NS
O O N N
B

OB BO
1. HCl (7% v) O O
N BTD COF : 4.1 nm
S
N 2. HO OH
O O
OB BO

B
NS SN
O O HO OH N N

HO OH
BO OB
O O
O O
B B
O O
N N
S

(b) (c)

Figure 2.6 (a) Synthesis of the BTD COF by in situ deprotection of boronic
pinacol ester, followed by the addition of HHTP via microwave heating,
(b) projection along the columns showing the hexagonal pore structures, and
(c) image of a crystal titled out of the columnar projection with a side view of
the pores. Scale bar: (b) 50 nm and (c) 20 nm. Republished with permission
of Royal Society of Chemistry, from Ref. [23], copyright (2013); permission
conveyed through Copyright Clearance Center, Inc.
32 Crystallization and Structural Linkages of COFs

A modulation concept has been developed to control the


structure and crystallinity of COFs, introducing a modulator in the
preparation that can compete with one of the building blocks during
solvothermal COF growth to form highly crystalline frameworks
with a large domain size and very high porosity. Bein and coworkers
utilized monoboronic acids as modulators in the solvothermal
method of the archetypical COF-5 to optimize the crystallinity,
domain size, and porosity of 2D COFs (Fig. 2.7). Furthermore, the
HO OH

+ (HO)2B B(OH)2 + (HO)2B R


HO OH
R : SH, CO2H
HO OH

HHTP BDBA modulator

Mestiylene / dioxane
100
R

B O O B
O O

O O
B B
O O

O O
B O O B

O B B O
O O

O O
R B B R
O COF-5-x O

O O
O B B O

B O O B
O O

O O
B B
O O

O O
B O O B

Figure 2.7 The modulation approach to prepare COF-5-x via the cross-
condensation of HHTP and 1,4-boronic acid with terminal monoboronic acids
with some functional groups to achieve higher crystallinity and external
functionalities.
b–O Linkages 33

addition of monoboronic acids also provides for the potential


construction of functional crystalline COFs [24]. The realization
of highly crystalline COFs with the option of additional surface
functionality will render the modulation concept beneficial for a
range of applications, catalysis, and optoelectronics.

2.1.3 Spiroborate-Linked COFs


HO K 3B 4O7(base)
O O
2 B K+ (2.3)
HO O O

There have been many studies on reactions between boric acid


and diols in connection with changes in conductivity, acidity, and
rotatory polarization [25]. Hermans pointed out that the changes in
these physical properties are due to the formation of a spiroborate
complex that is produced from those reactions in a solution [26].
The first isolation of the spiroborate complex was performed by
Böeseken and coworkers, who synthesized potassium biscatechol
spiroborate from the reaction of catechol with potassium borate
in water (Eq. 2.3) [25]. On the other hand, spiroborates are ionic
derivatives of boronic acid, which have been reported to exhibit high
resistance toward hydrolysis and stability in water, methanol, and
under basic conditions [27, 28]. A spiroborate linkage can be formed
readily through the condensation of polyols with alkali tetraborate
[29, 30–32] or boric acid [33–35] or through the transesterification
between borate and polyols [36] in a thermodynamic manner.
Recently, the condensation of diols with trialkyl borate in the
presence of basic catalysts has been explored for the synthesis of
spiroborate-linked COFs in which the negatively charged boron ions
are located on the edges with different cations as counters ions.
Zhang and coworkers constructed a novel type of spiroborate-linked
ionic COF (ICOF, Fig. 2.8) that contains sp3-hybridized boron anionic
centers and tunable countercations (lithium or dimethylammonium)
[37]. The prepared ICOFs (ICOF-1 and ICOF-2) show good thermal
stabilities and excellent resistance to hydrolysis, remaining nearly
intact when immersed in water or a basic solution for up to 2 days.
A scanning electron microscope (SEM) image supports the single-
crystalline morphology of both COFs. Powder X-ray diffraction
(PXRD) results also show multiple sharp peaks, which indicate the
34 Crystallization and Structural Linkages of COFs

orderliness of structures in the framework. However, the crystal


structures of ICOFs have not been characterized. These ICOFs
also have high BET surface areas (up to 1259 m2/g) and adsorb a
significant amount of H2 (up to 3.11 wt%, 77 K, 1 bar) and CH4 (up to
4.62 wt%, 273 K, 1 bar). The existence of permanently immobilized
ion centers in ICOFs enables the transportation of lithium ions with
RT lithium-ion conductivity of 3.05 × 10–5 S·L/cm and an average Li+
transference number of 0.80 ± 0.02.
R R

AcO OAc R R
EtCMo[NAr(t-bu)]3
LSi, CCl4, CHCl3
R = OAc
5 Å MS, 55 , overnight

R = OH
OH
LSi = Me R R
SiMe

R R

O O
O OB
OB O
O O

M+

O O
OB O
O OB
O O
M+ M+

B(OMe)3
MeNH or LiOH

+ M+
O M O
O OB
OB O
O O

ICOF-1 (M = [Me2NH2]+
ICOF-2 (M = Li+

O O
OB O
O OB
O O

Figure 2.8 Construction of spiroborate-linked COFs featuring anionic skeletons


with two counterions for high ionic conductivity.
b–O Linkages 35

More recently, Feng and coworkers have reported the first


3D anionic COFs based on flexible building blocks with different
counterions, where g-cyclodextrin (g-CD) molecules act as organic
struts that are covalently joined via spiroborate linkages. CD-COF-
Li as a 3D anionic COF coordinated with Li+ as a counterion was
synthesized by the condensation of g-CD and B(OMe)3 in the presence
of LiOH under microwave-assisted solvothermal conditions. The
PXRD of the COF shows a highly crystalline structure and a BET
surface area of 760 cm2/g [37]. The microwave-assisted solvothermal
synthesis ensures high production efficiency, good yields, and high
purity [11]. When the proton acceptor in the reaction was changed
to dimethylamine (DMA) or piperazine (PPZ), CD-COF-DMA and CD-
COF-PPZ with the corresponding cations were obtained. According
to structural analysis, all three CD COFs adopt an rra topology,
where the nodes of the net are substituted with building blocks
of the corresponding shape (Fig. 2.9). Owing to the high porosity,
flexible building blocks, and charge skeleton, CD COFs show great
potential in the fields of ion conduction and gas separation. The Li-
ion conductivity of CD-COF-Li is as high as 2.7 mS/cm at 30°C, and
this value is one of the highest Li-ion conductivities ever reported
for crystalline porous materials, including COFs and metal-organic
frameworks (MOFs).

OH
HO O O
O
OHHO OH O
O HO OH
O OH HO
HO O
OH HO
O O + B(OMe)3
OH HO
O OH HO O OH
O OH HO O
HO OHHO
O O
O OH
HO
-CD

CD COFs

Figure 2.9 Construction of CD COFs condensed by g-CD and B(OMe)3 with


LiOH, DMA, or PPz under microwave conditions. Reproduced with permission
from Ref. [37]. Copyright (2017), John Wiley and Sons.
36 Crystallization and Structural Linkages of COFs

2.1.4 Borazine-Linked COFs

- H6
3 NH2 BH 3 HB
N
BH (2.4)
N N
B
H

Borazine (H3B3N3H3), isolated by Stock and Pohland in 1926, is


often dubbed as “inorganic benzene” due to its similarities with
benzene: they are both liquid at RT, show equalized bond lengths
(1.40 Å for benzene and 1.44 Å for borazine, with the latter being
between B–N [single bond] at 1.51 and B=N [double bond] at
1.31 Å) and share a planar hexagonal structure. However, borazine
shows only a weakly aromatic character and displays a great
tendency to undergo hydrolysis to form boric acid and ammonia
in the presence of moisture. Another difference is the string polar
character of the B–N bonds resulting from the electron donation of
the nitrogen atoms to the electrophilic boron centers.
Previously, Sneddon et al. obtained boron nitride by pyrolysis
of polyborazylene polymer that had been obtained by thermal
polymerization of borazine B3N3H6 [38]. N- or B-substituted
borazines are also interesting compounds that may also be
polymerized. These polymers, via pyrolysis at high temperatures,
could lead to ceramic-like boron nitride and boron carbonitride
[38, 39]. N-substituted borazines are accessible by the thermolysis
of primary amine-borane complexes RNH2·BH3, usually prepared
by the reaction of lithium borohydride with primary amine salts
[40]. Manners and coworkers prepared borazines from NH3·BH3
or CH3NH2·BH3 at a low temperature (45°C) [41]. Kinetic studies
showed that the rate-determining step for both substrates is the
loss of the last H2 molecule, which have been demonstrated to be
fast only at high temperatures [42], leading to the dynamic covalent
chemistry (DCC) concept. As shown in Eq. 2.4, a 99% yield of tri-
N-phenylborazine could be obtained from PhNH2·BH3 only after 30
min. at 120°C.
Borazine has been mainly used for the fabrication of BN-based
ceramics or in organic optoelectronics [43]. However, to date, the
use of borazine as a building block for the preparation of porous
b–O Linkages 37

polymers remains scarce. EI-Kaderi and coworkers first reported the


targeted synthesis of the crystalline borazine-linked COF [BLP-2(H)]
(Fig. 2.10) [44] and investigated its structural aspects, porosity, and
performance in hydrogen storage. BLP-2(H) was prepared by the
thermal decomposition of 1,3,5-(p-aminophenyl)-benzene-borane
in a solvent mixture of mesitylene-toluene at 120°C/150 mTorr in
a sealed tube for 3 days, which gave a good yield of BLP-2(H) as a
white microcrystalline powder. The BET surface area and pore size of
this COF was 1178 m2/g and 6.4 Å, respectively. Using the Langmuir
model (P/P0 = 0.05–0.30), the surface area value was determined
to be 1564 m2/g. For comparison, the Langmuir surface area value
of BLP-2(H) is similar to the Connolly surface area predicted for
the eclipsed model (1840 m2/g) and is much lower than that of the
staggered model (3377 m2/g), which further supports the formation
of AA-eclipsed stacking. It can store up to 2.4 wt% of hydrogen at
77 K and 15 bar with isosteric heat of adsorption of 6.8 kJ/mol.

Figure 2.10 Construction of borazine-linked COF [BLP-2(H)] decomposed


by 1,3,5-(p-aminophenyl)-benzene-borane and SEM image of as-prepared
materials (inset). Reproduced with permission of Royal Society of Chemistry,
from Ref. [44], copyright (2012); permission conveyed through Copyright
Clearance Center, Inc.
38 Crystallization and Structural Linkages of COFs

2.2 Imine Linkages


O R N C R’
R NH2 +
H + H2 O (2.5)
‘R H

Nowadays, Schiff base chemistry or dynamic imine chemistry is


profusely utilized for the synthesis of COFs. Imine is a functional
group or chemical compound containing a carbon-nitrogen double
bond and is typically prepared by the condensation of primary
amines and aldehydes and less commonly ketones (Eq. 2.5). In terms
of mechanism, such reactions proceed via nucleophilic addition,
giving a hemiaminal –C(OH)(NHR)– intermediate, followed by the
elimination of water to give the imine (–N=C–) compound. The
equilibrium in this reaction usually favors the carbonyl compound
and amine, so azeotropic distillation or use of a dehydrating agent,
such as molecular sieves or magnesium sulfate, is required to push
the reaction in favor of imine formation. In recent years, several
reagents, such as tris(2,2,2-trifluoroethyl)borate, pyrrolidine, acetic
acid, and titanium ethoxide, have been shown to catalyze imine
formation. Especially, according to synthesis of COFs, aqueous acetic
acid as a catalyst is most useful.
Imine-based COFs constitute the largest amount of COFs based
on Schiff base chemistry. In general, they are quite stable in most
organic solvents and insensitive to water and acidic and basic
conditions. Imine-based COFs were discovered by a reversible
condensation of polytopic anilines with polybenzaldehydes or
polyketone, with the elimination of water by a catalyst of aqueous
acetic acid (Fig. 2.11). Generally, the crystallinity of the imine-based
COFs is lower compared with that of boronate ester-linked COFs
whereas the chemical stability in the presence of water, acids, and
bases is significantly enhanced.
The first imine-based COF was developed by Yaghi and coworkers
in 2009 [45]. They constructed an imine-based COF (COF-300),
which shows a 3D fivefold interpenetrating diamond-like skeleton.
COF-300 has a BET surface area of 1360 m2/g and a pore size of
7.8 Å, which provides better hydrolytic stability compared to
the boron-containing COFs. Furthermore, the broad range of
multifunctional amines and aldehydes gives a large number of
structural possibilities for imine-based COFs.
NH2
NH2

H2N NH2
H2N
H2N
NH2

H2N CHO CHO


HO OH

OHC CHO OHC CHO


OHC CHO OH
CHO OH
CHO

HO OH

N N
CHO OH
HO
CHO OHC CHO N N N N
HO OH
OH N
N
HO HO

OH

HO

BF-COF-1

N N

N O
N N

N N O O
N N
N NH
O
N O
N NH
N O H O
O HN
O O

O
COF-300
O
COF-320

BF-COF-2

3D-Py-COF

Figure 2.11 Construction of imine-linked COFs by co-condensation of tetratopic anilines—tetra-(4-anilyl)methane and


1,3,5,7-tetraaminoadamantane—with various ditopic, tetraaldehydes, and triangular aldehyde to form 3D networks.
Imine Linkages
39
40 Crystallization and Structural Linkages of COFs

Imine-based COF-320 was prepared via solvothermal


condensation of tetra(4-anilyl)-methane and 4,4¢-bisphneyldialdhyde
in 1,4-dioxane at 120°C. A ninefold interlaced diamond network was
formed, and the final crystalline material presented high porosity
with a Langmuir surface area of 2400 m2/g. The crystal structure
of the COF-320 solid was elucidated by single-crystal 3D electron
diffraction [46]. Recently, a pyridyl-functionalized version of
COF-320 was synthesized, namely LZU-301, in which the reversible
dynamic response upon guest accommodation and release was
uncovered. Thus, the design principle of 3D dynamic COFs was
claimed [47].
Wang and coworkers developed two new 3D pyrene-based COFs
(3D-Py-COF) by the condensation of tetra(p-aminophenyl)methane
and 1,3,6,8-tetrakis(4-formylphenyl)pyrene. This topology of a 3D
Py COF is pts, which was first reported in this work. To be built up by
pyrene in the 3D frameworks, this COF shows fluorescent properties
and can be used in the detection of explosives [48].
Novel 3D microporous functionalized COFs (BF-COF-1 and BF-
COF-2) were designed and synthesized by Yan and coworkers from
the reaction of a tetrahedral alkyl amine, 1,3,5,7-tetraaminoadaman-
tane, and 1,3,5-triformylbenzene (TFB) or triformylphloroglucinol
[49]. These BF COFs were used for Knoevenagel condensation reac-
tion with high conversion (BF-COF-1, 96%; BF-COF-2, 98%), highly
efficient size selectivity, and good recyclability. Adamantane as the
rigid building group is the key for forming targeted products. In con-
trast, the combination of tetra(p-aminophenyl) with TFB could not
produce a crystalline product.
A new strategy to enhance both the chemical stability and
crystallinity in 2D porphyrin COFs was proposed by Banerjee [50]
and Jiang [51] groups. In this method, the imine bond (–C=N–) in
the COF interior frame is protected by introducing –OH to the
Schiff base (–C=N) centers in COFs and creating an intramolecular
[–O–H···N=C–] hydrogen bond (Fig. 2.12) [50]. This hydrogen-
bonding interaction enhances the stability of imine bonds in the
presence of water and acid (3 N HCl). Otherwise, this hydrogen
bond in 2,3-dihydroxyterephthalaldehyde–5,10,15,20-tetrakis(4-
aminophenyl)-21H,23H-porphyrin (DhaTph) also could enhance the
crystallinity and porosity compared to the methoxy-substituted COF
DmaTph, in which this intramolecular hydrogen bond is absent.
NH N OMe NH N H O
N N
NH N NH N
N HN N HN
N N
MeO N HN CHO CHO O H N HN
MeO HO
H 2N NH2
N
OMe OH H
N CHO CHO N
N H O
OMe N Dma H Dha O N
N N H
OMe H O
MeO N O
H O
N MeO N O
H H
N H 2N NH2 N N

Tph hydrogen bonding

NH N OMe NH N H O
N NH N N NH N
N
N HN N HN N
MeO N HN O H N HN

Figure 2.12 Synthesis of DmaTph and DhaTph by the condensation of square planar Tph building unit and linear Dma and Dha building
units.
Imine Linkages
41
42 Crystallization and Structural Linkages of COFs

The formation of a crystalline structure via relatively weak


interactions between planar layers has been well defined and
designed in crystal engineering. Jiang and coworkers prepared
a series of 2D COFs locked with intralayer hydrogen-bonding
interactions (Fig. 2.13) [51]. The hydrogen-bonding interaction
sites were located on the edge units of the imine-linked tetragonal
porphyrin COFs, and the contents of the hydrogen-bonding sites in
the COFs were synthetically regulated by using a three-component
condensation system. The intralayer hydrogen-bonding interactions
suppress the torsion of the edge units and lock the tetragonal sheets
in a planar conformation. This planarization enhances the interlayer
interactions and triggers extended p-cloud delocalization over 2D
sheets. All pores in the COFs are 2.5 nm in size. These COFs have
an AA stacking layered structure. As a result, COFs with layered
2D sheets amplify these effects and strongly affect the physical
properties of the materials, including improving their crystallinity,
enhancing their porosity, increasing their light-harvesting capability,
and reducing their bandgap.
In addition to high stability and crystallinity of COFs by hydrogen-
bonding interaction, self-complementary p electronic forces provide
new opportunities for enhancing the crystallinity of COFs. Jiang and
Nagai groups reported a synthetic method to control the crystallinity
and porosity of COFs by managing interlayer interactions based on
self-complementary p electronic forces (Fig. 2.14).
A three-component condensation reaction of CuP and
1,4-diformylbenzene with 2,3,5,6-tetrafluoroterephthalaldehyde can
control the crystal structure of the CuP-Ph COF. Fluoro-substituted
aromatic units in different ratios were integrated into the edge units,
which will induce the self-complementary p electronic interactions
in the COFs.
This interaction increases the crystallinity and porosity of COFs
by maximizing the total crystal stacking energy and minimizing
the unit cell size. This work provides a new pathway to improve
the crystals of COFs by controlling the interlayer interactions [52].
The same authors have also developed a nice strategy to soften
the polarization influence of the C=N bond on the destabilization
of the layered structure in hexagonal 2D COFs based on imine
reactions [53]. In imine-linked COFs, the C=N bond is polarized to
yield partially positively charged carbon and negatively charged
H 2N NH2

N
N M N
N

H2 N NH2
MP
(M = Cu and Ni)

CHO CHO CHO


R HO CHO
HO
R OH
CHO CHO CHO OH
CHO
TA DHTA
TA (R = H) (100-x mol%) (x mol%: x = 25, 50, 75)
DMTA (R = OMe)

Non-H-Bonded Partially H-Bonded Fully H-Bonded

O H N N
R N N N N N M
N N
N N M N N N M
N M N N N
M N N N M N N N
N N H O
N N R N N N
H
O
N
N N H
H N O
N R N O H
H O
O O
R R O 2.5 nm O H 2.5 nm H
H N O N
R N O N H
H N
N N hydrogen bonding

O H N N
R N N N N N M
N M N M N N
N N N N N N
N N N N M N
M N M N N N H O
N N R N N

MP-Ph COF (R = H) MP-DHPhx COFs MP-DHPh COF


MP-DMPh COF (R = OMe) (anti-isomer x = 50)
Imine Linkages

Figure 2.13 Construction of 2D porphyrin COFs with designable content of hydrogen-bonding structures.
43
H 2N NH2
44

N
N Cu N
N

H2 N NH2
CuP

CHO CHO CHO CHO


F F F F

F F F F
CHO CHO CHO CHO
TA TA TFTA TFTA
(100-x mol%) ( x = 25, 50, 75)

F F N N
N N N N N Cu
N N
N N Cu N N N Cu
N Cu N N N
Cu N N N Cu N N N
Crystallization and Structural Linkages of COFs

N N F F
N N N N

N
N N
N F F
N N F F
F F F F
2.5 nm F F 2.5 nm 2.5 nm
F F
F F N
N F F N
N
N N

F F N N
N N N N N Cu
N Cu N Cu N N
N N N N N N
N N N N Cu N
Cu N Cu N N F F
N
N N N N

CuP-Ph COF CuP-TFPhx COFs CuP-TFPh COF


(anti-isomer x = 50)

Figure 2.14 Construction of 2D COFs integrated with self-complementary p electronic forces (CuP-TFPhX, where X = 25, 50, and 75 mol%)
and CuP-Ph and CuP-TFPh controls.
Imine Linkages 45

NH2

OHC
MeO
+
OMe
CHO

H 2N NH2 DMTA

TPBA

MeO
N

N
OMe

N OMe

MeO N
MeO OMe

N MeO -
N
HC - CH
-
N
-
OMe
Resonance effect of oxygen lone pairs
N
softene the interlayed chage repulsion

MeO OMe
N OMe

MeO N

MeO
N

N
OMe

TPB-DMTA COF

Figure 2.15 Synthesis of a TPB-DMTP COF via the condensation of DMTA and
TPBA. Inset: the structure of the edge units of the COF and the resonance effect
of oxygen lone pairs that weaken the polarization of the C=N bonds and soften
the interlayer repulsion in the COF.

nitrogen. Thus, in a hexagonal 2D COF, each macrocycle consists of


12 polarized C=N segments; the aggregation of a large number of
charge groups causes electrostatic repulsion and destabilizes the
layered structure, as predicted theoretically [54]. Therefore, in this
case, crystallinity and stability were enhanced by incorporating
46 Crystallization and Structural Linkages of COFs

methoxy groups in the pore walls of COFs. Introducing the two


electron-donating methoxy groups to each phenyl edge delocalizes
the two lone pairs from the oxygen atoms over the central phenyl
ring, which reinforces the interlayer interactions and so stabilizes
the COF and aids in its crystallization (Fig. 2.15).
On the other hand, the utilization of macrocyclic host derivatives
with a shallow rigid cavity as triformylcyclotrianisylene, a derivative
of the well-known cyclotrianisylene (CTV), seems to stabilize the
perfect eclipsed stack model rather than COFs based on planar
motifs. The columnar stacking of crown CTV motifs avoids sliding
between layers in CTV COFs (Fig. 2.16) [55].
(a) OMe MeO

MeO OMe
N N
N OMe n
MeO
N

N n
OHC OMe MeO n N OMe

90
+ H 2N NH2
EtOH/3M CH3CO2H
n
MeO CHO OMe MeO
OHC OMe MeO N OMe
N

n N
OMe MeO N n

n = 1 : CTV-COF 1
n = 2 : CTV-COF-2
MeO OMe
N N
OMe MeO
(b) n

Figure 2.16 (a) Synthetic route of CTV COFs and (b) stick view of CTV COF-1. All
hydrogen atoms are omitted for clarity. Different layers are shown as different
colors. Adapted with permission of Royal Society of Chemistry, from Ref. [55],
copyright (2014); permission conveyed through Copyright Clearance Center, Inc.

Regarding the size and shape of the pores in 2D COFs, it is


known that they can be well tailored by changing the structures
Imine Linkages 47

of monomers [16, 22, 56, 57]. In general, COFs with hexagonal and
tetragonal topologies are by far the most extensively investigated.
However, modifications of these topologies can introduce structural
changes in the COFs. Hence, COFs with a trigonal topology have
been rationally designed by using C6-symmetrical vertices as
hexaphenylbenzene (HPB) and hexabenzocoronene, thus providing
small pore sizes and high p column densities [58].
To achieve remarkable chemical stability, Banerjee and coworkers
explored a new method to enhance the chemical stability of imine
COFs by the transformation of combined reversible Schiff base
bonds to irreversible b-ketoenanime bonds (Fig. 2.17) [59]. They
synthesized the COFs TpPa-1 and TpPa-2 by the Schiff base reactions
of 1,3,5-triformylphloroglucinol (Tp) with p-phenylenediamine
(Pa-1) and 2,5-dimenthyl-p-phenylenediamine (Pa-2), respectively.
The expected enol-imine (OH) form underwent irreversible proton
tautomerization into irreversible ketoenamine form, which confers
outstanding stability to boiling water, aqueous acid (9 N HCl), and
base (9 N NaOH). This linkage has also been synthesized by a simple
solvent-free RT mechanochemical synthetic route [60]. The amide-
linked COFs show improved chemical stability relative to their imine
progenitors but a low degree of crystallinity. Recently, Yaghi and
coworkers developed a method for the chemical conversion of an
imine bond to an amine bond inside COFs to improve both chemical
stability (including base/acid stability) and crystallinity (Fig. 2.18).
They converted two layered imine COFs to amide-linked COFs (4PE-
1P-COF 1; 1¢ and 4PE-1P-COF 1¢; 2¢) without loss of their crystallinity
or topology. Fourier-transform infrared spectroscopy and 13C CP-
MAS NMR spectra indicated the successful conversion of COF-1 and
COF-2 to COF-1¢ and COF-2¢. The high crystal structures of COF-1¢
and COF-2¢ were examined by X-ray diffraction analysis. The BET
surface areas of COF-1¢ and COF-2¢ were compared to those of COF-1
and COF-2 to examine the increase in framework mass and decrease
in pore volumes before and after the reduction. This method offers a
new pathway to overcome the usual crystallization problem in COF
chemistry [61].
CHO
HO OH

OHC CHO
48
OH

NH2 H2N NH2


R N
NH2 NH2
R N N
Reversible Reversible Reversible Reversible
R Schiff base reaction Schiff base reaction Schiff base reaction Schiff base reaction
NH2
R NH2
NH2 NH2

HO OH
R HO OH
N HO OH
HO OH R N
R1 OH N N
N OH OH
OH N OH N
OH N
OH N N OH N
N
R1 N N N
R
R2 HO OH N HO OH N OH
HO OH R2 N OH N N
R
N HO OH N
N R
R1 N HO OH Enol form
HO OH N
R
N
R2 N N Enol form HO OH N
Enol form Enol form
N R2 HO OH
HO OH
HO OH R
R1 N OH N N
N
N R
OH N HO OH N OH
HO OH N
OH N R2 HO OH N N
R2 HO OH N
N R N OH
R1 N OH OH N N OH
OH N
N R OH
R1 OH R HO OH
HO OH
N N
R
HO OH

Irreversible
Irreversible tautomerism
tautomerism Irreversible Irreversible
tautomerism tautomerism
Crystallization and Structural Linkages of COFs

O O
O O O O H
O O R H N
R1 H H
N N N O HN
N R O
O
O HN O HN N N
O HN R1 O N N
N H
H R2 O O NH O
O O R2 NH O N O N N
H R
HN O O N HN
R H Kato form
R1 HN O O O O NH
O O
R
HN HN O O NH
R2 HN Kato form
NH Kato form R2 R NH
O O
Kato form O O H
O O O HN N
R1
NH
NH R N O
O HN R2 O O NH O HN O O
R2 H R H N N
O O NH N
N NH O
R1 NH O
O O O
O R N
N H H
H O HN N
R1 O O
O O
R O
O R
TpPa-1 (R1 = H, R2 = H) N N
R H H N
TpPa-2 (R1 = Me, R2 = Me) O O
TpPa-3 (R1 = H, R2 = NO2) H2N NH2
H2N NH2 N

(R = H)
N NH2
(R = Me) H2N N
(R = OMe)
(R = NO2) O
NH2

H2N
O

H2N NH2

Figure 2.17 Synthesis of secondary amine-linked COFs through tautomerism by linking Tp with ditopic or tritopic aniline-based COFs.
Hydrazone Linkages 49

N N
N O

H N NH
N H
N O
O O
HN

N O NH
N
N
NaClO2/2-methyl-2-butane O HN
N
Dioxane, AcOH O NH
NH O
r.t., 2 days
HN O
N

HN O

NH
N
N N O O
O
H N
HN N H

TPB-TP-COF 1
(COF 1)

O NH
N NH O
N

NH
N HN
O O
N
H O H O
N H O N
N N N N N
N N O H N O H
N O H
N
O NH O
NH NH NH
N N N O
N O

O O HN
HN HN O
N N HN O
N N

NaClO2/2-methyl-2-butane
Dioxane, AcOH
r.t., 2 days
NH O NH
O NH
N N NH O
N O
N

O O
HN HN HN
N O HN O
N N N
H O
H O N H O
N N N N
N N N O H N
N O H
N N O H

NH O
NH
N N O

O HN
N HN O
N

4PE-1P-COF 1
(COF 2)

Figure 2.18 Preparation of amide-linked COFs via postoxidation of imine-


linked COFs, which bypass the difficulty of crystallization for irreversible amide
bonds.

2.3 Hydrazone Linkages


H2N
O N
+ H2N NH2
R1 R2 R1 R2
+ H2O (2.6)
hydrazine

Hydrazones are valuable and versatile building blocks in synthetic


chemistry [62]. Owing to their similarities to carbonyl compounds,
hydrazones are a class of organic compounds with the structure
R1R2C=NHN2. They are related to ketones and aldehydes by the
replacement of the oxygen with the NNH2 functional group. They are
formed usually by the action of hydrazine on ketones or aldehydes,
as shown in Eq. 2.6.
50 Crystallization and Structural Linkages of COFs

Like the reaction of imine-linked COFs, the reaction of


benzaldehyde and hydrazide groups could yield hydrazine-linked
COFs (Fig. 2.19). Hydrazone-linked COFs represent good chemical
stability relative to imine-linked ones because of the hydrogen-
bonding interaction between the oxygen atom in the alkoxyl
chain and the hydrogens in amide (–CONH–) units. However, the
functional monomers available for the condensation of hydrazone-
based COFs are limited. Recently, the first hydrazone-based COFs
were developed by the condensation reaction of a difunctional
acylhydrazide (2,5-diethoxyterephthalohydraide) and trifunctional
aldehydes (1,3,5-triformylbenzene or 1,3,5-tris(4-formylphenyl)
benzene) to provide two crystalline mesoporous COFs, COF-42 and
COF-43 [63].
The exfoliation of COF-43 in various solvents was done by Dichtel
and coworkers to create few-layered 2D polymers. The obtained
exfoliated COF-43 showed by PXRD an apparent loss of crystallinity
but no apparent changes in its covalent linkages and high aspect
ratios [64].
Stegbauer et al. reported a hydrazone-based COF—1,3,5-
tris(4-formylphenyl)triazine (TFTP) COF—capable of visible-
light-driven hydrogen generation with Pt as a proton reduction
catalyst. The COF, which was constructed by TFPT and 2,5-diethoxy-
terephthalohydrazide building blocks, shows a layered structure
with a honeycomb-type lattice featuring mesopores of 3.8 nm and
a BET surface area of 1603 m2/g. On irradiation by visible light, the
Pt-doped COF continuously produces hydrogen from water, without
signs of degradation. This new application of COFs in photocatalysis
opens new pathways for heterogeneous catalysts [65].
A thioether-based hydrazone-linked COF (COF-LZU8) was
designed rationally by Ding and coworkers, and they used it for
highly sensitive detection and effective removal of Hg2+. Regarding
the high stability of the hydrazone linkages, together with a dense
distribution of the thioether groups and the straight channels in
COF-LZU8, the recycling of COF-LZU8 achieved the simultaneous
detection and removal of the toxic Hg2+ [66].
O
O HN NH2

H2N NH O
O

methylene
/dioxane 120
/acetic acid (aq) 72 h

CHO CHO CHO

OHC CHO
N N
N
OHC CHO OHC CHO

O
O HN N
N O
N NH O HN
O

N NH
HN O N N
O O O O N
O HN N O HN N N

N NH O N N NH O
O N O
N O N N O
O O N HN HN
NH HN
N O O
N NH N NH
HN O N HN O N
O O O O

O O N
N O NH O O O O N
HN N NH NH N
N N

O O
O N N N
NH O HN HN N HN
O N O O N O O
N
N NH O
O
O O O O
N O NH N O NH
HN N HN N
O O O O

NH NH
COF-42 O N O N N
O O N
O HN N O HN N N
N NH O N N NH O
O N O
N

COF-43 TFPT COF


Hydrazone Linkages

Figure 2.19 Synthesis of hydrazone-linked COFs (COF-42, COF-43, and TFPT COF).
51
52 Crystallization and Structural Linkages of COFs

2.4 Azine Linkages


R1
O H N
2
R1 H
+ H2N NH2
N H
+ 2 H2O (2.7)
hydrazine R1

Azines (–C=N–N=C–) are formed by the condensation of hydrazine


with aldehydes, as shown in Eq. 2.7. Azines are 2,3-diazosubstituted
analogues of 1,3-dienes. One therefore might expect that they would
show a noticeable p-electron conjugation between C=N bonds via
the N–N linkage.
Usually, azine-linked COFs have small micropores and are
stable in water, acid, and base (Fig. 2.20). Nagai et al. constructed
a highly crystalline 2D azine-linked Py COF (Py-azine COF) by the
condensation of hydrazine with 1,3,6,8-tetrakis(4-formylphenyl)
pyrene under solvothermal conditions [67]. The azines inside of a
Py-azine COF serve as Lewis basic sites to bind the guest and enable
the selective detection of 2,4,6-trinitrophenol, which exhibits a
potential application in chemosensing systems. Li et al. prepared
an azine-linked COF, ACOF-1, by the condensation of hydrazine
hydrate and 1,3,5-triformylbenzene. The high surface area
(1318 m2/g) and small pore size make it useful as a gas storage
medium for CO2 (177 mg/g at 273 K and 1 bar), H2 (9.9 mg/g at
273 K and 1 bar), and CH4 (11.5 mg/g at 273 K and 1 bar) [68, 69].
Smaldone and coworkers developed a novel azine-linked HPB-based
COF, HEX-COF-1, which shows excellent sorption capability for CO2
(20 wt%) and methane (2.3 wt%) at 273 K and 1 atm [70]. Lotsch
and coworkers reported a tunable water- and photostable azine
COF by hydrazine and triphenylarene aldehydes for visible light–
induced hydrogen generation [71]. Zhang et al. reported an azo-
containing COF by introducing a novel azobenzene monomer via the
borate ester formation reaction of azobenzene-4,4¢-diboronic acid
and HHTP. The azo COF has a hexagonal skeleton and permanent
porosity. The azo unit in azo COFs endow the COF material with
photoisomerization properties [72].
NH2-NH2

CHO
CHO OHC CHO
OHC CHO

X Y
Z OHC CHO
OHC CHO OHC CHO
CHO

X Y
Z
N
N
N
Y N N
N
N
X Z N
X N
N N
Z Y N N
N
N
N N
N
N N N
N
N
N N
N N N
Y Z N N
N N
N N
X
Z X
N
N Y
N
N N
N
N N
Z
X Y

N0-COF : X= Y= Z = H
N1-COF : X = Y = C-H, Z = H
N2-COF : X = C-H, Y = Z = N
N3-COF : X = Y = Z = N
Azine Linkages

Figure 2.20 Synthesis of azine-linked COFs.


53
54 Crystallization and Structural Linkages of COFs

2.5 Squaraine Linkages


O O O
2 H2N R R
HN NH
R
+ 2 H2O (2.8)
HO OH
O
squaric acid

Squaric acid and amines are used as building blocks for the synthesis
of squaraine-based COFs, which allow strong p conjugation from
donor and acceptor interaction into a squaraine bond (Eq. 2.8).
Nagai et al. reported a new reaction based on squaraine for the
synthesis of a new type of building block to construct a crystalline
2D conjugated COF (CuP-Sq COF) with a tetragonal mesoporous
skeleton (Fig. 2.21) [73]. The CuP-Sq COF possesses strong hydrogen
bond capabilities that are strengthened by the delocalization of the
lone pair of electrons in nitrogen into a four-membered ring to form
zwitterionic resonance structures and shows strong p conjugation,
as shown in Eq. 2.8. This COF shows a BET surface area of 539 m2/g
with a pore size of 2.1 nm. This research expands the types of COFs.

N N
NH O
Cu
N N
N N
O Cu
H 2N HN N
NH2 N

N O O O NH
N Cu N + o -PhCl2/n-BuOH
O
HO OH 85 , 7 days HN CuP-SQ COF NH
N O
Squaric acid (SQ)
HN O
H 2N NH2

N N
NH O
Cu
N N
N N
O Cu
HN N
N

Figure 2.21 Construction of squaraine-based COFs (CuP-SQ COF).

2.6 Imide Linkages


O O
H2N R + O R N + H2O (2.9)
O O
anhydride imide
Phenazine Linkages 55

In organic chemistry, an imide bond is a functional group consisting


of two acyl groups bound to nitrogen and is prepared by the
condensation reaction of anhydrides and primary amines (Eq. 2.9).
This imide-bonded compound is structurally related to acid
anhydride. However, imide bonds are thermally reversible and more
resistant toward hydrolysis.
Polyimides are formed by condensation of primary amines and
anhydrides (Fig. 2.22). Tan and coworkers prepared 3D porous
crystalline polyimide COFs (PI COFs) by choosing tetrahedral
building blocks of different sizes [74]. These PI COFs show high
thermal stability (more than 450°C) and large surface areas (more
than 1000 m2/g), as well as narrow pore sizes (13 Å for PI-COF-4
and 10 Å for PI-COF-5). They were the first COFs to be employed
in controlled drug delivery. The Tan group also prepared a series
of polyimide-based COFs with different pore sizes by changing the
length of triamines using the imidization reaction [75]. Among
them, PI-COF-3, synthesized by the condensation of pyromellitic
dianhydride and the extended triamine, 1,3,5-tris[4-amino(1,1-
biphenyl-4-yl)]benzene, has 5.3 nm wide pores and was loaded with
rhodamine B, a dye probe for biological applications.

2.7 Phenazine Linkages


NH2 O N
+ + 2 H2 O (2.10)
NH2 O N

Phenazines as one of azo-fused structures were pursued in making


conductive materials with the formula (C6H4)2N2 that have many
potential applications in electrochemical devices [76]. The DCC
reaction is led by the oxidation of dihydrophenazine, which is
prepared by heading pyrocatechin with o-phenylenediamine
(Eq. 2.10). However, the great challenge is how to make them highly
crystalline. Jiang and coworkers prepared a phenazine-linked COF
(CS COF, Fig. 2.23) via a ring-fusing reaction between quinone and
amine derivatives catalyzed by acidic forms. They used C3-symmetric
triphenylene hexamine and C2-symmetric tert-butylpyrene tetraone
as building blocks to prepare the crystalline phenazine–linked
CS COF under solvothermal conditions. The polygon skeletons of
the COF are highly p conjugated and chemically stable in various
solvents [77].
O O

O O
56
O O

mestylene / NMP / isoquinoline 200 for 5 days


NH2

NH2
NH2
NH2
NH2
H2 N
X NH2 H 2N
H2N
H 2N NH2
NH2

X: N and Ph H 2N NH2
O N O N
O O

X O N O O O
O O O O N O
O O
N N N N N
O O N
O N
O O O O O
N N O N O O
O N O
O O N
X X N
O O O
O N
O O
N O
O N O O N O O
O O N
N N O
O O
O O O O O O PI-COF-4 PI-COF-5
N N N N
O O

X
O O
N N
O O
O O
N N
Crystallization and Structural Linkages of COFs

O O
X

O N O O N O

PI-COF-1 (X = N, pore size = 3.3 nm) O O O O


N N
PI-COF-2 (X = Ph, pore size = 3.7 nm)

O O
N N
O O
O O
N N
O O

PI-COF-3 (pore size = 3.3 nm)

Figure 2.22 Development of 2D or 3D imide-linked COFs (PI-COF-1, PI-COF-2, PI-COF-3, PI-COF-4, and PI-COF-5) by the condensation of
benzene-1,2,4,5-tetracarboxylic dianhydriodes.
Triazine Linkages 57

N N
H 2N NH2 N N
N N

N N
H 2N NH2

H 2N NH2
TPHA Ehylene glycol / 3M AcOH aq, N N N N
120 , 3 days 1.6 nm
+
CS COF
O O N N N N

O N N
O
PT N N
N N

N N

Figure 2.23 Synthesis of a phenazine-linked COF.

2.8 Triazine Linkages

3 CN N N + 2 H2O (2.11)
N

Triazine as 1,3,5-trizaine is a class of nitrogen-containing


heterocycles and is synthesized by the condensation of aromatic
nitriles at high temperatures in molten salts or at RT in the presence
of a strong acid catalyst (Eq. 2.11). The crystalline, porous covalent
triazine frameworks (CTFs) are prepared by condensational growth
of aromatic nitriles in the presence of trifluoromethylsulfonic acid or
zinc chloride, as shown in Fig. 2.24.
The CTF was initially reported by Thomas and coworkers in
2008 and is prepared via the dynamic trimerization of aromatic
nitriles (1,3,5-triazines) using the ZnCl2 ionothermal method [78].
CTFs gave high surface areas (up to 2475 m2/g), low densities, and
outstanding thermal/chemical stability. However, CTFs show short-
range crystalline order and limited pore size distributions in some
cases [79].
58
NC

NC CN
O
NC
CN
CN O
CN
O CN

NC
ZnCl2 ZnCl2 ZnCl2
400 400 CF3SO3H
400
CHCl3
0

N N
N N N N
N
N N N
N
N N N O
N N N
N N N N N
N N N N N N O
Crystallization and Structural Linkages of COFs

O
N N O
O

N N N N O
N N N
CTF-0 N N N
N N N
N N
N N
N N N O
N
N
N N O
CTF-1
O

CTF-2

CTF-IP-10

Figure 2.24 Construction of crystalline covalent triazine frameworks (CTFs) by self-condensation of aromatic nitriles.
Multihetero Linkages in One COF Skeleton 59

Most CTFs were synthesized in molten ZnCl2 at high


temperatures (400°C–700°C). These conditions for CTF synthesis
provided less crystalline CTFs of limited functionality and only
from 1,4-dicyanobenzene and 2,6-dicyanobenzene [80]. Then,
Cooper et al. reported a microwave-assisted, acid-catalyzed method
to synthesize triazine networks with fewer discolored impurities
(P1M-P6M). Their structural and synthesized tunability developed
their specific applications, such gas storage, catalyst support, and
organic dye separation [81].
Thomas at al. reported the synthesis of organic CTFs (CTF-0)
by the trimerization of 1,3,5-trisyanobenzene in molten ZnCl2 [82].
This COF can present high CO2 uptakes and good catalytic activity
of CO2 cycloaddition. The high BET surface area and robust porous
structures of the organic framework were explored for highly
selective gas capture.
Ghosh and coworkers found a strategy to obtain a porous
CTF (CTF-IP-10) with bimodal functionality by acid-catalyzed RT
reaction of tricyanomonomer (PCN-M1) in CHCl3 [83]. The obtained
CTF-IP-10 consists of both an electron-deficient central triazine core
and electron-rich aromatic building blocks.

2.9 Multihetero Linkages in One COF Skeleton


To improve the functionality and stability of the COF structure with
b–O linkages, COFs connected by two or more types of covalent
bonds on orthogonal reaction were designed and synthesized.
First Zhao and group reported a 2D COF (NTU-COF-1, Fig. 2.25)
by the formation of imine group and boroxine (B3O3) rings, involving
the use of two building blocks: 4-formyl-phenylboronic acid (FPBA)
and 1,3,5-tris(4-aminophenyl)benzene. They also constructed a
ternary COF (NTU-COF-2, Fig. 2.25) via the formation of two types
of covalent bonds. The NTU-COF-2 has a high BET surface area
(1619 m2/g) and mesopores with a diameter of 2.6 nm. The design
and construction of COFs by this orthogonal reaction strategy is a
promising approach in rational COF development [84].
NH2 NH2
HO OH
60
CHO CHO

1 + 3 1 + 3 + 1
HO OH
B B
HO OH HO OH
HO OH
H 2N NH2 H 2N NH2

b) - n H2O
a) - n H2O

NH2
NH2
B O
O

O N
B O N B
O B O
B O
O
B O N
N

N
Crystallization and Structural Linkages of COFs

B O
B O N O
O B
NTU-COF-1 B O
O
B
NTU-COF-2 O

O
N
N B O

N
B O N
O B N
B O B O
O

O N
B
O

O
B O

Figure 2.25 Development of (a) NTU-COF-2 and (b) NTU-COF-2 by two types of covalent bonds.
Perspectives and Challenges 61

Qiu’s group prepared two extended 3D COFs (DL-COF-1 and


DL-COF-2) with dual linkages, imine groups and B3O3 boroxine rings
(Fig. 2.26), involving the utilization of two 1,3,5,7-tetraaminoada-
mantane and FPBA or 2-fluoro-4-formylphenyl boronic acid build-
ing blocks [85]. The obtained DL-COF-1 and DL-COF-2 showed high
BET surface areas (2259 m2/g and 2071 m2/g, respectively).
NH2

H 2N
NH2
H 2N

CHO CHO

F
B B
HO OH HO OH

(a) (b)
N N

F
N O B N O B F
B O B O
O B O B

N N

Figure 2.26 Synthesis of 3D COFs of (a) DL-COF-1 and (b) DL-COF-2 by two
types of covalent bonds from orthogonal reactions.

2.10 Perspectives and Challenges


The past decade has witnessed a boom in COF chemistry. Their
unique features, such as flexible molecular structure design,
permanent porosity, low density, modest to high crystallinity, and
controllable pore size (from ultraporous to mesoporous) and the
diversity of the available building blocks, easily make them targets
for structural and functional design and promise their great potential
as materials for application in various areas. However, there are still
a lot of challenges in the design and synthesis of COFs.
62 Crystallization and Structural Linkages of COFs

First, compared to the highly ordered structure of MOF materials,


further investigation on thermodynamically controlled solvothermal
reactions for the preparation of COF materials is quite important.
Such studies provide a general pathway for the preparation of
high-crystallinity COFs. Moreover, control over the defect sites,
morphology, and stacking mode of COF structures is also important
and effective methods to construct precise crystal COF structures
are necessary. At present, no direct analytical tools can be used to
visualize and analyze the stacking and topological structures of
COFs. Thus, the creation of single-crystal COFs is an unattainable
target.
Meanwhile, in the existing approaches, the methods to increase
the surface areas and porosity for COF materials are a big bottleneck.
The surface areas and porosity are the most important parameters
for COF materials for gas adsorption and gas separation applications.
How to design COF structures with improved surface area and these
properties is the key to realizing the potential application of COFs in
these fields. For 2D COFs, spatially enlarging the interlayered spaces
would increase the surface areas. For 3D COFs, the construction of
highly crystalline frameworks would lead to very high BET surface
areas, which can reach the theoretical SBET of over 6000 m2/g.

References
1. Korich, A. L., Inovine, P. M. (2010). Dalton Trans., 39, 1423–1431.
2. Scheleyer, P. V., Jiao, H. J., Hommes, N. J. R. V., Malkin, V. G., Malkina, O. L.
(1997). J. Am. Chem. Soc., 119, 12669–1779.
3. Brock, C. P., Minton, R. P., Niedenzu, K. (1987). Acta Crystallogr., C43,
1257–1268.
4. Synder, H. R., Kuck, J. A., Johnson, J. R. (1938). J. Am. Chem. Soc., 60,
105–111.
5. Morgan, A. B., Jurs, J. L., Tour, J. M. (2000). J. Appl. Polym. Sci., 76, 1257–
1268.
6. Li, Y., Ding, J., Day, M., Tao, T., Lu, J., D’iorio, M. (2003). Chem. Mater., 15,
4936–4934.
7. Tokunaga, Y., Ueno, H., Shimomura, Y., Seo, T. (2002). Heterocycles, 57,
787–790.
References 63

8. Côté, A. P., Benin, A. I., Ockwig, N. W., O’Keefe, M., Matzager, A. J., Yaghi,
O. M. (2005). Science, 310, 1166–1170.
9. EI-Kaderi, H. M., Hunt, J. R., Mendoza-Cortez, A. P., Côté, A. P., Taylor, R.
M., O’Keefe, M., Yaghi, O. M. (2007). Science, 316, 268–272.
10. Wan, S., Guo, J., Kim, J., Ihee, H., Jiang, D. (2009). Angew. Chem. Int. Ed.,
48, 5439–5442.
11. Campbell, N. L., Clowes, R., Ritchie, L. K., Cooper, A. I. (2009). Chem.
Mater., 21, 204–206.
12. Côté, A. P., Benin, A. I., Ockwig, N. W., O’Keefe, M., Matzager, A. J., Yaghi,
O. M. (2005). Science, 310, 1166–1170.
13. EI-Kaderi, H. M., Hunt, J. R., Mendoza-Cortez, A. P., Côté, A. P., Taylor, R.
M., O’Keefe, M., Yaghi, O. M. (2007). Science, 316, 268–272.
14. Côté, A. P., EI-Kaderi, H. M., Furukawa, H., Hunt, J. R., Yaghi, O. M. (2007).
J. Am. Chem. Soc., 129, 12914–12915.
15. Wan, S., Guo, J., Kim, J., Ihee, H., Jiang, D. (2009). Angew. Chem. Int. Ed.,
48, 5439–5442.
16. Tilford, R. W., Mugavero, S. J., Pellechia, P. J., Lavigne, J. J. (2008). Adv.
Mater., 20, 2741–2746.
17. Lanni, L. M., Tilford, R. W., Bharathy, M., Lavigne, J. J. (2007). J. Am.
Chem. Soc., 133, 13975–13983.
18. Dogru, M., Sonnauer, A., Gavryushin, A., Knochel, P., Bein, T. A. (2011).
Chem. Commun., 133, 13975–13983.
19. Spitler, E. L., Colson, J. W., Uribe-Romo, F. J., Woll, A. R., Giovino, M. R.,
Saldivar, A., Ditchel, W. R. (2012). Angew. Chem. Int. Ed., 51, 2623–2627.
20. Colson, J. W., Mann, J. A., DeBlase, C. R., Dichtel, W. R. (2015). J. Polym.
Sci. Polym. Chem., 53, 378–384.
21. Spitler, E. L., Dichtel, W. R. (2010). Nat. Chem., 2, 672–677.
22. Spitler, E. L., Giovino, M. R., White, S. L., Dichtel, W. R. (2011). Chem. Sci.,
2, 1588–1593.
23. Dogru, M., Sonnauer, A., Zimdars, S., Doblinger, M., Knochel, P., Bein, T.
(2013). Crystengcomm, 15, 1500–1502.
24. Calik, M., Sick, T., Dogru, M., Döblinger, M., Datz, S., VUdde, H., HArtschuh,
A., Auras, F., Bein, T. (2016). J. Am. Chem. Soc., 138, 1234–1239.
25. Böeseken, J. (1949). Adv. Carbohydrate Chem., 4, 189–210.
26. Hermans, P. H. (1925). Z. Anorg. Allgem. Chem., 142, 83–110.
27. (a) Xiao, L., Ling, Y., Alsbaiee, A., Li, D., Helbling, D. E., Dichtel, W. R.
(2017). J. Am. Chem. Soc., 139, 7689–7692; (b) Alsbaiee, A., Smith, B.
64 Crystallization and Structural Linkages of COFs

J., Ling, Y., Li, Y., Helbling, D. E., Dichtel, W. R. (2016). Nature, 529, 190–
194.
28. Smsldone, R. A., Forgan, R. S., Furukawa, H., Gassensmith, J. J., Slawin, A.
M., Yaghi, O. M., Stoddart, J. F. (2010). Angew. Chem. Int. Ed., 49, 8630–
8634.
29. Brown, H. C., Rangaishevi, M. V. (1988). J. Organimet. Chem., 358, 15–
30.
30. Van Duin, M., Peters, J. A., KIeboom, A. P. G., Van Bekkum, H. (1985).
Tetrahedron, 41, 3411–3421.
31. Van Duin, M., Peters, J. A., KIeboom, A. P. G., Van Bekkum, H. (1984).
Tetrahedron, 40, 2901–2911.
32. Yoshino, K., Kotaka, M., Okamoto, M., Kakihara, H. (1979). Bull. Chem.
Soc. Jpn., 52, 3005–3009.
33. Voisin, E., Maris, T., Wuest, J. D. (2008). Cryst. Growth Des., 8, 308–318.
34. Danjo, H., Hirata, K., Yoshiga, S., Azumaya, I., Yamaguchi, K. (2009). J.
Am. Chem. Soc., 131, 1638–1639.
35. Loewer, Y., Weiss, C., Biju, A. T., Froehilich, R., Glorius, F. (2011). J. Org.
Chem., 76, 2324–2327.
36. Abrahams, B. F., Price, D. J., Robson, R. (2006). Angew. Chem. Int. Ed.,
45, 806–810.
37. Zhang, Y., Duan, J., Ma, D., Li, P., Li, S., Li, H., Zhou, J., Ma, Z., Feng, Z.,
Wang, B. (2017). Angew. Chem. Int. Ed., 65, 16313–16317.
38. (a) Kim, D. P., Economy, L. G. (1994). Chem. Mater., 6, 395–400; (b)
Fazen, P. L., Remsen, E. E., Beck, J. S., Carroll, P. J., McGhie, A. R., Sneddon,
L. G. (1995). Chem. Mater., 7, 1942–1956.
39. (a) Lynch, A. T., Sneddon, J. G. (1989). J. Am. Chem. Soc., 111, 6201–
6209; (b) Riedel, R., Bill, J., Passing, G. (1991). Adv. Mater., 3, 551–552;
(c) Bonnetot, B., Guilhon, F., Viala, J. C., Mongeot, H. (1995). Chem.
Mater., 7, 299–303.
40. (a) Schaeffer, G. W., Anderson, E. R. (1949). J. Am. Chem., 71, 2143–
2145; (b) Hougn, W. V., Schaeffer, G. W., Dzurns, M., Stewart, A. C. J.
(1955). J. Am. Chem., 77, 864–865
41. Jaska, C. A., Temple, K., Lough, A. J., Manners, I. (2003). J. Am. Chem.,
125, 9424–9434.
42. Framery, E., Vaultier, M. (2000). Heteroat. Chem., 11, 218–225.
43. Wakamiya, A., Ide, T., Yamaguchi, S. (2005). J. Am. Chem., 127, 14859–
14866.
References 65

44. Jackson, K. T., Reich, T. E., EI-kaderi, H. M. (2012). Chem. Commun., 48,
8823–8825.
45. Uribe-Romo, F. J., Hunt, J. R., Furukawa, H., Klock, C., O’Keefe, M., Yaghi,
O. M. (2009). J. Am. Chem. Soc., 131, 4570–4571.
46. Zhang, Y. B., Su, J., Furukawa, H., Yun, Y. F., Gandara, F., Duong, A., Zou, X.
D., Yaghi, O. M. (2013). J. Am. Chem. Soc., 135, 16336–16339.
47. Ma, Y. X., Li, Z. J., Wei, L., Ding, S. Y., Zhang, Y. B., Wang, W. A. (2017). J.
Am. Chem. Soc., 139, 4995–4998.
48. Lin, G. Q., Ding, H. M., Yuan, D. Q., Wang, B. S. (2016). J. Am. Chem. Soc.,
138, 3302–3305.
49. Fang, Q. R., Gu, S., Zheng, J., Zhuang, Z. B., Qiu, S. L., Yan, Y. S. (2014).
Angew. Chem. Int. Ed., 53, 2878–2882.
50. Kandambeth, S., Shinde, D. B., Panda, M. K., Lukose, B., Heine, T.,
Banerjee, R. (2013). Angew. Chem. Int. Ed., 52, 13052–13056.
51. Chen, X., Addicoat, M., Jin, E., Zhai, L., Xu, H., Huang, N., Guo, Z., Liu, L.,
Irle, S., Jiang, D. (2015). J. Am. Chem. Soc., 137, 3241–3247.
52. Chen, X., Addicoat, M., Irle, S., Nagai, A., Jiang, D. (2013). J. Am. Chem.
Soc., 135, 546–549.
53. Xu, H., Gao, J., Jiang, D. (2015). Nat. Chem., 7, 905–912.
54. Ding, X., Chen, Y., Honsho, Y., Feng, X., Saengawang, O., Guo, J., Saeki, A.,
Seki, S., Irle, S., Nagase, S., Parasuk, V., Jiang, D. (2011). J. Am. Chem. Soc.,
133, 14510–14513.
55. Song, J. R., Sun, J., Liu, J., Huang, Z. T., Zheng, Q. Y. (2014). Chem. Commun.,
50, 788–791.
56. Lammi, L. M., Tilford, R. W., Bharathy, M., Lavigne, J. J. (2011). J. Am.
Chem. Soc., 133, 13975–13983.
57. Spitler, E. L., Koo, B. T., Novotney, J. L., Colson, J. W., Uribe-Romo, F. J.,
Gutierrez, G. D., Clancy, P., Dichtel, W. R. (2011). J. Am. Chem. Soc., 133,
19416–19421.
58. Dalapati, S., Addicoat, M., Jin, S., Sakurai, T., Gao, J., Xu, H., Irle, S., Seki,
S., Jiang, D. (2015). Nat. Commnun., 6, 7786.
59. Kandambeth, S., Mallick, A., Lukose, B., Mane, M. V., Heine, T., Banerjee,
R. (2012). J. Am. Chem. Soc., 134, 19524–19527.
60. Biswal, B. P., Chandra, S., Kandambeth, S., Lukose, B., Heine, T., Banerjee,
R. (2013). J. Am. Chem. Soc., 135, 5328–5331.
61. Waller, P. J., Lyle, S. J., Popp, T. M. O., Diercks, C. S., Reimer, J. A., Yaghi, O.
M. (2016). J. Am. Chem. Soc., 138, 15519–15522.
66 Crystallization and Structural Linkages of COFs

62. (a) Job, A., Janeck, C., Bettary, W., Peters, R., Enders, D. (2002).
Tetrahedron, 58, 2253–2329; (b) Sugiura, M., Kobayashi, S. (2005).
Angew. Chem. Int. Ed., 44, 5176–5186.
63. Uribe-Romo, F. J., Doonan, C. J., Furukawa, H., Oisaki, K., Yaghi, O. M.
(2011). J. Am. Chem. Soc., 1338, 11478–11481.
64. Bunk, D. N., Dichtel, W. R. (2013). J. Am. Chem. Soc., 135, 14952–14955.
65. Stegbauer, L., Schwinghammer, K., Lotsch, B. V. (2014). Chem. Sci., 5,
2789–2793.
66. Ding, S. Y., Dong, M., Wang, Y. W., Chen, Y. T., Wang, H. Z., Su, C. Y., Wang,
W. (2016). J. Am. Chem. Soc., 138, 3031–3037.
67. Dalapati, S., Jin, S. B., Gao, Y. T., Xu, T. C., Nagai, A., Jiang, D. L. (2013). J.
Am. Chem. Soc., 135, 17310–17313.
68. Li, Z. P., Feng, X., Zou, Y. C., Zhang, Y. W., Xia, H., Mu, Y. A. (2014). Chem.
Commun., 50, 13825–13828.
69. Li, Z. P., Zhi, Y. F., Feng, X., Ding, X. S., Zou, Y. C., Liu, X. M., Mu, Y. (2015).
Chem.-Eur. J., 21, 12079–12084.
70. Alahakoon, S. B., Thompson, C. M., Nguyen, A. X., Occhialini, G.,
McCandless, G. T., Smaldone, R. A. (2016). Chem. Commun., 52, 2843–
2845.
71. Vyas, V. S., Haase, F., Stegbauer, L., Savasci, G., Podjaski, F., Ochsenfeld,
C., Lotsch, B. V. (2015). Nat. Commun., 6, 8508–8516.
72. Zhang, J., Wang, L. B., Li, N., Liu, J. F., Zhang, Z. B., Zhou, N. C., Zhu, X. L.
(2014). Crystengcomm, 16, 6547–6551.
73. Nagai, A., Chen, X., Feng, X., Ding, X., Guo, Z., Jiang, D. (2013). Angew.
Chem. Int. Ed., 52, 3770–3774.
74. Fang, Q. R., Wang, J. H., Gu, S., Kaspar, R. B., Zhang, J., Wang, J. H., Qiu, S.
L., Tan, Y. S. (2013). J. Am. Chem. Soc., 137, 8352–8355.
75. Fang, Q. R., Zhuang, Z. B., Gu, S., Kaspar, R. B., Zhang, J., Wang, J. H.,
Zhang, J., Wang, J. H., Qiu, S. L., Tan, Y. S. (2014). Nat. Commun., 5, 4503–
4510.
76. Kou, Y., Xu, Y. H., Guo, Z. Q., Jiang, D, L. (2011). Angew. Chem. Int. Ed., 50,
8753–8757.
77. Guo, J., Xu, Y., Jin, S., Chen, L., Kaji, T,; Honsho, Y., Addicoat, M. A., Kim, J.,
Saeki, A., Ihee, H., Seki, S., Irle, S., Hiramoto, M., Gao, J., Jiang, D. (2013).
Nat. Chmmun., 4, 2736–2743.
78. Kuhn, P., Antoietti, M., Thomas, A. (2008). Angew. Chem. Int. Ed., 47,
3450–3453.
References 67

79. Kuhn, P., Forget, A., Su, D., Thomas, A., Antonietti, M. (2008). J. Am.
Chem. Soc., 130, 13333–13337.
80. Bojdys, M. J., Jermenok, J., Thomas, A., Antonietti, M. (2010). Adv.
Mater., 22, 2202–2205.
81. Ren, S. J., Bojdys, M. J., Dawson, R., Layboum, A., Khimyak, Y. Z., Adams,
D. J., Cooper, A. I. (2012). Adv. Mater., 24, 2357–2361.
82. Katekomol, P., Roeser, J., Bojdys, M., Weber, J., Thomas, A. (2013). Chem.
Mater., 25, 1542–1548.
83. Karmakar, A., Kumar, A., Chaudhari, A. K., Samanta, P., Desai, A. V.,
Krishna, R., Ghosh, S. K. (2016). Chem.-Eur. J., 22, 4931–4937.
84. Zeng, Y., Zou, R., Luo, Z., Zhang, H., Yao, X., Ma, X., Zou, R., Zhao, Y. (2015).
J. Am. Chem. Soc., 137, 1020–1023.
85. Li, H., Pan, Q., Ma, Y., Guan, X., Xue, M., Fang, Q., Yan, Y., Valtchev, V., Oiu,
S. (2016). J. Am. Chem. Soc., 138, 14783–14788.
Chapter 3

Gas Adsorption and Storage of COFs

COFs have been touted as a new class of porous crystalline materials


for gas storage applications due to their high inherent surface
areas, very low densities (being composed of light elements, such
as H, C, N, B, O, and Si), high stability from the formation of covalent
bonds, and tunable pore dimensions, ranging from ultramicropores
to mesopores. The gas storage capacity of COFs for hydrogen,
methane, and carbon dioxide has attracted increasing interest in the
past decades [1]. As a general rule, the gas adsorption capacity of
a COF depends primarily on the components and topologies of its
frameworks. Because of their larger surface areas and pore volumes,
3D COFs possess significantly higher uptake capacities than 2D COFs.
Moreover, COF-based thin film nanocomposite (TFN) membranes
and freestanding COF films were introduced, with application in gas
separation.

3.1 Gas Sorption


The true surface area, including surface irregularities and pore
interiors, cannot be calculated from particle size information but
is rather determined at the atomic level by the adsorption of an
unreactive, or inert, gas. The amount adsorbed, let’s call it X, is a
function not only of the total amount of exposed surface but also of
(i) temperature, (ii) gas pressure, and (iii) the strength of interaction

Covalent Organic Frameworks


Atsushi Nagai
Copyright © 2020 Jenny Stanford Publishing Pte. Ltd.
ISBN 978-981-4800-87-7 (Hardcover), 978-1-003-00469-1 (eBook)
www.jennystanford.com
70 Gas Adsorption and Storage of COFs

between gas and solid. Because most gases and solids interact
weakly, the surface must be cooled substantially in order to cause
a measurable amount of adsorption—enough to cover the entire
surface. As the gas pressure is increased, more is adsorbed on the
surface (in a nonlinear way). However, adsorption of a cold gas does
not stop when it has covered the surface in a complete layer one-
molecule thick (let’s call the theoretical monolayer amount of gas Xm).
As the relative pressure is increased, excess gas is adsorbed to form
“multilayers”; thus, gas adsorption as a function of pressure does
not follow a simple relationship, and we must use an appropriate
mathematical model for the surface area.

Figure 3.1 Schematic of gas adsorption.

3.2 Physical and Chemical Adsorption


Solid surfaces are not smooth in the microscopic sense owing to
irregular valleys and peaks distributed all over the area. The regions
of irregularity are particularly susceptible to residual force fields.
At these locations, the surface atoms of the solid may attract other
atoms or molecules in the surrounding gas or liquid phase. The
surfaces of pure crystal have nonuniform force fields due to the
atomic structure in the crystal. Such surfaces also have sites or active
centers where adsorption is favored.
Adsorption on a solid surface may be divided into two categories,
physical adsorption and chemical adsorption. Physical adsorption is
nonspecific and similar to condensation. The forces that attract the
fluid molecules to the solid surface are weak van der Waals forces.
Therefore, physical adsorption is also known as van der Waals
adsorption. The heat evolved during physical adsorption is low,
which usually lies between 2 and 25 kJ/mol. The energy of activation
Physical and Chemical Adsorption 71

for physical adsorption is also low (<5 kJ/mol). Equilibrium between


the solid surface and the gas molecules is usually attained rapidly,
and it is reversible because the energy requirements are small.
Multiple layers of adsorbed molecules are possible, especially near
the condensation temperature. The extent of physical adsorption
decreases with increasing temperatures. Physical adsorption is
useful for determining the surface area and pore size of a solid
catalyst.
In contrast, chemical adsorption, or chemisorption, is specific.
It involves forces that are much stronger than physical adsorption.
The amount of heat evolved in chemisorption is large (e.g., 50–
500 kJ/mol), which is similar to that in chemical adsorptions.
The term “chemisorption” was given by Taylor [2]. However, the
concept of chemisorption was proposed by Langmuir much earlier
[3]. According to him, in the interior of the solid material (i.e., the
adsorbent), the atoms have their force fields wholly satisfied by the
atoms that surround them. The atoms on the surface, however, are
not surrounded. These atoms have a residual force due to unshared
electrons toward the exterior. This residual force field leads to the
sharing of electrons with striking gas molecules (i.e., the adsorbate).
A sort of covalent linkage is formed between the gas molecules
and the atoms of the solid at the surface. However, this linkage is not
the same as the covalent bond that exists in a chemical compound. In
spite of the sharing of the electron with the gas molecule, the surface
atom remains bonded simultaneously with the other atoms in the
crystal lattice.
Langmuir observed that a stable film of oxide formed on tungsten
wires in the presence of oxygen, which was not the same as the normal
oxide, WO3, in terms of properties. This oxide was given off from the
surface upon desorption. This adsorbed compound was designated
as W•O3. Langmuir pointed out that because of the rapid falling off of
intermolecular forces with distance, it is probable that the adsorbed
layers are no more than a single molecule in thickness. Langmuir
proposed simple formulations for rates of adsorption and desorption
of gases on a solid surface [4]. These are also applicable to liquids.
Unlike physical adsorption, chemisorption does not take place on all
solids. It occurs with some chemically reactive gases. Chemisorption
72 Gas Adsorption and Storage of COFs

usually occurs at high temperatures. The surface coverage is limited


to a monolayer. The process is often irreversible. Chemisorption
is used for finding the active centers of a catalyst. Chemisorption
can be divided into two categories, activated chemisorption and
nonactivated chemisorption. In activated chemisorption, the rate
of chemisorption varies with temperature following Arrhenius law
with finite activation energy. In some situations, chemisorption
occurs rapidly, which suggests that the activation energy is nearly
zero. This type of adsorption is called nonactivated chemisorption.
Both activated and nonactivated chemisorption may take place at
different stages of an adsorption process. A comparison between
physical adsorption and chemisorption is presented in Table 3.1.

Table 3.1 Comparison between physical adsorption and chemisorption

Physical adsorption Chemisorption


All solid can be used as adsorbents. Some solids can be used as
adsorbents.
All gases below their critical Some chemically reactive gases can
temperatures can act as act as adsorbents.
adsorbents.
It occurs at low temperatures. It generally occurs at high
temperatures.
The heat of adsorption is low. The heat of reaction is high, which
is on the order of the heat of
chemical reactions.
The rate is high. Both low and high rates are
observed.
The activation energy is low. The activation energy is low for
nonactivated chemisorption and
high for activated chemisorption.
A multilayer is possible. Only a monolayer is formed.
The reaction is highly reversible. The reaction is often irreversible.
It is used for the determination of It is used for the determination
surface area and pore size. of the active center area and
elucidation of surface reaction
kinetics.
Brunauer–Emmett–Teller Theory 73

3.3 Brunauer–Emmett–Teller Theory


The theory of chemical adsorption proposed by Langmuir is based
on the formation of a monolayer. However, multilayers like covalent
organic frameworks (COFs) can form in physical adsorption. For
example, several layers of nitrogen molecules can get adsorbed on
top of each other on the surface of silica gel at 77 K below atmosphere
pressure. The isotherm in the case of a monolayer assumes the shape
shown in Fig. 3.2 (type I).
Saturation

Volume
Adsorbed
B

P0 P0 P0 P0 P0
Pressure

Figure 3.2 Adsorption isotherms.

The extent of adsorption increases with pressure and ultimately


reaches a limiting value, as predicted by the Langmuir theory.
However, for adsorption involving multilayer formation at low
temperature, at least four other types of adsorption isotherms can
be observed (types II–V in Fig. 3.1). Type II is perhaps the most
common for physical adsorption on relatively open surfaces, in which
adsorption proceeds progressively from submonolayer to multilayer;
the isotherm exhibits a distinct concave downward curvature at
some low relative pressure (P/P0) and a sharply rising curve at high
P/P0. Point B at the knee on the curve signifies completion of an
adsorbed monolayer. It forms the basis of the Brunauer–Emmett–
Teller (BET) model for surface area determination of a solid from the
assumed monolayer capacity, described below.
A type III isotherm signifies a relatively weak gas-solid
interaction, as exemplified by the adsorption of water and alkanes
on nonporous low-polarity solids such as polytetrafluoroethylene
(Teflon). In this case, the adsorbate does not effectively spread on
the solid surface. Type IV and V isotherms are the characteristic of
vapor adsorption reaching an asymptotic value as the saturation
pressure is approached. Adsorption of organic vapors on activated
carbon is typically type IV, whereas adsorption of water vapor on
74 Gas Adsorption and Storage of COFs

activated carbon is type V. The shape of the adsorption isotherm of a


solute from the solution is sensitive to the competitive adsorption of
the solvent and other components and may deviate greatly from that
of its vapor on the solid.
As a result, the pore sizes of porous materials can be decided
from isotherm types as follows:
∑ A type I isotherm is obtained when P/P0 < 1 and c > 1 in the
BET equation, where P/P0 is the partial pressure value and c
is the BET constant, which is related to the adsorption energy
of the first monolayer and varies from solid to solid. The
characterization of microporous materials, those with pore
diameters less than 2 nm, gives this type of isotherm.
∑ A type II isotherm is very difficult compared to the Langmuir
model. The flatter region in the middle represents the
formation of a monolayer. A type II isotherm is obtained when
c > 1 in the BET equation. This is the most common isotherm
obtained when using the BET technique. At very low pressures,
the micropores fill with nitrogen gas. At the knee, monolayer
formation and multilayer formation occur at a medium
pressure. At a higher pressure, capillary condensation occurs.
∑ A type III isotherm is obtained when c < 1 and shows the
formation of a multilayer. Because there is no asymptote in
the curve, no monolayer is formed and BET is not applicable.
∑ A type IV isotherm occurs when capillary condensation
occurs. Gases condense in the tiny capillary pores of the
solid at pressures below the saturation pressure of the gas.
In the lower-pressure regions, it shows the formation of a
monolayer followed by the formation of multilayers. The BET
surface area characterization of mesoporous materials, which
are materials with pore diameters between 2 and 50 nm, gives
this type of isotherm.
∑ Type V isotherms are very similar to type IV isotherms and
are not applicable to BET.

3.4 Hydrogen Gas Storage


Hydrogen gas storage undoubtedly attracts public interest because
it represents a future energy resource based on its high chemical
Hydrogen Gas Storage 75

abundance, high energy density, and environmentally friendly


characteristics.
The demand of the US Department of Energy (DOE) for hydrogen
storage is 5.5 wt% in gravimetric capacity and 40 kg/m3 in volumetric
capacity at an operating temperature of 233–333 K with a pressure
of 100 atoms by the year 2017. Recently, a large number of porous
materials have been explored for the application of hydrogen storage.
However, none of candidates developed so far has satisfied the DOE
target. COF materials, as new crystalline porous materials, have also
attracted increasing interest in hydrogen storage applications.
The capabilities of hydrogen storage are summarized by using
representative COFs in Table 3.2. COF-18 (SBET = 1263 m2/g, VP =
0.65 cm3/g, pore size = 1.8 nm) presents the highest hydrogen
uptake at a low pressure (1 bar and 77 K) among similar 2D
COFs with different alkyl chain lengths (COF-11Å, COF-14Å, and
COF-16Å). For context, 3D COFs (COF-102 and COF-103) showed
much higher hydrogen storage capacities because of the higher
surface areas and lower densities, relative to 2D COFs. Among
the different COFs for H2 uptake, the one with the largest storage
capacity is 3D COF-102 (Table 3.2, SBET = 3620 m2/g, VP =
1.55 cm3/g, pore size = 12 nm), which takes up 72 mg/g at 1 bar
and 77 K. This capacity is comparable to those of metal-organic
framework 177 (MOF-177, 75 mg/g, SBET = 4500 m2/g), MOF-5
(76 mg/g, SBET = 3800 m2/g), and the porous aromatic framework
PAF-1 (75 mg/g, SBET = 5600 m2/g). These capacities significantly
demonstrate the potential of COFs for use as hydrogen storage
materials.
However, hydrogen storage of COFs at ambient pressure and
temperature is still low and far from meeting the DOE requirements.
Adsorption simulations of COFs show that doping with charged
species can affect their capacity. Theoretical studies show that metal-
doped COFs can enhance hydrogen storage at a temperature range of
273 to 298 K [5, 6]. Recently, an improvement in the hydrogen storage
capacity has been theoretically predicted by lithium-doped COFs
due to the increased binding energy between the H2 and Li atoms.
These simulation studies indicate that hydrogen uptake by Li-doped
COF-105 and COF-108 (6.84 and 6.73 wt%, respectively, at 298 K
and 100 bar) was more than that by nondoped COFs and previously
76 Gas Adsorption and Storage of COFs

reported MOF materials. In spite of the fact that experimental and


theoretical results indicate that COFs are promising candidates for
hydrogen storage, practical and industrial use of COFs toward the
DOE target for hydrogen uptake is still far away.

Table 3.2 Selection of COFs with BET, porosity parameters, and H2 capture

BET surface Pore size Pore volume H2 uptake


COFs area (m2/g) (nm) (Vp, cm3/g) (wt%)
COF-1 750 0.9 0.3 1.48
COF-5 1670 2.7 1.07 3.58
COF-6 750 0.64 0.32 2.26
COF-8 1350 1.87 0.69 3.5
COF-10 1760 3.41 1.44 3.92
COF-102 3620 1.15 1.55 7.24
COF-103 3530 1.25 1.54 7.05
COF-1 628 0.9 0.36 1.28 (1 bar)
COF-11Å 105 1.1 0.05 1.22 (1 bar)
COF-14Å 805 1.4 0.41 1.23 (1 bar)
COF-16Å 753 1.6 0.39 1.4 (1 bar)
COF-18Å 1263 1.8 0.69 1.55 (1 bar)
CTC COF 1710 2.26 1.03 1.12 (0.5 bar)
Dual-pore COF 1771 0.73, 2.52 3.189 (P/P0) 1.37 (1 bar)
Source: Reproduced from Ref. [4] under the Creative Commons Attribution License
(https://creativecommons.org/licenses/by/4.0/).

3.5 Methane Gas Storage


Methane gas stands out as a potential vehicular fuel; however, the
lack of an effective, economic, and safe on-board storage system is a
major technical barrier that prevents methane-driven automobiles
from competing with the traditional ones. According to the US
DOE requirements, the current target for methane storage is 350
cm3STP/cm3 absorbent (u/u) and 0.5 gCH4/gadsorbent (699 cm3/STPg)
Methane Gas Storage 77

at 35 bar and 298 K (STP: standard temperature and pressure).


In Table 3.3, COF-1 as a 2D COF shows CH4 uptake of 40 mg/g at
298 K and 35 bar. However, upon increasing the pressure, no
increase in methane adsorption is observed because of the small
pores in COF-1. The pores in COF-1 are quickly occupied by CH4
molecules at low pressures and higher pressures may not compact
the gas any more. COF-102 has the largest pores as a 3D COF and
shows a higher CH4 storage capacity (187 mg/g). COF-103 possesses
a high methane uptake capacity (up to 175 mg/g). These values are
comparable to the highest methane uptake by an MOF (MOF-210;
210 mg/g). Theoretical simulation studies indicate that lithium-
doped COFs could significantly strengthen the building between
methane and Li+ cations, which will further improve the methane
capacity of COFs. In these results, methane storage by lithium-doped
COFs is double at 298 K and low pressures (>50 bar), as compared
to that by nondoped COFs.
The above-mentioned studies indicate the potential application
of COFs for methane storage at ambient temperatures and 35 bar.
However, the practical use of COFs to meet the requirements of the
DOE is a big challenge.

Table 3.3 BET surface area, porosity, parameters, and CH4 capture of selected
COFs

BET Surface Pore size Pore volume CH4 uptake


COFs area (m2/g) (nm) (VP, cm3/g) (mg/g)
COF-1 750 0.9 0.3 40
COF-5 1670 2.7 1.07 89
COF-6 750 0.64 0.32 65
COF-8 1350 1.87 0.69 87
COF-10 1760 3.41 1.44 80
COF-102 3620 1.15 1.55 187
COF-103 3530 1.25 1.54 175
ILCOF-1 2723 0.23 1.21 129 (L/L)
Source: Reproduced from Ref. [4] under the Creative Commons Attribution License
(https://creativecommons.org/licenses/by/4.0/).
78 Gas Adsorption and Storage of COFs

3.6 Carbon Dioxide Gas Storage


Carbon dioxide is the primary greenhouse gas that is responsible
for global warming, rising sea levels, and the increasing acidity
of the oceans. The possibility of capturing CO2 from industrial
emission sources has attracted broad interest. Among the current
strategies, adsorption by porous materials is energetically efficient
and technically feasible. In 2009, Yaghi and coworkers studied a
family of 2D COFs with 1D micropores (COF-1 and COF-6), 2D COFs
with mesopores (COF-5, COF-8, and COF-10), and 3D COFs with 3D
medium-sized pores (COF-102 and COF-103) on the application of
CO2 capture. Table 3.4 shows the storage of CO2 at 298 K and 55 bar,
in the order COF-102 > COF-103 > COF-10 > COF-5 > COF-8 > COF-6
> COF-1, which indicates that high-pressure CO2 capture is highly
related to the pore volumes and SABET of COFs.
Functional groups inside the channel walls of COFs can enhance
CO2 adsorption and separation. Fluorinated alkanes exhibit an
extraordinary affinity to CO2 molecules. Han and coworkers reported
fluorinated triazine–based COFs as FCFT-1 and FCTF-1-600 for CO2
capture by using tetrafluoroterephthalonitrile as a monomer [8].
Triazine-based COFs possess N-rich frameworks, which are favored
for CO2 adsorption. Meanwhile, the high electron negativity of F
enhances CO2 electrostatic interaction. The CO2 adsorption values of
these triazine-based COFs at 273 K and 1 bar were as follows: FCTF-
1-600 (124 cm3/g) > FCTF-1 (105 cm3/g) > CTF-1-600 (86 cm3/g)
> CTF-1 (55 cm3/g). These results demonstrate that the fluorinated
COFs show much better CO2 uptake under low pressure. Jiang et al.
converted a conventional hydroxy group–modified 2D COF ([HO]x%
-H2P COFs) into an outstanding CO2-capturing material, carboxyl
group–modified 2D COF ([HO2C]x%-H2P COFs), via the quantitative
ring opening reaction of succinic anhydride, as shown in Fig. 3.3 [9].
[HO2C]x%-H2P COFs show micropores ranging from 1.4 to 2.2 nm
instead of mesopores of [OH]x%-H2P COFs (2.5 nm). From the results
of Table 3.4, the [OH]x%-H2P COFs have a low CO2 capacity (between
23 and 32 cm3/g) at 273 K and 1 bar. However, [HO2C]x%-H2P COFs
exhibit enhanced CO2 adsorption between 49 and 89 cm3/g under
the same conditions.
Table 3.4 BET surface area, porosity parameters, and CO2 capture of selected COFs

COFs BET surface area (m2/g) Pore size (nm) Pore volume (VP, cm3/g) CO2 uptake (cm3/g)
Lowa Highb Ref.
COF-1 750 0.9 0.3 51 117 [7]
COF-5 1670 2.7 1.07 31 443
COF-6 750 0.64 0.32 85 158
COF-8 1350 1.87 0.69 33 321
COF-10 1760 3.41 1.44 27 514
COF-102 3620 1.15 1.55 34 611
COF-103 3530 1.25 1.54 38 606
FCTF-1 662 0.46, 0.54 - 105 [8]
FCTF-1-600 1535 0.46, 0.59 - 124
CTF-1 746 0.54 55
CTF-1-600 1553 - 80
[HO]25%-H2P COF 1054 2.5 27
[HO]50%-H2P COF 1089 2.5 23
[HO]75%-H2P COF 1153 2.5 26
Carbon Dioxide Gas Storage

[HO]100%-H2P COF 1284 2.5 32

(Continued)
79
80

Table 3.4 (Continued)

COFs BET surface area (m2/g) Pore size (nm) Pore volume (VP, cm3/g) CO2 uptake (cm3/g)
Lowa Highb Ref.
[HCO2]25%-H2P COF 786 2.2 49
[HCO2]50%-H2P COF 673 1.9 68
[HCO2]75%-H2P COF 482 1.7 80
Gas Adsorption and Storage of COFs

[HCO2]100%-H2P COF 364 1.4 89


[HC∫C]0-H2P COF 1474 2.5 0.75 37 [9]
[HC∫C]25-H2P COF 1413 2.3 0.71 27
[HC∫C]50-H2P COF 962 2.1 0.57 24
[HC∫C]75-H2P COF 683 1.9 0.42 22
[HC∫C]100-H2P COF 462 1.6 0.28 20
[Et]25-H2P COF 1326 2.2 0.55 28
[Et]50-H2P COF 821 1.9 0.48 23
[Et]75-H2P COF 485 1.6 0.34 21
[Et]100-H2P COF 187 1.5 0.18 19
[MeOAc]25-H2P COF 1238 2.1 0.51 43
[MeOAc]50-H2P COF 754 1.8 0.42 45
[MeOAc]75-H2P COF 472 1.5 0.31 42
COFs BET surface area (m2/g) Pore size (nm) Pore volume (VP, cm3/g) CO2 uptake (cm3/g)
Lowa Highb Ref.
[MeOAc]100-H2P COF 156 1.1 0.14 33
[AcOH]25-H2P COF 1252 2.2 0.52 48
[AcOH]50-H2P COF 866 1.8 0.45 60
[AcOH]75-H2P COF 402 1.5 0.32 55
[AcOH]100-H2P COF 186 1.3 0.18 49
[EtOH]25-H2P COF 1248 2.2 0.56 47
[EtOH]50-H2P COF 784 1.9 0.43 63
[EtOH]75-H2P COF 486 1.6 0.36 60
[EtOH]100-H2P COF 214 1.4 0.19 43
[EtNH2]25-H2P COF 1402 2.2 0.58 59
[EtNH2]50-H2P COF 1044 1.9 0.5 80
[EtNH2]75-H2P COF 568 1.6 0.36 68
[EtNH2]100-H2P COF 382 1.3 0.21 49
2
aCO uptake was measured at 1 bar and 273 K.

2
bCO uptake was measured at 55 bar and 298 K.
Carbon Dioxide Gas Storage

Source: Reproduced from Ref. [4] under the Creative Commons Attribution License (https://creativecommons.org/licenses/by/4.0/).
81
82

CHO
HO COOH

OH O
CHO NH N OH NH N O
N N
DHTA NH N NH N
N HN N HN
X% N N
HO N HN O N HN
O

CHO HOOC
N O N
O O
N N
2.5 nm
CHO
Gas Adsorption and Storage of COFs

PA N N
(100-X)%
N N
COOH

O
H2N NH2 N OH N O
NH NH
N NH N N NH N
N N
N HN N HN
N
H HO N HN O N HN
N N O
H
N
HOOC

H2 N NH2 X = 25: [HO2C]25%-H2P-COF


H 2P X = 25: [HO]25%-H2P-COF X = 50 : [HO2C]50%-H2P-COF
X = 50 : [HO]50%-H2P-COF X = 75 : [HO2C]75%-H2P-COF
X = 75 : [HO]75%-H2P-COF X = 100 : [HO2C]100%-H2P-COF
X = 100 : [HO]100%-H2P-COF

Figure 3.3 Synthesis of [OH]X%-H2P COFs with channel walls functionalized with carboxylic acid groups through the ring-opening reaction
of [OH]X%-H2P COFs with succinic anhydride.
Membrane Separation of COFs 83

Jiang et al. also reported a similar strategy to modify the pore


surface with different functional groups via a click reaction [10]. In
Table 3.4, it’s clear that different functions of pore walls of COFs can
induce a decrease in surface areas, pore sizes, and pore volumes.
However, the capacity of CO2 adsorption is highly dependent on
the structures of the functional groups. Ethyl-functionalized COFs,
[Et]x-H2P COFs, show a CO2 adsorption capacity similar to those of
[CH∫C]x-H2P COFs. The CO2 adsorption capacities of [CH∫C]x-H2P
COFs and [EtNH2]x-H2P COFs is enhanced with an increase in the BET
surface area and pore size. However, [MeOAc]x-H2P COFs, [AcOH]
x-H2P COFs, [EtOH]x-H2P COFs, and [EtNH2]x-H2P COFs exhibit much
higher CO2 adsorption capacities than those of [CH∫C]x-H2P COFs.
The adsorption capacities are in the order of [CH∫C]x-H2P COFs
ª [Et]x-H2P COFs << [MeOAc]x-H2P COFs < [AcOH]x-H2P COFs ª
[EtNH2]x-H2P COFs, which is because of the interaction between the
functional group and the CO2 molecule. As these results show, the
CO2 adsorption performance is highly related to the pore surface of
the COFs.

3.7 Membrane Separation of COFs


The implementation of the membrane technology in the context of
resource recovery and sustainable development has demonstrated
an ecofriendly potential for overcoming energy and environmental
challenges. Due to its low energy consumption, small carbon
footprint, and easy operation and scalability, membrane separation
has experienced rapid growth in the past few decades. Until now,
a variety of membranes with pore sizes ranging from several
micrometers to subnanometers have been fabricated from
conventional amorphous polymers, such as polyimide, polyamide
(PA), polyacrylonitrile, polyether sulfone (PES), polysulfone,
polyvinylidene fluoride (PVDF), and poly(amide-imide)-based cross-
linked polymers [11]. However, these polymers are not optimal
for transport because the resulting membranes lack ordered and
tunable pore channels, leading to a typical nonuniform pore size,
limited porosity, and poor interconnectivity. Novel membranes
fabricated from porous crystalline materials with uniform pore
channels are significantly advanced in selectivity and porosity.
84 Gas Adsorption and Storage of COFs

Crystalline materials such as zeolites and MOFs have demonstrated


a superior performance over conventional polymer membranes.
However, due to their relatively low adhesion to a polymeric support,
partly organic nature, and possible defects between crystals, the full
capability of MOF-based membranes is relatively difficult to realize.
Therefore, the development and design of new porous crystalline
polymers with ordered pore structures and tunable pore sizes are
important for the synthesis of advanced separation membranes.
Up to now, studies on COF-based membranes have cover a broad
range of topics, including gas separation membranes, separation
membranes for liquid phase, and proton exchange membranes in
fuel cells. This section is an introduction and discussion of membrane
design strategies based on COFs. The optimal method to synthesize a
high-performance COF membrane in terms of high perm-selectivity
will be highlighted. Especially, this section summarizes the
applications of COF membranes in gas separation.

3.7.1 Key Properties of COFs for Membrane Separation


Due to their well-defined and precise pore structure with relatively
large surface areas, COFs were first used for gas storage. Recently,
along with the emergence of well-designed and multifunctional
COFs, their application has gradually extended to membrane
technology for efficient and precise separation. Different from
gas storage, membrane separation covers a wide variety of areas,
including gas separation, water purification/treatment, organic
solvent nanofiltration, pervaporation, and fuel cells, in which the
membranes are exposed to various mild/harsh environments. This
makes the synthesis of suitable COFs for membrane separation more
complicated compared to other applications, requiring rational
selection criteria for fabricating COF-based membranes. Specifically,
the physical pore size defines the application of membranes, for
example, in reverse osmosis, nanofiltration, ultrafiltration, and
microfiltration. So the selection of COFs with specific pore sizes
is key for membrane design. In addition, membrane fabrication
uses mild or harsh solvents and mechanical treatments; thus, the
chemical and mechanical stability of COFs is another selection
criterion. Other properties, like hydrophilicity/hydrophobicity,
surface charge, and proton conductivity, also have a significant
Membrane Separation of COFs 85

influence on the performance of COF-based membranes in various


separations. Altogether, the selection of COFs has many variables that
determine the separation performance and future applications in
membrane technology. The best way to construct desired COF-based
membranes with a high perm-selectivity and long-term stability is
the rational selection of COFs on the basis of various combinations
of their properties.

3.7.2 Fabrication of COF-Based Membranes

3.7.2.1 Design principles


A rational design strategy bridging the gap between COFs and COF-
based membranes is the key to optimizing the advantages of COFs
for precise and rapid membrane separation. When compared with
other porous materials, COFs have orderly and closely arranged
pore structures, which is the main advantage of COFs in membrane
applications. Therefore, how to achieve the full potential of these
orderly arranged pore structures is the core challenge in fabricating
a COF-based membrane. Initially, COFs were blended into polymeric
matrix membranes as porous fillers to gain additional channels
for gas, water, and solvents to pass through. However, the pore
structures resulting from the phase inversion of polymer solutions
were still the dominating transport pathways. With advancing
membrane fabrication techniques and COF synthesis methods,
the pore structures in COFs are gaining increasing significance in
membrane separation. Continuous COF-based membranes were
synthesized from uninterrupted and pure COFs by in situ growth,
layer-by-layer stacking, and interfacial polymerization (IP). These
methods will be discussed in detail next.

3.7.2.2 Blending
Integrating COFs into common membrane fabrication methods,
including nonsolvent-induced phase inversion (NIPS) and IP, is a
simple and reproducible approach for fabricating COF-based hybrid
membranes. Compared to other classical inorganic particles, the
fully organic nature endows COFs with excellent compatibility with
86 Gas Adsorption and Storage of COFs

polymer matrices. Moreover, the pore structures of COFs provide


additional passages for gas, water, and organic solvent molecules to
permeate and enhance the perm-membrane fabrication methods.
The routes of introducing COFs in these two approaches are
different. In a typical NIPS, COFs are directly added in the organic
solvents with polymers to achieve a homogeneous dope solution.
Then a film of a certain thickness is formed by casing the solution on
a glass plate or nonwoven and subsequently experiences an induced
phase inversion while being immersed in water, to form a membrane.
The final COF-based membrane is obtained after the removal of the
leftover solvent by immersion in deionized water. These COF-based
membranes are also known as COF-based mixed matrix membranes
(MMMs). In the IP process, the incorporation of COFs in a thin PA
film is realized by dispersing COFs with the diamine monomers in
the water phase. During the reaction between diamine monomers
from the aqueous phase and the acid chloride monomers from
organic phase, the COFs dispersed in the aqueous phase are trapped
in the formed active thin layer. These two approaches of blending
COFs in polymer membranes will be highlighted in this section.
In the last three years, various MMMs have been constructed
by embedding COFs, such as SNW-1, COF-1, and functional COFs, in
polymers. In particular, 2D SNW-1 has been extensively investigated
because of its relatively low synthesis cost and high hydrophilicity.
Pioneering SNW-1-based membranes were fabricated by
incorporating COF SNW-1 particles with a diameter of 50–70 nm
into sodium alginate (SA) matrices for ethanol dehydration [12].
The incorporation of SNW-1 particles endowed the membranes with
an enhanced thermal/mechanical stability and good antiswelling
properties. The SA-SNW-1 (25) membrane with an SNW-1 loading
of 25 wt% achieved the optimal separation performance with a
permeation flux of 2397 g/m2 h and a separation factor of 1293 for
ethanol dehydration, which was superior to that of other porous
material–based membranes due to the porous structure of SNW-1.
In addition to SNW-1, ketoenamine-linked COFs, TpPa-1, TpPa-
2, TpHZ, TpBD, NUS-10, and NUS-9, have received tremendous
attention in synthesizing COF-based MMMs because of their
Membrane Separation of COFs 87

relatively high stability and hydrophilicity. Ketoenamine-linked


COFs shaped like hollow spheres, particles, and nanosheets were
explored in membrane fabrication and separation. Spherical flower–
like TpBD and TpPa-1 particles were used as the active phase in the
polymer (polybenzimidazole–5-tert-butylisophthalic acid [PBI-BuI])
for gas separation (Fig. 3.4) [13]. The compatibility between COFs
and PBI-BuI was significantly improved by creating intermolecular
interactions between H-bonded benzimidazole groups of PBI and
COFs; thus a considerably higher loading of COF (50%) in the polymer
matrix was achieved. The obtained membrane with a loading of 50%
TpBD showed a sevenfold elevation in the permeability of gases (H2,
N2, CH4, and CO2) compared with the pristine membrane, with small
deviations in selectivity. The improved permeation was mainly due
to the relatively low diffusion resistance achieved by incorporated
COFs, while the selectivity was still governed and maintained by
the polymer matrix. Assisted by a similar strategy, in 2018, TpBD
with hollow nanosphere structures was constructed by a facile
template-directed approach followed by mixing with SA matrices to
produce membranes for dehydration [14]. The structure of hollow
nanospheres was realized by etching the Fe3O4 core of TpBD@Fe3O4
obtained through the synthesis of TpBD on the template of an Fe3O4
nanocluster, in a HCl solution (Fig. 3.5). The incorporation of hollow
TpBD nanospheres into the polymer matrix allowed the prepared
membrane to have improved hydrophilicity and a large number of
microporous channels resulting from the pore structure of hollow
nanospheres and the free volume cavities between COFs and the
polymer. In addition, due to the highly porous structure of hollow
TpBD nanospheres, the porosity of TpBD-based MMMs was notably
boosted, resulting in a low flow resistance. COF nanosheets peeled
off from the bulk COF particles were also investigated for fabricating
MMMs. Cao et al. embedded 2D TpHz nanosheets with a lateral size
of 200–400 nm and a thickness of 4.5 nm in a PES matrix to fabricate
functional MMMs [15]. An interesting feature in this study was that
the content of the COFs on the membrane surface can be up to 50.90
vol% although only 12 wt% of TpHz was added in the dope solution;
this was primarily ascribed to the migration of COF nanosheets from
the matrix to the surface during the membrane formation.
88 Gas Adsorption and Storage of COFs

(A)

(B) (C)

Figure 3.4 COFs/PBI-BuI membrane. (A) Schematic representations of the


synthesis of COFs and their packing models, indicating the pore aperture and
stacking distances. (B) Overview of the solution-casting method for COF@PBI-
BuI hybrid membrane fabrication. (C) Digital photographs showing the flexibility
of TpPa-1 and TpBD(50)@PBI-BuI. Reproduced with permission from Ref. [13].
Copyright (2016), John Wiley and Sons.

Figure 3.5 The synthesis protocols and chemical structures of the H-TpBD.
Reprinted from Ref. [14], Copyright (2018), with permission from Elsevier.

Compared with pristine COFs, functionalized COFs allow for


the fabrication of versatile and tailored membranes with diverse
properties. Dan et al. reported the synthesis of COFs (NUS-9 and
NUS-10) with an intrinsic proton-conductive performance by
introducing sulfonic acid groups in organic linkers [16]. These two
Membrane Separation of COFs 89

functional COFs were then incorporated in PVDF to prepare proton


exchange membranes.
Unlike MMMs, TFN membranes are synthesized by a rapid IP of
two monomers from the organic and aqueous phase. The viscosity of
the aqueous phase is much lower than that of the dope solution for
the synthesis of MMMs. Thus achieving a good dispersion of COFs in
solution is much easier and not time consuming. Besides, COFs are
only incorporated in the thin layer of TFN membranes, which would
result in a relatively low consumption of COFs during membrane
fabrication. The first COF-incorporated TFN membrane was reported
by Wu et al. in 2016 [17]. In their study, SNW-1 particles bearing
secondary amine groups were introduced in the PA thin film via IP
of trimesoylchloride (TMC) and piperazine monomers on top of a
PES substrate. The –NH– groups of SNW-1 are involved in IP with
–COCl groups of TMC, so the interface compatibility between SNW-
1 and PA is thought to be high, which results in the high stability
of SNW-1 in the thin active film. The obtained COF-based TFN
membranes exhibited an increased pure water flux from 10 L m–2
h–1 bar–1 to 19.25 m–2 h–1 bar–1 while maintaining a Na2SO4 rejection
of over 80%. The enhanced performance was mainly attributed to
the improved hydrophilicity and the increased number of water
channels derived from the pore structure of SNW-1 and the voids at
the interface between SNW-1 and the polymer. Similarly, the same
COF, SNW-1, was embedded in a TFN membrane synthesized via
IP of m-phenylenediamine and TMC (Fig. 3.6) for organic solvent
nanofiltration [18]. The resultant membranes demonstrated an
improved surface hydrophilicity and a decreased skin layer thickness,
and thus a 46.7% increment of ethanol permeation and an increased
rhodamine B (476 Da) rejection of up to 99.4% were achieved
compared with COF-free membranes. Currently, although only SNW-
1 has been applied in fabricating COF-based TFN membranes, many
more different COFs can be explored in fabricating COF-based TFN
membranes due to their specific characteristics, such as porous
structures arranged in an orderly manner and relatively low density.
For instance, chemically stable COFs with pore sizes of around 1
nm—TpHz, ACOF-1, COF-300, TR-CIF-1, CTF-1, and COF-LIZ1—also
have great potential to be applied in synthesizing TFN membranes
with relatively high separation performance, although they have not
been explored yet.
90 Gas Adsorption and Storage of COFs

Figure 3.6 Schematic illustration of the fabrication of the TFN-OSN membrane


with COFs incorporated inside the Paskin layer. Reprinted from Ref. [18],
Copyright (2019), with permission from Elsevier.

3.7.2.3 In situ growth


Generally, discontinuous COF-based membranes cannot fulfill the
potential of the pure structures of COFs for membrane separation.
Thus, advanced membranes with high perm-selectivity are difficult
to synthesize by blending. Therefore, establishing methods of
preparing continuous COF-based membranes is highly necessary
for obtaining a higher separation performance. With advances in
COF synthesis, continuous COF membranes were grown in situ by
refining the current COF synthesis methods. Both freestanding COF
films and COF-based composite membranes can be designed by in
situ growth.
Kharul et al. reported that in situ growth of freestanding COF-
based membranes was pioneered in 2017 (Fig. 3.7) [19]. Briefly,
in the first step, the powered aromatic diamine the coreagent,
p-toluene sulfonic acid (PTSA), and water (PTSA·H2O) were mixed
in water to form an organic salt. The resultant salts and Tp were
shaken thoroughly using a vortex shaker to make the dough. Then, it
was knife-cast on a clean glass plate to generate a film. Finally, COF-
based membranes were produced via the in situ growth of COFs by
baking the film at 60°C–120°C in a programmable oven for 12–72 h.
The obtained continuous COF-based membrane (M-TpTD) showed
Membrane Separation of COFs 91

an acetonitrile flux of 260 L m–2 h–1 bar–1, which was 2.5 times higher
in magnitude than the PA-based literature-reported NF membranes
with similar solute rejection (~99%). This relatively high perm-
selectivity was mainly because the pores of COFs are fully employed
as the channels for solvent molecules to be transported. This
strategy was further extended to synthesize TpBD(Me)2-, TpAzo-,
and TpBpy-based proton exchange membranes by slightly changing
the casting of the reactant mixture on a mold instead of a glass plate
[20]. More recently, TpOMe-Pa1-, Tp-OMe-BD(NO2)2-, TpOMe-Azo-,
and TpOMe-Bpy-based COF membranes were synthesized [21].

Figure 3.7 In situ growth of COF membranes (COMs). (a) Schematic


representation of COM (M-TpBD) fabrication. (b) The pore apertures shown
inside the space-filling COM model in the panel have been calculated by
measuring the distance between the two opposite arms of the hexagon after
subtraction of van der Waals radius of the hydrogen atom. (c) Space-filling
packing model of M-TpBD hexagonal frameworks. (d) Comparison of N2
adsorption isotherms of all six COMs prepared in this work. Reproduced with
permission from Ref. [19]. Copyright (2017), John Wiley and Sons.

To address the issue of the low mechanical strength of the


freestanding COF-based membranes, COFs were grown in situ on
92 Gas Adsorption and Storage of COFs

modified or unmodified porous supporting membranes. Taking


advantage of this strategy, Caro et al. fabricated continuous and
high-quality COF-LUZ1 membranes supported on commercial
ceramic tube membranes for dye removal from water [22]. In their
research, the surface of the alumina tube was first modified with
3-amino-propyltriethoxysilane and was then further functionalized
with aldehyde groups by reaction with 1,3,5-triformylbenzene
(TFB) at 150°C for 1 h (Fig. 3.8). The final well-intergrown COF-
LZU1 layer on the alumina tube was realized by vertically placing the
modified alumina tube in the Teflon-lined stainless-steel autoclave
with reactant mixtures and the following reaction at 120°C for
72 h. This strategy was further extended to prepare continuous
COFs bilayers on a flat alumina porous substrate [23]. COF-LUZ1-
ACOF-1 membranes were fabricated by first synthesizing COF-LUZ1
by condensation of TFB with p-phenylenediamine (PDA) at room
temperature and then synthesizing high-crystallinity ACOF-1 by
condensation of TFB with hydrazine hydrate at a higher temperature.
The resultant COF-LUZ1-ACOF-1 bilayer membrane showed a much
higher separation selectivity for H2/CO2, H2/N2, and H2/CH4 gas
mixture than the individual pore networks, surpassing the Robeson
upper bounds. By utilizing a similar strategy, COF-MOF composite
membranes were also fabricated [24].

Figure 3.8 Synthesis of tubular COF-LZU1 membranes. Reproduced with


permission from Ref. [22]. Copyright (2018), John Wiley and Sons.

3.7.2.4 Layer-by-layer stacking


The route of membrane fabrication via layer-by-layer stacking of
nanosheets or monolayers originally came from the synthesis of
membranes from graphene and graphene oxide monolayers [25].
Membrane Separation of COFs 93

This involved exfoliation of bulk materials comprising monolayers


in water or solvents. Then, the nanosheets were stacked on a porous
substrate by pressure- or vacuum-assisted filtration, or dip coating,
to form a continuous thin membrane. This approach is applicable
to prepare continuous COF-based membranes from COF nanosheets.
Most of the prepared COFs have a 2D lamellar structure, resulting
from the planar organic linkers, and 2D COFs nanosheets were easily
obtained via the splitting of bulk COFs by breaking the van der Waals
force between the neighboring layers. Many different methods of
COF exfoliation have been developed, which have been summarized
by Tang et al. [26]. Different from the in situ growth approach, the
selective COF layer of the synthesized membranes is thinner. This
thin layer helps to achieve a low flow resistance when gases, water,
and solvents pass through the membrane. The thickness of this layer
is controllable by simply varying the amount of COF nanosheets.
Tsuru et al. fabricated an ultrathin COF membrane (~100 nm) for
gas separation by repeatedly dip-coating COF nanosheets on the
support membrane and then drying them at room temperature
(Fig. 3.9) [27]. Due to relatively low flow resistance and high
porosity, the synthesized membrane showed a high H2 permeance
(1 ¥ 10–6 mol m–2 Pa–1 s–1), which was approximately 1 magnitude
higher than that of most reported MOF membranes. By using this
approach, COF/graphene oxide composite membranes for H2/CO2
separation and functionalized COF-based membranes for ion sieving
were developed [28, 29].

Figure 3.9 Schematic illustration of the preparation of a COF-1 membrane via


the assembly of exfoliated COF-1 nanosheets. Reprinted with permission from
Ref. [27]. Copyright (2017) American Chemical Society.
94 Gas Adsorption and Storage of COFs

3.7.2.5 Interfacial polymerization


IP dominates the synthesis of thin film composite membranes with
PA active layers because of its scalability for commercial production
and capacity to produce thin separating layers with relatively high
water permeability. To achieve the full advantage of the ordered and
nanosized pore channels (1–2 nm) of COFs, this method was well
adapted to fabricate thin continuous COF-based membranes. In
2017, a pioneering fabrication of continuous COF-based membranes
using IP was reported by Banerjee et al. [30]. In this work, the
aldehyde organic linker Tp was dissolved in dichloromethane while
amine monomers were dissolved in water. The thin active layer was
formed by the polycondensation of Tp with diamine and triamine
at the interface of dichloromethane and water (Fig. 3.10). To avoid
the formation of amorphous polymers during the organic Schiff
base reaction, amines were first salt-mediated by PTSA. The formed
thin active layer was then transferred on porous grids for filtration
measurements. The H bonding in the PTSA-amine decreased the
diffusion rate of amine organic linkers, and the reaction rate was
slowed down with thermodynamically controlled crystallization.
Due to the highly porous structure, the TpBpy thin film membrane
displayed unprecedented acetonitrile performance of porous
339 L m–2 h–1 bar–1.
More recently, different from the traditional IP, a new approach
of confirming the polymerization of monomers at the oil/water
interface by using Sc(OTf)3 as the Lewis acid catalyst was developed
[31]. Specifically, the catalyst was preferentially dissolved in water
while the monomers terephthalaldehyde (PDA) and 1,3,5-tris(4-
aminophenyl)-benzene were dissolved in an organic phase. The
reaction was initialized by placing the oily organic solution on the
surface of the water phase containing catalysts and freestanding
COF-based membranes were formed at the interface.
Apart from the IP that occurred at the interface of liquid/liquid, IP
at the liquid/air interface was also established for the fabrication of
a continuous COF-based membrane. To achieve the polymerization
of monomers at the water/air interface, Lai et al. first synthesized
an amphiphilic diamine monomer, 9,9-dihexylfluorene-2,7-diamine
(DHF), by attaching two hexyl groups to 2,7-diaminofluorene [32].
Then, a layer of DHF and Tp was formed on the surface of water,
Membrane Separation of COFs 95

followed by the complete evaporation of toluene. The polymerization


of monomers at the interface was initiated by adding trifluoroacetic
acid to the water. The final yellow TPF-DHF thin membranes with
a thickness of 3 nm were obtained after a reaction of 48 h at room
temperature.

Figure 3.10 Synthesis scheme of COF thin films. (a) Schematic representation
of the interfacial crystallization process used to synthesize a TpBpy thin film. The
colorless bottom layer corresponds to aldehyde in a dichloromethane solution,
the blue layer contains only water as the spacer solution, and the yellow top
layer is the Bpy amine-PTSA aqueous solution. (b, c) SEM and AFM images (with
corresponding height profiles) of the TpBpy thin film synthesized as illustrated
in (a). (d) Chemdraw structures of all the COFs used for synthesizing the thin
films via an interfacial crystallization process. Reprinted with permission from
Ref. [30]. Copyright (2017) American Chemical Society.

3.7.3 Gas Separation of COF-Based Membranes


Conventional membranes fabricated from amorphous polymers
suffer from a disordered and inconsistent pore size and have
difficulties in achieving a perm-selectivity surpassing the current
96 Gas Adsorption and Storage of COFs

Robeson upper bounds. COFs with abundant and well-ordered in-


plane pores are particularly promising in realizing ultrafast and
highly selective molecular sieving. Therefore, COF-based membranes
for gas separation have been fabricated. The applications of COF-
based membranes in gas separation are summarized with particular
attention to the separation of H2 from other gases and CO2/CH4
separation due to the rise in energy consumption and because
environmental concerns result in demands for clean energy.
Hydrogen is an alternative clean energy source for the
replacement of conventional fuels in cars, but the current H2
production methods inevitably require the separation of unwanted
gases. Both discontinuous and continuous COF-based membranes
were applied for the separation of H2 from other gases. A
computational study demonstrated that the monolayer COF(CTF-
O)-based membrane can achieve an exceptionally high separation
factor at room temperature for the separation of H2/CO2, H2/N2, H2/
CO2, and H2/CH4 up to 9 ¥ 1013, 4 ¥ 1024, 1 ¥ 1022, and 2 ¥ 1036, which
theoretically proved the advantages of COF-based membranes in gas
separation. However, the highest separation factor for separation of
H2/CO2, H2/N2, H2/CH4 obtained from experimental studies were
31.4 [33], 140 [13], and 84 [23], respectively, which is much lower
than the theoretical value. This might be because the experimental
COF-based membranes are much thicker than the ideal continuous
COF membrane of one monolayer. It is worth pointing out that
continuous COF-based membranes normally have a much higher
H2, a relatively high porosity, and low resistance. For example, the
highest H2 permeability (1.7 ¥ 10–6 mol m–2 s–1 Pa–1) was realized
by an ultrathin continuous 2D-CTF-1 membrane (100 nm) [28].
The selectivity of H2/CO2 improved with the increase of membrane
thickness while the gas permeance reduced.
CH4 is another attractive substitute for petroleum due to its
abundant natural reserves and economic advantages. However, the
presence of CO2 in natural gas leads to a reduced heat value and
pipeline corrosion, and efficient removal of CO2 from CH2/CH4 is
highly desirable. To select COFs for CO2/CH4 separation, Zhong et
al. conducted high-throughput computational screening and design
of COFs (Fig. 3.11) [34] and 298 COFs with pore-limiting diameters
(PLDs) larger than 3.3 Å were evaluated. The membrane selectivity
generally increased with the decline of the PLD because of the size
Membrane Separation of COFs 97

exclusion effect. COFs with extremely small PLDs (3.3 Å < PLD < 3.8
Å; the kinetic diameter of CH4 is 3.8 Å), such as 2D COFs in staggered
stacking modes and 3D COFs with interpenetrating configurations,
displayed a relatively good performance in CO2/CH4 separation. The
thermodynamic factor also plays a significant role in the separation
performance of COFs. An interaction between the functional sites of
COFs and CO2 would enhance the selectivity between CO2 and CH4.
For example, COFs with CO2 favorable interaction sites, like TpMA
with a hydroxy group (–OH), TpPa-NO2 with a nitro group (–NO2),
and ATFGCOF and NUS-2 with a carbonyl group (–C=O–), showed
a great potential in CO2/CH4, and COFs with 10 different functional
group were synthesized. It was found that –NH2 and –CH3 contribute
little to improving the CO2/CH4 separation efficiency. In contrast, 13
COFs with decoration of –F and –Cl are dominant in the top 26 of the
modified COFs with CO2/CH4 separation factors over 20 because of
the great adsorption selectivity.

Figure 3.11 Screening and design of COF membranes for CO2/CH4 separation.
Reprinted with permission from Ref. [34]. Copyright (2019) American Chemical
Society.

Experimental studies were also performed on CO2/CH4 separation


by COF-based membranes [35, 36]. In 2016, COF-based MMMs were
prepared by the incorporation of ACOF-1 in the commercial polymer
Matrimid 5218 for CO2/CH4 separation. Compared with the pure
polymer, MMM with 16 wt% of ACOF-1 exhibited an increased CO2
permeability (5.01 ¥ 10–13 mol m–2 s–1 Pa–1) with a slight increase
98 Gas Adsorption and Storage of COFs

of selectivity (~32). The improved permeability was attributed to


the additional pathways introduced by the porous ACOF-1. More
recently, ACOF-1 was also in situ grown on a porous a-Al2O3 substrate
for CO2/CH4 separation [35]. This continuous COF-based membrane
displayed a good thermal stability and superhigh values of both
permeability (9.9 ¥ 10–9 mol m–2 s–1 Pa–1) and selectivity (86.4). The
overall performance surpassed the latest Robeson upper bounds
for the CO2/CH4 gas pair. The high performance is attributed to the
excellent CO2 adsorption capacity of ACOF-1 and the stacked pores
with narrower aperture sizes. From these two works, it is obvious
that continuous COF-based membranes are superior in CO2/CH4
separation compared with discontinuous membranes (COF-Based
MMMs) because the advantages of the pore structures of COFs are
fulfilled in separation by the continuous membrane.
CO2/N2 separation has also attracted tremendous attention to
the capture of carbon dioxide to address the issue of global warming.
The pioneering separation of CO2/N2 by COF-based membranes
was conducted by a computational study in 2016 [37]. This study
concluded that the interlayer passages formed between the stacked
nanosheets have a “gate closing” effect on the selective transport
performance of the ultrathin COF-based membranes. Tuning the
stacking modes of COF nanosheets resulted in a high permeability
of CO2 and a good CO2/N2 selectivity. In 2017, an experimental
study of CO2/N2 separation by COF-based MMMs showed that the
incorporation of SNW-1 in polymers of intrinsic microporosity could
enhance the permeability (2.5 ¥ 10–10 mol m–2 s–1 Pa–1) of CO2 and
selectivity (22.5) of CO2/N2 [38].
On the basis of these highlighted experimental studies and the
applications of COFs, it can be derived that COFs with pore sizes
of less than or around 1 nm, such as CTF-O, CTF-1, ACOF-1, NUS-
2, COF-320, and SNW-1, have potential in gas separation, with
high selectivities. This indicates that pore size is one of the key
factors determining the permeability and selectivity of a COF-
based membrane, which was proved by a computational study
[34]. Therefore, further development in COF-based membranes for
gas separation relies on the synthesis of new COFs with relatively
small pore sizes. In addition, although simulation studies reveal that
the functional groups on COFs have a significant influence on the
performance of gas separation, no experimental study has explored
this to date.
Outlook and Conclusions 99

3.8 Outlook and Conclusions


COFs, which are constructed from organic linkers, are a new class
of crystalline porous materials comprising periodically extended
and covalently bound network structures. The intrinsic structures
and tailorable organic linkers endow COFs with low densities, large
surface areas, tunable pore sizes and structures, and facilely tailored
functionality, attracting increasing interest from a lot of different
fields. In this chapter, gas adsorption and storage of COFs were
described, from the principle of gas adsorption to the separation of
gases. Especially, this chapter focused on gas adsorption, storage, and
separation for COF-based membranes. To apply COFs in membrane
separation, various strategies, including blending, in situ growth,
layer-by-layer stacking, and IP, have been established to synthesize
COF-based membranes. These membranes show a relatively high
performance in gas separation, displaying a huge potential in
membrane technology. It is believed that COF-based membranes
will be intensively studies in the field of membrane technology for
years to come. Specifically, with the advances in separation, other
properties of COFs, including polarity, gas favorable properties, and
hydrophobicity, are also expected to improve for organic solvent
nanofiltration and gas separation, and a combination of these
properties on one COF would be a breakthrough in fabricating
multifunctional COF-based membranes. In addition, compared to
blending, in situ growth, layer-by-layer stacking, and IP result in
continuous COF-based membranes, achieving the full capability of
COFs in separation. In particular, ultrathin COF-based membranes can
be synthesized from layer-by-layer stacking and IP, demonstrating
superhigh permeability, making these two methods promising for
the future development of COF-based membranes.
Challenges still remain for the future development and application
of COF-based membranes in academia and industrial sectors. COFs
with pore sizes less than 1 nm are required to construct gas separation
membranes with high selectivity by using methods such as blending,
in situ growth, layer-by-layer stacking, and IP. However, there is a
limitation to these synthetic methods. A further narrowing of the
pore sizes of COFs might be achieved by anchoring side groups in the
inner walls of COFs via postmodification. Finally, the relatively high
100 Gas Adsorption and Storage of COFs

cost of COFs and the complicated and time-consuming fabrication


methods might hinder large-scale fabrication of membranes.
Therefore, low-cost COFs and cost-effective fabrication strategies
should be developed to control the cost of COF-based membranes in
industrial applications.

References

1. Furukawa, H., Yaghi, O. M. (2009). J. Am. Chem. Soc., 131, 8875–8883.


2. Taylor, H. S. (1931). J. Am. Chem. Soc., 53, 578–597.
3. Langmuir, I. (1916). J. Am. Chem. Soc., 38, 2221–2295.
4. Langmuir, I. (1918). J. Am. Chem. Soc., 40, 1361–1403.
5. Cao, D., Lan, J., Wang, W., Smit, B. (2009). Angew. Chem. Int. Ed., 48,
4730–4733.
6. Klontzas, E., Tylianakis, E., Froudakis, G. E. (2009). J. Phys. Chem. C,
113, 21253–21257.
7. Uribe-Romo, F. J., Doonan, C. J., Furukawa, H., Oisaki, K., Yaghi, O. M.
(2011). J. Am. Chem. Soc., 1338, 11478–11481.
8. Zhao, Y., Yao, K. X., Teng, B., Zhang, T., Han, Y. A. (2013). Energy Environ.
Sci., 6, 3684–3692.
9. Huang, N., Chen, X., Krishna, R., Jiang, D. (2015). Angew. Chem. Int. Ed.,
54, 2686–2990.
10. Huang, N., Krishna, R., Jiang, D. (2015). J. Am. Chem. Soc., 137, 7079–
7082.
11. Karan, S., Jiang Z., Livingston, A. G. (2015). Science, 348, 1347–1351.
12. Yang, H., Wu, H., Pan, F., Li, Z., Ding, H., Liu, G., Jiang, Z., Zhang, P., Cao, X.,
Wang, B. (2016). J. Membr. Sci., 520, 583–595.
13. Biswal, B. P., Chaudhari, H. D., Banerjee R., Kharul, U. K. (2016). Chem.-
Eur. J., 22, 4695–4699.
14. Yang, H., Cheng, X., Cheng, X., Pan, F., Wu, H., Liu, G., Song, Y., Cao, X.,
Jiang, Z. (2018). J. Membr. Sci., 565, 331–341.
15. Yang, H. Wu, Z. Yao, B. Shi, Z. Xu, X. Cheng, F. Pan, G. Liu, Z. Jiang, Cao, X.
(2018). J. Mater. Chem. A, 6, 583–591.
16. Peng, Y., Xu, G., Hu, Z., Cheng, Y., Chi, C., Yuan, D., Cheng, H., Zhao, D.
(2016). ACS Appl. Mater. Interfaces, 8, 18505–18512.
17. Wang, C., Li, Z., Chen, J., Li, Z., Yin, Y., Cao, L., Zhong, Y., Wu, H. (2017). J.
Membr. Sci., 523, 273–281.
References 101

18. Li, C., Li, S., Tian, L., Zhang, J., Su, B., Hu, M. Z. (2019). J. Membr. Sci., 572,
520–531.
19. Kandambeth, S., Biswal, B. P., Chaudhari, H. D., Rout, K. C., Kunjattu, S.
H., Mitra, S., Karak, S., Das, A., Mukherjee, R., Kharul, U. K. (2017). Adv.
Mater., 29, 1603945.
20. Sasmal, H. S., Aiyappa, H. B., Bhange, S. N., Karak, S., Halder, A., Kurungot,
S., Banerjee, R. (2018). Angew. Chem. Int. Ed., 130, 11060–11064.
21. Halder, A., Karak, S., Addicoat, M., Bera, S., Chakraborty, A., Kunjattu,
S. H., Pachfule, P., Heine, T., Banerjee, R. (2018). Angew. Chem. Int. Ed.,
130, 5899–5904.
22. Fan, H., Gu, J., Meng, H., Knebel, A., Caro, J. (2018). Angew. Chem. Int. Ed.,
57, 4083–4087.
23. Fan, H., Mundstock, A., Feldhoff, A., Knebel, A., Gu, J., Meng, H., Caro, J.
(2018). J. Am. Chem. Soc., 140, 10094–10098.
24. Fu, J., Das, S., Xing, G., Ben, T., Valtchev, V., Qiu, S. (2016). J. Am. Chem.
Soc., 138, 7673–7680.
25. Abraham, J., Vasu, K. S., Williams, C. D., Gopinadhan, K., Su, Y., Cherian,
C. T., Dix, J., Prestat, E., Haigh, S. J., Grigorieva, I. V. (2017). Nat.
Nanotechnol., 12, 546.
26. Wang, H., Zeng, Z., Xu, P., Li, L., Zeng, G., Xiao, R., Tang, Z., Huang, D.,
Tang, L., Lai, C. (2019). Chem. Soc. Rev., 48, 488–516.
27. Li, G., Zhang, K., Tsuru, T. (2017). ACS Appl. Mater. Interfaces, 9, 8433–
8436.
28. Ying, Y., Liu, D., Ma, J., Tong, M., Zhang, W., Huang, H., Yang, Q., Zhong, C.
(2016). J. Mater. Chem. A, 4, 13444–13449.
29. Kuehl, V. A., Yin, J., Duong, P. H., Mastorovich, B., Newell, B., Li-Oakey, K.
D., Parkinson, B. A., Hoberg, J. O. (2018). J. Am. Chem. Soc., 140, 18200–
18207.
30. Dey, K; Pal, M., Rout, K. C., Kunjattu, S. H., Das, A., Mukherjee, R., Kharul,
U. K., Banerjee, R. (2017). J. Am. Chem. Soc., 139, 13083–13091.
31. Matsumoto, M., Valentino, L., Stiehl, G. M., Balch, H. B., Corcos, A. R.,
Wang, F., Ralph, D. C., Marin ̃as, B. J., Dichtel, W. R. (2018). Chem, 4,
308–317.
32. Shinde, D. B., Sheng, G., Li, X., Ostwal, M., Emwas, A.-H., Huang, K.-W.,
Lai, Z. (2018). J. Am. Chem. Soc., 140, 14342–14349.
33. Kang, Z., Peng, Y., Qian, Y., Yuan, D., Addicoat, M. A., Heine, T., Hu, Z., Tee,
L., Guo, Z., Zhao, D. (2016). Chem. Mater., 28, 1277–1285.
102 Gas Adsorption and Storage of COFs

34. Yan, T., Lan, Y., Tong, M., Zhong, C. (2018). ACS Sustainable Chem. Eng.,
7, 1220–1227.
35. Fan, H., Mundstock, A., Gu, J., Meng, H., Caro, J. (2018). J. Mater. Chem. A,
6, 16849–16853.
36. Shan, M., Seoane, B., Rozhko, E., Dikhtiarenko, A., Clet, G., Kapteijn, F.,
Gascon, J. (2016). Chem.-Eur. J., 22, 14467–14470.
37. Tong, M., Yang, Q., Ma, Q., Liu, D., Zhong, C. (2016). J. Mater. Chem. A, 4,
124–131.
38. Wu, X., Tian, Z., Wang, S., Peng, D., Yang, L., Wu, Y., Xin, Q., Wu, H., Jiang,
Z. (2017). J. Membr. Sci., 528, 273–283.
Chapter 4

Heterogeneous Catalytic Application of


COFs

Covalent organic frameworks (COFs) are defined as highly porous


and crystalline polymers, constructed and connected via covalent
bonds, extending along two or three dimensions. As compared
with other porous materials, such as zeolite and active carbon, the
versatile and alternative constituent elements, chemical bonding
types, and characteristics of the ordered skeleton and pores of COFs
enable the creation of more COFs for diverse applications, including
gas separation and storage, optoelectronics, proton conduction, and
energy storage and, in particular, “catalysis.” The representative
candidates of next-generation catalysis materials due to their large
surface areas and accessible and size-tunable open nanopores, COFs
are suitable for incorporating external useful active ingredients, such
as ligands, complexes, and metal nanoparticles, and for substrate
diffusion. This chapter will introduce and discuss examples of COFs
for application in heterogeneous catalysts.

4.1 Heterogeneous Catalysts of COFs for C–C


Bond Coupling Reactions
The catalytic C–C coupling method is the most important core of the
current organic synthesis chemistry field. Synthetic projects (<85%)

Covalent Organic Frameworks


Atsushi Nagai
Copyright © 2020 Jenny Stanford Publishing Pte. Ltd.
ISBN 978-981-4800-87-7 (Hardcover), 978-1-003-00469-1 (eBook)
www.jennystanford.com
104 Heterogeneous Catalytic Application of COFs

in fine chemical production, pharmaceutical synthesis, agricultural


chemicals, household chemicals, etc., are involved in catalyzed C–C
coupling movements. Nowadays, catalytic C–C bond formation often
resorts to homogeneous transition-metal complex catalysis systems
such as the famous Pd-catalyzed Suzuki–Miyaura and Heck cross-
coupling reactions [1]. Behind the success of these systems is the
fact that these d-block transition-metal complexes commonly have
empty p* antibonding orbitals for p back-bonding and an electron
pair for coordination to sp2- or sp-hybridized carbon–carbon
multiple bonds [2]. Moreover, the centers of these transition-
metal complexes are adapted to oxidative addition and reductive
elimination. The low-valent transition metals are stabilized by a
variety of ligands bearing lone pairs of electrons and p* antibonding
orbitals. Thus, it is easy to choose a number of different chiral
ligands to render the transition-metal complex asymmetric activity
[3]. These chiral organometallic compounds are often very powerful
catalysts for asymmetric C–C coupling reactions. The realization of
this asymmetric cross-coupling has greatly enlarged the application
scope of transition-metal-catalyzed reactions [4]. Although covalent
organic frameworks (COFs) are new materials with a short history,
the research on COFs is evolving rapidly. More and more research
articles have revealed that COFs possess many unique properties,
significantly differentiating them from organic-inorganic hybrids,
such as metal-organic frameworks (MOFs) [5]. Why do chemists
favor COFs specifically for a catalytic C–C coupling reaction? The
reactions can be summarized as follows:
∑ From the perspective of industrial applications, a Pd-based
homogeneous catalyst has an intrinsic limitation. Especially
for drug synthesis, the biotoxicity of the residue from noble
metals has long worried and been criticized by the public. And
the efficiency of metal separation is unsatisfactory. Moreover,
the cost of separation comprises a large portion of the total
cost of drug production. Thus, developing heterogeneous
catalysts, especially encapsulating catalytically active species
such as Pd into the pore space of porous materials like COFs
with uncompromising catalytic efficiency, will solve the issue
of separating the toxic metal residues.
Heterogeneous Catalysts of COFs for C–C Bond Coupling Reactions 105

∑ From the perspective of catalytic efficiency, which is the


core issue in catalysis, after binding metal active species,
the metal catalytic center is confined by the COF’s pores. On
the one hand, this increases the difficulty of the substrate
diffusing and approaching the catalytic sites and the
products leaving the catalytic center, which would decrease
the turn-over frequency (TOF). However, on the other hand,
if the COF constituents and pore structures are fine-tuned,
the COFs will repulse solvent molecules and accelerate
substrate adsorption and product desorption. Compared
with a homogeneous catalyst, this acceleration will greatly
increase the efficiency and selectivity due to the pore effect.
Certainly, realizing this point is challenging. However, this is
the most attractive property of COFs, that is, to search for the
optimal COF constituent and pore structure. On one hand, this
optimized COF constituent and pore structure will settle and
accommodate the metal catalytic center to exert its catalytic
ability of oxidative addition and reductive elimination to the
maximum. On the other hand, this will fine-tune the selectivity
of substrate adsorption and product desorption rate. Most of
the following examples in this chapter are the breakthroughs
and proceedings in these aspects.
∑ Compared with other 2D or 3D pore structure materials,
such as the most similar MOF counterparts, the advantage
of COFs in catalyzing C–C coupling reaction is the absence of
another external transition-metal center (metal in the MOF)
to influence the encapsulated Pd catalysis, which avoids as-
induced side reactions.
∑ As mentioned above, the stability to aqueous, acidic, basic,
and organic solvents renders COFs inert to solvation and
decomposition. Due to the short history of COFs, their
applications and adaptions for C–C coupling reactions require
further development and improvement. Even so, the currently
reported examples, that is, the following recommended
ones, have already displayed bright prospects for significant
application and development.
The following examples and discussion are divided according to
the C–C coupling reaction types and several typical COFs’ catalyst
synthesis and catalytic performance are recommended.
106 Heterogeneous Catalytic Application of COFs

4.1.1 Suzuki–Miyaura Reaction


In 2011, Wang and coworkers demonstrated the first example of
a COF for catalysis application [6]. An imine-linked 2D COF (COF-
LZU1) was prepared by the condensation of 1,3,5-triformylbenzene
and 1,4-diaminobenzene in a 1,4-dioxane/aqueous acetic acid
solution in solvothermal conditions after liquid nitrogen flash
freezing, evacuation, and flame-sealing treatment (Fig. 4.1). The
COF-LZU1 took a 2D layered-sheet structure, with eclipsed nitrogen
atoms in adjacent layers, at a distance of about 3.7 Å. This imine-
linked COF-LZU1 demonstrated high coordination affinity to
Pd(OAc)2 due to strong nitrogen-palladium interaction. Only by
simple impregnation, Pd(OAc)2 was effectively incorporated into
COF-LZU1 channels and pores. This palladium metal incorporation
has no effect on the long-ordered structure of COF-LZU1. However,
the intensity of powder X-ray diffraction (PXRD) and the Brunauer–
Emmett–Teller (BET) surface area were reduced to a certain degree.
The Pd-incorporated COF displayed very enhanced catalytic activity,
a shorter reaction time, and a lower catalyst load than Pd(OAc)2 and
its Pd COF analogues in a typical Suzuki–Miyaura cross-coupling
reaction (Fig. 4.2). Moreover, this COF did not lose its catalytic
activity at all even after four cycles of reuse. This superior catalytic
activity and stability after reutilization rendered it a very promising
catalyst for a classical Suzuki–Miyaura reaction.

N N
N

CHO N
N
N
N N
OHC CHO 3M AcOH
Dioxane OAc
1.8 nm N Pd(OAc) 2 Pd
OAc
+ ca. 3.7 Å
120 r& , 3d
N DCM, r.t. N N

H2 N NH2
N
Pd/COF-LZU1
N
N

N
N N N
N
N

COF-LZU1

Figure 4.1 Schematic representation of the synthesis of COF-LZU1 and Pd/


COF-LZU1.
Heterogeneous Catalysts of COFs for C–C Bond Coupling Reactions 107

0.5 mol% Pd/COF-LZU1


R-X + B(OH)2 R
K2CO3, p-xyxlene

Entry R X Time (h) Yield (%)

1 MeO I 3 96

2 O2N I 2 97

3 Br 3 97

4 O2N Br 3 97

5 Me Br 3 97

6 Br 2.5 98

7 Br 2.5 97

8 MeO Br 4 96

Figure 4.2 Catalytic activity test of Pd/COF-LZU1 in the Suzuki–Miyaura


coupling reaction.

After the report by Wang and group, Jiang et al. also reported
that a porphyrin-based H2P-Bph COF could incorporate Pd(OAc)2
species to efficiently catalyze a Suzuki reaction, with excellent
yields, ranging from 97.1% to 98.5% [7]. This COF was prepared via
condensation between 5,10,15,20-tetra(p-amino-phenyl)porphyrin
and 4,4¢-biphenyldialdehyde in an EtOH/mesitylene/acetic acid
aqueous solution at 120°C for 3 days in vacuum (Fig. 4.3). This
COF was nitrogen-rich due to the porphyrin unit’s tetrapyrrole
group and imine C=N bonds. These excess nitrogen groups aced as
effective docking sites for Pd(OAc)2 complexation. Solid state–13C–
nuclear magnetic resonance (NMR), Fourier-transform infrared
spectroscopy, X-ray photoelectron spectroscopy, and inductively
coupling plasma–atomic emission spectroscopy characterizations all
confirmed the inclusion of Pd(OAc)2. The porosity and crystallinity
decreased to a certain degree after Pd incorporation. This Pd-H2P
COF showed superior catalytic activity for Suzuki cross-coupling
108 Heterogeneous Catalytic Application of COFs

reactions for a variety of bromoarenes and phenylboronic acid,


forming biphenyl derivatives with yields ranging from 97.1% to
98.5%, surpassing the yields of its Pd MOF and Pd/C counterparts
(the yield of 4-methosybiphenyl product was 65.0%, while the
yield for Pd/MOF was 84.1% [8]). This report was inspiring and
encouraging since the authors showed that a porphyrin-based
catalyst could not only promote radical or carbene-based oxidation
but also accelerate the cross-coupling reaction by incorporating a Pd
complex into a COF support.

N N

H 2N NH2 N
N HN HN
N N
N N
NH NH N
N
N
H
N N
H
N

N
EtOH / Mesitylene/ 6M AcOH aq N
H2N NH2
120 , 3d
+

OHC CHO
N N

N HN N HN
N N
N N
NH NH N
N

N N

H2P-Bph-COF

N
OAc Pd(OAc)2, CH2Cl2
Pd
OAc
r.t. for 24 h
N

Pd/H2P-Bph-COF

Figure 4.3 The development of H2P-Bph COF and Pd/H2-Bph COF.

Apart from imine and porphyrin COFs, a triazine-based COF could


also serve as an efficient amphiphilic support for Pd(0) nanoparticle
(NP) deposition [9]. The triazine COF was synthesized by the
condensation reaction of 4,4¢,4¢¢-(1,3,5-triazine-2,4,6-triyl)tris(oxy)
tribenzenealdehyde and benzene-1,4-diamine in 1,4-dioxane/
mesitylene in the presence of aqueous acetic acid as a catalyst by
heating at 120°C for 72 h. The as-prepared triazine COF contained
both a long and flexible appendage and a nitrogen- and oxygen-rich
skeleton. The nitrogen-rich moiety was responsible for the facile in
Heterogeneous Catalysts of COFs for C–C Bond Coupling Reactions 109

situ reduction of Pd2+ to Pd(0) without any external oxidants. The


ether and imine moieties had strong interaction with Pd(0) NPs by
stabilizing and dispersing them on the COF. This triazine COF was
very effective in catalyzing the multifold Heck and Suzuki–Miyaura
cross-coupling reactions. Unprecedentedly high turn-over numbers
(TONs) and TOFs of the catalyst for multifold Heck reactions were
provided as compared with its homogeneous Pd, Pd MOF, and
Pd/C counterparts. Extremely short reaction times (1.5 h) and an
excellent isolated yield (up to 99%) was observed. The recyclability
of the catalyst showed apparently no loss in activity. This resulted in
a viable strategy to dock noble metal NPs into a COF without external
reducing agents, applying the triazine monomer itself as a reductant
for efficient Heck and Suzuki–Miyaura cross-coupling reactions by
appending long and flexible groups, leading to the formation of an
amphiphilic structure and incorporating more nitrogen and oxygen
atoms.
Two different kinds of Schiff-base COFs as triazine-based imine
and b-ketoenamine-linked COFs (TAT-DHBD and TAT-TFP) could
be synthesized from 1,3,5-tris(4-aminophenyl)triazine (TAT)
and 2,5-dihydroxybenezne-1,4-dicarboxaldehyde (DHBD) or
1,3,5-triformylphloroglucinol (TFP) under solvothermal conditions
in a dioxane/mesitylene mixture (Fig. 4.4) [10]. Including Pd(OAc)2
into the pore space and between the interlayer region of the 2D
sheets, Pd-loaded TAT-DHBD and TAT-TFA COFs were prepared.
Sequentially, the Pd-loaded COFs were synthesized by the NaBH4
reduction of Pd(II) to Pd(0) NPs. Four Pd-loaded COFs illustrated
very good catalytic activity in a Suzuki–Miyaura cross-coupling
reaction between bromobenzene and phenylboronic acid. The
best performance was obtained with Pd(0)-TAT-TFP COF catalyst,
which provided excellent conversion and yield for electron-rich and
electron-deficient bromobenzenes. The most appropriate substrate
was 4-cyanobromobenzene with phenylboronic acid, which showed
an almost quantitative formation of 4-cyanobisphenyl only after
4 h by this Pd(0)-TAT-TFA COF catalyst. As a result, these Pd COF
catalysts all displayed very good stability and reusability without
apparent leaching of Pd and loss of activity.
110
Heterogeneous Catalytic Application of COFs

Figure 4.4 Synthesis of TAT-DHBD (1) and TAT-TFP (2) and their Pd-embedded COFs. Conditions: (a) dioxane/mesitylene, 6 M AcOH, 120°C,
3 days; (b) Pd(OAc)2, CH2Cl2, 24 h, RT; (c) NaBH4, MeOH, 48 h, RT. Reproduced with permission from Ref. [10]. Copyright (2017), John Wiley
and Sons.
Heterogeneous Catalysts of COFs for C–C Bond Coupling Reactions 111

Figure 4.5 (a) Synthesis of thio-COF and (b) schematic representation of the
synthesis of thio-COF-supported PtNPs@COF. Top and side views of the energy-
minimized models of thio-COF (yellow, S; blue, N; gray, C; red, O) are shown
in (b). Reprinted with permission from Ref. [11]. Copyright (2017) American
Chemical Society.

More recently, a thioether-containing COF was reported to


give excellent support to ultrafine Pt and Pd NPs, providing a very
narrow size distribution and superior stability [11]. Inspired by their
designed shape-persistent thioether-containing organic cage, which
hosted ultrafine Pt and Pd NPs with a very narrow size distribution,
Zhang and coworkers elaborately designed and prepared the PtNPs@
COF and PdNPs@COF by condensing a trialdehyde and a thioether-
containing diamine in dioxane/mesitylene/aqueous acetic acid
solution at 120°C for 3 days (Fig. 4.5). The as-formed thio-COF was
further complexed with K2PtCl4 and K2PdCl4 in an aqueous solution
and then reduced by a methanolic NaBH4 solution to PtNPs@COF
and PdNPs@COF. According to a series of structural, morphological,
and compositional characterization, the authors demonstrated
that ultrafine Pd and Pt NPs were uniformly incorporated into the
pore space of the thio-COF. The thioester functional group provided
strong metal-sulfur interaction to stabilize the ultrafine noble metal
NPs to prevent them from aggregation. Moreover, the long-range
ordered pore-channel structure also assisted in the stabilization of
the residing external NPs. The as-formed PtNPs@COF and PdNPs@
COF illustrated very good catalytic activity toward 4-nitrophenol
reduction and Suzuki–Miyaura cross-coupling between a variety
of aryl halides and phenylboronic acid. PdNPs@COF provided
excellent NMR yields (up to 99%) for the cross-coupling of 4-methyl-
iodobenzene and phenylboronic acid to form 4-methylbiphenyl.
Furthermore, these two COF-based catalysts presented excellent
stability and recyclability after simple centrifugation or natural
settling for the cycling test, almost no decrease in conversion and
112 Heterogeneous Catalytic Application of COFs

yield in catalytic performance after the fifth cycle test, and no


noticeable leaching of metal NPs. This example is significant because
it was the first use of a thioether functional group to stabilize the
narrowly distributed ultrafine noble metal NPs for effective cross-
coupling reactions in a COF support.

4.1.2 Heck, Sonogashira, and Silane-Based Cross-


Coupling Reactions
A Heck cross-coupling reaction is the Pd-catalyzed coupling of
aromatic or vinylic halides with an unsaturated olefinic C=C bond,
as shown in Scheme 4.1 [12]. The catalytic cycle comprises oxidative
addition of aromatic halides, coordination with an olefinic C=C bond,
cis-insertion, cis-b-hydride elimination, and reductive elimination.
This reaction requires a Pd(0) active species. The advantage of a
Heck reaction is its high regio- and stereoselectivity. In contrast,
the disadvantage is the costly Pd catalyst. The Sonogashira coupling
reaction is the Pd/Cu cocatalyzed coupling of a terminal alkyne with
aryl or vinylic halides to a Pd(0) center. Then, a Cu-amine complex
mediates the transmetalation reaction. The last step is a reductive
elimination, releasing the coupling product and regenerating the
Pd(0) catalyst. A Cu-amine complex acts as a cocatalyst to assist in
deprotonating the alkyne substrate.

Scheme 4.1 Pd-catalyzed Heck reaction and its catalytic cycle [12].

Banerjee and coworkers demonstrated that the introduction of


a large number of nitrogen and oxygen atoms into the skeleton of a
COF would reinforce its stability when metal NPs or complexes are
deposited. They exhibited that the condensation reaction of TFP and
Heterogeneous Catalysts of COFs for C–C Bond Coupling Reactions 113

paraphenylenediamine in mixed solvents (mesitylene/dioxane) in


the presence of aqueous acetic acid at 120°C for 3 days in an inert
atmosphere would generate a nitrogen- and oxygen-rich COF (Fig.
4.6) [13]. The as-synthesized imine COF is further deposited with
Pd(II) complexes by immersing it in a methanol solution containing
Pd(OAc)2. Pd(0) NPs are generated from the in situ reduction of
Pd(II) COF with NaBH4. The authors demonstrated that Pd(0) and
Pd(II) inclusion does not greatly change the crystallinity and flower-
like morphology of the COF. Further, the Pd(0) COF shows superior
catalytic activity toward Heck- and Sonogashira-type reactions. The
Pd(II) COF demonstrated a considerably robust catalytic ability for
intramolecular C–H activation and further C–C coupling reaction,
synthesizing 9H-carbazole from diphenylaniline. This report
manifested its significance in incorporating Pd(0) and Pd(II) into the
same COF and applied the metal–COF composite in highly selective
C–C coupling and C–H activation transformations.

Figure 4.6 Development of Pd(II)- and Pd(0)-doped COFs (i.e., Pd(II)@TpPa-1


and Pd(0)@TpPa-1) and summary of their catalytic activity toward Sonogashira,
Heck, and oxidative biaryl couplings. The doped Pd(0) nanoparticles are probably
situated on the TpPa-1 surface. The scheme is to represent the synthesis and
organization of Pd nanoparticles on COF (TpPa-1) and it is not exactly fit to
scale. Republished with permission of Royal Society of Chemistry, from Ref. [13],
copyright (2014); permission conveyed through Copyright Clearance Center, Inc.
114 Heterogeneous Catalytic Application of COFs

Figure 4.7 Pore surface engineering strategy used to modulate the nitrogen
content of the 2D imine–linked COFs and scheme for regulated Pd(OAc)2
coordination on bipyridine and imine groups. Reprinted from Ref. [14], Copyright
(2016), with permission from Elsevier.

Chai and coworkers reported that two different nitrogen


ligands bipyridine and imine could be incorporated into a single
COF skeleton to provide differentiated Pd coordinating sites [14].
The authors designed and prepared X% bpy COF by condensing
x% 2,2-bipyridine-5,5¢-dicarbaldehyde, 100-X% 4,4¢-biphenyl
dialdehyde, and 4,4¢,4¢¢,4¢¢¢-(pyrene-1,3,6,8-tetrayl)tetraaniline
(PyTTA) building blocks in mesitylene/dioxane mixed solvents in
the presence of acetic acid (3 M) at 120°C for 3 days, as shown in
Fig. 4.7. The as-formed X% bpy COF contained two different kinds
of nitrogen ligands, bipyridine and imine. With further Pd(OAc)2
Heterogeneous Catalysts of COFs for C–C Bond Coupling Reactions 115

complexation experiment results, they demonstrated that Pd(OAc)2


coordinated with both bipyridine and imine units, but in different
regions. Pd(OAc)2 combined with bipyridine mainly dwelled in the
pore space, while Pd(OAc)2 joined with imine resided between the
adjacent layers of the 2D COF. Furthermore, the authors demonstrated
that these Pd@bpy COFs displayed a very good catalytic performance
toward the classical Pd-catalyzed Heck reaction between a series of
aryl halides and styrene. The Pd(II)@75% bpy COF showed the best
catalytic ability, providing >90% yield after four consecutive runs.
The superior activity for the Heck reaction of these Pd@bpy COF
catalysts was attributed to the uniform dispersion and the ultrahigh
loading of Pd(OAc)2.
On other occasions, Pd(0) NPs could be generated in situ by
choosing a predesigned metal-anchored building block. Initially,
a 2,2-bipyridine-5,5¢-diamine palladium chloride (Bpy-PdCl2)
complex was formed by a condensation reaction between PdCl2 and
2,2¢-bipyridine-5,5¢-diamine (Fig. 4.8) [15]. Then, via Schiff-base
condensation between Bpy-PdCl2 and Tp, a Pd@TpBpy COF was
prepared. This in situ generated Pd@TpBpy COF did not require
any external reducing agents for Pd(II) reduction to Pd(0) NPs.
The as-formed Pd@TpBpy COF represented excellent catalytic
performance toward a tandem C–C and C–O bond formation reaction
between 2-bromophenol and phenylacetylene. The Pd@TpBpy COF
heterogenous catalyst promoted tandem cyclization and provided
various substituents on the phenylacetylene or 2-bromophenol.
Moreover, this ketoenamine and bipyridine anchored Pd@TpBpy
COF showed very good stability and recyclability.
A triazine-based COF-SUD1-palladium hybrid could be an active
catalyst for the cross-coupling between silanes and aryliodides
[16]. The COF-SDU1 was prepared via an imine condensation from
tri-(4-formacylphenoxy)-1,3,5-triazine and p-phenylenediamine
in o-dichlorobenzene/n-butanol/6 M AcOH with heating at 85°C
for 7 days. The as-obtained COF contained two kinds nitrogen
ligand azine and imine, which are both suitable coordinating sites
for Pd(II). After a simple solution infiltration, monodispersing
Pd(II) ion was docked in the 2D COF. This Pd(II)-COF-SDU1 showed
excellent catalytic activity for a one-pot silane oxidation reaction.
The transformation of phenylsilanes with aryliodides to biphenyls
was effectively catalyzed in a methanolic solution. A variety of
116 Heterogeneous Catalytic Application of COFs

electron-rich and electron-deficient aryliodides were cross-coupled


with phenylsilanes and gave good to excellent yields. This COF also
displayed good recyclability and reusability without detectable Pd
leaching and loss of activity.

Figure 4.8 Various strategies are used to convert a homogeneous catalyst into
a reversible heterogeneous version ( a–c) Reported approaches and (d) new
approach. (e) Synthesis details of the in situ generation of highly dispersed Pd
nanoparticles in the TpBpy skeleton. The size of the Pd nanoparticles and the
COF pore organization are not exactly to scale. (f) Comparison of PXRD patterns
of simulated TpBpy (black) with experimental TpBpy (blue), Pd nanoparticles
(cyan), and experimental Pd@TpBpy (red). (g) Schematic representation of
tandem catalysis by Pd@TpBpy. Reprinted with permission from Ref. [15].
Copyright (2017) American Chemical Society.

The important point of COFs’ successful application in Heck and


Sonogashira cross-couplings are summarized as follows: low Pd
loading, high catalytic efficiency, and facile desorption of halide ions.

4.2 Chiral Heterogeneous Catalysts of COFs for


Asymmetric C–C Bond Coupling Reactions
In 2014, Jiang et al. developed a chiral-organocatalytic COF prepared
through pore surface engineering, whose system was proposed in
2011 [17, 18]. This organocatalytic COF was prepared by integrating
a chiral pyrrolidine unit into the main chain of the porphyrin-imine
COF, as shown in Fig. 4.9. Initially, they introduced an ethynyl group
into the imine moiety. By the facile alkyne-azide click reaction, they
Chiral Heterogeneous Catalysts of COFs for Asymmetric C–C Bond Coupling Reactions 117

anchored the triazole-substituted pyrrolidine ring to the imine by a


post-treatment catalyzed by CuI. The as-prepared COF demonstrated
its activity for organocatalysis and displayed a variety of advantages
in catalyzing an enantioselective asymmetric Michael addition
reaction. The conversion of the reactants was much accelerated by
this organocatalytic COF; however, moderate diastereoselectivity
(d.r.) and enantiomeric excess (e.e.) values were obtained (Fig. 4.10).
The most important was that this was the first time COFs realized
enantioselectivity control in a catalytic organic synthetic reaction.
Moreover, the design and synthesis of another organocatalytic
COF was also reported by Jiang et al. [19]. This was a milestone event
for a COF applied for an organic synthetic purpose. They discovered
that the introduction of a methoxy group in the edge unit would
greatly increase the stability of the COF against humidity, acidity, and
basicity since the methoxy group reinforced the interlayer interaction.
The COF was prepared by condensing triphenylbenzenetriamine
(TPB), 2,5-bis(2-propynyloxy)terephthalaldehyde (BPTA) and
2,5-dimenthoxyterephthaladehyde (DMTA) (Fig. 4.11). The [HC∫C]x-
TPB-DMTA COFs with alkynyl groups were further transformed to
[(S)-Py]x-TPB-DMTP COFs by a postsynthetic click reaction with
(S)-2-(azideomethyl)pyronylidine catalyzed by a CuI catalyst. The
as-formed [(S)-Py]x-TPB-DMTA COFs demonstrated extremely
strong stability when soaked in boiling water, 12 M HCl, and a
14 M NaOH solution. The COF displayed very little reduction in
crystallinity and porosity after these harsh condition treatments.
The intensities of the X-ray diffraction peaks showed no apparent
decrease and the BET and Langmuir surface area almost remained
unchanged. Besides extraordinary stability and uncompromised
crystallinity and porosity, the most significant point of this COF was
its functionality to catalyze chiral asymmetric organic reactions.
For a typical organocatalytic asymmetric Michael addition between
unactivated cyclohexanones and nitrostyrenes, this COF showed
superior catalytic activity compared with its homogeneous
counterpart (Fig. 4.12). The COF required only half the time required
by the organocatalyst. After a five-cycle reusability test, this COF did
not show any apparent loss in catalytic activity. The shining point
of this COF was its combination of unprecedented stability, good
crystallinity, and highly developed mesoporous structure with very
powerful catalytic ability, accelerated reaction kinetics, and excellent
yields and d.r. and e.e. values for chiral asymmetric Michael addition
in an aqueous solution (Fig. 4.11d).
118
Heterogeneous Catalytic Application of COFs

Figure 4.9 (A) The general strategy for the pore surface engineering of imine-linked COFs via a condensation reaction and click chemistries
(the case for X = 50 was exemplified). (B) A graphical representation of [Pyr]x-H2P COF with catalytic sites of different densities on the pore
walls (gray: carbon; red: nitrogen; green: oxygen; purple: carbon atoms of the pyrrolidine units; hydrogen is omitted for clarity). Republished
with permission of Royal Society of Chemistry, from Ref. [18], copyright (2014); permission conveyed through Copyright Clearance Center,
Inc.
Chiral Heterogeneous Catalysts of COFs for Asymmetric C–C Bond Coupling Reactions 119

Figure 4.10 Comparison of the pyrrolidine control, amorphous nonporous


polymers, and COFs as catalysts for a Michael addition reaction. Republished
with permission of Royal Society of Chemistry, from Ref. [18], copyright (2014);
permission conveyed through Copyright Clearance Center, Inc.

Apart from above-mentioned chiral organocatalyst–


incorporated COF for asymmetric organic transformations and
noble metal–incorporated COFs for achiral organic transformations,
a homochiral organocatalytic COF skeleton could also be a perfect
support for organometallic Pd species with chiral ligands [20]. The
Pd NPs dispersed in a chiral COF (CCOF) skeleton could effectively
be a heterogeneous catalyst for asymmetric Henry and reductive
Heck reactions, providing excellent isolated yields and e.e. values.
The CCOF was prepared by a condensing reaction between cyanuric
chloride and S-(+)-2-methylpiperazine with K2CO3 in a dioxane
solution. Further, Pd(0) NPs were included in the CCOF by in situ
reduction of a Pd(NO3)2 methanolic solution with NaBH4 in a CCOF
aqueous suspension (Fig. 4.13). The as-formed Pd@CCOF was
uniformly dispersed between the CCOF 2D layer, not residing in the
pore space due to the large size of the Pd NPs (2–5 nm) compared
with the CCOF micropore (1.5 nm). The incorporation of Pd NPs
greatly influenced the porous structure of the CCOF, enlarging its BET
surface area and pore size. Furthermore, the authors demonstrated
the synergistic catalytic activity by subjecting the Pd@CCOF catalyst
in typical Henry and reductive Heck reactions. To the authors’
delight, the Pd@CCOF catalyst displayed extremely superior
catalytic ability toward these two reactions, providing an excellent
yield (up to 99%) and a perfect e.e. value (up to 97%). To the best
of my knowledge, this was the first example of a heterogeneous
120 Heterogeneous Catalytic Application of COFs

Figure 4.11 Synthesis and structure of stable crystalline porous COFs. (a)
Synthesis of TPB-DMTP COF through the condensation of DMTA (blue) and
TAPB (black). Inset: The structure of the edge units of TPB-DMTP COF and the
resonance effect of the oxygen lone pairs that weaken the polarization of the
C=N bonds and soften the interlayer repulsion in the COF. (b) Graphic view of
TPB-DMTP COF (red, O; blue, N; gray, C; hydrogen is omitted for clarity). (c)
Synthesis of chiral COFs ([(S)-Py]x-TPB-DMTP COFs, x = 0.17, 0.34, and 0.50; blue,
DMTA; black, TAPB; red, BPTA; green, (S)-Py sites) via channel-wall engineering
using a three-component condensation followed by a click reaction. Reprinted
by permission from Springer Nature Customer Service Centre GmbH: Springer
Nature, Nature Chemistry, Ref. [19], copyright (2015).

Pd-catalyzed chiral asymmetric reductive Heck reaction. The Pd@


CCOF illustrated satisfactory recyclability and reusability after
five cycles of reuse, without apparent loss of catalytic activity,
Chiral Heterogeneous Catalysts of COFs for Asymmetric C–C Bond Coupling Reactions 121

providing an isolated yield of up to 93% and an e.e. value of 91%


for the fifth cycle test. This development of Pd@CCOF manifested
its importance in that it was the first COF combining a noble metal
catalyst with a chiral organocatalyst in a single COF carrier to fulfill
asymmetric transformations previously catalyzed by homogeneous
organometallic compounds with complex and elaborately designed
chiral ligands.

Figure 4.12  Scope  of  reactants.  Different  β-nitrostyrene  derivatives 


investigated for the Michael reactions catalyzed with chiral COFs, their products,
e.e. yields and d.r. values (red, cyclohexanone; green, newly formed C–C bond;
blue, nitrostyrene derivatives). R, substituent H, Cl, Br, Me, or OMe. Reprinted
by permission from Springer Nature Customer Service Centre GmbH: Springer
Nature, Nature Chemistry, Ref. [19], copyright (2015).

More recently, Cui and coworkers reported a multivariate


strategy to prepare CCOFs with controlled crystallinity and stability
for asymmetric catalysis [21], in which crystallizing mixtures of
triamines with and without chiral organocatalysts and a dialdehyde
produced a series of two- and three-component 2D COFs (Fig. 4.14).
These COFs were found to be efficient heterogeneous catalysts in
122 Heterogeneous Catalytic Application of COFs

Figure 4.13 (a) Henry reactions of nitromethane with different aromatic


aldehydes catalyzed by Pd@CCOF and (b) Pd NP-loaded homochiral COF for
heterogeneous asymmetric catalysis. Reprinted with permission from Ref. [20].
Copyright (2017) American Chemical Society.

catalyzing the asymmetric amino-oxylation reaction, aldol reaction,


and Diels–Alder reaction, with the stereoselectivity comparable
to or surpassing their homogeneous analogues. Moreover, they
developed a metal-directed synthesis strategy through which chiral
Zn(salen)-based COFs can be obtained by the imine-condensations
of enantiopure 1,2-diaminocyclohexane and C3-symmetric
trisalicylaldehydes with one or no 3-tert-butyl group, as shown
in Fig. 4.15. They found that the bulky tributyltrisalicylaldehydes
containing CCOF possessed superior stability. These two COFs
were further complexed with a variety of metal ions, such as Zn,
Fe, Mn, Cr, V, and Co. The Mn(salen) modules in the COFs, Mn-
salen COFs, demonstrated very good catalytic activity to a variety
of chiral asymmetric organic synthetic reactions, such as V-salen
COF–catalyzed cyanation reaction of aldehydes, Diels–Alder
cycloaddition reaction catalyzed by a Co-salen COF, epoxidation
Chiral Heterogeneous Catalysts of COFs for Asymmetric C–C Bond Coupling Reactions 123

catalyzed by Fe-salen and Mn-salen COFs, aminolysis opening of


epoxides catalyzed by a Cr-salen COF, and the tandem one-pot
heterogenous catalyst. These M(salen) CCOF–catalyzed reactions
not only provided satisfactory yields but also realized excellent
controls of enantioselectivity and diastereoselectivity. Moreover,
good recyclability and reusability was proved in the case of V-salen
COF–catalyzed cyanation of aldehydes, which showed almost no loss
of enantioselectivity and conversion after five runs of cycle tests.

Figure 4.14 Synthesis of chiral COFs. Reprinted with permission from Ref. [21].
Copyright (2017) American Chemical Society.

The construction of functional COFs by a bottom-up strategy is


a relatively difficult task because it must simultaneously meet the
requirements for crystallinity and functionality. Even so, several
attempts have been shown to be successful for the synthesis of COFs
bearing catalytic sites with this method. For example, a sulfonated
building lock, 2,5-diaminobenzenesulfonic acid (DABA), was used
to construct a sulfonated COF together with another building block,
TFP [23]. The as-synthesized TFP-DABA was found to be a highly
124 Heterogeneous Catalytic Application of COFs

efficient acid catalyst for fructose conversion with remarkable yields


(97% for 5-hydroxymethylfurfural and 65% for 2,5-diformylfuran),
good chemoselectivity, and good recyclability. Cui and coworkers
reported the synthesis of two 2D CCOFs by direct imine
condensations of enantiopure TADDOL-derived tetraaldehydes
with 4,4¢-diaminophenylmethane, which, after treatment with
Ti(OiPr)4 show highly catalytic activity for the asymmetric addition
of diethylzinc to aldehydes [24].

Figure 4.15 Synthesis of chiral COFs. Reprinted with permission from Ref. [22].
Copyright (2017) American Chemical Society.

In a similar way, Wang et al. reported a facile strategy for the


direct construction of chiral-functionalized COFs using chiral (S)-
4,4¢-(2-(pyrrolidine-2-yl)-1H-benzimidazole as a building block
(Fig. 4.16a) [25]. Two CCOFs, LZU-72 and LZU-76, are obtained based
on the building block, as shown in Fig. 4.16. The catalytic activity
of chemically stable LZU-76 was evaluated in the asymmetric aldol
reaction, and the reaction afforded the desired aldol products with
excellent enantioselectivity (88.4:11.6-94.0:60 e.r.), comparable
Heterogeneous Bimetallic or Bifunctional Catalysts of COFs 125

to the enantioselectivity (93.1:6.9) of a model catalyst, (S)-4,7-


diphenyl-2-(pyrrolidin-2-yl)-1H-benzimidazole (Fig. 4.16b).

Figure 4.16 (a) Schematic of the direct formation of chiral LZU-72 and LZU-
76 and (b) catalytic activity check of LZU-76 in an asymmetric aldol reaction.
Reprinted with permission from Ref. [25]. Copyright (2016) American Chemical
Society.

4.3 Heterogeneous Bimetallic or Bifunctional


Catalysts of COFs
A bifunctional organocatalytic COF was initially designed and
realized by Banerjee et al., which was stable in aqueous and acidic
conditions [26]. This COF was synthesized through the Schiff-base
condensation between 2,3-dihydroxyterephthalaldehyde (2,3-
Dha) and 5,10,15,20-tetrakis(4-aminophenyl)-21H,23H-porphyrin
unit (Tph) in the mixed solvents of dichlorobenzene (o-DCB) and
dimethylacetamide in the presence of 6 M acetic acid as a catalyst.
The as-synthesized COF showed unprecedented stability in water
and an acidic solution due to the catechol group, the presence of
trans conformation of imine bonds, and intermolecular hydrogen
bonding (–OH∑N=C; D = 2.579, d = 1.858 Å, y = 146.11), which have
been confirmed from the monomer crystal structure (Fig. 4.17).
126 Heterogeneous Catalytic Application of COFs

Figure 4.17 (a) The synthesis of 2,3-DhaTph and 2,3-DmaTph by the


condensation of Tph and 2,3-Dha/2,3-Dma. The catalytically active porphyrin
and catecholic –OH groups are shown in coral and cyan colors, respectively.
An ORETP diagram of 2,3-DhaTph and 2,3-DmaTph monomer units. (b) The
catalytic activity toward acid-base catalyzed reaction with various reactions.
Republished with permission of Royal Society of Chemistry, from Ref. [26],
copyright (2015); permission conveyed through Copyright Clearance Center, Inc.

Moreover, since the COF was considered from the Dha unit
with weak acidic catechol groups and Tph group containing basic
pyrrole groups and imine C=N bonds, this COF possessed acidic and
basic sites, providing it a promising bifunctional heterogeneous
catalyst. In a model cascade deacetalization–Knoevenagel reaction,
this COF demonstrated an excellent isolated yield up to 96%. The
deacetalization of benzaldehydedimethylacetal was catalyzed by the
acidic sites of this COF catalyst, while a further Knoevenagel reaction
between benzaldehyde and amlonitrile was effectively accelerated
by the basic sites of the DhaTph COFs. This discovery manifested its
significance in that it was the first stable bifunctional COF catalyst in
water and an acidic solution.
With the exception of single metal–deposited COFs, bimetallic
docked COFs were designed, synthesized, and applied as effective
catalysts for a Heck–epoxidation tandem reaction. Mn and Pd
bimetallic docking to a bipyridine-imine COF could be realized by
a programmed synthetic procedure (Fig. 4.18) [27]. First, a Py-2,2¢-
BPyPh COF skeleton was constructed via a Schiff-base condensation
reaction between PyTTA and 2,2¢-bipyridyl-5,5-dialdehyde. The as-
formed COF contained two different types of organic transformations,
that is, Pd-catalyzed Heck cross-coupling reaction and Mn-catalyzed
epoxidation reaction. The COF transformed iodobenzene and
styrene to trans-stilbene oxide in a tandem reaction. Initially,
Pd(OAc) incorporated in COF transformed iodobenzene and styrene
Heterogeneous Bimetallic or Bifunctional Catalysts of COFs 127

into trans-stilbene with an excellent yield (up to 95%), while the Mn


in COF catalyzed the epoxidation reaction of trans-stilbene to trans-
stilbene oxide in an almost quantitative yield (99%). The control
group experiments proved that Mn@Py-2,2¢-BPyPh COF and Pd@Py-
2,2¢-BPyPh COF alone could only catalyze the separate epoxidation
and Heck reaction. This finding was important since it demonstrated
that by choosing elaborately designed ligands, different metal
species could be incorporated into a single COF skeleton to fulfill
different genres of organic transformation by certain metals.

Figure 4.18 (a) Schematic representation of Mn/Pd bimetallic docked COFs


prepared via a programmed synthetic procedure, (b) top view and (c) side view
of Mn/Pd@Py-2,2¢-BPyPh COF, and (d) appearances and (e) PXRD pattern of the
COFs before and after metallic loading treatment. Republished with permission
of RSC Publishing, from Ref. [27], copyright (2016); permission conveyed
through Copyright Clearance Center, Inc.

Besides Mn and Pd codocking for a bimetallic COF catalyst, Rh


and Pd bimetallic docking could also be realized via this 2D BPy COF
(Fig. 4.19) [28]. By condensing PyTTA and 2,2¢-bipyridy-5,5-
dialdehyde and 100-X% 4,4¢-biphenyldialdehyde in a three-
component solvent, a series of structurally tunable 2D COFs were
prepared. The authors demonstrated that by a further solution-
infiltration method, Pd(OAc)2 and Ph(COD)Cl were sequentially
incorporated into the COF skeleton in a programmed synthetic
procedure. Pd(OAc)2 dispersed in the interlayer space coordinated
128 Heterogeneous Catalytic Application of COFs

with both imine units and bipyridine ligands, while the more
structurally rigid Rh(COD)Cl was deposited in the pore space and
complexed with bipyridine ligands. This Rh/Pd bimetallic–docked
Bph COF demonstrated superior catalytic activity toward a tandem
addition-oxidation reaction between phenylboronic acid and
benzaldehyde to initially form the intermediate diphenylmethanol
and further be oxidized to the final compound, benzophenone.
The authors found that the Rh(COD)Cl moiety in the COF was
accountable for the addition reaction between phenylboronic acid
and benzaldehyde, surpassing its homogeneous Rh(COD)Cl analogue
in catalytic activity. Pd(OAc)2 was responsible for the oxidation from
diphenylmethanol to benzophenone. The as-prepared Rh/Pd-Bph
COF showed excellent recyclability and reusability, providing isolated
benzophenone products up to 85% yield even after five cycles of
reuse, without noticeable leaching of the metal and apparent loss of
activity. This report manifested its significance in that it illustrated
that two different kinds of organometallic compounds could be
docked in a structurally its tunable COF by different coordinating
groups to render the COF materials different catalytic ability toward
totally differentiated reaction types in the first time.

Figure 4.19 (a) Use of a three-component condensation system to modulate


the nitrogen content of the 2D imine-type COFs, (b) designed strategies for the
monometallic (Route 1) and bimetallic docking (Route 2), (c) open channels
of the COFs, and (d) open channels of the metal-loaded COFs. Right: One-pot
cascade reactions using different homogeneous/heterogeneous catalysts.
Reproduced with permission from Ref. [28]. Copyright (2016), John Wiley and
Sons.
Heterogeneous Photo- and Electrocatalysts of COFs 129

4.4 Heterogeneous Photo- and Electrocatalysts


of COFs
Porous network structures, together with robustness and molecular
functionality with the possibility to load desirable species, favor the
generation of highly effective photocatalysts and electrocatalysts.
Lotsch et al. reported a new COF capable of visible-light-driven
hydrogen generation in the presence of Pt as a proton reduction
catalyst (PRC) [29]. A hydrazone-linked COF, 1,3,5-tris(4-
formylphenyl)triazine (TFPT) COF, was synthesized via the acetic
acid–catalyzed reversible condensation of TFPT and 2,5-diethoxy-
terephthalohydrazide as the building blocks in dioxane/mesitylene
at 120°C under solvothermal conditions (Fig. 4.20a). The COF
adopts a layered structure with a honeycomb-type lattice featuring
mesopores of 3.8 nm and the highest surface area among all

Figure 4.20 Synthesis of hydrazone-linked COF (TFPT COF) with honeycomb


mesopores. (a) Scheme showing the condensation of the two monomers to
form the TFPT COF and (b) top view with an eclipsed primitive hexagonal lattice
(gray: C; blue: N; red: O). Reproduced from Ref. [29] under a Creative Commons
Attribution 3.0 Unported Licence (https://creativecommons.org/licenses/
by/3.0/).
130 Heterogeneous Catalytic Application of COFs

hydrazone-based COFs (Fig. 4.20a). When illuminated with visible


light, the Pt-doped COF continuously produces hydrogen from water
without signs of degradation. This result indicates that photoactive
COFs are well-defined model systems with which to carry out
efficient photocatalytic processes.
CdS NPs have also been deposited on a highly stable 2D COF
matrix, and the generated hybrid was used as a photocatalyst for
visible-light-driven hydrogen production (Fig. 4.21). In particular, the
efficiency of a CdS COF hybrid with different COF contents has been
investigated. Upon the introduction of only 1 wt% of COF, a 10-fold
increase in the overall photocatalytic activity has been observed. The
hybrid with 10 wt% of COF content exhibits a marked H2 production
of approximately 3700 mmol h–1 g–1, which is significantly higher
than that for bulk CdS NPs (124 mmol h–1 g–1) [30]. This study means
an effective amalgamation of organic and inorganic materials within
a single hybrid, resulting in an improved activity compared to that of
each of the constituents.

Figure 4.21 Representation of (a) synthesis of the COF TpPa-2. (b) CdS COF
hybrid formation by the hydrothermal synthesis of CdS nanoparticles on the
COF matrix. Reproduced with permission from Ref. [30]. Copyright (2014), John
Wiley and Sons.
Heterogeneous Photo- and Electrocatalysts of COFs 131

A series of water- and photostable 2D azine-linked COFs have


been synthesized from hydrazine and triphenylarene aldehyde
with a varying number of nitrogen atoms [31]. The electronic and
steric variations in the precursors are transferred to the resulting
frameworks, leading to a progressively enhanced light-induced
hydrogen evaluation with increasing nitrogen content in the
frameworks. These results demonstrate that by the rational design
of COFs on a molecular level, it is possible to precisely adjust their
structural and optoelectronic properties, resulting in enhanced
photocatalytic activities. In addition, structure-property-activity
relationships in a pyridine unit containing an azine-linked COF
for photocatalytic hydrogen evolution have been elucidated [32].
Owing to their inherent porosity and well-ordered nanoscale
architectures, COFs are an especially attractive platform to tune
photocatalytic hydrogen evaluation, which is extended to a pyridine-
based photocatalytic active framework, where nitrogen substitution
in the peripheral aryl rings reverses the polarity compared to the
above description (Ref. [31]). Lotsch et al. demonstrated how
simple changes at the molecular level translate into significant
differences in atomic-scale structure, nanoscale morphology, and
optoelectronic properties, which greatly affect the photocatalytic
hydrogen evolution efficiency. In an effort to understand the
complex interplay of such factors, the authors carved out the
conformational flexibility of the PTP COF precursor and the vertical
radical anion stabilization energy as important descriptors to
understand the performance of COF photocatalysts. A benchmark
example of COF-based photocatalysts for solar fuel production
from CO2 has been realized. It is demonstrated that the visible-
light-harvesting capacity, suitable bandgap, and highly ordered p
electron channels contribute to the excellent performance of the
COF film photocatalyst [33]. In 2013, Jiang et al. first examined the
photocatalytic activity of squaraine-linked cupper porphyrin (CuP)
COF by using 1,3-diphenylisobenzofuran (DPBF) as a label for singlet
oxygen generation [34]. Further in 2015, Jiang et al. also investigated
the photocatalytic activity of an imine-linked CuP COF by using
DPBF for singlet oxygen generation. In comparison with other CuP
derivatives, the CuP-dihydroxyphenyl (DHPh) COF is exceptionally
active as a photocatalyst, exhibiting a 10- to 20-fold enhancement in
activity [35].
132 Heterogeneous Catalytic Application of COFs

Recently, a p-conjugated TpMA COF, which was composed of TFP


and melamine (MA), with triazine units and cyclic ketone units was
artfully designed and synthesized. This COF was found to exhibit an
excellent visible-photocatalytic capacity for the decomposition of
organic pollutants. The triazine units that were artfully integrated
into the COF skeleton served as photoactive centers, and the cyclic
ketone units served as electron-withdrawing moieties. Therefore,
the conjugated structure served as a photoelectron shift platform
[36]. Chemically stable CTFs have been also used as photocatalysts
for H2 evolution in the presence of Pt under visible-light irradiation,
and relatively low hydrogen evolution rates of approximately 200
mmol h–1 g–1 were obtained [37]. Subsequently, a fast and facile route
for the optimization of CTFs for photocatalytic hydrogen production
was presented by Thomas and coworkers [38]. They found that the
optimized CTF catalysts showed an average hydrogen evolution rate
of 1072 mmol h–1 g–1 under visible light (>420 nm). Very recently,
Lotsch et al. further demonstrated photocatalytic hydrogen evolution
using COF photosensitizers with molecular PRCs.
Using an azine-linked COF (N2) photosensitizer, a chloro(pyridine)
cobaloxime cocatalyst, and a triethanolamine donor, a H2 evolution
rate of 782 mmol h–1 g–1 and a TON of 54.4 was obtained in a water/
acetonitrile mixed solvent [39]. In addition to hydrogen evolution,
a 2D COF was found to be a highly efficient, metal-free recyclable
heterogeneous photocatalyst for oxidative C–H functionalization
under visible-light irradiation using O2 as a green oxygen source
[40]. In addition, 2D COFs can act as efficient type II photosensitizers
for photodynamic inactivation of bacteria [41].

4.5 Heterogeneous Catalysts of 3D COFs


In 2014, Yan and coworkers designed and prepared two 3D
microporous base–functionalized COFs, termed BF-COF-1 and
BF-COF-2, via the condensation reaction of a tetrahedral alkyl
amine, 1,3,5,7-tetraaminoadamantane (TAA), combined with
1,3,5-triformylbenzene (TFB), or TFP, as shown Fig. 4.22a [42].
Using these two COFs as catalysts, both BF COFs showed remarkable
conversion rates (96% for BF-COF-1 and 98% for BF-COF-2) and good
Heterogeneous Catalysts of 3D COFs 133

recyclability in base-catalyzed Knoevenagel condensation reactions.


Importantly, the COFs exhibited highly efficient size selectivity due
to the pore size effect of COFs (Fig. 4.22b). The strong interaction
between reactants and the COFs may lead to a high efficiency in size
selectivity.

Figure 4.22 Left: Schematic representation of the strategy for preparing


3D microporous base–functionalized COFs. (a) Model reaction of
1-adamantanamine with benzaldehyde to form the molecular N-(1-adamantyl)
benzaldehyde imine, (b) structure of 1,3,5,7-tetraaminoadamantane (TAA) as
a tetrahedral building unit, (c) structure of 1,3,5-triformylbenzene (R = H, TFB)
or triformylphloroglucinol (R = OH, TFP) as a triangular building unit, and (d)
condensation of tetrahedral and triangular building units to give a 3D network
with the symbol ctn (BF-COF-1 and BF-COF-2). Right: Catalytic activity of BF
COFs in the size-selective Knoevenagel condensation reaction. Reproduced
with permission from Ref. [42]. Copyright (2014), John Wiley and Sons.

Recently, Wang and coworkers described that an interpenetrating


dynamic 3D COF LZU-301 could be a Lewis-base catalyst for the
Knoevenagel condensation between malonitrile and three aromatic
aldehydes, as shown in Fig. 4.23 [43]. The authors discovered that
for the small-sized aldehyde, the 3D COF LZU-301 provided an
excellent yield of up to 72% in 4 h and 99% in 10 h. However, for
the larger 2-napthalenealdehyde and 9-anthracenealdehgyde, the
yield notably decreased to 21% and 5% because of a size effect
of the pore. The larger fuse-ring aromatic aldehyde could not be
accommodated into the pore space of LZU-301 and thus did not
134 Heterogeneous Catalytic Application of COFs

have enough interaction with its pyridine Lewis-base catalytic site,


leading to inferior catalytic performance. This example manifests
its significance in that LZU-301 was one of the few 3D COFs that
demonstrate considerable catalytic activities in meaningful organic
synthesis.

Figure 4.23 (a) Solvothermal synthesis of a 3D COF, LZU-301, via imine


condensation. For clarity, only the single framework of LZU-301 is shown.
(b) Side and top views of the porous crystalline structure of LZU-301, which
features a ninefold interpenetration of the underlying diamond net. Different
colors represent different penetrating frameworks from a side view under
reaction condensation. Reprinted with permission from Ref. [43]. Copyright
(2017) American Chemical Society.

The 3D COFs DL-COF-1 and DL-COF2 were prepared from the


dual linking reaction between TAA and 4-formylphenylboronic acid
(FPBA) or 2-fluoro-4-formylphenylboronic acid (FFPBA), forming
two kinds of linkages in the COF skeleton, boroxine and imine bonds
(Fig. 4.24) [44]. The as-synthesized 3D COFs displayed large specific
surface areas and incorporated both acidic boroxine sites and basic
imine sites. These two different sites rendered these 3D COFs as
versatile bifunctional heterogeneous catalysts. To demonstrate the
catalytic activity of DL-COF-1, a one-pot deacetalization-Knoevenagel
reaction was applied. DL-COF-1 exhibited excellent yields in both
acid-catalyzed deacetalization reaction (yield up to 100%) and
base-catalyzed Knoevenagel condensation reaction (yield up to
98%). The COF crystals can be recycled and reused three times with
almost no loss of activity and no identical change in PXRD and H2
uptake characterization. This experiment was very encouraging
since it produced the first bifunctional 3D COF with a large surface
area to fulfill heterogeneous catalytic applications.
Conclusions and Outlook 135

Figure 4.24 Left: Strategy for synthesizing 3D COFs with dual linkages (DL-
COF), (a) model reaction of AA with FPBA to form a triangular molecule with
dual linkages, (b) condensation of tetrahedral TTA and FFPBA to give 3D COFs
with dual linkages, DL-COF-1 or DL-COF-2, and (c) on the basis of triangular and
tetrahedral building units, both DL COFs show 3D networks with a ctn topology.
Right: Structural representations of 3D DL-COF-1 (a) and DL-COF-2 (b). Reprinted
with permission from Ref. [44]. Copyright (2016) American Chemical Society.

4.6 Conclusions and Outlook


Owing to the unique features of COFs and recent interesting
exploitation of their properties, COFs undoubtedly demonstrate
significant potential and advantages as catalysts. Therefore,
COF catalysis has attracted a great deal of attention, and some
progress in the field has been made in recent years. However, the
investigation of catalytic COF materials is still in its infancy, and COF
catalysis still faces many challenges in practical applications, for
example, how to further improve the chemical/thermal stability of
COFs so that more reactions can be carried out under very harsh
conditions (e.g., strong acidity or alkalinity, high temperature, and
high overpotential). The preparation of successful crystalline COFs
136 Heterogeneous Catalytic Application of COFs

has long been an intractable problem because there is a lack of


universal regulation to construct crystalline COFs and significant
time and effort must be spent on searching for the appropriate
reaction conditions. Therefore, there is a clear need to expand the
synthetic possibilities in different ways. Currently, the synthesis
of COFs is limited to the laboratory and it has been a challenge
to prepare COFs on an industrial scale for prospective practical
applications. Development of 3D COFs with versatile and cavities is
very interesting to provide active sites on the pore surface and to
transport reactants/products for formation of inner reactive vessels.
This would enable supramolecular catalysis or enzyme catalysis to
take place in an appropriate space under nanoscale confinement,
allowing multicomponent reactions and a wide range of substrates
from small molecules to large polyaromatics and carbohydrates.
In the future, the goal of COF catalysis is to develop heterogeneous
catalysts with high stability, low cost, and high conversion and
selectivity. New synthesis methods and process routes for large-
scale synthesis should be explored simultaneously, thus laying the
foundation for the commercialization and industrialization of COF
catalysts.

References

1. Beletskaya, I. P., Cheprakov, A. V. (2000). Chem. Rev., 100, 3009–3066.


2. Takemoto, T., Iwasa, S., Hamada, H., Shibatomi, K., Kameyama, M.,
Motoyama, Y., Nishiyama, H. (2007). Tetrahedron Lett., 48, 3397–3401.
3. Ozawa, F., Kubo, A., Hayashi, T. (1993). In Selectivity in Catalysis; ACS
Symposium Series 517; ACS: Washington, DC, USA, Vol. 517, 75–85.
4. Kamei, T., Sato, A. H., Iwasawa, T. (2011). Tetrahedron Lett., 52, 2638–
2641.
5. Liang, J., Liang, Z., Zou, R., Zhao, Y. (2017). Adv. Mater., 29, 1701139.
6. Ding, S.-Y., Gao, J., Wang, Q., Zhang, Y., Song, W.-G., Su, C.-Y., Wang, W.
(2011). J. Am. Chem. Soc., 133, 19816–19822.
7. Hou, Y., Zhang, X., Sun, J., Lin, S., Qi, D., Hong, R., Li, D., Xiao, X., Jiang, D.
(2015). Microporous Mesporous Mater., 214, 108–114.
8. Llabres i Xamena, F. X., Abad, A., Corma, A., Garcia, H. (2007). J. Catal.,
250, 294–298.
References 137

9. Mullangi, D., Nandi, S., Shalini, S., Sreedhala, S., Vinod, C. P.,
Vaidhyanathan, R. (2015). Sci. Rep., 5, 10876.
10. Kaleeswaran, D., Antony, R., Sharma, A., Malani, A., Murugavel, R.
(2017). Chempluschem, 82, 1253–1265.
11. Lu, S., Hu, Y., Wan, S., McCaffrey, R., Jin, Y., Gu, H., Zhang, W. (2017). J. Am.
Chem. Soc., 139, 17082–17088.
12. Beletskaya, I. P., Cheprakov, A. V. (2000). Chem. Rev., 100, 3009–3066.
13. Pachfule, P., Panda, M. K., Kandambeth, S., Shivaprasad, S. M., Diaz Daiz,
D., Banerjee, R. (2014). J. Mater. Chem. A, 2, 7944–7952.
14. Zhang, J., Peng, Y., Leng, W., Gao, Y., Xu, F., Chai, J. (2016). Chin. J. Cata.,
37, 468–475.
15. Bhadra, M., Sasmal, H. S., Basu, A., Midya, S. P., Kandambeth, S., Pachfule,
P., Balaraman, E., Banerjee, R. (2017). ACS Appl. Mater. Interface, 9,
13785–13792.
16. Lin, S., Hou, Y., Deng, X., Wang, H., Sun, S., Zhang, X. (2015). RSC Adv., 5,
41017–41024.
17. Nagai, A., Guo, Z., Feng, X., Jin, S., Chen, X., Ding, X., Jiang, D. (2011). Nat.
Commun., 2, 536, doi: 10.1038/ncomms1542.
18. Xu, H., Chen, X., Gao, J., Lin, J., Addicoat, M., Irle, S., Jiang, D. (2014).
Chem. Commun., 50, 1292–1294.
19. Xu, H., Cao, J., Jiang, D. (2015). Nat. Chem., 7, 905–912.
20. Ma, H. C., Kan, J. L., Chen, G. J., Chen. C. X., Dong, Y. B. (2017). Chem.
Mater., 29, 6518–6524.
21. Zhang, J., Han, X., Wu, Z. W., Liu, Y., Chu, Y. (2017). J. Am. Chem. Soc., 139,
8277–8285.
22. Han, X., Xia, Q., Huang, J., Liu, Y., Tan, C., Cui, Y. (2017). J. Am. Chem. Soc.,
139, 8693–8697.
23. Peng, Y. W., Hu, Z. G., Gao, Y. J., Yuan, D. Q., Kang, Z. X., Qian, Y. H., Yan, N.,
Zhao, D. (2015). ChemSusChem, 8, 3208–3212.
24. Wang, X. R., Han, X., Zhang, J., Wu, X. W., Liu, Y., Cui, Y. (2016). J. Am.
Chem. Soc., 138, 12332–12335.
25. Xu, H. S., Ding, S. Y., An, W. K., Wu, H., Wang, W. (2016). J. Am. Chem. Soc.,
138, 11489–111492.
26. Shinde, D. B., Kandambeth, S., Pachfule, P., Kumar, R. R. (2015). Chem.
Commun., 51, 310–313.
27. Leng, W., Ge, R., Dong, B., Wang, C., Gao, Y. (2016). RSC Adv., 6, 37403–
37406.
138 Heterogeneous Catalytic Application of COFs

28. Leng, W., Peng, Y., Zhang, J., Lu, H., Feng, X., Ge, R., Dong, B., Wang, B.,
Gao, Y. (2016). Chem. Eur. J., 22, 9087–9091.
29. Stegbauer, L., Schwinghammer, K., Lotsch, B. V. (2014). Chem. Sci., 5,
2789–2793.
30. Thote, J., Aiyappa, H. B., Deshpande, A., Díaz, D. D., Kurungot, S.,
Banerjee, R. (2014). Chem. Eur. J., 20, 15961–15965.
31. Vyas, V. S., Haase, F., Stegauer, L., Savasci, G., Podjaski, F., Ochenfeld, C.,
Lotsch, B. V. (2015). Nat, Commun., 6, 8508.
32. Haase, F., Banerjee, T., Savasci, G., Christian, O., Lotsch, B. V. (2017).
Faraday Discuss., 201, 247–264.
33. Yadav, R. K., Kumar, A., Park, N. J., Kong, K. J., Baeg, J. O. (2016). J. Mater.
Chem. A, 4, 9413–9418.
34. Nagai, A., Chen, X., Feng, X., Ding, X., Guo, Z., Jiang, D. (2013). Angew.
Chem. Int. Ed., 52, 3770–3774.
35. Chen, X., Addicoat, M., Jing, E. Q., Zhai, L. P., Xu, H., Huang, N., Guo, Z. Q.,
Liu, L. L., Irle, S., Jiang, D. L. (2015). J. Am. Chem. Soc., 137, 3241–9639.
36. He, S., Rong, Q., Niu, H., Cai, Y. (2017). Chem. Commun., 53, 9636–9639.
37. Bi, J., Fang, W., Li, L., Wang, J., Liang, S., He, Y., Liu, M., Wu, L. (2015).
Macromol. Rapid Commun., 36, 1799–1805.
38. Kuecken, S., Acharjya, A., Schwarze, M., Schomäcker, R., Thomas, A.
(2017). Chem. Commun., 53, 5854–5857.
39. Banerjee, T., Hasse, F., Savasci, G., Gottschling, K., Ochsenfeld, C., Lotsch,
B. V. (2017). J. Am. Chem. Soc., 139, 16228–16234.
40. Zhi, Y., Li, Z., Feng, X., Xia, H., Zhang, Y., Shi, Z., Mu, Y., Liu, X. (2017). J.
Mater. Chem. A, 5, 22933–22938.
41. Liu, T., Hu, X., Wang, Y., Meng, L., Zhou, Y., Zhang, J., Chen, M., Zhang, X.
(2017). J. Photochem. Photobiol. B, 175, 156–162.
42. Fang, Q. R., Gu, S., Zheng, J., Zhuang, Z. B., Qiu, D. L., Yan, Y. S. (2014).
Angew. Chem. Int. Ed., 53, 2878–2882.
43. Ma, Y. X., Li, Z. J., Wei, L., Ding, S. Y., Zhang,Y. B., Wang, W. (2017). J. Am.
Chem. Soc., 139, 4995–9091.
44. Li, H., Pan, Q., Ma, Y., Guan, X., Xue, M., Fang, Q., Yan,Y., Valtchev, V., Qiu,
S. (2016). J. Am. Chem. Soc., 138, 14783–14788.
Chapter 5

Energy Storage Applications of 2D COFs

Synthetic polymers with branched macromolecules and outstanding


functional group tolerance show diverse and useful properties that
influence most aspects of modern life. Extending polymerization
strategies to two dimensions allows precise integration of building
blocks into extended structures with periodic skeletons and ordered
nanopores. The construction principle of these frameworks is
the direct topological evaluation in a predictable manner with
controlled geometry, dimensions, and structural periodicity. This
unique designable feature of 2D covalent organic frameworks
(COFs) with versatile properties makes them an emerging material
platform with great relevance in areas such as gas storage and
separation (described in Chapter 3), catalysis (described in Chapter
4), and optoelectronics. This chapter focuses on the recent progress
in 2D COFs as optoelectronic materials, with an emphasis on their
semiconducting, energy conversion, and energy storage properties.

5.1 2D COFs for Optoelectronics and Energy


Storage
2D covalent polymers, such as covalent organic frameworks (COFs),
in which building blocks are precisely integrated into extended
structures with periodic skeletons and ordered pores, have some

Covalent Organic Frameworks


Atsushi Nagai
Copyright © 2020 Jenny Stanford Publishing Pte. Ltd.
ISBN 978-981-4800-87-7 (Hardcover), 978-1-003-00469-1 (eBook)
www.jennystanford.com
140 Energy Storage Applications of 2D COFs

distinctive advantages [1–3]. The unique topological diagram


directs the growth of the frameworks in a predictable manner,
and the geometry and dimensions of the building blocks govern
the structures of the resulting COFs. In 2D COFs, building units are
stacked via p-p interactions to form layered structures with a well-
defined alignment of p building units to their atomic layers and
segregated arrays of p columns. In such arrays, the intralayered
covalent bonds lock the frameworks whereas interlayer noncovalent
interaction controls the stability. Thus, various COFs can be achieved
by controlling pore design from a material-design perspective, by
controlling the skeleton design, or by using complementary designs
of both pores and skeleton. Such varying designable features of
2D COFs, with confined spaces in controllable 1D nanochannels,
offer the possibility to trigger interactions with excitons, electrons,
holes, spins, ions, and molecules. By this means, 2D COFs exhibit
unique properties and functions with outstanding applicability in
semiconductors [4], gas absorption [5, 6], proton conduction [7, 8],
luminescence [9, 10], and energy conversion and storage.

5.2 Semiconducting and Photoconducting 2D


COFs
Many conducting materials of organic compounds have been
designed by fine-tuning their energy gaps and highest occupied
molecular orbital/lowest occupied molecular orbital (HOMO/
LUMO) levels and can be efficiently produced on a large scale by
the versatile tools of organic chemistry [11]. The electronic devices
fabricated with these materials often exhibit lower stability and
efficiency than expected due to inefficient stacking or disordered
donor-acceptor interactions. Thus, it would be desirable to have
synthetic access to conducting materials with total control over their
nanoscale structures and orientations for the fabrication of nanoscale
devices and to expand our understanding of the fundamental
physics of 2D semiconducting materials. In this respect, 2D COFs, in
which modularly designed organic building blocks are periodically
linked with regular pores, offer the possibility of developing fully
conjugated 2D crystalline networks. Such periodically stacked
alignment develops columns in a direction that favors the transport
P-Type Semiconducting 2D COFs 141

of charge carriers and photoexcited states (excitons) produced from


excited building blocks through preorganized and built-in pathway.
This unidirectional charge transport improves the carrier mobility
of 2D COFs and provides a new platform for semiconducting and
photoconducting materials design. A selection of semiconducting
COFs is summarized in Table 5.1.

Table 5.1 Summarized semiconducting property of some COFs

Semiconducting Semiconducting
COFs property COFs Property
TP COF p-type DTP-ANDI COFs Ambipolar type
PPy COF p-type DTP-APyDI COFs Ambipolar type
MPc COFs (M = p-type Tp-P COFs Ambipolar type
Ni, Co, Cu, Zn)
COF-366 p-type ZnP COFs Ambipolar type
COF-66 p-type DMPc-APyDI COFs Ambipolar type
(M = Cu, Ni)
H2P COFs p-type DMPc-ANDI COFs Ambipolar type
(M = Cu, Ni, Zn)
T COFs p-type DMPc-APDI COFs Ambipolar type
(M = Cu, Zn)
TTF-Ph COFs p-type C2N-h2D Ambipolar type
TTF-Py COFs p-type [C60]y-ZnPC Ambipolar type
COFs
NiPc-BTTA COFs n-type [C60]-TT COFs Ambipolar type
CuP COFs n-type CS-COF-C60 Ambipolar type
D-A COFs Ambipolar type
TP, triphenylene; PPy, pyrene-2,7-diboronic acid; Pc, phthalocyanine; P, porphyrin;
T, thiophene; TTF, tetrathiafulvalene; Ph, phenyl; Py, pyrene; BTDA, benzothiazole;
D, donor; A, acceptor; PyDI, pyromellitic diimide; NDI, naphthalene diimide; PDI,
perylene diimide

5.3  P-Type Semiconducting 2D COFs


In general, COFs with extended p-conjugated systems, designed
by using more electron-rich building blocks, display p-type
142 Energy Storage Applications of 2D COFs

semiconducting behavior. Jiang and coworkers pioneered in this


field and first demonstrated the luminesce and semiconducting
behavior of COFs consisting of alternating pyrene and triphenylene
functionalities (named TP COFs) as shown in Fig. 5.1 [12]. The co-
condensation of 2,3,6,7,10,11-hexahydroxytriphenylene (HHTP)
and pyrene-2,7-diboronic acid (PDBA) adopts a belt-shaped
hexagonal crystalline framework with a pore diameter of 3.14 nm
and a specific surface area and a pore volume of 868 m2/g and
0.7907 cm3/g, respectively. Because of matching spectral profiles of
the building blocks, a TP COF harvests a wide wavelength range of
photons, allowing energy transfer and migration between them and
blue luminesce. Upon excitation of triphenylene units of a TP COF
at 340 nm, two bands were visible, with a weak emission maximum
at 402 nm and a strong emission maximum at 474 nm. In contrast,
the individual monomers on excitation show emission maxima at
402 nm (triphenyl unit) and 475 nm (pyrene unit). The difference
in fluorescence intensities at 474 and 402 nm (I474 nm/I402 nm = 16)
is viewed as evidence of electronic coupling of the building blocks,
which favors an energy transfer between them in the extended TP-
COF structure. In addition, fluorescence anisotropy measurements
illustrate that this excitation energy migrates randomly over the
crystalline belt within the lifetime of the excited state as well. To
measure the electrical conductivity, a disappeared COF solution
in acetone was drop-casted onto a Pt electrode with a width of 10
mm. In air at 25°C the I-V profile of the PT COF was almost linear,
irrespective of the voltage bias, while the gap itself was silent,
showing the semiconducting nature of this material. The stability of
this material was shown from continual on-off switching for many
cycles without significant weakening of the electric current. Upon
doping the framework with iodine, the electric current increased
from 4.3 to 20 nA at a bias voltage of 2 V, indicating the p-type
semiconducting character of the TP COF. They also reported the
continuous self-condensation of PDBA to afford a PPy COF with a
pore diameter of 1.73 nm and a surface area of 923 m2/g (Fig. 5.1)
[13]. The PPy COF displayed a photoconducting behavior with an
electrical conductivity of the same order as that of a TP COF. The
photoconductivity of the PPy COF was investigated using a sandwich
electrode, fabricated by casting a thin film (PPy COF/PMMA = 50/50,
wt%) with a thickness of 100 mm on an Al electrode and covering
P-Type Semiconducting 2D COFs 143

the film with a 30 nm thick layer of Au by vapor deposition. When


the PPy COF is irradiated with visible light (>400 nm) from the Au
side with a xenon lamp, a remarkable photocurrent is generated that
can be switched repetitively many times without any deterioration
at an on/off ratio of over 8 × 104.

Figure 5.1 Schematic representation of the chemical structures of TP COF and


PPy COF [12, 13].

Porphyrin and phthalocyanine are structurally related


macrocyclic compounds, ubiquitous in nature and serving as
important components of molecular materials with electronic,
144 Energy Storage Applications of 2D COFs

magnetic, and physicochemical properties. The incorporation of these


into 2D COFs, where p electron systems align in a precise order, opens
the possibility of developing promising photon drive devices. The
amalgamation of large metallophthalocyanine (MPc) p systems into
2D COFs based on a boronate esterification reaction was first reported
by Jiang et al. [14]. Under solvothermal conditions the condensation
of (2,3,9,10,16,17,23,24-octahydroxyphthalocyaninato)nickel(II)
and [(OH)8PcNi] with 1,4-benzenediboronic acid (BDBA) produced
the desired 2D NiPc COF with a 90% yield (Fig. 5.2). The resultant
COF exists as a layered structure of planar sheets, with a large
surface area of 624 m2/g and uniform microporous channels
with a diameter of 1.9 nm. This well-ordered eclipsed stacking

Electron

Hole

Figure 5.2 (a) Schematic representation of metallophthalocyanine 2D-covalent


organic frameworks (MPc COFs), (b) eclipsed stack of phthalocyanine 2D
sheets with microporous channels, (c) schematic representation of stacked
phthalocyanine p columns in MPc COFs for hole-carrier transport. Reproduced
with permission from Ref. [14]. Copyright (2011), John Wiley and Sons.
P-Type Semiconducting 2D COFs 145

conformation enhanced its light-harvesting capacity in the deep-


red visible and near-infrared (NIR) regions. Upon irradiation with
a xenon lamp (>400 nm), a significant increase in current from
20 nA to 3 mA at a 1 V bias was observed, which was reproducible
over many cycles without deterioration.
In addition, the wavelength-dependent on-off switching of the
photocurrent on irradiation with light passing through band path
filters (±5 nm) at the same bias revealed its extreme sensitivity toward
deep-red and NIR photons. When sandwiched between Au and Al
electrodes, the COF shows linear I-V profiles regardless of the bias
voltage, indicating its semiconducting nature. Laser flash-photolysis
time-resolved microwave conductivity (FP-TRMC) measurements
were used to evaluate the intrinsic charge-carrier mobility with the
value of fSm (Sm = the sum of carrier mobility, f = the charge-carrier
generation efficiency). The charge-carrier generation efficiency f was
determined by the integration of the time-of-flight (TOF) transients
at different bias voltages. The charge-carrier efficiency f of 3.0 ×
10–5 was estimated by the extrapolation of the bias at 0 V. Persistent
results on varying the atmosphere of the measurement signposts
indicate that carrier species predominantly originate from holes with
a mobility as high as 1.3 cm2 V–1 s–1. In a publication that followed,
they stretched their search and investigated the role of central ions
in controlling the photoelectric function of this COF framework [15].
They expanded the usefulness of phthalocyanine COFs (MPs COFs)
with various metal ions (M = Co, Cu, Zn), as shown in Fig. 5.2. In
terms of structural interpretation, all the MPc COFs are unique,
with highly ordered crystalline skeletons with periodic mesoporous
columns. The inherent charge-carrier mobility measured by the FP-
TRMC method under identical conditions revealed that a CoPc COF
has the highest fSm value (of 2.6 × 10–4 cm2 V–1 s–1) followed by a
ZnPc COF (2.2 × 10–4 cm2 V–1 s–1) and a CuPc COF (1.4 × 10–4 cm2 V–1
s–1). This trend matched well with the reported electron densities
of the phthalocyanine macrocycles, in the order CuPc COF < ZnPc
COF < CoPc COF [16, 17]. The photocurrent generation upon light
irradiation was examined on spin-coating the MPc COFs onto a
sandwich-type gap between Al and Au electrodes with an applied
bias voltage of 2 V. Although the random monomeric samples are
not photoconductive, the MPc COFs show high photoconductivity
with a generated photocurrent of approximately 110 nA for a CuPc
146 Energy Storage Applications of 2D COFs

COF, followed by 0.6 nA and 0.14 nA for a ZnPc COF and a CoPc
COF, respectively. This eccentricity between the order of produced
photocurrents and fSm values for different MPc COFs highlights the
importance of morphologies and boundaries of the objects as well
as the role of metal ions to control the performance of photoelectric
devices. More recently, two triangular COFs, hexaphenylbenzene
(HPB) COFs and hexabenzocoronene (HBC) COFs, were developed
by using two different C6-symmetric vertices HPB and HBC based
on the triangular topology [18]. This unique topology favors the
formation of supermicropores with pore sizes as low as 12 Å and
densely packed p columns of density 0.25 nm–2, which exceeds
those of the COFs and supramolecular p arrays reported to date.
These support intra- and interlayer p cloud delocalization on the
framework and display prominent photoconductivity, with carrier
mobilities as high as 0.7 cm2 V–1 s–1, which is among the highest
reported for 2D COFs and photographitic ensembles.
Yaghi and coworkers reported two unique COFs under
solvothermal condensation of porphyrin derivative, either through
boronate ester formation with tetrahydroxy anthracene (COF-66)
or through imine bond formation with tetraphthaladehyde (COF-
366) (Fig. 5.3) [19]. The electrical conductivity of both COFs was
examined across a gap of 2 mm between two Au electrodes. In air
at room temperature, both COFs display linear I-V curve profiles
and electrical conductivity with an electric current of 0.75 nA at
a 0.2 V bias voltage. The 1.5 mm thin film fabricated from COF/
poly(methyl methacrylate), 60/40 in wt%, between Al and indium
tin oxide (ITO) electrodes, showed hole conduction on TOF transient
current integration measurement. This 1D hole mobility of 8.1 cm2
V–1 s–1 and 3 cm2 V–1 s–1, respectively, from imine and boronate
ester COFs indicates that both are p-type semiconductors. This is
also attributed to the fact that the imine bond in COF-366 improves
the conjugation of the framework, responsible for the highest hole
mobility in semiconducting COFs up to now. Jiang et al. reported a
disk-shaped tetragonal H2P COF, with a porphyrin unit embedded
in the mesoporous framework (Fig. 5.4) [20]. The porphyrin unit
is aligned in an eclipsed stacking mode with high crystallinity
and a large surface area (1901 m2/g) and offers macrocycle-on-
macrocycle columnar porphyrin paths that favor condition pathways
for high-rate charge-carrier mobility. The charge-carrier mobility
P-Type Semiconducting 2D COFs 147

Figure 5.3 Schematic representation of the chemical structures of COF 266


and COF 66.

was evaluated with FP-TRMC methods after laser pulse irradiation


under a rapidly oscillating electric field. In both conditions (Ar and
SF6 atmosphere) the hole transport mobility displays the same value
(fmh = 1.8 × 10–4 cm2 V–1 s–1) for the H2P COF. The maximum quantum
yield (f) of photocarrier generation was 5 × 10–5, measured with the
TOF-transient current integration method. This implies a minimum
mh value of 3.5 cm2 V–1 s–1, which is twofold higher than that of
the previously reported phthalocyanine COF (1.6 cm2 V–1 s–1). The
photocurrent generation of approximately 0.01 nA was measured
upon irradiation with visible light (>400 nm) using a Xe lamp with
148 Energy Storage Applications of 2D COFs

an on-off ratio of 4. The unbalanced carrier (hole and electron)


transport was responsible for low photocurrent generation as well
as low on-off ratio.

Figure 5.4 (a) Chemical structures of MP COFs (M = H2, Zn, Cu), (b) schematic
graphs of metal-on-metal and macrocycle-on-macrocycle channels for electron
transport and hole transport, respectively, in stacked porphyrin columns of 2D
MP COFs, (c) schematic graphs of MP COFs with achiral AA stacking of 2D sheets
(C: light blue; N: deep blue; H: white; O: red; B: pink; Zn: green; Cu: violet).
Reproduced with permission from Ref. [20]. Copyright (2012), John Wiley and
Sons.

In the recent past, thiophene- and tetrathiafulvalene (TTF)-based


derivatives have been frequently utilized as strong electron-donor
molecules for the development of electrically conducting materials
[21]. This has led different research groups to use these derivatives
as basic building blocks for the synthesis of semiconducting
COFs. Dincã et al. isolated the first charge transfer (CT) complex
postsynthetically inside a COF embedded with thiophene,
bithiophene, and thienothiophene monomers to achieve electric
conductivity [22]. The condensation of HHTP with diboronic acid
of different thiophene derivatives in dioxane and mesitylene was
carried out to afford different thiophene-based COFs (T COFs) as off-
P-Type Semiconducting 2D COFs 149

white powders with a good yield (Fig. 5.4). All these T COFs preferred
to adopt an eclipsed stacking conformation with pore diameters of
2.04, 3.14, and 2.94 nm and surface areas of 927, 544, and 904 m2/g
for T-COF-1, T-COF-3, and T-COF-4, respectively. They performed
the doping experiment with these COF frameworks with a suitable
acceptor to provide a sufficient concentration of CT to induce electric
conductivity. A series of acceptor molecules were examined in search
for proper redox partners that promote full electron transfer without
disrupting the crystalline structure. Among these 2,3-dichloro-5,6-
dicyano-1,4-benzoquinone, chloranil, and I2 formed the CT complex
with these COFs but led to the amorphization of the frameworks. On
the other hand, a 0.1 mM solution of tetracyano quinodimethane
(TCNQ) in CH2Cl2 gave no reaction with T-COF-1 or T-COF-3. It
immediately formed a crystalline back precipitate with an adsorption
maximum centered at 850 nm with retained crystallinity in contract
with solid T-COF-4. Thus, thiophene derivatives with lower oxidation
potentials may offer better p-type hosts to form the CT complex in
such COF frameworks for the implementation of their materials in
electronic applications. Even more recently, Dincã et al. extended
their research and reported heavier chalcogen-based COFs in the
form of fused benzodiselenophenes and benzoditellurophenes, to
compare the electrical properties relative to the thiophene analogue
(Fig. 5.5) [23]. Two-point probe electrical conductivity measurements
of these COFs revealed values of 3.7 ± 0.4 × 10–10 S/cm, 8.4 ± 3.8 ×
10–9 S/cm, and 1.3 ± 0.1 × 10–17 S/cm for S, Se, and Te derivatives,
respectively. The enhanced orbital overlap afforded by the 3p and 4p
orbitals of selenium and tellurium atoms or the enhanced spin-orbit
coupling effects responsible for this trend of electrical conductivity
increase in the order of S > Se > Te, without significantly changing
the structure or the unit cell size of the resulting extended networks.
More than a few research groups used the important
electroactive TTF building block as a basic building unit to fabricate
semiconducting COFs. The C2 + C2 topological diagram to embed
TTF units in COF architectures with varying linker units was
used by Jiang et al. TTF-based COFs TTF-Ph COF and TTF-Py COF
were prepared using phenyldiamine and tetra(p-aminophenyl)-
pyrene as linkers under solvothermal conditions (Fig. 5.6a) [24].
Both COFs exist as layered lattices with periodic TTF columns and
tetragonal open nanochannels. Because of differences in the shape
150 Energy Storage Applications of 2D COFs

Figure 5.5 Schematic representation of partial structures of T-COF-1, T-COF-3,


T-COF-4, and heavier chalcogen-based COFs. Reproduced with from Ref. [22]
with permission from PNAS.

of the linkers, the phenyl units adopt a planar conformation with an


interlayer distance of 3.71 Å on the TTF-Ph COF, whereas the paddle-
shaped tetraphenylpyrene linker distorts the layer from a planar
conformation and expands the interlayer distance up to 3.87 Å.
Thus, the shape of the linker plays an important role in tuning the
periodicity and conformation of the 2D layer with the interlayer
distance, which are key to electric functions, such as carrier
transport and conductivity. The inherent charge-carrier mobility
measured by using the FP-TRMC method after 355 nm laser pulse
irradiation revealed values of 0.2 and 0.08 cm2 V–1 s–1 for the TTF-
P-Type Semiconducting 2D COFs 151

Ph COF and the TTF-Py COF, respectively. Furthermore, conductivity


measurements of the COFs were also performed on these 0.4 nm
wide and 0.05 nm thick COF films after doping them with iodine
vapor. The conductivities of the TTF-Ph COF and the TTF-Py COF
were estimated to be 10–5 and 10–6 W–1 cm–1, respectively. This
higher conductivity value on doping suggests that layered TTF
columns in the COF frameworks allow the formation of a CT complex
between TTF and the radical cation and the iodide radical anion
without disturbing the lattice structure. In addition, the difference
in the degrees of enhancement for the TTF-Ph COF and the TTF-Py
COF is likely related to the linker nature, which governs the layer
conformation and the interlayer distance.

Figure 5.6 (a) Schematic representation of the synthesis of mesoporous TTF-


Ph COF and microporous TTF-Py COF by a C2 + C2 topological diagram. Slipped
AA stacking structures of TTF-Ph COF in (b) top and (c) side views and eclipsed
stacking structures of TTF-Py COF in (d) top and (e) side views (yellow: S; blue:
N; gray: C; H was omitted for clarity). (f) Illustration of the mixed-valence state
in TTF COF. The “h” indicates inter-TTF-layer interactions. Reproduced with
permission from Refs. [24]. Copyright (2014), John Wiley and Sons. Republished
with permission of Royal Society of Chemistry, from Ref. [26], copyright (2014);
permission conveyed through Copyright Clearance Center, Inc.

The synthesis of the TTF-Ph COF using cyclic voltammetry (CV)


measurements was investigated in detail for its electrochemical
behavior [25]. The working electrode was modified due to low
152 Energy Storage Applications of 2D COFs

solubility of this COF in common organic solvents. The TTF-Ph COF


and carbon black in a ratio of 3:2 wt% were ground in an agate
mortar and pestle for 10 min. and sonicated in CH2Cl2. Then, the
sonicated solution was drop-casted onto the tip of the Pt working
electrode. The electron-donating nature of the TTF-Ph COF was
established from the CV profile with two reversible redox processes
at 0.69 V and 1.07 V. A shape peak at 2.0094, corresponding to
the TTF radical cation in the electron paramagnetic resonance
(EPR) spectrum, validated the formation of a CT complex upon I2
adsorption on the COF pore. The electrical conductivity of its doped
state contracts in a constant-voltage two-probe configuration. The
conductivity value at 25°C increased from 2.1 × 10–7 S/cm to 1.8
× 10–6 S/cm on increasing the doping time from one to two days,
indicating that the conducting property of the TTF-Ph COF could
be tuned by iodine doping. Zhang et al. reported the oriented thin
film of a TTF-Ph COF with an aim to enhance and tune its electrical
conductivity in the presence of a molecular dopant as a CT partner
[26]. The thin films of the TTF-Ph COF were grown in situ from the
liquid phase with a nominal thickness of around 150 nm on Si/
SiO2 substrates and transparent indium thin oxide (ITO)-coated
glass. The ordered polycrystalline nature of the thin film with a
preferential orientation of the columns normal to the substrate was
confirmed from grazing incidence wide-angle X-ray scattering. Thin
films on a Si/SiO2 substrate with prefabricated Au/Cr electrodes 3
mm long, 50 nm thick, and 125 nm apart determine the conductivity
of 1.2 × 10–4 S/cm. On exposing this thin film to I2 vapor in a
closed chamber, a significant time-dependent increase in electrical
conductivity with a maximum value of 0.28 S/cm occurred after 24
h. Similar conductivity enhancement, albeit of a lower magnitude,
was also observed when the film was doped with TCNQ. This highly
conductive nature is due to the formation of a CT complex with I2 or
TCNQ and correlates well with diagnostic signatures in ultraviolet-
visible-NIR and EPR spectra. Thin films on an ITO transparent glass
substrate were used to record the optical adsorption spectra in
transmission mode. The low-intensity NIR peak at 1200 nm of the
COF film progressively redshifts up to 1400 nm and 2000 nm over
the course of TCNQ and I2 exposure. These NIR adsorption bands
signify a mixed-valance TTF species formation between TTF and
TTF radical cation, where effective radical delocalization within
N-Type Semiconducting 2d COFs 153

mixed-valance TTF stacks surges upward upon the formation of a


CT complex (Fig. 5.6f). A small amount of paramagnetic TTF radical
as a dopant in an as-synthesized COF was found by a weak signal on
electron spin resonance (ESR) spectra. Different degrees of doping
on varying spin-spin interactions with cumulative concentrations
of TTF radicals also reflect the increased paramagnetic intensity of
approximately 1 and 3 orders of magnitude after exposure of TCNQ
and I2, respectively.

5.4 N-Type Semiconducting 2D COFs


The first report about the electron transporting 2D COF developed
by Jiang et al. highlights the importance of embedded p-electronic
building blocks in the conducting properties of the frameworks
(Fig. 5.7) [27]. In their established strategy, an electron-deficient
p-electronic building block such as benzothiazole (BTDA) was
embedded at the edges of a two-component tetragonal nickel(II)
phthalocyanine COF (Fig. 5.6). Integration of BTDA units at the edge
causes a drastic change in the carrier transport mode from the hole
transporting with a phenyl unit at the edge to electron-conducting
frameworks. Co-condensation of MPc with 1,4-benzothiazole-
diboronic acid (BTDADA) under solvothermal conditions leads to
the formation of a crystalline belt-shaped 2D NiPc-BTDA COF that
adopts an AA-type stacking arrangement with a tetragonal pore of
diameter 2.2 nm and a surface area of 877 cm/g. Due to its preferable
eclipsed stacking, it exhibits broad and enhanced absorbance up to
1000 nm, which motivated investigations of its photoconducting
behavior. Upon irradiation with visible light (>400 nm), it exhibits
an enhanced photocurrent from 250 mA to 15 mA at a 1 V bias, which
runs down after many sequences. The high sensitivity toward NIR
photons through a panchromatic response was investigated by using
different band pass filters upon selective excitation with different
wavelengths. FP-TRMC methods were utilized to investigate the
intrinsic charge-carrier mobility under atmospheres of O2, SF6, and
Ar after excitation with a 355 nm pulsed laser. The profile shows
a double exponential decay curve, fitting for electrons and holes,
under an Ar atmosphere and gives a fSm value of 5.8 × 10–4 cm2
V–1 s–1. The decay profile responsible for electrons was augmented
154 Energy Storage Applications of 2D COFs

considerably upon measurement under SF6 and O2 atmospheres,


which are well known as electron quenchers. In contrast, the decay
profile responsible for holes was unaltered with respect to the
measurement atmosphere. These results revealed a 2D NiPc-BTDA
COF to be an electron transporting framework with electron mobility
as high as 0.6 cm2 V–1 s–1, which is outstanding.

Figure 5.7 (A) Schematic representation of the synthesis of 2D NiPc-BTDA


COF with metallophthalocyanine at the vertices and BTDA at the edges of the
tetragonal framework and (B) top and side views of a graphical representation
of a 2 × 2 tetragonal grid showing the eclipsed stacking of 2D polymer sheets.
(Pc: sky blue; BTDA: violet; Ni: green; N: blue; S: yellow; O: red; B: orange; H:
white). Reprinted with permission from Ref. [27]. Copyright (2011) American
Chemical Society.

The same group reported one more n-type semiconducting


2D COF (CuP COF) with copper porphyrin units at the nodes of a
tetragonal mesoporous framework surrounded by phenyl units
at the edge (Fig. 5.4) [20]. The quantitative evolution of intrinsic
electron carrier mobility was obtained by using the FP-TRMC
technique in an Ar atmosphere. Upon laser flash, the rise and decay
profiles of the FP-TRMC signal gave a maximum value of 1.16 × 10–4
cm2 V–1 s–1 (fSm) for the CuP COF. The maximum quantum yield
(f) of photocarrier generation of the Cu COF was 6 × 10–4, which
Ambipolar Semiconducting 2D COFs 155

led to the minimum electron mobility of 0.19 cm2 V–1 s–1. This
distinct conducting nature of the CuP COF was also reflected in the
photoconductivity, with a generated photocurrent of 0.6 nA upon
irradiation with visible light. Thus, the central metal ion in this kind
of porphyrin-based COF governs the type of charge-carrier motion
within the frameworks. The eclipsed stacking of this COF offers the
formation of two channels for carrier motion, that is macrocycle
and metal channels. The presence of a copper ion at the core of the
porphyrin macrocycle leads to the formation of a ligand-to-metal
CT complex, which reduces the significant electron density on the
porphyrin wall. Consequently, the metal-on-metal ordering on a CuP
COF favors electron transport along the channel with high carrier
mobility.

5.5 Ambipolar Semiconducting 2D COFs


As to the development of organic electronics, in particular for efficient
solar cells with control over change dynamics, the macroscopic
crystallization of the donor and acceptor moieties of the separated
domains of donor and acceptor phases with control over their
morphology is one of the ongoing challenges. In common materials,
donor and acceptor moieties tend to stack one on top of the other to
form a disordered assembly, where change carriers are trapped and
readily annihilated through rapid combination. Such hurdles could
be easily overcome by embedding a donor and an acceptor moiety
into a 2D COF framework, which favors the production of periodic
and unidirectional donor-on-donor and acceptor-on-acceptor
columnar arrays with an ambipolar semiconducting behavior. To
achieve the ambipolar semiconducting nature in COFs, various
research groups either used postsynthesis functionalization or used
donor and acceptor monomers directly in the synthesis.
At first, Jiang et al. demonstrated that donor and acceptor
molecules can be embedded in a 2D COF framework, where self-
sorting and co-crystallization between them leads to the formation
of vertical unidirectional donor-on-donor and acceptor-on-
acceptor columnar arrays [28]. These crystalline self-sorting and
bicontinuous segregation alignments offer periodic independent
pathways that allow ambipolar electron and hole conductions via
156 Energy Storage Applications of 2D COFs

acceptor-on-acceptor and donor-on-donor columns. Solvothermal


condensation of HHTP as donor and BTDADA as acceptor leads to
the formation of a donor-acceptor COF as a crystalline orange belt
(Fig. 5.8a). Different characterization details exposed at the
macroscopic level of the AA stacking mode shape the covalent
2D sheets into a framework with a pore diameter of 28 Å and a
surface area of 2021 m2/g. The distinct signature in the electronic
absorption spectra at 425 nm with a decreasing HOMO-LUMO gap
from 1.66 eV for the monolayer to 1.49 eV for the AA stacking forms
indicates the charge-transporting ability in the stacked structure. FP-
TRMC techniques in an Ar atmosphere were used to measure hole
mobilities (fSm), including both electrons and holes, while a SF6
atmosphere was employed for the measurement of hole mobilities
(mh). Accordingly, the hole and electron mobilities of a 2D donor-
acceptor COF are 0.01 and 0.04 cm2 V–1 s–1, respectively. A significant
increase in current from 0.8 pA to 10.1 nA with a linear I-V curve was
observed upon irradiation with visible light, which can be switched
on and off many times without deterioration. They further highlight
the choice of an appropriate donor-acceptor pair to explain the role
of the lattice structure in CT and separation in donor-acceptor COFs
with a pore size as large as 5.3 nm [29]. By exploiting triphenylene
as a common vertex and two different diimides as edge units, Jiang
et al. achieved two hexagonal triphenylene-diimide COFs (DTP-ANDI
COF and DTP-APyDI COF) with a high surface area and crystallinity
(Fig. 5.8b). In comparison with individual counterparts, steady-state
absorption spectra of the DTP-APyDI COF showed a clear CT band,
whereas no such band was observed for the DTP-ANDI COF. This implies
that, independent of their lattice structures, the DTP-APyDI COF forms
a through-bond CT complex, whereas in the DTP-ANDI COF the donor
and acceptor columns are independent at the ground state, giving
rise to a neutral system ready for charge separation. Time-resolved
ESR (TR-ESR) spectroscopy uncovered a rapid increase in intensity
(up to 2.5 ms) for the DTP-ANDI COF as a result of charge separation.
In sharp contrast, the DTP-APyDI COF did not show any signals in the
TR-ESR measurements under otherwise identical conditions. The
fluorescence lifetime of the DTP-ANDI COF measured by using time-
resolved fluorescence spectroscopy revealed a lifetime of 0.82 ns
for an estimated charge-separation efficiency of 84%. Thus, the
appropriate choice of the donor and acceptor pair is crucial to form
Ambipolar Semiconducting 2D COFs 157

the charge-separated state in COFs, a key process to achieve photo-


energy conversion.

Figure 5.8 (a) Schematic representation and partial chemical structure of 2D


donor-acceptor COF with self-sorting periodic donor-acceptor ordering and
bicontinuous conducting channels. (b) Partial chemical structure of DTP-ANDI
COF and DTP-APyDI COF. Reproduced with permission from Ref. [28]. Copyright
(2012), John Wiley and Sons. Republished with permission of Royal Society
of Chemistry, from Ref. [29], copyright (2013); permission conveyed through
Copyright Clearance Center, Inc.

Bein et al. reported a porphyrin- and triphenylene-containing


TP-P COF featuring ordered columns of donor and acceptor moieties
within its framework that promote paths for exciton and charge-
carrier migration upon photoexcitation of either building block
(Fig. 5.9) [30]. Photoinduced absorption spectroscopy measured
on the TP-P COF film grown on fused silica substrates exhibits two
characteristic bands centered at 700 and 960 nm, like the sum of the
158 Energy Storage Applications of 2D COFs

Figure 5.9 (a) Partial chemical structure of TP-P COF with an illustration of
the TP-P COF highlighting the alternating columns of triphenylene (red) and
porphyrin (blue) subunits. (b) Frontier orbital energy diagrams of the two
COF subunits measured by DPV in solution and a schematic illustration of
the photoinduced charge transfer. (c) Photoinduced absorption spectroscopy
spectrum of the TP-P COF film after excitation at 470 nm (blue squares; the
blue line) together with the radical ion absorption spectra a 1:1 ratio of the
two species. After photoexcitation, the TP-P COF film shows two absorption
bands in the range of the free radical ion absorption, indicating electron
transfer from the donor to the acceptor moiety within the network. Reprinted
with permission from Ref. [30]. Copyright (2014) American Chemical Society
<https://pubs.acs.org/doi/10.1021/ja509551m>. Further permissions related
to the material excerpted should be directed to the ACS.

spectra of electron acceptor (P–) and donor (TP+) free radical ions
(Fig. 5.9c). The energy-level diagram calculated from differential
pulse voltammetry (DPV) techniques further supports this claim
and indicates the possibility of efficient electron transfer within
the framework (Fig. 5.9b). Vertically oriented thin 50 nm COF film
sandwiched between ITO/MoOx and ZnO/Al electrodes revealed the
short-circuit current about 30 times higher than that of a reference
device based on a randomly intermixed blend of the two building
blocks. Upon illuminating the device with simulated solar light, the
external quantum efficiency could be boosted to more than 30%
at 350 nm and well above 10% up to 450 nm under a reverse bias.
Quantum efficiency measurements in the presence of an external
Ambipolar Semiconducting 2D COFs 159

collection field show the potential of this novel device concept,


provided that recombination losses can be minimized from further
improvements in the electron and hole transport properties of these
materials. Jiang and coworkers also highlighted a 2D porphyrin COF
(ZnP COF), where the porphyrin moiety coordinated with zinc ion
revealed an ambipolar conducting nature, as shown in Fig. 5.4 [20].
Intrinsic charge-carrier mobility of the ZnP COF measured by FP-
TRMC methods shows comparable carrier mobilities for both holes
and electrons (fme = 5.4 × 10–5 cm2 V–1 s–1, fmh = 3.36 × 10–5 cm2 V–1
s–1). Transient absorption spectroscopic measurements exhibited
the maximum quantum yield for photocarrier generation (f) to be
1.9 × 10–3, and the me and mh values were 0.016 and 0.032 cm2 V–1
s–1, respectively. Photoconductivity measurements upon irradiation
with visible light generated a photocurrent of approximately 26.8
nA with a large on-off ratio (5 × 104). Thus, an ambipolar conducting
nature with a balanced hole and electron transporting ability drives
the production of a large photocurrent as well as a high on-off ratio,
as compared to unbalanced carrier transport.
Expanding on the intracrystalline donor-acceptor theme, Jiang et
al. further elaborated their search on COF CT and separation, with a
series of 2D COFs embedded with different metallophthalocyanates
and diimides as electron-donating and -accepting building blocks
(Fig. 5.10) [31, 32]. A series of COFs, namely DMPc-ADI COFs, were
prepared by using solvothermal reactions, and it was shown that
polycondensation is widely applicable for various MPcs, including
copper, nickel, and zinc, with diimide and perylene diimide.
Experimental and theoretical structural parameters were studied
in all these COF frameworks, and it was found that MPc donor and
diimide acceptor units are exactly linked and interfaced to form
ordered segregated yet bicontinuous periodic p arrays.
All these COFs have light absorption over a wide range of the
visible spectrum and into the NIR up to 1100 nm. The absence of
the absorption characteristic after 1500 nm for DMPc-ADI COFs ruled
out the possibility of ground-state interactions between MPc and
DI to form a CT complex and support the existence of independent
p columns. This type of structural feature offers two classes of p
columns—four proximate directly linked central donor pairs and
eight remote pairs without direct linkage—with different spatial
center-to-center distances (<2 nm for the proximate pair and >4
nm for the remote pair). Thus, in such a framework four available
160 Energy Storage Applications of 2D COFs

equivalent acceptors are stacked in proximate columns for accepting


the electron from an excited donor unit. Furthermore, the discrete
center-to-center distances on DMPc-ADI COFs allow independent
pathways for exciton migration and hole and electron transport
along with fine screening to achieve long-lived charge separation.
Time-resolved spectroscopy measured a dispersed solution of DMPc-
ADI COFs in benzonitrile at room temperature upon irradiation of
singlet spin and charge migration following the charge separation.

Figure 5.10 Schematic representation of the donor-acceptor COFs (DMPc-


ADI COFs) with covalently linked phthalocyanine diimide (MPc-DI) structures.
Reprinted with permission from Ref. [31]. Copyright (2015) American Chemical
Society.

The lifetime values for charge separation were relatively longer


on DCuPc-ADI COFs compared to those on DNiPc-ADI COFs and DZnPc-ADI
COFs. The higher oxidation states of CuPc compared to NiPc and ZnPc
may cause charge recombination to occur by providing much higher
larger driving forces for back electron transfer, responsible for this
trend. A TR-ESR signal up to 1.35 ms was observed for DZnPc-APDI
COFs, in sharp contrast to DZnPc-APDI COFs, where the paramagnetic
nature of copper complicates the resolution of the spectral profile.
In addition, the DPV measurements with a dispersion of DZnPc-
APDI COFs with respect to a Ag/AgNO3 reference electrode showed
Ambipolar Semiconducting 2D COFs 161

oxidation of the ZnPc at 0.42 eV and a reduction of the naphthalene


diimide (NDI) at 0.52 eV, which the authors interpreted as a large
exothermic driving force for electron transfer. Thus, the structural
feature in these donor-acceptor COFs and resulting insights into
the mechanistic aspects of charge separation make it possible to
extract holes and electrons for electric current production as well as
constitute a basis for developing COFs for photovoltaic applications.

Figure 5.11 (a) Schematic representation of the reaction between HAB and
HKH to provide the C2H-h2D crystal and their partial chemical structures. (b)
An atomic force microscopy (AFM) image of the C2H-h2D crystal on Cu (111).
(c) AFM image of the C2H-h2D crystal with height profile along the cyan-blue
line (scale bar: 7 mm). (d) Optical microscopy image of a C2H-h2D crystal FET
prepared on a SiO3 (300 nm)/n++ Si wafer. The inset is an optical microscopy
image taken before the deposition of Au electrodes on the crystal. Reproduced
from Ref. [33] under a Creative Commons Attribution 4.0 International Licence
(https://creativecommons.org/licenses/by/4.0/).

Recently, Baek and coworkers reported a 2D-layered network


structure consisting of evenly distributed periodic holes and
nitrogen atoms with a C2N stoichiometry in its basal plane called
“nitro-generated holey graphene” (C2H-h2D), as shown in Fig.
162 Energy Storage Applications of 2D COFs

5.11 [33]. The wet chemical reaction between hexaaminobenzene


trihydrochloride (HAB) and hexaketocyclohexane octahydrate
(HKH) in N-methyl-2-pyrrolidone in the presence of sulfonic acid
(catalytic amount) leads to the formation of a holey graphene
framework, where phenyl rings are bridged by pyrazine rings.
Density functional theory (DFT) calculations revealed a remarkably
high potential energy (89.7 kcal U mol–1) gain by the aromatization
drive in this polycondensation reaction, which favors the formation
of a 2D network structure. Further, the magnitude of the bandgap
and the existence of flat bands near the Fermi levels support that
the C2H-h2D as the active layer with a thickness of 8 nm, showing
ambipolar charge transport with an electron mobility of 13.5 cm2
V–1 s–1 and a hole mobility of 20.6 cm2 V–1 s–1. The unintentional
doping effects by the trapped impurities and/or absorbed gases on
the open framework are responsible for this behavior and further
indicate that the electronic properties are tunable.
Jiang et al. further developed a strategy to spatially confine
electron acceptor COFs with the open channels of electron-donating
frameworks by exploring covalent click chemistry to achieve donor-
acceptor COFs with an ambipolar conducting nature [34]. In such
systems photoinduced electron transfer and charge separation, along
with carrier concentration efficiency, largely depend on the acceptor
content in the open pore. They developed a three-component
topological design in conjunction with click chemistry for adapting
an open-lattice COF into ordered donor-acceptor arrays with a
photoelectric structure. A series of electron-donating ZnPc COFs
were synthesized by three-component solvothermal condensation
with Zn-phthalocyanine as the vertex and varying the molar ratio
of BDBA and 2,5-bis(azidomethyl)-1,4-phenylenediboronic acid
as edges (Fig. 5.12). The click reaction between the azide moieties
of the channel wall with the alkyne derivatives of fullerene leads
to channel confinements with the lattice altering into an ordered
donor-acceptor system. A significant increase on the absorption
band at 451 nm responsible for the fullerene moiety observed with
increasing molar percentage of the azide unit on ZnPc COFs further
supports this strategy. Upon the laser flash at 680 nm, the electron-
donating ZnPc COF revealed a broad TR-ESR signal generated from
the excited triplet states without the generation of charge-separated
states. A drastic change of this on integration of electron-accepting
Ambipolar Semiconducting 2D COFs 163

fullerene molecules into the channel walls indicates light irradiation


trigger electron transfer from the cation radical ZnPc•+ to the anion
radical C60•– and delocalization of the radical cations and anions in the
ZnPc and C60 columns. A significant increase in the signal intensity
with increasing fullerene contents indicates that the radical species
are increased in the framework, which increases the efficiency of the
photoinduced electron transfer and charge separation.

Figure 5.12 (A) Partial chemical structure of ZnPc COFs and the reaction to
integrate a C60 molecule on the channel wall. (B) Schematic representation
of top and side views of ZnPc COFs with C60 integrated on the channel walls.
Reprinted with permission from Ref. [34]. Copyright (2014) American Chemical
Society.

As to an alternative strategy, two groups researched a


supramolecular approach by spatially confining fullerene molecules
as electron acceptors within the open channels of electron-donating
frameworks, for converting the open lattice of COFs into donor-
acceptor photoelectric structures. Knochel et al. reported a COF-
based photovoltaic device where light-induced charge separation and
164 Energy Storage Applications of 2D COFs

collection is clearly feasible with an electron donor in the walls and


the complementary electron acceptor in ordered periodic channels
(Fig. 5.13) [35]. Under solvothermal conditions, the condensation of
thieno[3,2-b]thiophene-2,5-diyldiboronic acid with HHTP leads to
the formation of thienothiophene-based TT COF with a surface area
of 1810 m2/g and a pore diameter of 3 nm. This open framework
allows the uptake of electron acceptor [6,6]-phenyl-C60-butyric
acid methyl ester (PCBM) and favors the formation of a periodic
ordered donor-acceptor network. Spectroscopic results containing

Figure 5.13 (a) Chemical structure of TT COF. (b) Schematic representation of


the host-guest complex and a PCBM molecule to scale (C: gray; O: red; B: green;
S: yellow). Reproduced from Ref. [35]. Copyright (2013), John Wiley and Sons.
Ambipolar Semiconducting 2D COFs 165

absorption, emission, and time-resolved data demonstrate light-


induced efficient CT within the lifetime of the excitons from the
photoconductive TT COF donor network to the encapsulated PCBM
phase thickness into a ITO/TT-COF:PCBM/Al device structure, which
reveals a linear I-V profile with a power conversion efficiency of
0.053% under illumination with simulated solar light. The maximum
external quantum efficiency of 3.4% at 405 nm matches well with
the absorption characteristics of the COF film, whereas the tail at
wavelengths above 700 nm indicates that the characteristic feature
of PCBM with both interpenetrating networks is photoelectrically
active.

Figure 5.14 (a) Synthetic scheme of a CS COF. (b) Schematic representation


of the synthesis of CS-COF-C60 by sublimed crystallization of fullerene in the
open 1D channels (white: carbon; red: nitrogen; purple: fullerene). Reprinted
by permission from Springer Nature Customer Service Centre GmbH: Springer
Nature, Nature Communications, Ref. [36], copyright (2013).
166 Energy Storage Applications of 2D COFs

More recently, Jiang et al. developed the synthesis of a crystalline


phenazine-linked CS COF with an ordered stable structure and
delocalized p clouds by topological ring fusion of triphenylene
hexamine and tert-butylpyrene tetraone as the monomers
(Fig. 5.14) [36]. The FP-TRMC method showed exceptional mobility
of 4.2 cm2 V–1 s–1, suggesting that the CS COF is a high-rate hole-
conducting framework, which is ranked among the top-class hole
transporting organic semiconductors. Furthermore, the micropores
of the CS COF allow complementary functionalization; physically
filling the pore with fullerene molecules converts it to a bicontinuous
order donor-acceptor system (CS-COF-C60). Slipped AA stacking
with bulky tert-butyl groups on the channel walls tolerates the
accommodation of only one C60 molecule in the pore. To explore
the possibility of photoenergy conversion, PCBM was used as
a glue to fabricate 1 × 1 cm2 ITO/active layer/Al cells using spin-
coated CS-COF-C60 films with an adjustable thickness of 100 nm. The
sandwiched devices with CS-COF-C60 as the photoconductive layer
display a rapid response to light irradiation with an on-off ratio of up
to 5.9 × 107 and a power conversion efficiency of 0.9%.

5.6 Lithium-Ion Batteries Using 2D COFs as


Electrodes
Now, Li-ion batteries (LIBs) have attracted a wide range of interest
because they can directly store energy, have great success in portable
electronic products, and are a clean energy source that is expected
to accelerate the development of electric vehicles [37, 38]. However,
most commercially available LIBs are transition metal oxide cathodes,
which will place a heavy burden on the ecological environment. For
sustainable development, the battery cathode should preferably be
made of highly efficient and environmentally friendly materials. For
example, Pang et al. used a sulfur cathode—for example, in lithium-
sulfur (Li-S) batteries—to take the place of conventional insertion
cathode materials, or used organic cathode materials instead
of transition metal oxide cathodes because they do not contain
transition metal species [39, 40]. Porous COF materials possess high
surface areas, and current efforts toward utilizing porous organic
polymers in energy storage applications mainly focus on capacity
Lithium-Ion Batteries Using 2D COFs as Electrodes 167

improvement by the formation of electrochemical double layers,


and no solid-state redox reactions take place within the framework
itself. In this case, COFs as an electrode material in energy storage
have attracted attention [41]. The structural advantages of 2D COFs,
with ultrahigh surface areas, tunable pore sizes and shapes, and the
ability to exhibit nanoscale effects on the pore wall, offer advantages
for application of these kinds of materials for energy conversion and
storage devices. This section on LIBs focuses on the application of
2D COF materials in battery anodes and cathodes, demonstrating
the superior properties of this material in terms of electrical energy
storage and being an environmentally friendly alternative.

5.6.1 Battery Cathode Application


Li-S batteries have relatively high theoretical energy density and
gravimetric capacity and low production cost. They are considered
to be promising energy storage devices, have received extensive
attention, and are worthy of vigorous development [42, 43].
However, Li-S batteries possess the following problems in practical
applications [44–47]. In the course of repeated charge and discharge
cycles, a succession of structural and morphological alternations can
occur in cyclo-S8. This issue leads to the formation of soluble lithium
polysulfides Li2Sx and insoluble Li2S2/Li2S. The polysulfides will
dissolve in the electrolyte during the charging process of the battery,
may shuttle back and forth between the anode and the cathode, and
will have a side reaction with the anode and a reoxidation reaction
with the cathode. These possible reactions will result in poor battery
efficiency and stability [41, 43, 47].
Two-dimensional COFs are ordered structures with rigid
covalent bond networks and pores. When applied to Li-S batteries,
they may be conducive to the transport of electrolyte ions. In
addition, the conductivity of sulfur can be increased by using the
conductivity of 2D COFs. Wang et al. utilized CTF-1 (pore size of 1.23
nm) as a new host material for the insertion of sulfur to capture
soluble polysulfides and improve the stability of the battery [48].
Figure 5.15A shows a sulfur-containing COF obtained by the melting
and diffusion methods. The main experimental object of this
experiment was CTF-1/S@155°C, which was obtained by heating a
mixture of CFT-1 and sulfur (3:2) at 155°C for 15 h. Simple physical
168 Energy Storage Applications of 2D COFs

mixing of CTF-1 and sulfur, a mixed product labeled CTF-1/S@RT,


was used for easy comparison [49, 50].
The XRD patterns of sulfur, CTF-1, and CTF-1/S@155°C were
examined, as shown in Fig. 5.15B. Unlike sulfur and CTF-1, which
have a number of strong peaks, CTF-1/S@155°C showed featureless
and weak bands. CTF-1/S@155°C disappears on XRD patterns
and the elements that make up CTF-1 are lightweight. This may
indicate that sulfur is well dispersed within CTF-1 at the nanoscale
and subnanometer levels. In addition, Pang et al. observed that the
specific surface area (789–1.6 m2/g) and pore volume (0.3–0.0036
cm3/g) decrease sharply. This further confirmed that the pores of
CTF-1 have been filled with sulfur and that sulfur occupies almost all
of the pores [50].
The discharge curve of a typical CTF-1/S@155°C cathode
(0.1 C, 1.1–3 V) is shown in Fig. 5.15C. The stage potential profile
distribution is generated at the cathode of CTF-1/S@155°C. A
shorter plateau is in a higher potential (2.3 V), which is consistent
with the first reduction of sulfur to lithium polysulfides (Li2Sn, 2 <
n < 8), and the longer plateau is at a lower potential (~2.1 V), which
is consistent with the solid lithium sulfides (Li2S2 and LiS) that
are produced by further reduction of polysulfide. In Fig. 5.15C, the
potential hysteresis between discharge and charge (0.2 V) shows
the rapid kinetics of the transition between lithium polysulfide and
lithium sulfide in the pores of the CTF-1 substrate. Obviously, the
electrical touch of both embedded sulfur and the CTF wall facilitates
these fast kinetics [50].
Figure 5.15D exhibits the cycling performance of the two ma-
terials acting as cathodes. In the first cycle, compared to the cath-
ode discharge capacity for CTF-1/S@RT, CTF-1/S@155°C displayed
higher values, indicating that the diffusion process is favored at high
temperatures. Also, the S8 molecule may be strongly adsorbed in the
nanopore of CTF-1 and the charge–discharge coulombic efficiency
of the composite material is observed to be close to ~97%, as a re-
sult of which the soluble polysulfide intermediate can effectively
inhibit the shuttle effect. After the third cycle, the capacity value be-
comes relatively stable, and after 20 cycles, the value is maintained
at approximately 1000 mAh g–1. It can be stated that the stable cycle
performance of the CTF-1/S@155°C cathode is determined by the
CTF-1 substrate and sulfur loading in CTF-1 pores can significantly
increase the sulfur electrode cycle performance.
Lithium-Ion Batteries Using 2D COFs as Electrodes 169

Figure 5.15 (A) Schematic diagram of the composite synthesis from CTF-1
by impregnation of molten sulfur (C: gray; N: blue; H: red; S: yellow) and SEM
image of (a) the corresponding elemental mapping of sulfur (b) for the CTF-
1/S@155°C composite. (B) Galvanostatic discharge and charge profiles of the
CTF-1/S@155°C composite at a 0.1 C rate. (C) XRD patterns of sublimed sulfur
(a), CTF-1 (b), and the CTF-1/S@155°C composite (c). (D) Cycling performance
of the CTF-1/S@155°C composite and the CTF-1/S@RT composite at a 0.1 C
rate. (E) Discharge capacity for the CTF-1/S@155°C composite at different
rates. Republished with permission of Royal Society of Chemistry, from Ref. [48],
copyright (2014); permission conveyed through Copyright Clearance Center, Inc.
170 Energy Storage Applications of 2D COFs

Even if the current density increases, the discharge capacity does


not drop rapidly, as shown in Fig. 5.15E. Although when the high
current is high (1 C), the CTF-1/S@155°C cathode exhibits excellent
rate performance, it can provide a reversible capacity of 541 mAh
g–1. More importantly, if the current density of 0.1 C is again applied,
the capacity value will return to approximately 750 mAh g–1.
Apparently, due to the presence of CTF-1 substrates with ordered
nanoporous structures, composite materials yield excellent rate
performance, greatly improve the sulfur conductivity, and provide a
tunnel for electrolyte penetration. This work offered possibilities for
the development of 2D COFs in new energy-related applications.
With carbon nanotubes (CNTs) as conducting materials, Jiang et
al. demonstrated a new method to employ the CNT–2D COF composite
as an electrode [51]. Through an in situ polycondensation process,
the redox-active COF (Fig. 5.16, DTP-ANDI COF), which contained
NDI units, could be formed on the surface of CNT wires. Due to the
redox properties of NDI units, the as-formed 2D COFs and DTP-ANDI-
COF@CNTs composite were redox-active. From the XRD analysis
of the composite, it was found that the 2D COFs formed ordered
mesoporous channels and adopted an AA stacking mode. Using this
composite as the cathode, the performance was investigated. After
100 cycles, the LIBs with DTP-ANDI-COF@CNT cathodes could present
high energy-storage stability and nearly 100% coulombic efficiency.
In the charge–discharge curves presented as shown in Fig. 5.16g,
no obvious polarization could be observed as the profiles retained
similar shapes. This indicated that both electrons and ions inside the
batteries could be transported efficiently, which could be beneficial
for the rapid charge–discharge processes. Despite the relatively low
capacity, the battery showed good stability. When the current density
increased to 12 C, the battery could still retain 85% of the capacity
value at 2.4 C. At the rate of 2.4 C, the long-term stability was tested
(Fig. 5.16i). Both capacity and coulombic efficiency were almost
unchanged even after 700 cycles, indicating its excellent stability. The
above examples imply that the pores inside 2D COFs could be used
to trap energy-related species such as sulfur, polysulfides, and ions.
What’s more, when the 2D COFs are used in batteries, the advantage
of possible conductivity against other p-conjugated materials could
contribute to the final device performance.
Lithium-Ion Batteries Using 2D COFs as Electrodes 171

Figure 5.16 (a) Representative chemical structure of the DTP-ANDI COF


composite. (b) Redox-active positions on the naphthalene diimide unit. (c)
Pictures of the prepared coin-type lithium-sulfur batteries. (d) Schematic
representation of the composite (gray color represents the CNTs). (e) Charge–
discharge profiles of the composite cathodes (2.4 C; red is the 1st cycle, while
blue is the 80th cycle). (f) Capacities of the composite cathodes and CNT (black
dotted line) batteries and coulombic efficiency of the composite cathodes for
100 cycles. (g) Charge–discharge profiles of the composite cathodes at various
charge–discharge rates. (h) Discharge capacity results of the composite cathodes
at different rates. (i) Long-time stability result of the composite cathodes at the
rate of 2.4 C. Reprinted by permission from Springer Nature Customer Service
Centre GmbH: Springer Nature, Scientific Reports, Ref. [51], copyright (2015).
172 Energy Storage Applications of 2D COFs

5.6.2 Battery Anode Application


Besides 2D COFs as cathode electrodes in LIBs, they could also serve
as anode materials. For example, Zhao et al. recently reported two
novel 2D crystalline COFs that could exhibit good selectivity for
different gases (15:1 and 7:1 for H2 to N2 and CO2 to N2, respectively).
More interestingly, when used as anodes, the porous 2D COFs also
showed superior performance in Li-ion storage [52]. The average
diameters of channels for N2 COF and N3 COF were measured to be
23 Å and 16 Å, respectively (Fig. 5.17). Owing to the porosity and
high conductivity, the 2D COFs could exhibit better performance
than other related porous materials, such as silicon and metal-
organic frameworks (MOFs). In the charge–discharge curves
presented in Figs. 5.18a and 5.18b, the 2D COF–based batteries
didn’t exhibit obvious polarization, which could be beneficial to both
charge and discharge processes. For the batteries based on N2 COF
(Fig. 5.18c), the charge capacity was 497 mA h g–1 at the rate of 5 C.
The capacity could increase to 607 mA h g–1 when a current density
of 2 C was adopted. For the N3-COF-based batteries (Fig. 5.18d),
a much better performance was observed. The charge capacity
values were maintained at ~520 mA h g–1 at the relatively high rate
of 5 C. Furthermore, the capacity value of ~610 mA h g–1 could be
recovered when the current density of 2 C was applied again. The
cycling experiments (Fig. 5.18e,f) disclosed that the stabilities of
both N2-COF- and N3-COF-based batteries were excellent.
Good conductivity is vital for electrode materials in LIBs with
high performance. In the reported results, 2D COFs were mainly
constructed through covalent connections based on several building
blocks, such as boronate esters, borazine, enamine, imine, hydrazone,
and triazine [53–57]. Generally, the conductivity of the as-formed
COFs is poor compared to that of the conductive polymers. Porphyrin
is a universal unit with unique optoelectronic properties, and it has
also been incorporated into several 2D COFs. However, limited by
the covalent linkage with boronic acids, its conductivity is not high
enough for its application as an electrode material. With the purpose
of further enhancing the conductivity of porphyrin-based 2D COFs,
(4-thiophenephenyl)porphyrin (TThPP) was used as the building
units for application as electrode materials (Fig. 5.19) [58] since
thiophene is a universal unit in organic electronics, which could
Lithium-Ion Batteries Using 2D COFs as Electrodes 173

increase the planarity and electrical conductivity of the as-prepared


materials.

Figure 5.17 Structural presentation of the 2D COFs (a, b) and (c, d) the layered
stacking mode (top view) based on theoretical DFT calculations (C: blue; N:
red; H: gray-white). (e, f) STM images for the two 2D COFs. Republished with
permission of Royal Society of Chemistry, from Ref. [52], copyright (2016);
permission conveyed through Copyright Clearance Center, Inc.

Figure 5.18 Charge–discharge curves of the 2D COFs (a) N2 COF and (b) N3
COF; current density of 1 Ag-1 was applied. Rate capability results at different
current densities: (c) N2 COF and (d) N3 COF. Cycling performance of N2 COF (e)
and N3 COF (f). The red profiles show the efficiency that was derived from the
ratio of charge and discharge capacities. Republished with permission of Royal
Society of Chemistry, from Ref. [52], copyright (2016); permission conveyed
through Copyright Clearance Center, Inc.
174 Energy Storage Applications of 2D COFs

Figure 5.19 (a) Scheme of synthesizing TThPP. (b) Performance of the 2D


COF–based anode; charge–discharge capacities at various rates. (c) Cycling
experiment results at the rate of 1 C. (d) Charge–discharge curves of the anode
from 1 to 200 cycles. (e) CV curves of the TThPP-based anode for 1 to 3 cycles.
Reprinted with permission from Ref. [58]. Copyright (2016) American Chemical
Society.

Using the chemical activity of thiophene units, the 2D COFs could


be prepared via an oxidation polymerization reaction [59–61]. When
the as-synthesized COFs were used as anode materials in batteries,
the capacity value was 666 mA h g–1 at the current density of 0.2 C.
At the rate of 1 C, a capacity value of 384 mA h g–1 could be observed.
Lithium-Ion Batteries Using 2D COFs as Electrodes 175

When the rate changed from 5 to 0.2 C, the capacity value could be
recovered to nearly 100% of the original capacity, indicating the
stability of the batteries. The cycling experiment results are shown
in Fig. 5.19b. After the charge–discharge process, the coulombic
efficiency was calculated to be 99.3% while the discharge capacity of
384 mA h g–1 could be maintained even after 200 charge–discharge
cycles.
Besides COFs, it is to be noted that Tarascon et al. pioneered
the usage of MOFs as electrode materials in Li-S batteries [62].
The MOF MIL100(Cr) mixed with sulfur as the composite cathode
could improve the cycling performance, even though the sulfur
content was limited. Inspired by this, Xiao et al. demonstrated
that Ni MOF (Ni6(BTB)4(BP)3 (BTB: benzene-1,3,5-tribenzoate;
BP: 4,4′-bipyridyl) could remarkably immobilize the sulfur and
polysulfide species within the cathode structure [63]. As a host
material, the Ni MOF/S composite-based batteries exhibited a
discharge capacity of 689 mA h g–1 at 0.1 C. At a current density of
2 C (Table 5.2), the composite was capable of delivering a discharge
capacity of 287 mA h g–1. Li et al. studies several MOFs as host
materials in Li-S batteries [64]. Results showed that the nanosized
ZIF-8 and MIL-53 could serve as good host materials to produce a
fast sulfur cathode with a long cycle life. Under a constant rate of 0.5
C, the maximum discharge capacities for S/HKUST-1, S/ZIF-8, and S/
MIL-53 were 431, 738, and 793 mA h g–1, respectively.
Use of graphene composites in LIBs has been extensively
pursued for a long time. The LiFePO4/graphene composite as a
cathode electrode exhibited discharge capacities of 165 and 88
mA h g–1 at 0.1 C and 10 C, respectively [65]. Explored by Choi et
al., vanadium pentoxide (V2O5) mixed with graphene as a cathode
electrode showed a discharge capacity of 94.4 mA h g–1 at 10 C,
with ultrahigh stability (100,000 cycles) [66]. Instead of cathode
materials, as anode materials, they generally showed better
performance [73]. Cheng et al. reported that composites of graphene
nanosheets (GNSs) and Fe3O4 could exhibit 580 and 520 mA h g–1 at
current densities of 0.7 and 1.75 C, respectively [67]. Furthermore,
the analogue of borocarbonitrides could also show a capacity of 150
mA h g–1 at a current density of 1 C [69]. The inorganic counterparts
MoS2 and WS2 could also be functionalized as electrode materials.
For example, Cho et al. reported MoS2 nanoplates that contained
176 Energy Storage Applications of 2D COFs

disordered graphene-like layers that could work as high-rate lithium


battery anode materials. The capacity was 700 mA h g–1 even at the
rate of 50 C [70].

Table 5.2 Performance of Li-ion batteries with representative 2D COFs and


related materials

Capacity Current
Category Materials Electrode (mA h g–1) density (C) Ref.
2D COF CTF-1/ Cathode 541 1 [50]
S@155°C
2D COF Por-COF/S Cathode 670 1 [42]
2D COF DTP-ANDI- Cathode 58 12 [51]
COF@CNT
2D COF N2 COF Anode 497 5 [52]
2D COF N3 COF Anode ~520 5
2D COF TThPP-based Anode 384 1 [58]
COF
MOF Ni MOF/S Cathode 287 2 [63]
MOF S/HKUST-1 Cathode 431 0.5 [64]
MOF S/ZIF-8 Cathode 738 0.5 [64]
MOF S/MIL-53 Cathode 793 0.5 [64]
Graphene Graphene- Cathode 88 10 [65]
wrapped
LiFePO4/C
Graphene V2O5/graphene Cathode 94.4 10 [66]
Graphene rGO/Fe3O4 Anode 520 1.75 [67]
Graphene GNS/C60 Anode 784 0.05 [68]
BxCyNz BxCxNz (CVD) Anode 150 1 [69]
MoS2 Nanosheets Anode 700 50 [70]
MoS2 TO/GS Anode 199 1 [71]
composite
WS2 Composite Anode 596 1.4 [72]

On performance comparison, it can be seen that 2D COF–based


electrodes can show capacity comparable to that of other similar
materials (NOFs, graphene, BxCyNz, and MoS2). 2D COFs are promising
Summary and Perspective 177

candidates as electrode materials in LIBs. However, how to achieve


both high capacitance and long cycling performance is still the major
bottleneck in using COFs as electrode materials for energy storage,
probably due to the structural defects, low energy level matching,
and insufficient device optimization during the charge–discharge
process. From this viewpoint, further exploration to prepare new
2D COFs and gain ideas from other better-performing materials is
highly desirable.

5.7 Summary and Perspective


This chapter introduced the recent progress in terms of some
representative 2D COFs for energy-related applications, including
conductivity study, semiconductor, lithium batteries, and capacitive
storage. To fully understand the advantages and disadvantages
of 2D COFs, a comparison study with other related materials (e.g.,
MOFs, graphene, borocarbonitrides, and MoS2) was also conducted.
Although research on 2D COFs has already attracted some attention,
poor solubility is still the main obstacle for their usage beyond gas
adsorption and storage (described in Chapter 3). In fact, according
to literature results, energy-related applications of 2D COFs are still
in their infancy. However, with more innovation, these unique 2D
materials should have more potential applications. For example, in
rechargeable lithium batteries, incorporation of sulfur into the pores
of 2D COFs is a smart way to make them useful as electrode materials
in energy-related applications. Also, utilizing the effective click
reaction to introduce redox groups into conventional COFs would be
a good way to provide 2D COFs with more chances for application in
capacitive energy storage.
Finally, the true structure of the COF and the layer stacking
sequence (interlaced or overlapping) remains controversial.
Furthermore, the chemical stability of crystalline and ordered
COF materials is poor, primarily because the formation reaction is
limited by the reversible reaction of the formation of borate and
imine double bonds. Therefore, another important aspect that needs
to be explored is the development of a novel reaction and joining
method to synthesize a class of COFs with long-range order and high
chemical stability.
178 Energy Storage Applications of 2D COFs

Although many challenges still exist, numerous possibilities are


offered within this fascinating area and the authors of this book
envision a bright future for the development of COF materials for
application across fields, which will grow into an important research
area with great potential.

References

1. Huang, N., Wang, P., Jiang, D. (2016). Nat. Rev. Mater., 1, 16068.
2. Dogru, M., Bein, T. (2014). Chem. Commun., 50, 5531–5546.
3. Feng, X., Ding, X., Jiang, D. (2012). Chem. Soc. Rev., 41, 6010–6022.
4. Mahmood, J., Lee, E. K., Jung, M., Shin, D., Choi, H.-J., Seo, J.-M., Jung,
S.-M., Kim, D., Li, F., Lah, M. S., Park, N., Shin, H -J., Oh, J. H., Baek, J.-B.
(2016). Proc. Natl. Acad. Sci., 113, 7414–7419.
5. Huang, N., Krishna, R., Jiang, D. (2015). J. Am. Chem. Soc., 137, 7079–
7082.
6. Huang, N., Chen, X., Krishna, R., Jiang, D. (2015). Angew. Chem. Int. Ed.,
54, 2986–2990.
7. Xu, H., Tao, S., Jiang, D. (2016). Nat. Mater., 15, 722–726.
8. Ma, H., Liu, B., Li, B., Zhang, L., Li, Y.-G., Tan, H.-Q., Zang, H.-Y., Zhu, G.
(2016). J. Am. Chem. Soc., 138, 5897–5903.
9. Dalapati, S., Jin, E., Addicoat, M., Heine, T., Jiang, D. (2016). J. Am. Chem.
Soc., 138, 5797–5800.
10. Ding, S.-Y., Dong, M., Wang, Y.-W., Chen, Y.-T., Wang, H.-Z., Su, C.-Y., Wang,
W. (2016). J. Am. Chem. Soc., 138, 3031–3037.
11. Ostroverkhova, O. (2016). Chem. Rev., 116, 13279–13412.
12. Wan, S., Guo, J., Kim, J., Ihee, H., Jiang, D. (2008). Angew. Chem. Int. Ed.,
47, 8826–8830.
13. Wan, S., Guo, J., Kim, J., Ihee, H., Jiang, D. (2009). Angew. Chem. Int. Ed.,
48, 5439–5442.
14. Ding, X., Guo, J., Feng, X., Honsho Y., Guo, J., Seki, S., Maitarad, P., Saeki,
A., Nagase, S., Jiang, D. (2011). Angew. Chem. Int. Ed., 50, 1289–1293.
15. Ding, X., Feng, X., Saeki, A., Seki S., Nagai, A., Jiang, D. (2012). Chem.
Commun., 48, 8952–8954.
16. Liao M.-S., Scheiner, S. (2001). J. Chem. Phys., 114, 9780–9791;
17. Liao M.-S., Watts, J. D., Huang, M.-J., Gorun S. M., Kar, T., Scheiner, S.
(2005). J. Chem. Theory Comput., 1, 1201–1210.
References 179

18. Dalapati, S., Addicoat, M., Jin, S., Sakurai, T., Gao, J., Xu, H., Irle, S., Seki,
S., Jiang, D. (2015). Nat. Commun., 6, 7786.
19. Wan, S., Gμndara, F., Asano, A., Furukawa, H., Saeki, A., Dey, S. K., Liao,
L., Ambrogio, M. W., Botros, Y. Y., Duan, X., Seki, S., Stoddart, J. F., Yaghi,
O. M. (2011). Chem. Mater., 23, 4094–4097.
20. Feng, X., Liu, L., Honsho, Y., Saeki, A., Seki, S., Irle, S., Dong, Y., Nagai, A.,
Jiang, D. (2012). Angew. Chem. Int. Ed., 51, 2618–2622.
21. Segura, J. L., MartÌn, N. (2001). Angew. Chem. Int. Ed., 40, 1372–1409.
22. Bertrand, G. H. V., Michaelis, V. K., Ong, T.-C., Griffin, R. G., Dincã, M.
(2013). Proc. Natl. Acad. Sci., 110, 4923–4928.
23. Duhović, S., Dincã, M. (2015). Chem. Mater., 27, 5487–5490.
24. Jin, S., Sakurai, T., Kowalczyk, T., Dalapati, S., Xu, F., Wei, H., Chen, X., Gao,
J., Seki, S., Irle, S., Jiang, D. (2014). Chem. Eur. J., 20, 14608–14613.
25. Ding, H., Li, Y., Hu, H., Sun, Y., Wang, J., Wang, C., Wang, C., Zhang, G.,
Wang, B., Xu, W., Zhang, D. (2014). Chem. Eur. J., 20, 14614–14618.
26. Cai, S.-L., Zhang, Y.-B., Pun, A. B., He, B., Yang, J., Toma, F. M., Sharp, I. D.,
Yaghi, O. M., Fan, J., Zheng, S.-R., Zhang, W.-G., Liu, Y. (2014). Chem. Sci.,
5, 4693–4700.
27. Ding, X., Chen, L., Honsho, Y., Feng, X., Saengsawang, O., Guo, J., Saeki, A.,
Seki, S., Irle, S., Nagase, S., Parasuk, V., Jiang, D. (2011). J. Am. Chem. Soc.,
133, 14510–14513.
28. Feng X., Chen, L., Honsho, Y., Saengsawang, O., Liu, L., Wang, L., Saeki, A.,
Irle, S., Seki, S., Dong, Y., Jiang, D. (2012). Adv. Mater., 24, 3026–3031.
29. Jin, S., Furukawa, K., Addicoat, M., Chen, L., Takahashi, S., Irle, S.,
Nakamura, T., Jiang, D. (2013). Chem. Sci., 4, 4505–4511.
30. Calik, M., Auras, F., Salonen, L. M., Bader, K., Grill, I., Handloser, M.,
Medina, D. D., Dogru, M., Lçbermann, F., Trauner, D., Hartschuh, A., Bein,
T. (2014). J. Am. Chem. Soc., 136, 17802–17807.
31. Jin, S., Supur, M., Addicoat, M., Furukawa, K., Chen, L., Nakamura, T.,
Fukuzumi, S., Irle, S., Jiang, D. (2015). J. Am. Chem. Soc., 137, 7817–
7827.
32. Jin, S., Ding, X., Feng, X., Supur, M., Furukawa, K., Takahashi, S., Addicoat,
M., El-Khouly, M. E., Nakamura, T., Irle, S., Fukuzumi, S., Nagai, A., Jiang,
D. (2013). Angew. Chem. Int. Ed., 52, 2017–2021.
33. Mahmood, J., Lee, E. K., Jung, M., Shin, D., Jeon, I.-Y., Jung, S.-M., Choi, H.-
J., Seo, J.-M., Bae, S.-Y., Sohn, S.-D., Park, N., Oh, J. H., Shin, H.-J., Baek, J.-B.
(2015). Nat. Commun., 6, 6486.
34. Chen, L., Furukawa, K., Gao, J., Nagai, A., Nakamura, T., Dong, Y., Jiang, D.
(2014). J. Am. Chem. Soc., 136, 9806–9809.
180 Energy Storage Applications of 2D COFs

35. Dogru, M., Handloser, M., Auras, F., Kunz, T., Medina, D., Hartschuh, A.,
Knochel, P., Bein, T. (2013). Angew. Chem. Int. Ed., 52, 2920–2924.
36. Guo, J., Xu, Y., Jin, S., Chen, L., Kaji, T., Honsho, Y., Addicoat, M. A., Kim, J.,
Saeki, A., Ihee, H., Seki, S., Irle, S., Hiramoto, M., Gao, J., Jiang, D. (2013).
Nat. Commun., 4, 2736.
37. Xie, J., Gu, P., Zhang, Q. (2017). ACS Energy Lett., 2, 1985–1996.
38. Xie, J., Wang, Z., Xu, Z. J., Zhang, Q. (2018). Adv. Energy Mater., 8,
1703509.
39. Zhan, X., Chen, Z., Zhang, Q. (2017). J. Mater. Chem. A, 5, 14463.
40. Yang, X., Dong, B., Zhang, H., Ge, R., Gao, Y., Zhang, H. (2015). RSC Adv.,
5, 86137.
41. Zhang, Y., Riduan, S. N., Wang, J. (2017). Chem. Eur. J., 23, 16419.
42. Liao, H., Wang, H., Ding, H., Meng, X., Xu, H., Wang, B., Ai, X., Wang, C.
(2016). J. Mater. Chem. A, 4, 7416–7421.
43. Yang, X., Dong, B., Zhang, H., Ge, R., Gao, Y., Zhang, H. (2015). RSC Adv.,
5, 86137.
44. Li, N., Zheng, M., Lu, H., Hu, Z., Shen, C., Chang, X., Ji, G., Cao, J., Shi, Y.
(2012). Chem. Commun., 48, 4106.
45. Zhao, Z., Wang, S., Liang, R., Li, Z., Shi, Z., Chen, G. (2014). J. Mater. Chem.
A, 2, 13509.
46. Gu, P.-Y., Zhao, Y., Xie, J., Binte Ali, N., Nie, L., Xu, Z. J., Zhang, Q. (2016).
ACS Appl. Mater. Interfaces, 8, 7464.
47. Ghazi, Z. A., Zhu, L., Wang, H., Naeem, A., Khattak, A. M., Liang, B., Khan,
N. A., Wei, Z., Li, L., Tang, Z. (2016). Adv. Energy Mater., 6, 1601250.
48. Liao, H., Ding, H., Li, B., Ai, X., Wang, C. (2014). J. Mater. Chem. A, 2,
8854.
49. Kandambeth, S., Shinde, D. B., Panda, M. K., Lukose, B., Heine, T.,
Banerjee, R. (2013). Angew. Chem. Int. Ed., 52, 13052.
50. Liao, H., Ding, H., Li, B., Ai, X., Wang, C. (2014). J. Mater. Chem. A, 2,
8854.
51. Xu, F., Jin, S., Zhong, H., Wu, D., Yang, X., Chen, X., Wei, H., Fu R., Jiang, D.
(2015). Sci. Rep., 5, 8225–8230.
52. Bai, L., Gao Q., Zhao, Y. (2016). J. Mater. Chem. A, 4, 14106–14110.
53. Kuhn, P., Antonietti M., Thomas, A. (2008). Angew. Chem. Int. Ed., 47,
3450–3453.
54. Ding, S., Gao, J., Wang, Q., Zhang, Y., Song, W., Su, C., Wang, W. (2011). J.
Am. Chem. Soc., 133, 19816–19822.
References 181

55. Uribe-Romo, F. J., Doonan, C. J., Furukawa, H., Oisaki K., Yaghi, O. M.
(2011). J. Am. Chem. Soc., 133, 11478–11481.
56. Kandambeth, S., Mallick, A., Lukose, B., Mane, M. V., Heine, T., Banerjee,
R. (2012). J. Am. Chem. Soc., 134, 19524–19527.
57. Jackson, K. T., Reich T. E., El-Kaderi, H. M. (2012). Chem. Commun., 48,
8823–8825.
58. Yang, H., Zhang, S., Han, L., Zhang, Z., Xue, Z., Gao, J., Li, Y., Huang, C., Yi,
Y., Liu H., Li, Y. (2016). ACS Appl. Mater. Interfaces, 8, 5366–5375.
59. Jlassi, K., Mekki, A., Zayani, M. B., Singh, A., Aswai D. K., Chehimi, M. M.
(2014). RSC Adv., 4, 65213–65222.
60. Fukuoka, T., Tonami, H., Maruichi, N., Uyama, H., Kobayashi, S.,
Higashimura, H. (2000). Macromolecules, 33, 9152–9155.
61. Zhan, H., Lamare, S., Ng, A., Kenny, T., Guernon, H., Chan, W. K., Djurisic,
A. B., Harvey, P. D., Wong, W. Y. (2011). Macromolecules, 44, 5155–5167.
62. Demir-Cakan, R., Morcrette, M., Nouar, F., Davoisne, C., Devic, T.,
Gonbeau, D., Dominko, R., Serre, C., F ́erey G., Tarascon, J. M. (2011). J.
Am. Chem. Soc., 133, 16154–16160.
63. Zheng, J., Tian, J., Wu, D., Gu, M., Xu, W., Wang, C., Gao, F., Engelhard, M.
H., Zhang, J., Liu J., Xiao, J. (2014). Nano Lett., 14, 2345–2352.
64. Zhou, J., Li, R., Fan, X., Chen, Y., Han, R., Li, W., Zheng, J., Wang, B., Li, X.
(2014). Energy Environ. Sci., 7, 2715–2724.
65. Shi, Y., Chou, S., Wang, J., Wexler, D., Li, H., Liu H., Wu, Y. (2012). J. Mater.
Chem., 22, 16465–16470.
66. Lee, J. W., Lim, S. Y., Jeong, H. M., Hwang, T. H., Kang, J. K., Choi, J. W.
(2012). Energy Environ. Sci., 5, 9889–9894.
67. Zhou, G., Wang, D., Li, F., Zhang, L., Li, N., Wu, Z., Wen, L., Lu, G., Cheng,
H. (2010). Chem. Mater., 22, 5306–5313.
68. Yoo, E., Kim, J., Hosono, E., Zhou, H., Kudo, T., Honma, I. (2008). Nano
Lett., 8, 2277–2282.
69. Sen, S., Moses, K., Bhattacharyya, A. J., Rao, C. N. R. (2014). Chem. Asian
J., 9, 100–103.
70. Hwang, H., Kim, H., Cho, J. (2011). Nano Lett., 11, 4826– 4830.
71. Li, N., Liu, G., Zhen, C., Li, F., Zhang, L., Cheng, H. (2011). Adv. Funct.
Mater., 21, 1717–1722.
72. Xu, X., Rout, C. S., Yang, J., Cao, R., Oh, P., Shin, H. S., Cho, J. (2013). J.
Mater. Chem. A, 1, 14548–14554.
73. Zhu, J., Yang, D., Yin, Z., Yan, Q., Zhang, H. (2014). Small, 10, 3480–3498.
Chapter 6

Biomedical Applications of COFs

Distinctive features, like large surface areas, tunable porosity, and p


conjugation of covalent organic frameworks (COFs), undergo unique
photoelectric properties and will enable COFs to serve as a promising
platform for drug delivery, bioimaging, biosensing, and theranostic
applications. In this chapter, the resent advanced research on COFs
is introduced in the biomedical and pharmaceutical sectors and the
challenges and opportunities regarding COFs are discussed in terms
of biomedical purposes. Although currently still in their infancy,
COFs as an innovative source have paved a new way to meet future
challenges in human healthcare and disease theranostic.

6.1 Introduction of Biomedical Application


Recent decades have witnessed a surge of exploration related
to biomedical applications of biosensing [1], bioimaging [2, 3],
chemotherapy [4, 5], gene therapy [6–9], immunotherapy [10, 11],
photodynamic therapy (PDT) [12], photochemical therapy [13, 14],
tissue engineering [15, 16], and others [17]. Multifarious materials
have been widely developed to achieve these objects, including
organic (liposomes [18–21], polymers [22–24], dendrimers [25, 26],
etc.), inorganic (metals [27], metallic oxides [28, 29], carbon [30,
31], mesoporous silica [32–34], etc.), and hybrid [35–38] materials.

Covalent Organic Frameworks


Atsushi Nagai
Copyright © 2020 Jenny Stanford Publishing Pte. Ltd.
ISBN 978-981-4800-87-7 (Hardcover), 978-1-003-00469-1 (eBook)
www.jennystanford.com
184 Biomedical Applications of COFs

To further improve their therapeutic efficacy, researchers endow


these biomaterials with stimuli responsiveness [39, 40] and targeted
delivery [41, 42]. Some agents have been approved by the Food and
Drug Administration (FDA) of the United States or in clinical trials,
such as Doxul. However, there are still more biomaterials limited
to clinical trials due to their unsatisfactory efficiency and safety
[43]. Organic materials usually suffer from low in vivo stability and
loading capacities [44], while inorganic substances often possess
undesirable toxicity and poor degradability [45]. Metal-organic
frameworks (MOFs), which are hybrid materials, have high surface
areas, large pore sizes, and good biodegradability, but their chemical
stability and toxicity are unsatisfactory since they are constructed by
metal coordination bonds [46]. These issues need to be addressed.
Exploitation of new materials is one of the possible methods.
Covalent organic frameworks (COFs) are crystalline porous
polymers with periodic skeletons and large surface areas [47].
Similar to MOFs, COFs have abundant regular pores with controllable
sizes and shapes and are easy to modify, but they are linked by
dynamic covalent bonds as molecular building blocks, which makes
them not only more stable than conventional organic materials
because of the molecular self-assembly process but also more
adaptive than inorganic particles formed by ionic or metallic bonds
[48]. In addition, COFs are often composed of light atoms, avoiding
the toxicity of metal ions. This kind of composition also features
COFs with low density. For example, the density of COF-18, a typical
boronate ester COF, is only 0.17 g/cm–3 [49].
Recently, COFs were developed for diagnoses and treatment [50,
51] since they exhibit many excellent properties, including stimuli
responsiveness, biodegradability, high surface areas, large pore
volumes, tunable pore structures, p-conjugated systems, unique
photoelectric properties, outstanding modifiability, and so on. The
next section introduces the applications of COFs in drug delivery,
photochemical therapy, PDT, biosensing, bioimaging, and others.

6.2 COF Properties of Biomedical Applications


COFs possess multiple prominent advantages that allow them to be
used in theranostic applications.
COF Properties of Biomedical Applications 185

∑ Dynamic covalent linkages for stimuli responsiveness


and biodegradation
Instead of traditional covalent bonds, coordinate bonds, and
physical interactions, COFs utilize dynamic covalent bonds
as linkages, which can preserve their structures in normal
conditions and they can be broken by stimulants such as acids
[52]. Their characters make COFs stable enough to reach the
target tissues and finish their task before they biodegrade.
Moreover, they can rapidly release cargos under stimulation,
especially for imine and its analogue linkage frameworks.
Polymer-covered boronate ester and spiroborate-linked
COFs can also be promising candidates for drug delivery,
as the reversibility and the dynamic nature of the boronate
ester offer not only pH-responsive and glucose-responsive
(dual responsiveness) features but also adaptive mechanical
behaviors, such as shear-thinning protonation of injection
[53–55]. Recently, covalent triazine frameworks (CTFs) were
found to possess pH responsiveness due to the protonation
and biodegradation of the triazine units [56]. Apparently, the
stimuli responsiveness and biodegradation make COFs ideal
carriers for therapy.
∑ High BET surface area and large pore volume for high
capacity
COFs possess high surface areas and large pore volumes
because of their unique framework skeletons. For instance,
the Brunauer–Emmett–Teller (BET) surface area of COF-102
is 3472 m2/g and of COF-103 is 4210 m2/g; the pore volume
of COF-103 is 1.66 m3/g and of COF-102 is 1.44 m3/g [57, 58].
These features are much better than those of most porous
materials, such as mesoporous silica and carbon [59, 60].
Furthermore, COFs are constructed by aromatic compounds
and can be used to load aromatic monomers through p–p
stacking. Thus, COFs possess a high loading capacity for drugs
and biological molecules.
∑ Tunable pore character for controlled release
As described in Chapter 1, the pore sizes and shapes of COFs
can be adjusted by changing the building blocks or after
186 Biomedical Applications of COFs

modification. Since pore sizes and shapes have an effect on


the diffusion of guest monomers, COFs with tunable pore
structures can be used to carry different pharmaceutical
agents and control their release rates.
∑ Wonderful photoelectronic properties for biosensing and
bioimaging
Mostly, COFs contain p-conjugated systems and laminated
structures, which endow them with excellent photoelectronic
properties. For example, polycyclic aromatic units (e.g.,
tetraphenylethenes and pyrenes) can make COFs emit
fluorescence and electron donors and acceptors units (e.g.,
naphthalimides, porphyrins, and phthalocyanines) can
endow COFs with proton conductivity. On the basis of these
properties, COFs show a great potential in biosensing and
bioimaging.
∑ Unique modifiability for versatile functionalities
Since COFs consist of building blocks with high tailorability,
extensive materials can be designed as building blocks and
introduced into their skeletons, enabling COFs to perform as
desired, such as producing singlet oxygen for PDT [61]. What’s
more, tailored attributes of COFs pave the way for the study of
structure-activity relationship.
∑ Outstanding modifiability for versatile functionalities
Modification is an effective way to enhance the applicability
of COFs. Inspired by other nanoparticles, surface modification
can improve their biocompatibility and targeting ability
[62, 63]. Since MOFs can regulate drug release profiles
by incorporating different substituents on the pore walls
[64], COFs may achieve similar results. Even biological
probes and drugs can be linked to COFs by a postsynthesis
method. With proper ligands, inorganic pharmaceutical
agents (e.g., cisplatin) may be connected to COFs. There
are also multifarious possible functionalities waiting for
further exploration. These properties open the door for the
application of COFs in biomedicine.
Biomedical COF Applications 187

6.3 Biomedical COF Applications


Due to unique features that endow them with great potential in
diagnosis and treatment, COFs have been explored in drug delivery,
photochemical therapy, PDT, biosensing, bioimaging, and other
biomedical applications. Even though in their early stages, COFs
have attracted a lot of researcher attention and a surge of studies
have been reported in the past several years.

6.3.1 Drug Delivery


The first example of application of COFs in drug delivery was
reported by Yan and coworkers [65], who designed two new 3D
polyimide (PI) COFs with different pore sizes, PI-COF-4 (13 Å) and
PI-COF-5 (10 Å). Ibuprofen (IBU; molecular size 5Å × 10 Å) was
selected as a model drug as it can be entrapped by the pores of
PI COFs. Thermogravimetric analysis (TGA) proved that the drug
loading efficacy (DLE) was as much as 24% (PI-COF-4) and 20% (PI-
COF-2). What’s more, the drug release was sustained for more than
6 days. Besides IBU, PI COFs are also suitable to deliver captopril and
caffeine, achieving similar results to that of IBU.
Porphyrin-based covalent triazine frameworks (PCTFs) were
also used in drug delivery [66]. The authors believed that the acid
group decorated in IBU could interact with the triazine rings of COFs,
which would have a positive effect on drug loading and controlled
release. TGA illustrated that the DLE was about 19% for metal-free
PCTFs and 23% for Mn-coordinated PCTFs. For both PTCFs, the total
release amount was beyond 90% within 48 h. These results are no
better than those of the above-mentioned PI COFs, which may be
because PCTFs have irregular morphologies and wider pore size
distribution because of the poor crystallinity of triazine-based COFs
[67].
Although a number of studies were conducted to investigate in
vitro drug release from various COFs, few works are focused on the
in vivo biocompatibility and cytotoxicity. Zhao et al. carried out cell
experiments with two imine-linked 2D COFs, PI-2 COFs and PI-3
COFs [68]. Both COFs exhibited high drug loading capacities for
5-FU, captopril, and IBU, even reaching 30 wt%. The release rates
188 Biomedical Applications of COFs

of 5-FU were similar for both PI-n COFs, and most of the drug was
released within 3 days (Fig. 6.1c).

Figure 6.1 (a) Confocal images of cells before and after the treatment of
drug-loaded COFs, (b) quantitative MIT analysis showing the cell viability when
incubated with bare COFs, 5-FU, and 5-FU-loaded COFs for 24 h. (c) Release
profiles of two 5-FU-loaded COFs at 100 mg/mL and corresponding fitting
curves. Republished with permission of Royal Society of Chemistry, from Ref.
[68], copyright (2016); permission conveyed through Copyright Clearance
Center, Inc.

Importantly, these COFs were uniform spherical nanoparticles


with a diameter of 50 nm when dispersed in a phosphate-buffered
saline (PBS) solution with a bit of dimethyl sulfoxide as an additive.
This feature made COFs carriers of a suitable size for cell uptake
and in vivo drug delivery. Confocal images proved the effective
uptake of drug-loaded nanoparticles by the MCF-7 cells (Fig. 6.1a).
Quantitative 3-(4,5-dimethylthiazol-2-yl)-2.5-diphenyltetrazolium
bromide (MIT) analysis indicated that both COFs showed good
biocompatibility. However, the viability of the cells was decreased to
10% when they were treated with 5-FU-loaded hybrids for 48 h, as
shown in Fig. 6.1b. Lotsch and coworkers demonstrated a new imine-
based TTI COF to deliver quercetin (Fig. 6.2a,b) [69]. The TTI COF
could retain its morphology in water and weak acids for more than
Biomedical COF Applications 189

two days. Molecular dynamic simulations proved that the quercetin


was bound to the pore wall by C–H–p and H bond interactions (Fig.
6.2c,d). For this reason, the TTI COF seemed to be able to carry more
drugs. However, trifluoroacetic acid (TFA) analysis failed to calculate
the accurate drug loading rate since the quercetin decomposed
slowly over a wide temperature range. What’s worse, an in vitro
drug-release study was also not successful due to the fact that the
quercetin was hard to dissolve in PBS and oxidized easily [70]. Cell
experiments indicated that the drug-loaded COF successfully killed
most of the human breast carcinoma MDA-MB-231 cells while the
bare COFs showed no cytotoxicity. Regardless of some inadequacies,
this work carefully discussed every detail and shared with us
valuable experiences to design this new carrier.

Figure 6.2 (a) Synthesis of a TTI COF from TT-ald and TT-am, (b) structure of
quercetin, (c) side view of the modeled COF pore showing interaction energies
between the quercetin and the model hexagon layers, and (d) top view of the
modeled COF pore showing the interaction of quercetin with the pore wall.
Reproduced with permission from Ref. [69]. Copyright (2016), John Wiley and
Sons.

Stimuli responsiveness is one of most fascinating properties to


achieve real-time monitoring and on-demand drug release [71, 72].
190 Biomedical Applications of COFs

Besides the inherent acid sensitivity of boronate- and imine-based


COFs, researchers have endowed COFs with more environment-
responsive features to control the drug delivery and release. Huh
et al. developed pH-responsive CTF nanoparticles (NCTPs) via a
Friedel–Crafts reaction [73]. Doxorubicin (DOX) was loaded in
NCTP as a model drug, and the hydrophobic interactions between
DOX and NCTP were weakened. Thus, the release rate was faster
than that under a PBS solution of pH 7.4. More interestingly, COFs
as drug carriers can provide more flexibility and additional option
for access to controllable drug release because of their unique
photoelectronic properties. For example, Lei and coworkers applied
self-condensation of 4,4¢-phenylazobenzyl diboronic acid (ABBA) to
build single-layered photoresponsive COFs on the surface of highly
oriented pyrolytic graphite [74]. Under ultraviolet (UV) irradiation,
the frameworks would be destroyed and release the guest, copper
phthalocyanine (CuPc). Furthermore, the destroyed COFs could be
recovered through annealing. This finding offers a possible on/off
switch for drug release.
In 2015, Banerjee and coworkers reported a series of explorations
of imine COFs for drug loading and release [75–77]. They prepared
hollow spherical COFs via a template-free method [76]. The drug
loading capacity was only 0.35 mg/g for DOX, calculated by UV-visual
absorbance spectra. The release rate was very slow, and more than
50% of the drug remained after 7 days in a pH 5 phosphate buffer.
After that, they designed self-standing porous COF membranes
to sieve larger molecules (more than 1 nm) such as rose Bengal,
tetracycline, and curcumin [77]. Since the enhanced permeability
and retention (EPR) effect is not effective as expected and only about
5% of the administered nanoparticles can reach the target tumor,
modifying with target groups is a possible way to improve the use
of nanocarriers [42]. Banerjee and coworkers took the first step
regarding COFs for targeted drug delivery (Fig. 6.3). They modified
covalent organic nanosheets with folic acid (naming it “TpASH-
FA”) through the postsynthetic modification method, and 5-FU was
chosen as the model drug [75]. However, the loading efficiency of
these nanosheets (only 12%) was lower than that of many other
COFs [78–81]. They exhibited pretty good anticancer activity,
and only 14% of MDA-MB-231 cells survived under a dosage of
50 mg/mL.
Biomedical COF Applications 191

Figure 6.3 (a) Schematic representation of targeted drug delivery by CONs


(sheet-like material denotes CONs here), (b) drug loading study of 5-FU by UV-
vis spectra, (c) MTT assay on MDA-MB-231 cell lines showing cellular viability,
and (d) comparison of cellular migration study between control and TPASH-FA-
5-FU-treated sets. Reprinted with permission from Ref. [75]. Copyright (2017)
American Chemical Society.

Apart from carrying anticancer drugs, functional COFs can kill


tumor cells by themselves. Bhaumik et al. reported an interesting
work in which phloroglucinol-contained COFs were utilized as
anticancer agents [82]. They applied 2,4,6-triformylphloroglucinol
and 4,4¢-ethylenediamine to construct nanofiber-like COFs
(EDTFP-1) with a diameter of 22–30 nm, as shown in Fig. 6.4a.
The phloroglucinol derivative that was created could accelerate
reactive oxygen species generation and caused the apoptosis of
cancer cells [83]. This cell experiment indicated that EDTFP-1
successfully induced mitochondrial-dependent apoptosis associated
with DNA fragmentation, mitochondrial membrane potential loss,
phosphatidylserine externalization, and pro- and antiapoptotic
protein imbalance, as shown in Fig. 6.4b. Thus, EDTFP-1 showed
obvious cytotoxicity against cancer cells, like HCT 116, HepG2, A549,
and MIA-Paca2. This finding will open a door for COFs to work as
anticancer agents.
192 Biomedical Applications of COFs

Figure 6.4 Schematic representations of the preparation of EDTFP-1 (A) and


its induced apoptotic pathway (B). Reprinted with permission from Ref. [82].
Copyright (2017) American Chemical Society.

The applications of drug delivery using COFs are summarized


in Table 6.1. As per the table, most drug-loading COFs are linked
by imine (C–N) bonds, which are believed to be stable in water.
Regardless, for COFABBA built by boroxine bonds, neither loading
efficacy nor release rate was studied. The pore sizes are mostly
designed to fit the model drug, but these carriers have varied
instead of uniform nanoparticles, which needs further exploration.
In general, the DLE of COFs was rather high and the drugs and the
COFs showed good biocompatibility. Several limitations markedly
hamper their clinical translation. However, COFs have shown their
pharmaceutical potential. The intricate structural characteristics
demand careful engineering of the COFs in order to realize the
desirable effect. The complexity of the COF preparations is a key
issue that needs to be addressed to scale up the production or to
ensure batch-to-batch reproducibility. Challenges still exist in
terms of delivery of the cargo to the targeted site as well as efficient
clearance of the COFs once they have finished their mission in vivo.
Although several studies have tested the efficacy of COF formulation
and their safety, few studies have been carried out on their long-
term accumulation and degradation profiles. However, I believe that
continuous improvement in the design and detailed investigations
on their in vivo behavior will help to tale COFs as nanocarriers from
bench to besides.
Table 6.1 Examples of COFs as drug carriers

Pore sizes
Year COFs Linkages (nm) Morphologies Model drug Characters Ref.

2015 PI-COF-4 Imide 1.3 Rectangular; IBU, caffeine, DLEa: 24 wt %; DRR: 95% for 6 [65]
PI-COF-5 1.0 length hundreds and captopril days. DLEa: 20 wt %; DRR: 95%

2015
of nanometers for 6 days.

COF-DhaTab Imine 3.7 Submicron DOX DLEb: 0.35 mg/g; DRR: 42% [76]

2016
hollow spheres after 7 days at pH 5.

PI-2 COF Imine 1.4 Spherical 5-FU, IBU, DLEa: 30 wt %; DRR: 85% for 5 [82]
PI-3 COF 1.1 nanoparticles; and captopril days; good biocompatibility.

2016
50 nm

COFABBA Boroxine 2.1 Single layer CuPc Photoresponsive release; no [74]


DLE or DRR measured.

2016 TTI COF Imine 2.4 Elongated rods Quercetin DLEb: Failed to measure; [69]
DRR: Failed to measure; good

(Continued)
biocompatibility.
Biomedical COF Applications
193
194

Table 6.1 (Continued)

Pore sizes
Year COFs Linkages (nm) Morphologies Model drug Characters Ref.

2016 NCTP Triazine 1.21 Spherical DOX DLEa: 20 wt%; DRR: 60% for [79]
nanoparticles; 2 days at pH 7.4 and 80%
Biomedical Applications of COFs

50–70 nm for 2 days at pH 4.8; good

2017 b-ketoenamine
biocompatibility.

EDTFP-1 1.5 Nanofibers; TFP TFP works as a model drug as [82]


diameter 22–30 well as a building block; cancer

2017 b-ketoenamine
nm cells killed by COFs themselves.

TpASH-FA 1.3 Nanosheets 5-FU DLEb: 12 wt%; DRR: 50% for 75 [75]
h at pH 7.4 and 75% for 75 h at

2017
pH 5; good biocompatibility.

PCTF Triazine 0.8–2.7 Irregular IBU DLEa: 19 wt%; DRR: 90% for [80]
PCTF-Mn 0.74.2 nanoparticles; 48 h. DLEa: 23 wt%; DRR: 95%
plate shaped for 48 h.
aMeasured by TFA.
bMeasured by UV.
DLE; drug loading efficacy; DDR: drug release rate; IBU: ibuprofen; DOX: doxorubicin; 5-FU: 5-fluorouracil; CuPc: copper phthalocyanine; TFP:
2,4,6-triformylphloroglucinol.
Biomedical COF Applications 195

6.3.2 Photothermal and Photodynamic Therapy


Besides chemotherapy, COFs can also be utilized for photothermal
therapy (PTT) [84] and PDT [85, 86].
Prompted by conjugated microporous polymers [87], Guo et
al. attempted to cover imine-linked COFs on the surface of Fe3O4
nanoclusters for PTT, as shown in Fig. 6.5a [84]. The resulting Fe3O4@
COF microspheres increased the system temperature by 25°C (Fig.
6.5b), and the photochemical conversion efficiency (21.5%) was
comparable to that of some widely studied photosensitizers, such
as Au nanorods [88, 89]. Moreover, the Fe3O4 core featured the
microspheres with magnetic target characteristic.

Figure 6.5 Schematic representations of the preparation (a) and


photochemical effect (b) of Fe3O4@COF. Reproduced with permission from Ref.
[84]. Copyright (2016), John Wiley and Sons.

Porphyrin and its derivatives can generate singlet oxygen (1O2)


under photoirradiation and are widely studied as photosensitizers
for PDT [90]. Using porphyrin derivatives as building blocks and 3D
structures, the construction of photosensitive COFs is a promising
method to avoid the aggregation (such as H- or J-aggregates)
of porphyrin macrocycles in aqueous media and improve their
efficiency in vivo. Wang et al. designed 3D porphyrin-based COFs
with excellent photosensitivity, resulting in 3D effects enabling the
suppression of intermolecular and intramolecular p–p stacking
interaction. Authors could be used to provide reactive singlet oxygen
under photoirradiation [86]. The photosensitive properties of COFs
could be tuned by the incorporation of metal ions into the center of the
porphyrin macrocycle. Xie and coworkers grew imine-linked COFs
196 Biomedical Applications of COFs

on the surface of an amine-decorated MOF to prepare porphyrin-


based COFs with uniform nanoscale structures, as shown in
Fig. 6.6a–c [91]. The amine-modified MOF was a regular octahedron
around 165 nm. The hybrid particles (referred to as UNM) were
nearly spherical and had an average size of 176 nm, determined by
dynamic light scattering. Confocal laser scanning microscopy images
showed that the UNM could be endocytosed by HeLa cells. As shown
in Fig. 6.6b, with 2¢,7¢-dichlrodihydrofluorescein diacetate DCFH-DA
as indicators, bright-green fluorescence could be observed under
light irradiation, which was attributed to the production of singlet
oxygen. MTT assay proved that UNM had significant cytotoxicity
for both HepG2 and HeLa cells under irradiation but nearly no
cytotoxicity can be found without light.

Figure 6.6 (a) Synthesis and preparation of UNM nanoparticles, (b) structure
of POP on UNM nanoparticles, (c) the cellular uptake and photodynamic
therapy in cells, and (d) the photodynamic effect of MOF@COF. Reprinted with
permission from Ref. [91]. Copyright (2017) American Chemical Society.

6.4 Biosensing and Bioimaging


The p-conjugated system and photoelectrical properties of COFs
have been widely studied in catalysis [92, 93], proton conduction
[94, 95], and energy storage [96, 97]. Lately, these features have
been used for biosensing and bioimaging [98–100].
One of the biosensors was designed on the basis of the
electrochemical activity of COF films. Fang et al. prepared imine-
linked COF films on amino-functionalized silicon wafers, and the
Biosensing and Bioimaging 197

presence of amino groups can endow the hybrid films with a positive
charge [98]. Negatively charged biomolecules, for example, BSA
and probe DNA, would be adsorbed on the surface by electrostatic
interactions to strengthen the electrochemical activity of the
COF films. Electrochemical impedance spectroscopy was used to
detect this change. Thus, the biological signals were converted into
electric signals. The other biosensing application is based on the
p-conjugated system of COFs [99, 100]. Utilizing the p–p stacking
effect, a fluorescent dye was quenched via fluorescence resonance
energy transfer [100]. When the target DNA combined with probes,
the interaction between the DNA probes and COF films weakened
and the fluorescence was recovered. Similarly, a carboxyfluorescein-
labeled probe (FAM probe) was adsorbed on the b-ketoenamine-
linked COFs by hydrogen bond and p–p stacking interactions [99].
As shown in Fig. 6.7, when the target biomolecules interacted
with the FAM probes, the fluorescence was enhanced. This new
platform showed highly sensitive and selective DNA and adenosine
5¢-triphosphate.
Besides biomolecules, COFs can also be applied to the detection
of monomers and metal ions. Jiang and coworkers designed azine-
linked pyrene-based COFs in which the azine units extended p
delocalization over the 2D skeleton, enabling the framework to
emit a yellow light [101]. Furthermore, azine units have lone
pairs of electrons and provide docking sites for hydrogen-bonding
interaction with guest molecules. Thus, these frameworks can be
utilized to detect 2,4,6-trinitrophenol with high sensitivity and
selectivity since phenol units can form hydrogen bonds with azine
units and nitro groups can quench the fluorescence. Similarly,
another 3D pyrene-based COF was used for chemosensing [102].
Furthermore, Yang and coworkers prepared two kinds of polyimide-
based COFs that emitted strong fluorescence in solution, and the
fluorescence would be quenched by Fe3+ [103]. This property can be
used for the selective luminescence of Fe3+.
Triazine-based frameworks of NTCTPs showed a strong
emission property and can be used for bioimaging due to extended p
conjugation in the frameworks [104]. What’s more, the aggregation-
induced emission (AIE) effect and blue luminescence of conjugated
COFs can hopefully be developed for bioimaging [105–107].
198 Biomedical Applications of COFs

Figure 6.7  (a) Time courses for the fluorescence quenching of the probe (50 


nM) upon TpTta addition (10 mg) (black curve) and for the fluorescence recovery 
of the probe/TpTta upon target DNA addition (400 nM DNA) (red curve). (b)
Dependence  of  the  fluorescence  of  the  probe  (50  nM)  on  the  concentration 
of the target DNA (0, 10, 20, 40, 60, 80, and 100 nM). (c) Plot of the recovered
fluorescence intensity (DF) against the target DNA concentration. (d) Selectivity 
of this DNA assay. The fluorescence intensity was measured at 528 nm under 
excitation at 493 nm. The error bars represent the standard deviation of three
replicate measurements. Reproduced of Royal Society of Chemistry, with
permission from Ref. [99], copyright (2014); permission conveyed through
Copyright Clearance Center, Inc.

6.5 Other Biomedical Applications


With their selective adsorption ability, COFs can be applied in protein
immobilization, biomolecular adsorption, and hazardous substance
removal. Immobilization of enzymes plays an important role in
biomedical industries [108]. Banerjee and coworkers designed
hollow spherical COFs (COF-DhaTab) for trypsin immobilization
[109]. Although trypsin is a globular protein with a diameter of 3.8
nm, which is a bit larger than the pore size of COF-DhaTab, it could
be immobilized due to its soft character. Figure 6.8 shows various
methods that were used to demonstrate the immobilization of
trypsin by COFs and the loading capacity was as high as 15.5 mmol/g.
Cai et al. designed core-shell-structured Fe3O4@COFs by way of room
temperature synthesis [110]. On the basis of the hydrophobic effect,
Other Biomedical Applications 199

the composite nanoparticles adsorbed hydrophobic peptides and


repelled hydrophilic peptides. With the assistance of the magnetic
responsiveness of Fe3O4 cores, hydrophobic peptides were separated
from the mixture system. Further, these particles were used for the
selective enrichment of peptides and exclusion of proteins. This
feature has great potential in real-world applications, such as in the
analysis of human serum. When amine groups were incorporated
onto the pore walls, COFs were used for the adsorption of lactic acid
[111]. Other works show that when modified with sulfur derivatives,
COFs were used to remove Hg2+, Hg0, Pd2+, and Cu2+ [112, 113].

Figure 6.8 (a) Schematic representation of trypsin immobilization, (b) amount


of trypsin loading in COF-DhaTab; evidence of trypsin immobilization through
(c) PXRD patterns, (d) N2 adsorption isotherms, (e, f) TEM images, and (g)
confocal Z stacks. Reprinted by permission from Springer Nature Customer
Service Centre GmbH: Springer Nature, Nature Communications, Ref. [109],
copyright (2015).

COFs can be used for antimicrobial applications.


1,3,5-triformylphloroglucinol and guanidinium halide were used
to build self-exfoliated ionic covalent organic nanosheets (iCONs)
[114]. Since guanidinium units can form hydrogen bonds with the
oxoanions of phosphate, positively charged iCONs can break the
negatively charged phospholipid bilayers of bacteria, as shown in
Fig. 6.9. Antibacterial studies testified that these nanosheets show
excellent antimicrobial activity against both gram-positive and gram-
negative bacteria. Furthermore, they mixed iCONs and polysulfone
200 Biomedical Applications of COFs

(PSF) to fabricate an iCONs@PSF membrane as an antimicrobial


coating. Moreover, COFs are expected to deliver bioactive gases, like
nitric oxide (NO), as MOFs do [115].

Figure 6.9 (a) SEM images and (b) TEM images of control and TpTGCl-treated
S. aureus. (c) SEM images and (d) TEM images of control and TpTGCl-treated
E. coli. (e) Schematic representation for the mode of action between bacteria
and iCONs. (f) Digital image of a TpTGCl@PSF mixed matrix membrane. (g) SEM
image of the TpTGCl@PSF mixed matrix membrane. Antibacterial property of
TpTGCl@PSF mixed matrix membrane by growth of (h) S. aureus and (i) E. coli
on it. Reprinted with permission from Ref. [114]. Copyright (2016) American
Chemical Society.

6.6 Conclusions of Biomedical Applications


COFs, a representative material of porous organic polymers, have
been widely explored in various fields and play an important role in
nanotechnology. In this chapter, some biomedical examples of COFs
were introduced to pave the way for their porous frameworks to be
used for biomedical purposes. Most importantly, the complementary
functional design of skeletons and high flexibility in the control of
pore geometry may provide a powerful means of exploring COFs for
challenging biomedicine issues. Specifically because of the unique
photoelectric properties, COFs are expected to provide an ideal
theranostic platform since it is easy to integrate the drug therapy
and photoelectric diagnostic function in one COF-based system.
References 201

So, it is no wonder that though in its infancy, COFs’ application in


drug delivery, PTT, PDT, biosensing, bioimaging, and other possible
diagnosis has significantly increased and received increasing
attention with the development of new COF materials. However,
despite their many favorable characteristics and great progress,
most of the currently available COFs face many challenges in terms of
preparation complexity, regulated morphology, hydrolytic stability,
targeted delivery, controlled release, and long-term biocompatibility,
which indeed impairs their practical application in biomedical and
pharmaceutical fields. As a result, considerable effort is required
to design and fabricate COFs with well-defined morphologies and
desirable properties. Especially, future work may need to focus on
the following in the biomedical field:
∑ Scaling up the preparation of COFs with a controlled and
uniform morphology
∑ Ensuring hydrolytic and hemodynamic stability
∑ Ensuring biocompatibility and addressing systemic toxicity
Although there is still a long way to go before COFs can be
applied in clinical settings, the rapid development of these materials
will help them become a potential platform for biomedical and
pharmaceutical purposes, opening up some exciting new avenues to
improve human welfare.

References
1. Kim, S.-J., Choi, S.-J., Jang, J.-S., Cho, H.-J., Kim, I.-D. (2017). Acc. Chem.
Res., 50, 1587–1596.
2. Wolfbeis, O. S. (2015). Chem. Soc. Rev., 44, 4743–4768.
3. Smith, B. R., Gambhir, S. S. (2017). Chem. Rev., 117, 901–986.
4. Huang, P., Liu, J., Wang, W., Zhang, Y., Zhao, F., Kong, D., Liu, J., Dong, A.
(2016). Acta Biomater., 40, 263–272.
5. Huang, P., Wang, W., Zhou, J., Zhao, F., Zhang, Y., Liu, J., Liu, J., Dong, A.,
Kong, D., Zhang, J. (2015). ACS Appl. Mater. Interfaces, 7, 6340–6350.
6. Riley, M. K., Vermerris, W. (2017). Nanomaterials, 7, 94.
7. Zhou, J., Wu, Y., Wang, C., Cheng, Q., Han, S., Wang, X., Zhang, J., Deng, L.,
Zhao, D., Du, L. (2016). Nano Lett., 16, 6916–6923.
8. Wang, C., Du, L., Zhou, J., Meng, L., Cheng, Q., Wang, C., Wang, X., Zhao,
202 Biomedical Applications of COFs

D., Huang, Y., Zheng, S. (2017). ACS Appl. Mater. Interfaces, 9, 32463–
32474.
9. Tatiparti, K., Sau, S., Kashaw, S. K., Iyer, A. K. (2017). Nanomaterials, 7,
77.
10. Li, P., Zhou, J., Huang, P., Zhang, C., Wang, W., Li, C., Kong, D. (2017).
Regen. Biomater., 4, 11–20.
11. Li, P., Song, H., Zhang, H., Yang, P., Zhang, C., Huang, P., Kong, D., Wang,
W. (2017). Nanoscale, 9, 13413–13418.
12. Croce, R., Van Amerongen, H. (2014). Nat. Chem. Biol., 10, 492–501.
13. Cherukula, K., Manickavasagam Lekshmi, K., Uthaman, S., Cho, K., Cho,
C. S., Park, I. K. (2016). Nanomaterials, 6, 76.
14. Pattani, V. P., Tunnell, J. W. (2012). Lasers Surg. Med., 44, 675–684.
15. Amezcua, R., Shirolkar, A., Fraze, C., Stout, D. A. (2016). Nanomaterials,
6, 133.
16. Chieruzzi, M., Pagano, S., Moretti, S., Pinna, R., Milia, E., Torre, L., Eramo,
S. (2017). Nanomaterials, 6, 134.
17. Song, G., Cheng, L., Chao, Y., Yang, K., Liu, Z. (2017). Adv. Mater., 29,
1700996.
18. Deng, H., Song, K., Zhao, X., Li, Y., Wang, F., Zhang, J., Dong, A., Qin, Z.
(2017). ACS Appl. Mater. Interfaces, 9, 9315–9326.
19. Lombardo, D., Calandra, P., Barreca, D., Magazu, S., Kiselev, M. A. (2016).
Nanomaterials, 6, 125.
20. Campani, V., Salzano, G., Lusa, S., De Rosa, G. (2016). Nanomaterials, 6,
131.
21. Al-Jamal, W. T., Kostarelos, K. (2011). Acc. Chem. Res., 44, 1094–1104.
22. Wang, Y., Li, P., Truong-Dinh Tran, T., Zhang, J., Kong, L. (2016).
Nanomaterials, 6, 26.
23. Amirmahani, N., Mahmoodi, N. O., Mohammadi Galangash, M.,
Ghavidast, A. (2017). J. Ind. Eng. Chem., 55, 21–34.
24. Gong, J., Chen, M., Zheng, Y., Wang, S., Wang, Y. (2012). J. Control. Release,
159, 312–323.
25. Kesharwani, P., Lyer, A. K. (2015). Drug Discov. Today, 20, 536–547.
26. Sikwal, D. R., Kalhapure, R. S., Govender, T. (2017). Eur. J. Pharm. Sci.,
97, 113–134.
27. Boisselier, E., Astruc, D. (2009). Chem. Soc. Rev., 38, 1759–1782.
28. Pankhurst, Q., Thanh, N., Jones, S., Dobson, J. (2009). J. Phys. D: Appl.
Phys., 42, 224001.
References 203

29. Ulbrich, K., Hola, K., Subr, V., Bakandritsos, A., Tucek, J., Zboril, R.
(2016). Chem. Rev., 116, 5338–5431.
30. Mendes, R. G., Bachmatiuk, A., Büchner, B., Cuniberti, G., Rümmeli, M. H.
(2013). J. Mater. Chem. B, 1, 401–428.
31. Hong, G., Diao, S., Antaris, A. L., Dai, H. (2015). Chem. Rev., 115, 10816–
10906.
32. Martinez-Carmona, M., Colilla, M., Vallet-Regi, M. (2015). Nanomaterials,
5, 1906–1937.
33. He, Q., Shi, J., Chen, F., Zhu, M., Zhang, L. (2010). Biomaterials, 31, 3335–
3346.
34. Tang, F., Li, L., Chen, D. (2012). Adv. Mater., 24, 1504–1534.
35. Wu, M.-X., Yang, Y.-W. (2017). Adv. Mater., 29, 1606134.
36. Lismont, M., Dreesen, L., Wuttke, S. (2017). Adv. Funct. Mater., 27,
1606314.
37. He, C., Liu, D., Lin, W. (2015). Chem. Rev., 115, 11079–11108.
38. Zhao, F., Yao, D., Guo, R., Deng, L., Dong, A., Zhang, J. (2015).
Nanomaterials, 5, 2054–2130.
39. Liu, X., Yang, Y., Urban, M. W. (2017). Macromol. Rapid Commun., 38,
1700030.
40. Li, F., Lu, J., Kong, X., Hyeon, T., Ling, D. (2017). Adv. Mater., 29, 1605897.
41. Bar-Zeev, M., Livney, Y. D., Assaraf, Y. G. (2017). Drug Resist. Updates,
31, 15–30.
42. Toy, R., Bauer, L., Hoimes, C., Ghaghada, K. B., Karathanasis, E. (2014).
Adv. Drug Deliv. Rev., 76, 79–97.
43. Bobo, D., Robinson, K. J., Islam, J., Thurecht, K. J., Corrie, S. R. (2016).
Pharm. Res., 33, 2373–2387.
44. Blanco, E., Shen, H., Ferrari, M. (2015). Nat. Biotechnol., 33, 941–951.
45. Deng, C., Jiang, Y., Cheng, R., Meng, F., Zhong, Z. (2012). Nano Today, 7,
467–480.
46. Yang, P., Gai, S., Lin, J. (2012). Chem. Soc. Rev., 41, 3679–3698.
47. Huxford, R. C., Della Rocca, J., Lin, W. (2010). Curr. Opin. Chem. Biol., 14,
262–268.
48. Huang, N., Wang, P., Jiang, D. (2016). Nat. Rev. Mater., 1, 16068.
49. Yaghi, O. M. (2016). J. Am. Chem. Soc., 138, 15507–15509.
50. El-Kaderi, H. M., Hunt, J. R., Mendoza-Cortés, J. L., Côté, A. P., Taylor, R.
E., O’Keeffe, M., Yaghi, O. M. (2007). Science, 316, 268–272.
204 Biomedical Applications of COFs

51. Fang, Q., Wang, J., Gu, S., Kaspar, R. B., Zhuang, Z., Zheng, J., Guo, H., Qiu,
S., Yan, Y. (2015). J. Am. Chem. Soc., 137, 8352–8355.
52. Mitra, S., Sasmal, H. S., Kundu, T., Kandambeth, S., Illath, K., Diaz Diaz,
D., Banerjee, R. (2017). J. Am. Chem. Soc., 139, 4513–4520.
53. Mura, S., Nicolas, J., Couvreur, P. (2013). Nat. Mater., 12, 991–1003.
54. Chen, J., Su, Q., Guo, R., Zhang, J., Dong, A., Lin, C., Zhang, J. (2017).
Macromol. Chem. Phys., 218, 1700166.
55. Guo, R., Su, Q., Zhang, J., Dong, A., Lin, C., Zhang, J. (2017).
Biomacromolecules, 18, 1356–1364.
56. Zhao, F., Wu, D., Yao, D., Guo, R., Wang, W., Dong, A., Kong, D., Zhang, J.
(2017). Acta Biomater., 64, 334–345.
57. El-Kaderi, H. M., Hunt, J. R., Mendoza-Cortés, J. L., Côté, A. P., Taylor, R.
E., O’Keeffe, M., Yaghi, O. M. (2007). Science, 316, 268–272.
58. Côté, A. P., El Kaderi, H. M., Furukawa, H., Hunt, J. R., Yaghi, O. M. (2007).
J. Am. Chem. Soc., 129, 12914–12915.
59. Hong, G., Diao, S., Antaris, A. L., Dai, H. (2015). Chem. Rev., 115, 10816–
10906.
60. Martinez-Carmona, M., Colilla, M., Vallet-Regi, M. (2015). Nanomaterials,
5, 1906–1937.
61. Lin, G., Ding, H., Chen, R., Peng, Z., Wang, B.: Wang, C. (2017). J. Am.
Chem. Soc., 139, 8705–8709.
62. Bar-Zeev, M., Livney, Y. D., Assaraf, Y. G. (2017). Drug Resist. Updates,
31, 15–30.
63. Greenwald, R. B., Choe, Y. H., McGuire, J., Conover, C. D. (2003). Adv.
Drug Deliv. Rev., 55, 217–250.
64. Dong, Z., Sun, Y., Chu, J., Zhang, X., Deng, H. (2017). J. Am. Chem. Soc.,
139, 14209–14216.
65. Fang, Q., Wang, J., Gu, S., Kaspar, R. B., Zhuang, Z., Zheng, J., Guo, H., Qiu,
S., Yan, Y. (2015). J. Am. Chem. Soc., 137, 8352–8355.
66. Luo, Y., Liu, J., Liu, Y., Lyu, Y. (2017). J. Polym. Sci. Part A: Polym. Chem.,
55, 2594–2600.
67. Kuhn, P., Forget, A., Su, D., Thomas, A. (2008). J. Am. Chem. Soc., 130,
13333–13337.
68. Bai, L., Phua, S. Z., Lim, W. Q., Jana, A., Luo, Z., Tham, H. P., Zhao, L., Gao,
Q., Zhao, Y. (2016). Chem. Commun., 52, 4128–4131.
69. Vyas, V. S., Vishwakarma, M., Moudrakovski, I., Haase, F., Savasci, G.,
Ochsenfeld, C., Spatz, J. P., Lotsch, B. V. (2016). Adv. Mater., 28, 8749–
8754.
References 205

70. Wybranowski, T., Kruszewski, S. (2014). Acta Phys. Pol. A, 125, A57–
A60.
71. Mura, S., Nicolas, J., Couvreur, P. (2013). Nat. Mater., 12, 991–1003.
72. Lu, Y., Aimetti, A. A., Langer, R., Gu, Z. (2016). Nat. Rev. Mater., 2, 16075.
73. Rengaraj, A., Puthiaraj, P., Haldorai, Y., Heo, N. S., Hwang, S. K., Han, Y.
K., Kwon, S., Ahn, W. S., Huh, Y. S. (2016). ACS Appl. Mater. Interfaces, 8,
8947–8955.
74. Liu, C., Zhang, W., Zeng, Q., Lei, S. (2016). Chemistry (Easton), 22, 6768–
6773.
75. Mitra, S., Sasmal, H. S., Kundu, T., Kandambeth, S., Illath, K., Diaz Diaz,
D., Banerjee, R. (2017). J. Am. Chem. Soc., 139, 4513–4520.
76. Kandambeth, S., Venkatesh, V., Shinde, D. B., Kumari, S., Halder, A.,
Verma, S., Banerjee, R. (2015). Nat. Commun., 6, 6786.
77. Kandambeth, S., Biswal, B. P., Chaudhari, H. D., Rout, K. C., Kunjattu, H.
S., Mitra, S., Karak, S., Das, A., Mukherjee, R., Kharul, U. K., Banerjee, R.
(2017). Adv. Mater., 29, 1603945.
78. Mitra, S., Sasmal, H. S., Kundu, T., Kandambeth, S., Illath, K., Diaz Diaz,
D., Banerjee, R. (2017). J. Am. Chem. Soc., 139, 4513–4520.
79. Rengaraj, A., P uthiaraj, P., Haldorai, Y., Heo, N. S., Hwang, S. K., Han, Y.
K., Kwon, S., Ahn, W. S., Huh, Y. S. (2016). ACS Appl. Mater. Interfaces, 8,
8947–8955.
80. Luo, Y., Liu, J., Liu, Y., Lyu, Y. (2017). J. Polym. Sci. Part A: Polym. Chem.,
55, 2594–2600.
81. Bai, L., Phua, S. Z., Lim, W. Q., Jana, A., Luo, Z., Tham, H. P., Zhao, L., Gao,
Q. (2016). Chem. Commun., 52, 4128–4131.
82. Bhanja, P., Mishra, S., Manna, K., Mallick, A., Das Saha, K., Bhaumik, A.
(2017). ACS Appl. Mater. Interfaces, 9, 31411–31423.
83. Zhang, Y., Luo, M., Zu, Y., Fu, Y., Gu, C., Wang, W., Yao, L., Efferth, T. (2012).
Chem. Biol. Interact., 199, 129–136.
84. Tan, J., Namuangruk, S., Kong, W., Kungwan, N., Guo, J., Wang, C. (2016).
Angew. Chem. Int. Ed. Engl., 55, 13979–13984.
85. Zheng, X., Wang, L., Pei, Q., He, S., Liu, S., Xie, Z. (2017). Chem. Mater.,
29, 2374–2381.
86. Lin, G., Ding, H., Chen, R., Peng, Z., Wang, B., Wang, C. (2017). J. Am.
Chem. Soc., 139, 8705–8709.
87. Tan, J., Wan, J., Guo, J., Wang, C. (2015). Chem. Commun., 51, 17394–
17397.
88. Pattani, V. P., Tunnell, J. W. (2012). Lasers Surg. Med., 44, 675–684.
206 Biomedical Applications of COFs

89. Tan, J., Wan, J., Guo, J., Wang, C. (2015). Chem. Commun., 51, 17394–
17397.
90. Croce, R., Van Amerongen, H. (2014). Nat. Chem. Biol., 10, 492–501.
91. Zheng, X., Wang, L., Pei, Q., He, S., Liu, S., Xie, Z. (2017). Chem. Mater.,
29, 2374–2381.
92. Kamiya, K., Kamai, R., Hashimoto, K., Nakanishi, S. (2014). Nat.
Commun., 5, 5040.
93. Lin, S., Diercks, C. S., Zhang, Y.-B., Kornienko, N., Nichols, E. M., Zhao, Y.,
Paris, A. R., Kim, D., Yang, P., Yaghi, O. M. (2015). Science, 349, 1208–
1213.
94. Xu, H., Tao, S., Jiang, D. (2016). Nat. Mater., 15, 722–726.
95. Chandra, S., Kundu, T., Kandambeth, S., Babarao, R., Marathe, Y., Kunjir,
S. M., Banerjee, R. (2014). J. Am. Chem. Soc., 136, 6570–6573.
96. Chandra, S., RoyChowdhury, D., Addicoat, M., Heine, T., Paul, A.,
Banerjee, R. (2017). Chem. Mater., 29, 2074–2080.
97. Deng, W., Li, Y., Zheng, S., Liu, X., Li, P., Sun, L., Yang, R., Wang, S., Wu, Z.,
Bao, X. (2017). Angew. Chem. Int. Ed., 139, 8194–8199.
98. Wang, P., Kang, M., Sun, S., Liu, Q., Zhang, Z., Fang, S. (2014). Chin. J.
Chem., 32, 838–843.
99. Li, W., Yang, C. X., Yan, X. P. (2017). Chem. Commun., 53, 11469–11471.
100. Peng, Y., Huang, Y., Zhu, Y., Chen, B., Wang, L., Lai, Z., Zhang, Z., Zhao, M.,
Tan, C., Yang, N. (2017). J. Am. Chem. Soc., 139, 8698–8704.
101. Dalapati, S., Jin, S., Gao, J., Xu, Y., Nagai, A., Jiang, D. (2013). J. Am. Chem.
Soc., 135, 17310–17313.
102. Guo, J., Xu, Y., Jin, S., Chen, L., Kaji, T., Honsho, Y., Addicoat, M. A., Kim, J.,
Saeki, A., Ihee, H. (2013). Nat. Commun., 4, 2736
103. Wang, T., Xue, R., Chen, H., Shi, P., Lei, X., Wei, Y., Guo, H., Yang, W. (2017).
New J. Chem., 41, 14272–14278.
104. Rengaraj, A., Puthiaraj, P., Haldorai, Y., Heo, N. S., Hwang, S. K., Han, Y.
K., Kwon, S., Ahn, W. S., Huh, Y. S. (2016). ACS Appl. Mater. Interfaces, 8,
8947–8955.
105. Dalapati, S., Jin, E., Addicoat, M., Heine, T., Jiang, D. (2016). J. Am. Chem.
Soc., 138, 5797–5800.
106. Wan, S., Guo, J., Kim, J., Ihee, H., Jiang, D. (2008). Angew. Chem. Int. Ed.
Engl., 47, 8826–8830.
107. Baldwin, L. A., Crowe, J. W., Shannon, M. D., J aroniec, C. P., McGrier, P. L.
(2015). Chem. Mater., 27, 6169–6172.
References 207

108. Lei, C., Shin, Y., Liu, J., Ackerman, E. J. (2002). J. Am. Chem. Soc., 124,
11242–11243.
109. Kandambeth, S., Venkatesh, V., Shinde, D. B., Kumari, S., Halder, A.,
Verma, S., Banerjee, R. (2015). Nat. Commun., 6, 6786.
110. Lin, G., Gao, C., Zheng, Q., Lei, Z., Geng, H., Lin, Z., Yang, H., Cai, Z. (2017).
Chem. Commun., 53, 3649–3652.
111. Lohse, M. S., Stassin, T., Naudin, G., Wuttke, S., Ameloot, R., DeVos, D.,
Medina, D. D., Bein, T. (2016). Chem. Mater., 28, 626–631.
112. Sun, Q., Aguila, B., Perman, J., Earl, L. D., Abney, C. W., Cheng, Y., Wei, H.,
Nguyen, N., Wojtas, L., Ma, S. (2017). J. Am. Chem. Soc., 139, 2786–2793.
113. Huang, N., Zhai, L., Xu, H., Jiang, D. (2017). J. Am. Chem. Soc., 139, 2428–
2434.
114. Mitra, S., Kandambeth, S., Biswal, B. P., Khayum, M. A., Choudhury, C. K.,
Mehta, M., Kaur, G., Banerjee, S., Prabhune, A., Verma, S. (2016). J. Am.
Chem. Soc., 138, 2823–2828.
115. Xiao, B., Wheatley, P. S., Zhao, X., Fletcher, A. J., Fox, S., Rossi, A. G.,
Megson, I. L., Bordiga, S., Regli, L., Thomas, K. M. (2007). J. Am. Chem.
Soc., 129, 1203–1209.
Index

5-fluorouracil (5-FU), 187–88, benzothiazole (BTDA), 141,


190–91, 193–94 153–54
5-FU, see 5-fluorouracil BET, see Brunauer–Emmett– Teller
1,3,5-triformylbenzene (TFB), 40, BET constant, 74
92, 132–33 BET equation, 74
1,3,5-triformylphloroglucinol (TFP), BET surface area, 24, 34–35,
109, 112, 123, 132–33, 194 37–38, 47, 50, 54, 59, 61–62,
1,4-benzenediboronic acid (BDBA), 74, 76–77, 79–81, 83, 119
4, 32, 144, 162 BF COFs, 39–40, 132–33
1,4-benzothiazole-diboronic acid biocompatibility, 186, 193–94, 201
(BTDADA), 153, 156 long-term, 201
biodegradability, 184
AA stacking, 42, 148, 151, 153, biodegradation, 185
156, 170 bioimaging, 183–84, 186–87,
acceptor moieties, 155, 157–58 196–97, 201
acids, 3, 5–6, 22, 24, 27–28, 33, 38, biomaterials, 184
40, 52, 54, 108–9, 111, 128, biomedicine, 186
141–42, 188–90 biomolecules, 197
adsorption, 37, 69–75, 78, 83, 91, biosensing, 183–84, 186–87,
149, 152, 199 196–97, 201
biomolecular, 198 biotoxicity, 104
competitive, 74 boronate esters, 4, 6, 28, 30, 146,
substrate, 105 172, 185
vapor, 73 boronic acids, 21–22, 25, 27–31,
AFM, see atomic force microscopy 33, 172
aggregation-induced emission boronic esters, 21, 24, 28
(AIE), 197 Brunauer–Emmett– Teller (BET),
AIE, see aggregation-induced 13, 24, 73–74, 76, 106, 117,
emission 185
aldehydes, 5, 38–39, 49–50, 52, BTDA, see benzothiazole
94–95, 122–23, 131, 133 BTDADA, see 1,4-benzothiazole-
Arrhenius law, 72 diboronic acid
atomic force microscopy (AFM),
95, 161 carbon footprint, 83
azine-linked COFs, 52–53, 131–32 carbon nanotubes (CNTs), 170–71
catalysts, 2–3, 10, 13, 38, 59, 94,
BDBA, see 1,4-beneznediboronic 104, 106, 108–9, 112, 117,
acid 119, 125, 132, 135
210 Index

acid, 57, 124 drug-loaded, 188–89


active, 115 drug-loading, 192
basic, 33 dynamic, 40
effective, 126 electron acceptor, 162
homogeneous, 105, 116 ester-linked, 38
homogeneous/heterogeneous, fluorinated, 78
128 functional, 86, 89, 123, 191
noble metal, 121 functionalized, 40, 88, 132–33
optimized CTF, 132 hexagonal, 28
porphyrin-based, 108 high-crystallinity, 62
proton reduction, 50, 129 lithium-doped, 75, 77
solid, 71 metal-doped, 75
strong Brønsted acid, 14 metal-loaded, 128
tandem one-pot heterogenous, monolayer, 96
123 multifunctional, 84
catalytic activities, 59, 106–7, 109, nondoped, 75, 77
open-lattice, 162
111, 113, 115, 117, 119–20,
oxygen-rich, 113
122, 124–26, 128, 133–34
photoactive, 130
CCOFs, see chiral covalent organic
photoconductive, 24
frameworks
photosensitive, 195
charge transfer (CT), 148–49,
semiconducting, 141, 146,
151–53, 156, 159
148–49
photoinduced, 158
single-crystal, 62
chemical stability, 4, 6–9, 14, 28,
ternary, 59
38, 40, 47, 177, 184
tunable, 128
good, 50 COF catalysts, 126, 136
high, 177 bimetallic, 127
improved, 47 stable bifunctional, 126
chemisorption, 71–72 COF crystals, 3, 23, 42, 61, 105,
activated, 72 130, 134
nonactivated, 72 COF films, 29, 69, 90, 131, 151–52,
chiral covalent organic frameworks 157–58, 165, 196–97
(CCOFs), 119–24 COF materials, 15, 52, 62, 75, 128,
clean energy source, 166 167, 177–78, 201
alternative, 96 COF membranes, 84–86, 90–91,
CNTs, see carbon nanotubes 93–100, 190
COF, see covalent organic COF nanosheets, 87, 93, 98
framework conjugated COFs, 5, 54, 197
bulk, 14, 93 CONs, see covalent organic
chiral, 119–21, 123–24 nanosheets
crystalline mesoporous, 50 covalent bonds, 2, 4, 10, 21, 59–61,
donor-acceptor, 156–57, 160–62 69, 71, 103
Index 211

dynamic, 1, 4, 184–85 moderate, 117


irreversible, 2 Diels–Alder cycloaddition reaction,
traditional, 185 122
covalent organic framework (COF), differential pulse voltammetry
1–6, 9–15, 21–26, 28–40, 42, (DPV), 158, 160
44–52, 54–56, 58–62, 64, DLE, see drug loading efficacy
69–70, 72–100, 102–36, docking sites, 107, 197
138–78, 183–88, 190–202 donor moieties, 155, 157–58
covalent organic nanosheets DOX, see doxorubicin
(CONs), 14, 190–91, 199 doxorubicin (DOX), 190, 194
covalent triazine frameworks DPV, see differential pulse
(CTFs), 5, 13–14, 57–59, voltammetry
78–79, 89, 98, 132, 167–70, drug delivery, 1, 55, 183–85,
185 187–88, 190–92, 201
crystalline COFs, 12, 31, 33, drug loading, 187, 190–91
135–36, 172 drug loading efficacy (DLE), 187,
crystallinity, 3–4, 6–9, 14, 21, 24, 192–94
28, 32, 38, 40, 42, 45, 47, 50, drug release, 186–87, 190, 194
61, 107, 113, 117, 121, 146, controllable, 190
149, 187 on-demand, 189
crystallization, 2–3, 11–12, 21, 23, in vitro, 187
46, 49, 94–95, 155 drug release rate (DDR), 194
CS COF, 55, 57, 165–66 dynamic covalent bonds (DCBs), 1,
CT, see charge transfer 4–5, 125, 184–85
CTFs, see covalent triazine dynamic covalent chemistry (DCC),
frameworks 2–4, 21, 36, 55
organic, 59 dynamic light scattering, 196
stable, 132
CV, see cyclic voltammetry eclipsed stacking, 144, 149, 151,
cyclic voltammetry (CV), 151–52, 153–55
174 electron paramagnetic resonance
(EPR), 152, 190
DCBs, see dynamic covalent bonds electron spin resonance (ESR), 153
DCC, see dynamic covalent electron transfer, 149, 158,
chemistry 160–61, 163
DDR, see drug release rate efficient, 158
density functional theory (DFT), photoinduced, 162–63
162, 173 enantioselectivity, 123–25
design and synthesis, 2–12, 14, 16, energy storage, 103, 139, 167, 177,
18, 61, 117 196
DFT, see density functional theory energy transfer, 142
diastereoselectivity, 123 fluorescence resonance, 197
212 Index

EPR, see electron paramagnetic versatile bifunctional, 134


resonance hexahydroxyltriphenylene (HHTP),
ESR, see electron spin resonance 4, 11, 25, 28, 30–32, 52, 142,
148, 156, 164
FDA, see Food and Drug hexabenzocoronene (HBC), 146
Administration hexaphenylbenzene (HPB), 47, 146
flash-photolysis time-resolved HHTP, see
microwave conductivity (FP- hexahydroxyltriphenylene
TRMC), 145, 147, 150, 153–54, highest occupied molecular orbital
156, 159, 166 (HOMO), 140
fluorescence, 142, 156, 196–98 hollow nanospheres, 87
Food and Drug Administration HOMO, see highest occupied
(FDA), 184 molecular orbital
force fields, 71 HPB, see hexaphenylbenzene
nonuniform, 70 hydrazone-linked COFs, 50–51,
residual, 70–71 129
FP-TRMC, see flash-photolysis hydrogen bonding, 5, 41, 43
time-resolved microwave intermolecular, 5, 125
conductivity hydrophilicity, 84, 86–87, 89
Friedel–Crafts alkylation, 14, 190 hydrophobicity, 84, 99
Friedel–Crafts reaction, 190
ICOFs, see ionic covalent organic
gas separation, 1, 35, 62, 69, 84, frameworks
87, 93, 96, 98–99, 103 iCONs, see ionic covalent organic
Gibb free energy, 3 nanosheets
GNSs, see graphene nanosheets charged, 199
graphene, 29, 92, 161, 175–77 mixed, 199
graphene nanosheets (GNSs), 175 imine-linked COFs, 6, 14, 39, 42,
49–50, 118, 195
HBC, see hexabenzocoronene indium tin oxide (ITO), 146, 152,
heat, 70–72 158, 165
isosteric, 37 interactions, 24, 42, 69, 83, 97,
Heck cross-coupling reactions, 134, 140, 189, 197
104, 109, 112, 115, 127 donor-acceptor, 54, 140
Heck reaction electronic, 42
reductive, 119–20 electrostatic, 197
Henry reaction, 122 gas-solid, 73
asymmetric, 119 ground-state, 159
heterogeneous catalysts, 50, hydrogen-bonding, 40, 42, 50,
103–4, 119, 132, 136 197
bifunctional, 126 hydrophobic, 190
efficient, 121 intermolecular, 87
Index 213

inter-TTF-layer, 151 lowest occupied molecular orbital


physical, 185 (LUMO), 140
relatively weak, 42 luminescence, 10, 140, 197
spin-spin, 153 LUMO, see lowest occupied
strong, 109, 133 molecular orbital
strong metal-sulfur, 111
strong nitrogen-palladium, 106 MC, see mechanochemical
traditional, 4 MC COFs, 14
interface, 89, 94–95 mechanochemical (MC), 12–14
liquid/air, 94 mesopores, 50, 59, 69, 78, 129
oil/water, 94 honeycomb, 129
water/air, 94 metal ions, 9–10, 122, 145–46,
interfacial polymerization (IP), 184, 195, 197
85–86, 89, 94, 99 metal NPs, 109, 111–12
ionic covalent organic frameworks metal-organic frameworks (MOFs),
(ICOFs), 33–34
2, 9–10, 35, 62, 75–77, 84,
ionic covalent organic nanosheets
104–5, 172, 175–77, 184, 186,
(iCONs), 199–200
196, 200
IP, see interfacial polymerization
Michael addition elimination
irradiation, 50, 145, 147, 153,
reactions, 5
155–56, 159–60, 190, 196
Michael addition reaction, 119
laser pulse, 147, 150
Michael reactions, 121
visible-light, 132
micropores, 52, 74, 78, 166
ITO, see indium tin oxide
mixed matrix membranes (MMMs),
Knoevenagel condensation, 40, 86–87, 89, 97
133 MMMs, see mixed matrix
base-catalyzed, 133–34 membranes
Knoevenagel reaction, 126 model drug, 187, 190, 192, 194
MOFs, see metal-organic
Langmuir model, 37, 74 frameworks
Langmuir surface area, 37, 40, 117 amine-decorated, 196
Langmuir theory, 73 amine-modified, 196
layered structures, 42, 45, 50, 129, novel, 2
140, 144 monolayers, 72–74, 92–93, 96, 156
extended staggered, 24 adsorbed, 73
Lewis acid, 3–4, 30, 94 graphene oxide, 92
Lewis-base catalyst, 133 monomers, 3, 10, 12–13, 47, 50,
LIB, see Li-ion battery 52, 86, 89, 94–95, 125, 129,
Li-ion battery (LIB), 166–67, 170, 142, 148, 155, 185–86
172, 175–77 MP COFs, 145, 148
loading capacities, 184–85, 187,
198 nanocarriers, 190, 192
loading efficacy, 192 nanochannels, 140, 149
214 Index

nanoclusters, 87, 195 phosphate-buffered saline (PBS),


nanofiltration, 84, 89, 99 188–89
nanoparticles (NPs), 108–9, 111, photocatalysis, 50
113, 115, 119, 186, 190, photocatalysts, 129–32
193–94 photoconductivity, 142, 145–46,
composite, 199 155, 159
drug-loaded, 188 photocurrent, 143, 145–48, 153,
pH-responsive CTF, 190 155, 159
uniform, 192 photoinduced absorption
uniform spherical, 188 spectroscopy spectrum,
nanopores, 1, 103, 139, 168 157–58
nanosheets, 87, 92–93, 98, 190, photosensitizers, 132, 195
199 photothermal therapy (PTT), 195,
near-infrared (NIR), 145, 152–53, 201
159 phthalocyanine COFs, 29, 145, 147,
NIPS, see nonsolvent-induced 153
phase inversion physical adsorption, 70–73
NIR, see near-infrared PLDs, see pore-limiting diameters
NMR, see nuclear magnetic polyether sulfone (PES), 83, 87, 89
resonance polyimide COFs, 55, 187
nonsolvent-induced phase polymerization, 10, 94–95
inversion (NIPS), 85–86 interfacial, 85, 94
NPs, see nanoparticles thermal, 36
nuclear magnetic resonance polymers, 2–3, 36–37, 50, 83,
(NMR), 22, 107, 111 85–87, 89, 97–98, 154, 183
amorphous, 83, 94–95
one-pot cascade reactions, 128 amorphous nonporous, 119
one-pot deacetalization- conductive, 172
Knoevenagel reaction, 134 cross-linked, 83
one-pot silane oxidation reaction, cross-linking, 3
115 hyperblanching, 3
optoelectronics, 1, 33, 103, 139 microporous, 2
organic, 36 organic, 13, 166, 200
organocatalytic COF, 116–17, 125 polyborazylene, 36
polysulfone (PSF), 200
PBS, see phosphate-buffered saline polyvinylidene fluoride (PVDF),
PCTFs, see porphyrin-based 83, 89
covalent triazine frameworks pore-limiting diameters (PLDs),
metal-free, 187 96–97
PDBA, see pyrene-2,7-diboronic pore sizes, 9–10, 25, 27–30, 37–38,
acid 47, 52, 54–56, 71–72, 74–77,
permeability, 87, 94, 96–99, 190 79–81, 83–84, 98–99, 185–87,
PES, see polyether sulfone 192–94, 198
Index 215

controllable, 61 PTT, see photothermal therapy


inconsistent, 95 PVDF, see polyvinylidene fluoride
large, 24, 184 PXRD, see powder X-ray diffraction
narrow, 55 pyrenes, 40, 52, 141–42, 149, 186
physical, 84 pyrene-2,7-diboronic acid (PDBA),
tunable, 84, 99, 167 142
typical nonuniform, 83
pore space, 104, 109, 111, 115, quenching, 3, 198
119, 128, 133
pore structures, 10, 84–87, 89, 98, recyclability, 109, 111, 115, 120,
105, 184, 186 128, 133
porosity, 10, 28, 32, 37, 40, 42, 62, reversible condensation, 38, 129
76–77, 79, 83, 87, 107, 117, reversible dimerization, 5
131, 172 room temperature (RT), 12, 14, 22,
controllable, 9 34, 36, 57, 110
limited, 83 RT, see room temperature
permanent, 52, 61
tunable, 1, 183 Schiff base reactions, 47–48
ultrahigh, 14 organic, 94
porous materials, 1–2, 9–10, selectivity, 2, 83, 87, 92, 96–98,
74–75, 78, 85–86, 103–4, 172, 105, 136, 172, 197–98
185 self-condensation, 4, 10, 23–24, 58,
porous structures, 10, 59, 86–87, 142, 190
89, 94, 119 self-healing feedback, 3
porphyrin-based covalent triazine singlet oxygen, 131, 186, 195–96
frameworks (PCTFs), 187, 194 solvothermal condensation, 40,
porphyrin COFs, 40, 42–43, 108, 146, 156
159 solvothermal conditions, 24, 28,
porphyrins, 5, 40, 107, 125–26, 35, 52, 55, 106, 109, 129, 144,
131, 141, 143, 146, 155, 149, 153, 164
157–59, 172, 186, 195 solvothermal reactions, 62, 159
powder X-ray diffraction (PXRD), solvothermal synthesis, 12, 35, 134
33, 35, 50, 106, 116, 127, 134, Sonogashira coupling reaction, 112
199 Sonogashira cross-coupling, 116
PRC, see proton reduction catalyst Sonogashira-type reaction, 113
proton reduction catalyst (PRC), spectroscopy, 22, 156
50, 129 infrared, 107
PSF, see polysulfone time-resolved fluorescence, 156
PT COFs, 142 spiroborate-linked COFs, 33–34,
p-toluene sulfonic acid (PTSA), 185
90, 94 spiroborates, 6, 21, 33
PTSA, see p-toluene sulfonic acid squaraine-based COFs, 54
216 Index

standard temperature and trigonal, 47


pressure (STP), 77 toxicity, 184, 201
STP, see standard temperature and TP COFs, 141–43
pressure TR-ESR, see time-resolved electron
Suzuki cross-coupling, 107 spin resonance
Suzuki–Miyaura coupling reaction, trifluoroacetic acid (TFA), 189, 194
107 TTF, see tetrathiafulvalene
Suzuki reaction, 107 TTI COF, 188–89, 193
turn-over frequency (TOF), 105,
tautomerism, 48 109
T COFs, 25, 141, 148–50 turn-over numbers (TONs), 109,
tetrathiafulvalene (TTF), 141, 132
148–53
TFA, see trifluoroacetic acid 189, ultrafiltration, 84
194 ultraviolet (UV), 152, 190–91, 194
TFB, see 1,3,5-triformylbenzene UNM, 196
TFN, see thin film nanocomposite UV, see ultraviolet
TFP, see
1,3,5-triformylphloroglucinol vacuum, 12, 107
TGA, see thermogravimetric vacuum-assisted filtration, 93
analysis van der Waals adsorption, 70
thermogravimetric analysis (TGA), van der Waals forces, 70, 93
187 van der Waals radius, 91
thin film nanocomposite (TFN), 69, vapor deposition, 143
89–90 visible light, 50, 52, 130, 132, 143,
time-of-flight (TOF), 145–47 147, 153, 155–56, 159
time-resolved electron spin
resonance (TR-ESR), 156, 160, wide-angle X-ray scattering, 152
162
TMC, 89 X-ray crystallographic analysis, 22
TOF, see turn-over frequency XRD, see X-ray diffraction
TOF, see time-of-flight X-ray diffraction (XRD), 47, 117,
TONs, see turn-over numbers 168–70
topologies, 9–11, 40, 47, 69, 146
ctn, 135 zeolites, 2, 9, 84, 103
new, 21 microporous, 9
rra, 35 ZnPc COFs, 145–46, 162–63
tetragonal, 47 electron-donating, 162
triangular, 146 ZnP COF, 159

You might also like