Sie sind auf Seite 1von 318

CHEMICAL ENGINEERING METHODS AND TECHNOLOGY

COORDINATION POLYMERS AND


METAL ORGANIC FRAMEWORKS:
PROPERTIES, TYPES
AND APPLICATIONS

No part of this digital document may be reproduced, stored in a retrieval system or transmitted in any form or
by any means. The publisher has taken reasonable care in the preparation of this digital document, but makes no
expressed or implied warranty of any kind and assumes no responsibility for any errors or omissions. No
liability is assumed for incidental or consequential damages in connection with or arising out of information
contained herein. This digital document is sold with the clear understanding that the publisher is not engaged in
rendering legal, medical or any other professional services.
CHEMICAL ENGINEERING
METHODS AND TECHNOLOGY

Additional books in this series can be found on Novas website


under the Series tab.

Additional E-books in this series can be found on Novas website


under the E-book tab.

MATERIALS SCIENCE AND TECHNOLOGIES

Additional books in this series can be found on Novas website


under the Series tab.

Additional E-books in this series can be found on Novas website


under the E-book tab.
CHEMICAL ENGINEERING METHODS AND TECHNOLOGY

COORDINATION POLYMERS AND


METAL ORGANIC FRAMEWORKS:
PROPERTIES, TYPES
AND APPLICATIONS

OSCAR L. ORTIZ AND LUIS D. RAMREZ


EDITORS

Nova Science Publishers, Inc.


New York
Copyright 2012 by Nova Science Publishers, Inc.

All rights reserved. No part of this book may be reproduced, stored in a retrieval system or
transmitted in any form or by any means: electronic, electrostatic, magnetic, tape, mechanical
photocopying, recording or otherwise without the written permission of the Publisher.

For permission to use material from this book please contact us:
Telephone 631-231-7269; Fax 631-231-8175
Web Site: http://www.novapublishers.com

NOTICE TO THE READER


The Publisher has taken reasonable care in the preparation of this book, but makes no expressed or
implied warranty of any kind and assumes no responsibility for any errors or omissions. No
liability is assumed for incidental or consequential damages in connection with or arising out of
information contained in this book. The Publisher shall not be liable for any special,
consequential, or exemplary damages resulting, in whole or in part, from the readers use of, or
reliance upon, this material. Any parts of this book based on government reports are so indicated
and copyright is claimed for those parts to the extent applicable to compilations of such works.

Independent verification should be sought for any data, advice or recommendations contained in
this book. In addition, no responsibility is assumed by the publisher for any injury and/or damage
to persons or property arising from any methods, products, instructions, ideas or otherwise
contained in this publication.

This publication is designed to provide accurate and authoritative information with regard to the
subject matter covered herein. It is sold with the clear understanding that the Publisher is not
engaged in rendering legal or any other professional services. If legal or any other expert
assistance is required, the services of a competent person should be sought. FROM A
DECLARATION OF PARTICIPANTS JOINTLY ADOPTED BY A COMMITTEE OF THE
AMERICAN BAR ASSOCIATION AND A COMMITTEE OF PUBLISHERS.

Additional color graphics may be available in the e-book version of this book.

Library of Congress Cataloging-in-Publication Data

Coordination polymers and metal organic frameworks : properties, types, and applications /
editors, Oscar L. Ortiz and Luis D. Rammrez.
p. cm.
Includes bibliographical references and index.
ISBN  ((%RRN)
1. Coordination polymers. I. Ortiz, Oscar L. II. Rammrez, Luis D.
QD382.C67C66 2011
547'.7--dc23
2011027339

Published by Nova Science Publishers, Inc. New York


CONTENTS

Preface i
Chapter 1 Hybrid Vanadates, towards Metal-Organic Frameworks 1
Edurne S. Larrea, Roberto Fernndez de Luis, Jos L. Mesa,
Jos L. Pizarro, M. Karmele Urtiaga, Tefilo Rojo
and Mara I. Arriortua
Chapter 2 Structure and Magnetic Properties of Mono - and Polynuclear
Complexes Containing Rhenium(IV) 59
Carlos Kremer and Ral Chiozzone
Chapter 3 The Applications of Metal Organic Frameworks in the Fields of
Hydrogen Storage and Catalysis 99
Yaoqi Li, Ping Son, Yan Li and Xingguo Li
Chapter 4 MOF-Based Mixed-Matrix-Membranes for Industrial
Applications 129
Hoang Vinh-Thang and Serge Kaliaguine
Chapter 5 Coordination Polymers: Opportunities in Heterogeneous
Catalysis 169
Francesc X. Llabrs i Xamena
Chapter 6 High Pressure Gas Storage on Porous Solids: A Comparative
Study of MOFs and Activated Carbons 197
A. Linares-Solano, D. Cazorla-Amors, J. P. Marco-Lozar
and F. Surez-Garca
Chapter 7 Metal-Organic Frameworks for CO2 Capture: What are
Learned from Molecular Simulations 225
Jianwen Jiang
Chapter 8 Halogen Bonding in the Assembly of High-Dimensional
Supramolecular Coordination Polymers 249
Ya-Juan Fan, Shuang-Quan Zang, Yong-Li Wei,
Chen-Xia Du and Hong-Wei Hou
vi Contents

Chapter 9 Subtractive Approach for Introducing Functional Groups onto


MetalOrganic Framework 277
Teppei Yamada and Hiroshi Kitagawa
Chapter 10 Performance of Metal-Organic Framework MIL-101 in the
Liquid Phase Adsorption of Heterocyclic Nitrogen Compounds 291
Alexey L. Nuzhdin, Konstantin A. Kovalenko,
Vladimir P. Fedin and Galina A. Bukhtiyarova
Index 297
PREFACE
In this book, the authors present topical research in the study of coordination polymers
and metal organic frameworks. Topics discussed include hybrid vanadates and metal organic
frameworks; structure and magnetic properties of mono- and polynuclear complexes
containing Re(IV)l; metal organic framework applications in the fields of hydrogen storage
and catalysis; MOF-Based mixed-matrix-membranes for industrial applications; coordination
polymers in heterogeneous catalysis; high pressure gas storage on porous solids; metal
organic frameworks for CO2 capture and halogen bonding in the assembly of high-
dimensional supramolecular coordination polymers. (Imprint: Nova)
Chapter 1 - The combination of metal-organic polymers with different inorganic
oxoanions has become a great strategy to obtain highly complex crystal architectures. Those
materials present metal-organic and inorganic subnets combined in the same crystal structure.
In this sense, hybrid vanadates exhibit a vast crystal chemistry, ranging from structural
archetypes comparable to that of the aluminophosphates or transition metal phosphates to
flexible structures similar to MOFs. Hybrid vanadates with first-row transition metals exhibit
several structural archetypes according to the metal center, the geometry of the ligand and the
vanadium oxide subunit. Vanadium shows a wide variety of oxidation states, each of them
taking different coordination environments. In addition, vanadium polyhedra have a great
ability to polymerize, giving rise to clusters, rings, chains, layers and three-dimensional
substructures. The polymer grade is closely related to the synthetic conditions and, specially,
to the pH during the reaction. The geometry of the ligand and the coordination environment
of the metal centers also plays an important role in the final complexity, dimensionality and
functionality of the crystal structures of hybrid vanadates. Consequently, a deep analysis of
the crystal archetypes observed in hybrid vanadates has allowed us to propose a classification
based on the metal-organic and inorganic substructure dimensionalities. The properties of
these materials are directly related to the structural characteristics, depending directly on the
synergetic interaction between the metal-organic and vanadium oxide subunits. This way, the
loss of solvent in hybrid vanadates could generate a flexible, dynamical and reversible
response of the crystal structure, as in some MOFs, or rigid behaviors, without significant
structural changes, as in inorganic zeolites. Likewise, in the hybrid vanadates, the loss of
coordinated water molecules bonded to the metal centers gives rise to irreversible structural
transformations with a drastic reduction of the crystallinity. The magnetic properties in hybrid
vanadates depend on the connectivity between the metal centers. The magnetic exchange can
also take place through the vanadate oxoanion, giving rise to dimmeric or one-dimensional
magnetic behaviors. The catalytic and photocatalitic tests of several hybrid vanadates reveal
ii Oscar L. Ortiz and Luis D. Ramrez

that they could be active materials in oxidation reactions or for the decomposition of
pollutants.
Chapter 2 - One of the major goals in inorganic supramolecular chemistry today is the
design of polynuclear coordination arrays and the study of their magnetic properties. With the
generation of well-defined architectures it is possible to understand the different factors which
determine the exchange coupling between spin carriers. Most of the results found in the
literature are focused on polynuclear complexes containing metal ions belonging to the first
transition series. Once the magnetic interaction between 3d metal ions is well understood, the
study of those systems containing 4d or 5d metal ions becomes very interesting. In this
review we revise the structure and magnetic properties of Re(IV) complexes. Rhenium(IV), a
5d3 ion, usually forms octahedral complexes which are reasonably stable against redox
processes and inert to ligand substitution. This is the basis for the preparation of mononuclear
species that can act as ligands towards first-row transition metal ions. For example,
complexes containing dicarboxylic ligands, [ReX4(ox)]2 and [ReX4(mal)]2 (X = Cl, Br; ox =
oxalato; mal = malonato), or N-donor ligands, [ReCl5(pyz)] (pyz = pyrazine) have been used
as building blocks to construct heteropolynuclear complexes. The different designed
structures, from discrete binuclear complexes to extended chain-like compounds, are
reviewed in this work. In addition, the magneto-structural studies of these mono- and
polynuclear complexes are also included and discussed.
Chapter 3 - MOFs (metal organic frameworks) are porous frameworks constructed by the
coordination centers of metal ions and polyatomic organic bridging ligands. Nowadays,
MOFs have attracted much attention as they have been widely investigated for hydrogen
storage, gas separation and catalysis. The microporous porosity, the large specific surface
area and especially the controllable framework have made MOFs superior to traditional
inorganic porous materials such as zeolites and activated carbon. For gas storage and
separation, physisorption on MOFs is one significant approach for future application. By
crystal engineering, both the pore size and the electronic and chemical nature of the interior
surface, on which gas molecules will be adsorbed, can be modified by careful designs. Not
only the porous structure, but also the metal sites and the organic linkers in the frameworks
should greatly affect the interactions between the gas molecules and MOFs. Comparing with
other microporous materials, the low framework density, the high specific surface area,
especially the controllable crystal structure have made MOFs be a favorable research interest.
Furthermore, the frameworks of MOFs contain various structures and large amounts of
essential metal ions, which should be helpful to promote molecular separations and chemical
reactions. Therefore, MOFs can be potentially applied as a heterogeneous catalyst. MOFs
have been used as good precursors of catalysts and special substrates for dispersed active
sites. The existence of abundant metal ions, the large surface areas and the tailorable
microporous structures in MOFs clearly help to obtain highly efficient catalysts.
Chapter 4 - Modern membrane gas separation technology has enjoyed a rapid
development of its commercial applications in the chemical, petrochemical, semiconductor,
food, pharmaceutical, biotechnology and environmental industries, due to their low energy
consumption, compelling low cost and ease of large-scale operation.
Mixed-matrix membranes (MMMs) combine some of the assets of polymer membranes
with the increased separation selectivities associated with the presence of a load of inorganic
particles. In this chapter the potential advantages of metal-organic frameworks (MOFs) as the
discrete phase in MMMs are reviewed.
Preface iii

Chapter 5 - Coordination Polymers (CPs) are an emerging class of materials that are
attracting considerable interest in recent years. Their unique properties make these materials
very promising for applications in a number of fields, including heterogeneous catalysis. In
this chapter, with 126 references, we will revise the main strategies that have been
specifically developed for introducing catalytic active sites in these materials. The enormous
possibilities of this class of materials will be outlined throughout selected examples taken
from the recent literature. I hope that this chapter will be useful either as an introductory
lecture to those who approach the field of CPs or heterogeneous catalysis for the first time, as
well as an updated state-of-the-art vision for all scientists working in this field.
Chapter 6 - Porous materials provide an alternative for satisfying gas storage demands for
on-board storage in transportation technologies (e.g. CH4 and H2) and for capture, storage and
transport (e.g. CO2). The principle of their storages is the use of a high pressure adsorption
process (or physisorption), as a supercritical gas (e.g. H2 and CH4) or as a subcritical one (e.g.
CO2). Such adsorption process has some advantages as: its high storage capacity (very much
depending on the surface area, porosity and pore size of the material), its fast kinetic of
storage and release (reversibility), its short refueling time, its low heat evolution and its
efficient cyclability. Additionally, the porous solid (the adsorbent) presents advantages;
different types are available (e.g., zeolites, porous carbons, MOFs, all of them with a large
variety of materials), its porosity, morphology, size and shape are tunable. Among them, two
types of porous solids stand out: 1) the cl assical activated carbons and 2) the r ecent and
new type of porous materials (i.e. MOFs and COFs). Most of the papers report the gas
storage capacity of an adsorbent refereed per unit of weight (i.e. gravimetric basis). However,
for applications where the volume of the tank is an important controlling factor (e.g. in
transportation), the gas storage capacity should also be reported per unit of volume
(volumetric basis). And, the density of material used should be consistently measured (i.e. tap
or packing). Unfortunately, this is not always the case and very frequently (as it happens with
recent and new porous materials) calculated density (e.g. crystal density) is used. Using
such crystal density, impressive volumetric storage capacities have been reported for MOFs
(also COFs), claiming that they can achieve higher storage capacities for H2, CH4 and CO2
than other porous materials such as zeolites and porous carbons. In our opinion, such claimed
superior gas storage capacity of MOFs in relation to activated carbons needs further
evaluations. In this chapter, we comparatively analyses the adsorption capacity of two
activated carbons (ACs) and MOF-5 for storing gases (H2, CH4 and CO2) at different
temperatures (77K and RT) and pressures (from 0.1MPa to 20MPa) both on gravimetric and
volumetric basis paying attention to the data reported in the literature as well as on the
suitability of different densities employed. We advance that, from the data presented and
discussed in this chapter, the outstanding adsorption capacities of MOFs in relation to ACs on
volumetric basis, frequently claimed in the literature, is mainly due to the use of an unrealistic
high density (crystal density) which, not including the inter-particle space of the adsorbents,
gives an apparently high volumetric gas storage capacity. Using a density measured similarly
in both types of adsorbents (e.g. tap density) MOF presents, on volumetric basis, and for all
gases and conditions studied, lower adsorption capacities than ACs, due to its lower inherent
density.
Chapter 7 - CO2 capture is currently a topical issue in environmental protection and
sustainable development. This chapter reviews the recent molecular simulation studies for
CO2 capture in metal-organic frameworks (MOFs). Emerged as an intriguing class of
iv Oscar L. Ortiz and Luis D. Ramrez

nanoporous materials, MOFs have been considered versatile candidates for storage,
separation, catalysis and other widespread potential applications. However, the number of
MOFs synthesized to date is extremely large, experimentally testing and screening of ideal
MOFs for high-performance CO2 capture is formidable and time-consuming. With ever-
growing computational resources and advance in mathematical techniques, molecular
simulations have become an indispensable tool for materials characterization, screening and
design. At a molecular level, simulations can provide microscopic insights from the bottom-
up and establish structure-function relationships. Here, representative simulation studies are
summarized for CO2 capture in MOF sorbents and membranes respectively, strategies
(catenation, functionalization, exposed metals, ionic frameworks and metal doping) are
discussed for improving capture performance, and the effects of water on CO2 capture are
also considered. The chapter is concluded with the key insights learned from simulations and
the outlook for future endeavors.
Chapter 8 - The investigation of supramolecular assemblies based on halogen bonding
(XB) has been a field with rapid growth because a large variety of novel architectures which
were constructed through halogen bonding have been reported to possess potential
applications. Halogen bonding as well as related halogenhalogen and halogen
intermolecular interactions found in a given crystalline is valuable to inorganic chemists on
their study and poses an interesting challenge. This chapter will give a concise overview on
recent developments in the syntheses and preparations of high-dimensional supramolecular
coordination architectures based on halogen-related interactions. The interplay of
coordination bonds and such intermolecular forces highlights the complexity and challenge in
supramolecular assembly of high-dimensional coordination polymers.
Chapter 9 - The aim of this work is to apply a new procedure in the synthesis of metal
organic frameworks (MOFs), where the organic moiety bears functional groups that can be
involved in the proton conduction mechanism. Several method can be applied for introducing
functional groups onto MOFs, including a naive method which simply uses a ligand having
acidic functional groups for constructing a MOF, and several papers succeeded in introducing
it. However, acidic functional groups have ability to coordinate to metal site as well as other
functional groups for coordinating, and the obtained MOF structure is a sort of creation of a
chance.
Chapter 10 - The heterocyclic nitrogen compounds containing in liquid hydrocarbons
streams poison many industrial catalysts. The selective removal of the nitrogen species from
refinery streams by adsorption at ambient temperature is a promising approach. Remarkable
adsorption capacity and selectivity towards heterocyclic nitrogen compounds were observed
for metal-organic framework [Cr3O(C8H4O4)3F(H2O)2] (MIL-101) under the sorption from
isooctane and the hydrotreated gas oil. The adsorption capacity of MIL-101 towards
heterocyclic nitrogen compounds is significantly higher than the capacity of conventional
adsorbents such as activated carbons, activated alumina, silica-based adsorbents and zeolite-
based materials especially under the sorption from hydrocarbon with low nitrogen content.
The spent MIL-101 can be regenerated and reused in the next adsorption cycle without loss of
its adsorption capability. The very high adsorption capacity and selectivity of metal-organic
framework MIL-101 for the nitrogen compounds, along with its good regenerability, indicate
that MIL-101 may be the promising adsorbent for deep denitrogenation of liquid hydrocarbon
streams.
In: Coordination Polymers and Metal Organic Frameworks ISBN: 978-1-61470-899-5
Editors: O. L. Ortiz and L. D. Ramirez, pp. 1-58 2012 Nova Science Publishers, Inc.

Chapter 1

HYBRID VANADATES, TOWARDS


METAL-ORGANIC FRAMEWORKS

Edurne S. Larrea,a Roberto Fernndez de Luis,a


Jos L. Mesa,b Jos L. Pizarro,a M. Karmele Urtiaga,a
Tefilo Rojob and Mara I. Arriortuaa
a
Departamento de Mineraloga y Petrologa and bDepartamento de Qumica
Inorgnica, Facultad de Ciencia y Tecnologa. Universidad del Pas Vasco,
UPV/EHU. Apdo. 644, E-E-48080 Bilbao. Spain

ABSTRACT
The combination of metal-organic polymers with different inorganic oxoanions has
become a great strategy to obtain highly complex crystal architectures. Those materials
present metal-organic and inorganic subnets combined in the same crystal structure.
In this sense, hybrid vanadates exhibit a vast crystal chemistry, ranging from
structural archetypes comparable to that of the aluminophosphates or transition metal
phosphates to flexible structures similar to MOFs.
Hybrid vanadates with first-row transition metals exhibit several structural
archetypes according to the metal center, the geometry of the ligand and the vanadium
oxide subunit. Vanadium shows a wide variety of oxidation states, each of them taking
different coordination environments. In addition, vanadium polyhedra have a great ability
to polymerize, giving rise to clusters, rings, chains, layers and three-dimensional
substructures. The polymer grade is closely related to the synthetic conditions and,
specially, to the pH during the reaction. The geometry of the ligand and the coordination
environment of the metal centers also plays an important role in the final complexity,
dimensionality and functionality of the crystal structures of hybrid vanadates.
Consequently, a deep analysis of the crystal archetypes observed in hybrid vanadates has
allowed us to propose a classification based on the metal-organic and inorganic
substructure dimensionalities. The properties of these materials are directly related to the
structural characteristics, depending directly on the synergetic interaction between the
metal-organic and vanadium oxide subunits. This way, the loss of solvent in hybrid

Corresponding author: e-mail: maribel.arriortua@ehu.es; phone: +34946012534; fax: +34946013500.


2 Edurne S. Larrea, Roberto Fernndez de Luis, Jos L. Mesa et al.

vanadates could generate a flexible, dynamical and reversible response of the crystal
structure, as in some MOFs, or rigid behaviors, without significant structural changes, as
in inorganic zeolites. Likewise, in the hybrid vanadates, the loss of coordinated water
molecules bonded to the metal centers gives rise to irreversible structural transformations
with a drastic reduction of the crystallinity. The magnetic properties in hybrid vanadates
depend on the connectivity between the metal centers. The magnetic exchange can also
take place through the vanadate oxoanion, giving rise to dimmeric or one-dimensional
magnetic behaviors. The catalytic and photocatalitic tests of several hybrid vanadates
reveal that they could be active materials in oxidation reactions or for the decomposition
of pollutants.

1. INTRODUCTION
Materials Science and Solid State Chemistry have contributed significantly to the design,
development and optimization of new materials with physicochemical properties suitable for
specific and direct applications.
The zeolites, alumino-silicates, alumino-phosphates or transition metal phosphates,
arsenates, phophites, germanates are classic porous materials that possess inorganic rigid
scaffolds built around different structural templates. [1] The development of zeotype
materials during the last decades has allowed obtaining a great variety of porous materials,
some of them with direct applications in industrial processes. Many structures of zeolitic
inorganic solids with an anionic skeleton often collapse during the extraction of the cationic
template, owing to the strong electrostatic host-guest interactions, which energetically
represent an important contribution to the lattice energy.
Nowadays, the scientific research has focused on the synthesis of new coordination
polymers, also referred to as metal-organic frameworks (MOFS). [2] Coordination polymers
contain two central components, connectors and linkers. These are defined as starting
reagents, by which the main framework of the coordination polymer is built. In addition, there
are other auxiliary components, such as blocking ligands, counter-anions, and non-bonding
guest or template molecules. The structural integrity of the building blocks is maintained
throughout the reaction, which allows their use as modules bricks in the assembly of extended
structures. The final crystal structure is based on the connectivity between the metal centers
or inorganic clusters through the organic ligands. [3] The key to success is the choice and/or
des ign of the molecular building blocks which would direct the formation of the desired
structural, chemical, and physical properties of the resulting materials.
Consequently, the structures and properties of coordination polymers can be well-
designed and systematically tuned by the judicious choice of metal-based building blocks and
organic linkers, in principle.
An increasing number of coordination polymers have been studied for their interesting
properties, including optic, magnetic and electronic properties, as well as for their various
potential applications such as catalysis, ion exchange, gas storage, separation, sensing,
polymerization, and drug delivery. [4]
With respect to the porosity, there are four types or porous structures, 0D cavities (dots),
1D channels, 2D layers and 3D intersecting channels. On the other hand, the response of the
crystal framework to the loss of solvent defines three groups of compounds: first generation
materials, which collapse due to the removal of guest molecules, second generation
Hybrid Vanadates, towards Metal-Organic Frameworks 3

compounds, which have a stable and robust porous framework, such as the zeotypes, and
finally, third generation materials, whose crystal framework show a flexible and dynamical
response to external stimuli, such as light, temperature, pressure, electric field, guest
molecules, changing the shape and/or the size of their channels or pores reversibly. Structural
reasons must exist for explaining such behavior: i) the host-guest interactions (hydrogen
bonds, VdW forces, - interactions) and ii) the intrinsic flexibility of the framework itself,
induced by the existence of weak points within the skeleton. [5]
This chapter is focused on the crystal chemistry and properties of transition metal hybrid
vanadium oxides, which are halfway between zeolites and MOFs. The structural archetypes
of these materials are discussed in the function of the metal-organic and inorganic
substructure dimensionalities. The coordination of the metal centers, the use of chelating or
bridging ligands, the great diversity of vanadium oxide subunits, the flexibility or rigidity of
the ligands, the existence of solvent or coordinated species all give rise to a vast crystal
chemistry in which the crystal architectures and, hence, the physicochemical properties range
from those typical for zeolites to near those of the metal-organic frameworks (MOF),
exhibiting rigid or flexible structural response to the loss of solvent, in function of the
structural archetypes.

2. VANADIUM OXIDE CRYSTAL CHEMISTRY


The rich crystal chemistry of inorganic and hybrid vanadates is based on two major
reasons: i) the different coordination environments and oxidation states of vanadium, [6] and
ii) the great variety of different vanadium oxide subunits, such us cycles, chains, sheets. The
structural diversity of vanadium oxides and inorganic transition metal vanadates is greatly
increased with the introduction of inorganic components in the system. The task of this
section is describing the different vanadium oxide subunits. In the following sections, their
possible combinations with organic cations, metal-organic discrete units, or more complex
metal-organic nets, such us chains, layers or 3D substructures will be described.

2.1. Coordination Environment of Vanadium

Vanadium commonly adopts V(V), V(IV) or V(III) oxidation states. However, under the
hydrothermal conditions used for the synthesis of hybrid vanadates, mildly reducing
conditions are attained, and, consequently, only V(IV) and/or V(V) oxidation states are
observed.
The coordination environments, as well as the V-O bond distances, are clearly related to
the vanadium oxidation state. V(IV) exhibits five or six coordination environments, while
V(V) ranges from six to four coordination. Figure 1 presents the most common coordination
environments for the vanadium atom.
Taking into account the V-O bond distances, three kind of V-O bonds can be defined, the
vanadyl type or V=O terminal bond is one which has a short bond length in the range of 1.57-
1.68 ; it is a multiple bond with a -component arising from electron flow from O(p) to
V(d) orbitals. When there are two vanadyl bonds present in the same polyhedron, they are in
4 Edurne S. Larrea, Roberto Fernndez de Luis, Jos L. Mesa et al.

a cis arrangement. In five and six coordinated (V(IV)On) and (V(V)On) polyhedra, equatorial
bonds occur in a cis arrangement to the vanadyl bonds, and they are longer than the vanadyl
ones. In (V(IV)O6) and (V(V)O6) polyhedra, the sixth ligand is trans to a vanadyl bond. This
trans bond is usually longer than the equatorial bonds. The number of vanadyl, equatorial and
trans bonds defines the different coordination environments for the vanadium atoms.

Figure 1. Coordination environment of vanadium atoms for the oxidations states +3, +4 and +5. The
double line represents the vanadyl bonds, and the dash line trans longer bonds.

As it has been previously mentioned, for V(IV), two different coordination environments
are observed. The distorted octahedron, denoted as [1+4+1], containing four intermediate
equatorial bonds (1.86-2.16 ), one axial V=O terminal bond, and one long trans axial bond
(2.20-2.32 ). The five coordinated V(IV) is observed as [4+1] square pyramidal geometry,
that is, one short vanadyl bond and four longer equatorial bonds (1.80-2.10 ).
For V(V), four different coordination environments are observed. Four coordinated
regular tetrahedron: The statistical analysis of the CSD database [7] confirms that the V-O
bond distances vary depending on the oxygen ligand connectivity (Figure 2). Three kinds of
V-O bonds can be distinguished, the vanadyl bond (1.561.72 ), the V-O bonds with the
oxygen atom shared by the vanadium atom and a metal center (1.60 to 1.76 ), and the V-O
bonds with the oxygen ligand shared by two VO4 tetrahedron (1.74-1.88 ). Thus, a
tetrahedron sharing three vertices will exhibit one short and three longer bond distances,
while a tetrahedron sharing two vertices present two vanadyl short distances and two longer
bonds.
The five coordinated V(V) is present in [4+1] square pyramidal geometry, and [3+2]
trigonal bipyramid. The coordination geometry depends on the number of vanadyl bonds. The
square pyramidal geometry possesses one vanadyl bond (1.50-1.66 ) and four equatorial
ones (1.74-2.06 ). The trigonal bipyramid presents two vanadyl bonds (1.54-1.78 ),
occupying two equatorial positions, while the longer bonds (1.80-2.06 ) are located in one
equatorial and two axial positions.
Hybrid Vanadates, towards Metal-Organic Frameworks 5

Figure 2. VO bond distances histogram of four coordinated vanadium for hybrid vanadium oxides
(data retrieved from the CSD database). Reprinted from Ref. [40]. Copyright (2010), with permission
from RSC.

Finally, the six coordinated V(V) displays the [1+4+1] or [2+2+2] geometries, depending
on the existence of one or two vanadyl bonds, respectively. The [1+4+1] bond distribution is
similar to that of the tetravalent vanadium. The [2+2+2] geometry presents two short vanadyl
bonds in cis orientation, two long bonds (2.10-2.30 ), trans to the vanadyl ones, and two
intermediate bond lengths (1.85-2.05 ).

2.2. Vanadium Oxide (VxOy) Subunits

Vanadium polyhedron may fuse to provide different vanadium oxide subunits, such us
discrete oligomers (dimmers, cycles or clusters), 1D chains or 2D layers constructed from the
same polyhedra or mixtures of polyhedral types.
The most comprehensive study of inorganic structures is provided by A.F. Wells in the
book Structural Inorganic Chemistry, [8] but no specific information about vanadium oxides
is included. Later works of P.Y. Zavalij and M.S. Whittingham, [9] and J. Zubieta et al. [10]
are extensive studies of the structural chemistry of vanadium oxides with open framework
and the influence of organic components on vanadium oxide architectures, respectively. P.Y:
Zavalij and M.S. Whittingham deeply describe, and systematize the vanadium oxide subunits,
classifying them according to the coordination environment of vanadium polyhedron and the
connectivity between them. The incorporation of structure directing organic molecules or
secondary metal-organic subunits provides unique crystal architectures not seen in pure
vanadium oxides or inorganic vanadates. The role of the organic or metal-organic moiety is
determinant in the classification of hybrid vanadates proposed by J. Zubieta et al. However,
this work is mainly focused on the construction of hybrid vanadates from organic components
that usually acts chelating the secondary metal centers. Therefore, a generalized review
comprising the wide range of hybrid vanadates containing extensive metal-organic
substructures, and even a descriptive systematization of these architectures is urgent.
However, and in order to introduce the reader to this interesting crystal-chemistry world,
a brief review of the most common structural subunits is going to be made. The rich crystal
6 Edurne S. Larrea, Roberto Fernndez de Luis, Jos L. Mesa et al.

chemistry of the different vanadium oxides subunits arises, in part, from the linkage modes
flexibility of the various vanadium polyhedra through V-O-V bridging interaction.
As discrete cyclic units, four different cycles has been observed in the crystal chemistry
of vanadium, depending on the number of polyhedra that compose them, {V4O12}, {V5O15},
{V6O18}, {V12O36}.
The {V4O12} is the most common and versatile structural unit observed in hybrid
vanadates, due to the different possible linkages of the {V4O12} cycle with other vanadium
polyhedra or with different metal-organic substructures. In this respect, the {V4O12} cycles
could act as a two, four or six connectors between metal centers or other vanadium oxide
subunits.
Vanadium oxide subunits possess a certain degree of adaptability to the crystal
environment, due to their ability for reorientation and reorganization of the vanadium
polyhedra. Interesting examples are the metavanadate chains constructed from corner-sharing
VO4 tetrahedra. Figure 4 shows some metavanadate chains belonging to different vanadates.
The number of tetrahedra that, by translation, give rise to the whole chain increase
progressively from two, in Ba(VO3)2H2O, to twelve, in {Ni(en)3}(VO3)2.
The structural diversity of the four coordinate chains can be translated to the chains
constructed from edge-sharing five and six coordinated polyhedra (VO5, VO6). Figure 5
depicts the different vanadate chains constructed from five and six coordinated vanadium.
Some chains exist as isolated units in real structures, but in others, they are linked to each
other or to other vanadium subunits giving rise to layers or frameworks. The main difference
between the chains lies in the connectivity between the (VO5) or (VO6) polyhedra, but all of
them are sharing edges.

Figure 3. Vanadium oxide cycles observed in hybrid vanadates.

Figure 4. Metavanadate chains of corner-shared VO4 tetrahedra. (a) Ba(VO3)2H2O, [11] (b)
NH3(CH2)4NH3(VO3)2, [12] (c) M(Hdpa)V4O12 (M=Co, Ni), [13](d) Cu(2,2-Bpy)V2O6, [14] (e)
Co3(bpypr)4V6O182H2O, [15] (f) {Ni(en)3}(VO3)2. [16].
Hybrid Vanadates, towards Metal-Organic Frameworks 7

Figure 5. (a) (d) Chains of edge- sharing VO5 polyhedra. (e) (g) Chains of edge- sharing VO6
octahedra.

For each kind of five coordinated vanadium chains (Figure 4 (a)-(c)), the orientation of
the vanadyl bond (perpendicular to the equatorial ones) upwards or downwards the chain,
generates different subgroups of chains. Curiously, there are scarce examples of transition
metal hybrid vanadates containing six coordinated vanadium, being the tetrahedral and five
coordinated vanadium, or the combinations between them, more common. Probably, it is due
to the difficulty to obtain highly condensed vanadium oxide subunits combined when organic
molecules are present in the structure, because they promote the crystallization of more open
crystal architectures.
The structural diversity of inorganic-organic vanadates is not only based on the great
variety of vanadium oxide subunits, because these can link to each other giving rise to more
complex one, two- or three-dimensional architectures. There are several vanadates whose
crystal structure is constructed from the combination of previously described vanadium oxide
subunits. Examples of those are the chains constructed from corner-sharing {V6O18} cycles,
the 1D chains formed by the connectivity between {V4O12} cycles and {V2O7} dimmers,
sheets of corner-sharing (VO4) polyhedra or metavanadate chains combined with {V2O7}
dimmers.

Figure 6. Sheets of edge-shared VO5 polyhedra. (a) VO20.5H2O, [17] (b) V2O5, (c)
[VO2(Terpy)](V4O10), [18] (d) (NMe)4V4O10.
8 Edurne S. Larrea, Roberto Fernndez de Luis, Jos L. Mesa et al.

A very important group of vanadium subunits is the 2D vanadium oxides and vanadates,
whose sheets are described according to the previously defined structural subunits. The
interlayer spaces of this group of crystal structures could be occupied by water molecules,
several inorganic or organic cations, such as Li+, Na+, Mg2+, Ag+, tetramethylammonium,
ethylenediamonium, among others, or isolated metal-organic complexes. The crystal structure
of 2D vanadates is defined by the sheets morphology and the interlayer space. Examples of
these sheets are those constructed from corner- or edge-sharing (VO5) square pyramids
(Figure 6). These layers are defined by the disposition of the apical oxygen atoms upwards or
downwards the sheet, and the existence of ordered vacancies of (VO5) polyhedra.

3. HYDROTHERMAL SYNTHESIS OF HYBRID VANADATES


Usually, the hybrid vanadates are synthesized under hydrothermal and or solvothermal
conditions at low temperatures (<250C) and pressures (10-20 bar). The most relevant
chemical parameters of the synthesis are, the pH, the reagent concentrations, the reaction
time, the reactor filling factor or the reagents initial stoichiometry, among others.
The lower viscosity of water under hydrothermal conditions enhances the diffusion
processes, so that solvent extraction of solids and crystal growth from solution is favored.
This fact has allowed the use of organic molecules, which could act as structural directing
agents or being an integral part of the crystal framework as a ligand. However, under these
hydrothermal conditions, the dielectric properties of the solvent changes. This means that an
extrapolation of the conditions applied at room temperature is not valid yet. [19]
Studying the relationship between the hydrothermal synthesis conditions and the crystal
structures obtained is an essential task. From this point of view, the knowledge of
hydrothermal synthetic conditions-crystal structure relationship implies a feedback process,
because the careful analysis of the final outcome of the reaction, and concretely, the crystal
structure of the obtained compounds, give decisive information about what are the initial
synthesis parameters to be modified.
In that respect, the hydrothermal synthesis can be focused to the crystallization of similar
structural archetypes, modifying the geometry and/or length of the organic ligand, or selecting
different metal centers. In the specific case of vanadates, the systematic modification of the
initial conditions, may allow a partial control of vanadium species in solution, and hence, a
partial control of the final crystal framework in which they are included. The studies carried
out in mineral vanadates, reveal the importance of the pH value and temperature in the
polymerization degree of the resulting vanadium oxide subunits. [20] In a similar way, the
studies carried out with 51V NMR have allowed to detect the different species of V(V) in
solution at room temperature according to the pH and concentration. [21]
The condensation of discrete units, such as [HnVO4](3-n)-, present in solution at high pH
values, can be generated by the hydrolysis of V-OH groups. A progressive fall of the pH
value gives rise to pirovanadates, cycle-vanadates, metavanadates, or decavandates. The
stability fields of vanadium species in solution can be drastically influenced by the
introduction of organic molecules, the temperature and/or pressure related with the
hydrothermal synthesis conditions, the presence of different salts (nitrates, sulfates, chloride)
or organic ligands The piro, cycle and metavanadate species are common structural units in
Hybrid Vanadates, towards Metal-Organic Frameworks 9

hybrid vanadates crystal structures, and hence, these species must be present in solution
during the hydrothermal reaction.
The importance of the pH value in the polymerization degree of the vanadium oxide has
been confirmed in different hydrothermal systems. Chirayil et al. [22] studied the
tma/V2O5/LiOH (tma= tetramethylammonium) system keeping constant the temperature (185
C), the time of reaction, the initial stoichiometry and modifying the pH value with an acetic
acid solution. The obtained crystal structures show an increase in the coordination
environment of vanadium, and an increase in the polymerization degree of the VxOy
vanadium oxide subunit when the reactions conditions are more acidic.

3.1. High-Throughput (HT) Methods: Compositional Space Diagrams

The current interest of the applications of HT (High-throughput) concerns the research of


the parameter of solvothermal reactions in solid state science, the discovery of new
compounds, the optimization of reactions as well as the identification of reaction trends, and
particularly, the influence of pH and water content.
Nowadays, it is difficult to extrapolate any rigorous conclusion in the relationship
between the obtained crystal structures and the initial hydrothermal synthesis conditions.
Usually, hydrothermal reaction conditions reported in scientific literature correspond to the
best ones, with regard to the quality of the obtained crystal or the reaction efficiency.
However, the uns uccessful reactions also possess a lot of information that allows to suggest
some general trends of the crystallization process, specifically for hybrid vanadates to relate
partially to the obtained crystal architectures or vanadium oxide subunits with the initial pH
value, concentration, temperature, etc.
To report the variables involved in these hydrothermal reactions pH/concentration,
pH/stoichiometry, or in specific complex systems, the compositional space diagrams should
be developed. Compositional space diagrams are similar to ternary phase diagrams, in which
the products are directly related to initial reagent mole fractions, while other variables such as
temperature and amount of solvent are held constant. The outcome of the reactions are
carefully analyzed by polarizing light microscope, X-ray diffraction, IR spectroscopy,
chemical analyses, in order to characterize the different compound(s) obtained after the
hydrothermal reaction. The compositional space diagrams are not phase diagrams. Only the
crystalline products are analyzed, while products which are amorphous or remain in solution
are neglected.
As examples, the hydrothermal synthesis and possible crystallization pathways of three
different systems, (1) Ni/4Bpy/VxOy, (2) Ni/Bpe/VxOy, and (3) Ni/Bpa/VxOy will be discussed.
In order to obtain extended hybrid frameworks, we choose Ni2+ due to the octahedral
coordination preferences of this metal, and bidentate bridging ligands such as pyrazine (Pz),
4,4-bipyridine (4Bpy), 1,2-di(4-pyridyl)ethane (Bpa) or 1,2-di(4-pyridyl)ethylene (Bpe).
The selection of these ligands is based on different criteria; from 4Bpy to Bpe is an
increase of the ligand length, maintaining almost constant the ligand rigidity (Figure 7).
Additionally, the Bpe ligand could adopt two different orientational conformations, and the
10 Edurne S. Larrea, Roberto Fernndez de Luis, Jos L. Mesa et al.

interconversion between them can take place through pedal motion mechanism. [23] In
contrast, Bpa has almost the same length as Bpe but the ethane provides more flexibility for
the ligand, and hence for the crystal structure of the resulting phases. The Bpa ligand could
also adopt trans or gauche conformations.

Figure 7. General characteristics of ligands used in the Ni/4Bpy/V xOy, Ni/Bpe/VxOy and Ni/Bpa/VxOy
systems.

The final outcome of the hydrothermal synthesis of these three systems was carefully
analyzed and represented in compositional diagrams. Despite the difficulty to establish a
precise crystallization sequence, and concrete relationships between the obtained crystal
structures and the initial conditions, some general trends are observed.

3.1.1. Ni/4Bpy/ VxOy System (4Bpy = 4,4-Bipyridine)


The hydrothermal reactions of the Ni(NO3)26H2O/4Bpy/NaVO3 system were carried out
at 170C during three days. After the reaction time, the reaction containers were cooled at
room temperature.

Figure 8. Compositional space diagram for the Ni/4Bpy/V xOy system. Adapted from Ref [24].
Copyright (2008), with permission from Elsevier.
Hybrid Vanadates, towards Metal-Organic Frameworks 11

The initial stoichiometry of the reactants was fixed to 1:1:1 molar ratio, increasing
progressively the concentration of the system from 0.26 mmol/30 ml to 1.21 mmol/30 ml.
The reagents were mixed in 30 ml of water under vigorous stirring and the pH value was
adjusted with a 1M HNO3 solution. The final results are depicted in the concentration vs. pH
diagram shown in Figure 8.
Four different stability fields are distinguished. Compound (1),
[{Ni2(H2O)2(4Bpy)3}(V4O12)]2.5H2O, crystallizes as fiber like pale green single crystals.
The stability field is restrained at high concentrations and neutral or slightly acidic conditions.
The phase [{Ni8(4Bpy)16}(V24O68)]8.5H2O, (2), is obtained at neutral or slightly acidic
conditions at low concentrations. When the concentration of the system is increased, the
stability field of (2) is moved to lower pH values. The stability field of (2) is strongly
overlapped with that for compound (3) at high concentrations. In those conditions, in which
the stability fields are strongly overlapped, the single crystals of (2) present inclusions and a
dusty appearance.
The compound [Ni6(H2O)10(4Bpy)6}(V18O51)]0.5H2O, (3), crystallizes at acidic or
slightly acidic conditions as a yellow prismatic single crystal, with no inclusions observed in
the single crystals.
At lower pH values, low crystalline red powder crystallizes. The chemical analyses and
the IR spectrum confirm the presence of the organic ligand and nickel cation into the crystal
structure. The red color of the sample and several absorption bands of the IR spectrum
suggest the existence of five coordinated vanadium.
The strong overlapping of the stability fields for compounds (2) and (3) suggest a
sequential crystallization process. Probably, in a first step, the single crystals of (3) begin to
crystallize, promoting changes in the concentration or in the reagent stoichiometric ratio
which favors subsequent crystallization of (2). This sequential crystallization process,
explains why the single crystals of (2) presents inclusions, while those of (3) are not included.

Figure 9. Vanadium oxide subunits and metal-organic substructure for the compounds obtained in the
Ni/4Bpy/VxOy system. Adapted from Ref [24]. Copyright (2008), with permission from Elsevier.

Figure 9 depicted the vanadium oxide subunits and metal-organic substructure for the
three hybrid vanadates obtained in this system. For (1) and (2), the vanadium oxide subunits
possess four coordinated vanadium. For compound (3), a combination of five and four
12 Edurne S. Larrea, Roberto Fernndez de Luis, Jos L. Mesa et al.

coordination vanadium is observed. In an opposite way, the compounds obtained at neutral or


slightly acidic conditions present more complex metal-organic substructures, the
interpenetration of two rectangular metal-organic layers for (1) and a poly-catenation of a 3D
cadmium sulphate cds like topology with two square-like metal-organic layers for (2).
Compound (3) contains metal-organic chains.
The crystal structures of (1) and (2) are more open architectures, with crystallization
water molecules included in their channels, while (3) possesses a compact inorganic 3Dl
inorganic framework. This fact is reflected in the density values obtained by single crystal X-
ray diffraction, 1.86 and 1.67 grcm-3 for (1) and (2), respectively, and 2.11gcm-3 for (3).

3.1.2. Ni/Bpe/VxOy System Bpe = 1,2 Di(4-Pyridyl)Ethylene


When 4Bpy is replaced by Bpe, the hydrothermal conditions at which the hybrid
vanadates are obtained change. Then, the working temperature is 140 C, although the same
compounds can also be obtained at 120 and 100 C. The reduction of the reaction time from
three days to one day does not drastically influence the outcome of the reaction. Several tests
have been carried out at four hours of reaction time and the final compounds are also
observed, but with an appreciable decrease of the crystallinity.
Figure 10 (a) shows the initial stoichiometry vs. pH value diagrams for a concentration of
0.26 mmol/30ml. The hydrothermal reactions were carried out with 30 ml of water at 140C
during three days. Two stability fields are distinguished. Compound {Ni(Bpe)}(VO3)2 (1)
crystallizes in all of the studied conditions. However, at acidic conditions, or when increasing
the NaVO3 molar ratio in the initial reactants, the co-crystallization of compound
[{Ni(H2O)2(Bpe)}(V4O11)]0.5H2O (2) is observed. The increase of the global concentration
(Figure 10 (b)) displaces the stability field of (2) to less acidic conditions. In the same way as
in the previously described system, the reduction of the pH value favors the presence of five
coordinated vanadium present in the crystal structure of compound (2). [24]

Figure 10. pH value vs. initial stoichiometry compositional space diagrams for the Ni/Bpe/VxOy system.
Adapted from Ref [24]. Copyright (2008), with permission from Elsevier.

Compounds (1) and (2) crystallize as emerald block-like single crystals and red prism-
like shape single crystals, respectively. The single crystals of (1), obtained in the conditions at
which both compounds crystallize, possess inclusions of the red single crystal of (2) (Figure
Hybrid Vanadates, towards Metal-Organic Frameworks 13

11(a)). By contrast, the single crystals of (2) obtained at the same conditions are completely
clean.

Figure 11. (a) Single crystals of (2) included into the single crystals of {Ni(Bpe)}(VO3)2 (1). Single
crystals of [{Ni(H2O)2(Bpe)}(V4O11)]0.5H2O (2).

This fact suggests that when the pH is acidic or the NaVO3 molar ratio is high enough the
crystallization of (2) is favored. This compound has a Ni:V ratio of 1:4 in its formula, so the
crystallization of compound (2) produces a progressive decrease of the vanadium
concentration in the solution. At a certain point of the hydrothermal reaction, compound (1)
begins to crystallize, including the single crystals of (2) into it.
As occurs in the Ni/4Bpy/VxOy system, the crystal structure of (1), obtained at neutral or
slightly acidic conditions, possesses a more complex metal-organic substructure, which
consists of the inclined interpenetration of rhombic-like sheets, instead of the metal-organic
chains observed for the compound (2) (Figure 12). For the vanadium oxide subunits, the same
trend is also observed, with more condensed vanadium oxide subunit for the compound (2),
obtained at more acidic conditions or at higher rates of vanadium in the initial stoichiometry.

Figure 12. Vanadium oxide subunit and metal-organic substructures observed in the Ni/Bpe/VxOy
system. Adapted from Ref [24]. Copyright (2008), with permission from Elsevier.

3.1.3. Ni/Bpa/VxOy System Bpa = 1,2-Di(4-Pyridyl)Ethane


The compositional space diagrams for Ni/Bpa/VxOy system are shown in Figure 13. The
[{Ni(H2O)(Bpa)}(VO3)2]2H2O, (1), {Ni(H2O)2(Bpa)2}(VO3)2]2H2O, (2),
[{Ni3(H2O)3(Bpa)4}(V6O18)]8H2O, (3), and {Ni2(H2O)2(Bpa)2}(V6O17), (4), compounds are
14 Edurne S. Larrea, Roberto Fernndez de Luis, Jos L. Mesa et al.

obtained at different mole ratios of Ni(NO3)26H2O, NaVO3, and Bpa at 120C. Compounds
(1) and (3) crystallize as rhombic prisms and plate-like single crystals, suitable for the crystal
structure determination by X-ray diffraction. Compound (2) is obtained as a low crystalline
pale green powder. Finally, the compound (4) crystallizes as yellow hexagonal prism single
crystals. The chemical formula of (2) was obtained based on chemical analysis, IR spectra
and TG measurements.

Figure 13. Compositional space diagrams for the Ni/Bpa/VxOy system.

Table 1 summarizes the Ni:V and Ni:Bpa ratios used to obtain the different compounds
observed in the Ni/Bpa/VxOy system.

Table 1. Ni:V and Ni:Bpa ratio for the compounds obtained in the Ni/Bpa/VxOy system

Compound Ni : V Ni : Bpa
[{Ni(H2O)(Bpa)}(VO3)2]2H2O (1) 1:2 1:1
[{Ni(H2O)2(Bpa)2}(VO3)2]2H2O (2) 1:2 1:2
[{Ni3(H2O)3(Bpa)4}(V6O18)]8H2O (3) 1:2 3:4
{Ni2(H2O)2(Bpa)2}(V6O17) (4) 1:3 1:1

First of all, the compositional space diagram constructed from hydrothermal reactions
with 0.256 mmol of Ni(NO3)26H2O, and pH=6.5 (Figure 13 (a)) is going to be discussed.
This compositional space diagram for the Ni/Bpa/VxOy system shows three major
crystallization fields. From the bottom right corner to the center of the space diagram, the
crystallization field related with compound (4) is located (yellow area in Figure 13). This
compound crystallizes from initial NaVO3:Bpa molar ratios equal or higher than 1:1, and the
crystallization field is strongly overlapped with those of compounds (2) and (3).
The crystallization of compound (4), which has the higher Ni/V ratio in formula of this
system, is favored by NaVO3/:Bpa molar ratios equal or higher than 1:1 ratio in the initial
stoichiometry. The crystallization field related to the compound (2), is located from the top to
the center of the compositional space diagram, and hence involves NaVO3:Bpa initial ratios
equal or lower than 1:1. The higher concentration of organic ligand in the hydrothermal
reaction favors the crystallization of the compound (2). Finally, at NaVO3/Bpa ratios equal or
Hybrid Vanadates, towards Metal-Organic Frameworks 15

close to 1:1, the hydrothermal reactions always give rise to mixtures of (2) + (3) or (2) + (3) +
(4). Hence, the crystallization field of compound (3) is completely overlapped with the field
of compound (2), and partially overlapped with that of compound (4).
The crystallization field of compound (1) can not be clearly established, because this
phase is always obtained as a minor product of the hydrothermal reaction. We have marked
into the compositional diagram the different conditions in which this minor phase is observed.
The increase of the concentration (Figure 13 (b)) does not drastically affect the crystallization
fields of the compositional space. The crystallization field of compound (4) is slightly
displaced to the top of the compositional diagram.
One of the fundamental questions about any system with two or more products has to do
with the identification of the factors which control the crystallization of one compound over
another. Taking into account the overlapping of the crystallization fields, and the Ni/V and
Ni/Bpa ratios in the compounds formula, we can suggest a tentative crystallization sequence
shown in Figure 14. Five different crystallization processes, in function of the NaVO3:Bpa
initial ratio, and its evolution during the hydrothermal reaction are depicted. This is a
qualitative explanation of the studied hydrothermal reactions. Probably, there are several
factors that could influence in the final outcome. However, the compositional diagrams
suggest that one of the most important factors is the NaVO3:Bpa initial ratio.
Pathway (A): For an initial NaVO3/Bpa ratio equal to 3:1, compound (4) is obtained. We
have simulated the evolution of the NaVO3/Bpa ratio if it is taken into account that the
crystallization of one mmol of (4) reduce 6 mmol of vanadium(V) and 2 mmol of Bpa in
solution, hence the NaVO3/Bpa ratio is maintained constant in solution, and the crystallization
of (4) is progressively stabilized during the course of the reaction. The final outcome of the
hydrothermal reaction is the compound (4) as an isolated phase.
Pathway (B): For an initial NaVO3/Bpa = 2:1 ratio, the final product of the reaction is
also compound (4). The NaVO3/Bpa ratio is progressively reduced during the course of the
hydrothermal reaction, but not enough to give rise to the crystallization of compound (3).

Figure 14. Tentative crystallization sequence for the Ni/Bpa/VxOy system.


16 Edurne S. Larrea, Roberto Fernndez de Luis, Jos L. Mesa et al.

Pathway (C): For an initial NaVO3/Bpa ratio equal to 1.5:1, the compositional diagrams
show that a mixture of (4) + (3) + (2) is obtained after the hydrothermal synthesis. Compound
(4) begins to crystallize and the NaVO3/Bpa ratio in solution is progressively reduced. For a
NaVO3/Bpa ratio equal or near equal to one the compositional diagrams, it is shown that
compounds (3) and (2) are obtained. So probably, during the course of the hydrothermal
reaction, when the NaVO3/Bpa ratio reach a value near to 1 the crystallization of compound
(3) should be favored. This compound has a V/Bpa = 6/4 ratio in its formula, and its
crystallization also reduce progressively the NaVO3/Bpa ratio in solution. In a third step, the
NaVO3/Bpa ratio gets values near 1:2. The compositional diagrams show that for this
NaVO3/Bpa ratio, compound (2) is obtained, and therefore, at this point of the hydrothermal
reaction, the crystallization of (2) should start.
Pathway (D): The increase of Bpa molar rate with regard to NaVO3 in the initial
stoichiometry favors the crystallization of (3) instead of (4). The NaVO3/Bpa ratio in solution
is progressively reduced during the hydrothermal reaction, and in a second step, compound
(2) begin to crystallize. This compound possesses 1 vanadium(V) and 1 Bpa molecule per
formula, and hence, the NaVO3/Bpa ratio in solution is held constant during the reaction.
Pathway (E): For initial NaVO3/Bpa equal or lower than 1:2, only compound (2) is
obtained. The NaVO3/Bpa ratio is constant during the course of the reaction and then, only
compound (2) is obtained.
Some general trends are observed in the system; the increase of the organic ligand ratio in
the initial stoichiometry favors the crystallization of compounds with a higher Ni:Bpa ratio.
Unfortunately, the crystal structure of (3) cannot be solved, in order to establish a complete
comparison between the compounds present in the system.

4. STRUCTURAL ARQUETYPES: FROM ZEOTYPES TO MOFS


The synergetic interaction between the organic and inorganic components allows a partial
degree of crystal engineering by exploiting fundamental aspects of the structural
organization:

(i) Coordination preferences and oxidation states of the metal centers.


(ii) The length of the spacer, geometry, degree of flexibility and relative position of the
ligand donor groups.
(iii) The structural influence of the anion on the framework organization and assembly
adopted by the different metalorganic moieties.

Bipyridyl dipodal ligands have been widely used since they have two nitrogen atoms that
can connect metal atoms to form polymeric compounds. These polytopic organicamino
ligands serve to link the metal sites propagating the structural information expressed by the
metal coordination preferences through the extended structure. The length, geometry and
position of the nitrogen donor atoms of the ligand determine the role of the ligand in the
construction of the final crystal structure. Chelating ligands, such as 2,2-bipyridine, 1,10-
phenanthroline or terpyridine, gives rise to 1D or 2D architectures. On the other hand, the
choice of dipodal or multi-podal ligands such as 4,4-bipyridine, pyrazine, or 1,2-di(4-
Hybrid Vanadates, towards Metal-Organic Frameworks 17

pyridyl)ethylene allows the possibility to obtaining extended metal-organic nets and high
dimensionality crystal structures. Usually, the vanadium oxide subunits are located in the
channels of the metal-organic substructure, and linked to the metal centers sharing vertexes.
The combination of metal-organic nets based on nitrogen donor dipodal ligands and
vanadium oxide subunits, has become a great chemical strategy to synthesize self-catenated
and/or highly connected nets. [25,26] Despite the difficulties to control the final crystal
architectures in hybrid vanadates, the analysis of the synergetic interaction between the
inorganic and organic components allows a partial degree of cr ystal engineering, towards
more open or condensed structural archetypes, and to a more rigid or flexible response of the
crystal structure to the loss or uptake of the solvent.
During the last decades, a great number of transition metal hybrid vanadates have been
synthesized. The works of P.Y. Zavalij and M.S.Whittingham and J. Zubieta et. al. [9,10]
summarizes the crystal chemistry of these materials. They propose two different but
complementary classifications. P.Y. Zavalij and M.S. Whittingham based their classification
of the open framework vanadium oxides on the coordination environment of vanadium atom.
Seven classes can be distinguished according to the presence of tetrahedra (T), square
pyramid (SP) and octahedra (O) into the crystal structures.
As a complementary point of view, J. Zubieta et al. take into account the role of the
metal-organic or organic moiety into the crystal structure, distinguishing four different classes
of hybrid vanadates:

a) Compounds in which the organic component or a metal-organic discrete coordination


complex acts as an isolated cation, compensating the charge of the vanadium oxide
host.
b) Phases in which the organic molecule serve as a ligand directly attached to the oxide
array.
c) Transition metal hybrid vanadates in which the organic ligand is coordinated to the
transition metal and this secondary metal linked to the vanadium oxide subunits
through oxo-bridging groups. The organic ligand coordinates only one metal center,
preventing the propagation of the crystal structure through the organic ligand.
d) Hybrid vanadates constructed from extended metal-organic nets (one-, two- or three-
dimensional) linked through the metal center to VxOy vanadium oxide subunits.

The discussion of the crystal architectures will be focused in the (d) group, because since
the work of J. Zubieta et al., a great number of transition hybrid vanadium oxides containing
polymeric metal-organic substructures have been reported, being necessary to do a specific
overview of this interesting subclass of vanadates. Despite the hybrid vanadium oxides
constructed from chelating ligands are deeply described in the literature, several common
structural characteristics for the (c) and (d) groups are going to be pointed out during the
description of the different crystal structures.
The geometry and position of the nitrogen atoms in the organic ligands allows obtaining
extended metal-organic substructures in the hybrid vanadates crystal structures. In that regard,
the transition hybrid vanadates constructed from dipodal ligands has common characteristics
with MOFs.
The dimensionality and topology of the metal-organic substructure depends on: i)
coordination environment of the metal center that acts as the node of the different metal-
18 Edurne S. Larrea, Roberto Fernndez de Luis, Jos L. Mesa et al.

organic substructures, and ii) the connectivity between the metal centers through the organic
ligand. In the same way that in coordination polymers, the one-, two- or three-dimensional
metal-organic nets, could be interpenetrated or polycatenated, increasing the complexity of
the crystal framework.
During the structural description, several terms concerning to the topology, such as,
interpenetration or polycatenation of the metal-organic subnets, will be used. More
information about topology can be found in the works of M. OKeeffe. et al. [27] In the same
way, there are several systematic studies about the interpenetration, polycatenation,
polythreading and polyknotting of the nets in the works of S. Batten, [28] and D.M. Proserpio
et al. [29]

Figure 15. Structural archetypes in hybrid vanadates.


Hybrid Vanadates, towards Metal-Organic Frameworks 19

The introduction of vanadium oxide subunits in conjunction with the different metal-
organic subnet, greatly increases the complexity of the resulting crystal architectures. Usually,
the vanadium oxide subunits share oxygen atoms with the metal centers, giving rise to an
inorganic scaffold within the structure. The combination of the metal-organic subnets and
vanadium oxide subunits usually increase the dimensionality of the crystal structure, and
could give rise to phenomena such as self-catenated or highly connected nets.
One of the most remarkable characteristics of transition metal hybrid vanadates is that the
crystal structures could combine both 3D inorganic and metal-organic substructures.
Moreover, the metal-organic subnet could be interpenetrated or polycatenated. During the
description of the crystal structures, they are going to be classified according to the metal-
organic substructure, the inorganic substructure and the overall crystal structure
dimensionalities. For example, a compound denoted as 1D + 2D 3D possesses a 1D
inorganic substructure generated by the connectivity between the metal centers and the
vanadium oxide subunit and 2D metal-organic layers, being the whole crystal structure 3D. In
the same way, when an interpenetration or polycatenation of the metal-organic subnets is
observed, this will be denoted, for example, as (2D+1D), that would be a crystal structure that
possess metal-organic layers polycatenated with metal-organic chains.
The systematic analysis of the hybrid vanadates of the literature, according to their crystal
structures, made necessary a classification which includes all of the architectures observed.
We have classified the hybrid vanadate structures according to the dimensionalities of the
inorganic and the metal-organic subtructures. The inorganic substructure is defined by the
linkage between the vanadium oxide and the secondary metal centers, while the metal-organic
substructure is built up from the metal centers and the organic ligand.
Taking into account the dimensionality of the inorganic substructure, four major classes
of hybrid vanadates structures are described, A, B, C, and D, which respectively have 3D, 2D,
1D, and 0D inorganic substructures. The dimensionality of the metal-organic substructure is
denoted with the codes, 0, 1, 2, or 3, after the corresponding letter, for 0D, 1D, 2D or 3D
metal-organic substructures. Therefore, a crystal structure exhibiting inorganic layers and
metal-organic chains is described as a B1 architecture. In Figure 15, a schematic
representation of this classification is shown.
Figures 16 and 17 shows the most common dipodal and multi-podal organic ligand used
in the hydrothermal synthesis of the hybrid vanadates belonging to (d) class.

Figure 16. The most common dipodal organic ligands observed in the class (d) hybrid vanadates: Pz
(pyrazine), 4Bpy (4,4-bipyridine), Bpe (1,2-di(4-pyridyl)ethylene), Bpa (1,2-di(4-pyridyl)ethane), Bbp
(1,3-bis(4-pyridyl)propane, Bbi (1,4-Bis(N-imidazolyl)butane), Btp (1,4-bis(Triazol-1-
ylmethyl)benzene).
20 Edurne S. Larrea, Roberto Fernndez de Luis, Jos L. Mesa et al.

Figure 17. The most common multipodal organic ligands observed in the class (d) hybrid vanadates:
Trz: (Triazolate), Ptrz: (5-(Pyrimidin 2yl)tetrazolato, Btpt: bis(2,4,6-tris(pyrid-4yl)-1,3,5-triazine-N),
Tpp: (2,3,5,6-tetrakis(2-pyridyl)pyrazine), Bpq: m2-6`,6``-bis(2-pyridyl)-2,2`:4`,4``:2``,2```-
quaterpyridyl).

In Tables 2 to 6, most of the hybrid vanadates constructed from dipodal and multi-podal
organic ligands are included. The vanadium oxide subunit and metal-organic substructure are
briefly described and the crystal structure is classified according to the inorganic and metal-
organic substructure dimensionalities. In a parallel way, each compound is classified with a
code described in the classification proposed below in this section and sumarized in Figure
15. The CSD Cam bridge Structural Database reference codes for the compounds are also
shown when available.

Table 2. Hybrid vanadium oxides constructed from pyrazine

Compound Vanadium oxide subunit Metal-organic subnet Class. Ref.


Pyrazine (Pz)
POCNIF [30]
{M(Pz)}(VO3)2 Metal-Pz crossed chains A1
Metavanadate chains QOXFEO
M= Ni, Co (1D) 3D + 1D3D
[31]
Vanadium oxide layers XEHXUE
Metal-Pz chains parallel
{M(Pz)}(V4O10), constructed from corner- A1 [32]
to the vanadium oxide
M=Ni, Co, Zn shared double metavanadate 3D + 1D3D XEHYAL
layers (1D)
chains [32]
{Ni2(H2O)2(Pz)}(V4 Ni2NxOx(H2O)X-Pz B1 GICDEC
{V4O12} cycles
O12) chains (1D) 2D + 1D3D [33,34]
Metavanadate chains of
corner-sharing VO4 Ni-Pz, trans, trans, cis A1 GICDAY
{Ni3(Pz)3}(V8O23)
tetrahedra, linked through chains (1D) 3D + 1D3D [33]
V2O7 dimmers
Vanadium oxide layers
Cu(Pz)2 square like
[{Cu(Pz)2}(V6O16)] constructed from corner- A2
metal-organic layers IJUCIZ [35]
0.22H2O sharing double 3D + 2D3D
(2D)
metavanadate chains
{Cu2(Pz)} discrete units A0 MUJSAL
{Cu2(Pz)}(V4O12) {V4O12} cycles
(0D) 3D + 0D3D [36]
Hybrid Vanadates, towards Metal-Organic Frameworks 21

Table 3. Hybrid vanadium oxides constructed from 4,4-bipyridine

Compound Vanadium oxide subunit Metal-organic subnet Class. Ref.


4,4-Bipyridine (4Bpy)
Interpenetrated
[{M2(H2O)2(4Bpy)3}(VO Metavanadate chains of A2
{M2(H2O)2(4Bpy)3} square OCAQOY [37]
3)4]2.5H2O corner-sharing VO4 3D + (2D + 2D)
like metal-organic layers FIYDOH [38]
M= Ni, Co tetrahedra 3D
(2D+2D)
Metavanadate chains of {Ni(H2O)2(4Bpy)} and
{Ni6(H2O)10(4Bpy)6}(V18 edge-sharing VO5 {Ni(H2O)(4Bpy)} crossed A1
OHUQUE [39]
O51)]1.5 H2O polyhedra, corner linked to metal-organic chains 3D+ 1D3D
V2O7 dimmers. (1D)
Polycatenation of
{Ni(4Bpy)2} cadmium
V5O15 cycles corner linked A3
sulfate like 3D metal-
[{Ni8(4Bpy)16}(V24O68)] thought single VO4 3D +
organic subnet with two DUXKEN [40]
8.5H2O tetrahedra, generating (2D+2D+3D)
square like {Ni(4Bpy)2}
chains 3D
metal-organic layers.
(3D+2D+2D)
Co(4Bpy) parallel metal-
B1
{Co(4Bpy)}(VO3)2 {V4O12} cycles organic chains QOXFAK [41]
2D + 1D3D
(1D)
[{Ag(4Bpy)}4(V4O12)]2 {Ag(4Bpy)} parallel chains B1
{V4O12} cycles EGOGUD [42]
H2O (1D) 2D + 1D3D
Vanadium oxide layers
{Metal(4Bpy)}parallel
{M(4Bpy)}(V4O10) constructed from corner- A1 WIMHUW [43
chains
M= Cu, Ag sharing double 3D + 1D3D WIMJAE [43]
(1D)
metavanadate chains
{Cu2(4Bpy)2} double
[{Cu(4Bpy)}4(V4O12)]2 C1
{V4O12} cycles chains PAVDAS [44]
H2O 1D + 1D3D
(1D)
Metavanadate chains of
[{Mn(4Bpy)}(VO3)2]1.1 {Mn(4Bpy)} layers B2
edge-sharing VO5 [45]
6H2O (2D) 2D+2D 3D
polyhedra
Metavanadate chains of {Mn(4Bpy)0.5} discrete
[{Mn(4Bpy)0.5}(VO3)2]0. B2
edge-sharing VO5 layers [45]
62H2O 2D + 2D 3D
polyhedra (2D)

Table 4. Hybrid vanadium oxides constructed from Bpe


(1,2-di(4-pyridyl)ethylene) and Bpa (1,2-di(4-pyridyl)ethane

Compound Vanadium oxide subunit Metal-organic subnet Class. Ref.


1,2-Di(4-pyridyl)ethylene (Bpe)
Interpenetrated {Ni(Bpe)}
Metavanadate chains of B2
rhombic like metal-organic TIGHOH
{Ni(Bpe)}(VO3)2 corner-sharing VO4 2D + (2D + 2D)
layers [46]
tetrahedra 3D
(2D+2D)
Two corner sharing
metavanadate chains,
generating a double Crossed {Ni(H2O)2(Bpe)}
[{Ni(H2O)2(Bpe)}(V4O11) A1 OHURAL
metavanadate chain. The metal-organic chains
]0.5 H2O 3D + 1D 3D [39]
single chains are (1D)
constructed from edge-
sharingVO5 polyhedra
DIWXUD
{Co(HBpe)2} discrete units C0
{Co(HBpe)2}(V4O12) {V4O12} cycles [47]
(0D) 1D + 0D 1D
22 Edurne S. Larrea, Roberto Fernndez de Luis, Jos L. Mesa et al.

Table 4. (Continued).

Compound Vanadium oxide subunit Metal-organic subnet Class. Ref.


1,2-Di(4-pyridyl)ethane (Bpa)
{Ag(Bpa)} chains corner
[{Ag(Bpa)}4(V4O12)]4H2 linked giving rise to metal- B2 EGOHAK
{V4O12} cycles
O organic layers 2D + 2D 3D [42]
(2D)
[{Ni(H2O)(Bpa)}(VO3)2] {Ni(H2O)(Bpa)} chains B1
{V4O12} cycles [48]
2H2O (1D) 2D + 1D 3D
Three interpenetrated A3
[{Ni3(H2O)3(Bpa)4}(V6O1
{V6O18} cycles mot like 3D nets 3D + (3D +3D + [48]
8)]8H2O
(3D+3D+3D) 3D) 3D
{V4O12} cycles linked to Ni(H2O)(Bpa) parallel
{Ni2(H2O)2(Bpa)2}(V6O17 A1
{V2O7} dimmers generatingchains [48]
) 3D + 1D 3D
1D chains (1D)
{Co(Bpa)}parallel chains B1
{Co(Bpa)}(VO3)2 {V4O12} cycles [48]
(1D) 2D + 1D 3D

Table 5. Hybrid vanadium oxides constructed from Bbp (1,3-bis(4-pyridyl)propane, Bbi


(1,4-bis(Imidazol-1yl)butane) and Btp (1,4-bis(Triazol-1-ylmethyl)benzene)

Compound Vanadium oxide subunit Metal-organic subnet Class. Ref.


{Cd(HBbi)2} discrete
C0
{Cd(HBbi)2}(V4O12) {V4O12} cycles units AHAXOX [49]
1D + 0D 1D
(0D)
Metavanadate chains of Parallel corrugated
A1
{Co(H2O)(Bbi)}(VO3)2 corner-sharing VO4 {Co(H2O)(Bbi)} chains TOHQAJ [50]
3D + 1D 3D
tetrahedra (1D)
Metavanadate chains of Parallel VUTLAY
{M2(H2O)2(Bbi)}(VO3)4 B1
corner sharing VO4 {M2(H2O)2(Bbi)} chains VUTLEC
M = Co, Mn 2D + 1D 3D
tetrahedra (1D) [51]
Two 3D {M2(Bbi)3}
interpenetrated metal- VIRYOL
[{M2(Bbi)3}(V4O12)]4H2
organic subnets. The C3 [52]
O {V4O12} cycles
metal nodes are 1D + (3D+3D) 3D VUTLIG
M= Ni, Co
(MN6O4) dimmers [51]
(3D+3D)
Metavanadate chains of Perpendicular TOHPUC
{Ni2(H2O)4(Bbi)2}(VO3)4 A1
corner-sharing VO4 {Ni(H2O)2(Bbi)} chains XOJCUV
)]2H2O 3D + 1D 3D
tetrahedra (1D) [50]
[{Cu(Bbi)2}(V4O12)]4H2 Cu(Bbi)2 layers D2 NUXXUA
{V4O12} cycles
O (2D) 0D + 2D 2D [53]
{Cu(Bbi)2} layers (2D) C2 VIRYIF
{Cu2(Bbi)3}(V4O12) {V4O12} cycles
+ {Cu(Bbi)} (1D) chains 1D + (2D+1D) 2D [52]
A2 SAZSOC [54]
{M2(Bbp)4}(V4O12) Two interpenetrated
{V4O12} cycles 3D + (2D + 2D) DIXNII
M= Co, Ni M(Bbp)2 layers (2D+2D)
3D [55]
A1
Metavanadate chains of Perpendicular Zn(Bbp) HISWUC
{Zn(Bbp)}(VO3)2 3D + (1D+1D+1D)
corner-sharing VO4 chains (1D) [56]
3D
{Cu(Bbp)} metal-
[{Cu4(Bbp)4}(V4O12)]3H D3 LIGPUN
{V4O12} cycles organic chains
2O 0D + 1D 3D [57]
(1D)
Polycatenated
{V8O23} units A2
{Co2(H2O)(Btp)3} and MUTZEH
{Co3(H2O)(Btp)5}(V8O23) constructed from two 3D + (2D + 2D)
{Co(Btp)2} layers [58]
{V5O15} fussed rings 3D
(2D+2D)
Hybrid Vanadates, towards Metal-Organic Frameworks 23

Table 6. Hybrid vanadium oxides constructed from Trz: (Triazolate), Ptrz:


(5-(Pyrimidin 2yl)tetrazolate, Btpt: bis(2,4,6-tris(pyrid-4yl)-1,3,5-triazine-N),
Tpp: (2,3,5,6-tetrakis(2-pyridyl)pyrazine) and Bpq: (m2-6`,6``-bis(2-pyridyl)-
2,2`:4`,4``:2``,2```-quaterpyridyl)

Compound Vanadium oxide subunit Metal-organic subnet Class. Ref.


[{Cu2(Bpq)}(V4O12)]2 {Cu2(Bpq)} discrete C0
{V4O12} cycles TIJHAW [59]
H2O units (0D) 1D +0D 2D
{Zn2(Btpt)} discrete C0 ATAGUX
{Zn2(Btpt)}(V4O12) {V4O12} cycles
units (0D) 1D +0D 2D [60]
{Co2(Btpt)2}discrete C0 ATAGEH
{Co2(Btpt)2}(V4O12) {V4O12} cycles
units (0D) 1D + 0D 1D [60]
{Cu2(Btpt)2} discrete D0 MUJSEP
{Cu2(Btpt)2}(V4O12) {V4O12} cycles
units (0D) 0D + 0D 0D [61]
{Cu4(Btpt)2} Chains D1 ATAGOR
{Cu4(Btpt)2}(V4O12) {V4O12} cycles
(1D) 0D + 1D 2D [60]
Metavanadate chains of {Zn(Tpp)2} discrete B0 EHODEK
{Zn(Tpp)2}(VO3)2
corner-sharing VO4 units (0D) 2D + 0D 2D [62]
Two parallel
A2
[{Cu3(H2O)2(Tpp)2}(V Metavanadate chains of interpenetrated EHODOU
3D + (2D + 2D)
8O23)]3H2O corner-sharing VO4 {Cu3(H2O)2(Tpp)2} [62]
3D
layers (2D+2D)
Parallel and corrugated A2
[{Zn3(H2O)2(Tpp)2}(V Metavanadate chains of EHODIO
{Zn3(H2O)2(Tpp)2} 3D + (2D+2D)
O3)6]6H2O corner-sharingVO4 [62]
layers (2D+2D) 3D
B1 MUJSIT
{Cu3(Trz)2}(V4O12) {V4O12} cycles Cu3(Trz)2 chains (1D)
2D + 1D 3D [61]
Metavanadate chains of C2 CIXPAD
{Cu3(Ptrz)2}(VO3)4 Cu3(Ptrz)2 layers (2D)
corner-sharing VO4 1D + 2D 3D [63]

4.1. A Architectures: Hybrid Vanadates Archetypes Containing Three-


Dimensional Inorganic Scaffolds

The A type crystal architectures are characterized by three-dimensional inorganic


substructures constructed from the corner linkages between the secondary metal centers and
the vanadium oxide subunits. According to the metal-organic substructure dimensionality,
four subclasses of A crystal structures can be distinguished: A0, A1, A2 and A3 architectures.
Among the A0 architectures, there are two different types of structures which are denoted
A0a and A0b. A0a is a special type of hybrid vanadate in which the organic molecule is not
bonded to the inorganic substructure, so, in fact, the metal-organic substructure, does not
exist. In this type of structures, the organic molecule is protonated and acts compensating the
charge of the inorganic skeleton. These crystal structures are very similar to those observed
for aluminophosphates or transition metal phosphates, arsenates, phosphites, germanates, etc.
The vast crystal chemistry of A0a materials has been deeply described during the last
decades.[1] Few examples of hybrid vanadates belonging to this group are known, such as
{Mn3(en)}(V4O16), [64] and [(N(CH3)4]2[Co(H2O)4(V12O28)], [65]
[{Ni(enMe)2}0.5(V6O14)](H 2enMe)0.5. [66]
In the A0b crystal structures, the organic molecule acts as a chelating ligand. This means
that the ligand is bonded to the metal center by more than one donor atom, and, therefore, the
dimensionality of the metal-organic substructure is 0D. In [{Ni(Ht)}(VO3)2]0.33H2O,
24 Edurne S. Larrea, Roberto Fernndez de Luis, Jos L. Mesa et al.

[{Cu(Ht)}(VO3)2]0.33H2O, (Ht= 5,5,7,12,12,14-hexamethyl-1,4,8,11-


tetraazacyclotetradecane) [67] the equatorial plane of the nickel(II) and copper(II) atoms
consist of four nitrogen atoms belonging to the Ht organic ligand. The axial oxygen atoms of
the secondary metal center are shared with the metavanadate chains, generating porous 3D
frameworks, whose channels are filled by crystallization water molecules (Figure 18).
Another example of A0b architecture is the {Cu2(Pz)}(V4O12) [36] hybrid vanadate, in which
the two copper(II) atoms are linked through the Pz ligand generating discrete metal-organic
units.

Figure 18. Crystal structure of [{Ni(Ht)}(VO3)2]0.33H2O. [67].

When the ligand acts as a bridge between two or more metal centers, the A1, A2 and A3
crystal architectures are generated. The linkage between the metal centers through the organic
ligand generates metal-organic chains for the A1 crystal architectures, metal-organic layers in
A2-type crystal structures, and 3D metal-organic nets in A3 archetypes.
Among the A1 crystal structures, four different subtypes can be distinguished, focusing
on the disposition of the metal-organic chains with respect to each other and to the vanadate
substructure. The metal-organic chains could present a crossed disposition to each other, and
be perpendicular to the vanadium oxide subunit, as in {M(Pz)}(VO3)2 M= Ni, Co,
[{Ni6(H2O)10(4Bpy)6}(V18O51)]1.5 H2O, and [{Ni(H2O)2(Bpe)}(V4O11)]0.5 H2O hybrid
vanadates, shown in Figure 19. [30,3139]. The {Ni2(H2O)4(Bbi)2}(VO3)4)]2H2O, [50]
compounds exhibit a similar arrangement, however, the metal-organic chains are undulated
due to the flexibility of Bbi ligand. In other structures, a parallel disposition of the metal-
organic chains to each other, being perpendicular to the vanadium oxide chains, is observed.
In the specific case of {Ni3(pz)3}(V8O23) [33] hybrid vanadate, the metal-organic chains
possess a trans-trans-cis connectivity, instead the linear linkage observed for the previously
described compounds. The same structural arrangement is observed for
{Co(H2O)(Bbi)}(VO3)2 [50] hybrid vanadate, with {Co(Bbi)} undulated metal-organic
chains. An intermediate structure between the previously described archetypes is the
{Zn(Bbp)}(VO)3 [56] compound, which exhibits both parallel and crossed arrangement of the
metal-organic chains.
Hybrid Vanadates, towards Metal-Organic Frameworks 25

Figure 19. Crystal structures, metal-organic substructure and vanadium oxide subunit for
{Ni(Pz)}(VO3)2, ,M=Ni, Co [{Ni6(H2O)10(4Bpy)6}(V18O51)]1.5 H2O and
[{Ni(H2O)2(Bpe)}(V4O11)]0.5 H2O. [30,31,39].

The third example of A1 archetype is the {Ni2(H2O)2(Bpa)2}(V6O17) compound, [48]


whose crystal structure has metal-organic chains parallel to each other and to the vanadium
oxide subunit. In its structure, the Ni(II) octahedra are directly coordinated to the Bpa ligands
generating {NiBpa}n chains parallel to the vanadium oxide chains (Figure 20). Each metal-
organic chain is connected to vanadium oxide chains in three different directions, while the
vanadium oxide subunit, chains of {V4O12} cycles and {V2O7} dimmers, are linked to six
different metal-organic chains (Figure 20 (b)).

Figure 20. (a) Crystal structure of {Ni2(H2O)2(Bpa)2}(V6O17) (metal centers (blue), vanadate chains
(green)). (b) Connectivity between the {V4O12} cycles and {V2O7} dimmers and the metal-organic
chains. The Bpa ligand has been simplified as a line. [48]

Finally, there are some A1 architectures exhibiting vanadium oxide layers connected by
the metal centers. These belong to the metal-organic chains which are arranged parallel
between the vanadate layers. Examples of this architecture are {M(Pz)}(V4O10) (M= Ni(II),
26 Edurne S. Larrea, Roberto Fernndez de Luis, Jos L. Mesa et al.

Co(II), Zn(II)) and {M(4Bpy)}(V4O10) (M= Cu(I), Ag(I)) (Figure 21) hybrid vanadates.
[32,43]

Figure 21. (a) Crystal structure of {M(4Bpy)}(V4O10) (M= Cu, Ag). (b) Connectivity between the
metavanadate chains and the metal-organic chains. [43]

Among the A2 archetypes, three different arrangements of the metal-organic layers have
been observed. In the first one, the metal centers are connected in two different directions
through the dipodal ligand, generating square-like ({M2(Bbp)4}(V4O12) (M=Ni,Co)
{Co3(H2O)(Btp)5}(V8O23)) or rectangular-like ([{M2(H2O)2(4Bpy)3}(V4O12)]2.5H2O
M=Ni(II), Co(II)) inclined interpenetrated metal-organic sheets. [54,55,58,37,38] Figure 22
shows the crystal structure and the interpenetrated metal-organic sheets of
[{M2(H2O)2(4Bpy)3}(V4O12)]2.5H2O, M=Ni(II), Co(II), hybrid vanadate. In all of these
compounds, the metal centers are also connected to the vanadium oxide subunits in two
different directions, generating the 3D inorganic substructure characteristic for the A-type
crystal structures.

Figure 22. 3D porous crystal structure of [{M2(H2O)2(4Bpy)3}(VO3)4]2.5H2O (M= Ni, Co) (left) and
metal-organic substructure constructed from the interpenetration of two rectangular-like layers. [37,38]

On the other hand, the [{Cu3(H2O)2(Tpp)2}(V8O23)]3H2O (Figure 23 (a)) and


[{Zn3(H2O)2(Tpp)2}(VO3)6]6H2O [62] compounds exhibit parallel interpenetration of
hexagonal like metal-organic layers (3D+(2D + 2D) 3D), instead of the previously
described inclined interpenetration (Figure 23 (b)). The hexagonal-like metal-organic
substructures are clearly related with the geometry of the Tpp ligand. The
[{Cu3(H2O)2(Tpp)2}(V8O23)]3H2O is the unique hybrid vanadate that exhibits 2D vanadium
oxide layers constructed exclusively from four coordinated vanadium (Figure 23 (c) and (d)).
Hybrid Vanadates, towards Metal-Organic Frameworks 27

Figure 23. (a) Crystal structure of [{Cu3(H2O)2(Tpp)2}(V8O23)]3H2O. (b) Parallel interpenetration of


hexagonal-like metal-organic layers. (c) and (d) Vanadium oxide layers constructed from corner linked
VO4 tetrahedra. [62]

In the third A2 architecture observed, there are no interpenetrated layers. By contrast, the
matel-organic layers are disposed between the vanadium oxide layers and parallel to them. In
the [{Cu(Pz)2}(V6O16)]0.22H2O structure, [35] the Cu(II) cations are linked in two directions
through the organic ligands, generating square-like metal-organic layers.
The A3 archetypes are characterized by the coexistence of both 3D inorganic and metal-
organic substructures. The [{Ni8(4Bby)16}(V24O68)]8.5H2O [40] and
[{Ni3(H2O)3(Bpa)4}(V6O18)]8H2O [48] are examples of this architecture. The metal-organic
substructure of [{Ni8(4Bby)16}(V24O68)]8.5H2O consists of the polycatenation of cds like
3D metal-organic subnet with two square-like metal-organic layers. The vanadium oxide
chains are located in the channels of the metal-organic substructure and linked via corners to
the Ni(II) cations. The metal-organic substructure of [{Ni3(H2O)3(Bpa)4}(V6O18)]8H2O
consists of the interpenetration of three 3D m ot-like metal-organic nets. The {V12O36}
vanadate cycles connect the metal centers of the different metal-organic nets. giving rise to
six and ten connected porous crystal structures.

4.2. B Architectures: Hybrid Vanadates Archetypes Containing


Two-Dimensional Inorganic Layers

In B architectures, the inorganic substructure is 2D. Depending on the dimensionality of


the metal-organic substructures, four subtypes can be described, B0, B1, B2 and B3.
However, only examples of the first three are reported.
Two subtypes of B0 architectures can be distinguished, B0a and B0b. B0a subype is
characterized for having discrete metal-organic complexes between the vanadate layers. In
B0b crystal structures, the secondary metal is bonded to the vanadate layers and the ligand
acts chelating the metal, generating 2D inorganic-organic layers stabilized by weak
interactions. Examples of these type of architectures are: {Co(Phen)3}(V10O27), [68]
{Co2(2Bpy)2}(V6O17), [69] .{Co2(Phen)4}(V8O23). [70] and {Zn(Tpp)2}(VO3)2.
28 Edurne S. Larrea, Roberto Fernndez de Luis, Jos L. Mesa et al.

In B1 and B2 architectures, the vanadium oxide subunit is linked with the metal centers
generating bimetallic inorganic layers. The ligand is directly linked to the metal centers and
acts as a pillar between the inorganic layers. The main difference between the B1 and B2
archetypes is the dimensionality of the metal-organic substructure; B1 crystal structures
contain metal-organic chains, while B2 architectures has 2D metal-organic layers.
The {Ni2(H2O)2(Pz)}(V4O12), {Co(4Bpy)}(VO3)2, {Co(Bpa)}(VO3)2,
[{Ni(H2O)(Bpa)}(VO3)2]2H2O, {M2(H2O)2(Bbi)}(VO3)4 (M=Mn(II), Co(II)),
[{Ag(4Bpy)}4(V4O12)]2H2O and {Cu3(Trz)2}(V4O12) hybrid vanadates belong to the B1
group (Figure 24). [33, 34, 41, 42, 48, 51, 61] The main differences between these
compounds are the coordination environment of the metal centers, the vanadium oxide
subunit and the connectivity between them to generate the inorganic layers.

Figure 24. (a.1), (b.1) and (c.1): Crystal structures of {Co(4Bpy)}(VO 3)2, {Co(Bpa)}(VO3)2 and
[{Ni(H2O)(Bpa)}(VO3)2]2H2O. (a.2) and (b.2) Inorganic layers constructed from CoN2O3 and {V4O12}
cycles. (c.2) Inorganic layers constructed from Ni(H2O)N2O3 octahedra and {V4O12} cycles. [41,48]

The B2 compounds present metal-organic layers instead of the metal-organic chains of


the B1 architectures. For [{Mn(4Bpy)}(VO3)2]1.16H2O and
[{Mn(4Bpy)0.5}(VO3)2]0.62H2O, [{Ag(Bpa)}4(V4O12)]4H2O [42, 45], the metal centers are
also linked to each other within the inorganic layers, and hence the connectivity through the
organic ligand generates parallel metal-organic layers. In the specific case of the
{Ni(Bpe)}(VO3)2 [46, 80] hybrid vanadate, the metal centers are sharing edges with each
other, generating dimmeric units. Each dimmeric unit is connected to another four through the
organic ligand, giving rise to an interpenetration of two square-like metal-organic layers
within the crystal structure. The metal centers are also connected through the metavanadate
chains, giving rise to a ten-connected and self-catenated crystal structure.
The B3 architecture is a hypothetical crystal structure combining inorganic layers and 3D
metal-organic nets. This architecture has not been observed in any known hybrid vanadate.
The B1 and B2 crystal archytectures are usually observed in MOFs, such as in cobalt(II)
and nickel(II) succinates. [71] This fact indicates the structural similarities between the hybrid
vanadates containing chain or layer metal-organic substructures and the classic MOFs.
Hybrid Vanadates, towards Metal-Organic Frameworks 29

4.3. C Architectures: Hybrid Vanadates Archetypes Containing One-


Dimensional Inorganic Chains

There are very few examples of hybrid vanadates with dipodal ligands that contain 1D
inorganic chains combined with different metal-organic chains, layers or nets. The presence
of dipodal ligands seems to favors the crystallization of the previously described structural
archetypes instead of crystal structures containing one-dimensional inorganic substructures.
Despite this fact, this type of hybrid vanadates shows very interesting crystal structures, and
shares more characteristics with the classic MOFs than the previous.
In C0 archetypes, the ligand chelates the metal centers, giving rise to discrete metal-
organic units. Those can be linked to metavanadate chains, or connected through discrete
vanadium oxide units giving rise to bimetallic inorganic chains. Examples of C0 architectures
are the {Co(HBpe)2}(V4O12), [47] {Cd(HBbi)2}(V4O12), [49] [{Mn(Phen)}(V4O12)]0.5H2O,
[72] {Co(2Bpy)2}(V4O12) [73] compounds, among others.
Another possible C0 architecture is observed in [{Cu2(Bpq)}(V4O12)]2H2O,
{Zn2(Btpt)}(V4O12) and {Co2(Btpt)2}(V4O12), hybrid vanadates. [59, 60] In these structures,
the metal centers of adjacent inorganic chains are connected through the organic ligand. The
metal-organic substructure is 0D, but the connectivity through the organic molecule generates
an increase of the crystal structure dimensionality 1D+0D = 2D.
The C1 architectures consist of inorganic chains connected by the ligand. The linkage
between the metal centers and the organic molecule also gives rise to 1D metal-organic
chains. Consequently, an increase in the dimensionality of the crystal structure is observed. A
hypothetical archetype in hybrid vanadates, commonly observed in MOFS, [74], would be
that in which the adjacent inorganic chains are linked through the organic ligand, giving rise
to inorganic-organic layers (1D+1D = 2D), which are stacked by weak interactions (hydrogen
bonds, Van der Walls). This archetype is still not reported for hybrid vanadates.
The [{Cu(4Bpy)}4(V4O12)]2H2O compound [44] is the unique example of hybrid
vanadate exhibiting a C1 architecture. (Figure 25) In this concrete case, each inorganic chain
(Figure 25 (a) and (b)) is connected to another six through the organic ligand, generating a 3D
crystal structure 1D+1D = 3D (Figure 25(c)).

Figure 25. (a) and (b) inorganic chains and (c) crystal structure of [{Cu(4Bpy)}4(V4O12)]2H2O. [44]

The C2 crystal structures combine both inorganic chains and metal-organic layers. Only
two examples are known in hybrid vanadates, {Cu2(Bbi)3}(V4O12), [52] and
30 Edurne S. Larrea, Roberto Fernndez de Luis, Jos L. Mesa et al.

{Cu3(Ptrz)2}(VO3)4. [63] The inorganic-organic layer of {Cu2(Bbi)3}(V4O12) hybrid vanadate


has copper(II) cations linked to two {V4O12} cycles via vertexes, generating 1D inorganic
chains (Figure 26 (a.1)). The copper(II) cations of adjacent inorganic chains are connected via
the Bbi ligand (Figure 26 (a.2)). The metal-organic layer is a {Cu(Bbi)2} square-like sheet
(Figure 26 (b.1)). The interpenetration of these two kinds of layers generates a 3D framework
without an increase of the dimensionality for the inorganic substructure because the metal-
organic layers are not directly attached to the inorganic-organic ones, being the
interpenetration what generates the increase of the dimensionality (Figure 26 (b.2) and (b.3)).

Figure 26. (a.1) Inorganic chains. (a.2) Inorganic-Organic layers. (b.1) {Cu(Bbi)2} metal-organic layers
of {Cu2(Bbi)3}(V4O12). (b.2) and (b.3) interpenetration of the metal-organic and inorganic-organic
layers. [52]

{Cu3(Ptrz)2}(VO3)4 compound combines {Cu3(PTrz)2} undulated layers (Figure 27 (b))


linked through metavanadate chains (Figure 27 (a)). The copper(II) metal centers connect the
metavanadate chains, generating also 1D inorganic chains (Figure 27 (c)). The crystal
structure is also described as 1D+2D = 3D.

Figure 27. (a) Crystal Structure of {Cu3(PTrz)2}(VO3)4 (b) {Cu3(PTrz)2}metal-organic layers. (c)
Inorganic chains. [63]
Hybrid Vanadates, towards Metal-Organic Frameworks 31

The C3 crystal architecture is similar to the C2 one, however, the metal centers are
connected through the organic ligand generating 3D metal-organic substructure. In that
respect, the [{M2(Bbi)3}(V4O12)] 4H2O, M= Ni(II), Co(II) [51,52] hybrid vanadate possesses
two interpenetrated 3D six-connected metal-organic nets (Figure 28 (a) and (b)). The metal
centers, belonging to the metal-organic substructure, are connected through {V4O12} cycles
giving rise to the inorganic chains (Figure 28 (c)) and an eight-connected self catenated
crystal structure shown in Figure 28 (d).

Figure 28. (a) Six-connected dimmeric units. (b) Two interpenetrating 3D metal-organic substructure.
(c) Inorganic chains. (d) Eight-connected self-catenated structure constructed from inorganic chains and
metal-organic net of[{M2(Bbi)3}(V4O12)] 4H2O. M=Ni(II), Co(II). [51,52]

4.4. D Architectures: Hybrid Vanadates Archetypes Containing Discrete


Inorganic Units

In the D-type architectures, the inorganic substructure is 0D and acts as the node
connected through the organic ligand, like in MOFs. Four different types are described in
function of the metal-organic dimensionality, D0, D1, D2 or D3 for discrete, one-
dimensional, two-dimensional or three-dimensional metal-organic substructures.
The D0 architectures are discrete coordination complexes commonly generated from
chelating ligands. The metal centers are also linked to discrete vanadium oxide subunits.
Concretely, for hybrid vanadates, this kind of structural archetype is deeply described by J.
Zubieta et al. [10] An example of this architecture is the [{Cu2(Btpt)2}(V4O12)] compound.
[61]
The D1 crystal structures contain metal-vanadate discrete units linked through the
organic ligand. The connectivity between the metal centers and the organic ligand generates
metal-organic chains. Only one example belonging to this group is known,
{Cu4(Btpt)2}(V4O12) [60, 57] The D2 architectures are very common in MOFs, and the
discrete units are linked in two different directions generating metal-organic layers, however,
only one example is known in hybrid vanadates; the [{Cu(Bbi)2}(V4O12)]4H2O hybrid
32 Edurne S. Larrea, Roberto Fernndez de Luis, Jos L. Mesa et al.

vanadate consists of {Cu(Bbi)2} metal organic layers linked through weak interations to
V4O12 cycles.
The D3 crystal structures are extensively represented in MOFs. In that respect, the
combination of metal centers or metal clusters and carboxylic ligands generates different
families of coordination polymers exhibiting the same topology and very interesting
physicochemical properties. [75] However, with respect to the hybrid vanadates, only one
example is known, [{Cu4(Bbp)4}(V4O12)]3H2O. [57] The Cu4(V4O12) inorganic units are
connected to another eight, generating an unprecedented eight-connected self-catenated
crystal structure (Fig 29).

Figure 29. (a) Cu4V4O12 inorganic clusters present in [{Cu4(Bbp)4}(V4O12)]3H2O structure. (b) Eight
connected nodes. [57]

5. PHYSICOCHEMICAL PROPERTIES OF HYBRID


VANADIUM OXIDES
The physicochemical properties of hybrid materials are directly related to their crystal
structures. In that respect, the present section will be focused on the thermal and magnetic
properties and the catalytic activity of hybrid vanadium oxides, and specifically on dipodal
and multi-podal-based vanadates.
The thermal properties of hybrid vanadates will be centered on the 3D and 2D crystal
structures. Unfortunately, most of the thermal studies carried out on hybrid vanadates consist
on thermogravimetric measurements, which only give information about the ranges of
temperatures in which the different decomposition processes, such as, loss of crystallization,
removal of coordination water molecules or calcination of the organic ligand, occurs. There
are few examples in which thermodiffractometric studies have been carried out. These
measurements complement the thermogravimetric studies, and give a lot of information about
the thermal response of the crystal structure to the loss of solvent or coordinated species, or
simply, to the increase of temperature (thermal expansion or structural transitions).
The magnetic properties of hybrid vanadates are directly related to the connectivity of the
metal centers through the vanadium oxides subunits and the organic ligands. Usually, the
magnetic exchange is not very strong, however, there are several examples exhibiting very
interesting magnetic properties, such as crystal structures that contain connected metal
centers, giving rise to magnetic discrete units (dimmers) or 1D chains, or hybrid vanadates
with short aromatic dipodal ligands, which favor the magnetic exchange between metal
centers.
Hybrid Vanadates, towards Metal-Organic Frameworks 33

On the other hand, vanadium is one of the most reactive elements. Its good catalytic
activity is well-documented in the literature [76]. Since the discovery of vanadate-dependent
enzymes from various algae [77] and terrestrial fungi [78], the coordination chemistry of this
element has received increasing interest from researchers. However, very few hybrid
vanadates were tested as catalysts.

5.1. Thermal Response of the Crystal Structure to the Loss of Crystallization


and Coordinated Water Molecules

The thermal response of the crystal structures to the loss of crystallization and
coordinated water molecules depends on the crystal architecture and the rigidity or flexibility
of the organic ligand. Several examples of hybrid vanadates exhibiting different thermal
behaviors are going to be discussed.

5.1.1. Rigid A1 Crystal Architectures Containing Crystallization and Coordinated


Water Molecules: [{Ni6(H2O)10(4Bpy)6}(V18O51)]1.5H2O and [{Ni(H2O)2(Bpe)}
(V4O11)]0.5H2O
The 3D structures of both compounds consist on 3D rigid inorganic frameworks [39].
The organic ligands are directly coordinated to the metal centers and acts as pillars that
stabilize the crystal framework. The crystallization water molecules are encapsulated between
the inorganic scaffold and the organic ligand, and the coordination environment of the metal
centers is completed by coordinated water molecules (see Section 4.1.).
The thermogravimetric measurements (Figure 30) show that the loss of co-crystallized
water molecules in these materials finished at high temperatures (180C). This is not
surprising because these are encapsulated between the organic ligands and the inorganic
framework. The loss of coordinated water molecules is a continuous process that takes place
in the 180-280 C temperature range. At higher temperatures, up to 370C, the calcination of
the organic ligand occurs.

Copyright (2009) Wiley-VCH Verlag GmbH and Co. KGaA. Reproduced with permission.

Figure 30. ATD and DSC curves for [{Ni6(H2O)10(4Bpy)6}(V18O51)]1.5H2O and


[{Ni(H2O)2(Bpe)}(V4O11)]0.5H2O. Reproduced with permission from Ref. [39].
34 Edurne S. Larrea, Roberto Fernndez de Luis, Jos L. Mesa et al.

The thermal stability was also studied by X-ray thermodiffractometry in air atmosphere
(Figure 31). No structural response of the crystal structures to the loss of crystallization water
molecules is observed, probably due to the combination of a rigid 3D inorganic framework
and also rigid ligands, such as 4Bpy and Bpe. In that respect, the crystal structures possess a
second generation structural response to the loss of crystallization water molecules. The loss
of coordinated water molecules gives rise to a strong reorganization of the crystal structures,
promoting the crystallization of the anhydrous phases, up to 210C (Figure 31). Moreover, the
crystallinity is drastically reduced during the transformation, hindering the study of the
anhydrous phase by X-ray diffraction.

Figure 31. Thermodiffractometric measurements for [{Ni6(H2O)10(4Bpy)6}(V18O51)]1.5H2O and


[{Ni(H2O)2(Bpe)}(V4O11)]0.5H2O. Reproduced with permission from Ref. [39]. Copyright (2009)
Wiley-VCH Verlag GmbH and Co. KGaA. Reproduced with permission.

For [{Ni(H2O)2(Bpe)}(V4O11)]0.5H2O, the position of the diffraction maxima in the


hydrated compound can be related to the position of these corresponding to the anhydrous
phase. Taking into account that the crystal structure of the anhydrous compound probably is
related to the initial one, the unit cell has to be similar. However, the intensity of the
reflections changes drastically during the removal of coordination water molecules,
suggesting a strong structural reorganization during the transformation. The position of the
(020) reflection is displaced from 7.9 to 7.4 in 2 while the (011) maximum changes its
position from 8.3 to 9.0 (Figure 32).
The initial values of the b and c parameters were calculated from the position of
(020) and (110) reflections. The obtained cell parameters (Table 7) have to be considered
qualitatively, because the pattern of the anhydrous phase exhibits poor crystallinity, with only
few and broad maxima in the measured 2 range.
Hybrid Vanadates, towards Metal-Organic Frameworks 35

Copyright (2009) Wiley-VCH Verlag GmbH and Co. KGaA. Reproduced with permission.

Figure 32. Displacement of the (020) and (011) reflections during the structural transformation of
[{Ni(H2O)2(Bpe)}(V4O11)]0.5H2O. Reproduced with permission from Ref. [39].

Table 7. Cell parameter for [{Ni(H2O)2(Bpe)}(V4O11)]0.5H2O and the anhydrous one

a () b () c () ()
[{Ni(H2O)2(Bpe)}(V4O11)]0.5H2O 7.2721(3) 23.062(2) 12.4826(6) 91.245(4)
{Ni(Bpe)}(V4O11) 7.722(9) 24.18(1) 10.824(3) 91.3(1)

Probably, the removal of coordination water molecules is compensated by the generation


of two new Ni-O bonds with the terminal oxygen atoms of the VO5 polyhedra (Figure 33).
This connectivity shorts the distances between the adjacent vanadate chains, and hence,
reduces the c parameter. The same connectivity between the nickel(II) ions and the
vanadate chains is observed in the inorganic vanadate Ni(VO3)24H2O [79]. The cooperative
movements of the structural subunits imply a strong reorganization of the crystal structure,
and hence, promote a reduction of the crystal domains, lowering the crystallinity of the
patterns.

Figure 33. Qualitative model for the transformation due to the loss of coordinated water molecules.
Reproduced with permission from Ref. [39]. Copyright (2009) Wiley-VCH Verlag GmbH and Co.
KGaA. Reproduced with permission.
36 Edurne S. Larrea, Roberto Fernndez de Luis, Jos L. Mesa et al.

The infrared spectra also show a progressive reduction of the band related with the
stretching vibration of the O-H bonds (Figure 34 (a)), in good agreement with the loss of
crystallization (25-125C) and elimination of the coordinated (175-250C) water molecules.
The removal of co-crystallized water molecules do not modify the bands related with the
organic molecule (Figure 34 (b)) and vanadium oxide subunit (Figure 34 (c) and (d)).

Figure 34. Infrared spectra for the original sample and heated ones at 125, 150, 175, 200, 225, 250 C
during one hour. Reproduced with permission from Ref. [39]. Copyright (2009) Wiley-VCH Verlag
GmbH and Co. KGaA. Reproduced with permission.

After the removal of coordinated water molecules, the vibrational stretching C=C band is
split in two signals (Fig 34 (b)), and the absorption maxima related to the breathing of the
pyridyl rings are broadening (Figure 34 (c)). However, the most important differences in the
IR spectra are related to the vibrational modes of the vanadium oxide subunit (Figure 34 (c)
and (d)). Concretely, the position and intensity of the bands related with the V=O and V-O-V
stretching vibrations (Figure 34 (c)) changes drastically during the transformation. In both
compounds, the changes of the IR spectra suggest a strong reorganization of the vanadium
oxide subunit and minor changes in the organic molecules during the removal of coordinated
water molecules.

5.1.2. Flexible A3 Crystal Architectures: The Effect of the Rigidity-Flexibility of the


Organic Ligand, [{Ni8(4Bpy)16}(V24O68)]8.5H2O and {Ni3(H2O)3(Bpa)4}(V6O18)]8H2O
The described vanadates are the unique examples of A3 architectures, and both of them
contain crystallization water molecules located in the channels of the crystal structures (see
section 4.1). The thermogravimetric measurements show two processes for
[{Ni8(4Bpy)16}(V24O68)]8.5H2O corresponding to the loss of crystallization water molecules
and the calcination of the organic ligands (Figure 35 (a)). For
[{Ni3(H2O)3(Bpa)4}(V6O18)]8H2O the release of the coordinated water molecules is also
observed (Figure 56 (b)). [40, 48]
Hybrid Vanadates, towards Metal-Organic Frameworks 37

Figure 35. TG and DSC curves for (a) [{Ni8(4Bpy)16}(V24O64)]8.5H2O and


(b)[{Ni3(H2O)3(Bpa)4}(V6O18)]8H2O. Adapted from Ref. [40]., Copyright (2010), with permission
from RSC.

The thermodiffractometric experiment for [{Ni8(4Bpy)16}(V24O68)]8.5H2O indicates that


the initial compound is stable up to 370C. However, the crystallinity decreases dramatically
by 355C. The Rietveld refinement of the crystal parameters with respect to the temperature
show a 0.6% contraction for the unit cell volume (Figure 36 (d)) due to the release of the
crystallization water molecules (30-100C).

Figure 36. Thermal evolution of the cell parameters for [{Ni8(4Bpy)16}(V24O68)]8.5H2O. Reprinted
from Ref. [40]. Copyright (2010), with permission from RSC.
38 Edurne S. Larrea, Roberto Fernndez de Luis, Jos L. Mesa et al.

The thermal evolution of the crystal parameters and the cell angles also indicates the
existence of a distortion and compression of the metalorganic (Figure 36). Despite the
existence of a 3D inorganic scaffold, and a rigid ligand such as the 4Bpy, the crystal structure
is contracted during the loss of crystallization water molecules.

Figure 37. Two representative intervals of the thermodiffractometry of ({Ni3(H2O)3(Bpa)4}


(V6O18))8H2O.

Figure 38. Thermal variation of the position and intensity for (110) (a), (002) (b) and (100) maxima for
[{Ni3(H2O)3(Bpa)4}(V6O18)]8H2O.
Hybrid Vanadates, towards Metal-Organic Frameworks 39

The [{Ni3(H2O)3(Bpa)4}(V6O18)]8H2O framework possesses a more dynamic structural


response to the loss of solvent, probably due to the presence of the Bpa organic ligand in the
crystal structure. [48] The thermodiffractometric measurement shows important
displacements of the diffraction maxima position during the heating process (Figure 37).
The peak fit of the (110), (002) and (100) maxima allow us to determine the thermal
variation of intensity and 2() position for these reflections (Figure 38). Both the shift of the
(110) maximum and the cell parameters evolution suggest three different tendencies during
the heating process, in good agreement with the TG/ATD measurement.
The first process occurs between room temperature and 85C, and gives rise to an
important increase of the (110) reflection intensity and an important shift of the (002)
maximum from 6.70 6.88 in 2() (Figure 38(a)). The temperature range is related to the
loss of crystallization water molecules. This process generates the decrease of the c
parameter, in good agreement with the (002) maximum displacement. This fact also indicates
a contraction of the [100] channels in which the crystallization water molecules were located.
The (110) reflection intensity increase also suggests important structural changes of the host
during the release of crystallization water molecules. Above 85C, and up to 155C, the (110),
(002) and (100) reflections shift its positions from 5.55 5.75 , 6.80 6.95 and 6.90
7.40 in 2, respectively. The (002) maximum also lost intensity progressively. According to
the thermal displacements of the three strongest reflections, the crystal structure continues its
contraction after the removal of the crystallization water molecules. This fact suggests that the
dehydrated compound exhibits negative thermal expansion behavior. In the 200C to 275C
temperature range, the diffraction patterns show a continuous loss of crystallinity of the
maxima located at 2() values higher than 6. The loss of coordinated water molecules
during this temperature range generates an irreversible and strong structural transformation,
and probably implies a strong structural reorganization promoting a reduction of the crystal
domains, and lowering the crystallinity of the patterns.

Figure 39. Diffraction patterns of [{Ni3(H2O)3(Bpa)4}(V6O18)]8H2O samples after being heated at 100
(A), 150 (B), 200 (C) and 250C (D) and exposed to water at room temperature.
40 Edurne S. Larrea, Roberto Fernndez de Luis, Jos L. Mesa et al.

In order to corroborate the reversibility of the heating response of the crystal structure,
patterns were recorded at 100, 150, 200, 250C and after lowering the temperature to 30C
and adding distilled water to the pre-heated samples. The results are summarized in Figure 39.
The patterns are completely identical to the initial phase when the sample is heated up to
150C or below. This fact indicates that both the loss of crystallization water molecules and
the negative thermal expansion between 85 and 155C are completely reversible. The samples
heated at 200C partially recover the position of the diffraction maxima. However, the pattern
recorded at 30C of the sample heated at 250C do not recover the position and intensity of
the maxima, revealing that the loss of coordinated water molecules is an irreversible process.
The main difference between the thermal response of the
[{Ni8(4Bpy)16}(V24O68)]8.5H2O and [{Ni3(H2O)3(Bpa)4}(V6O18)]8H2O compounds to the
loss of solvent is the range of cell volume contraction. Both compounds exhibit a reversible
and dynamical structural response, however while the crystal framework of
[{Ni8(4Bpy)16}(V24O68)]8.5H2O responses with a 0.6% of cell volume reduction, the
[{Ni3(H2O)3(Bpa)4}(V6O18)]8H2O compound exhibits a contraction of approximately 6%.
Both compounds possess complex 3D structures combining both 3D inorganic and metal-
organic substructures, however, the flexibility of the Bpa ligand is larger than the 4Bpy one.
While Bpa organic molecules could adopt its conformation, giving rise to a reduction of the
distance between the nitrogen donor atoms, the 4Bpy ligand acts as a rigid pillar.

5.1.3. B2 Crystal Architectures: The Reorganization of Layered-Like Hybrid Vanadates


due to the Loss of Crystallization and Coordinated Water Molecules:
[{Mn(4Bpy)}(VO3)2]1.16H2O, [{Mn(4Bpy)}0.5(VO3)2]0.62H2O, {Ni2(H2O)2(Pz)}(V4O12)
An amazing example of dynamical and reversible response of the crystal structures due to
the loss of crystallization water molecules are the manganese hybrid vanadates
[{Mn(4Bpy)}(VO3)2]1.16H2O and [{Mn(4Bpy)}0.5(VO3)2]0.62H2O. [45] The crystallization
water molecules are located between the organic pillars and the inorganic layers (Figure 41).
For both compounds, two important displacements and intensity changes of some reflections
take place in the heating process [{Mn(4Bpy)}(VO3)2]1.16H2O at 28 and 40 C;
[{Mn(4Bpy)}0.5(VO3)2]0.62H2O at 26 and 42 C) (Figure 40).
The structural changes due to the removal of crystallization water molecules give rise to
important reductions of the a parameter for both compounds. For
[{Mn(4Bpy)}0.5(VO3)2]0.62H2O, an appreciable change in the c parameter is also
observed. The thermal evolution of the cell parameters allows a qualitative description of the
structural changes. Thus, the loss of solvent gives rise to a tilting of the 4Bpy pillars between
the inorganic layers, and hence, an increase of the angle value. However, surprisingly, the
major structural changes are directly associated with the compression of the inorganic layers,
which generates the reduction of the a parameter. This way, the opposite tilt of the vanadate
and manganese octahedra chains explains, qualitatively, the variation in the unit cell
parameters for both compounds due to the removal of the solvent (Figure 41).
The {Ni2(H2O)2(Pz)}(VO3)2 hybrid vanadate exhibits an irreversible structural
transformation due to the loss of the coordinated water molecules. In that respect, the
thermodiffractometric measurement indicates that the transformation occurs up to 260C,
with a drastic change in the diffraction maxima (Figure 42). [33, 34]
Hybrid Vanadates, towards Metal-Organic Frameworks 41

Figure 40. Two representative intervals of the thermodiffractometry of (a)


[{Mn(4Bpy)}(VO3)2]1.16H2O and (b) [{Mn(4Bpy)}0.5(VO3)2]0.62H2O. Reprinted with permission
from Ref [45]. Copyright 2010 American Chemical Society.

Figure 41. Qualitative description of the crystal response to the loss of crystallization water molecules
in [{Mn(4Bpy}(VO3)2]1.24H2O and (b) [{Mn(4Bpy}0.5(VO3)2]0.62H2O. Reprinted with permission
from Ref [45]. Copyright 2010 American Chemical Society.
42 Edurne S. Larrea, Roberto Fernndez de Luis, Jos L. Mesa et al.

Figure 42. Thermodiffractometry of {Ni2(H2O)2(Pz)}(V4O12). Reprinted from Ref [33]. Copyright


(2008), with permission from Elsevier.

The loss of coordinated water molecules generates a very important structural


reorganization, and also a drastic reduction of the crystallinity, probably associated to a
reduction of the crystalline domains. The crystal structure of the anhydrous compound could
not be solved, however a structural transformation pathway can be proposed, taking into
account the disposition of the metal centers and {V4O12} cycles within the inorganic layers of
the compound (Figure 43).

Figure 43. Structural transformation pathway proposed for the dehydration of {Ni2(H2O)2(Pz)}(V4O12).

The coordinated water molecules, linked to the nickel(II) cations, establish an intralayer
hydrogen bond, Ow(7)-Hw(7)O3. It is very likely that the six-coordination environment of
the nickel(II) atoms is maintained during the structural transformation. So, taking into account
this fact, a slight rotation of the nickel dimmers along the direction of the aforementioned
hydrogen bond could allow the generation of a new Ni-O(3) bond and compensate the loss of
coordinated water molecules. This rotation would produce the approach between the layer,
which is in good agreement with the shift of the first maxima to higher 2 values observed in
the thermodiffractometry.
An amazing example of a combined loss of crystallization and coordinated water
molecules at low temperature is the [{Ni(H2O)(Bpa)}(V4O12)]2H2O hybrid vanadate (see
Hybrid Vanadates, towards Metal-Organic Frameworks 43

section 4.1, Figure 20). [48] The thermodiffractometric measurement shows that the
compound loss the crystallization and coordinated water molecules at 50C, with a complete
structural reorganization, according to the change in the maxima position and intensity. The
structural transformation is a reversible process with an important change in the compound
color (Figure 44).

Figure 44. Thermodiffractometry of [{Ni(H2O)(Bpa)}(V4O12)]2H2O.

A good structural model for the nickel(II) anhydrous compound cannot be obtained from
the powder patterns. However, the structural similarities observed for the hydrated compound,
[{Ni(H2O)(Bpa)}(V4O12)]2H2O, and the {Co(Bpa)}(V4O12) vanadate, [48] allow a
comparative study in order to obtain the cell parameters of the anhydrous phase. The analysis
suggests a supercell with a propagation vector of (0 0.5 0.5), with regard to the
{Co(Bpa)}(V4O12) cell parameters. Therefore, the b and c parameters are doubled.
Despite that there are few thermal studies carried out in hybrid vanadates, it is possible to
conclude that the thermal response of the crystal structures depends on the crystal structure
itself. The thermal contraction during the release of water molecules is a continuous process
in the A1 and A3 crystal architectures; however, the existence of a significant structural
contraction depends on the rigidity of the crystal framework. In that respect, the
[{Ni6(H2O)10(4Bpy)6}(V18O51)]1.5H2O and [{Ni(H2O)2(Bpe)}(V4O11)]0.5H2O do not show
any structural response because the inorganic crystal substructures are rigid frameworks
stabilized by rigid ligands. The [{Ni8(4Bpy)16}(V24O68)]8.5H2O possess a more open crystal
framework, and exhibits a slight contraction of the crystal structure. Despite that, the 4Bpy
ligands act as rigid pillars, the inorganic scaffold is constructed from corner-linked
polyhedra, and hence, the crystal structure possesses some degree of freedom to respond to
the loss of the solvent.
On the other hand, when a flexible ligand is introduced
([{Ni3(H2O)3(Bpa)4}(V6O18)]8H2O), the structural response to the loss of crystallization
water molecules is also continuous, but the contraction is ten orders larger than for the 4Bpy
compound. For B2 architectures, the loss of crystallization water molecules, and hence, the
crystal structure response, is restrained to concrete temperature ranges. The structural
transformations take place near room temperature.
44 Edurne S. Larrea, Roberto Fernndez de Luis, Jos L. Mesa et al.

5.1.4. Negative Thermal Expansion in {Ni(Bpe)}(VO3)2


The thermal behavior and thermal stability of this compound was studied by
thermogravimetry and thermodiffractometry (Figure 45). The thermodiffractometric
experiments were carried out both at room pressure and under vacuum. [48,80] The
thermogravimetric and thermodiffractometric measurements establish the thermal stability of
{Ni(Bpe)}(VO3)2 at 400C. However, above 385C, the diffraction patterns show a
progressive loss of intensity associated with the crystal structure collapse due to the
calcination of the organic ligand.

Figure 45. Thermodiffractometry for {Ni(Bpe)}(VO3)2.Adapted from Ref. [80]., Copyright (2010), with
permission from RSC.

The Rietvel analysis indicates three different tendencies during the heating process:

(i) 25 100C: Structural contraction due to negative thermal expansion.


(ii) 100C 315C: Thermal expansion of the crystal structure.
(iii) Above 315C: Structural collapse due to the progressive calcination of the organic
ligand.

The thermal evolution of the cell parameters was determined by Rietveld analysis of the
diffraction patterns. The relative thermal expansion for the cell parameters and volume was
calculated taking into account the R.E.(%)=[(Pt(i)/Pt(0))-1]x100 formula ((Pt(0)= the parameter
value at room temperature and Pt(i)= the parameter value at a given temperature). The results
are shown in Figure 46.
The thermal evolution of the cell volume between 25 and 100C (-0.3%) confirms the
negative thermal expansion of the crystal structure. The thermal behavior is closely related
with the b and c parameters. In order to determine the nature of the negative thermal
expansion more precisely, thermodiffractometric experiments were carried out at room
pressure and under vacuum between room temperature and 150C. The thermal expansion of
the cell parameters and volume were determined by Rietveld analysis of the patterns. The
evolution of the cell parameters for the measurements carried out at room pressure and under
vacuum is depicted in Figure 47. The thermal evolution of the b and c is strongly
dependent on the pressure. The b parameter shows a negative thermal expansion at room
pressure and positive thermal expansion under vacuum conditions. The negative thermal
expansion of the c parameter is appreciably reduced in vacuum conditions.
Hybrid Vanadates, towards Metal-Organic Frameworks 45

Figure 46. Relative thermal expansion for cell parameters and volume in {Ni(Bpe)}(VO3)2. Adapted
from Ref. [80]., Copyright (2010), with permission from RSC.

Figure 47. Relative expansion of the cell parameters and volume at room pressure and under vacuum
for {Ni(Bpe)}(VO3)2.

The thermal expansion of the materials is related to [81,82]:

(i) The increase of thermal motion along the chemical bonds.


(ii) The transversal vibrational motion.
46 Edurne S. Larrea, Roberto Fernndez de Luis, Jos L. Mesa et al.

The transversal motion could generate opposite rotation of the adjacent polyhedra and a
decrease of the M-M distance. In that respect, the transversal motion is the mechanism
responsible of the negative thermal expansion of different oxides, inorganic compound and
zeolites.[83 - 85] In MOFs or cyanide-based compounds, the oxygen bridge is replaced by a
cyanide group or an organic ligand.[86 - 89] The thermal structural response is related to the
relative rotation of the octahedra and organic linkers, which act as a rigid blocks.
In Figure 48, the simplification of the metal-organic subnets within the crystal structure
of {Ni(Bpe)}(VO3)2 is shown. The nickel(II) atoms belonging to the nickel dimmers have
been represented as spheres and the organic ligands as lines. The interpenetration angle
between the metal-organic layers is defined by the diagonal to the a and b parameters
(Figure 48 (a)). As shown in Figure 48 (b), the dimmers of the metal-organic nets are slightly
inclined with respect to the Bpe ligand. The reduction of the c parameter during the
negative thermal expansion can be attributed to a slight cooperative rotation of dimmers and
organic ligand (Figure 48 (b) and (c)). The qualitative model depicted in Figure 48 (b) and (c)
shows the reduction of the c parameter for a rotation of 15 of the metal centers with respect
to the organic ligand.

Figure 48. (a) Simplification of the metal-organic layers interpenetration in the (a,b) plane. (b) and (c)
Reduction of the c parameter due to a cooperative rotation of the dimmers.

This qualitative model only explains the thermal contraction of the


c parameter. For the
a and b parameters, another mechanism has to be considered. In that respect, the a and
b parameters are closely related to the interpenetration angle of the metal-organic layers,
defined by angle (Figure 49). The reduction of the interpenetration angle generates an
important negative thermal expansion of the b parameter, and a slight thermal expansion of
the a one.
From the model depicted in Figure 50, the theoretical values for the a and b
parameters have been calculated and compared with the real data obtained from the X-ray
diffraction experiments (Figure 50). In fact, only a reduction of -0.2 of the interpenetration
angle is necessary to explain the observed negative thermal expansion.
The proposed model only considers the metal-organic substructure, however VO4
tetrahedra belonging to the metavanadate chains also have to adapt their orientation during
the negative thermal expansion process.
Hybrid Vanadates, towards Metal-Organic Frameworks 47

Figure 49. Influence of the interpenetration angle of the metal-organic layers in the thermal evolution of
the a and b parameters for {Ni(Bpe)}(VO3)2.

Figure 50. Relative thermal expansion of a and


b parameters for {Ni(Bpe)}(VO3)2. Points:
Experimental data. Lines: Data simulated from the model depicted in Figure 70.

The model is in good agreement with the experimental data, however, one of the possible
reasons that could explain the negative thermal expansion is the pedal motion of the Bpe
ligand. The organic molecule presents two different conformations in the crystal structure. As
in previously reported works, [90] the increase of the temperature promotes the conformer
interconversion. In fact, the crystal data recorded at 150K and 293K presents slightly different
values for the occupation factor of the conformers: 150K 0.63/0.37, 293K: 0.55/0.45.

5.2. Magnetic Properties

The magnetic properties of hybrid vanadium oxides are directly related to the
connectivity between the secondary metal centers through the vanadium oxide subunits or
organic ligand.
There are several examples of hybrid vanadium oxides with edge-sharing or corner-
sharing metal centers, giving rise to discrete dimmeric units or 1D chains. In that respect, the
{Ni2(H2O)2(Pz)}(V4O12) and {Ni(Bpe)}(VO3)2 crystal structures contains inorganic layers
48 Edurne S. Larrea, Roberto Fernndez de Luis, Jos L. Mesa et al.

interconnected by metal-ligand chains. The nickel(II) octahedra share one edge, giving rise to
dimmeric units in both crystal structures. Figure 51 shows the connectivity of the dimmeric
units and the vanadium oxide subunits for {Ni2(H2O)2(Pz)}(V4O12) and {Ni(Bpe)}(VO3)2.
[33, 34, 78, 46, 80]

Figure 51. Inorganic sheets of (a) {Ni2(H2O)2(Pz)}(V4O12) and {Ni(Bpe)}(VO3)2.

The magnetic susceptibility shows a broad maxima located at 38K and 100K for
{Ni2(H2O)2(Pz)}(V4O12) and {Ni(Bpe)}(VO3)2, respectively (Figure 52). The temperature
dependence of the magnetic susceptibility is typical of an intradimmer anti-ferromagnetic
interaction, because the distances between adjacent dimmers are considerably longer than the
intradimmer separation.

Figure 52. Thermal evolution of the magnetic susceptibility (m) for (a) {Ni2(H2O)2(Pz)}(V4O12) and (b)
{Ni(Bpe)}(VO3)2. The lines are the susceptibility fitted to a isotropic dimmer model. (a) Reprinted from
Ref [33]. Copyright (2008), with permission from Elsevier. (b) Reprinted from Ref. [80]., Copyright
(2010), with permission from RSC.

The molar magnetic susceptibilities were fitted to a dimmer isotropic model, considering
the contribution at low temperatures of a minor paramagnetic impurity. The best fitting
parameters are shown in Table 8.
Hybrid Vanadates, towards Metal-Organic Frameworks 49

Table 8. Parameters obtained from the magnetic susceptibility fit

J/k (K) g p(%) TIP (cm3K/mol)


{Ni2(H2O)2(Pz)}(V4O12) -22.4 2.29 5.0 3.2 10-4
{Ni(Bpe)}(VO3)2 -59.4 2.08 4.4 7.610-4

The anti-ferromagnetic behavior for both compounds is similar to those observed in


octahedrally coordinated Ni(II) complexes with similar binuclear cores. The magneto-
structural correlations show that the J exchange parameter is directly connected with the Ni-
O-Ni bridging angle with a crossover value of 97 between the ferromagnetic and anti-
ferromagnetic behavior. [91] The Ni-O-Ni angle for {Ni2(H2O)2(Pz)}(V4O12) (95.95) and
{Ni(Bpe)}(VO3)2 (92.80) are into the values reported for a ferromagnetic behavior.
However, the out-of-plane angle is also as important as Ni-O-Ni angle to determine the sign
and direction of the exchange value. In that respect, the Ni-O-Ni angle for
{Ni2(H2O)2(Pz)}(V4O12) and {Ni(Bpe)}(VO3)2 may favor the ferromagnetic interaction
between metal centers, but the out-of-plane Ni-O-V angles and the connectivity between the
metal centers and the vanadium atoms have a great influence in the final magnetic behavior,
as observed in the susceptibility measurements. [92]
The magnetic exchange between metal centers through the vanadium oxide subunits is
also possible, giving rise to dimmeric or 1D systems. As it has been described previously, the
{Co(4Bpy)}(VO3)2 [41] and {Co(Bpa)}(VO3)2 [48] hybrid vanadates contains inorganic
{CoV2O6} layers interconnected by Co(4Bpy) and Co(Bpa) chains. The magnetic exchange
across the organic ligand is expected to be very weak, so the overall anti-ferromagnetic
interaction should be mainly attributed to the superexchange coupling within the inorganic
layers (see Figure 24, section 4.2).
The thermal evolution of the magnetic susceptibilities for {Co(4Bpy)}(VO3)2 and
{Co(Bpa)}(VO3)2 show broad maxima located at 36 and 20K, respectively. The m curves
increase below the maxima due to the existence of minor paramagnetic impurities (Figure
53).

Figure 53. m and mT curves for {Co(Bpa)}(VO3)2.

The shortest CoCo distance within the layers are 4.582 for {Co(4Bpy)}(VO3)2 and
4.825 for {Co(Bpa)}(VO3)2, and the connectivity between the metal centers across the O-
50 Edurne S. Larrea, Roberto Fernndez de Luis, Jos L. Mesa et al.

V-O atoms suggest the existence of dimmer like anti-ferromagnetic behavior. The cobalt(II)
atoms of these dimmer magnetic units are separated by 5.45 and 5.91 for
{Co(4Bpy)}(VO3)2 and 5.45 and 5.81 for {Co(Bpa)}(VO3)2 from the adjacent dimmers.
Despite the fact that the inter-dimmer distances are not very large, the magnetic pathway
involucres five atoms (Co-O-V-O-V-O-Co), and for this reason, the magnetic interaction
between the dimmeric units is not expected to be very strong. So, the thermal evolution of the
mT curve can be explained based on a dimmeric anti-ferromagnetic interactions plus the
spin-orbit coupling typical for Co(II) d7 cations.
Another example of magnetic exchange through the vanadium oxide subunit is the
[{Ni(H2O)2(Bpe)}(V4O11)]0.5H2O compound [39]. The crystal structure has been previously
described (see Section 4.1). The magnetic susceptibility shows a continuous increase of its
value in the 300 to 2K temperature range (Figure 54).

Copyright (2009) Wiley-VCH Verlag GmbH and Co. KGaA. Reproduced with permission.

Figure 54. m and mT curves for [{Ni(H2O)2(Bpe)}(V4O11)]0.5H2O. The lines are the best fit for a
Heisenberg isotropic antiferromagnetic S=1 chain model. Reproduced with permission from Ref. [39].

Reproduced with permission.

Figure 55. Magnetic exchange pathway for [{Ni(H2O)2(Bpe)}(V4O11)]0.5H2O. Reproduced with


permission from Ref. [39]. Copyright (2009) Wiley-VCH Verlag GmbH and Co. KGaA.
Hybrid Vanadates, towards Metal-Organic Frameworks 51

The NiNi shortest distance is of 6.26 through one VO5 polyhedra. The magnetic data
was first analyzed with a zero-field splitting model, typical for Ni(II) magnetically isolated d8
high spin axially distorted cations. The obtained values were out of range in comparison with
other systems exhibiting isolated Ni(II) cations, indicating the existence of a magnetic
exchange between the Ni(II) cations through the VO5 polyhedra. The VO5 polyhedra and
Ni(H2O)2N2O2 octahedra are corner-sharing, giving rise to the 1D chains observed in Figure
55.
The magnetic susceptibility was analyzed with Heisenberg isotropic anti-ferromagnetic
S=1 chain model. The best fitting gives rise to g=2.27 and J=-5.83K values, corroborating the
existence of a weak anti-ferromagnetic interaction between the metal centers through the
vanadate subunit.

5.3. Catalytic and Photocatalytic Activity of Hybrid Vanadates

The catalytic properties of the isostructural vanadates with formula M(HAep)2(VO3)4,


M= Co(II), Ni(II), Cu(II), Aep= 1-(2-amoethyl)piperazine, were studied. The phases
crystallize in the monoclinic space group P21/c, with very similar cell parameters. [93] These
compounds show a 2D crystal structure (Figure 56(a)). Vanadium(V) atoms are in tetrahedric
coordination environments. Each VO4 tetrahedra is linked to two other tetrahedra through two
of their vertex. This joint gives rise to metavanadate chains in the [010] direction. On the
other hand, the metallic ions, Co(II), Ni(II) or Cu(II), are octahedrally coordinated, corner
linked to two oxygen atoms of two vanadate chains in the axial positions and to four nitrogen
atoms of two HAep+ cations, in the equatorial plane. The organic molecule is chelating the
metal centers via one nitrogen atom of the pyperazine ring and the nitrogen of the
aminoethylic radical. The other nitrogen atom of the ring is protonated.

Figure 56. (a) Inorganic organic layers and (b) crystal structure of M(HAep)2(VO3)4, M= Co2+, Ni2+,
Cu2+ compounds. [93]

The joint between the octahedra and the vanadate chains through the oxo groups gives
rise to the sheets, parallel to the (100) plane (Figure 56(a)). These sheets are linked to each
others via hydrogen bonds between the N-H group belonging to the 1-(2-
aminoethyl)pyperazonium ligand and the oxygen atoms of the metavanadate chain. There are
intralayer hydrogen bonds also(Figure 56(b)).
52 Edurne S. Larrea, Roberto Fernndez de Luis, Jos L. Mesa et al.

The activity of this family of hybrid vanadates was probed towards oxidation of sulfides.
To note, this reaction is interesting both for the preparation of sulfoxides and sulfones, [94] as
well as for the desulfurization of fuels. [95] Various catalytic essays were performed with
different oxidizing agents (H2O2 and TBHP) and substrates (methyl phenyl sulfide, ethylbutyl
phenyl sulfide, p-chlorophenyl methyl sulfide and methyl p-tolyl sulphide).
First of all, the catalysts were tested for the oxidation of methyl phenyl sulfide with H2O2
as oxidizing agent. The behavior of all of the catalysts is very similar, reaching conversions
between 83 and 64 %, with selectivity rates towards the sulfoxide over 94 % (Figure 57).

Figure 57. Kinetic profile for the oxidation reaction of methyl phenyl sulfide with M(HAep)3(VO3)4,
M= Co2+, Ni2+, Cu2+. Reprinted from Ref. [93]. Copyright (2011), with permission from RSC.

The influence of the oxidizing agent was studied comparing the kinetic profiles of the
oxidation reactions of methyl phenyl sulfide with H2O2 and TBHP. The kinetic profiles are
very similar to each other, observing no loss in the conversion rate when increasing the steric
hindrance of the oxidizing agent.
On the other hand, when the steric hindrance of the substrate is increased, the kinetic
profile and the conversion and selectivity rates change. When a substituent is introduced in
the phenyl ring of the substrate (Cl, Me), an increase on the turn-over frequency (TOF) and
conversion rate is observed. However, the kinetic profile for ethylbutyl phenyl sulfide shows
an important induction period due to the steric hindrance of this substrate, which makes the
access to the active sites more difficult.
Some silver hybrid vanadates also present photocatalytic properties. A synergistic effect
between silver ions and vanadates that could improve their catalytic selectivity was observed
in these compounds.
H. Lin and P.A. Maggard have synthesized three silver vanadates with 4Bpy, Bpa and
pyrazinedicarboxylate (Pzc) as ligands, [42] whose formulas are:
[{Ag(4Bpy)}4(V4O12)]2H2O, [{Ag(Bpa)}4(V4O12)]4H2O and {Ag4(Pzc)2}(V2O6). As it is
described above, [{Ag(4Bpy)}4(V4O12)]2H2O and[{Ag(Bpa)}4(V4O12)]4H2O crystal
structures are composed of neutral {Ag4V4O12}n layers that are pillared via coordination by
the organic ligands, 4Bpy or Bpa, to the Ag(I) sites. The layers consist of isolated {V4O12}
rings that are linked together by eight Ag+ cations per ring. The third compound is composed
of a 3D {Ag2(Pzc)+}n metal-organic network that contains {VO3-}n double chains within its
open channels (Figure 58).
Hybrid Vanadates, towards Metal-Organic Frameworks 53

Figure 58. (a) Crystal structure of {Ag4(Pzc)2}(V2O6). (b) Inorganic chains of metavanadate and silver
edge shared chains.

On the other hand, Y. Han et al., have recently obtained a new silver(I) hybrid vanadate
with 1,4-Bis(N-imidazolyl)butane, Bbi, [Ag(Bbi)][{Ag(Bbi)}4{Ag3(V4O12)2}].2H2O. [96]
The phase crystallizes in the triclinic space group P-1 and the most fascinating feature is the
1D chains containing {Ag7}7+ clusters with AgAg interactions based on dicyclic rings of
[V4O12]4-.
All of these silver(I) vanadates show photocatalytic activity and were tested for the
decomposition of methylene blue (MB) under ultraviolet and/or visible light. MB is
commonly used as a representative of widespread organic dyes that contaminate textile
effluents and that are very difficult to decompose in waste streams. In Figure 59, changes in
the concentration of the aqueous MB solution vs. irradiation time with ultraviolet light were
plotted. All of them are active for the decomposition of MB under UV light irradiation
(Figure 59).

Figure 59. Photocatalytic decomposition of MB solution under UV light with the use of hybrid silver(I)
vanadates.
54 Edurne S. Larrea, Roberto Fernndez de Luis, Jos L. Mesa et al.

The photocatalytic rates of {Ag4(Pzc)2}(V2O6) and [Ag(bbi)][{Ag(bbi)}4{Ag3(V4O12)2}].


2H2O are notably higher than either [Ag(4Bpy)]4V4O12.2H2O or [Ag(dpa)]4V4O12.4H2O.
{Ag4(Pzc)2}(V2O6) is also active under visible light, decomposing an 80% of the MB after 3
h.

CONCLUSION
The crystal chemistry of hybrid vanadium oxides ranges from rigid frameworks, similar
to the zeotypes, to flexible and dynamic frameworks, comparable to MOFs. The analysis of
the crystal structures indicates that the use of multi-podal bridging ligands favors the
formation of 3D crystal structures. In these compounds, three- and two-dimensional inorganic
substructures are frequent. Hybrid vanadates containing 1D or discrete inorganic units are
scarce; however, their linkage through the organic ligand generates crystal structures similar
to those observed in MOFs.
With regard to the metal-organic substructure, the hybrid vanadates include more
commonly 1D chains or 2D layers. The interpenetration or polycatenation of the nets is
frequent in hybrid vanadates containing metal-organic layers or nets. In fact, the unique
examples of hybrid vanadates containing 3D metal-organic subnets exhibits polycatenated
and interpenetrated metal-organic scaffolds.
The great crystal chemistry diversity in hybrid vanadium oxides arises, in part, in the
possibility of obtaining different crystal architectures from the same synthetic system, by
modifying the hydrothermal conditions. The secondary metal center also influences the
obtained crystal architectures. In fact, usually isostructural compounds for nickel(II),
cobalt(II) and manganese(II) cations or for copper(I) and silver(I) have been obtained. The
choice of the organic ligand is determinant. Despite that the obtained crystal architectures are
similar for the different vanadates with multi-podal ligands, the thermal properties of these
materials change drastically depending on the flexibility or rigidity of the organic part. The
magnetic properties of these materials, directly depend on the connectivity between the metal
centers to each other and through the vanadium oxide and/or the ligand. Hybrid vanadates
could be active catalysts in oxidation reactions and for the decomposition of pollutants, such
as, methylene blue.

ACKNOWLEDGMENTS
This work has been financially supported by the M inisterio de Ciencia e Innovacin,
MICINN (MAT2010-15375) and the G obierno Vasco (IT-177-07), which we gratefully
acknowledge. Edurne S. Larrea and Roberto Fernndez de Luis wish to thank the Gobierno
Vasco and M inisterio de Ciencia e Innovacin for funding.

REFERENCES
[1] Natarajan, S.; Mandal, S. Angew. Chem. Int. Ed. 2008, 47, 4798-4828.
Hybrid Vanadates, towards Metal-Organic Frameworks 55

[2] Frey, G. Chem. Soc. Rev. 2008, 37, 191-214.


[3] Yaghi, O.M.; OKeefee, M.; Ockwig, N.W.; Chae, H.K.; Eddaoudi, M.; Kim, J. Nature,
2003, 423, 705-714.
[4] Janiak, C.; Vieth, J.K. New. J. Chem. 2011, 34, 2366-2388.
[5] Kitagawa, S.; Kitaura, R. Noro, S.I. Angew. Chem. Int. Ed. 2004, 43, 2334-2375.
[6] Schindler, M.; Hawthorne, F.C., Baur, W.H. Chem. Mater. 2000, 12, 1248-1259.
[7] (a) Allen, F.H.; Kennard, O. Chem. Des. Autom. News, 1993, 8, 31-37. (b) Allen, F.H.;
Motherwell W.D.S. Acta Crystallogr. 2002, B58, 407-422.
[8] Wells, A.F. Structural Inorganic Chemistry, Claredon Press, Oxford, 1984.
[9] Zavalij, P.Y.; Whittingham, M.S. Acta. Crystallogr. 1999, B55, 627-663.
[10] Hargrman, P.J.; Finn, R.C., Zubieta, J. Solid State Sciences, 2001, 3, 745-774.
[11] Ulick, B.L.; Pavelck, F.; Huml, K. Acta Crystallogr. 1987, C43, 2266-2268.
[12] Riou, D.; Frey, G. J. Solid State Chem. 1996, 124, 151-154.
[13] La Duca, R.L.; Rarig, R.S.; Zubieta, J. Inorg. Chem., 2001, 40, 607-612.
[14] Debord, J.R.D.; Zhang, Y.; Haushalter, R.C.; Zubieta, J.; O`Connor, C. J. Solid State
Chem. 1996, 122, 251-258.
[15] La Duca, R.L.; Ratkoski, R.; Rarig, R.S.; Zubieta, J. Inorg. Chem. Commun. 2001, 4,
621-625.
[16] Shi-Xiong, L.; Bi-Zhou, L.; Shen, L. Inorg. Chim. Acta 2000, 304, 33-37.
[17] Hargman, D.; Zubieta, J.; Warren, C.J.; Meyer, L.M.; Treacy, M.M.; Haushalter, R.C.
J. Solid State Chem. 1998, 138, 178-182.
[18] Hargman, P.; Zubieta, J. Inorg. Chem. 2000, 39, 3252-3260.
[19] Gerardin, C.; In, M.; Allouche, L.; Haouas, M.; Taullete, F. Chem. Mater. 1999, 11,
1285-1292.
[20] ivaje, J. Coor. Chem. Rev. 1998, 178, 999-1018.
[21] Bouhedja, L.; Steunou, N.; Maquet, J.; Livage, J. J. Solid State Chem. 2001, 162, 315-
321.
[22] Chirayil, T.; Zavalij, P.Y.; Whittingham, M.S. Chem. Mater. 1998, 10, 2629-2640.
[23] Harada, J.; Ogawa, K. Chem. Soc. Rev. 2009, 38, 2244-2252.
[24] Fernndez de Luis, R.; Urtiaga, M.K.; Mesas, J.L.; Rojo, T.; Arriortua, M.I. J. Alloy
Compd. 2009, 480, 54-56.
[25] Qu, X.; Xu, L.; Gao, G.; Li, F.; Yang, Y. Inorg. Chem. 2007, 46, 4775-4777.
[26] Long, D.L.; Blake, A.J.; Champness, N.R.; Wilson, C.; Schrder, M. Angew. Chem. Int.
Ed. 2001, 40, 2443-2447.
[27] Delgado-Friedrichs, O.; OKeeffe, M. J. Solid State Chem. 2005, 178, 2480-2485.
[28] Batten, S. Cryst. Eng. Comm., 2001, 3, 37-72.
[29] Carlucci, L.; Ciani, G.; Proserpio, D.M. Coord. Chem. Rev. 2003, 246, 247-289.
[30] Jiang, Y.-Q.; Xu, Z.-H.; Xie, Z.-X. J. Coord. Chem. 2008, 61, 1575-1581.
[31] Zheng, L.-M.; Wang X.; Wang, Y.; Jacobson, A.J. J. Mater. Chem. 2001, 11, 1100-
1105.
[32] Yan, B.; Luo, J.; Dube, P.; Sefat, A.S.; Greedan, J.E.; Maggard, P.A. Inorg. Chem.
2006, 45, 5109-5118.
[33] Larrea, E.S.; Mesa, J.L.; Pizarro, J.L.; Arriortua M.I.; Rojo, T. J. Solid State Chem.
2007, 180, 1149-1157.
56 Edurne S. Larrea, Roberto Fernndez de Luis, Jos L. Mesa et al.

[34] Larrea, E.S. Doctoral thesis, New Transition Metal Hybrid Vanadates. Hydrothermal
Synthesis, Structural Study and of their Spectroscopic and Magnetic Properties,
Universidad del Pas Vasco (UPV/EHU), 2009.
[35] Maggard, P.A.; Boyle, P.D. Inorg. Chem. 2003, 42, 4250-4252.
[36] Devi, R.N.; Rabu, P.; Golub, V.O.; OConnor, C.J.; Zubieta, J. Solid State Sci. 2002, 4,
1095-1102.
[37] Yang, L.; Hu, C.; Naruke, H.; Yanase, T. Acta Crystallogr. Sect. C: Cryst. Struct.
Commun. 2001, 57, 799-801.
[38] Khan, M.I.; Yohannes, E.; Nome, R.C.; Ayes, S.; Golub V.O.; OConnor, C.J.;
Doedens, R.J. Chem. Mater. 2004, 16, 5273-5279.
[39] Fernndez de Luis, R.; Mesa, J.L.; Urtiaga, M.K.; Rojo, T.; Arriortua, M.I. Eur. J.
Inorg. Chem. 2009, 4786-4794.
[40] Fernndez de Luis, R., Urtiaga, M.K.; Mesa, J.L.; Aguayo, A.T.; Rojo, T.; Arriortua,
M.I. Cryst. Eng. Comm. 2010, 10, 1880-1886.
[41] Zheng, L.-M.; Wang, X.; Wang, Y.; Jacobson, A.J. J. Mater. Chem. 2001, 11, 1100-
1105.
[42] Lin, H.; Maggard, P.A. Inorg. Chem. 2008, 47, 8044-8052.
[43] Yan, B.; Maggard, P.A. Inorg. Chem. 2007, 46, 6640-6646.
[44] Zhang, C.-D.; Liu, S.-X.; Xie, L.-H.; Gao, B.; Sun, C.-Y.; Li, D.-H. J. Mol. Struct.
2005, 753, 40-44.
[45] Fernndez de Luis, R.; Urtiaga, M.K.; Mesa, J.L.; Vidal, K.; Lezama, L.; Rojo, T.;
Arriortua, M.I. Chem. Mater. 2010, 22, 5543-5553.
[46] Wang, C.-C. Acta Crystallogr., Sect. E: Struct. Rep. Online 2007, 63, m2233.
[47] Wang, C.-C.; Yin, C. Z. Kristallogr. New Cryst. Struct. 2008, 223, 13-15.
[48] Fernndez de Luis, R. Doctoral thesis Heterometallic Vanadates Based on Dipodal
Ligands Self-Assembly Universidad del Pas Vasco (UPV/EHU), 2009.
[49] Wang, X.; Hu, H.-L.; Chen, B.-K.; Liu, G.-C.; Li, J. Z. Anorg. Allg. Chem. 2009, 635,
2692-2696.
[50] Lan, Y.-Q.; Li, S.-L.; Wang, X.-L.; Shao, K.-Z.; Du, D.-Y.; Su, Z.-M.; Wang, E.-B.
Chem. Eur. J. 2008, 14, 9999-10006.
[51] Wang, X.-L.; Chen, B.-K.; Liu, G.-C.; Lin, H.-Y.; Hu, H.-L. J. Organomet. Chem.
2010, 695, 827-832.
[52] Qi, Y.-F.; Xiao, D.; Wang, E.; Zhang, A.; Wang, X. Aust. J. Chem. 2007, 60, 871-878.
[53] Wang, X.; Chen, B.; Liu, G.; Lin, H.; Hu, H.; Chen, Y. J. Inorg. Organomet. Polym.
Mater. 2009, 19, 176-180.
[54] Khan, M.I.; Deb, S.; Doedens, R.J. Inorg. Chem. Commun. 2006, 9, 25-28.
[55] Zhang, L.-J; Hu, Z.-J.; Zang, Z.; Gou, H.-Y. Chin. J. Chem. 2007, 25, 566-569.
[56] Qu, X.; Xu, L.; Li, F.; Gao, G.; Yang, Y. Inorg. Chem. Commun. 2007, 10, 1404-1408.
[57] Qu, X.; Xu, L.; Gao, G.; Li, F.; Yang, Y. Inorg. Chem. 2007, 46, 4775-4777.
[58] Qi, Y.-F.; LV, C.-P.; Li, Y.-G.; Wang, E.-B.; Li, J.; Song, X.-L. Inorg. Chem. Commun.
2010, 13, 384-387.
[59] Ouellette, W.; Zubieta, J. Solid State Sci. 2007, 9, 658-663.
[60] Ouellette, W.; Burkholder, E.; Manzar, S.; Bewley, L.; Rarig, R.S.; Zubieta, J. Solid
State Sci. 2004, 6, 77-84.
[61] Devi, R.N.; Rabu, P.; Golub, V.O.; OConnor, C.J.; Zubieta, J. Solid State Sci. 2002, 4,
1095-1102.
Hybrid Vanadates, towards Metal-Organic Frameworks 57

[62] Rarig, R.S Jr.; Zubieta, J. Dalton Trans. 2003, 1861-1868.


[63] Li, J.-R.; Yu, Q.; Sanudo, E.C.; Tao, Y.; Song, W.-C; Bu, X.-H. Chem. Mater. 2008,
20, 1218-1220.
[64] Willians, I.D.; Law, T.S.; Sung, H.H.-Y.; Wen, G.-H.; Zhang, X.-X. Solid State Chem.
2000, 2, 47-55.
[65] Wang, X.; Liu, L.; Jacobson, A.J.; Ross, K. J. Mater. Chem. 1999, 9, 859-861.
[66] Shi, Z.; Zhang, L.; Zhu, G.; Yan, G.; Huar, J.; Ding, H.; Feng, S. Chem. Mater. 1999,
11, 3565-3570.
[67] Ou. G.-C.; Jiang, L.; Feng, X.-L.; Lu, T.-B. Dalton Trans. 2009, 71-76.
[68] Lu, J.; Wang, E.; Chen, J.; Qi, Y.; Hu, C.; Xu, L.; Peng, J. J. Solid State Chem. 2004,
177, 946-950.
[69] Xiao, D.; Wang, E.; An, H.; Xu, L.; Hu, C. J. Molec. Struc. 2004, 707, 77-81.
[70] Xiao, D.; Xu, Y.; Hou, Y.; Wang, E.; Wang, S.; Li, Y.; Xu, L.; Hu, C.; Eur. J. Inorg.
Chem. 2004, 1385-1388.
[71] Guillou, N.; Livage, C.; Ferey, G. Eur. J. Inorg. Chem. 2006, 24, 4963-4978.
[72] Lu, Y.; Wang, E.; Yuan, M.; Li, Y.; Xu, L.; Hu, C.; Hu, N.; Jia, H. Solid State Science
2002, 4, 449-453
[73] Liu, C.-M.; Zhnag, D.-Q.; Xiong, M.; Dai, M.-Q.; Hu, H.-M., Zhu, D.-B. J. Coord.
Chem. 2002, 55, 1327-1335.
[74] Gao, E.-Q.; Cheng, A.-L.; Xu, Y.-X.; He, M-Y.; Yan, C.-H. Inorg. Chem. 2005, 44,
8822-8835.
[75] Yaghi, O.M.; OKeeffe, M.; Ockwig, N.W.; Chae, H.K.; Eddaoudi, M.; Kim, J. Nature,
2003, 423, 705-714.
[76] (a) Shiels, R.A.; Venkatasubbaiah, K.; Jones, C.W. Adv. Synth. Catal., 2008, 350,
2823-2834; (b) Zhang,A.; Zhang, J.; Cui, N.; Tie, X.; An, Y.; Li, L. J. Mol. Catal. A-
Chem., 2009, 304, 28-32; (c) Katsoulis, D.E. Chem. Rev., 1998, 98, 359-387; (d) Centi,
G.; Trifiro, F. Appl. Catal. A: General, 1996, 143, 3-16.
[77] (a) Weyand, M.; Hecht, H.J.; Kiess, M.; Liaud, M.F.; Vilter, H.; Schomburg, D., J.
Mol. Biol., 1999, 293, 595-611; (b) Carter-Franklin, J.N.; Parrish, J.D.; Tchirret-Guth,
R.A.; Little, R.D.; Butler, A., J. Am. Chem. Soc., 2003, 125, 3688-3689; (c) Rechder,
D.; Santoni, G.; Licini, G.M.; Schilzke, C.; Meier, B. Coord. Chem. Rev., 2003, 237,
53-63.
[78] Butler, A. Coord. Chem. Rev., 1999, 187, 17-35.
[79] Avtamonova, N.V.; Trunov, V.K.; Bezrukov, I.Ya. Izvestiya Akademii Nauk SSSR,
Neorganicheskie Materialy, 1990, 26, 346-349.
[80] Fernndez de Luis, R.; Mesa, J.L.; Urtiaga, M.K.; Lezama, L.; Arriortua, M.I.; Rojo, T.
New. J. Chem. 2008, 32, 1582-1589.
[81] Barrera, G.D.; Bruno, J.A.O.; Barron, T.H.K.; Allan, N.L. J. Phys.: Condens. Matter.
2005, 17, R217-R252.
[82] Miller, W.; Smith, C.W.; Mackenzie, D.S.; J. Mater. Sci. 2009, 44, 5441-5451.
[83] Evans, J.S.O. J. Chem. Soc., Dalton Trans. 1999, 33173326.
[84] Evans, J.S.O.; Mary, T.A.; Sleight, A.W. Physica B, 1998, 311-316.
[85] Amri, M.; Walton, R.I.W. Chem. Mater. 2009, 21, 33803390.
[86] Matsuda, T.; Kim, J.E.; Ohoyama, K.; Moritomo, Y. Physical Review B, 2009, 79,
172302-1 -172302-4.
[87] Chapman, K.W.; Chupas, P.J. J. Am. Chem. Soc. 2007, 129, 10090-10091.
58 Edurne S. Larrea, Roberto Fernndez de Luis, Jos L. Mesa et al.

[88] Han, S.S.; Goddard, W.A.; J. Phys. Chem. C, 2007, 111, 15185-15191.
[89] Wu, Y.; Kobayashi, A.; Halder, G.J.; Peterson, V.K.; Chapman, K.W.; Lock, N.;
Southon, P.D.; Kepert, C. J. Angew. Chem. Int. Ed. 2008, 47, 8929 8932.
[90] Harada, J.; Ogawa, K. Chem. Soc. Rev. 2009, 38, 22442252.
[91] Nanda, K.K.; Thompson, L.K.; Bridson, J.N.; Nag, K. J. Chem. Soc., Chem. Commun.
1994, 1337-1341.
[92] Zeng, M.H.; Wang, B.; Wang, X.Y.; Zhang, W.X.; Chen, X.M.; Gao, S. Inorg. Chem.
2006, 45, 7069-7076.
[93] Larrea, E.S.; Mesa, J.L.; Pizarro, J.L.; Iglesias, M.; Rojo, T.; Arriortua, M.I. Dalton
Trans. DOI: 10.1039/C1DT10971E.
[94] Conte, V.; Bortolini, O. in The Chemistry of Peroxides Transition Metal Peroxides.
Synthesis and Role in Oxidation Reactions, ed. Rappoport, Z. Wiley Interscience, 2006.
and references cited therein; Conte, V.; Floris, B. Inorg. Chim. Acta, 2010, 363, 1935-
1946.
[95] Song, C. Catal. Today, 2003, 86, 211-263.
[96] Hu, Y.; Luo, F.; Dong, F. Chem. Commun. 2011, 47, 761-763.
In: Coordination Polymers and Metal Organic Frameworks ISBN: 978-1-61470-899-5
Editors: O. L. Ortiz and L. D. Ramirez, pp. 59-98 2012 Nova Science Publishers, Inc.

Chapter 2

STRUCTURE AND MAGNETIC PROPERTIES


OF MONO - AND POLYNUCLEAR COMPLEXES
CONTAINING RHENIUM(IV)

Carlos Kremer and Ral Chiozzone


Ctedra de Qumica Inorgnica, Departamento Estrella Campos, Facultad
de Qumica, Universidad de la Repblica, Montevideo, Uruguay.

ABSTRACT
One of the major goals in inorganic supramolecular chemistry today is the design of
polynuclear coordination arrays and the study of their magnetic properties. With the
generation of well-defined architectures it is possible to understand the different factors
which determine the exchange coupling between spin carriers. Most of the results found
in the literature are focused on polynuclear complexes containing metal ions belonging to
the first transition series. Once the magnetic interaction between 3d metal ions is well
understood, the study of those systems containing 4d or 5d metal ions becomes very
interesting. In this review we revise the structure and magnetic properties of Re(IV)
complexes. Rhenium(IV), a 5d3 ion, usually forms octahedral complexes which are
reasonably stable against redox processes and inert to ligand substitution. This is the basis
for the preparation of mononuclear species that can act as ligands towards first-row
transition metal ions. For example, complexes containing dicarboxylic ligands,
[ReX4(ox)]2 and [ReX4(mal)]2 (X = Cl, Br; ox = oxalato; mal = malonato), or N-donor
ligands, [ReCl5(pyz)] (pyz = pyrazine) have been used as building blocks to construct
heteropolynuclear complexes. The different designed structures, from discrete binuclear
complexes to extended chain-like compounds, are reviewed in this work. In addition, the
magneto-structural studies of these mono- and polynuclear complexes are also included
and discussed.

INTRODUCTION
The use of transition metal complexes to construct predictable multidimensional infinite
networks is an area of chemistry which has received increasing attention over the last years
60 Carlos Kremer and Ral Chiozzone

[1, 2]. The formation of such polynuclear complexes is a self-assembling process based on
complementary and specific interactions between building blocks which generate the final
product. Mononuclear complexes with different metal ions, ligands, geometries, etc. are used
as building blocks in the so-called com plexes as ligands approach. The polynuclear
complexes have found many applications in several areas of interest. The variety of solid
networks that can be designed and modulated allowed the synthesis of metal-organic
frameworks (MOFs) with specific properties in gas storage, catalysis, molecular recognition,
etc. [3-8].
On the other hand, the interaction of the metal ions in the supramolecular network
produces functional solid materials by a cooperative effect. The design of solids with a
negative thermal expansion, molecular based magnets, light-conversion devices and nonlinear
optical materials are good examples [9].
Magnetism in polynuclear complexes can be implemented by incorporating paramagnetic
metals. If these spin carriers present some kind of interaction (exchange), a cooperative
magnetic phenomenon can be obtained. This is often achieved in polynuclear complexes and
MOFs [8, 10-12]. So far the majority of magnetic frameworks are those containing first-row
transition metals (V, Cr, Mn, Fe, Co, Ni, Cu) and most of the study of the interaction between
magnetic centers has been restricted to them [13]. Heavier elements have been largely ignored
in this respect. The large splitting of d orbitals and the smaller interelectronic repulsion in
their mononuclear complexes account for the low-spin configuration they exhibit. This means
that ions with an even number of d electrons are usually diamagnetic. The polynuclear
complexes are also frequently diamagnetic because of the tendency to form metal-metal
bonds which pairs the electrons. These facts account for the less explored magnetic properties
of 4d and 5d metal ions. In spite of this, the greater diffuseness and spin-orbit coupling of the
magnetic orbitals compared to those of 3d metal ions, make the study of heavy metal
polynuclear complexes a very interesting field.
This work reviews the structural aspects and magnetic properties of Re(IV) complexes. In
octahedral environments, the ground state of this metal ion is a 4A2g term with three unpaired
electrons (5d3 electronic configuration). Re(IV) complexes are reasonably stable against
redox processes and are reluctant to ligand substitution. These conditions together with the
adequate choice of bridging ligands allow the preparation of polynuclear complexes.
Mononuclear Re(IV) complexes are then used as ligands towards 3d paramagnetic ions. The
structure and magnetic interaction of the heteropolynuclear compounds are also
comparatively discussed in this review.

RHENIUM(IV)
Coordination compounds of rhenium with the metal ion in the oxidation states +1 up to
+7 have been characterized and reported. Different reviews cover the relevant rhenium
coordination chemistry [14-16].
Rhenium(IV) complexes are relatively scarce if compared with those of rhenium(III) or
the most representative rhenium(V). One of the reasons is the tendency to hydrolyze when
exposed to water. This fact has been recognized and studied many years ago in the simple
hexahalorhenate(IV) complexes [ReX6]2 (X = Cl, Br) which finally yields ReO2 rather
Structure and Magnetic Properties of Mono - and Polynuclear Complexes 61

quickly [17]. The stepwise formation constants (K6) of hexacloro- and hexabromorhenate(IV)
at 15 C, I = 3M HClO4 are 2.20 x 106 and 1.8x105 respectively, indicating a better stability
for the chloro analogue. In this line, the exchange rates for halide isotopic exchange at 60 C
are very low, 1.29 x 104 and 81.6 x 104 M1 h1 respectively. The relative stability
(thermodynamic and kinetic) seems to be originated in the stronger Re-X bond interaction for
X = Cl (stretching force constants 1.62 and 1.35 mdyne 1 for Re-Cl and Re-Br,
respectively) [18].
On the other hand, this is an intermediate oxidation state with the possibility to be either
oxidized or reduced. The standard reduction potentials E (ReO4/ReO22H2O) and E
(ReO22H2O/Re) are 0.49 and 0.29 V respectively [19]. This accounts for the general
tendency of Re(IV) complexes to be oxidized and/or reduced under mild conditions. ReO4 is
a common undesirable byproduct in the synthesis of Re(IV) complexes if they are not
performed under inert atmosphere [20]. The easy reduction to Re(III) complexes can be also
verified, as reported for the compounds [ReX4(bpym)] (X = Cl, Br; bpym = 2,2-
bipyrimidine). The reversible reduction of [ReIVX4(bpym)] to [ReIIIX4(bpym)] occurs at
+0.19 V (vs. NHE) in acetonitrile. These Re(IV) complexes can also be oxidized to Re(V)
compounds in a non reversible process [21].
The most numerous Re(IV) complexes are octahedral halo species of the type [ReX6]2,
[ReX5L], [ReX4L2], and [ReX4(L)], where L and L are monodentate and bidentate neutral
ligands, respectively [14-16]. The heptacyano anion [Re(CN)7]3 is a remarkable exception
containing seven-coordinate metal ion [22, 23].
Mononuclear Re(IV) complexes can be prepared by different synthetic routes. Many of
them include a redox process. Examples are the reduction of Re(VII) or Re(V) compounds,
and the oxidation of lower state rhenium complexes [14-16]. Indeed, Re(IV) complexes are
often isolated as unexpected byproducts in many reactions of Re(V) or Re(III) compounds.
The reaction of [ReVOCl2(C2H5O)(PPh3)2] (Ph = phenyl) with acetylacetone gives
[ReIVCl2(acac)2] (acac = acetylacetonate) [24]. Treatment of the dinuclear [ReII2Cl4(PEt3)4]
(Et = ethyl) with H2 in dichloromethane produces a mixture of complexes containing trans-
[ReIVCl4(PEt3)2] and [PEt4][ReIVCl5(PEt3)] [25]. [ReIVCl4(bipy)] (bipy = 2,2-bipyridine) is
directly obtained from ReCl5 [26]. cis-[ReIV(NCS)4(Ph3P)(Ph3PO)] is formed during the
reaction of [ReVOCl3(Ph3P)2] and Me3SiNCS (Me = methyl) [27], while [ReIVCl4(dppom-
P,O)] (dppom-P,O = Ph2PCH2P(=O)Ph2) and [ReIVCl4(OPPh3)2] are also obtained from the
same Re(V) precursor [28, 29]. The reaction of [ReVNCl2(PPh3)2] with Me3SiNCS produces
trans-[ReIV(OH)(NCS)3(Ph3P)2]MeOH [30]. Other starting compounds have been mer-
[ReVOBr3(bipy)] [31] and KReO4 [32].
The simpler anions [ReX6]2 have been recently exploded as precursors of rhenium(IV)
complexes, in spite of their known inertness to substitution [23, 33-47]. The reactions are
performed in nonaqueous solvents with excess of the incoming ligand. The complete halogen
substitution is a kinetically hindered event, hence mixed-ligand complexes are normally
obtained. Alternatively, direct ligand substitution reactions on other Re(IV) parent complexes
as cis-[ReCl4(CH3CN)2], trans-[ReCl4(PPh3)2] or cis-[ReCl4(thf)2] (thf = tetrahydrofurane)
have been accomplished as well [48-53].
62 Carlos Kremer and Ral Chiozzone

SYNTHESIS AND STRUCTURE OF HEXAHALORHENATES


Hexachlororhenate is the most common Re(IV) complex. It can be conveniently prepared
by the reduction of [ReO4] with H3PO2 in HCl [54, 55]. Different salts have been isolated
and characterized. Those whose structure has been determined either by x-ray or neutron
diffraction include:

a) salts with discrete small cations, K2[ReCl6] [56, 57], Cs2[ReCl6] [58], (NH4)2[ReCl6]
[59, 60], Se4[ReCl6] [61], Ag2[ReCl6] [62],
b) salts with larger organic cations, (NR4)2[ReCl6] (R = Me, Et) [63, 64],
(PPh4)2[ReCl6]2CH3CN [65], (MePPh3)2[ReCl6] [66], (PPN)2[ReCl6] (PPN =
(triphenylphosphine)iminium) [67], (DTF)2[ReCl6]Cl23H2O (DTF =
dithiobisformamidium) [68], -(BEDT-TTF)4[ReCl6]C6H5CN (BEDT-TTF =
bis(ethylenedithio)tetrathiafulvalene) [69], (p-rad)2[ReCl6] and (m-rad)2[ReCl6] (p/m-
rad = 2-(4/3-N-methylpyridinium)-4,4,5,5-tetramethyl-4,5-dihydro-1H-imidazol-1-
oxyl-3-N-oxide) [70],
c) salts with protonated bases as cations, (HB)2[ReCl6] (B = p-toluidine, pyridine,
quinoline, 2,2-bipyridine, 18-crown-6) [71-74], (H4cyclam)[ReCl6]Cl24dmso
(cyclam = tetraazacyclotetradecane; dmso = dimethylsulfoxide) [75],
(H4biim)[ReCl6]4H2O (H2biim = 2,2-biimidazole) [76], and
d) salts with complex cations, [OsIIICl(NH3)5][ReCl6] [77], [MIIICl(NH3)5]2[ReCl6]Cl2
(M = Ru, Ir [78, 79], [Fe(C5R5)2]2[ReCl6] (R = H, Me) [80],
[ReIIICl2(CH3CN)4]2[ReCl6]2CH3CN [60], and [cis-Re(CH3CN)4(CO)2]2[ReCl6]
[81].

[ReBr6]2 can also be prepared by reduction with H3PO2 [54]. The reported structures
with this anion is limited to M2[ReBr6] (M = K, Rb, Cs, NH4, PPh4) [58, 82-85],
[Fe(C5H5)2]2[ReBr6] [80], [RhIIICl(NH3)5][ReBr6] [86], and
(C15H13N4)4[ReVBr4O(H2O)]2[ReBr6]Br42H2O [87].
Different salts of hexaiodorhenate have been prepared by direct reduction of perrhenate
or perrhenic acid with concentrated HI [88-90]. Only two compounds with this anion are fully
characterized, K2[ReI6] and (NH4)2[ReI6] [90].

Figure 1. Structure and average distance Re-X in the [ReX6]2 anions.


Structure and Magnetic Properties of Mono - and Polynuclear Complexes 63

The hexafluororhenate analogue has been very scarcely studied. M2[ReF6] salts (M = Na,
K, NH4, Rb, Cs) have been prepared by a solid-state reaction in a mixture M2[ReI6]:MHF2
[91]. K2ReF6 is the only crystal structure reported for this anion [92].
From a structural point of view, the [ReX6]2 compounds present the rhenium(IV) center
surrounded by six halide anions in a regular or slightly distorted octahedral environment. The
average Re-X distances follow the expected trend, being larger for I (Figure 1).

MAGNETIC PROPERTIES OF HEXAHALORHENATE SALTS


Hexahalorhenate(IV) salts are among the first examples of magnetic studies on
rhenium(IV) complexes. Actually, they comprise a very numerous family of compounds that
behave magnetically in many different ways. For the purpose of ordering the presentation of
individual cases, we discuss first the magnetic properties of the hexahalorhenate salts of
diamagnetic cations and later we consider the possibility of additional magnetic contributions
from the presence of paramagnetic counter ions.

Hexahalorhenate(IV) Salts of Diamagnetic Cations with Magnetically


Isolated Anions

This situation has been found in a group of M2[ReX6] salts with bulky cations, namely
(HNMe3)2[ReCl6], (NMe4)2[ReCl6] [93], (NBu4)2[ReCl6] (Bu = n-butyl), (AsPh4)2[ReCl6]
[33], (PPh4)2[ReCl6]2MeCN [65], (PMePh3)2[ReCl6] [66], (HNMe3)2[ReBr6],
(NMe4)2[ReBr6] [94], (NMe4)2[ReI6] [95] and (AsPh4)2[ReI6] [90]. The cations in these salts
are big enough as to magnetically dilute the Re(IV) centers and to preclude any significant
interaction between them, even at the lowest studied temperatures. The characteristic
magnetic behavior of these compounds is shown in Figure 2, where the usual MT product is
plotted against temperature T for a polycrystalline sample of (AsPh4)2[ReCl6], M being the
molar susceptibility. At room temperature, the value of MT is close to 1.66 cm3 mol1 K, as
expected for an isolated mononuclear Re(IV) ion with g 1.85, calculated as:

N 2 2 (1)
MT g S ( S 1)
3k

where N is the Avogadro number, is the Bohr magneton, k is the Boltzmann constant and S
is the spin quantum number equal to 3/2. Here, the influence of the orbital momentum has
been incorporated into the effective g-factor, related with the gyromagnetic ratio ge for the
free electron as:

(2)
g g e 1 4

with ' being the total spin-orbit coupling constant reduced by electron delocalization and
the crystal field splitting parameter.
64 Carlos Kremer and Ral Chiozzone

Figure 2. Thermal dependence of MT for (AsPh4)2[ReCl6] (circles) and K2[ReCl6] (squares). The inset
shows the low temperature range for (AsPh4)2[ReCl6]. The solid lines are the calculated curves (see
text) [33].

The MT values remain practically constant upon cooling until 30 K, but at lower
temperatures, they decrease and tend to a finite value of ca. 1 cm3 mol1 K at 0 K. This
temperature dependence of the magnetic moment can be fully explained in terms of the zero-
field splitting [33].
The combined effect of second-order spin-orbit coupling and the presence of a ligand-
field component of symmetry lower than cubic responsible for the zero-field splitting are
represented in Figure 3, where the simplest case of tetragonal distortion of the octahedral
geometry is shown.

Figure 3. Schematic representation of the term splitting in a Re(IV) ion under tetragonal distortion of
the octahedral geometry and spin-orbit coupling. The number of microstates is shown in parenthesis.
Structure and Magnetic Properties of Mono - and Polynuclear Complexes 65

For a d3 electronic configuration in an octahedral environment, the first excited state is


the T2g term, arising from the 4F free-ion ground state. Under a tetragonal distortion, this
4

state is split into an orbital singlet 4B2 and an orbital doublet 4E at energies || and ,
respectively. Then, the fourfold degeneracy of the 4A2g is partially removed into the MS =
1/2 and MS = 3/2 levels, with the latter 2D in energy above the former. In this case, the 4T2g
splitting is related to the zero-field splitting of the ground state through Eq. 3, where is the
spin-orbit coupling parameter:

1 1 (3)
2D 8 2

||

The zero-field splitting is a source of paramagnetic anisotropy, that is, of susceptibilities that
differ as the external magnetic field is rotated with respect to the principal axis of the
molecule, and formally, the corresponding spin Hamiltonian can be written as:

H zfs S D S (4)

D being a real, symmetric tensor, that can be diagonalized if the coordinates x, y, z are chosen
parallel to its eigenvectors. Then the Hamiltonian takes the form:

1 (5)
H zfs D S 2z S ( S 1) E (S 2x S 2y )
3

where Sx, Sy, Sz are the spin operators, and D, E are the axial and rhombic zero-field splitting
parameters, respectively, which can be related to the diagonal elements of the D-tensor
through:

3 and 1 (6)
D Dzz E ( Dxx Dyy )
2 2

with

E 1
0
D 3

In axial symmetry, E = 0 and only the D parameter is needed to express the energies of the
(2S + 1) spin levels of the S multiplet. At this level of approximation, the |MS> and |MS>
states remain degenerate in pairs, which are called Kramer doublets. D can be positive or
negative: in the former case the levels with lowest |MS| are the most stable. The spin
Hamiltonian taking into account the Zeeman perturbation in axially distorted surroundings is
then:
66 Carlos Kremer and Ral Chiozzone

1 (7)
H D S2z S ( S 1) g||H zS z g ( H xS x H yS y )
3

where Hx, Hy, Hz are the directions of the magnetic field and g|| and g are the components of
the g-factor in the parallel (z) and perpendicular (x, y) directions.
The expressions of the magnetic susceptibilities in Eqs. 8 and 9 are easily derived from
the Hamiltonian in Eq. 7:

Ng ||2 2
|| F|| (8)
4k T

Ng 2 2
F (9)
4k T

where

1 9 exp( 2 D kT ) (10)
F||
1 exp( 2 D kT )

and

4 (3kT D)[1 exp( 2 D kT )] (11)


F
1 exp( 2 D kT )

The average magnetic susceptibility is finally approximated by:

1
M ( || 2 ) (12)
3

It is interesting to note that the limit of MT when T approaches zero, substituting Eqs. 8 - 11
in Eq. 12 and letting g = g|| = g, is:

3N 2 2
MT g (13)
4k

a value close to 1.00 cm3 mol1 K if g is taken as 1.88.


Least-squares fitting of the experimental data through Eq. 12 leads to |D| = 13(2) cm1
and g|| = g = 1.87(1) for (AsPh4)2[ReCl6]. Unfortunately, the sign of D cannot be
unambiguously determined from magnetic susceptibility measurements of microcrystalline
powder samples. As can be noted in Figure 2, the quality of the fit for (AsPh4)2[ReCl6] is very
good even at very low temperatures without considering any intermolecular magnetic
interactions. Values of D and g for other salts of this group are resumed in Table 1.
Structure and Magnetic Properties of Mono - and Polynuclear Complexes 67

The high value of the zero-field splitting is a recurrent signature of all Re(IV) compounds
and it contrasts with the lower values found for the first-row transition metal ions. This
difference can be mainly attributed to the large value of the spin-orbit parameter of this third-
row transition element when compared with the same parameter for lighter metals ( values
of 90 and 1100 cm1 for the d3 Cr(III) and Re(IV) free ions, respectively) [96].

Table 1. Calculated magnetic data for some magnetically isolated hexahalorhenates

Compound |D|/cm1 g|| g g Ref.


(AsPh4)2[ReCl6] 13(2) 1.87(1) 1.87(1) - [33]
(NBu4)2[ReCl6] 9(1) 1.87(1) 1.87(1) - [33]
(PMePh3)2[ReCl6] 14(2) - - 1.83a [66]
(AsPh4)2[ReI6] 24.9 - - 1.84 [90]
a
Calculated from the room temperature MT value using Eq. 1.

Hexahalorhenate(IV) Salts of Diamagnetic Cations with Significant Magnetic


Interaction between Anions

The difference between the magnetic behavior of K2ReCl6 and (AsPh4)2[ReCl6] is well
illustrated in Figure 2. For the former compound the MT value (and so the magnetic moment)
at room temperature is much lower than expected for isolated S = 3/2 ions and tend to vanish
as temperature approaches to zero [33]. This kind of experimental dependence of the
magnetic moment with the temperature, was early noted for several simple salts of the
[ReX6]2 anions (X = F, Cl, Br, or I) in the range 80300 K. In these compounds, the
magnetic moments were found to continuously decrease with lowering temperature and
antiferromagnetic interactions between Re(IV) centers were then proposed as the origin of
this behavior [96].
The thermal dependence of M and its inverse M1 for K2[ReCl6] is shown in Figure 4
[97]. The maximum in the susceptibility curve at Tmax = 13.0 0.5 K suggest the onset of
antiferromagnetic ordering, what was confirmed by a -shape anomaly in the heat capacity at
11.9 0.1 K [98]. The extrapolated value of M at T = 0 is about 0.88 its value at Tmax, in the
range expected for a three-dimensional antiferromagnet [99]. Neutron diffraction
measurements have found that below the Neel temperature, the magnetic moments are
ferromagnetically aligned in (100) planes, and the magnetic moments in adjacent planes are
oriented antiparallel to each other [100]. At temperatures well above Tmax the susceptibility
obeys a simple Curie-Weiss law with a negative value:

T
1 (14)
C

The negative value supports that a significant antiferromagnetic interaction takes place in
K2ReCl6. But even if the magnetic exchange is the dominant effect, the zero-field splitting is
still present. Therefore the parameter includes both intermolecular interactions and the zero-
68 Carlos Kremer and Ral Chiozzone

field splitting. Unfortunately, the independent evaluation of both factors is not possible due to
the lack of an adequate theoretical model to analyze this kind of cases.

Figure 4. Thermal dependence of M (circles) and M1 (squares) for K2[ReCl6]. The solid line is the
Curie-Weiss fit at temperatures well above Tmax [97].

Magnetic relevant data for several salts are included in Table 2. The reported Tmax values
correspond to maxima in the thermal dependence of the susceptibility, and they are not
necessarily associated with a cooperative-type transition to an antiferromagnetic ordered
state, which should be verified or ruled out by additional experimental measurements.
Comparison of these Tmax suggest that magnetic interactions roughly increase in the order
[ReCl6]2 < [ReBr6]2 < [ReI6]2, even when different ionic salts may have different crystal
structures. The intermolecular contacts of the type Re-XX-Re mediate the magnetic
interaction between the paramagnetic centers. The magnetic exchange in these systems
depends on the tendency of the metal electrons to be transferred to the X ligand and also on
the XX separation between the anions in the lattice. The magnetic coupling can be
understood taking into account that a great degree of -bonding occurs in the third-row
transition metal ions, which is accompanied by an increased chance of finding the magnetic
electron on the ligand atoms. In this respect, a larger degree of spin delocalization on the
halogen atoms will be expected as their atomic number increases, and this would account for
the trend previously mentioned. A relative large spin delocalization on the peripheral ligands
in the compounds with third-row transition metal ions when compared with the compounds
containing first-row transition metal cations, was first suggested by Owen and Steven in 1953
on the basis of EPR measurements [102] and corroborated later by polarised neutron
diffraction measurements on neutral tetrachlorobis(N-phenylacetamidine)rhenium(IV) [103,
104] and DFT calculations of spin densities on the [ReCl4(ox)]2 ion (ox = oxalato) [33]. This
effect accounts for the relatively important through space magnetic interactions in this class
of compounds.
On the other hand, the identity of the cation in this group of salts also influences the
magnetic properties, as they determine the distances of neighbor spin carriers in the crystal
Structure and Magnetic Properties of Mono - and Polynuclear Complexes 69

structure. A neat example is the lowering of the critical temperature as cation size increases in
the series Li2[ReI6] Na2[ReI6] K2[ReI6] Rb2[ReI6] Cs2[ReI6]. However, salts with
non-spherical cations may not exhibit so simple trends.

Table 2. Magnetic data for hexahalorhenates with anion-anion


antiferromagnetic interaction

Compound Tmax/K /K Ref.


13.0 0.5 100 5 [97]
K2[ReCl6]
12.4 0.5 55 5 [101]
Ag2[ReCl6] 26 102.6 [62]
(MeNH3)2[ReCl6] 3.8 23.2 [93]
(Me2NH2)2[ReCl6] 9.8 9.8 [93]
(Hpy)2[ReCl6] 4.25 9.78 [72]
(Hqy)2[ReCl6] 3.57 9.13 [72]
(Hacr)2[ReCl6] 3.25 5.38 [72]
K2[ReBr6] 15.3 0.5 55 5 [101]
(MeNH3)2[ReBr6] 5.6 33.2 [94]
(Me2NH2)2[ReBr6] 17.8 20.2 [94]
Li2[ReI6] 28 b
[90]
Na2[ReI6] 27 b
[90]
K2[ReI6] 24 a b
[90]
Rb2[ReI6] 21 b
[90]
Cs2[ReI6] 16 b
[90]
20 b
[90]
(NH4)2[ReI6]
17.5 b
[95]
(MeNH3)2[ReI6] 5.5 b
[95]
(Me2NH2)2[ReI6] 20.0 b
[95]
(Me3NH)2[ReI6] 21.5 b
[95]
py = pyridine; qy = quinoline; acr = acridine. a Temperature of the frequency-independent maximum in
the M in-phase ac signal. b Value not reported.

The case of K2[ReI6] is very special and requires a further comment. Its magnetic
properties in the form of a MT versus T plot are shown in Figure 5. The shape of the curve in
the high-temperature range is similar to that observed for the other compounds of the family.
At room temperature, MT is ca. 1.24 cm3 mol1 K, well below the expected value for an
isolated Re(IV) ion. Upon cooling, this value continuously decreases until ca. 25 K, where it
70 Carlos Kremer and Ral Chiozzone

exhibits an abrupt increase, and later it drops again with T below 20 K. This increase is more
pronounced at lower fields, where saturation effects are minimized.
The high-temperature behavior of K2[ReI6] reveals that a relatively large
antiferromagnetic coupling between [ReI6]2 anions actually occurs. On the other hand, the
abrupt increase of MT in the low-temperature range has been attributed to spin canting, with
a canting angle estimated on 1.2.

Figure 5. Thermal variation of the MT product for K2[ReI6] under an applied magnetic field of 1 T. The
inset shows the in-phase ac signal at two different frequency values [90].

The onset of long range magnetic ordering below 24 K was confirmed by field-cooled
magnetization measurements and by the presence of a maximum at this temperature in the in-
phase signal of the ac susceptibility on polycrystalline samples at different frequency values.
Well defined magnetic hysteresis loops have been obtained below Tmax (Figure 6) and the
magnetization at an external field of 5 T is far from the saturation value, in agreement with
the presence of a weak ferromagnetism [90]. The spin canting requires the lack of inversion
center in the crystal structure. In addition, the expected large anisotropy of the Re(IV), which
has already been mentioned, may contribute to significant spin canting through increasing the
possible antisymmetric exchange interaction [12, 105, 106].
Structure and Magnetic Properties of Mono - and Polynuclear Complexes 71

Figure 6. Hysteresis loop for K2ReI6 at 15 K [90].

Finally, the special case of (H4biim)[ReCl6]4H2O also deserves to be mentioned. Its


crystal structure comprises [ReCl6]2 and (H4biim)2+ ions and tetrameric rings of water
molecules. Their magnetic properties are shown in Figure 7. MT at room temperature is close
to that expected for magnetically isolated Re(IV). Upon cooling, MT reaches a minimum
value of 1.26 cm3 mol1 K at 6 K, and then it steeply increases. Given that the minimum value
of MT at low temperature is higher than 1 cm3 mol1 K, the occurrence of antiferromagnetic
exchange is ruled out. So, the features observed were interpreted in terms of ferromagnetic
coupling due to an unusual Re-Cl(H2O)Cl-Re pathway and the decrease of MT values at
higher temperatures was attributed exclusively to the zero-field splitting effect. The field
dependence of the magnetization supports the presence of a weak ferromagnetic interaction
but no hysteresis loops were observed [76]. The experimental magnetic data were then
analyzed by means of the following expression for the magnetic susceptibility derived from
the Hamiltonian of Eq. 7:

Ng 2 2 1
M ( F|| 2 F ) (15)
4k (T ) 3

where F|| and F have been defined in Eqs. 10 and 11, a term under the form of (T ) was
included to account for the magnetic coupling between the anions, and g|| = g = g to avoid
overparametrization. The results of the fitting procedures are |D| = 15.3(1) cm1, g = 1.81(1)
and = +0.58 K. The positive small value indicates the weak ferromagnetic interaction.
72 Carlos Kremer and Ral Chiozzone

Figure 7. Thermal dependence of MT for (H4biim)[ReCl6]4H2O. The inset shows a detail of the low
temperature range. Reproduced from Ref. [76] with authorization of the Royal Chemical Society.

Hexahalorhenate(IV) Salts of Paramagnetic Cations without Significant


Magnetic Interactions between Oppositely Charged Ions

This third group of M2[ReX6] compounds comprises a reduced number of salts shown in
Table 3. In a general scheme, their magnetic properties can be interpreted as the addition of
two different contributions: the zero-field splitting of Re(IV) and the paramagnetism of the
cation. In some cases, additional intermolecular interactions between the cations and/or
between the anions in the lattice will be a third contribution.
An interesting example is seen in the [ReCl6]2 salts of the organic radical cations p-rad
and m-rad. They crystallize in very different spatial groups, but in both cases the nitronyl
nitroxide cations adopt cyclic hexagonal arrangements in the crystal structures. The observed
magnetic properties are just the sum of magnetically isolated [ReCl6]2 anions (with their
characteristic zero-field splitting) and weak but significant antiferromagnetic exchange
between the radical cations [70]. The spin Hamiltonian is defined in Eqs. 16 and 17, as:

1
H H Re H radrad (16)
3

i 6
H radrad J (S1S 2 S 2S 3 S 3S 4 S 4S 5 S 5S 6 S 6S1 ) g rad H z S zi (17)
i 1
Structure and Magnetic Properties of Mono - and Polynuclear Complexes 73

where J is the isotropic exchange parameter as shown in Figure 8 for the six-membered ring
of spin doublets Si and HRe is as defined in Eq. 7. The results of the analysis of their magnetic
data are presented in Table 3.

Figure 8. Schematic representation of the cyclic arrangement of cations in (p-rad)2[ReCl6] with the
intermolecular interaction J [70].

A different situation is found in the [Fe(C5H5)2]+ salts of [ReCl6]2 and [ReBr6]2. In the
crystal lattices, the [ReX6]2 octahedra are arranged along one direction forming chains of Re
and X atoms, ReXXReXXRe, where the intermolecular XX distances are
smaller than the sum of van der Waals radii. These contacts allow the occurrence of weak
antiferromagnetic coupling between the hexahalorhenate(IV) units, in addition to the presence
of a large zero-field splitting of the Re(IV) cation and the paramagnetic contribution of the
ferrocenium cations [80]. After subtracting the last contribution, the average magnetic
susceptibility is expressed as in Eq. 15. The results of the fitting procedures are also given in
Table 3.
In the case of the [Fe(C5Me5)2]2[ReX6] series, very weak antiferromagnetic interactions
between anions are also present in the chloro and bromo compounds, while there is no
magnetic coupling in the [ReI6]2 salt. Only the crystal structure of [Fe(C5Me5)2]2[ReCl6] is
known and it was found that the anions are separated by bulky decamethylferrocenium
cations, the ClCl distances between adjacent anions along the c axis being close to the van
der Waals value. This would account for the lower antiferromagnetic coupling in this
compound compared with the ferrocenium analog. Although the structure of
[Fe(C5Me5)2]2[ReBr6] is unknown, it is probably the same as the [ReCl6]2 salt, taking into
account the similarity of their magnetic curves and that the corresponding ferrocenium
derivatives are isostructural. In the case of the iodo compound, a different structure is
anticipated in the lack of magnetic interactions [80].
The last example of this class of salts is -(BEDT-TTF)4[ReCl6]C 6H5CN, whose
structure consists of alternate layers of BEDT-TTF and of [ReCl6]2C 6H5CN. Half of the
BEDT-TTF molecules carry a charge close to +1 and the other half are approximately neutral.
In this case, the cation forms antiferromagnetic chains, although the magnetic properties are
mainly due to the dominant behavior of the isolated anions [69].
74 Carlos Kremer and Ral Chiozzone

Table 3. Calculated magnetic data for hexahalorhenates of paramagnetic cations


without significant cation-anion magnetic interactions

Compound |D|/cm1 gRe /K grad J/cm1 a Ref.


[Fe(C5H5)2]2[ReCl6] 16.6 1.82 4.1 - - [80]
[Fe(C5H5)2]2[ReBr6] 10.2 1.84 5.9 - - [80]
[Fe(C5Me5)2]2[ReCl6] 16.0 1.82 1.8 - - [80]
[Fe(C5Me5)2]2[ReBr6] 14.5 1.85 1.9 - - [80]
[Fe(C5Me5)2]2[ReI6] 12.0 1.83 0 - - [80]
(p-rad)2[ReCl6] 12 1.86 - 2.0 2.4 [70]
(m-rad)2[ReCl6] 13 1.85 - 2.0 4.5 [70]
a
J is the isotropic exchange constant describing the radrad interaction.

Hexahalorhenate(IV) Salts of Paramagnetic Cations with Magnetic


Interactions between Oppositely Charged Ions

The small number of compounds of this group is presented in Table 4. The first example
in the list, is [Fe(C5H5)2]2[ReI6] which exhibit a ferromagnetic coupling between anions and
cations, as evidenced by the shape of the T vs T plot and by the fact that the magnetization at
the highest available magnetic field tends to 5 BM (due to ferromagnetic coupling between
SRe = 3/2 and two S = 1/2 from two ferrocenium cations). Unfortunately, the lack of a crystal
structure for this iodo compound precludes a detailed analysis of the exchange pathway for
the observed ferromagnetic interaction. Clearly, its structure has to be different from that of
the chloro and bromo analogs mentioned in the previous section, and most likely, the anions
should be well separated hindering any antiferromagnetic coupling between them [80].
[CuL1][ReCl6]H 2O, (L1 = 6,13-bis(dodecylaminomethylidene)-1,4,8,11-
1
tetrazacyclotetradeca-4,7,11,14-tetraene), [CuL ][ReBr6] [107] and [Cu(tren)][ReCl6] (tren =
tris(2-aminoethyl)amine) [108] are examples of bimetallic Re-Cu compounds that behave as
ferrimagnetic chains. At room temperature, the values of MT for all of them are close to the
expected ones for uncoupled Re(IV) and Cu(II) ions. As the temperature is lowered, the MT
values smoothly decrease and reach a rounded minimum at Tmin = 13, 20, and 5 K,
respectively. The MT values at Tmin are 1.27, 1.31 and 0.93 cm3 mol1 K, below the sum of
those of the magnetically isolated Cu(II) and [ReCl6]2 units at these temperatures, supporting
the presence of antiferromagnetic coupling between both metallic centers. Upon further
cooling, the values of MT increase until a maximum is reached at 4.7, 5.5 and 2.8 K,
respectively, and finally they drop.
The occurrence of such minima in the MT values has also been found in other one-
dimensional ferrimagnetic chain-like rhenium(IV) compounds (vide infra). The maxima at
lower temperatures could be seen as weak antiferromagnetic exchange between chains or be
ascribed to saturation effects. At low temperatures, the magnetization steeply rises at low
fields supporting the second explanation in these cases. In addition, the magnetization vs.
temperature curves reveal magnetic transitions at 4.2, 6.2 and 3.15 K, respectively, which
agrees with a ferrimagnetic ordering below these temperatures. [CuL1][ReBr6] also exhibits a
second transition temperature at 3.6 K, suggesting that more than one magnetic phase can be
Structure and Magnetic Properties of Mono - and Polynuclear Complexes 75

present. Clearly, further studies are required for a full understanding of the magnetic behavior
of these compounds.
In the absence of structural details, their magnetic properties were analyzed as alternating
bimetallic chains with a single intrachain magnetic coupling parameter J, the spin
Hamiltonian being:

H { JS iz S iz1 D[(S iz ) 2 5 4] (g || ReS iz g||CuS iz1 )H z


i (18)
(g ReS ix g CuS ix1 )H x (g ReS iy g CuS iy1 )H y }

Assuming that zero-field splitting of Re(IV) is large and dominant over to the exchange
coupling, the following expressions for the magnetic susceptibility have been proposed:

N 2 2
|| (g||Cu g||2Re F|| )|| (19)
4k T

N 2 2
(g Cu g 2 Re F ) (20)
4k T

where F|| and F have been defined in Eqs. 10 and 11, and

exp J kT R 2 exp J kT
|| (21)
1 R2

2kT J 1 2 J
tanh sech (22)
J 4kT 2 4kT
with

g Re g Cu
R (23)
g Re g Cu

The high temperature range is then described by two contributions, the susceptibility of the
4
A2g term with zero-field splitting and that for uncoupled copper(II) ion, whereas in the low-
temperature region (but above the ordering temperature) the system is described as an
effective spin-1/2 Ising chain with different local g values [33, 110, 111]. The results of least-
squares fitting of the experimental data through these expressions at temperatures higher than
Tmin are also shown in Table 4. The model requires that D is much greater than J. The
obtained J are referred to a Seff = 1/2, and they should be reduced by a factor of about 3/5 for
comparing them with the D values (referred to a real S = 3/2) [112].
By contrast, the solvated form [Cu(tren)][ReCl6]2CH3OH behaves simply as an isolated
dinuclear entity with antiferromagnetic interaction between Re(IV) and Cu(II). The presence
76 Carlos Kremer and Ral Chiozzone

of solvent molecules could result in longer intermolecular distances, leading to the observed
change in the magnetic properties [108]. The corresponding spin Hamiltonian in this case is:

H J SReSCu D[S2z Re 5 4] g ReHSRe g CuHSCu (24)

where J is the exchange coupling parameter between the quartet Re(IV) and doublet Cu(II)
local spins, and the last terms account for the Zeeman effects of the two metal ions with
average gRe and gCu values.
Other interesting bimetallic system is that of [Ni(tetren)][ReCl6] (tetren =
tetraethylenepentamine) [109]. As the temperature is lowered, MT smoothly decreases,
reaches a rounded minimum at 4.4 K with MT = 0.91 cm3 mol1 K, then rises to a maximum
at 3.5 K and finally drops. Below 3.9 K, spontaneous magnetization appears, suggesting the
occurrence of 3D-ferrimagnetic ordering. Hysteresis loops characteristic of a soft magnet are
well visible at 1.7 K. For the analysis of the magnetic data, it was assumed valid the weak
exchange-coupling limit and that the magnetic anisotropy of the Ni(II) centers is much lower
than that of the Re(IV) because of the values of their spin-orbit coupling constants. Then, the
magnetic susceptibility above 6 K has been described in terms of g||Re, gRe and gNi values and
the isotropic exchange coupling constant J. The zero-field splitting parameter has been
constrained through the relation ship:

Re
DRe (g||Re g Re ) (25)
2

On the other hand, the solvated form [Ni(tetren)][ReCl6]CH3OH behaves like isolated
dinuclear entities with weak antiferromagnetic interaction between Re(IV) and Ni(II) [109].
Once again, the crystal structure of these rhenium-nickel compounds is unknown and
therefore it is impossible to go further in the analysis of their properties.

Table 4. Calculated magnetic data for hexahalorhenates of paramagnetic


cations with magnetic interactions between cations and anions

Compound |D|/cm1 gRe gM /K J/cm1 Ref.


[Fe(C5H5)2]2[ReI6] 13.0 1.84 - +1.9 - [80]
1
[CuL ][ReCl6]H 2O 26.7 1.91 a
2.06 a
- 9.1 b
[107]
1
[CuL ][ReBr6] 31.8 1.83 a
2.09 a
- 12.2 b
[107]
[Cu(tren)][ReCl6] 26.3 1.72 a 2.12 a - 9.27 b [108]
[Ni(tetren)][ReCl6] 28.0 1.46 a
2.109 - 2.35 [109]
[Cu(tren)][ReCl6]2CH3OH 13.1 1.81 2.11 - 2.3 [108]
[Ni(tetren)][ReCl6]CH3OH 26.5 1.45 a
2.489 - 3.25 [109]
a
Average g value calculated as ( g||2 2 g 2 ) / 3 . J value referred to a Seff = .
b
Structure and Magnetic Properties of Mono - and Polynuclear Complexes 77

OTHER RHENIUM(IV) MONONUCLEAR COMPLEXES


In a previous section of this chapter we discussed the general methods for the preparation
of Re(IV) mononuclear complexes. During the last years the preparative methods based on
halide substitution on [ReX6]2 (X = Cl, Br) complexes in non-aqueous media have been
extensively used [20]. Many novel complexes have been prepared and fully characterized.
Coordination number six for the rhenium atom is maintained with the exception of
[Re(CN)7]3 which exhibits a pentagonal bipyramidal geometry [23]. In the case of the
classical octahedral complexes, some distortion can be observed due to the steric
requirements of bidentate ligands like oxalato, malonato, catecholate or bipyrimidine. It
should be noted that most of them (with the single exception of (NBu4)2[Re(NCS)6] [37]) are
mixed-ligand complexes. The inertness of the central atom precludes a full substitution.
The magnetic properties of mononuclear Re(IV) compounds can be related to those of
some of the groups of hexahalorhenate salts previously mentioned. In fact, all of the
mononuclear Re(IV) complexes are expected to show a significant magnetic anisotropy, and
depending on the way the molecules arrange in the solid state, their magnetic behaviors span
from magnetically isolated to strongly correlated spin centers.
We can consider, for example, the list of compounds in Table 5 which comprises a group
of salts of anionic rhenium complexes with bulky diamagnetic cations. All of them behave
essentially as isolated or weakly interacting Re(IV) centers and no maximum of susceptibility
is observed for them in the temperature range explored. The thermal variation of the MT
product for any of these compounds can be mainly attributed to the large zero-field splitting
of Re(IV). Then, their magnetic susceptibilities have been adequately fitted by Eq. 15. Large
D values imply a noteworthy axial anisotropy and thus, there should be different values for g||
and g. However, from measurements of average magnetic susceptibility of powder samples,
it is not always possible to find reliable and unique values for both g-components. The fit of
the experimental data considering all parameters may result in several mathematically
acceptable solutions. Nevertheless, in all of them the absolute value of D and the mean value
of g remain practically constant.
As can be seen in Table 5, the |D| values for complexes containing monodentate ligands
as pyz or dmf, are very similar to the values for the hexahalorhenates. In contrast, the |D|
values for complexes with bidentate ligands as oxalato or malonato, are quite higher than the
previous ones. In these cases, the greater distortion of the octahedral geometry would account
for this difference.
The anions in the lattices are in general well-separated by the cations, but weak
antiferromagnetic interactions can be noted in some cases, as revealed by the negative sign of
. The magnitude of these interactions can be roughly related to the XX intermolecular
distances in the crystal.
One last rare example of magnetically isolated anionic rhenium complex is that of
[(CuL2)2Cl][ReCl4(ox)]Cl (L2 = N-dl-5,7,7,12,12,14-hexamethyl-1,4,8,11-
tetraazacyclotetradeca-4,14-diene). Here the cations [(CuL2)2Cl]3+ are paramagnetic and the
two Cu(II) ions are antiferromagnetically coupled through the chloro bridge [115]. In this
case, a |D| value of 45.6 cm1 has been calculated for the Re(IV) center, assuming that the
magnetic susceptibility of the compound can be expressed as the sum of that of the dinuclear
78 Carlos Kremer and Ral Chiozzone

copper unit as specified by the Bleaney-Bowers equation [116] and that of the isolated
[ReCl4(ox)]2 ion.

Table 5. Calculated magnetic and related structural data for anionic Re(IV)
mononuclear complexes (other than hexahalorhenates) with bulky diamagnetic cations

Compound |D|/cm1 g /K dmin(XX)/ Ref.


(NBu4)[ReCl5(pyz)] 9.4 1.83 0 5.932(1) [38]
(NBu4)[ReCl5(pym)] 6.2(2) 1.803(1) 2.33(5) 4.361(1) [36]
(NBu4)[ReCl5(pyd)] 14.1(1) 1.818(1) 0.75(1) 4.627(1) [36]
(NBu4)[ReCl5(dmf)] 10.1 1.85 0 4.263 [113]
(NBu4)[ReBr5(dmf)] 19.6 1.88 0 4.236 [113]
(NBu4)[ReBr5(pyzCOOH)] 12.0(1) 1.90(1) 7.85(3) 3.708 [35]
(NBu4)[ReBr4(OCN)(dmf)] 20.8 1.84 0.3 4.284(7) [42]
(NBu4)2[ReBr(OCN)2(NCO)3] 19.6 1.88 0 - [42]
(AsPh4)2[ReCl4(ox)] 60(5) 1.85 a 0 7.209 [33]
(NBu4)2[ReCl4(ox)] 53 1.85 a 0 6.210 [114]
(AsPh4)(HNEt3)[ReCl4(mal)] 54 1.82 0 5.514 (47) [34]
(AsPh4)1.5(HNEt3)0.5[ReCl4(mal)] 57 1.79 0 6.891(30) [34]
(PPh4)2[ReBr4(mal)] 35(2) 1.79(1) 0 b
[41]
(NBu4)(HNEt3)[ReCl4(cat)]H2cat 95(5) 1.998(6) 0.27(4) 6.942(2) [39]
(NBu4)trans-[ReCl4(CN)2]2dma 14.4 c
1.66 c
0 c
7.633(2) [53]
(PPh4)[ReCl5(tcm)] 15.2 1.65 0.78 d b
[43]
(AsPh4)2[Re(NCS)6]2H2O 7.2 1.77 0.38 d - [44]
pyz = pyrazine; pym = pyrimidine; pyd = pyridazine; dmf = N, N-dimethylformamide; pyzCOOH = 2-
pyrazinecarboxylic acid; mal = malonato; H2cat = catechol; dma = N, N-dimethylacetamide; tcm =
tricyanomethanide. a Average g value calculated as ( g||2 2 g 2 ) / 3 . b Unknown structure. c Value
obtained from the fitting of low-temperature magnetization data collected under various applied dc
fields. d zJ' value (in cm1) calculated as i = i [1(2zJ/N2gi2)i)]1, where i is the exchange-
corrected susceptibility in the i direction, and zJ is a measure of the exchange interaction [12].

The rhenium complexes in Table 6 are anionic or neutral, and short contacts between
halogen atoms of neighboring molecules lead to magnetic exchange from moderate to strong.
Most of these compounds exhibit a maximum in their susceptibilities at low temperatures. In
some cases, where the decrease of MT values observed at higher temperatures can be
attributed exclusively or mainly to the zero-field splitting, the experimental data have been
adequately fitted through the Eq. 15, leading to a set of D, g and parameters.
Alternatively, for compounds in which experimental evidence suggest a three-
dimensional antiferromagnetic ordering, their magnetic behaviors have been described by a
Curie-Weiss law at temperatures above Tmax. In these cases, the reported Weiss constants
correspond to the upper value of the intermolecular antiferromagnetic interactions, given that
they include the expected magnetic anisotropy.
Structure and Magnetic Properties of Mono - and Polynuclear Complexes 79

Table 6. Calculated magnetic data for Re(IV) mononuclear complexes


(other than hexahalorhenates) exhibiting significant magnetic exchange

Compound Tmax/K /K |D|/cm1 g /K Ref.


(NH2Me2)[ReCl5(pyz)] 10 - 6.7 1.79 11.0 [38]
(NH4)[ReCl5(pyz)]0.75H2O 12 - 7.4 1.86 43.0 [38]
[Ni(cyclam)][ReCl5(pyz)]2 a 7.5 - 5.0 1.90 14.0 [38]
trans-[ReCl4(py)2] 9.2 43.5 - 1.93 - [46]
[ReCl4(H2biim)]2dmf 2.8 - - 1.90 5.5 [45]
[ReCl4(pyim)]dmf 2.8 - - 1.88 9.6 [45]
[ReCl4(bipy)] 5.6 21.2 - 1.87 - [45]
(HMepy)[ReCl5(py)]4,4-bipy 5.5 8.97 - 1.93 - [47]
[ReCl4(bpym)] 7.0 b 15.3 - 1.86 - [117]
[ReBr4(bpym)] 20.0 b 22.5 - 1.84 - [117]
cis-[ReCl4(NHCMeNHPh)2] 9.7(1) - 10(1) 1.86 - [103]
pyim = 2-(2-pyridyl)biimidazole ; HMepy = 2-methylpyridinium; 4,4-bipy = 4,4-bipyridine. a

Square-planar [Ni(cyclam)]2+ is diamagnetic. Temperature of magnetic ordering.


b

The compounds [ReCl4(bpym)] and [ReBr4(bpym)] behave as weak ferromagnets at


temperatures below 7.0 and 20.0 K respectively, due to spin canting. Well defined magnetic
hysteresis loops have been obtained at 2.0 K for the two compounds, and the calculated
canting angles are 1.8 and 2.5, respectively [117]. Although it is not possible to predict the
spin canting occurrence, the non-centrosymmetric space group of the crystal structures and
the large Re(IV) anisotropy are again responsible of this phenomenon.
Other interesting behavior can be observed in cis-[ReCl4(NHCMeNHPh)2]. At very low
temperatures, this compound behaves as a metamagnet. This means that an external magnetic
field can cause a sudden transition from an antiferromagnetic to a ferromagnetic phase. In this
metamagnetic system, a phase transition is induced if a high field is applied along b direction,
but not along a nor c. After application of the field, the magnetic structure changes from one
with the magnetic moments aligned along c to a highly canted ferrimagnetic structure with
moment components along b [103, 104].
Finally, it should be mentioned that the magnetic properties of (NBu4)3[Re(CN)7] are
consistent with the 2E1'' ground state expected for a D5h environment, with a low-spin d3
electron configuration. The magnetic moment at room temperature indicates S = 1/2 and g =
2.33. The EPR spectrum of powder at 20 K shows a significant magnetic anisotropy, with g||
= 3.66 and g = 1.59 [23].

POLYNUCLEAR RHENIUM(IV) COMPLEXES


The preparation and full characterization of stable Re(IV) mononuclear complexes has
been used to design novel heterobimetallic compounds with different structural motifs and
interesting magnetic properties. This is the case for the anions [ReX4(ox)]2, [ReX4(mal)]2 (X
= Cl, Br), [ReBr5(pyzCOOH)], [ReCl5(pyz)], trans-[ReCl4(CN)2]2 and [Re(CN)7]3. The
structure of the complexes (Figure 9) shows the presence of ligands which can act as donor
80 Carlos Kremer and Ral Chiozzone

toward a second metal ion. The second ion can be fully solvated or as the central atom of a
preformed complex with the coordination sphere partially blocked. This is the com plex as
ligand strategy that has been used to prepare different Re(IV) heteropolynuclear complexes.
Their structures include well defined dinuclear complexes, tri-, tetra- and pentanuclear
compounds and chains.

Figure 9. Mononuclear Re(IV) complexes used to build heteropolynuclear complexes.

Dinuclear Re(IV)-M(II) Complexes

Table 7 gives a list of compounds that behave magnetically as isolated dinuclear Re(IV)-
M(II) units. Their magnetic properties can be well described by the Hamiltonian in Eq. 26:

H J SRe SM DRe[S2z Re 5 / 4] DM [S 2zM n (n 2) / 12] (g ReS Re g MS M )H (26)

where SM, gM and n stands for the spin operator, the mean g value and the number of unpaired
electrons on M(II), respectively. DM is a magnetic anisotropy parameter for M(II), the energy
gap between two successive sublevels being given by [MS2 (MS 1)2]DM. This spin-
Hamiltonian formalism is applicable only to metal ions with orbitally nondegenerate ground
terms in an octahedral environment. This condition is satisfied by manganese(II) (6A1g), and
nickel(II) (3A2g) but not by iron(II) (5T2g) and cobalt(II) (4T1g). To analyze the magnetic data in
these cases, it is necessary to introduce some simplifications. Actually, the orbital
contribution to the MT value should be lower than expected due to the distorted environment
of the M(II) ions. The crystal field component of lower symmetry splits the T term giving a
nondegenerate ground term. It was considered that this term reflects the electronic structure of
the complex in all the temperature range investigated and that it is further splitted by second-
order spin-orbit coupling, as for Re(IV). Obviously, this is a justified approach only at very
low temperatures [119]. When M = Cu, this general Hamiltonian is reduced to the simpler
case of Eq. 24. The least-squares fitting of the experimental results lead to the values given in
Table 7.
Structure and Magnetic Properties of Mono - and Polynuclear Complexes 81

Table 7. Calculated magnetic data for isolated dinuclear Re(IV)-M(II) complexes

DRe/cm DM/cm
Compound 1 1 gRe gM J/cm1 Ref.

[118]
[ReCl4(-ox)Cu(phen)2]CH 3CN 24(1) - 1.88(1) 2.02(1)
0.90(2)
[ReCl4(-ox)Ni(dmphen)2]CH 3CN [119]
a 44 6.1 1.82 2.08 +5.9

[ReCl4(-ox)Co(dmphen)2]CH 3CN 49 6.2 1.85 2.49 +5.2 [119]


[ReCl4(-ox)Fe(dmphen)2]CH 3CN [119]
a 48 14 1.83 2.20 +2.8

[ReCl4(-ox)Mn(dmphen)2]CH 3CN 45 0 1.85 2.0 0.1 [119]


[ReCl4(-ox)CuL3] a 62.7 - 1.772 b 2.470 0.71 [120]
[ReCl4(3-mal)Cu(bipy)2] 60 - 1.79 2.13 0.09 [34]
[ReCl4(-mal)Cu(phen)2]CH 3CN 44 - 1.78 2.12 0.39 [34]
[ReCl4(-mal)Cu(terpy)] a 57 - 1.7 2.08 +1.51 [34]
[ReCl4(- [121]
52(2) 16.1(2) 1.81(1) 2.20(2)
mal)Ni(dmphen)2]CH 3CN a 0.65(2)
[ReCl4(- [121]
mal)Ni(dmphen)(CH3CN)2(H2O)] 52(1) 8.2(2) 1.80(1) 2.15(1) 6.8(1)
(CH 3CN)0.5(H 2O)0.5
[ReCl4(- [121]
57(2) 27.1(3) 1.80(1) 2.48(2)
mal)Co(dmphen)2]CH 3CN 0.50(2)
[ReCl4(- [121]
58(2) 9.1(2) 1.81(1) 2.09(1)
mal)Fe(dmphen)2]CH 3CN 0.44(2)
[41]
[ReBr4(-mal)Cu(bipy)2] a 17(1) - 1.80(1) 2.15(1)
0.26(1)
[41]
[ReBr4(-mal)Cu(phen)2]0.25H2O 29(1) - 1.80(1) 2.16(2)
1.83(2)
[41]
[ReBr4(-mal)Ni(dmphen)2] a 30(2) 15(2) 1.80(1) 2.17(2)
1.37(2)
[ReBr4(- [41]
59(2) 18(1) 1.80(1) 2.30(1)
mal)Co(dmphen)2]CH 3CN 0.90(3)
[41]
[ReBr4(-mal)Mn(dmphen)2] a 48(1) 0 (fix) 1.80(1) 2.00(1)
1.29(1)
[ReCl4(CN)(-CN)FeL4] (see text for details) [122]
phen = 1,10-phenanthroline; dmphen = 2,9-dimethyl-1,10-phenanthroline; terpy = 2,2:6,2-
terpyridine; L3 = N-rac-5,12-dimethyl-7,14-diethyl-1,4,8,11-tetraazacyclotetradeca-4,11-diene; L4
= 6,6'-(pyridin-2-ylmethylazanediyl)bis-(methylene)bis(N-tert-butylpicolinamide. a Unknown
structure. b Average g value calculated as ( g||2 2 g 2 ) / 3 .

In [ReCl4(-ox)Cu(phen)2]CH3CN, both metal centers are bridged by an asymmetric


bis(bidentate) oxalato ligand, with two long Cu(II)-O bonds. The weak intramolecular Cu(II)-
Re(IV) antiferromagnetic coupling (J = 0.90 cm1) can be understood by considering the
symmetry and poor overlap between the metal-centered magnetic orbitals through bridging
oxalato. To visualize this interaction, in Figure 10 the dx2y2-type natural magnetic orbital
centered on Re(IV) and the dz2-type magnetic orbital centered on Cu(II) are represented. Both
orbitals are delocalized toward the bridge and they would be strictly orthogonal in C2v
symmetry. However, due to the asymmetry of the bridge and the distorted coordination
82 Carlos Kremer and Ral Chiozzone

geometry of Cu(II) the orthogonality between these orbitals is broken, but the overlap is
predicted to be very small resulting in a weak antiferromagnetic coupling between the metal
ions [118].

Figure 10. Schematic representation of symmetry adapted magnetic orbitals in [ReCl4(-ox)Cu(phen)2].

The family of compounds of formula [ReCl4(-ox)M(dmphen)2] (M = Ni, Co, Fe, Mn)


consist of neutral units in which the M(II) and Re(IV) metal ions exhibit distorted octahedral
coordination geometries, being bridged by a bis(bidentate) oxalato ligand. Their magnetic
behavior has been investigated over the temperature range 2.0-300 K. A very weak
antiferromagnetic coupling between Re(IV) and Mn(II) was observed, while a significant
ferromagnetic interaction occurs between Re(IV) and M(II) in the other three compounds
[119].
The nature and relative magnitude of these magnetic interaction can again be understood,
qualitatively at least, through orbital symmetry considerations. Assuming an octahedral
symmetry for the metal ions, Ni(II) holds two unpaired electrons in its eg magnetic orbitals,
which are orthogonal with the t2g magnetic orbitals of Re(IV), and hence all the bielectronic
interactions are ferromagnetic, giving a positive J value. High-spin Co(II), has an additional
unpaired electron in a t2g orbital, and consequently, an antiferromagnetic interaction with
Re(IV) arises, which opposes the others, giving a less positive net J value.
The antiferromagnetic contribution is greater in ReFe units, since high-spin Fe(II) has
two unpaired electrons in t2g orbitals. Eventually, in Mn(II) the number of antiferromagnetic
contributions is greater than that of the ferromagnetic ones, leading to quasi-compensation
and a very small and negative J value results.
It is important to emphasize at this point the difference between crystal structure and
magnetic behavior. A clear example of this is [ReCl4(3-mal)Cu(bipy)2]. Well-defined chains
are formed in the solid, where both axial positions of the Cu(II) ions are occupied by oxygen
atom belonging to the malonato of adjacent [ReCl4(3-mal)]2 units. However, its magnetic
properties suggest extremely weak magnetic interaction between the metal ions. Then,
considering only the shortest exchange pathway Cu(II)-O as significant, the compounds
behaves magnetically as a Cu(II)-Re(IV) dinuclear complex and its magnetic properties can
be reproduced theoretically by means of the Hamiltonian of Eq. 24 [34].
The remaining [ReX4(-mal)]2-based compounds in Table 7 whose crystal structure is
known, also consist of dinuclear neutral units where the metal ions are linked through a
malonato ligand, which adopts simultaneously the bidentate (at Re(IV)) and monodentate (at
M(II)) coordination modes.
Structure and Magnetic Properties of Mono - and Polynuclear Complexes 83

The M atom can be six- or five-coordinated, the latter being surrounded by four nitrogen
atoms of two bidentate ligands and one oxygen atom of the malonato ligand. The values of
the magnetic couplings show the existence of weak antiferromagnetic interactions through an
antisyn carboxylate bridge in most of the compounds. The antiferromagnetic contributions
dominate due to the low symmetry of the coordination sphere of M(II) ions that reduces the
possibility of orthogonality between the magnetic orbitals [41, 121].
Other interesting recent example of dinuclear unit is provided by [ReCl4(CN)(-
CN)FeL4]. Re(IV) is briged to Fe(II) through one cyanide ligand and the iron atom is in a
pentagonal bipyramidal environment with S = 2. As the temperature is lowered, the MT
values rise monotonically, reaching a maximum value 6.56 cm3 mol1 K at 16 K, indicative of
ferromagnetic exchange [122]. The data have been modeled according to the simple
Heisenberg-Dirac-Van Vleck spin Hamiltonian:

H J S Re S Fe (27)

A fit to the data in the range 30-300 K provided an exchange constant J = +6.0 cm1, giving
rise to an S = 7/2 ground state. Low-temperature magnetization data at various fields were
further collected and the non-superimposable isofield curves were analyzed through:

H DS 2z gSH (28)

where the ground state is finally characterized by the axial zero-field splitting parameter D7/2
= 2.3 cm1 and g7/2 = 1.84.

Discrete-Size Heterometallic Clusters

Trinuclear compounds of the type Re(1)MRe(2) are presented in Table 8. The


proposed spin Hamiltonian for the analysis of their magnetic properties is:

H J (S Re1 S M S Re 2 S M ) DRe1[S 2z Re1 5 / 4] DRe 2 [S 2z Re 2 5 / 4]


(29)
DM [S 2zM n (n 2) / 12] (g Re1S Re1 g Re 2S Re 2 g MS M )H

where the terms have the same meaning as in Eq. 26, and it is assumed that DRe1 = DRe2 = DRe
based on the molecular geometry. The magnetic data of (NBu4)2[{ReCl4(-ox)}2Co(Him)2]
have not been fitted, but intramolecular ferromagnetic interactions between Re(IV) and Co(II)
have been confirmed [123].
Unlike what has been seen with the set of trinuclear compounds, the situation with the
tetranuclears is much more diverse. Compound [ReCl4(-ox)Cu(terpy)(H2O)][ReCl4(-
ox)Cu(terpy)(CH3CN)] (Figure 11) is a tetranuclear entity of the type
Cu(1)Re(1)Cu(2)Re(2), where three J values are needed to describe the three non
equivalent intramolecular interactions [118].
84 Carlos Kremer and Ral Chiozzone

Table 8. Calculated magnetic data for trinuclear Re(1)MRe(2) complexes

Compound DRe/cm1 gRe gM J/cm1 /K Ref.


[{ReCl5(-pyz)}2Cu(dmf)4] 34.3 1.79 2.10 +11.8 - [38]
(NBu4)2[{ReCl4(-ox)}2Cu(Him)2] 45.8 1.86 2.11 +7.7 1.6 [123]
(NBu4)2[{ReCl4(-ox)}2Ni(Him)2] 63.5 1.82 2.14 +14.2 - [123]
(NBu4)2[{ReCl4(-ox)}2Co(Him)2] a a a
>0 a
[123]
(NBu4)2[{ReCl4(-ox)}2Mn(Him)2] 35.7 1.90 2.00 0.35 - [123]
Him = imidazole. a Value not determined.

Figure 11. Perspective drawing of the tetranuclear entity in [ReCl4(-ox)Cu(terpy)(H2O)][ReCl4(-


ox)Cu(terpy)(CH3CN)] [118].

Having into account its structure, the magnetic data were analyzed through the
Hamiltonian of Eq. 30:

H J11S Re1 S Cu1 J12 S Re1 S Cu2 J 22 S Re 2 S Cu2 DRe1[S 2z Re1 5 / 4] DRe 2 [S 2z Re 2 5 / 4] (30)
(g Re1S Re1 g Cu1S Cu1 g Re 2S Re 2 g Cu2S Cu2 )H

Least-squares fit leads to J11 = 0.83(2), J12 = +0.70(2), J22 = +5.6(3) cm1, DRe1 = 32(1), DRe2
= 34(1) cm1, gRe1 = 1.83(1), gRe2 = 1.85(1), gCu1 = 2.00(1), gCu2 = 2.01(1). The fit is very
good, however, the overparametrization cannot be excluded because of the large number of
variable parameters considered.
An interesting point is the simultaneous presence of antiferromagnetic and ferromagnetic
interactions within the molecule. The first Re(1)Cu(1) path is weakly antiferromagnetic, as
it was seen in [ReCl4(-ox)Cu(phen)2]CH3CN, where also an asymmetric bis(bidentate)
oxalato bridges the two metal ions. On the other hand, there are ferromagnetic interactions
through the chloro bridge between Re(1) and Cu(2) and the oxalato bridge between Re(2) and
Cu(2). The ferromagnetic couplings in these cases can be also explained in terms of the
negligible overlap between the intervening magnetic orbitals [118].
A different pattern of magnetic interactions are seen in [ReBr4(-ox)Cu(bipy)2] [125] and
[ReBr5(-pyzCOOH)M(dmphen)2]2CH3CN (M = Ni, Co) [35]. Their crystal structures are
made up of neutral heterodinuclear units, but there are close BrBr contacts that allow the
Structure and Magnetic Properties of Mono - and Polynuclear Complexes 85

establishment of significant through-space interactions between pairs of molecules (Figure


12).

Figure 12. Perspective drawing of [ReBr5(-pyzCOOH)Ni(dmphen)2]2CH3CN. Only the metal ion


environments and the bridging ligand are shown. Dashed lines represent the shortest BrBr
interactions [35].

Therefore, these compounds behave magnetically as tetranuclear


M(1)Re(1)Re(2)M(2) units, with one intra- JReM and one inter-dinuclear JReRe
exchange parameters. The appropriate Hamiltonian takes the form of Eq. 31 for compounds
with M = Cu, Ni:

H J Re M (S Re1 S M1 S Re 2 S M2 ) J Re Re S Re1 S Re2 DRe[S 2z Re1 S 2z Re 2 5 / 2]


(31)
DM [S 2zM1 S 2zM2 4 / 3] [g Re (S Re1 S Re 2 ) g M (S M1 S M2 )]H

with DM = 0 for Cu(II). In the case of the cobalt-containing compound a slightly more
complicated Hamiltonian was proposed, given the presence of the orbitally degenerate 4T1g
ground term of the high spin six-coordinate Co(II):

H J Re Co (S Re1 S Co1 S Re 2 S Co2 ) J Re Re S Re1 S Re2 DRe[S 2z Re1 S 2z Re 2 5 / 2]


Co (L Co1 S Co1 L Co2 S Co2 ) Co[L2zCo1 L2zCo2 4 / 3]
(32)
[g Re (S Re1 S Re 2 ) g e (S Co1 S Co2 ) (L Co1 L Co2 )]H

In this equation, = kA, where k is the orbital reduction factor (0 k 1) and A describes the
strength of the crystal field (A = 3/2 and 1 for the weak and strong crystal field limits,
respectively), Co is the spinorbit coupling constant of the Co(II) ion and Co represents the
86 Carlos Kremer and Ral Chiozzone

energy gap between the 4E and 4A2 levels arising from an axial distortion of the ground term
4
T1 of the Co(II) ion in octahedral symmetry [124]. The fit of the magnetic susceptibility
through Hamiltonian in Eq. 32 gives the values shown in the footnotes in Table 9. The quality
of the fit in the whole temperature range can be seen in Figure 13.

Figure 13. Thermal variation of the MT product for [ReBr5(-pyzCOOH)Co(dmphen)2]2CH3CN. The


solid line is the calculated curve (see text) [35].

These examples stress the need for an accurate structural knowledge of the rhenium(IV)-
containing complexes due to the importance of the intermolecular interactions in the magnetic
coupling between the metal ions.
Finally, the remaining compounds in Table 9 comprise a family of tetrametallic stars of
the type [{ReCl4(-ox)}3M]4. One central M atom and three peripheral Re atoms are
interconnected through bis-bidentate oxalato ligands, the four metal atoms being coplanar.
The anions are well separated from each other by bulky NBu4+ cations. The general
Hamiltonian for this case is:

H J Re M S M1 i 1 S Rei DRe i 1 (S 2z Rei 15 / 4) DM [S 2zM n(n 2) / 12]


3 3

(33)
[g Re (i 1 S Rei ) g MS M ]H
3

where M stands for Cu, Ni, Fe and Mn (DCu = 0). As in the previous example, the third term
of Eq. 33 was replaced by those in Eq. 34 to analyze the magnetic properties of the Co(II)
complex:

H' CoLCo SCo Co[L2zCo 2 / 3] LCoH (34)


Structure and Magnetic Properties of Mono - and Polynuclear Complexes 87

As in all the aforementioned heteronuclear oxalato compounds, antiferromagnetic coupling


between Re(IV) and Mn(II), and ferromagnetic coupling between Re(IV) and Ni(II), Co(II) or
Fe(II) are present [35, 126].

Figure 14. Thermal variation of the MT product for (NBu4)4[{ReCl4(-ox)}3Ni] under an applied
magnetic field of 100 G. The inset shows the frequency and temperature dependence of the out-of-
phase ac signal under external applied dc magnetic fields of 2000 (top) and 0 G (bottom) in a 1 G
oscillating field and in the frequency range 100-1400 Hz. Reproduced from Ref. [127] with
authorization of the American Chemical Society.

Table 9. Calculated magnetic data for tetranuclear Re(IV)-M(II) complexes

DRe/cm DM/cm JReM/cm JReRe/cm


Compound 1 1 gRe gM 1 1 Ref.

[ReCl4(-ox)Cu(terpy)(H2O)]
(see text for details) [118]
[ReCl4(-ox)Cu(terpy)(CH3CN)]
[ReBr4(-ox)Cu(bipy)2] 83.8 - 1.80 2.10 0.65 1.51 [125]
[ReBr5(-
8.84(5) 4.46(5) 1.85(1) 2.28(1) +0.66(1) 0.12(2) [35]
pyzCOOH)Ni(dmphen)2]2CH3CN
[ReBr5(-
pyzCOOH)Co(dmphen)2]2CH3CN 8(1) - 1.89(1) - +0.60(2) 0.02(1) [35]
a

(NBu4)4[{ReCl4(-ox)}3Cu] 14.3 - 1.93 2.20 +4.64 - [126]


(NBu4)4[{ReCl4(-ox)}3Ni] 2.8 0 (fixed) 1.87 2.18 +16.3 - [126]
(NBu4)4[{ReCl4(-ox)}3Co]b 45 - 1.95 - +3.0 - [126]
(NBu4)4[{ReCl4(-ox)}3Fe] 40.9 7.77 1.89(1) 2.30(1) +1.62 - [126]
(NBu4)4[{ReCl4(-ox)}3Mn]c 23.1 0.00 1.95(1) 2.00(1) 1.30 - [126]
a
= 1.40(1), Co = 130(2) and Co = +715(10) cm1. b = 1.28, Co = 150 and Co = 200 cm1. c

Unknown structure.

(NBu4)4[{ReCl4(-ox)}3Ni] is the first example of a Re(IV)-containing compound that


behaves as a single-molecule magnet with an S = 11/2 ground-state and D11/2 = 0.8(1) cm1.
88 Carlos Kremer and Ral Chiozzone

The negative D value was best estimated from high-frequency electron paramagnetic
resonance studies on polycrystalline samples at 4.2 K. The compound exhibits frequency-
dependent out-of-phase ac signals at very low temperatures, indicating a system with slow
relaxation of magnetization (Figure 14). For H = 0, no maxima of ''M are observed above 1.9
K in the frequency range explored. However, maxima are observed at T > 1.9 K for H > 500
G, their position being shifted to greater temperatures with increasing field [127]. Additional
micro-SQUID magnetization measurements on single crystals at temperatures down to 0.04 K
have shown that extremely fast tunneling occurs at zero applied dc field. However, the
tunneling is switched off when H > 0 and temperature- and sweep-rate-dependent coercitivity
appears [126].
The final example is provided by the pentanuclear cluster
[(PY5Me2)4Mn4Re(CN)7](PF6)56H2O (PY5Me2 = 2,6-bis(1,1-bis(2-pyrydil)ethyl)pyridine).
Its structure consists of a central [Re(CN)7]3 complex connected through cyanide bridges to
four surrounding [(PY5Me2)Mn]2+ units. The arrangements of the four Mn(II) centers can be
described as a slightly distorted square, with two of the metals binding axial cyanide ligands
and the other two binding non-neighboring equatorial cyanide ligands. Magnetic
susceptibility data show the presence of ferromagnetic interactions between the S = 1/2
[Re(CN)7]3 and the four S = 5/2 Mn(II) centers. Assuming an exchange Hamiltonian of the
form:

H J ReM S Re i1 S Mni
4
(35)

the values J = +4.6 cm1 and g = 2.00 have been calculated. Variable-field magnetization data
at low temperatures are also consistent with a high-spin ground state (S = 21/2) with
significant zero-field splitting (D21/2 = 0.44 cm1). The frequency dependence of the out-of-
phase component of the ac susceptibility data confirms single-molecule magnet behavior,
with an effective spin relaxation barrier 33 cm1 [128].

Magnetic Chains

[ReCl4(-ox)Cu(bipy)2] is the first example of a Re(IV)-Cu(II) ferrimagnetic chain. The


[ReCl4(ox)]2 anions are coordinated to the [Cu(bipy)2]2+ cations through one oxalato-oxygen,
giving neutral heterometallic dinuclear units. These units arrange in such a way that a
chlorine atom of one of them points toward the copper atom of the neighboring one, resulting
in the formation of helical chains. In spite of the nonbonding CuCl distance, the compound
behaves as an alternating bimetallic chain with two weak intrachain magnetic coupling
through the oxalato-oxygen atom and through the chloro br idge [33]. A similar alternating
chain structure has been found with [ReCl4(-ox)Cu(pyim)2] although the oxalato bridges do
not adopt the same coordination mode towards the copper ion [129].
The approach used to analyze the magnetic susceptibility data of this kind of compounds
was similar to that described in previous examples of ferrimagnetic chains, as an effective
spin-1/2 Ising chain with different local g values and two alternating coupling parameters J
and j [33]. The expression for the relevant spin Hamiltonian is:
Structure and Magnetic Properties of Mono - and Polynuclear Complexes 89

H i { J S 2z i1 S 2z i j S 2z i S 2z i1 D[(S 2z i ) 2 5 / 4] (g Re||S 2z i g Cu||S 2z i1 )H z


(36)
(g ReS 2xi g CuS 2xi1 )H x (g ReS 2yi g CuS 2yi1 )H y }

and the parallel and perpendicular susceptibilities are given by the Eqs. 19 and 20 with:

exp J kT R 2 exp J 2kT (37)


||
(1 R 2 ) cosh( J / 2kT )

2kT j 1 2 j
tanh sech (38)
j 4kT 2 4kT

and

J j, g g Cu , g (39)
J g Re R
2 2 g

This approach has allowed the theoretical reproduction of the experimental susceptibility data
in the whole temperature range. Unfortunately, it is not easy to assign the J and j values to the
corresponding bridging pathway.
Other examples of ferromagnetic Re(IV)-Cu(II) chains are provided by [ReCl4(3-
ox)CuL5]dmf (L5 = N-dl-5,7,7,12,14,14-hexamethyl-1,4,8,11-tetraazacyclotetradeca-4,14-
diene) [130] and [ReCl4(ox)CuL6] (L6 = N-d-7,14-diisopropyl-5,12-dimethyl-1,4,8,11-
tetraazacyclotetradeca-4,11-diene) [131]. In the former compound the [ReCl4(ox)]2 anions
coordinate from above and below to the planar [CuL5]2+ cations through one oxalato-oxygen
(Figure 15). There is then only one magnetic pathway through the oxalato bridge and its
magnetic susceptibility is well described with the Hamiltonian of Eq. 18. In addition,
[ReCl4(3-ox)CuL5]dmf also shows magnetic ordering below 3.5 K. The crystal structure of
the second compound has not been resolved, but its magnetic behavior can be adequately
described by the same approach.

Figure 15. Crystal structure of [ReCl4(3-ox)CuL5]dmf [130].

[ReCl4(3-mal)Mn(dmphen)(H2O)2]dmphen(CH3CN)(H2O) is a ferrimagnetic chain


with regular alternating Re(IV) and Mn(II) cations. The malonato group adopts
90 Carlos Kremer and Ral Chiozzone

simultaneously the bidentate (towards the rhenium atom) and bis-monodentate (towards two
manganese atoms) coordination modes. The approach used to treat such a spin chain was
based in the assumption that the high temperature domain would obey the sum of the
magnetic contributions of a pair of magnetically non-interacting Re(IV) and Mn(II) cations
because of the weak intrachain antiferromagnetic interaction observed. Then, at lower
temperatures the problem is reduced to the interaction between an effective spin doublet of
the Re(IV) and the spin sextet of the high-spin Mn(II) where FJ was derived through the
Hamiltonian:

H J S eff [(1 a)S Mn i (1 a)S Mn ( i1) ] (40)


i

and the expression used to describe the magnetic susceptibility is as follows:

( Re Mn ) FJ (41)

2
g eff [ Seff ( Seff 1)(1 P) 2QR] 2Seff G(Q R) G 2 (1 P)
FJ (42)
(1 P)[G 2 g eff
2
Seff ( Seff 1)]

where:

J J eff [SMn (SMn 1)]1 / 2 , G g Mn [SMn (SMn 1)]1 / 2 , x J / kT,

A1 x[(1 a) B0 (1 a) B1 x[(1 a) B0 (1 a) B1
P , Q , R ,
A0 A0 A0

Seff
A0 (2 / 2 ) [ exp( ) /
Seff
2
]( 1)

Seff
A1 ( / 4 ) [ exp( ) /
Seff
4
][ 33 3 22 (6 22 ) 22 6]

Seff
B0 (2 / 2 ) exp( )
Seff

Seff
B1 ( / 4 ) [ exp( ) /
Seff
2
][ 22 2 2 22 ]

2 x, a , 2 2 x 2 (1 a 2 ) 2 x 2 (1 a 2 ) (43)
Structure and Magnetic Properties of Mono - and Polynuclear Complexes 91

with a classical SMn =5/2 and a quantic Seff =1/2 interacting local spins [121]. It has to be
stressed again that in order to compare the J values from all of these ferrimagnetic chains with
J values from other molecules, the real spin S = 3/2 is related to the effective spin Seff = 1/2
within the Kramers doublet by S = 5/3 Seff. So, the obtained J should be reduced by a factor of
about 3/5 [112].
The analogue bromo compound [ReBr4(-mal)Mn(dmphen)(H2O)2]dmphen(CH3CN)
(H2O) also shows a ferrimagnetic chain behavior, but its crystal structure was not resolved. In
order to get an estimation of the magnitude of the antiferromagnetic magnetic coupling, its
magnetic data have been analyzed as a dinuclear Re(IV)Mn(II) pair through the
Hamiltonian in Eq. 26. The calculated curve closely follows the experimental data from room
temperature until ca. the temperature of the minimum MT value. The disagreement between
the experimental and calculated data in the very low temperature domain is most likely due to
the magnetic interactions between the considered Re(IV)Mn(II) pair and its nearest
neighbors within the chain [41].

Figure 16. Crystal structure of [ReCl4(-CN)2Mn(dmf)4] [53].

Finally, single-chain magnet behavior has been found in a family of compounds of


formula [ReCl4(-CN)2M(dmf)4] (M = Mn, Fe, Co, Ni) whose structure is sketched in Figure
16. Intrachain antiferromagnetic exchange coupling between Re(IV) and Mn(II) and
ferromagnetic exchange between Re(IV) and the other ions are present in these compounds.
The MT vs T curve for the Mn(II) compound shows the typical minimum for ferrimagnetic
chains at 35K, and below this temperature MT climbs to a maximum at 6 K and then turns
down sharply, owing to field saturation of the magnetization. The corresponding curves for
the other chains show monotonically increasing MT values when temperature is decreased,
until a maximum is reached at very low temperatures. In order to quantify the strength of
intrachain exchange coupling, the data were modeled according to a spin Hamiltonian for a
chain comprised of alternating Heisenberg classical spins:


H J S Rei S Mi S Mi S Rei 1 (44)
i

Then the susceptibility is given by:

N 2 2 1 P 2 1 P
(45)
( M Re M M ) 1 P ( M Re M M ) 1 P
6k T
92 Carlos Kremer and Ral Chiozzone

where

M i gi Si (Si 1) (46)

and

J S Re ( S Re 1) S M ( S M 1) kT (47)
P coth
k T J S Re ( S Re 1) S M ( S M 1)

The experimental data have been fitted above the temperature of the maximum of MT to give
the results shown in Table 10. In the case of the Fe, Co and Ni compounds, it was assumed
that gRe = gM, because the fitting procedure was unable to independently determine the two g
parameters.
Variable-temperature ac susceptibility measurements reveal a strong frequency
dependence of both in-phase 'M and out-of-phase ''M components (Figure 17). From the ''M
data, the relaxation times were extracted for each peak through the expression = 1/2,
where is the switching frequency of the ac field. At low temperatures, all of the compounds
exhibit thermally-activated slow relaxation of the magnetization. The energy barriers to
relaxation are as high as 56 cm1 for the iron chain. Notably, this compound presents a
significant hysteresis effect at 1.8 K, with a coercive field of 1.0 T and a remnant
magnetization of 3.77 BM, thus demonstrating magnet-like behavior in this one-dimensional
system [53].

Table 10. Calculated magnetic data for chain-like Re(IV)-M(II) complexes

DRe/cm DM/cm
Compound 1 1 gRe gM J/cm1 j/cm1 Ref.


25(1)
[ReCl4(-ox)Cu(bipy)2] 53(5) - 1.83 a 2.12 a e 13.0(5) [33]
e

[ReCl4(-ox)Cu(pyim)2] 54.8 - 1.80 2.29 7.8 e 6.0 e [129]


[ReCl4(3-ox)CuL5]DMF 49.5 - 1.91 2.27 3.36 e - [130]
[ReCl4(ox)CuL6] 49.5 - 1.91 2.27 26.6 e - [131]
[ReCl4(3-
3.0(1)
mal)Mn(dmphen)(H2O)2] 49(3) 0 (fixed) 1.80(1) 1.98(1) e - [121]
dmphen(CH 3CN)(H 2O)
[ReBr4(-
1.80(1) 2.00(1)
mal)Mn(dmphen)(H2O)2] 46(1) c 0 (fixed) c c 2.60(2) - [41]
c
dmphen(CH 3CN)H 2O b
[ReCl4(-CN)2Mn(dmf)4] - - 1.80(6) 1.96(2) 10.8(8) - [53]
[ReCl4(-CN)2Fe(dmf)4] - - 1.96(6) d +9.6(8) [53]
[ReCl4(-CN)2Co(dmf)4] - - 2.11(3) d +4.8(2) [53]
[ReCl4(-CN)2Ni(dmf)4] - - 2.04(4) d +7.4(6) [53]
a
Average g value calculated as ( g||2 2 g 2 ) / 3 . Unknown structure. Approximate value calculated
b c

as for dinuclear Re(IV)Mn(II) pairs. d


gRe was set equal to gM in the fitting procedure. e
J (j)
value referred to Seff = .
Structure and Magnetic Properties of Mono - and Polynuclear Complexes 93

Figure 17. Variable-temperature in-phase (top) and out-of-phase (bottom) ac magnetic susceptibility
data for [ReCl4(-CN)2Fe(dmf)4] collected in a 4 Oe ac field oscillating at various frequencies.
Reproduce from Ref. [53] with authorization of the American Chemical Society.

CONCLUSION
Rhenium(IV)-containing heterobimetallic species have been synthesized by using the
rational complex as ligand approach, where a fully solvated transition metal ion or a
preformed complex whose coordination sphere is unsaturated is allowed to react with an inert
Re(IV) complex. The efforts done during the last years to prepare and characterize stable
Re(IV) mononuclear complexes have provided of several novel building blocks which have
been used to design polynuclear compounds with different structural motifs.
Mononuclear Re(IV) complexes are usually octahedral and they exhibit two very
important characteristics from a magnetic point of view: a large degree of spin delocalization
on the ligands in its complexes (because of covalency effects) and its remarkable magnetic
anisotropy which is due to the high value of the spin-orbit coupling parameter.
The magneto-structural studies of all of these mono- and polynuclear compounds have
provided a reasonable understanding of the magnetic behavior of the Re(IV) mononuclear
complexes as well as those of the heterometallic Re(IV)M(II) species. The detailed
interpretation of their magnetic properties is not an easy task due to the extremely large axial
zero-field splitting parameter (D) for the Re(IV) precursors.
Both, ferro- and antiferromagnetic interactions have been observed in different structural
units, and the results have been qualitatively explained in terms of the symmetry of the
interacting magnetic orbitals. Finally, single molecule magnet and single chain magnet
behavior has been observed in discrete-size clusters and chain compounds, respectively.
94 Carlos Kremer and Ral Chiozzone

ACKNOWLEDGEMENT
This chapter was supported by CSIC (Comisin Sectorial de Investigacin Cientfica,
Uruguay), project number 653.

REFERENCES
[1] Thompson, L. K. Coord. Chem. Rev. 2002, 233-234, 193-206.
[2] Frey, G. Chem. Soc. Rev. 2008, 37, 191-214.
[3] Uemura, T.; Yanai, N.; Kitagawa, S. Chem. Soc. Rev. 2009, 38, 1128-1236.
[4] Ma, L.; Abney, C.; Lin, W. Chem. Soc. Rev. 2009, 38, 248-1256.
[5] Czaja, A. U.; Trukhan, N.; Mller, M. Chem. Soc. Rev. 2009, 38, 1284-1293.
[6] Murray, L. J.; Dinca, M.; Long, J. R. Chem. Soc. Rev. 2009, 38, 1294-1314.
[7] Allendorf, M. D.; Bauer, C. A.; Bhakta, R. K.; Houk, R. J. T. Chem. Soc. Rev. 2009, 38,
1330-1352.
[8] Kurmoo, M. Chem. Soc. Rev. 2009, 38, 1353-1379.
[9] Steed, J. W.; Atwood, J. L. Supramolecular Chemistry; 2nd. Ed., John Wiley and Sons
Ltd., West Sussex, 2009.
[10] Verdaguer, M. Polyhedron 2001, 20, 1115-1128.
[11] Miller, J. S. Dalton Trans. 2006, 2742-2749.
[12] Carlin, R. L. Magnetochemistry; Springer, Berlin, 1986.
[13] Kahn, O. Molecular Magnetism; VCH, New York, 1993.
[14] Rouschias, G. Chem. Rev. 1974, 74, 531-566.
[15] Connor, K. A.; Walton, R. A. Rhenium. In Comprehensive Coordination Chemistry,
Wilkinson, G., Gillard, R. D., McCleverty, J. A. Eds. Elsevier, Oxford, 1985.
[16] Abram, U. Rhenium. In Comprehensive Coordination Chemistry II, J. A. McCleverty,
J. A., Meyer, T. J. Eds. Elsevier, 2004.
[17] Schwochau, K. Z. Naturforsch. 1965, 20a, 1286-1289.
[18] Schwochau, K. Z. Naturforsch. 1973, 28a, 89-97.
[19] Schweitzer, G. K.; Pesterfield, L. L. The Aqueous Chemistry of the Elements; Oxford
University Press, 2010.
[20] Chiozzone, R.; Cuevas, A.; Gonzlez, R.; Kremer, C.; Armentano, D.; De Munno, G.;
J.; Faus, Inorg. Chim. Acta 2006, 359, 2194-2200.
[21] Chiozzone, R.; Gonzlez, R.; Kremer, C.; Cerd, M. F.; Armentano, D.; De Munno, G.;
Martnez-Lillo, J.; Faus, J. Dalton Trans. 2007, 653-660.
[22] Griffith, W. P.; Kiernan, P. M. J. Chem. Soc., Dalton Trans. 1978, 1411-1417.
[23] Bennett, M. V.; Long, J. R. J. Am. Chem. Soc. 2003, 125, 2394-2395.
[24] Brown, I. D.; Lock, C. J. L.; Wan, C. Can. J. Chem. 1973, 51, 2073-2076.
[25] Bucknor, S.; Cotton, F. A.; Falvello, L. R.; Reid, A. H.; Schmulbach, C. D. Inorg.
Chem. 1987, 26, 2954-2959.
[26] Hermann, W. A.; Thiel, W. R.; Herdtweck, E. Chem. Ber. 1990, 123, 271-276.
[27] Hbener, R.; Abram, U. Inorg. Chim. Acta, 1993, 211, 121-123.
[28] Rossi, R.; Marchi, A.; Marvelli, L.; Magon, L. Peruzzini, M.; Casellato, U.; Graziani,
R. Inorg. Chim. Acta, 1993, 204, 63-71.
Structure and Magnetic Properties of Mono - and Polynuclear Complexes 95

[29] Chen, X.; Femia, F. J.; Babich, J. W.; Zubieta, J. Inorg. Chim. Acta 2000, 306, 113-116.
[30] Hbener, R.; Abram, U.; Strhle, J. Acta Cryst. Sect. C 1995, 51, 876-878.
[31] Machura, B.; Dziegielewski, J. O.; Kruszynski, R.; Bartczak, T. J.; Kusz, J. Polyhedron,
2003, 22, 2573-2580.
[32] Hahn, F. E.; Imhof, L.; Lgger, T. Inorg. Chim. Acta, 1997, 261, 109-112.
[33] Chiozzone, R.; Gonzlez, R.; Kremer, C.; De Munno, G.; Cano, J.; Lloret, F.; Julve,
M.; Faus, J. Inorg. Chem. 1999, 38, 4745-4752.
[34] Cuevas, A.; Chiozzone, R.; Kremer, C.; Suescun, L.; Mombr, A.; Armentano, D.; De
Munno, G.; Lloret, F.; Cano, J.; Faus, J. Inorg. Chem. 2004, 43, 7823-7831.
[35] Cuevas, A.; Kremer, C.; Hummert, M.; Schumann, H.; Lloret, F.; Julve, M.; Faus,
Dalton Trans. 2007, 342-350.
[36] Arizaga, L.; Gonzlez, R.; Chiozzone, R.; Kremer, C.; Cerd, M. F.; Armentano, D.; De
Munno, G.; Lloret, F.; Faus, J. Polyhedron, 2008, 27, 552-558.
[37] Gonzlez, R.; Barboza, N.; Chiozzone, R.; Kremer, C.; Armentano, D.; De Munno, G.;
Faus, J. Inorg. Chim. Acta 2008, 361, 2715-2720.
[38] Martnez-Lillo, J.; Armentano, D.; Marino, N.; Arizaga, L.; Chiozzone, R.; Gonzlez,
R.; Kremer, C.; Cano, J.; Faus, J. Dalton Trans. 2008, 4585-4594.
[39] Cuevas, A.; Geis, L.; Pintos, V.; Chiozzone, R.; Sanchz, J.; Hummert, M.; Schumann,
H.; Kremer, C. J. Mol. Struct. 2009, 921, 80-84.
[40] Pintos, V.; Cuevas, A.; Onetto, S.; Seoane, G.; Denis, P. A.; Gancheff, J. S.; Faccio, R.;
Mombr, A. W.; Kremer, C. J. Mol. Struct. 2010, 963, 9-15.
[41] Cuevas, A.; Kremer, C.; Suescun, L.; Mombr, A. W.; Lloret, F.; Julve, M.; Faus, J.
Dalton Trans. 2010, 39, 11403-11411.
[42] Martnez-Lillo, J.; Armentano, D.; De Munno, G.; Lloret, F.; Julve, M.; Faus, J. Inorg.
Chim. Acta 2006, 359, 4343-4349.
[43] Malecka, J.; Kochel, A.; Mroziski, Materials Science 2002, 20, 91-96.
[44] Malecka, J.; Kochel, A.; Mroziski, J. Polish J. Chem. 2002, 76, 1509-1512.
[45] Martnez-Lillo, J.; Armentano, D.; De Munno, G.; Faus, J. Polyhedron 2008, 27, 1447-
1454.
[46] Mroziski, J.; Kochel, A.; Lis, T. J. Mol. Struct. 2002, 610, 53-58.
[47] Kochel, A.; Transition Met. Chem. 2010, 35, 1-5.
[48] Rouschias, G.; Wilkinson, G. J. Chem. Soc. A 1968, 489-496.
[49] [49] Middleton, A. R.; Masters, A. F.; Wilkinson, G. J. Chem. Soc., Dalton Trans.
1979, 542-546.
[50] Duatti, A.; Rossi, R.; Magon, L.; Mazzi, U. Transition Met. Chem. 1983, 8, 170-174.
[51] Bertolasi, V.; Ferretti, V.; Gilli, G.; Duatti, A.; Marchi, A.; Magon, L. J. Chem. Soc.,
Dalton Trans. 1987, 613-617.
[52] Sawusch, S.; Schilde, U.; Uhlemann, E. Z. Naturforsch. 1997, 52b, 61-64.
[53] Harris, T. D.; Bennett, M. V.; Clrac, R.; Long, J. R. J. Am. Chem. Soc. 2010, 132,
3980-3988.
[54] Watt, G. W.; Thompson, R. J. Inorg. Synth. 1963, 7, 189-192.
[55] Rulfs, C. L.; Meyer, R. J. J. Am. Chem. Soc. 1955, 77, 4505- 4507.
[56] Grundy, H. D.; Brown, I. D. Can. J. Chem. 1970, 48, 1151-1154.
[57] Takazawa, H.; Ohba, S.; Saito, Y. Acta Cryst. 1990, B46, 166-174.
[58] Sperka, G.; Mautner, F. Cryst. Res. Technol. 1988, 23, K109-K111.
[59] Lisher, E. J.; Cowlam, N.; Gilliot, L. Acta Cryst. 1979, B35, 1033-1038.
96 Carlos Kremer and Ral Chiozzone

[60] Wolff von Gudenberg, D.; Frenzen, G.; Massa, W.; Dehnicke, K. Z. Anorg. Allg. Chem.
1995, 621, 525-530.
[61] Beck, J.; Desgroseilliers, A.; Mller-Buschbaum, K.; Schlitt, K. J. Z. Anorg. Allg.
Chem. 2001, 628, 1145-1151.
[62] Martnez-Lillo, J.; Armentano, D.; De Munno, G.; Lloret, F.; Julve, M.; Faus, J. Cryst.
Growth Des. 2006, 6, 2204-2206.
[63] Loris, R.; Maes, D.; Lisgarten, J.; Bettinelli, M.; Flint, C. Acta Cryst. 1993, C49, 231-
233.
[64] Bettinelli, M.; Di Sipio, L.; Valle, G.; Aschieri, C.; Ingletto, G., Z. Kristallogr. 1989,
188, 155-160.
[65] Malecka, J.; Jger, L.; Wagner, C.; Mroziski, J. Polish J. Chem. 1998, 72, 1879-1885.
[66] Holyska, M.; Korabik, M.; Lis, T. Acta Cryst. 2006, E62, m3178-m3180.
[67] Chau, C.N.; Wardle, R. W. M.; Ibers, J. A. Acta Cryst. 1988, C44, 751-753.
[68] Lis, T.; Starynowicz, Acta Cryst. 1985, C41, 1299-1302.
[69] Kepert, C. J.; Kurmoo, M.; Day, P. J. Mater. Chem. 1997, 7, 221-228.
[70] Gonzlez, R.; Romero, F.; Luneau, D.; Armentano, D.; De Munno, G.; Kremer, C.;
Lloret, F.; Julve, M.; Faus, J. Inorg. Chim. Acta, 2005, 358, 3995-4002.
[71] Adman, E.; Margulis, T. N. Inorg. Chem. 1967, 6, 210-214.
[72] Mroziski, J.; Kochel, A.; Lis, T. J. Mol. Struct. 2002, 641, 109-117.
[73] Englert, U.; Koelle, U.; Nageswara, R. N.; Z. Kristallogr. 1994, 209, 780-780.
[74] Barbour, L. J.; MacGillivray, L. R.; Atwood, J. L. J. Chem. Crystallogr. 1996, 26, 59-
61.
[75] Blake, A. J.; Greig, J. A.; Schrder, M. Acta Cryst. 1990, C46, 322-324.
[76] Martnez-Lillo, J.; Armentano, D.; De Munno, G.; Marino, N.; Lloret, F.; Julve, M.;
Faus, J. Cryst. Eng. Comm. 2008, 10, 1284-1287.
[77] Yusenko, K. V.; Korolkov, I. V.; Gromilov, S. A.; Korenev, S. V. J. Struct. Chem.
2007, 48, 379-382.
[78] Martynova, S. A.; Yusenko, K. V.; Korolkov, I. V.; Baidina, I. A.; Korenev, S. V. J.
Struct. Chem. 2009, 50, 120-126.
[79] Gromilov, S. A.; Korenev, S. V.; Korolkov, I. V.; Yusenko, K. V.; Baidina, I. A. J.
Struct. Chem. 2004, 45, 482-489.
[80] Gonzlez, R.; Chiozzone, R.; Kremer, C.; Guerra, F.; De Munno, G.; Lloret, F.; Julve,
M.; Faus, J. Inorg. Chem. 2004, 43, 3013-3019.
[81] Cotton, F. A.; Daniels, L. M.; Schmulbach, C. D. Inorg. Chim. Acta 1983, 75, 163-167.
[82] Hauck, J.; Rssler, K. Acta Cryst. 1977, B33, 2124-2128.
[83] Kochel, A. Acta Cryst. 2007, E63, m596-m597.
[84] Mrozinski, J.; Tomkiewicz, A.; Hartl, H.; Brdgam, I.; Villain, F. Polish J. Chem.
2002, 76, 285-293.
[85] Berthold, H. J.; Jakobson, G. Angew. Chem. 1964, 76, 497-497.
[86] Yusenko, K. V.; Gromilov, S. A.; Baidina, I. A.; Korolkov, I. V.; Korenev, S. V. J.
Struct. Chem. 2005, 46, 109-115.
[87] Kochel, A. Acta Cryst. 2007, E63, m1968-m1968.
[88] Briscoe, H. V. A.; Robinson, P. L.; Rudge, A. J.; J. Chem. Soc. 1931, 3218-3220.
[89] Chakravorti, M. C.; Gangopadhyay, T.; In Inorganic Synthesis, Vol. 27, Ginsberg, A.P.
Ed. Wiley, New York, 1990.
Structure and Magnetic Properties of Mono - and Polynuclear Complexes 97

[90] Gonzlez, R.; Chiozzone, R.; Kremer, C.; De Munno, G.; Nicolo, F.; Lloret, F.; Julve,
M.; Faus, J. Inorg. Chem. 2003, 42, 2512-2518.
[91] Peacock, R. D. J. Chem. Soc. 1956, 1291-1293.
[92] Clark, G. R.; Russell, D. R. Acta Cryst. 1978, B34, 894-895.
[93] Mrozinski, J. Bull. Acad. Polon. Sci. Ser. Chim., 1978, 26, 789-798.
[94] Mrozinski, J. Bull. Acad. Polon. Sci. Ser. Chim., 1980, 28, 559-567.
[95] Tomkiewicz, A.; Villain, F.; Mrozinski, J. J.Mol. Struct. 2000, 555, 383-390.
[96] Figgis, B. N.; Lewis, J.; Mabbs, F. E. J. Chem. Soc. 1961, 3138-3145.
[97] Chiozzone, R. Ph. D. Thesis, Facultad de Qumica, Universidad de la Repblica,
Montevideo, 2000.
[98] Busey, R. H.; Dearman, H. H.; Bevan, R. B., Jr. J. Phys. Chem. 1962, 66, 82- 89.
[99] Anderson, P. W. Phys. Rev. 1950, 79, 705- 710.
[100] Smith, H. G.; Bacon, G. E. J. Appl. Phys. 1966, 37, 979-980.
[101] Busey, R. H.; Sonder, E. J. Chem. Phys. 1962, 36, 93- 97.
[102] Owen, J.; Stevens, K. W. H. Nature 1953, 171, 836-836.
[103] Reynolds, P. A.; Moubaraki, B.; Murray, K. S.; Cable, J. W.; Engelhardt, L. M.; Figgis,
B. N. J. Chem. Soc., Dalton Trans. 1997, 263-267.
[104] Reynolds, P. A.; Figgis, B. N.; Martn y Marero, D. J. Chem. Soc., Dalton Trans. 1999,
945-950.
[105] Dzyaloshinsky, I. J. Phys. Chem. Solids, 1958, 4, 241-255.
[106] Moriya, T. Phys. Rev. 1960, 120, 91- 98.
[107] Tomkiewicz, A.; Korybut-Daszkiewcz, B.; Zygmut, A.; Mrozinski, J. J. Mol. Struct.
2002, 613, 115-119.
[108] Tomkiewicz, A.; Zygmut, A.; Mrozinski, J. J. Mol. Struct. 2003, 644, 97-103.
[109] Tomkiewicz, A.; Boca, R.; Nahorska, J.M.; Mrozinski, J. J. Mol. Struct. 2005, 734,
143-148.
[110] Coronado, E.; Drillon, M.; Nugteren, P. R.; de Jongh, L. J.; Beltrn, D. J. Am. Chem.
Soc. 1988, 110, 3907- 3913.
[111] Fischer, E. J. Math. Phys. 1963, 4, 124- 135.
[112] Lines, M. E. J. Chem. Phys. 1971, 55, 2977- 2984.
[113] Martnez-Lillo, J.; Armentano, D.; De Munno, G.; Lloret, F.; Julve, M.; Faus, J. Inorg.
Chim. Acta 2006, 359, 3291-3296.
[114] Tomkiewicz, A.; Bartczak, T. J.; Kruszynski, R.; Mrozinski, J. J. Mol. Struct. 2001,
595, 225-231.
[115] Mroziski, J.; Bienko, A. Chem. Pap. 2009, 63, 306-312.
[116] Bleaney, B.; Bowers, K. D. Proc. Roy. Soc. (London) Ser. A 1952, 214, 451-465.
[117] Martnez-Lillo, J.; Lloret, F.; Julve, M.; Faus, J. J. Coord. Chem.. 2009, 62, 92-99.
[118] Chiozzone, R.; Gonzlez, R.; Kremer, C.; De Munno, G.; Armentano, D.; Cano, J.;
Lloret, F.; Julve, M.; Faus, J. Inorg. Chem. 2001, 40, 4242-4249.
[119] Chiozzone, R.; Gonzlez, R.; Kremer, C.; De Munno, G.; Armentano, D.; Lloret, F.;
Julve, M.; Faus, J. Inorg. Chem. 2003, 42, 1064-1069.
[120] Bienko, A.; Klak, J.; Mroziski, J.; Kruszynski, R.; Bienko, D. C.; Boca, R.
Polyhedron, 2008, 27, 2464-2470.
[121] Cuevas, A.; Kremer, C.; Suescun, L.; Russi, S.; Mombr, A. W.; Lloret, F.; Julve, M.;
Faus, Dalton Trans. 2007, 5305-5315.
[122] Harris, T. D.; Soo, H. S.; Chang, C. J.; Long, J. R. Inorg. Chim. Acta 2011, 369, 92-96.
98 Carlos Kremer and Ral Chiozzone

[123] Martnez-Lillo, J.; Delgado, F. S.; Ruiz-Prez, C.; Lloret, F.; Julve, M.; Faus, J. Inorg.
Chem. 2007, 46, 3523-3530.
[124] Lloret, F.; Julve, M.; Cano, J.; Ruiz-Garca, R.; Pardo, E. Inorg. Chim. Acta 2008, 361,
3432-3445.
[125] Chiozzone, R.; Gonzlez, R.; Kremer, C.; Armentano, D.; De Munno, G.; Julve, M.;
Lloret, F. Inorg. Chim. Acta 2011, 370, 394-397.
[126] Martnez-Lillo, J.; Armentano, D.; De Munno, G.; Wernsdorfer, W.; Clemente-Juan, J.
M.; Krzystek, J.; Lloret, F.; Julve, M.; Faus, J. Inorg. Chem. 2009, 48, 3027-3038.
[127] Martnez-Lillo, J.; Armentano, D.; De Munno, G.; Wernsdorfer, W.; Julve, M.; Lloret,
F.; Faus, J. J. Am. Chem. Soc. 2006, 128, 14218-14219.
[128] Freedman, D. E.; Jenkins, D. M.; Iavarone, A. T.; Long, J. R. J. Am. Chem. Soc. 2008,
130, 2884-2885.
[129] Martnez-Lillo, J.; Armentano, D.; De Munno, G.; Lloret, F.; Julve, M.; Faus, J. Dalton
Trans. 2008, 40-43.
[130] Tomkiewicz, A.; Mrozinski, J.; Brdgam, I.; Hartl, H. Eur. J. Inorg. Chem. 2005, 1787-
1793.
[131] Mrozinski, J.; Bienko, A. Chem. Papers 2009, 63, 306-312.
In: Coordination Polymers and Metal Organic Frameworks ISBN: 978-1-61470-899-5
Editors: O. L. Ortiz and L. D. Ramirez, pp. 99-127 2012 Nova Science Publishers, Inc.

Chapter 3

THE APPLICATIONS OF METAL ORGANIC


FRAMEWORKS IN THE FIELDS OF HYDROGEN
STORAGE AND CATALYSIS

Yaoqi Li, Ping Son, Yan Li and Xingguo Li


Beijing National Laboratory for Molecular Sciences (BNLMS),
(The State Key Laboratory of Rare Earth Materials Chemistry and Applications),
College of Chemistry and Molecular Engineering, Peking University,
Beijing, China.

ABSTRACT
MOFs (metal organic frameworks) are porous frameworks constructed by the
coordination centers of metal ions and polyatomic organic bridging ligands. Nowadays,
MOFs have attracted much attention as they have been widely investigated for hydrogen
storage, gas separation and catalysis. The microporous porosity, the large specific surface
area and especially the controllable framework have made MOFs superior to traditional
inorganic porous materials such as zeolites and activated carbon. For gas storage and
separation, physisorption on MOFs is one significant approach for future application. By
crystal engineering, both the pore size and the electronic and chemical nature of the
interior surface, on which gas molecules will be adsorbed, can be modified by careful
designs. Not only the porous structure, but also the metal sites and the organic linkers in
the frameworks should greatly affect the interactions between the gas molecules and
MOFs. Comparing with other microporous materials, the low framework density, the
high specific surface area, especially the controllable crystal structure have made MOFs
be a favorable research interest. Furthermore, the frameworks of MOFs contain various
structures and large amounts of essential metal ions, which should be helpful to promote
molecular separations and chemical reactions. Therefore, MOFs can be potentially
applied as a heterogeneous catalyst. MOFs have been used as good precursors of catalysts
and special substrates for dispersed active sites. The existence of abundant metal ions, the
large surface areas and the tailorable microporous structures in MOFs clearly help to
obtain highly efficient catalysts.

E-mail: xgli@pku. Phone and Fax: +86-10-62765930.


100 Yaoqi Li, Ping Son, Yan Li et al.

1. INTRODUCTION
Metal-organic frameworks (MOFs), which are constructed by the coordination centres of
metal ions and polyatomic organic bridging ligands, have been greatly focused on in recent
years. Metal-organic frameworks (MOFs) have attracted much attention as they have been
widely investigated for hydrogen storage, gas separation and catalysis.[1] The microporous
porosity, the uniform structure, the large specific surface area and especially the controllable
framework have made MOFs superior to traditional inorganic porous materials such as
zeolites and activated carbon.[1b, 2]

2. MOFS USED AS HYDROGEN STORAGE MATERIALS


In the past few decades, global dependency on the fossil fuel energy has caused
tremendous ecological crisis. Nowadays, worldwide interest is focused on using a clean-
burning substitute for fossil fuels, due to both economic and environmental reasons. Among
the various alternative energy strategies, utilizing hydrogen energy is a promising method due
to its high efficiency, reversibility and environmental-friendliness. Hydrogen is a promising
clean carrier for storage, transportation and conversion of energy in the 21st century.
However, the storage of hydrogen in a safe and economical manner is considered to be a
bottleneck in the application of an on-board power supply. The storage of this lightweight
fuel is one of the most important challenges impeding its practical application, which calls for
storing and releasing hydrogen in the proper capacity and condition with fast kinetics and
favorable reversibility [3]. The US Department of Energy 2010 targets (revised in 2009 Feb.)
for a hydrogen storage system are: a capacity of 4.5 wt% and 28 g H2/L, a lifetime of 1000
cycles at ambient conditions; The 2015 targets are: a capacity of 5.5 wt% and 40 g H2/L, a
lifetime of 1500 cycles. [4] By far, there exist mainly two kinds of hydrogen storage
materials. One of them is storage based on chemisorption, including metal hydrides and
complex hydrides. The other kind is based on physisorption, including active carbon, zeolites
and metal organic frameworks (MOFs). The chemisorption materials, such as magnesium,
usually have large hydrogen storage capacities but show bad reversibility and weak kinetic
properties. While the Physisorption materials is another significant approach for storing
hydrogen [5]. Comparing with the chemisorption materials, the physisiorption materials show
rapid kinetic properties and excellent reversibility at low temperatures because of the weak
interaction between hydrogen molecules and the materials. Physisorption on large surface-
area materials is considered as a promising approach for storing hydrogen. Furthermore,
comparing with other porous materials such as zeolites and active carbon, some MOFs
(Metal-organic frameworks) show promising results for hydrogen adsorption at low
temperatures under appropriate pressures [5]. The low framework density, the high specific
surface area, especially the controllable crystal structure have made MOFs be a favorable
research interest [1b, c, 2b, 5d, 5g, 6]. By crystal engineering7, both the pore size [19] and the
electronic and chemical nature of the interior surface, on which H2 molecules will be
adsorbed, can be modified by careful designs.
Importantly, the pores in the as-synthesized MOFs are filled with guest molecules
(usually the solution molecules) after the preparation process. Therefore, to use MOFs as
The Applications of Metal Organic Frameworks 101

hydrogen storage materials, we must first let the MOFs get rid of guest molecules, that is, the
activation process. Usually, MOFs with guest molecules are treated at high temperature and
under vacuum to empty the pores. However, MOFs, built up by connecting inorganic parts
with various organic linkers, have an important characteristic that the porosity of the
framework may not exist under high temperature evacuation treatment. The thermal stability
varies a lot for different MOF structure. In the research on MOFs, TGA (Thermogravimetric
analysis) result not only shows the thermal stabilities of MOF, but also turns out to be a
decisive data to ensure an optimized temperature of activation to remove the guest molecules.
It has been proved that the activation is the crucial process determining the gas sorption
property of MOFs [8]. For many MOFs, activations at different temperatures result in
remarkable differences in gas sorption isotherms and specific surface areas. At operating
temperatures lower than the optimal activation temperature, the solvent guest molecules could
not be removed from the porous framework completely. The obtained BET surface area and
corresponding pore volume is logically smaller than the actual data. However, if the operating
temperatures are higher than the optimal activation temperature, the partial collapse and
change of the framework structure take place. In some cases, if the guest molecules are too
big to get out of the pores, it is reasonable to use another small guest molecules to replace the
origin molecules in the pores (for example, use methanol to replace N,N-
dimethylformamide) , in order to promote the activation process. This exchange can be done
by immersing the MOFs into the solvent of replacing molecules. High temperature may
accelerate the exchange of guest molecules; however, some MOF structures with low stability
will not be kept during the heating process. Only after the careful and complete activation
process, the MOFs with empty microporous pores are ready for the gas uptake tests.
For all hydrogen storage materials, adsorbing more hydrogen to increase the hydrogen
storage density is one of the most important targets. Researchers have proposed that the size
of pores, the chemical nature of the porous framework, the rolling surface of the pores and the
exposed metal sites, may affect the hydrogen uptake properties of MOFs [14,5d, 9]. Yaghi
and his partners reported the gas adsorption of a series of microporous metal organic
frameworks (MOFs) consisting of tetrahedral [Zn4O]6+ units linked by linear
aryldicarboxylates [1b, 10], which shows a significant way to achieve the success of
hydrogen storage. In general, the hydrogen adsorption of the framework increases with its
specific surface area, which is in accordance with the physisorption theory. However, with the
research deepened, some new phenomena were reported. For example, in the work of Yaghi,
sample MOF-177 with larger specific surface area adsorbed less hydrogen [11], which
indicates the relationship between hydrogen and the microporous pores still needs to be
investigated further.

2.1. [M(Pyz) { Ni(CN)4 }] (M =Fe, Co, Ni)

To illustrate the relationship between pore size and gas uptake, we have reported the
hydrogen and nitrogen adsorptions of [ M(Pyz) { Ni(CN)4 }] (M =Fe, Co, Ni) (Pyz =
pyrazine) (for short, they are named FePy, CoPy, NiPy, respectively). The 3D coordinating
compound [ Fe(Py) ] was rst reported by Kitazawa et al. [12]. Thereafter, its analogues were
widely studied due to their superior spin crossover properties [13]. In the series of [ M(Py) {
Ni(CN)4 }] (M = Fe, Co, Ni), coordinating planes are formed by Ni(CN)42 ions tetragonally
102 Yaoqi Li, Ping Son, Yan Li et al.

coordinated with M2+13c. At each side of the plane, the coordinating vacancies are occupied by
the nitrogen atoms of pyrazines. The coordinating polymers form 3D pores with tetragonal
crystal structure. The coordinating ability of M2+ ions to the ligands increases in the following
order: Fe2+ < Co2+ < Ni2+. This difference determines the crystallization and stability of the
frameworks, as shown in Figure 1. The XRD (X-ray diffraction) peaks of each curve move
right in the order: FePy < CoPy < NiPy. This fact is in accordance with the prediction because
the unit cell parameters decrease with the radii of M2+ ions, and the diffraction peaks move
right according to Bragg equation. The series were prepared in aqueous solution and the pores
of the framework are lled with solution molecules, water molecules. According to TGA
results, which illustrate the thermal stability of the Py compounds, the activation of this series
was carried out at 493 K. After the activation and hydrogen storage test, the XRD patterns of
the series are tested again to monitor the structure change. The patterns are almost identical to
that of the as-prepared ones, which indicates that the series keeps the structure during the
dehydration and the hydrogen sorption process.

Figure 1. Left: XRD patterns of the as-synthesized FePy, CoPy and NiPy (There exist some unknown
impurities in FePy); Right: TGA curves of FePy, CoPy and NiPy.

Figure 2. Left: Nitrogen adsorption curves of FePy, CoPy and NiPy at 77 K; Right: Hydrogen
adsorption isotherms of FePy, CoPy and NiPy at 77 K.
The Applications of Metal Organic Frameworks 103

Nitrogen adsorptions were tested at liquid nitrogen temperature and the isotherms are
shown in Figure 2 (Left). The nitrogen isotherms show typical type- sorption behaviors.
According to BET (Brunauere-Emmette-Teller) equation, the specic surface areas of the
series increase in the order of FePy (331 m2g-1) < CoPy (497 m2g-1)< NiPy (518 m2g-1), which
should be attributed to the increasing crystallization (FePy < CoPy < NiPy) and purity of the
compounds. For hydrogen storage at 77 K, the max storages of Py series at 7.7 MPa are 1.46,
2.19, 2.34 wt% respectively, as shown in Figure 2 (Right) and Table 1. According to BET
theory, gas molecules are adsorbed layer by layer on the adsorbent during physisorption.
Based on the monolayer hypothesis, the hydrogen adsorption amount on unit specic surface
area is estimated to be 2.27 wt% per 1000 m2g-1. Importantly, the capacity of Py series is very
high and the storages at 7.7 MPa are 4.41% for FePy, 4.41% for CoPy, 4.52% for NiPy per
1000 m2g-1, respectively, which are almost twice as much as the monolayer adsorption. For
comparison, the hydrogen storage properties of a Prussian blue analogue, FeFe(CN)6, is
measured and the data is shown in Figure 2 (Right) and Table 1. The storage capacity of
FeFe(CN)6 is 2.25 per 1000 m2g-1, only half of Py series. For further discussion, the gas
molecules adsorbed into the unit cell of the samples are calculated and the data are listed in
Table 1. According to the results, about two hydrogen molecules get into the unit cell of Py
series, while only one nitrogen molecule gets in the unit cell of Py compounds. Differently, in
FeFe(CN)6, only one hydrogen molecule can get into the unit cell under the same condition.
Importantly, the van der Waals diameters of nitrogen and hydrogen molecules are about 0.29
and 0.24 nm, respectively. In FeFe(CN)6, the pore diameter is about 0.23 nm, which is
determined by four tetragonally coordinated cyanides. The pore either permits one nitrogen
molecule to get in or allows only one hydrogen molecule to get in. Differently, in Py series
the pore diameter is determined by two adjacent pyrazines, which is about 0.42 nm, nearly
twice as big as that of FeFe(CN)6. Here, still only one nitrogen molecule can get into the hole,
while two hydrogen molecules can get into the same pore. The pore that contains one nitrogen
molecule in FeFe(CN)6 can contain only one hydrogen molecule, while the pore of
NiPy[Ni(CN)4] (Py = pyrazine) may hold one nitrogen molecule or hold two hydrogen
molecules in the diagonal direction. It is noticeable that hydrogen/nitrogen ratio adsorbed in
FeFe(CN)6 is 0.87.

Table 1. The properties of FePy, CoPy, NiPy and FeFe(CN)6

Sample FePy CoPy NiPy FeFe(CN)6


BET surface area (m2g-1) 331 497 518 405
Hydrogen storage at 7.7 MPa 1.46 2.19 2.34 0.91
(wt%)
Hydrogen storage per 1000 m2g-1 4.41 4.41 4.52 2.25
(wt%)
Nitrogen molecules per unit cell 1.28 1.90 1.95 5.54
(P/P0 = 0.2)
Hydrogen molecules per unit cell 2.16 3.28 3.5 4.84
(P = 7.7 MPa)
Ratio of hydrogen and nitrogen 1.69 1.73 1.79 0.87
molecules
104 Yaoqi Li, Ping Son, Yan Li et al.

Although the ratio of Py series is twice of FeFe(CN)6s, it is about 2. FeFe(CN)6 can only
absorb about half of the hydrogen that Py series absorb on per 1000 m2g-1, that is to say, pores
with too small diameter are not favorable for hydrogen storage because no more than one
hydrogen molecule can get into them. We suggest that the pores should be larger than 0.42
nm to obtain high hydrogen storage density.

2.2. Prussian Blue Analogues

To obtain further insights about the relationship between pore size and the hydrogen
sorption, microporous pores with different diameters were achieved by synthesizing a series
of Prussian blue analogues (metal-cyanide framework).

Figure 3. Structure of the analogues. With the rotation of water molecules, the positions of hydrogen
atoms are very hard to identify, so red balls are used to represent the water molecules. The structures of
the analogues shown in the figure are based on investigations and analysis.
The Applications of Metal Organic Frameworks 105

Prussian blue analogues were synthesized by solution method in this research. Prussian
blue analogues possess the structure that is based on face-center cubic (fcc) Mm+x[Mn+(CN)6]y
unit, in which [Mn+(CN)6] anions are linked by octahedrally coordinated nitrogen-bonded
Mm+ ions. When m + n = 6, x=y=1 and the unit cell is intact, for example Fe[Fe(CN) 6]. When
the sum is not 6, some sites of the unit cell are absent, which is more prevalent in the
analogues [14]. By changing the valences of the metallic ions in the analogues, we can make
the compounds to show different pore structures (Figure 3), which provide a good
opportunity to investigate the interaction between gas adsorbents and pores. In the XRD
patterns of the as-synthesized analogues, the positions of the peaks are almost the same,
which indicates that these samples possess quite similar unit cells. The ions in the unit cell of
the analogues are listed in table 2 based on these structures, in which half of Ni3[Fe(CN)6]2s
is calculated for easy comparison. TGA is used to determine the stability of the compounds,
the compounds lose its adsorptive molecules (mainly water molecules) at about 423 K and
keep their structures to about 503 K or higher. After the activation process carried out at 473
K, XRD patterns of the analogues are tested again after the hydrogen storage measurements
to confirm that the structures have not been changed.
As shown in table 2, based on BET equation, their specific surface areas are calculated
and the data are listed in Table 2. The specific surface areas range from 5 m2/g of
K2Ni[Fe(CN)6] to 718 m2/g of Ni3[Fe(CN)6]2. As mentioned above, the analogues possess
similar unit cells. Therefore, the differences of specific surface areas should be attributed to
the different fillings of ions in the unit cells. In the analogues, transitional metal ions must be
coordinated hexagonally either by cyanides or by other ligands. The coordination is
preferably performed by the nitrogen atoms in cyanides, but when cyanide ions are not
enough in most cases, water molecules take their places. Therefore, the compound with fewer
molecules in the unit cell has bigger porosity and shows higher specific surface area, for
example Ni3[Fe(CN)6]2. When there are similar amount of molecules in the unit cell, the
compound possessing less water molecules has bigger specific surface area, because water
molecule is made up of three atoms and is bigger than cyanide ion in volume, and also
because the metal ions are very small compared to the negative ions. The regular change of
specific surface areas of Ni2[Fe(CN)6], Fe[Fe(CN)6] and Fe4[Fe(CN)6]3 gives a good proof of
this conclusion. The hydrogen adsorption curves are shown in Figure 4, and the storages at 4
MPa are listed in Table 2, which ranges from 0.014 wt% of K2Ni[Fe(CN)6] to 2.45 wt% of
Ni3[Fe(CN)6]2.

Table 2. Basic data of the analogues

Hydrogen Hydrogen storage on


BET surface
Compounds storage at 4 1000 m2/g specific Ions in the unit cell
area (m2/g)
MPa (wt%) surface area (wt%)
Ni3[Fe(CN)6]2 718 2.45 3.41 5M2+, 12CN-, 6H2O
Ni2[Fe(CN)6] 282 0.97 3.44 6M2+, 12CN-, 12H2O
K2Ni[Fe(CN)6] 5 0.014 8M2+, 8K+, 24CN-
Fe[Fe(CN)6] 405 0.85 2.09 8M2+, 24CN-
Fe4[Fe(CN)6]3 320 1.03 3.22 7M2+, 18CN-, 6H2O
106 Yaoqi Li, Ping Son, Yan Li et al.

In order to discuss at a molecular level, the nitrogen and hydrogen adsorptions of the
analogues are converted into molecules in the unit cells for more convenient study, which are
listed in Table 3. For nitrogen, the adsorption at P/P0= 0.2 is taken, because the monolayer
adsorption has finished and the multilayer adsorption doesnt happen at P/P0= 0.2. For
hydrogen uptake, the hydrogen adsorptions at 4 MPa are taken. Hydrogen molecules adsorbed
decrease in the order: Ni3[Fe(CN)6]2>Fe4[Fe(CN)6]3> Fe[Fe(CN)6]> Ni2[Fe(CN)6]>
K2Ni[Fe(CN)6]. The amount correlates well with the change of specific surface areas or the
nitrogen molecules adsorbed except sample Fe[Fe(CN)6], which has smaller adsorption than
the analogues as mentioned above. In terms of hydrogen storage on unit specific surface area,
Fe[Fe(CN)6] adsorbs hydrogen of only 2.0 wt%, while other samples adsorb more 3.0 wt%.
As stated, the unit cell with small occupation can hold more adsorptive hydrogen molecules,
which is similar to the situation of nitrogen adsorption. The pore structures are investigated in
order to get further understanding about the adsorptive behavior of the analogues. There are
mainly four types of pores in the analogues. Pore I consists of 4 cyanides with vertexes of
metal ions. This type of pore is tetragonal and the van der Waals pore size is about 0.17 nm.
Pore type I is the basic structure of the crystal, which exits in all the analogues. It is known
that the hydrogen molecule has the van der Waals size of 0.24 nm in diameter and 0.30 nm in
length, and nitrogen molecule 0.31 nm and 0.42 nm respectively. As a result, pore type I is
too small to be an equilibrium position for either nitrogen or hydrogen, and is only a tunnel
for the adsorbents to get into the unit cell. Pore II is a cubic pore and composed of 12
cyanides with 8 vertexes of metal ions. The van der Waals pore size is about 0.38 nm. It
occupies the tetrahedral passions of the fcc unit cell, seeing the structure of Fe[Fe(CN)6],
which is an equilibrium position for adsorbents. This type of pores exits in all the analogues
except K2Ni[Fe(CN)6]. In compound K2Ni[Fe(CN)6], all this type of pores are filled by
potassium ions. As a result, compound K2Ni[Fe(CN)6] that have only type I pores behaves
like non-porous materials in the adsorption measurement. While in the cases of
Ni3[Fe(CN)6]2, Ni2[Fe(CN)6] and Fe4[Fe(CN)6]3, the cyanides were partly replaced by water
molecules, and subsequently the van der Waals pore size of type II will become smaller due
to the bigger size of three-atom water molecules. Pore III is octahedral and composed of six
water molecules vertexes. The van der Waals pore size is about 0.32 nm. It occupies the
octahedral position of the original fcc unit cell, seeing the structure of Fe4[Fe(CN)6]3 and
Ni2[Fe(CN)6]. Pore IV consists of a group of ions, which possesses the biggest van der pore
size in the analogues, about 0.68 nm. This big pore takes two original fcc unit cell to form
one, as shown in Figure 3, seeing the structure of Ni3[Fe(CN)6]2. Due to the large pore size,
pore IV can hold about two nitrogen molecules and about 3 hydrogen molecules. Therefore,
the ratio of H2/N2 in compound Ni3[Fe(CN)6]2 is 1.3, also bigger that 1, which is close to
compound Fe[Fe(CN)6] and Ni2[Fe(CN)6] but is smaller than NiPy[Ni(CN)4] (1.79).
Particularly, for the specail compound Fe[Fe(CN)6], Pore II has van der Waals pore
diameter of 0.38 nm. Comparing the pore size with the gas size, this type of pores can hold
one hydrogen molecule, and combined with pore I, it can hold one nitrogen molecule too.
Therefore, the ratio of H2/N2 in the unit cell is 0.82 for Fe[Fe(CN)6], about 1. Due the fact that
the adsorption of hydrogen molecule is much weaker than that of nitrogen, the ratio of H2/N2
adsorbed is difficult to reach 1. Particularly, pore III has the van der Waals pore size of 0.36
nm, which is similar to pore II. However, different from the isolated pore II in Fe[Fe(CN)6],
pore III is combined with eight pores II in Fe4[Fe(CN)6]3 and is combined to other pores III
and pores II in Ni2[Fe(CN)6]. Therefore, it can hold one nitrogen molecule as pore II does, but
The Applications of Metal Organic Frameworks 107

it may hold more than one hydrogen molecules. Therefore, the ratio of H2/N2 in compound
Fe4[Fe(CN)6]3 and Ni2[Fe(CN)6] is 1.26 and 1.40, which is bigger than 1, but doesnt double
the ration of Fe[Fe(CN)6].
This difference of hydrogen and nitrogen uptakes, as shown in the researches Py series
and Prussian blue analogues series, may be considered as another kind of molecule-sieve
effect for gas molecules with different sizes. This phenomenon is similar to molecular sieve
effect, in which the difference is between ads orb for the small and notadsorb for the big.
While in these researches, the difference is between ads orb more for the small and ads orb
less for the big. This phenomenon can be regarded as molecular sieve effect of a higher
level, which can be considered for the application of gas separation.
According to adsorption theory, the microporous pore provides some extra enhancement
to the adsorption in porous materials, and the enhancement reduces dramatically with the ratio
of pore-diameter/molecular-diameter [15]. That means the frameworks for hydrogen storage
should not be designed with pores of too large diameter, which has been proved by other
researchers [5b]. However, the specific surface area increases with the pore diameter in
general, which is also crucial for the hydrogen adsorption. Therefore, it is very important to
achieve an optimal pore size to obtain best hydrogen adsorption. It has been reported that
when the van der Waals pore diameter is about the adsorbent size, the adsorptive heat is 3
times of that on flat surface, and even when the pore diameter is 3 times of the adsorbent size
[15c], the adsorptive heat is still 0.5 times higher than that on flat surface, and for hydrogen,
that means 0.72 nm. In this research, Ni3[Fe(CN)6]2 with 0.72 nm van der Waals pore
diameter still have very strong hydrogen adsorption, which confirms the above analysis. It is
in accordance with the reported data, that is, the 0.7 nm wide slit-shaped porous materials can
show one layer of hydrogen molecules absorbed on opposite surfaces and show max van der
Waals potential [16] .

2.3. (M(HBTC)(4,4-Bipy).3DMF(M = Ni and Co)

Although large pore volumes and surface areas are significant factors for the
physisorption of hydrogen [17], it is obviously difficult to create a MOF with extremely large
surface area and pore volume because of Aristotles observation that nat ure abhors a
vacuum. Besides the pore size and the specific surface area, the interaction between
hydrogen molecules and the MOFs can also greatly affect the hydrogen uptake in MOF
systems. The metal-organic framework without exposed metal sites has the average
adsorption enthalpy of 6 kJmol-1. However, as reported [16], material used under pressure
between 1.5 bar and 30 bar at 298 K should need an optimal adsorption enthalpy (Hopt ) of
22~25 kJmol-1. Therefore, it is great challenge to increase the hydrogen adsorption enthalpy
to increase the operating temperature of MOFs (from 77 K to room temperature). It has been
reported that open metal coordination sites on the surface of MOFs can enhance the
interaction between MOF and hydrogen molecules. For example, the Mn3[(Mn4Cl)3(BTT)8]2
(H3BTT = benzene-1,3,5-tris(1H-tetrazole) shows a large isosteric adsorption heat (10.1
kJmol-1) at zero coverage, because of the exposed open Mn2+ sites [5d]. The interaction
between hydrogen and MOFs is greatly affected by the metal sites, while the different organic
ligands affect the structure and chemical nature of the frameworks and also affect the
interaction. To obtain some insights of the interaction between MOFs and hydrogen
108 Yaoqi Li, Ping Son, Yan Li et al.

molecules, two compounds (M(HBTC)(4,4-bipy).3DMF, M = Ni and Co, H3BTC = 1,3,5-


benzenetricarboxylic acid, 4,4-bipy = 4,4-bipyridine, DMF = N,N-dimethylformamide)
were synthesized by one-pot solution reaction and solvothermal method respectively, as
shown in Figure 4. To form the framework assembled by octahedral M2+ centers and two
kinds of connecting ligands, one of the bridging ligands (HBTC) posses -2 charge, the other
(4,4-bipy) should be neutral. M2+ ions (M = Ni and Co), bridged by HBTC2- divalent ions,
form 2D sheets. In these 2D sheets, one of the BTC units is in bidentate mode, while the other
two units are in monodentate fashion. The sheets are further pillared by 4,4-bipyridine along
c axis to form the porous structure. Two kinds of channels exist in the porous framework of
M(HBTC)(4,4-bipy).3DMF. One is normal rectangle channels running along a and b axis,
and the other is honeycomb channels generated by M(HBTC) layers. The XRD peaks of
Co(HBTC)(4,4-bipy).3DMF shift to slightly lower angle than the ones of Ni(HBTC)(4,4-
bipy).3DMF, because the unit cell parameter increases with the radii of M2+ ion, and the
diffraction peaks shift to low angle according to Bragg equation. As shown in the XRD
patterns (Figure 5), the crystallinity of the MOF structure remains unchanged under the
pressure and temperature conditions during the activation and gas adsorption tests.

Figure 4. Paking diagram viewed along the c-axis in Ni(HBTC)(4,4-bipy). Ni: green; O: red; C: black;
H: white.

Figure 5. Left: XRD patterns of (a) the simulation, (b) the as-synthesized sample, (c) the guest-free
sample evacuated at 453 K for 2 h and (d) the regenerated sample in DMF for Ni(HBTC)(4,4-
bipy).3DMF. Right: XRD patterns of (a) the as-synthesized sample for Ni(HBTC)(4,4-bipy).3DMF,
(b) the as-synthesized sample for Co(HBTC)(4,4-bipy).3DMF, (c) the guest-free sample evacuated at
433 K for 2 h and (d) the regenerated sample in DMF for Co(HBTC)(4,4-bipy).3DMF.
The Applications of Metal Organic Frameworks 109

Thermogravimetric analysis (Figure 6) shows the weight loss in M(HBTC)(4,4-


bipy).3DMF in N2 atmosphere. The framework of Ni(HBTC)(4,4-bipy).3DMF does not
collapse up to 503 K. Due to the decreased coordination ability of Co2+, the decomposing
temperature of Co(HBTC)(4,4-bipy).3DMF is about 473 K. The porous structures are
prepared in DMF solution and the pores are filled with DMF molecules. The first major
weigh loss step34%corresponds to the removal of three DMF molecules from the
framework (the calculated weight change for three DMF molecules per M(HBTC)(4,4-
bipy).3DMF is 34%). The following step, in the region of high temperatures, corresponds to
the decomposition of the structure. The Nitrogen sorption tests reveal that the BET surface
area is 1590 m2/g for Ni(HBTC)(4,4-bipy).3DMF (activated at 453 K) and 887 m2/g for
Co(HBTC)(4,4-bipy).3DMF (activated at 433 K). The corresponding total pore volumes are
0.81 mL/g and 0.54 mL/g, respectively.

Figure 6. TGA curves of (a) Ni(HBTC)(4,4-bipy).3DMF (dashed) and (b) Co(HBTC)(4,4-


bipy).3DMF (solid line).

Hydrogen sorption isotherms at 77 K, 90 K and 98 K have been measured and shown in


Figure 7. All isotherms show full reversibility without hysteresis. At 77 K, Ni(HBTC)(4,4-
bipy).3DMF has the hydrogen capacity of 3.42 wt% and Co(HBTC)(4,4-bipy).3DMF has the
saturated capacity of 2.05 wt%. To investigate the interaction between H2 and the framework,
Clausius-Clapeyron equation, used in physical adsorption analysis [18], has been adopted in
the calculations of adsorption heats. To extract the coverage dependent adsorption heat for the
compound, the data measured at 77 K, 90 K and 98 K are caculated by the equation:

Qst
ln P n C
RT

where P is the pressure, n is the amount adsorbed, T is the temperature, R is the universal gas
constant and C is a constant. Hydrogen adsorption heat in M(HBTC)(4,4-bipy).3DMF is
obtained by calculating the slope of the lnP-(-1/RT) curve. The heats of Ni(HBTC)(4,4-
bipy).3DMF and Co(HBTC)(4,4-bipy).3DMF show gradual decrease in the values as a
function of the adsorbed amounts of H2. This decrease indicates that the H2 molecules are
combined with stronger biding sites preferentially, and it is possibly attributed to different
interacting sites located at different types of channels in M(HBTC)(4,4-bipy).3DMF.
110 Yaoqi Li, Ping Son, Yan Li et al.

Particularly, different from the results at low temperatures of 77 K, 90 K and 98 K,


M(HBTC)(4,4-bipy).3DMF presents a linear increasing trend on the hydrogen adsorption
curve up to 1.20 wt% for Ni(HBTC)(4,4-bipy).3DMF and 0.96 wt% for Co(HBTC)(4,4-
bipy).3DMF under 7.20 MPa at 298 K. The hydrogen storage capacity is proportional to the
applied pressure from 0.01 MPa to 7.20 MPa. The slopes are 0.162 wt% per MPa and 0.133
wt% per MPa, respectively. No saturation is observed at room temperature and this
phenomenon is consistent with previous reports [1b, 5c, d, 17, 19]. It has been published that
IRMOF-1 showed 0.45 wt% linear hydrogen uptake under 60 bar (0.075 wt% per MPa) at
298 K and IRMOF-8 presented the capacity of 0.4 wt% at 30 bar (0.133 wt% per MPa) [20].
Considering the smaller surface areas and lower hydrogen capacities at nitrogen temperature
of M(HBTC)(4,4-bipy).3DMF, it is notable that M(HBTC)(4,4-bipy).3DMF presents even
higher hydrogen uptake and steeper slopes of linear isotherms than those of IRMOFs at 298
K. The high capacities of M(HBTC)(4,4-bipy).3DMF at 298 K are probably attributed to its
specific framework. Compared to the common rectangle channels in IRMOFs, the
M(HBTC)(4,4-bipy).3DMF contains nonlinear honeycomb channels with only 0.5 nm at the
narrowest and 0.8 nm at the widest spacing and rectangle channels with the size of 0.7 nm
0.6 nm [18c]. The honeycomb-liked channels with rolling surface in M(HBTC)(4,4-
bipy).3DMF should affect the interaction of adsorbent and hydrogen and increase the
hydrogen adsorption properties at room temperature. Researchers have also proposed that the
curvature of the framework, has also been a factor to determine hydrogen uptake [14,9]. The
nonlinear honeycomb channels may contribute to the interaction between adsorbent and
hydrogen molecules, which is significant for increasing the hydrogen capacity of MOFs at
ambient temperatures. Furthermore, the specific rolling surface of the channels with the
appropriate size for hydrogen sorption should induce high capacity of hydrogen. The channel
sizes in M(HBTC)(4,4-bipy).3DMF are appropriate for hydrogen molecules with the van der
waals diameter of 0.24 nm to enter the porous structure and be adsorbed. Besides, considering
the organic ligands (H3BTC, bipy) and the framework structure, it is notable the entire surface
exposed to gas is the electron cloud of -bands. The interaction of -bands with hydrogen
molecules might also contribute to the adsorption [21]. Recent studies on neutron diffraction,
inelastic neutron scattering (INS) spectroscopy, and computational methods have also shown
that aromatic ligands provide strong interaction sites with hydrogen [22].

Figure 7. Left: Hydrogen adsorption (shaded symbols) and desorption (open symbols) isotherms for
Ni(HBTC)(4,4-bipy).3DMF at (a) 98 K, (b) 90 K and (c) 77 K. Right: Hydrogen adsorption (shaded
symbols) and desorption (open symbols) isotherms for Co(HBTC)(4,4-bipy).3DMF at (a) 98 K, (b) 90
K and (c) 77 K.
The Applications of Metal Organic Frameworks 111

2.4. Ni2(BTEC)(Bipy)33DMF2H2O and Ni2(BDC)2(Dabco)4DMF1.5H2O

Moreover, our group have synthesized two pillared-layer MOFs with transition metal
ions and mixed-ligand systems, Ni2(BTEC)(bipy)33DMF2H2O and
Ni2(BDC)2(dabco)4DMF1.5H2O (BTEC = 1,2,4,5-benzenetetracarboxylate; bipy = 4,4-
bipydine; BDC = 1,4-benzenedicarboxylate; dabco = 1,4-diazabicyclo[2.2.2]octane; DMF =
N,N-dimethylformamide). In Ni2(BTEC)(bipy)33DMF2H2O, the layer is formed by infinite
Ni-BTEC ladder chains interlinked by one type of bipy and further pillared by the other type
of bipy to form a 3D structure. The windows of the two type channels (along [100] and [010])
are both 0.7 0.7 nm2, with the channel along [100] direction much more open [23]. The
layers in Ni2(BTEC)(bipy)33DMF2H2O are formed by both O- and N- donor coordination
ligands and much thicker than the ones in Ni2(BDC)2(dabco)4DMF1.5H2O. The 3D
structure in Ni2(BDC)2(dabco)4DMF1.5H2O is formed by square-grid Ni2(BDC)2 layers
further interlinked by dabco, and the two windows of interlacing channels are supposed to be
of similar sizes with other M2(BDC)2(dabco) (Co2(BDC)2(dabco): 0.76 0.76 nm2 and 0.51
0.37 nm2 [24]; Zn2(BDC)2(dabco) and Cu2(BDC)2(dabco): 0.75 0.75 nm2 and 0.480.32
nm2 [25]). The BET surface area of Ni2(BTEC)(bipy)3 is 766 m2g-1 and Ni2(BDC)2(dabco)
showed a BET surface area (1809 m2/g). For hydrogen uptake, the excess isotherms reach the
saturation at 12.0 (Ni2(BTEC)(bipy)3) and 28.1 (Ni2(BDC)2(dabco)) bar with a capacity of
1.61 and 3.76 wt%, respectively. Compound Ni2(BTEC)(bipy)3 and Ni2(BDC)2(dabco) shows
a hydrogen uptake of 1.28 and 1.99 wt% at 77K 1 bar (Figure 9), respectively. The heats of
hydrogen adsorption of Ni2(BTEC)(bipy)3 and Ni2(BDC)2(dabco) were calculated with the
Clausius-Clapeyron equation. At zero coverage, the adsorption heats of Ni2(BTEC)(bipy)3
and Ni2(BDC)2(dabco) are 7.08 and 5.83 kJ/mol, and decrease as more hydrogen adsorbed.
As shown in Table 3, the hydrogen uptake of Ni2(BDC)2(dabco) at 77 K 1 bar is higher than
Cu2(BDC)2(dabco), but lower than Zn2(BDC)2(dabco) and Co2(BDC)2(dabco).

Figure 8. Left: Paking diagram in 1 viewed along the a (a) and c (b) axis; Right: Paking diagram in
Co2(BDC)2(dabco) viewed along the a (c) and c (d) axis24. Ni, green; Co, deep blue; N, blue; O, red; C,
black; H, white.
112 Yaoqi Li, Ping Son, Yan Li et al.

The Qst of M2(BDC)2(dabco) (M= Zn, Cu, Ni) were 5.3-5.0, 6.1-4.9 and 5.3-5.0 kJmol-1.
This indicates that the hydrogen affinity of the three MOFs is almost the same under low
hydrogen coverage. It is concluded the unexposed metal ions have little effect on hydrogen
adsorption properties in M2(BDC)2(dabco), which is consistent with the infrared adsorption
spectroscopy measurements reported by Nijem N. et al. [26] This is probably because metal
ions in the guest-free samples of M2(BDC)2(dabco) are fully coordinated.

Table 3. Comparisons of M2(BDC)2(dabco) (M = Ni, Co, Zn, Cu)

Metal sites BET surface H2 uptake at Qst Reference


area (m2/g) 77K 1 bar (wt%) (kJ/mol)
Ni 1809b 1.99 5.3-5.0 This book
(0.01-0.7) a
Co 1707b 2.27 Not reported Wang, H. et al.24
Zn 1794c 2.1 5.3-5.0 Lee J.Y. et al.25
(0.01-0.7)
Cu 1461c 1.8 6.1-4.9 Lee J.Y. et al.25
(0.02-0.6)
a
: the numbers in brackets indicate the range of hydrogen coverage; b: BET (Brunauere-Emmette-Teller)
surface area estimated by nitrogen adsorption isotherms at 77 K in this work; c: BET surface area
estimated by argon adsorption isotherms at 87 K by Lee J.Y. et al.[25].

Figure 9. Lower-pressure hydrogen adsorption and desorption isotherms for 1 and 2 at 77 and 88 K.
The shaded symbols and open symbols represent adsorption and desorption.
The Applications of Metal Organic Frameworks 113

Therefore, in the MOFs with saturated metal ions such as M2(BDC)2(dabco), the metal
ions might have little effect, and the strongest hydrogen adsorption sites provide multi
interactions. An effective means of increasing the adsorption enthalpy is the introduction of
distorted microporous structures and rolling surfaces inside MOFs, which strengthens the
affinity of hydrogen molecules through multi interactions between hydrogen molecules and
aromatic ligands.
As mentioned above, MOFs can be potentially applied as hydrogen storage materials.
The microporous porosity, the large specific surface area and especially the controllable
framework have made MOFs superior to traditional inorganic porous materials. The hydrogen
uptake properties are greatly affected by the various exposed metal sites, the coordination
ligands and the corresponding crystal structures. By chemical engineering, the chemical
nature and the pore structure of MOFs can be well adjusted. That is to say, we can improve
the hydrogen sorption properties of MOFs by choosing the optimal MOFs. However, the
practical hydrogen storage materials call for the combination of high gravimetric/volumetric
hydrogen density, adequate working temperature, proper kinetics, reversibility, low cost and
low toxicity. One existing problem of using MOFs as hydrogen storage materials is the low
operating temperature. Therefore, it is crucial to increase the interaction between MOFs and
hydrogen in order to enhance the hydrogen storage capacities of MOFs at room temperatures.
Either exposed metal sites or the rolling/distorted surface should be considered to enhance
this interaction.

3. MOFS USED AS CATALYST PRECURSORS


Simultaneously, metal-organic frameworks (MOFs) have also been focused on in the
field of catalysis in recent years. The high specific surface areas, the uniform microporous
pores, and the framework structures have made MOFs promising candidates for catalysis.
MOFs contain abundant microporous pores and channels, which offer convenient conditions
for small molecules to access the interior surface and each metal site. The MOFs in principle
should promote the formation of active catalytic metal sites and provide porous support.
Furthermore, according to crystal engineering7, the MOFs family contains different
frameworks with various structures, metal ions and organic ligands. This huge diversity
should make MOFs possess a variety of catalytic activities in the chemical reactions.
On the other hand, among various chemical hydrides of interest, nowadays ammonia
borane (NH3BH3) has become a very promising candidate for the chemical hydrogen-storage
application. Among the possible classes of hydrogen storage materials, ammonia borane
(NH3BH3, AB) has attracted increasing attention as an efficient and lightweight storage
medium for hydrogen [27], owing to its high capacity of H2 (19.6 wt% by weight and 0.145
kgL-1 by volume) [28], the low molecular weight (30.7 gmol-1), the stability in ambient
atmosphere [27a, 27e] and the exothermic nature of the decomposition process [29]. NH3BH3
is a colourless solid at room temperature. Both pure NH3BH3 and the corresponding aqueous
solution are very safe and stable at room temperature. H2 can be generated from NH3BH3
through thermal decomposition, [30] catalyzed dehydrocoupling in organic solutions, [31]
and solvolysis (hydrolysis, methanolysis et al.). [30a, 32]
The hydrolysis of NH3BH3 can be briefly expressed as follows:
114 Yaoqi Li, Ping Son, Yan Li et al.

NH3BH3 + 2H2O NH4 + + BO2 + 3H2

The thermal decomposition of NH3BH3 can be briefly expressed as follows:

For hydrolysis system, the NH3BH3 aqueous solution is very stable at room temperature. For
NH3BH3 aqueous solution without catalysts, the sluggish kinetics of hydrogen release from
the hydrolysis reaction at room temperature greatly affects the practical application of this
hydrolysis system. Development of efficient catalysts to improve the kinetics of hydrogen
generation from aqueous NH3BH3 solution at room temperatures is of extreme significance
for the future application.
For thermal decomposition, as shown above, the thermal analysis results show that during
the thermolysis, NH3BH3 can release H2 via three steps [27b, 29, 33]. Below 200 C,
approximately 2 equiv. H2 can be released from 1 equiv. AB. The two decomposition steps
are associated with the formation of solid products polyaminoborane (NH2BH2)n and
polyiminoborane (NHBH)n. Other undesirable volatile by-products generated from thermal
decomposition are dependent on the rate of temperature increase [27b, 27e]. The final
decomposition of [NHBH]n occurs at very high temperatures ( 500 C) and thus is not
considered in practical hydrogen storage. Currently, the application of NH3BH3 thermal
decomposition has been greatly limited by the long induction period and the high kinetic
barrier, which result in the sluggish H2 releasing rate of neat NH3BH3 at 100 C. Other
drawbacks include the emission of a small fraction of volatile by-products (borazine,
(N3B3H6), et al.) and the lack of economical viable methods for spent fuel regeneration. [27b,
34] To overcome these difficulties, tremendous efforts have recently been devoted to enhance
the hydrogen release from AB. Autrey T. and coworkers [35] investigated the thermal
decomposition of NH3BH3 dispersed in mesoporous silica (AB: silica = 1: 1 by weight) and
obtained a remarkably improved kinetics. By dissolving NH3BH3 in ionic liquid, Sneddon
L.G. and co-workers [36] observed the release of 0.95 equiv. H2 within 3 h at 85 C.
Furthermore, Heinekey D.M. [31a], Baker R. T. [31b] and Manners I. et al. [37] examined the
dehydrocoupling of NH3BH3 and suggested that the transition metals such as Rh, Ir and Ni-
based catalysts could effectively accelerate the dehydrogenation of NH3BH3 dissolved in
organic solvents. These advances have led to significant improvements of both the rate and
the extent of H2 release from AB. However, in most NH3BH3 systems, the modified H2
generation requires the utilization of ionic liquid, organic solvent or porous supporting
material, which inevitably brings high additional weight to AB. The H2 density of the whole
system is considerably lowered. The catalyzed direct thermal decomposition of solid-state
NH3BH3 with only a small amount of additional highly active catalysts should be greatly
attractive to obtain a system with high energy density.
As mentioned above, efficient catalysts are essential for further development of both the
hydrolysis and thermal decomposition of AB. In our group, we use MOFs as catalyst
precursors to synthesize a kind of highly effective MOF-based catalysts. The novel MOF-
The Applications of Metal Organic Frameworks 115

based catalysts show great catalytic activities in both NH3BH3 hydrolysis system and
NH3BH3 thermal decomposition system, the detailed research are shown below.

3.1. MOF-Based Catalyst Used in the NH3BH3 Hydrolysis System

In NH3BH3 hydrolysis system, the MOF of Ni(4,4-bipy)(HBTC) (MOF 1, H3BTC =


1,3,5-benzenetricarboxylic acid; 4,4-bipy = 4,4-bipyridine) [8] has been adopted [32b]. The
MOF, which has honeycomb channels (0.8 nm at the wide and 0.5 nm at the narrow spacing)
and rectangle channels (0.7 nm 0.6 nm), is used as the support and precursor of the catalyst.
The catalyst (MOF1cat) has been simply synthesized by immersing the MOF 1 into NH 3BH3
methanol solution in glove box at room temperature. NH3BH3 has been proved to be an
efficient and mild reducing reagent [30b, 32b, 38]. In our research, NH3BH3 is a mild
reducing agent in methanol for Ni2+ in MOF 1. During the synthesis process, the green
crystals of Ni(4,4-bipy)(HBTC) gradually turn into black powders, as shown in Figure 10.
The XPS (X ray photoelectron spectroscopy) results of the as-synthesize catalyst display the
peaks which can be ascribed to Ni, C, N and O elements. The Ni 2p XPS peaks can be
demonstrated into two peaks at 852.2 eV and 855.5 eV respectively, corresponding to the
presence of metallic Ni and divalent Ni. The C, N, O elements and divalent Ni should belong
to the reserved framework. During the preparation, NH3BH3 molecules diffuse into the pores
and the channels of Ni(4,4-bipy)(HBTC) (MOF 1) in porous structure. Part of Ni2+ ions have
been reduced by NH3BH3. The as-synthesized MOF1cat contains both Ni2+ in the remaining
MOF 1 and metallic Ni supported by the MOF 1. The transmission electron microscopy
(TEM) image shows that the MOF-based catalyst is composed of particles with an average
diameter of 100 nm. The particles have rough and mesoporous surface, which should be
caused by the in situ reducing reaction in the tiny pores and channels of the MOF. Moreover,
the catalyst particles form a good suspension in the NH3BH3 methanol solution. NH3BH3 has
been used effectively as both a reducing agent and a dispersing agent during the synthesis
process.
The hydrogen generation kinetics of this MOF-based catalyst (MOF1cat) from NH3BH3
hydrolysis system is shown in Figure 11a. No detectable H2 release is observed without the
catalyst. The normal nickel powder mixed with green MOF crystals by grinding method is
found inefficient to accelerate the hydrolysis of NH3BH3 at room temperature. However, with
MOF1cat (the molar ratio of MOF/NH3BH3 = 0.10, the concentration of NH3BH3 = 0.32 M),
the hydrogen generation reaches saturation within only 5 min at 298 K in air. The molar ratio
of the generated hydrogen to initial NH3BH3 is close to 3.0 (average value of 20 cycles),
indicating that the complete hydrogen generation corresponds to approximately 8.9 wt% of
the reactants (NH3BH3 and H2O, excluding solvent water). As shown in Figure 11b, the
hydrolysis of NH3BH3 is completed in approximately 11, 7, 6 and 5 min at MOF/ NH3BH3
molar ratios of 0.02, 0.05 and 0.08, 0.10 respectively. Further increase of MOF/NH3BH3 ratio
has almost no effect on the reaction rate. For all NH3BH3 concentrations from 0.16 M to 0.97
M, the amount of H2 generated increases nearly linearly with time before saturation (Figure
11c), exhibiting a quasi zero-order behavior. Different from other nano-sized catalysts, the
formed boracic precipitate seems not to block the active sites supported on the MOF 1 and the
high activity of MOF1cat can be kept at high NH3BH3 concentrations. The rate constant k of
hydrogen generation is determined in the presence of the MOF-based catalyst at different
116 Yaoqi Li, Ping Son, Yan Li et al.

temperatures (Figure 11d). According to the Arrhenius equation, the activation energy for the
hydrogen generation reaction catalyzed by the MOF-based catalysts has been calculated to be
26 kJmol-1. The high reaction rate and low activation energy suggest the high catalytic
activity of MOF-based catalyst, which is comparable with that of expensive platinum-based
catalysts. Furthermore, the catalyst can be recycled by centrifugal separation and reused up to
20 cycles without obvious loss of activity.
Compared to the reported Ni/-Al2O3 catalyst [39], Ni1-xPtx hollow spheres catalyst [40]
and nanopowders by using Ni2+ salts as the precursor [32a], in this research, MOF1cat shows
unexpected high catalytic activity in kinetic properties. Recently, the in situ synthesized Fe
nanoparticles have been reported to show excellent catalytic activity for the hydrolysis of
NH3BH3[20]. The hydrolysis catalyzed by Fe particles has been accomplished in 8.5 minutes.
For the MOF-based catalyst, under the same condition (the concentration of NH3BH3 is 0.16
M; at room temperature; in the air), the complete hydrogen release has been completed in 2
minutes.

Figure 10. Photos of a vial during the synthesis process: (a) MOF in NH3BH3 methanol solution; (b), (c)
and (d) MOF reacted with NH3BH3 for 15 min, 30 min and 45 min; (e) and (f) the as-synthesized
catalyst dispersed in methanol and aqueous solutions.

Figure 11. Hydrogen generation from the hydrolysis of NH3BH3 (0.32 M, 1.0 mL) in the presence of (a)
different catalysts and without any catalyst at 298 K in air; (b) MOF-based catalyst with MOF/NH3BH3
molar ratios of 0.02, 0.05, 0.08 and 0.10. (c) Hydrogen release from aq. NH 3BH3 solution (1.0 mL) with
different concentrations (0.16 M, 0.32 M, 0.65 M and 0.97 M) in the presence of MOF-based catalyst
(20 mg) and (d) hydrogen generation from aq. NH3BH3 (0.32 M, 1.0 mL) in the presence of MOF-
based catalyst (MOF/ NH3BH3 = 0.10) at 273, 298 and 313 K.
The Applications of Metal Organic Frameworks 117

The unexpected outstanding catalytic performance of the MOF-based catalyst should be


ascribed to both the special generantion procedure of the metallic Ni and the microporous
structure of Ni(4,4-bipy)(HBTC). The porous framework with large BET (Brunauer-
Emmett-Teller) surface area (1590 m2/g) not only offers enough channels for NH3BH3 to
access the interior surface, but also contains abundant metal and organic sites to adsorb
NH3BH3. Since the Ni2+ sites are well separated by the organic ligands in the MOF structure
and are accessed by NH3BH3 molecules independently. The in situ generated metallic Ni,
which is reduced by NH3BH3 and synthesized from the framework structure, is very tiny
clusters that are highly active in catalysis. In addition, the microporous structure of the MOF
support results in much higher density of active sites on the surface (both exterior and
interior) than normal nickel nanoparticles. Furthermore, the in situ generated Ni clusters are
stabilized by the MOF 1 framework and the conglomeration, which is almost inevitable for
very small metal clusters, is inhibited to great extent in the present catalysis system. As a
result, the high catalytic activity can be maintained during repeated cycles.

3.2. MOF-Based Catalyst Used in the NH3BH3


Thermal Decomposition System

In our efforts to promote the thermal decomposition of AB [41], we use two efficient
MOF-based catalysts (MOFcats). We report adopt the mild reduction method as mentioned
above to synthesize MOFcats from Ni(4,4-bipy)(HBTC) (MOF 1) and Ni(pyz)[Ni(CN)4]
(MOF 2). The MOF 1 (Ni(4,4-bipy)(HBTC)), MOF 2 (Ni(pyz)[Ni(CN)4]) and MOF 3
(Ni3[Fe(CN)6]2) used in this work, as shown in Figure 12, are characterized by large BET
surface areas (1570 m2g-1, 518 m2g-1 and 718 m2g-1), good thermal stabilities and uniform
microporous structures. As discussed previously, the microporous MOFs serve as excellent
precursors and effective supports of the as-synthesized catalysts. The porous channels in
MOFs in principle allow for the high dispersion of reactant molecules and therefore promote
the synthesis of the MOFcats. In MOFcats, both the mild synthesis and the confinement of the
catalytic sites within microporous framework may result in unexpected high catalytic activity
and stability. In this study, MOF-1, 2, 3 based catalysts (MOF1cat, MOF2cat and MOF3cat)
are prepared respectively from MOF 1, 2, 3 by the reduction reaction in NH3BH3 methanol
solution. During the preparation, the NH3BH3 methanol solution could be infiltrated into the
internal pores of MOFs. The special structures of MOFs containing both channels and well
separated Ni2+ sites in promote the NH3BH3 molecules to access and react with each Ni2+ ion
in the frameworks independently. When treated with NH3BH3 methanol solution, the green
crystals of MOF 1 and the purple powders of MOF 2 gradually turn into black powders.
Importantly, the Ni 2p XPS peaks of the as-synthesized MOF1cat and MOF2cat can be
demonstrated into two peaks at 852 eV and 856 eV (Figure 13 Left), indicating the presence
of both metallic Ni (852 eV) and divalent Ni (856 eV). The results reveal that in MOF1cat
and MOF2cat, part of Ni2+ ions in the original MOFs have been reduced to zero valent Ni,
which has been proven to be catalytically active in the hydrolysis system. The Ni2+ peak (856
eV) should be attributed to the remaining frameworks of MOF 1 and MOF 2. The comparison
between the peak area of Ni0 (852 eV) and that of Ni2+ (856 eV) in XPS spectra indicates that
more Ni2+ has been reduced to Ni0 in MOF1cat than in MOF2cat. Besides, the as-prepared
MOF 1cat and MOF2cat also display the XPS peaks that could be assigned to elemental C, N,
118 Yaoqi Li, Ping Son, Yan Li et al.

O and Ni. The C, N ,O elements, which should belong to the organic ligands constructing the
framework, confirm the existing of the reserved MOF structures in MOF1cat and MOF2cat.
Different from the results of MOF1cat, which reveal that MOF1cat is in the amorphous phase,
the XRD patterns of MOF2cat are alike those of pure MOF 2, indicating that in MOF2cat,
both the crystal structure and the microporous framework of MOF 2 have been unexpected
well kept during the synthesis. The XRD results demonstrate that the MOF2cat consists of
amouphous metallic Ni supported on the well reserved microporous MOF 2. The difference
of the framework crystallinity between MOF1cat and MOF2cat, should be attributed to the
reduction of more Ni2+ sites to Ni0 during the preparation of MOF1cat, as mentioned
previously. Owing to the special precursors (MOFs) and the mild reduction method,
MOF1cat and MOF2cat with metallic Ni0 sites and framework supports have been
successfully synthesized. The MOF1cat with amorphous metallic Ni and mesoporous surface
should be anticipated to have high concentration of active Ni0 sites. In MOF2cat, the metallic
Ni supported on well reserved MOF 2 framework should be anticipated to have modified
catalytic activity due to the supporting effect of the remaining MOF as well as specially
synthesized Ni0. Both the amorphous Ni0 sites and the framework support in MOF1cat and
MOF2cat should greatly enhance catalytic activity. Such features can clearly promote the
easier release of H2 and enhance the dehydrogenation kinetics evidenced below.
For MOF3cat, the results are quite different. No significant colour change of MOF 3 has
been observed during the synthesis process. Moreover, the Ni 2p XPS spectrum of the as-
prepared sample shows only one peak at 856 eV, indicating only divalent Ni constructing the
framework. It should be ascribed to the strong bonds between metal ions and the CN ligands
which prevent the reduction reaction of Ni2+ in MOF 3. The XRD patterns of MOF3cat
further prove that there is no significant change of MOF 3 after treated with the reducing
reagent of AB.
1.0 mol% MOFcat (MOF1cat and MOF2cat) was mixed with NH3BH3 by hand in an
agate mortar. The only added weight to NH3BH3 is 1.0 mol% catalytic additive. Therefore,
the total hydrogen capacity of the system has not been sacrificed significantly. Figure 13
(Right) shows the differential scanning calorimetry (DSC) curves of the neat and the catalyst-
doped NH3BH3 samples. Identical to what was reported [27b, 27e], neat NH3BH3 released the
first equivalent of H2 at ~ 114 oC and the second equivalent of H2 with a board DSC peak
centred at ~ 161 oC with a heating rate of 5 oCmin-1. Compared with the neat AB,
NH3BH3/MOF1cat releases H2 with one peak centred at a low temperature of 80 oC and
another peak at 101 oC. For NH3BH3/MOF2cat, the decomposition occurs with a desorption
DSC peak at 102 oC and gives a peak of very low intensity at around 150 oC. MOF1cat lower
the dehydrogenation onset temperatures more effectively than MOF2cat. It should possibly be
attributed to the existing abundant catalytic Ni0 sites in MOF1cat. Both MOF1cat and
MOF2cat consist of zero-valent Ni and the remaining framework support. The Ni0
synthesized from MOFs, as reported in the hydrolysis system [32b], could present unexpected
high catalytic activity. It is probable that the large amount of amorphous Ni0 sites in MOF1cat
greatly enhance the kinetics and lower the reaction onset temperature. The effect of Ni0 sites
has been further proved by MOF3cat. As mentioned previously, no Ni0 in MOF3cat has been
observed. Different from those of NH3BH3/MOF1cat and NH3BH3/MOF2cat, DSC curve of
NH3BH3/MOF3cat shows no significant difference from that of neat AB, indicating that only
with active amorphous Ni0 sites, the MOFcats can possess high catalytic activity and
efficiently lower the dehydrogenation temperatures.
The Applications of Metal Organic Frameworks 119

Moreover, neat NH3BH3 generates an endothermic dip at ~ 108 oC, which is ascribed to
the melting phase transition of AB, whereas for NH3BH3/MOF1cat and NH3BH3/MOF2cat,
there was no endothermic peak in DSC curves. It should be caused by the catalytically
dehydrogenation of NH3BH3/MOFcats occurring at temperatures prior to the melting point of
AB. DSC results of NH3BH3 and NH3BH3/MOFcats also illustrate that the decomposition
reaction enthalpies from NH3BH3/MOF1cat and NH3BH3/MOF2cat (H = - 4.3 kJmol-1 and
7.9 kJmol-1) are significantly less than that from the neat NH3BH3 (H = - 21 kJmol-135, 42).
The differences of H further indicate that the MOFcats with Ni0 sites lead to a change in the
thermodynamics of the thermal decomposition, which might be the reason that the
NH3BH3/MOFcats can generate H2 at lower temperatures.
Isothermal volumetric H2 release measurements also present distinct features among the
neat and the MOFcats doped NH3BH3 samples. As shown in Figure 14 (Left), very little
hydrogen generation has been detected from the neat NH3BH3 after being kept at 80 oC for
2.5 h, which is probably due to the long induction period for the formation of the initiator
(diammoniate of diborane, DADB) [30a, 33a] and/or the high kinetic barrier in
dehydrogenation. In contrast, NH3BH3/MOF1cat can release ~ 7.5 wt% H2 within 2 h and
NH3BH3/MOF2cat generates ~ 6.0 wt% H2 within only 40 minutes at 80 oC.

Figure 12. The frameworks of (a) MOF 1, (b) MOF 2 and (c) MOF 3.

Figure 13. Left : Ni 2p X-ray photoelectron spectra of MOF1cat, MOF2cat and MOF3cat. Right: The
DSC results of NH3BH3/MOF1cat, NH3BH3/MOF2cat, NH3BH3/MOF3cat and neat AB.
120 Yaoqi Li, Ping Son, Yan Li et al.

Importantly, the NH3BH3/MOF1cat does not suffer from the long induction period for the
release of H2 that is present in neat NH3BH3 (Figure 14). Moreover, the kinetic behaviours of
NH3BH3/MOF2cat is unexpected different. After a short induction period of ~ 20 minutes, the
NH3BH3/MOF2cat evolves H2 vigorously at a very high releasing rate of 0.48 wt% per
minute. There are several possible reasons for the remarkable accelerated kinetics and
differences in the dehydrogenation of NH3BH3/MOF1cat and NH3BH3/MOF2cat. One is the
active Ni0 sites. As proved by the DSC results, the enhanced kinetic behaviours of
NH3BH3/MOFcats systems should be attributed to the prepared Ni0 sites. As mentioned
previously, in MOF1cat, there exist lots of active metallic Ni0 sites, which have been proved
quite effective to lower the reaction onset temperatures and enhance the kinetics. It is quite
possible that the abundant amorphous Ni0 sites in MOF1cat lead to the fast H2 evolution
without any induction period in NH3BH3/MOF1cat. Another possibility is the reserved
framework, which provides a good porous support for the active sites and exerts catalytic
influence on the dehydrogenation. The stable microporous structure of MOFs is by nature a
good substrate for highly dispersed tiny metal clusters. Compared with MOF1cat, MOF2cat
contains better reserved framework after the reduction synthesis. In NH3BH3/MOF2cat
sample, the Ni0 sites might not be sufficient enough to reach a very low reaction onset
temperature as discussed previously and a kinetic process without induction. However, the
isothermal experimental results indicates that the Ni0 sites specially dispersed and supported
on the well remained porous MOF 2 can greatly speed up the reaction rate after the short
induction period. The supporting effect of well reserved MOF combined with amorphous Ni0
sites leads to a further significant acceleration of the kinetics in NH3BH3/MOF2cat. Although
a short induction period seems inevitable for the whole process in NH3BH3/MOF2cat due to
the lack of sufficient Ni0 sites, the H2 releasing rate of NH3BH3/MOF2cat is apparently high
(0.48 wt% H2 per minute).
For the systems of NH3BH3/MOFcats, the overall features of hydrogen release are clearly
different the sigmoidal kinetic behaviour of the pristine NH3BH3 which has been reported35,
43
. It reveals that the mechanism of the catalytic dehydrogenation of NH3BH3/MOFcats might
be different from the initiated thermal decomposition. It has also been supported by the
hypothesis of different thermodynamics in NH3BH3/MOFcat system in the first
dehydrogenation DSC step discussed above. To obtain further insights of the enhanced
kinetics, the activation energy has been determined by the Kissinger equation, namely

ln( / TP 2 ) ln( AR / Ea ) Ea / RTP

is the heating rate, Tp is the temperature of the maximum reaction rate peak, A is the pre-
exponential factor, R is the gas constant and Ea is the activation energy. DSC tests have been
performed with various heating rates and the dependence of ln( / TP ) and 1000/T has been
2
p
given in Figure 14 (Right). The activation energy of the hydrogen release of
NH3BH3/MOF1cat and NH3BH3/MOF2cat is determined to be ~ 131 kJmol-1 and ~ 1605
kJmol-1, respectively, which are lower than that of the neat NH3BH3 (184 kJmol-1)35. The
obvious decreases of Ea in NH3BH3/MOFcats further provide direct evidence for the
possibility of enhancing kinetics of the NH3BH3 decomposition with MOFcats. Compared
The Applications of Metal Organic Frameworks 121

with NH3BH3/MOF2cat, MOF1cat shows even lower Ea, which should be attributed to the
abundant Ni0 sites as discussed previously. The amorphous Ni0 sites serve as active metal
centres, which efficiently lower the kinetic barrier.

Figure 14. Left: The volumetrically measured H2 release from NH3BH3/MOF1cat, NH3BH3/MOF2cat
and neat NH3BH3 at 80 oC. Right: Kissinger plots of the MOF1cat-doped() and MOF2cat-doped()
NH3BH3 samples.

The above results suggest that MOF1cat and MOF2cat can provide a very effective
strategy to significantly modify the enthalpy of decomposition, speed up the kinetics and
lower the activation barrier in the thermal dehydrogenation of AB. The unexpected
outstanding catalytic performances of the MOF-based catalysts should be ascribed to both the
mild generation procedure of the metallic Ni0 sites and the special framework support. The
compositional and structural diversity of the MOFs (MOF 1, MOF2 and MOF 3) leads to
obviously distinct catalytic activities and behaviours of the corresponding MOFcats during
the thermal decomposition of AB. The obviously distinct catalytic activities between the three
catalysts should be attributed to the different structures and ligands of original MOFs. For the
MOFs constructed by ligands with relatively strong coordination ability to the corresponding
metal ions, it should be difficult to obtain catalytically active metallic sites (Ni0) from Ni2+
ions, as discussed previously (MOF3cat). However, if the coordination between ligands and
metal ions in MOFs is weak, as illustrated in MOF1cat, the framework support can not be
reserved properly after the reduction process of Ni2+. For future development, considering the
crystal engineering, by a proper choice of metal ions, organic ligands and reducing agent, we
may find a suitable MOF and the corresponding MOFcat containing both sufficient active Ni0
sites and properly reserved MOF support. This new MOFcat should in principle further
enhance the catalytic activity and kinetics in the dehydrogenation of AB. These results not
only present a novel kind of efficient catalysts, but also provide new insights into the
application of MOFs in the field of catalysis. More importantly, for future development, both
the framework and the metal sites in MOFs can be adjusted and modified to obtain novel and
effective MOF-based catalysts for various catalytic processes.
A good catalyst is a big stride towards the application of the many hydrogen storage
systems. Importantly, the in situ generation of metal clusters in the MOF structure described
here provides a new strategy to produce highly efficient catalysts. The stable microporous
structure of MOFs is by nature a good substrate for highly dispersed tiny metal clusters. By
122 Yaoqi Li, Ping Son, Yan Li et al.

proper choice of metal ions, organic ligands and the reducing agents, a large number of highly
efficient MOF-based catalysts can be created for various catalysis applications.

CONCLUSION
As a summary, MOF, which is a novel kind of functional material, can be widely used in
the filed of gas storage, catalysis and so on. The researches of the hydrogen storage properties
of MOFs present potential for the application of MOFs in the filed of gas storage. By altering
the metal sites, the pore diameter and the porous structure of MOFs, it is quite possible to
obtain MOF materials with nice hydrogen storage properties for further research and practical
application. The discovery of the unexpected high catalytic effect of MOF-based catalysts
appears very promising and implies a new kind of catalysts that can be exploited. Considering
the large componential and structural diversity of MOFs, further development of the MOFcats
should be hopeful to obtain new catalysts with even better catalytic properties.

REFERENCES
[1] (a) Kaye, S. S.; Long, J. R., Hydrogen storage in the dehydrated Prussian blue
analogues M-3[Co(CN)(6)](2) (M = Mn, Fe, Co, Ni, Cu, Zn). J. Am. Chem. Soc. 2005,
127 (18), 6506-6507; (b) Rosi, N. L.; Eckert, J.; Eddaoudi, M.; Vodak, D. T.; Kim, J.;
O'Keeffe, M.; Yaghi, O. M., Hydrogen storage in microporous metal-organic
frameworks. Science 2003, 300 (5622), 1127-1129; (c) Kitagawa, S.; Kitaura, R.; Noro,
S., Functional porous coordination polymers. Angew. Chem., Int. Ed. 2004, 43 (18),
2334-2375; (d) Cho, S. H.; Ma, B. Q.; Nguyen, S. T.; Hupp, J. T.; Albrecht-Schmitt, T.
E., A metal-organic framework material that functions as an enantioselective catalyst
for olefin epoxidation. Chem. Commun. 2006, (24), 2563-2565.
[2] (a) Ma, B. Q.; Mulfort, K. L.; Hupp, J. T., Microporous pillared paddle-wheel
frameworks based on mixed-ligand coordination of zinc ions. Inorg. Chem. 2005, 44
(14), 4912-4914; (b) Rowsell, J. L. C.; Millward, A. R.; Park, K. S.; Yaghi, O. M.,
Hydrogen sorption in functionalized metal-organic frameworks. J. Am. Chem. Soc.
2004, 126 (18), 5666-5667.
[3] (a) Schlapbach, L.; Zuttel, A., Hydrogen-storage materials for mobile applications.
Nature 2001, 414 (6861), 353-358; (b) Grochala, W.; Edwards, P. P., Thermal
decomposition of the non-interstitial hydrides for the storage and production of
hydrogen. Chem. Rev. 2004, 104 (3), 1283-1315.
[4] Ma, S. Q.; Zhou, H. C., Gas storage in porous metal-organic frameworks for clean
energy applications. Chemical Communications 2010, 46 (1), 44-53.
[5] (a) Latroche, M.; Surble, S.; Serre, C.; Mellot-Draznieks, C.; Llewellyn, P. L.; Lee, J.
H.; Chang, J. S.; Jhung, S. H.; Ferey, G., Hydrogen storage in the giant-pore metal-
organic frameworks MIL-100 and MIL-101. Angewandte Chemie-International Edition
2006, 45 (48), 8227-8231; (b) Wong-Foy, A. G.; Matzger, A. J.; Yaghi, O. M.,
Exceptional H-2 saturation uptake in microporous metal-organic frameworks. J. Am.
Chem. Soc. 2006, 128 (11), 3494-3495; (c) Panella, B.; Hirscher, M.; Putter, H.; Muller,
The Applications of Metal Organic Frameworks 123

U., Hydrogen adsorption in metal-organic frameworks: Cu-MOFs and Zn-MOFs


compared. Adv. Funct. Mater. 2006, 16 (4), 520-524; (d) Dinca, M.; Dailly, A.; Liu, Y.;
Brown, C. M.; Neumann, D. A.; Long, J. R., Hydrogen Storage in a Microporous
Metal-Organic Framework with Exposed Mn2+ Coordination Sites. J. Am. Chem. Soc.
2006, 128 (51), 16876-16883; (e) Lin, X.; Jia, J. H.; Zhao, X. B.; Thomas, K. M.;
Blake, A. J.; Walker, G. S.; Champness, N. R.; Hubberstey, P.; Schroder, M., High H-2
adsorption by coordination-framework materials. Angew. Chem., Int. Ed. 2006, 45 (44),
7358-7364; (f) Loiseau, T.; Lecroq, L.; Volkringer, C.; Marrot, J.; Ferey, G.; Haouas,
M.; Taulelle, F.; Bourrelly, S.; Llewellyn, P. L.; Latroche, M., MIL-96, a porous
aluminum trimesate 3D structure constructed from a hexagonal network of 18-
membered rings and mu(3)-oxo-centered trinuclear units. J. Am. Chem. Soc. 2006, 128
(31), 10223-10230; (g) Surble, S.; Millange, F.; Serre, C.; Duren, T.; Latroche, M.;
Bourrelly, S.; Llewellyn, P. L.; Ferey, G., Synthesis of MIL-102, a chromium
carboxylate metal-organic framework, with gas sorption analysis. Journal of the
American Chemical Society 2006, 128 (46), 14889-14896.
[6] (a) Chae, H. K.; Siberio-Perez, D. Y.; Kim, J.; Go, Y.; Eddaoudi, M.; Matzger, A. J.;
O'Keeffe, M.; Yaghi, O. M., A route to high surface area, porosity and inclusion of
large molecules in crystals. Nature 2004, 427 (6974), 523-527; (b) Rowsell, J. L. C.;
Millward, A. R.; Park, K. S.; Yaghi, O. M., Hydrogen sorption in functionalized metal-
organic frameworks. Journal of the American Chemical Society 2004, 126 (18), 5666-
5667; (c) Collins, D. J.; Zhou, H. C., Hydrogen storage in metal-organic frameworks. J.
Mater. Chem. 2007, 17 (30), 3154-3160; (d) Noro, S.; Kitaura, R.; Kondo, M.;
Kitagawa, S.; Ishii, T.; Matsuzaka, H.; Yamashita, M., Framework engineering by
anions and porous functionalities of Cu(II)/4,4 '-bpy coordination polymers. J. Am.
Chem. Soc. 2002, 124 (11), 2568-2583; (e) Seo, J. S.; Whang, D.; Lee, H.; Jun, S. I.;
Oh, J.; Jeon, Y. J.; Kim, K., A homochiral metal-organic porous material for
enantioselective separation and catalysis. Nature 2000, 404 (6781), 982-986; (f) Janiak,
C., Engineering coordination polymers towards applications. Dalton Trans. 2003, (14),
2781-2804; (g) Rao, C. N. R.; Natarajan, S.; Vaidhyanathan, R., Metal carboxylates
with open architectures. Angew. Chem., Int. Ed. 2004, 43 (12), 1466-1496.
[7] Braga, D.; Brammer, L.; Champness, N. R., New trends in crystal engineering.
Crystengcomm 2005, 7, 1-19.
[8] Li, Y. Q.; Xie, L.; Liu, Y.; Yang, R.; Li, X. G., Favorable Hydrogen Storage Properties
of M(HBTC)(4,4 '-bipy)center dot 3DMF (M = Ni and Co). Inorganic Chemistry 2008,
47 (22), 10372-10377.
[9] (a) Chun, H.; Dybtsev, D. N.; Kim, H.; Kim, K., Synthesis, X-ray crystal structures, and
gas sorption properties of pillared square grid nets based on paddle-wheel motifs:
Implications for hydrogen storage in porous materials. Chem.Eur. J 2005, 11 (12),
3521-3529; (b) Chapman, K. W.; Chupas, P. J.; Maxey, E. R.; Richardson, J. W., Direct
observation of adsorbed H-2-framework interactions in the Prussian Blue analogue Mn-
3(II)[Co-III(CN)(6)](2): The relative importance of accessible coordination sites and
van der Waals interactions. Chem. Commun. 2006, (38), 4013-4015.
[10] (a) Li, H.; Eddaoudi, M.; O'Keeffe, M.; Yaghi, O. M., Design and synthesis of an
exceptionally stable and highly porous metal-organic framework. Nature 1999, 402
(6759), 276-279; (b) Eddaoudi, M.; Kim, J.; Rosi, N.; Vodak, D.; Wachter, J.; O'Keeffe,
124 Yaoqi Li, Ping Son, Yan Li et al.

M.; Yaghi, O. M., Systematic design of pore size and functionality in isoreticular MOFs
and their application in methane storage. Science 2002, 295, 469-472.
[11] Wong-Foy, A. G.; Matzger, A. J.; Yaghi, O. M., Exceptional H-2 saturation uptake in
microporous metal-organic frameworks. Journal of the American Chemical Society
2006, 128 (11), 3494-3495.
[12] Kitazawa. T; Gomi. Y; Takahashi. M; Takeda. M; M, E., Spin-Crossover Behavior of
Coordination Polymer FeII(C6H5N)2NiII(CN)4. J. Mater. Chem. 1996, 6, 119-121.
[13] (a) Bonhommeau, S.; Molnar, G.; Galet, A.; Zwick, A.; Real, J. A.; McGarvey, J. J.;
Bousseksou, A., One shot laser pulse induced reversible spin transition in the spin-
crossover complex Fe(C4H4N2 {Pt(CN)(4)} at room temperature. Angewandte
Chemie-International Edition 2005, 44 (26), 4069-4073; (b) Niel, V.; Martinez-Agudo,
J. M.; Munoz, M. C.; Gaspar, A. B.; Real, J. A., Cooperative spin crossover behavior in
cyanide-bridged Fe(II)-M(II) bimetallic 3D Hofmann-like networks (M = Ni, Pd, and
Pt). Inorganic Chemistry 2001, 40 (16), 3838-+; (c) Molnar, G.; Niel, V.; Real, J. A.;
Dubrovinsky, L.; Bousseksou, A.; McGarvey, J. J., Raman spectroscopic study of
pressure effects on the spin-crossover coordination polymers Fe(pyrazine) M(CN)(4)
center dot 2H(2)O (M = Ni, Pd, Pt). First observation of a piezo-hysteresis loop at room
temperature. Journal of Physical Chemistry B 2003, 107 (14), 3149-3155.
[14] (a) Goodwin, A. L.; Chapman, K. W.; Kepert, C. J., Guest-dependent negative thermal
expansion in nanoporous Prussian Blue analogues (MPtIV)-Pt-II(CN)(6)center dot
x{H2O} (0 <= x <= 2; M = Zn, Cd). Journal of the American Chemical Society 2005,
127, 17980-17981; (b) Ayrault, S.; LoosNeskovic, C.; Fedoroff, M.; Garnier, E.; Jones,
D. J., Compositions and structures of copper hexacyanoferrates(II) and (III):
Experimental results. Talanta 1995, 42 (11), 1581-1593.
[15] (a) Brunauer, S.; Emmett, P. H.; Teller, E., Adsorption of gases in multimolecular
layers. Journal of the American Chemical Society 1938, 60, 309-319; (b) Brunauer, S.;
Deming, L. S.; Deming, W. E.; Teller, E., On a theory of the van der Waals adsorption
of gases. Journal of the American Chemical Society 1940, 62, 1723-1732; (c) Everett,
D. H.; Powl, J. C., adsorption in slit-like and cylindrical micropores in henrys law
region - model for microporosity of carbons. Journal of the Chemical Society-Faraday
Transactions I 1976, 72, 619-636.
[16] Murray, L. J.; Dinca, M.; Long, J. R., Hydrogen storage in metal-organic frameworks.
Chemical Society Reviews 2009, 38 (5), 1294-1314.
[17] Banu Kesanli, Y. C. M. R. S. E. W. B. B. C. B. W. L., Highly Interpenetrated Metal-
Organic Frameworks for Hydrogen Storage. Angew. Chem., Int. Ed. 2005, 44 (1), 72-
75.
[18] (a) Lin, X.; Jia, J. H.; Hubberstey, P.; Schroder, M.; Champness, N. R., Hydrogen
storage in metal-organic frameworks. Crystengcomm 2007, 9 (6), 438-448; (b)
Krungleviciute, V.; Lask, K.; Heroux, L.; Migone, A. D.; Lee, J. Y.; Li, J.; Skoulidas,
A., Argon adsorption on Cu-3(Benzene-1,3,5-tricarboxylate)(2)(H2O)(3) metal-organic
framework. Langmuir 2007, 23 (6), 3106-3109; (c) Gao, C. Y.; Liu, S. X.; Xie, L. H.;
Ren, Y. H.; Cao, J. F.; Sun, C. Y., Design and construction of a microporous metal-
organic framework based on the pillared-layer motif. Crystengcomm 2007, 9 (7), 545-
547; (d) Zhou, W.; Wu, H.; Hartman, M. R.; Yildirim, T., Hydrogen and methane
adsorption in metal-organic frameworks: A high-pressure volumetric study. Journal of
Physical Chemistry C 2007, 111 (44), 16131-16137.
The Applications of Metal Organic Frameworks 125

[19] Fang, Q. R.; Zhu, G. S.; Xue, M.; Zhang, Q. L.; Sun, J. Y.; Guo, X. D.; Qiu, S. L.; Xu,
S. T.; Wang, P.; Wang, D. J.; Wei, Y., Microporous metal-organic framework
constructed from heptanuclear zinc carboxylate secondary building units. Chem.Eur.
J. 2006, 12 (14), 3754-3758.
[20] (a) Rowsell, J. L. C.; Yaghi, O. M., Effects of Functionalization, Catenation, and
Variation of the Metal Oxide and Organic Linking Units on the Low-Pressure
Hydrogen Adsorption Properties of Metal-Organic Frameworks. J. Am. Chem. Soc.
2006, 128 (4), 1304-1315; (b) Dailly, A.; Vajo, J. J.; Ahn, C. C., Saturation of hydrogen
sorption in Zn benzenedicarboxylate and Zn naphthalenedicarboxylate. J. Phys. Chem.
B 2006, 110 (3), 1099-1101.
[21] Li, Y.; Liu, Y.; Wang, Y. T.; Leng, Y. H.; Xie, L.; Li, X. G., Hydrogen storage
properties of [M(Py){Ni(CN)(4)}] (M = Fe, Co, Ni). Int. J. Hydrogen Energy 2007, 32
(15), 3411-3415.
[22] Mulder, F. M.; Assfour, B.; Huot, J.; Dingemans, T. J.; Wagemaker, M.; Ramirez-
Cuesta, A. J., Hydrogen in the Metal-Organic Framework Cr MIL-53. Journal of
Physical Chemistry C 2010, 114 (23), 10648-10655.
[23] Gao, C. Y.; Liu, S. X.; Xie, L. H.; Sun, C. Y.; Cao, J. F.; Ren, Y. H.; Feng, D.; Su, Z.
M., Rational design microporous pillared-layer frameworks: syntheses, structures and
gas sorption properties. Crystengcomm 2009, 11 (1), 177-182.
[24] Wang, H.; Getzschmann, J.; Senkovska, I.; Kaskel, S., Structural transformation and
high pressure methane adsorption of Co-2(1,4-bdc)(2)dabco. Microporous and
Mesoporous Materials 2008, 116 (1-3), 653-657.
[25] Lee, J. Y.; Olson, D. H.; Pan, L.; Emge, T. J.; Li, J., Microporous metal-organic
frameworks with high gas sorption and separation capacity. Advanced Functional
Materials 2007, 17 (8), 1255-1262.
[26] Nijem, N.; Veyan, J. F.; Kong, L. Z.; Li, K. H.; Pramanik, S.; Zhao, Y. G.; Li, J.;
Langreth, D.; Chabal, Y. J., Interaction of Molecular Hydrogen with Microporous
Metal Organic Framework Materials at Room Temperature. J. Am. Chem. Soc. 2010,
132 (5), 1654-1664.
[27] (a) Marder, T. B., Will we soon be fueling our automobiles with ammonia-borane?
Angew. Chem., Int. Ed. 2007, 46, 8116-8118; (b) Stephens, F. H.; Pons, V.; Baker, R.
T., Ammonia - borane: the hydrogen source par excellence? Dalton Trans. 2007, (25),
2613-2626; (c) Stephens, F. H.; Pons, V.; Baker, R. T., Ammonia - borane: the
hydrogen source par excellence? Dalton Transactions 2007, (25), 2613-2626; (d) Peng,
B.; Chen, J., Ammonia borane as an efficient and lightweight hydrogen storage
medium. Energy Environ. Sci. 2008, 1 (4), 479-483; (e) Hamilton, C. W.; Baker, R. T.;
Staubitz, A.; Manners, I., B-N compounds for chemical hydrogen storage. Chem. Soc.
Rev. 2009, 38 (1), 279-293.
[28] Bowden, M.; Autrey, T.; Brown, I.; Ryan, M., The thermal decomposition of ammonia
borane: A potential hydrogen storage material. Current Applied Physics 2008, 8 (3-4),
498-500.
[29] Hu, M. G.; Geanangel, R. A.; Wendlandt, W. W., Thermal-Decomposition of
Ammonia-Borane. Thermochim. Acta 1978, 23 (2), 249-255.
[30] (a) Bowden, M.; Autrey, T.; Brown, I.; Ryan, M., The thermal decomposition of
ammonia borane: A potential hydrogen storage material. Curr. Appl. Phys. 2008, 8 (3-
4), 498-500; (b) He, T.; Xiong, Z. T.; Wu, G. T.; Chu, H. L.; Wu, C. Z.; Zhang, T.;
126 Yaoqi Li, Ping Son, Yan Li et al.

Chen, P., Nanosized Co- and Ni-Catalyzed Ammonia Borane for Hydrogen Storage.
Chem. Mater. 2009, 21 (11), 2315-2318.
[31] (a) Denney, M. C.; Pons, V.; Hebden, T. J.; Heinekey, D. M.; Goldberg, K. I., Efficient
catalysis of ammonia borane dehydrogenation. J. Am. Chem. Soc. 2006, 128 (37),
12048-12049; (b) Keaton, R. J.; Blacquiere, J. M.; Baker, R. T., Base metal catalyzed
dehydrogenation of ammonia-borane for chemical hydrogen storage. J. Am. Chem. Soc.
2007, 129 (7), 1844-1845.
[32] (a) Kalidindi, S. B.; Sanyal, U.; Jagirdar, B. R., Nanostructured Cu and Cu@Cu2O core
shell catalysts for hydrogen generation from ammonia-borane. Phys. Chem. Chem.
Phys. 2008, 10 (38), 5870-5874; (b) Li, Y. Q.; Xie, L.; Li, Y.; Zheng, J.; Li, X. G.,
Metal-Organic-Framework-Based Catalyst for Highly Efficient H-2 Generation from
Aqueous NH3BH3 Solution. Chem.Eur. J. 2009, 15 (36), 8951-8954.
[33] (a) Stowe, A. C.; Shaw, W. J.; Linehan, J. C.; Schmid, B.; Autrey, T., In situ solid state
B-11 MAS-NMR studies of the thermal decomposition of ammonia borane:
mechanistic studies of the hydrogen release pathways from a solid state hydrogen
storage material. Phys. Chem. Chem. Phys. 2007, 9 (15), 1831-1836; (b) Sit, V.;
Geanangel, R. A.; Wendlandt, W. W., The Thermal-Dissociation of Nh3bh3.
Thermochim. Acta 1987, 113, 379-382.
[34] Baker, R. T.; Staubitz, A.; Manners, I., B-N compounds for chemical hydrogen storage.
Chem. Soc. Rev. 2009, 38 (1), 279-293.
[35] Gutowska, A.; Li, L. Y.; Shin, Y. S.; Wang, C. M. M.; Li, X. H. S.; Linehan, J. C.;
Smith, R. S.; Kay, B. D.; Schmid, B.; Shaw, W.; Gutowski, M.; Autrey, T.,
Nanoscaffold mediates hydrogen release and the reactivity of ammonia borane. Angew.
Chem., Int. Ed. 2005, 44 (23), 3578-3582.
[36] Bluhm, M. E.; Bradley, M. G.; Butterick, R.; Kusari, U.; Sneddon, L. G., Amineborane-
based chemical hydrogen storage: Enhanced ammonia borane dehydrogenation in ionic
liquids. J. Am. Chem. Soc. 2006, 128 (24), 7748-7749.
[37] (a) Jaska, C. A.; Temple, K.; Lough, A. J.; Manners, I., Transition metal-catalyzed
formation of boron-nitrogen bonds: Catalytic dehydrocoupling of amine-borane adducts
to form aminoboranes and borazines. J. Am. Chem. Soc. 2003, 125 (31), 9424-9434; (b)
Jaska, C. A.; Temple, K.; Lough, A. J.; Manners, I., Rhodium-catalyzed formation of
boron-nitrogen bonds: a mild route to cyclic aminoboranes and borazines. Chem.
Commun. 2001, (11), 962-963.
[38] Yan, J. M.; Zhang, X. B.; Han, S.; Shioyama, H.; Xu, Q., Iron-nanoparticle-catalyzed
hydrolytic dehydrogenation of ammonia borane for chemical hydrogen storage. Angew.
Chem., Int. Ed. 2008, 47 (12), 2287-2289.
[39] Xu, Q.; Chandra, M., Catalytic activities of non-noble metals for hydrogen generation
from aqueous ammonia-borane at room temperature. Journal of Power Sources 2006,
163 (1), 364-370.
[40] Cheng, F. Y.; Ma, H.; Li, Y. M.; Chen, J., Ni1-xPtx (x=0-0.12) hollow spheres as
catalysts for hydrogen generation from ammonia borane. Inorganic Chemistry 2007, 46
(3), 788-794.
[41] Li, Y. Q.; Song, P.; Zheng, J.; Li, X. G., Promoted H 2 Generation from NH 3 BH 3
Thermal Dehydrogenation Catalyzed by MetalOrganic Framework Based Catalysts
Chem.Eur. J. 2010, 16 (35), 10887-10892.
The Applications of Metal Organic Frameworks 127

[42] Wolf, G.; Baumann, J.; Baitalow, F.; Hoffmann, F. P., Calorimetric process monitoring
of thermal decomposition of B-N-H compounds. Thermochim. Acta 2000, 343 (1-2),
19-25.
[43] Heldebrant, D. J.; Karkamkar, A.; Hess, N. J.; Bowden, M.; Rassat, S.; Zheng, F.;
Rappe, K.; Autrey, T., The Effects of Chemical Additives on the Induction Phase in
Solid-State Thermal Decomposition of Ammonia Borane. Chem. Mater. 2008, 20 (16),
5332-5336.
In: Coordination Polymers and Metal Organic Frameworks ISBN: 978-1-61470-899-5
Editors: O. L. Ortiz and L. D. Ramirez, pp. 129-168 2012 Nova Science Publishers, Inc.

Chapter 4

MOF-BASED MIXED-MATRIX-MEMBRANES
FOR INDUSTRIAL APPLICATIONS

Hoang Vinh-Thang and Serge Kaliaguine


Department of Chemical Engineering, Laval University,
Qubec (Qubec) Canada.

ABSTRACT
Modern membrane gas separation technology has enjoyed a rapid development of its
commercial applications in the chemical, petrochemical, semiconductor, food,
pharmaceutical, biotechnology and environmental industries, due to their low energy
consumption, compelling low cost and ease of large-scale operation.
Mixed-matrix membranes (MMMs) combine some of the assets of polymer
membranes with the increased separation selectivities associated with the presence of a
load of inorganic particles. In this chapter the potential advantages of metal-organic
frameworks (MOFs) as the discrete phase in MMMs are reviewed.

1. INTRODUCTION
Mixed matrix membranes (MMMs) are mostly useful in gas separation processes. Today
some industrially relevant membrane separation processes are already implemented at
commercial scale. Some others are at more or less advanced development stage whereas some
should desirably be developed.
In the first category, the Monsanto hydrogen separation based on the PRISM membrane
was launched in the early 80s [1]. Many such plants have been installed since that time, with
one of the most significant applications being the recovery of hydrogen from ammonia plants
purge-gas streams. Also in the 80s Dow produced the GENERON first commercial
membrane for separation of nitrogen from air. Further developments of this technology by
Dow, Ube and Dupont/Air liquide resulted in a rapid expansion of the process. Over 10,000

Corresponding author. Phone: 1-418-656-2708; Fax: 1-418-656-3810. E-mail: Serge.Kaliaguine@gch.ulaval.ca.


130 Hoang Vinh-Thang and Serge Kaliaguine

such systems have already been installed worldwide. The production of oxygen enriched air,
however still requires better membranes to become commercial [2].
Dried cellulose acetate membranes were first produced and utilized in commercial plants
by Cynara, Separex and Grace Membrane Systems for the removal of carbon dioxide from
methane in the purification of natural gas. Further developments of this technology involved
polyimide hollow-fibre membranes by Air Liquide [3]. Significant improvements of this
technology would be required for its large scale application to purification of biogas. This
resource is still underutilized mostly due to the cost of current purification technologies.
Some research activities are conducted worldwide toward that objective [4].
Commercial membrane separation plants for vapour separation from air have been
installed by MTR, GKSS, NIHO Denko since the early 90s. Vaperma has developed recently
the SIFTEK membrane for gas phase separation of water vapour from ethanol [5].
Several petrochemical and refinery processes generate streams of nitrogen or hydrogen
gas with sizeable contents of small hydrocarbons. The use of membrane separation units for
the recovery of these hydrocarbons is receiving special attention currently. These industries
would also certainly benefit from membrane processes for the separation of the numerous
organic mixtures they generate. Large scale processes would for example be needed in
vapour-vapour separations of ethylene from ethane, propylene from propane, n-butane from
isobutene [6].
Other applications of membrane gas separation are also actively seeked for in the area of
air pollution control for the recovery of VOC from contaminated atmospheres.

2. MIXED-MATRIX-MEMBRANES
Membrane-based separation processes are widely implemented commercially in
petrochemical, chemical, food, pharmaceutical, semiconductor, biotechnological, and
environmental industries. Compared to the earlier techniques such as distillation, absorption
and adsorption their advantages lie in both low capital cost and high energy efficiency.
Polymer membranes are the most common commercial membranes for gas separations based
on their low cost, high processability, good mechanical stability and excellent transport
properties [7-21]. Both rubbery and glassy polymers including poly(dimethyl siloxane),
silicone rubber, nitril-butadiene, ethylene-propylene and polychloroprene rubbers, cellulose
acetate, polysulfone, polyethersulfone,polyamides, polyimides, polyetherimides,
polypropylene, poly(vinyl chloride), poly(vinyl fluoride), have been used for the preparation
of polymer membranes [20-21]. However, a poor contaminant resistance, low chemical and
thermal stability, and a limit in the trade-off between permeability and selectivity (polymer
upper bound limit) are among some of their disadvantages. Figure 1 shows the polymer upper
limits for O2/N2 and CO2/CH4 separations demonstrated by Robeson in 2008 [22].
On the other hand, inorganic membranes such as zeolite and carbon molecular sieve
(CMS) membranes are overcoming some drawbacks of the polymer membranes. The unique
properties of microporous molecular sieves make inorganic membranes very attractive for gas
separations by taking their excellent thermal and chemical stability, good erosion resistance
and high gas flux and selectivity. However, certain aspects still require further attention such
MOF-Based Mixed-Matrix-Membranesfor Industrial Applications 131

as mechanical resistance, reproducibility, long-term stability, up-scaling and, of course,


fabrication cost [2,23-27].

Figure 1. Upper bound correlation for O2/N2(left) and CO2/CH4(right) separations [22].

Mixed matrix membranes (MMMs) are heterogeneous membranes consisting of


inorganic fillers such as zeolites or carbon molecular sieves, dispersed in a continuous
polymer matrix. They can exceed the polymeric upper-bound trade-off curves. MMMs
combine the potential advantages of both inorganic and polymer membranes such as the
superior permeability and selectivity of the former with the good processability and
mechanical property of the latter. Traditionally, both zeolitic and non-zeolitic inorganic
materials have been used as fillers for synthesizing MMMs. Non-zeolitic inorganic fillers are
non-porous or porous materials including non-porous and porous silica nanoparticles,
alumina, non- and functionalized activated carbon, carbon molecular sieves,
TiO2nanoparticles, layered materials, and mesoporous molecular sieves [28-35]. On the other
hand, zeolitic inorganic particles are conventional zeolites, AlPO- and SAPO- molecular
sieves. Mesoporous materials were employed as additive fillers in order to enhance the filler-
polymer interaction through polymer chain penetration of the mesopores, but no selectivity
enhancement was obtained since gas diffusion in mesopores is non-selective. To improve the
selectivity, some approaches were developed such as to incorporate micropores into the
mesoporous materials [36-39] or to create mesopores in pure zeolites [40-43], and to modify
the interface composition by functionalizing with organic groups, i.e., periodic mesoporous
organosilicas (PMOs) [44]. However, there are several drawbacks to fabricate MMMs using
inorganic materials, such as the formation of non-selective voids at the inorganic-polymer
interface, the partial blockage of the pore of inorganic fillers by polymer chains, and the
limited number of possible structure and composition [45-46].
Moreover, modeling of the permeation properties of MMMs has received much attention
since representing gas transport through MMMs is a complex mathematical problem. Up to
now, there are several theoretical permeation models available in literature to predict the gas
separation performances of MMMs as functions of the permeabilities of the dispersed fillers
and polymer matrix [2,20,28-30,33]. A comparative model for permeation through
heterogeneous MMMs developed by Petropolous [47], and later by Bouma et al. [48] is a
generalization of a particularly useful equation derived by Maxwell in 1873 for a system of
132 Hoang Vinh-Thang and Serge Kaliaguine

conducting spheres [49]. This model predicts reasonably well the permeabilities in
composites at low filler loadings. However, the effect of particle size distribution, particle
shape, aggregation of particles, and blocked surface are not considered. The Bruggeman
model is a Maxwell models improvement, which covers a broader range of the volume
fraction of the fillers, as compared to the maximum loading of 20 wt.% of the Maxwell
model, and still has above-mentioned Maxwell models limitation and cannot predict the
permeability at the maximum volume fraction [50]. The Lewis-Nielsen [51-52] and Pal [53-
54] models can be used to resolve the later disadvantage of the Bruggeman model and to take
into account the effects of morphology of the dispersed phase such as particle size
distribution, particle shape, and aggregation of particles. However, both models are divergent
when the permeability ratio marches forward the infinity value. As a development, the
modified Maxwell [29,55] and Felske [56] models introduce the influence of interfacial layer
for permeability prediction, but have the same limitations as that of the Maxwell model. Also,
there are other models available in the literature such as those of Bottcher [57], Higuchi and
Higuchi [58], Cheng and Vachon [59], and Agari and Uno [60].
Recently, Matsuura and co-researchers developed a new theoretical gas permeability
model for MMM systems [61-62]. To develop their model, the authors
classifiedthemorphology of MMMs in to four groups: (i) ideal contact betweenthe polymer
matrixandthe dispersed phase, knownas ideal MMM; (ii) rigidied
polymerlayeraroundthellerparticle,knownasrigidied MMM, (iii)
sieveincagemorphologyarisesfromtheweakinteractionbetween polymermatrixand the particle,
knownas voidin MMM; and (iv) blocked pore surfacemorphology, knownas
blockageinMMM. Then, due to the ideal MMM morphology being hardly achievable, this
new model focuses on considering the presence of interphase thickness around the dispersed
particle. This model can cover the volume fraction parameter under effects of rigidied
polymerlayer, non-selective void and pore blockage.As verified using seven cases through the
published experimental data, a high degree of agreement between model predictions and
experimental results is reported.
However, there are many other additional parameters affecting the phase transport
processes through MMMs, i.e., the particle pore size and its distribution, dispersed particle
and polymer properties and its interaction, and gas feed composition. In addition, the
operating conditions such as temperature and pressure have strong effects on the transport
properties of gases over the dispersed phase and the polymer matrix. It is clear from the above
discussion that if a predictive model can describe the true trends of gas transport in MMMs is
still a matter of debate.

3. MOF-BASED MEMBRANES
In the last decade, hybrid metal-organic materials have appeared as a new class of
crystalline and porous materials formed by self-assembly of complex subunits comprising
transition metal centers connected by various poly-functional organic ligands to form one-,
two-, and three-dimensional structures. These hybrids are usually labeled as Metal-Organic
Frameworks (MOFs). Since they first appeared in the literature in 1995 [63], several thousand
MOF materials have been synthesized with some derived acronyms such as Porous
MOF-Based Mixed-Matrix-Membranesfor Industrial Applications 133

Coordination Polymers (PCPs), Coordination Polymers (CPs), Porous Coordination Networks


(PCNs), Polymer Organic Frameworks (POFs), Innite Coordination polymer Particles
(ICPs), Metal Organic Polyhedra (MOPs), Metal-Organic Macrocycles (MOMs), Metal-
Organic Coordination Networks (MOCNs), Porous Metal-Organic Frameworks (PMOFs),
Microporous Metal-Organic Frameworks (MMOFs), Isoreticular Metal-Organic Frameworks
(IRMOFs), Materials Institute Lavoisier (MILs), Covalent Organic Frameworks (COFs),
Covalent Triazine-based Frameworks (CTFs), Zeolitic Imidazolate Frameworks (ZIFs),
Zeolitic Tetrazolate Frameworks (ZTFs),and other names [64-89].
These materials have interesting properties such as framework regularity, large surface
areas, high porosities, low densities, and enormous flexibility in pore size, shape and
structure. Compared to other porous materials, which associate with only limited metals,
MOFs accept almost all the cations up to tetravalent. A huge choice of organic linkers
containing O or N donors such as carboxylates, phosphonates, sulfonates, cyanides, pyridine,
imidazoles, etc. gives rise to an opportunities for functionalization and grafting. Moreover,
the carbon subnetwork can be itself functionalized with halogeno- or aminogroups. MOFs
may be non-porous or porous materials. Non-porous MOFs are those in which the organic
ligands are filling their cavities. In contrast, the blocked channels of porous MOFs can be
opened by means of thermal or chemical treatments. Although most of porous MOFs are
microporous solids, some MOFs have intrinsic mesoporous channels or mesostructured
structures [90-93]. MOFs have the mostly apparent disadvantage that they are stable up to
temperatures typically above 200 oC, which eliminates them in any high
temperatureapplications.Fortunately, a number of ZIFs and COFs exhibit exceptional thermal,
hydrothermal and chemical stability up to 400 oC. Moreover, some MOFs are weakly stable
in water environment, which causes some limitations for their use in membrane synthesis
[65,68].
While current MOFs studies are mostly dealing with the synthesis and characterization
[64-89,94-100], an increasing number of MOFs are now being explored for their promising
potential for gas adsorption and separation [67,69-70,72-73,75-79,83-85,87-89,101-119],
heterogeneous catalysis [67,76-78,84-85,87-89,112,120-131], semiconductor [132],
luminescence [67,70,76-78,83-85,88-89,133-134], magnetism [70,74-78,84,89,135-136], thin
films [88,137-138], biological and medical applications [76-78,84,89,139-140], and more.
All these properties make MOFs promising candidates for their use in the construction of
mixed-matrix membranes with superior performances. The interface morphology between
MOFs and polymer matrix is more easily controlled due to their organic linkers having better
affinity with polymer chains, and their surface properties can be modified by
functionalization. This review focuses on current investigations of MOF-based MMMs on
their synthesis, characterization and industrial applications. Before this main topic is
presented, the section below will provide a short description of the situation for solely MOF-
based membranes.
To date, MOF-based membranes are generally fabricated as thin films using a number of
synthesis techniques including in situ growth after support surface modification (metal-
organic chemical vapor deposition, MOCVD) with either organosilane molecules [142-
143,147,150-151] or organic ligands [156-157,159-160], solvothermal synthesis [144-
145,154,159,161,168,170,173,176], twin copper source [149], spin coating [155,177], layer-
by-layer growth [148,164-165], colloidal deposition [162,166-169], microwave-induced
thermal deposition [141,145,168,171-172,175], electrochemical synthesis [146], gel-layer
134 Hoang Vinh-Thang and Serge Kaliaguine

approach [153], seed layer deposition [158], reactive seeding [163], and solvent evaporation
[152,174]. For this kind of membranes, many types of MOF particles have been used for their
syntheses such as MOF-5 [141-145], Cu3(BTC)2 or HKUST-1 [146-158], Zn(bdc)2(dabco)
[150], Fe-MIL-88B [153], Cu(hfipbb)(H2hfipbb)0.5[159], Mn(COO)2 [160], SIM-1 [161],
MIL-89 [162], MIL-53 [163], SURMOFs [164-165], ZIF-7 [166-170], ZIF-8 [168,171-175],
ZIF-22 [168], ZIF-69 [176] and Ln(BTC) (Ln=Dy3+, Eu3+ or Tb3+) [177].

Table 1. MOF-based thin film membranes

MOFs Supports Synthesis methods Major applications Ref.


MOF-5 -Alumina Microwave-induced thermal - [141]
deposition
-Alumina MOCVD with organosilane - [142]
Silica
Au-SAMs MOCVD with organosilane - [143]
-Alumina Solvothermal H2, CO2, CH4, O2 [144]
-Alumina Microwave-induced thermal H2, CO2, CH4, O2 [145]
deposition
Solvothermal
Cu3(BTC)2 Cu Electrochemical QCM-sensors [146]
Au-SAMs MOCVD with organosilane - [147]
Au-SAMs Layer-by-layer - [148]
Cu Twin copper source H2/CH4, H2/CO2, [149]
H2/N2
-Alumina MOCVD with organosilane - [150]
Silica
Au-SAMs MOCVD with organosilane - [151]
HKUST-1 Silicon Solvent evaporation - [152]
Glass Gel-layer - [153]
-Alumina Solvothermal H2/CH4, H2/CO2, [154]
H2/N2
-Alumina Spin coating Water [155]
Au-SAMs MOCVD with organic ligands Sensors [156]
Au-SAMs MOCVD with organic ligands Pyridine diffusion [157]
-Alumina Seed layer deposition - [158]
Cu(hfipbb)(H2hfipbb)0.5 -Alumina Solvothermal H2/N2 [159]
MOCVD with organic ligands
Mn(COO)2 -Alumina MOCVD with organic ligands - [160]
Porous graphite
SIM-1 -Alumina Solvothermal CO2 [161]
MIL-89 Silicon Colloidal deposition - [162]
Fe-MIL-88B Glass Gel-layer - [153]
MIL-53 -Alumina Reactive seeding N2, H2, CO2, CH4 [163]
Zn(bdc)2(dabco) -Alumina MOCVD with organosilane - [150]
Silica
SURMOFs Si wafer Layer-by-layer - [164]
Au-SAMs Layer-by-layer N2, H2, O2, CO2, [165]
CH4
ZIF-7 -Alumina Colloidal deposition - [166]
-Alumina Colloidal deposition H2 [167]
-Alumina Colloidal deposition H2/CH4, H2/CO2, [168]
H2/N2, H2/O2
-Alumina Colloidal deposition H2/CH4, H2/CO2, [169]
H2/N2
-Alumina Solvothermal N2, H2, O2, CO2, [170]
MOF-Based Mixed-Matrix-Membranesfor Industrial Applications 135

MOFs Supports Synthesis methods Major applications Ref.


CH4
ZIF-8 Titania Microwave-induced thermal CO2/CH4 [171]
deposition
-Alumina Microwave-induced thermal H2/CH4, H2/CO2, [168]
deposition H2/N2, H2/O2
Titania Microwave-induced thermal N2, H2, O2, CO2, [172]
deposition CH4
-Alumina Solvothermal CO2/CH4 [173]
Glass Solvent evaporation Sensors [174]
Silicon
Titania Microwave-induced thermal Ethene/Ethane [175]
deposition
ZIF-22 TiO2 Solvothermal H2/CH4, H2/CO2, [168]
H2/N2, H2/O2
ZIF-69 -Alumina Solvothermal CO2/CO [176]
Ln(BTC) Glass Spin coating Luminescent [177]
Ln = Dy3+, Eu3+, or Tb3+

These dense thin films of MOFs are deposited on various supports such as -alumina
[141-142,144-145,150,154-155,158-161,163,165-170,173,176], silica [142,150], Au-SAMs
(Au self-assembled monolayers) [143,147-148,151,156-157,165], pure Cu [146,149], silicon
[152,162,174], glass [153,174,177], porous graphite [160],Si wafer [164], and titania
[168,171-172,175]. General details in the synthesis and performance of these MOF
membranes are summarized in Table 1. Moreover, there are several other modelling studies
on the MOF-based membranes using simulation method to predict the gas permeability and
selectivity [175,178-182].
Gas adsorption and separation performances are most important potentials of intrinsic
MOFs as well as their membranes for industrial applications. Although MOF membranes
have demonstrated potential applications in di erent elds like luminescence [177], and
QCM-based sensors [146,156,174],a number of investigations for gas separations have been
reported [144-145,149,154,159,161,163,166,168-173,176]. For example, Cu3(BTC)2 on a
400-mesh copper net thin films provided an excellent performance for H2 and separation
factors of 6-7for H2/N2 and H2/CH4 mixtures, which exceed those expected from Knudsen
diffusion [149]. Another example is a Cu(hfipbb)(H2hfipbb)0.5 membrane that shows very
high selectivity for H2/N2 separation [159]. ZIF-7/alumina [167] and ZIF-8/titania [172]
membranes exhibit good separation factors of 5.9 and 11.2for H2/CH4, respectively, while the
former membrane gives a separation factor of 7.7 for H2/N2 [167].
The descriptions above demonstrate that MOF-based membranes are promising materials
for gas separation and storage, but are still very much in their infancy. All drawbacks of the
zeolites-based membranes are unlikely dealt with MOF-based membranes, except for the
unlimitation in creating new types of MOFs.
The first incorporation of MOFs into a polymer matrix for the preparation of MMMs was
reported by Yehia et al. [183].Three dimensional copper(II) biphenyl dicarboxylate-
triethylenediamine MOF was used to develop MMMs with poly(3-acetoxyethylthiophene)
and Matrimid polymers. Compared to the pure polymer, these MMMsexhibit enhanced
methane permeability and selectivity. At loadings of 20 and 30 wt.%, an increase in methane
permeability from 0.08 to 0.12 and 0.22 Barrers was observed.
Table 2. MOF-based mixed-matrix membrane performances

MMMs Synthesis procedure Major Operating Measurement Example performance Ref.


MOF Polymer (solvent) application condition method Permeability Selectivity
(loading, wt.%) (Barrer)
Zn-IRMOF-1 Matrimid 5218 Solvent evaporation 50 oC, Constant- =38.8 =29.2 [184,
(up to 20) (dioxolane, NMP) 100 psig volume =1.33 =86.4 185]
=114.9
Zn-IRMOF-1 Ultem 1000 Solvent evaporation 50 oC, Constant- =2.97 =26.3 [184,
(10 and 20) 100 psig volume =0.11 =149.3 185]
=16.9
Zn-MOF-5 Matrimid 5218 Solvent evaporation 35 oC, Constant- =20.2 =44.7 [186]
(1030) (chloroform) 2 atm volume =0.45 =0.87
Time-lag =0.52
CuTPA Poly(vinyl acetate) Solvent evaporation, 35 oC, Constant- =3.26 =40.4 [187]
(15) (toluene) 1.35 psig volume =0.08
Cu(hfipbb)- Matrimid 5218 - 35 oC, Maxwell and =920 =95 [188]
(H2hfipbb)2 2 atm Bruggeman
(up to 30) models
Zn(bdc)(ted)0.5 Polytrimethylsilypropyne, - 35 oC, Maxwell - [189]
(up to 30) Hyflon AD60X, 2 atm model =187426.8
Teflon AF-2400,
Sulfonated polyimide, =52118.8
6FDA-mMPD,
6FDA-DDBT
Cu3(BTC)2 Polydimethylsiloxane - - - =20003000 =33.6 [190]
(up to 40) Polysulfone =68 =721
Cu3(BTC)2 Matrimid 5218 Solvent evaporation, 50 oC, Constant- =22.1 =29.8 [185]
(up to 30) (dioxolane, NMP) 100 psig volume =0.74 =90.3
=66.9
Cu3(BTC)2 Matrimid/Polydimethyl- Phase inversion, Rose Bengal RT, Constant- P=0.50.9 - [191]
(up to 20) siloxane (THF, NMP, from 1 atm volume (l/m2.h.bar)
hexane) isopropanol
Cu3(BTC)2 Matrimid/Polysulfone Phase inversion, 35 oC, Constant- =1018 =19.528 [192]
(up to 30) (NMP, dioxolane) 10 bar volume =719 =1327
MMMs Synthesis procedure Major Operating Measurement Example performance Ref.
MOF Polymer (solvent) application condition method Permeability Selectivity
(loading, wt.%) (Barrer)
Cu3(BTC)2 Poly(amic acid) Solvent evaporation 0.5-5atm Constant- =1266 =240 [193]
(up to 20) (PMDA, DMAc and volume =163
ODA) =28
Dry/wet spinning =42
Mn(HCOO)2 Polysulfone - - - =1010.5 =1426 [190]
(up to 40)
Cu-BPY-HFS Matrimid 5218 Solvent evaporation 35 oC, Constant- =7.8115.06 =25.5531.93 [194]
(up to 30) (chloroform) 2 bar volume =0.240.59 =45.3869.15
=16.7526.74
MOP-18 Matrimid 5218 Solvent evaporation 35 oC, Constant- =9.415.6 =16.4723.19 [195,
(up to 80) (chloroform) 1000 torr volume =0.410.95 =23.5244.55 196]
=17.822.3
ZIF-8 Matrimid/Polydimethyl- Phase inversion, Rose Bengal RT, Constant- P=0.50.8 - [191]
(up to 20) siloxane (THF, NMP, from 1 atm volume (l/m2.h.bar)
hexane) isopropanol

ZIF-8 Matrimid 5218 Solvent evaporation 35 oC, Constant- up to 24.55 up to 124 [197,
(up to 80) (chloroform) 2000 torr volume up to 0.89 up to 427 198]
Time-lag up to 71.22 up to 50
up to 0.52
ZIF-8 Poly(1,4-phenylene ether- Solvent evaporation 5 oC Constant- =626 - [199]
(up to 30) ether-sulfone) (chloroform) 010 bar volume
Time-lag
PFG NMR
V-MIL-47 Matrimid/Polydimethyl- Phase inversion, Rose Bengal RT, Constant- P=0.50.7 - [191]
(up to 20) siloxane (THF, NMP, from 1 atm volume (l/m2.h.bar)
hexane) isopropanol
Al-MIL-53 Matrimid/Polydimethyl- Phase inversion, Rose Bengal RT, Constant- P=0.50.7 - [191]
(up to 20) siloxane (THF, NMP, from 1 atm volume (l/m2.h.bar)
hexane) isopropanol
ZIF-90 Ultem 1000, Solvent evaporation 25 oC, Constant- =590720 =3437 [200]
(15) Matrimid 5218, (dichloromethane) 2 atm volume
6FDA-DAM polyimide
138 Hoang Vinh-Thang and Serge Kaliaguine

Up to now, several MOFs were used to prepare MMMs [184-200]. Although these MOF-
based MMMs exhibit a high degree interaction between MOFs and polymer matrix, but their
gas separation performance is still in undesired values. Detailed summarization on the
synthesis and performances of the MOF-based MMMs are listed in Table 2.

4. MOF-5 OR IRMOF-1BASED
MIXED-MATRIX-MEMBRANES
MOF-5 is the first type of MOF materials initially developed by Yaghis group in 1999,
synthesized by mixing inexpensive benzene-1,4-dicarboxylic acid as linker and zinc nitrate
hexahydrate in N,N-dimethylformamide under mild condition [65]. MOF-5 and IRMOF-1 are
isoreticular. They exhibit a face-centered 3D cubic crystal structure. MOF-5 has a pore size of
0.8nm that makes a surface area up to 3000 m2/g and possesses high thermal stability up to
400 oC (Figure 2). MOF-5 is most attractive among the MOF family due to its potential
applications in gas storage and separation [64,68].
The first report on the gas separation application of MOF-5 based MMMs was the patents
of Liu et al. in 2006 [184-185]. In these patents, Matrimid 5218 polyimide and Ultem 1000
polyetherimide were used as membrane matrix. Up to 20 wt.% IRMOF-1 particles, an
isoreticular MOF-5 structure, were dispersed as filler phase in the polymer matrix. Pure gas
separation experiments on these IRMOF-1 based MMMs showed a dramatical improvement
in gas permeability properties for both CO2 and H2 (i.e., 2-3 fold) compared to pure polymers
without significant loss in ideal gas selectivities (Tables 3-4).

Figure 2. (a) Construction of the MOF-5 framework. Top, the Zn4(O)O12C6 cluster. Left, as aball and
stick model (Zn, blue; O, green; C, grey). Middle, the same with the Zn4(O)tetrahedron indicated in
green. Right, the same but now with the ZnO4 tetrahedraindicated in blue. Bottom, one of the cavities in
the Zn4(O)(BDC)3, MOF-5, framework and (b) Representation of a {100} layer of the MOF-5
framework [65].
MOF-Based Mixed-Matrix-Membranesfor Industrial Applications 139

Table 3. Pure gas permeation results of IRMOF-1/Matrimid MMMs [185]

(Barrer) (Barrer) increase


Matrimid 5218 10.0 0.355 - 28.2
20 wt.% MMMs 38.8 1.33 288% 29.2
(Barrer) (Barrer) increase
Matrimid 5218 0.355 33.1 - 93.2
20 wt.% MMMs 1.33 114.9 247% 86.4

Table 4. Pure gas permeation results of IRMOF-1/Ultem MMMs [185]

(Barrer) (Barrer) increase


Ultem 1000 1.95 0.0644 - 30.3
10 wt.% MMMs 2.81 0.101 44% 27.8
20 wt.% MMMs 2.97 0.113 52% 26.3
(Barrer) (Barrer) increase
Ultem 1000 0.0644 11.2 - 174.4
20 wt.% MMMs 0.113 16.9 51% 149.3

Figure 3. SEM images of the surface (a, d, and g), cross-section at low magnification (b, e, and h), and
cross-section at high magnification (c, f, and i) of 10, 20, and 30 wt.% MOF-5/MatrimidMMMs,
respectively [186].

In late 2007, Perez et al. reported a study on the synthesis of MOF-5/Matrimid MMMs
for the separation of gas and binary mixtures [186]. Up to 30 wt.% MOF-5 nanocrystals with
high surface area (3000 m2/g) and high thermal stability (up to 400 oC) were incorporated into
140 Hoang Vinh-Thang and Serge Kaliaguine

Matrimid polymides in chloroform solution to form MMMs. The synthesized MOF-5


particles were prepared by modifying typical published procedures [65] by directly adding
pure organic amine such as triethylamine (TEA) under strong stirring in order to obtain MOF
crystals in nano-size scale [201]. The membrane cross-section SEM images provided an
evidence of the formation of circular cavities and polymer veins with increased deformation
of the polyimides (Figure 3). At higher MOFs loading, the agglomeration of MOF-5
nanocrystals in polymer matrixis more evident.
Similar to previous results in zeolites containing MMMs [27,32], gas permeability values
of all gases such as H2, N2, O2, CO2 and CH4 summarized in Table 5 increased proportionally
to the MOF-5 loading, however, the ideal selectivities remained unchanged. For example, 30
wt.% MOF-5/Matrimid MMMs showed a 55% increase in CO2 permeability, 100% increase
in CH4 diffusivity, but only a 6% higher in CO2/CH4 selectivity, therefore suggesting an
enhanced CH4 transport upon MOF-5 incorporation.

Table 5. Gas permeabilities and ideal selectivities of MOF-5/Matrimid MMMs [186]

wt.% MOF-5 (Barrer) (Barrer) (Barrer) (Barrer) (Barrer)


0 24.40.1 9.00.1 1.900.01 0.250.04 0.220.02
10 29.94.8 11.11.4 2.300.30 0.280.08 0.220.04
20 38.38.8 13.82.8 2.900.60 0.400.01 0.340.04
30 53.83.9 20.21.4 4.120.37 0.520.04 0.450.06
wt.% MOF-5
0 113.09.3 41.73.3 7.61.0 2.710.01 0.860.06
10 137.443.6 51.014.6 8.41.3 2.680.10 0.860.37
20 112.012.2 40.53.5 7.21.2 2.760.10 0.850.07
30 120.07.7 44.73.0 7.90.1 2.660.01 0.870.05

Table 6. Gas transport properties of pure PVAC and 15 wt.%


CuTPA/PVAc MMMs [187]

(Barrer) (Barrer) (Barrer) (Barrer) (Barrer)


PVAc 15.10.8 0.5140.034 0.07830.0064 0.06970.0034 2.440.32
MMMs 19.00.5 0.6240.026 0.09120.0032 0.08060.0035 3.260.23

PVAc 2121.75 6.570.143 1.090.0433 32.11.36 34.02.86


MMMs 23711.8 6.790.136 1.140.0191 35.41.68 40.42.45

PVAc 6.210-96.810-10 4.610-81.910-9 1.210-84.010-10 2.910-92.710-10


MMMs 5.110-91.510-9 2.010-81.010-8 4.110-92.910-9 5.710-101.510-
10

Recently, Adams et al. dispersed two-dimensionally coordinated copper containing


MOFs particles in poly(vinyl acetate) polymer (PVAc) to create 15 wt.% CuTPA/PVAc
MMMs [187]. CuTPA is a type of MOF-5 structure synthesized from copper nitrate
trihydrate and terephthalic acid (TPA), also called benzene-1,4-dicarboxylic acid (BDA or
BDC), in N,N-dimethylformamide (DMF). The MMMs were prepared via the polymer
solution method using toluene as solvent. SEM images of MMMs showed a good dispersion
MOF-Based Mixed-Matrix-Membranesfor Industrial Applications 141

of CuTPA particles in polymer matrix and the absence of void defects at filler-polymer
interfaces. Gas transport properties of 15 wt.% CuTPA/PVAc MMMs as compared with those
of pure PVAc polymer are summarizes in Table 6.
As summarized in Table 6, pure gas permeabilities and selectivies of 15 wt.%
CuTPA/PVAc MMMs are substantially enhanced over pure PVAc properties. A large
enhancement in both permeability and selectivity values is observed for CO2 gas, giving a
potential application of these MMMs in selective removal of carbon dioxide from natural gas.
On the other hand, the apparent diffusivities are reduced with the incorporation of Cu-TPA
particles. This suggests a substantial increase in solubility presumably associated with the
high adsorption capacity of this MOF for both CO2 and CH4.
Very recently, a molecular simulation study of gas mixture separations in MOF-5-type
based MMMs was reported by Keskin and Sholl [188]. In their paper, using experimental data
measured by Perez et al. for MOF-5 (or IRMOF-1)/Matrimid MMMs [186], an agreement
between prediction models and experimental data was reported using Maxwell and
Bruggeman models (Figure 4). IRMOF-1 also is a Zn-containing MOF-5 structure. Both
models slightly overestimated the gas permeability data compared to the experimentalresults.
The Bruggeman model predicted higher gas permeabilities than the Maxwell model,
especially at higher MOFs loadings. On the other hand, the ideal selectivities of gas pairs
calculated using the Maxwell model are slightly lower than those of experimental values.

Figure 4. (a) Comparison of pure gas permeabilities of CH 4, N2, CO2 and H2 for IRMOF-1/Matrimid
MMMs and (b) predictions of the Maxwell model for ideal selectivities as function of MOFs loading.
Closed and open symbols are the predictions of the Maxwell and Bruggeman models, respectively
[188].

5. CU(HFIPBB)(H2HFIPBB)0.5BASED MIXED-MATRIX-MEMBRANES
As a continuation of the computational study of CO2/CH4 separation over IRMOF-
1/Matrimid MMMs reported in the previous section, Keskin and Sholl described the predicted
performance of Cu(hpbb)(H2hpbb)0.5/Matrimid MMMs by Maxwell model [188].
However, no data of the membrane synthesis and characterization were showed.
Cu(hpbb)(H2hpbb)0.5 has a 3D interpenetrating framework containing unique and perfectly
142 Hoang Vinh-Thang and Serge Kaliaguine

ordered microporous 1D open channels (microtubes) built upon a Cu2(hfipbb)4(H2hfipbb)2


paddle-wheel building unit. H2hfipbb is 4,4-(hexafluoroisopropylidene) bis(benzoic acid). Its
medium size cages (5.1 5.1 ) are connected by small 3.5 3.2 windows based on
atom to atom distances (Figure 5) [202]. Based on the unique shape and size of these
channels, interesting separation properties of Cu(hpbb)(H2hpbb)0.5 crystals have been
reported to selectively adsorb smaller normal alkanes (C1-C4) while rejecting linear alkanes
larger than C4 and all branched alkanes [203]. It is also a promising sorbent for H2 [204] and
should provide high selectivity in CO2/CH4separation [180].
As shown in Figure 6, at higher loading of Cu(hpbb)(H2hpbb)0.5 in the MMMs, both
the ideal selectivity and permeability of CO2 increased relative to pure Matrimid membrane.
For example, increasing the Cu-MOF volume fraction from 0 to 30% increases the
permeability of CO2 from 9 to 20 Barrer, and ideal selectivity for CO2 reach a value of 95,
twice higher than that of pure polyimide (factor of 41). H2/CH4 ideal selectivity of MOF-
filled MMMs is four times higher than that of unfilled membranes. On the other hand,
H2/CO2 ideal selectivity was predicted a reduction when MOF particles were incorporated.

Figure 5. Structure of Cu(hpbb)(H2hpbb)0.5 in (010) direction (left) and (101) direction (right) [202].

Figure 6. Predictions of the Maxwell model (closed symbol) and Bruggeman model (open symbol) for
ideal selectivities of Cu(hfipbb)(H2hfipbb)0.5/Matrimid MMMs: (a) in the separation of CO2/CH4
mixture, and (b) as function of MOF volume fraction in the MMMs [188].
MOF-Based Mixed-Matrix-Membranesfor Industrial Applications 143

6. ZN(BDC)(TED)0.5BASED MIXED-MATRIX-MEMBRANES
In early 2010, as a continuation in his separation computational study, Keskin represented
predicted self-diffusivities of CH4/H2 mixtures in Zn(bdc)(ted)0.5/polymers MMMs using
Maxwellmodel [189]. Zn(bdc)(ted)0.5 has a tetragonal 3D porous structure containing
interlacing channels of two different sizes (7.5 7.5 , 4.8 3.2 ), space group P4/ncc
(Figure 7).Zn(bdc)(ted)0.5is usually synthesized by solvothermal reaction of a mixture
ofzinc(II) nitrate tetrahydrate, terephthalic acid (so-called BDC) and triethylenediamine
(TED) in DMF [205-207].

Figure 7. Zn(bdc)(ted)0.5 structure viewed along c axis. Zn (green), O (red), C (dark gray), H (light
gray), and N (blue) [207].

Figure 8. Maxwell model predictions for H2/CH4 selectivity and permeability of Zn(bdc)(ted)0.5 based
MMMs composed of different types ofpolymers. The circles represent selectivity/permeability
characteristics of pure polymers. The diamond shows H2/CH4selectivity and permeability of pure
Zn(bdc)(ted)0.5 membrane. The stars are the predictions of Maxwell model for MMMs with volume
fractions of Zn(bdc)(ted)0.5varying from 0.1 to 0.5. The lines represent prior and present upper bounds
established by Robeson [189].
144 Hoang Vinh-Thang and Serge Kaliaguine

In Keskins study, there is neither any information concerning the synthesis and
characterization of Zn(bdc)(ted)0.5 based MMMs nor any experimental data for the gas
separation properties. A series of polymers such as polytrimethylsilypropyne, Hyflon
AD60X, Teflon AF-2400, and sulfonated, 6FDA-mMPD and 6FDA-DDBT polyimides were
considered as the continuous phase in these predictions. As showed in Figure 8, the predicted
results of the Maxwell model suggested an increase in H2 permeability in all types of
polymers with Zn(bdc)(ted)0.5 loadings, similar to those previously reported for MOF-based
MMMs. Adding Zn(bdc)(ted)0.5 particles into polymers is not predicted to make a significant
improvement in H2/CH4 selectivity compared to pure polymers.

7. CU3(BTC)2OR HKUST-1BASED MIXED-MATRIX-MEMBRANES


[Cu3(BTC)2] or HKUST-1 (BTC = benzene-1,3,5-tricarboxylate) is a 3D porous MOF
with a zeolite-like structure. [Cu3(BTC)2] possesses a main channel of 0.9nm diameter
surrounded by tetrahedral pockets of 0.5nm diameter. The latter are connected with the main
channel by triangular windows of 0.35nm diameter (Figure 9). High BET surface areas up to
1500 m2/g have been reported, depending on synthesis procedures and conditions [208-210].
Among the various MOFs, Cu3(BTC)2 is one of the most studied for gas adsorption and
storage, such as H2, N2, CO2, and CH4 [211-214].
The first report on the synthesis and separation application of Cu3(BTC)2 based MMMs
was carried out in 2006 [190]. Car et al. synthesized two different types of MMMs including
Cu3(BTC)2 and Mn(HCOO)2 dispersed in polydimethylsiloxane (PDMS) and Cu3-
(BTC)2/polysulfone (PSf). The gas separation performance for single and mixed gases of H2,
N2, O2, CH4 and CO2 was measured using time lag apparatus with the MOFs loading up to
40 wt.% (Figure 10). Incorporation of Cu-MOF in both polymers showed high permeability
and selectivity for CO2 and H2, while Mn(HCOO)2 showed high sorption for H2. Higher
loading reduced the gas solubility, but increased the selectivity. The authors explained their
results by the presence of voids at the interface of MOF particles and polymers. However, no
SEM images of these MOF-based MMMs were presented.

Figure 9. The 3DstructureofCu3(BTC)2 [214].


MOF-Based Mixed-Matrix-Membranesfor Industrial Applications 145

Figure 10. Gas separation properties of Cu 3(BTC)2/polymers MMMs: (a) permeability and (b)
selectivity of Cu3(BTC)2/polydimethylsiloxane (PDMS), (c) permeability and (d) selectivity of Cu 3-
(BTC)2/polysulfone (PSf) MMMs [190].

Table 7. Pure gas permeation results of Cu3(BTC)2/Matrimid MMMs [185]

(Barrer) (Barrer) increase


Matrimid 5218 10.0 0.355 - 28.2
30 wt.% MMMs 22.1 0.741 121% 29.8
(Barrer) (Barrer) increase
Matrimid 5218 0.355 33.1 - 93.2
30 wt.% MMMs 0.741 66.9 102% 90.3

Gas separation performance of Cu3(BTC)2/Matrimid MMMs was also reported in the


patent of Liu et al. [185]. No membrane synthesis and characterization results were presented.
The CO2 permeability of 30 wt.% loading samples increased 121% without loss of CO2/CH4
selectivity compared to pure Matrimid membrane (Table 7). The same trend was obtained for
H2/CH4 separation (increase 102%).
In 2009, Basu et al. also reported the synthesis of Cu3(BTC)2/polydimethylsiloxane
(PDMS) MMMs and their application in the separation of Rose Bengal (RB) from
isopropanol (IPA) [191]. After the preparation of Cu3(BTC)2particles, in order to increase the
adhesion between polymers and fillers, these authors first modified the MOFs surface by
silylation with N-methyl-N-(trimethylsilyl) triuoroacetamide (MSTFA). The MSTFA
modied llers showed a better compatibility with the polymer and resulted in defect-free
membranes. The synthesis procedure of MMMs was slightly complicated including several
steps. First, a polyimide (PI) Matrimid 9725 porous support layer was prepared from a
146 Hoang Vinh-Thang and Serge Kaliaguine

polymer dope solution with the phase inversion technique, and then post-treated by
immersing in isopropanol (IPA) and then in a mixture of 4-methyl-2-pentanone (MIBK),
toluene and mineral oil. Second, a coating Cu3(BTC)2/PDMS solution with filler loading up
to 20 wt.% was prepared by adding the fillers in a 10 wt.% PDMS solution in hexane
containing RTV 615A and B crosslinkers. Finally, Cu3(BTC)2/PDMS solution was coated on
top of the polymer support at 60oC.
As shown in Figure 11a, the permeance of unmodified Cu3(BTC)2/PDMS MMMs
increased with the loading and is better than that of MSTFA functionalized one. However, the
RB retention of MSTFA functionalized MMMs reached values up to 97% (Figure 11b), and
is more than twice larger than that of unmodified samples.
Recently, asymmetric Cu3(BTC)2/Matrimid and Cu3(BTC)2/Matrimid-polysulphone
blends MMMs were preparedviaphaseinversion as reported by Basu et al. [192]. Up to 30
wt.% Cu-MOFs micro-particles were firstly dispersed in a mixture of N-methylpyrrolidinone
(NMP) and dioxolane. Then, the polymers were partly added to the mixture under mechanical
stirring. The overall morphology of Cu3(BTC)2 based MMMs with 30 wt.% loading showed
that the micro-particles (around 10 m) are embedded/wrapped within/by the polymer matrix
immediately below the thin selective layer (Figure 12). The improvement in thermal and
mechanical propertiesof the MOF-based MMMs with the Cu3(BTC)2 content was confirmed
by TGA, dynamic mechanical storage and Youngs modulus, and tensile strength and break at
elongation test results.

Figure 11. (a) Permeance of Cu3(BTC)2/PDMS MMMs loaded with different contents of unmodified
and MSTFA functionalized MOFsand (b) permeance (, l/m2.h.bar) and retention (, %) of MSTFA
functionalized Cu3(BTC)2/PDMS MMMs [191].

Figure 12. SEM cross-section images of (a) 30 wt% [Cu3(BTC)2] in polyimide with PDMS coated on
top (b) 30 wt% [Cu3(BTC)2] in polyimide at low magnification [192].
MOF-Based Mixed-Matrix-Membranesfor Industrial Applications 147

In order to evaluate gas separation performance of the Cu3(BTC)2/polymers MMMs, the


authors investigated several effective parameters such as CO2/CH4 and CO2/N2 composition,
CO2 feed content, and filler loading. CO2/CH4 selectivity increased with filler loading as a
function of CO2 concentration at 10 bar and 35 oC. In contrast, a decreasing selectivity with
increasing CO2 content in the feed gas was observed. The CO2 permeance of Cu3-
(BTC)2/polymers MMMs is better than that of un-loaded polyimide, and increased with Cu3-
(BTC)2 content. For example, an increase up to 80% in the CO2 permeance was reported for
the 30 wt.% Cu-MOF loading sample, as compared to the un-filled membranes. The addition
of polysulphone blends in the membrane synthesis also improved the gas permeability.
Very recently, Hu et al. prepared mixed-matrix membrane hollow fibers containing Cu3-
(BTC)2 and poly(amic acid) solution, and tested their separation and adsorption performance
for several gases such as H2, N2, CO2, O2 and CH4 [193]. In the membrane preparation, a
highly dispersed and viscous Cu3(BTC)2/poly(amic acid) solution was synthesized by mixing
up to 20 wt.% lab-synthesized Cu3(BTC)2 particles possessing a surface area of 1396 m2/g
with a solution of pyromellitic dianhydride (PMDA), dimethylacetamide (DMAc and 4,4-
oxydianiline (ODA). Then,Cu3(BTC)2/poly(amic acid) hollow fiber MMMs were prepared by
the dry/wet spinning method with the spinning temperature in the range of 30-45 oC, N2
extrusion pressure of 0.1-0.3 MPa and a detailed temperature-programmed procedure. As
illustrated in SEM pictures of Cu3(BTC)2/poly(amic acid) MMMs (Figure 13), the hollow
fibers were divided into three portions with different structures including an inner dense layer
(A), fingerlike voids (B) and spongelike structure (C) from the outside to the inner surface,
respectively.

Figure 13. SEM images of hollow fiber Cu3(BTC)2/poly(amic acid)MMMs: (a,b) crosssection, (c) outer
surface, (d) surface of fingerlike void, and (e) spongelikestructure [193].
148 Hoang Vinh-Thang and Serge Kaliaguine

The gas performance of 3 and 6 wt.% Cu3(BTC)2 based hollow fiber MMMs compared
with poly(amic acid) polyimide for H2, O2, N2, CO2 and CH4 are summarized in Table 8.
Although up to 20 wt.% of Cu3(BTC)2 powders were loaded into polyimide membranes, only
3 and 6 wt.% samples were tested. The MMMs with MOFs loading higher than 10 wt.% are
brittle to spin into hollow fibers. The H2 permeability increased with the filler loading, and
reached a value 1.5 times larger than that of pure polyimide. However, the CO2 permeability
of 6 wt.% Cu3(BTC)2 based MMMs is less than one-half that of pure polyimide. A similar
trend was observed for microporous fillers containing MMMs, the selectivities of H2/CO2,
H2/CH4, H2/N2, and H2/O2 increased with the MOFs loading and are twice as large compared
to those of pure polyimide.

Table 8. Permeation (and selectivity of hollow-fiber Cu3(BTC)2/poly(amic acid)


MMMs at 1 MPa and 298K [193]

3 wt/% Cu3- 6 wt/% Cu3-


Poly(amic acid)
(BTC)2/poly(amic acid) (BTC)2/poly(amic acid)
polyimide
MMMs MMMs
CO2 permeance (GPU*) 87.6 64.9 37.2
H2 permeance (GPU) 876 934 1270
H2/CO2 selectivity 10.0 18.0 27.8
H2/CH4 selectivity 128 138 240
H2/N2 selectivity 86.3 107 163
H2/O2 selectivity 16.3 24.3 41.7
* GPU: gas permeation unit; 1 GPU = 10-6cm3(STP)/cm2.s.cmHg.

8. CU-BPY-HFSBASED MIXED-MATRIX-MEMBRANES
Cu-BPY-HFS based mixed-matrix-membrane is another kind of MOF-based MMM
developed by Musselman and Balkuss group [194]. Free standing lms containing a
microporous metal-organic framework Cu-4,4-bipyridinehexauorosilicate (Cu-BPY-HFS)
combined with Matrimid5218 polyimidwere synthesized in chloroform solution. The 3D
structure of Cu-4,4-bipyridine-hexauorosilicate (Cu-BPY-HFS) involves a square network
of copper bipyridyl complexes that are pillared by SiF62 ions, which possesses a rectangular
channel with dimension of 0.8 x 0.2 nm along the a and b axes, and a large square pore with
diameter of 0.8 x 0.8 nm along the c axis (Figure 14).This MOF is known to be insensitive to
water and exhibits a high methane uptake of 145.6 cm3 (STP)/g at 25oC and high pressure of
3.6 (MPa) [139]. In this report, the synthesized Cu-BPY-HFS MOFs are block-like crystals
ranging from 200 to 300 nm in diameter. The BET surface areas of these MOFs are up to
2000 m2/g, much higher than the 1200 m2/g in the results reported by Noro et al. [215].
In the membrane synthesis, the Cu-BPY-HFS powders up to 40% (w/w) slurried into
chloroform were mixed, stirred and sonicated with Matrimid polymer to obtain a
homogeneous suspension. The SEM images of Cu-BPY-HFS/Matrimid MMMs show good
interfacial contact between the Cu-BPY-HFS particles and Matrimid polyimide and absence
of voids (Figure 15). Good dispersion was achieved when the loading is lower than 30%
(w/w). However, particle aggregation was observed at 40% (w/w) loading.
MOF-Based Mixed-Matrix-Membranesfor Industrial Applications 149

Figure 14. Channel structures of Cu-BPY-HFS [215].

Figure 15. Cross-section view of Cu-BPY-HFS/Matrimid MMMs with differentloadings: (A) 10 wt%,
(B) 20 wt%, (C) 30 wt% and (D) 40 wt% [194].

The permeability properties of Cu-BPY-HFS/Matrimid MMMs weretested for the single


gases H2, N2, O2, CH4, and CO2 and the gas mixtures CO2/CH4, H2/CO2 and CH4/N2 (Table
9). The results showed that the pure gas permeabilities of the MMMs generally increased with
loading. At 40% loading, a substantial increase in the permeabilities was observed, which was
attributed to the sieving effect of MOFs nanoparticles. However, no significant change in
ideal selectivities of O2/N2, H2/CO2, and CH4/N2 as compared with the 30 and 20% loadings
was observed. For the CH4/N2 mixture, the ideal selectivity increasedfrom 0.95 up to 1.7 with
loading, while the H2/CO2, H2/CH4, H2/N2, and CO2/CH4ideal selectivity values showed the
reverse trend.Owingto the high Cu-BPY-HFSs affinity towards CH4, and its huge
surfacearea, the solubility of CH4 in the MOF-based MMMs increased, which led to higher
selectivity towards CH4.
150 Hoang Vinh-Thang and Serge Kaliaguine

Table 9. Permeability properties of Cu-BPY-HFS/Matrimid MMMs[194]

Permeability (Barrers)
H2 N2 O2 CH4 CO2
Matrimid 17.50 0.22 1.46 0.21 7.29
10 wt.% 16.91 0.24 1.44 0.24 7.81
20 wt.% 16.75 0.31 1.77 0.36 9.88
20 wt.% 16.87 0.32 1.81 0.37 10.02
30 wt.% 20.34 0.31 1.98 0.38 10.36
40 wt.% 26.74 0.49 3.06 0.59 15.06
Ideal selectivity
H2/N2 O2/N2 CO2/CH4 H2/CH4 H2/CO2 CH4/N2
Matrimid 79.55 6.64 34.71 83.33 2.40 0.95
10 wt.% 71.04 6.04 31.93 69.15 2.17 1.03
20 wt.% 54.46 5.76 27.62 46.82 1.70 1.16
20 wt.% 52.72 5.66 27.08 45.59 1.68 1.16
30 wt.% 65.23 6.33 27.45 53.89 1.96 1.21
40 wt.% 54.78 6.27 25.55 45.38 1.78 1.21

9. MOP-18BASED MIXED-MATRIX-MEMBRANES
In their 2008s report, Musselman and Balkuss group used MOP-18 as new inorganic
filler particles [195]. MOP-18 is a metal-organic great rhombicuboctahedron functionalized
with extended alkyl chains, which was developed by Yaghis group [216]. It has 12 rigid
square-shaped Cu2(CO2)4 paddle-wheel building units and 24 5-dodecoxybenzene-1,3-
dicarboxylate links. MOP-18 is highly soluble in a variety of organic solvents, such as
chloroform, toluene, tetrahydrofuran, ethyl acetate, N,N-dimethylacetamide, and hot DMF,
but insoluble in dimethyl sulfoxide (DMSO), acetonitrile, methanol, 1-butanol (BuOH), and
isoamyl alcohol. The solubilization of MOP-18 made it possible to overcome the dispersion
of the aggregates and to improve the affinity between the additive and the polymer thus
minimizing phase separation.

Figure 16. (a) Cu2(CO2)4 paddle-wheel unit and dodecoxyl organic link (5-OC12H25-H2mBDC), (b) a
great rhombicuboctahedral framework of 13.8 diameter void (yellow sphere), and (c) X-ray single-
crystal structure of MOP-18. Cu, blue; O, red; C, black; all hydrogen atoms and terminal ligands on the
paddle-wheel units are omitted for clarity [216].
MOF-Based Mixed-Matrix-Membranesfor Industrial Applications 151

Figure 13. SEM images of the surface (a, d, g, and j), cross-section at low magnification (b, e, h, and k),
Figure 17. SEM images of the surface (a, d, g, and j), cross-section at low magnification (b, e, h, and k),
and cross-section at high magnification (c, f, i, and l) of 0, 23, 33, and 45 wt% MOP-18/Matrimid
and cross-section at high magnification (c, f, i, and l) of 0, 23, 33, and 45 wt% MOP-18/Matrimid
mixed-matrix membranes, respectively. The cross-sections show no aggregation of the MOP-18 crystals.
mixed-matrix membranes, respectively [195-196].

40

Figure 14. SEM image of a 80 wt% MOP-18/Matrimid, a) surface, b) cross-section low magnification,
Figure 18. SEM image of 80 wt.% MOP-18/Matrimid MMMs, a) surface, b) cross-section low
and c) cross-
magnification, andsection high magnification.
c) cross- section high magnification [195-196].
152 Hoang Vinh-Thang and Serge Kaliaguine

The synthesis of MOP-18 was accomplished by linking the paddle-wheel unit, a square,
with a suitably functionalized dicarboxylate link, which provides the required 120 angles for
producing a great rhombicuboctahedra. MOP-18 has a tetragonal unit cell containing four
rhombicuboctahedra with disordered guest molecules. Typically, MOP-18 was obtained by
mixing 5-dodecoxybenzene-1,3-dicarboxylic acid (5-OC12H25-H2mBDC) and Cu(CH3CO2)2
in DMF solvent [216].
In this report, a series of up to 80 wt% MOP-18/Matrimid MMMs was fabricated using
CHCl3 as solvent [195]. SEM images of the surfaces and cross-sections are shown in Figures
17-18. Interestingly, the surfaces of the MMMs showed no aggregation of the crystals as the
loading of MOP-18 increased. Thermogravimetric results indicated that these MMMs are
stable up to 300 C.
Single gas permeation experiments were performed with several gases such as H2, CO2,
O2, N2, CH4, C3H6, and C3H8 (Tables 10-11). In general, the permeability of the gases
increased proportionally as the content of MOP-18 increased.In the gas mixture tests, the
membranes are more selective for CH4 in the CO2/CH4 and CH4/N2 separations. The ideal
selectivities of CH4/N2(1.62), and C3H6/N2 (5.79) are 1.5 and 3 times higher than those of
pure polymer (CH4/N2=0.92, C3H6/N2=1.61),respectively. These results are above the
Knudsen values (CH4/N2 = 1.32, C3H6/N2 = 0.81), indicating the gas transport follows the
solubility/diffusivity model. Although the ideal selectivities of H2/CH4 (44.5), H2/N2 (53.65)
and CO2/CH4=0.6) are also larger than the Knudsen values (H2/CH4=2.82, H2/N2=3.74 and
CO2/CH4=0.60), respectively, they are lower than those of pure Matrimid membrane and
decreased with the loading. However, no comparison with other MOF-based MMMs was
reported.

Table 10. Permeability (Barrer) of permanent gases and small organic vapors
at 35 C and 2000 Torr in MOP-18/Matrimid MMMs[195-196]

wt.% H2 CO2 O2 N2 CH4 Propylene Propane


0 17.1 1.5 7.3 0.7 1.49 0.16 0.24 0.03 0.22 0.02 0.39 0.03 0.3 0.01
23 17.8 1.9 9.4 1.2 1.8 0.2 0.34 0.06 0.41 0.05 0.66 0.04 0.13 0.04
33 22.3 0.6 14.0 1.3 2.6 0.2 0.61 0.01 0.65 0.12 1.22 0.33 0.23 0.04
45 22.3 0.6 15.6 0.9 2.9 0.2 0.60 0.05 0.95 0.04 3.40 0.23 1.60 0.42

Table 11. Ideal selectivities of permanent gases and small organic vapors
at 35 C and 2000 Torr inMOP-18/MatrimidMMMs[195-196]

wt.% H2/CH4 H2/N2 CO2/CH4 H2/O2 O2/N2


0 76.64 0.47 70.46 2.15 32.86 0.07 11.47 0.14 6.14 0.11
23 44.55 0.55 53.65 4.35 23.19 0.31 9.60 0.11 5.59 0.39
33 35.01 5.35 36.66 0.67 21.87 2.00 8.53 0.34 4.30 0.25
45 23.52 0.38 38.10 2.31 16.47 0.20 7.74 0.31 4.92 0.10
Knudsen 2.82 3.74 0.60 4.00 0.93
wt.% H2/CO2 CH4/N2 C3H6/C3H8 C3H6/N2 C3H8/N2
0 2.33 0.02 0.92 0.02 1.29 0.11 1.61 0.07 1.25 0.16
23 1.90 0.05 1.22 0.08 5.45 1.20 1.99 0.25 0.37 0.04
33 1.60 0.10 1.06 0.18 5.24 0.48 2.00 0.57 0.38 0.07
45 1.43 0.04 1.62 0.07 2.18 0.44 5.79 0.13 2.71 0.48
Knudsen 4.70 1.32 1.02 0.81 0.79
MOF-Based Mixed-Matrix-Membranesfor Industrial Applications 153

10. ZIF-8BASED MIXED-MATRIX-MEMBRANES


ZIF-8 is one of the most studied prototypical ZIF compounds with large pores of 11.6
connected through small apertures of 3.4 , high surface areas of 1300-1600 m2/g, and an
amazing thermal stability up to near 400oC [74,173,217]. The first MOF-based MMMs using
ZIF-8 as fillers in polydimethylsiloxane (PDMS) membranes were investigated by Basu et al.
[191].In this work, the ZIF-8 crystals are commercial Basolite Z1200TM products as obtained
from Sigma-Aldrich. Due to a poor adhesion between polymers and fillers, ZIF-8 surface was
also modified by silylation with N-methyl-N-(trimethylsilyl) triuoroacetamide (MSTFA).
The membrane synthesis is similar to that of Cu3(BTC)2 based MMMs also reported in this
chapter (see section 7).
The separation performance of ZIF-8/PDMS MMMs was tested in solvent resistant
nanofiltration (SRNF), in particular in the separation of Rose Bengal (RB) from isopropanol,
and compared with other MOFs-based PDMS MMMs (Cu3(BTC)2, V-MIL-47 and Al-MIL-
53 as fillers). The results have shown that the retention values of the silylated ZIF-8/PDMS
MMMs (92-93%) are significantly higher than that of the unfilled PDMS membranes (87%),
but are still slightly lower than those of the other MOFs-based PDMS MMMs (up to 95-
98%). This was explained by the effect of reduced polymer swelling and size exclusion of
ZIF-8 nanoparticles, and by the small pores of ZIF-8 as compared toCu3(BTC)2, V-MIL-47
and Al-MIL-53, respectively (Figure 19).
Recently, ZIF-8/Matrimid MMMs with loadings up to 80% (w/w) were also synthesized
by Balkuss group [197-198]. Using nanometer-sized crystals of ZIF-8 achieved by adding a
base triethylamine (TEA) during synthesis [12], ZIF-8/Matrimid MMMs were obtained by
solution blending in chloroform. As seen in SEM results (Figure 20), as the ZIF-8 loading
increased, the interfacial contact between ZIF-8 particles and Matrimid polymer got worse
and the aggregation of ZIF-8 particles increased. This was resulting in a loss of mechanical
strength, and as consequence, a decrease in flexibility.
Then, the permeability and selectivity properties of these MMMs with ZIF-8 loadings up
to 60% (w/w) were tested with single gases of H2, CO2, O2, N2, CH4, and C3H8, and gas
mixtures of H2/O2, H2/CO2, H2/CH4, CO2/CH4, CO2/C3H8, and H2/C3H8 (Figure 21). The
permeabilities showed an increase as the ZIF-8 loading increased to 40% (w/w) and a
decrease at higher loadings of 50% and 60% (w/w). This was explained by the increasing
distance between polymer chains creating more polymer free volume when ZIF-8
nanoparticles were added to Matrimid polyimides.

Figure 19. Filtration performance of MSTFA unmodified and functionalized ZIF-8/PDMS MMMs:(a)
RB permeance (, l/m2.h.bar) and (b) RB retention (, %) [191].
154 Hoang Vinh-Thang and Serge Kaliaguine

Figure 20. SEM images of air-surfaces (left) and cross-sections (right) of ZIF-8/Matrimid MMMs: (a)
30%, (b) 50%, and (c) 80% (w/w) ZIF-8 [197].

Figure 21. Gas separation performances of ZIF-8/Matrimid MMMs at 0%, 20%, 30%, 40%, 50%, and
60% (w/w) ZIF-8 loadings: (a) permeability for N2, CH4, and C3H8; (b) permeability for O2, CO2, and
H2, (c) selectivity for O2/N2, CH4/N2, H2/O2, H2/CO2, and H2/C3H8; and (d) selectivity for CO2/CH4,
H2/CH4, H2/N2, and H2/C3H8 [197].
MOF-Based Mixed-Matrix-Membranesfor Industrial Applications 155

The selectivities of small gases such as H2, CO2, O2, and N2 were also increased. In
contrast, the diffusivities of CH4 and C3H8 decreased continuously with ZIF-8 loading due to
their larger diameters as compared to the pore aperture of ZIF-8 (3.4 ).
Very recently, Diaz et al. presented a study on the CO2 diffusion in ZIF-8/PPEES
(poly(1,4-phenylene ether-ether-sulfone)) MMMs using pulsed file gradient NMR techniques
at room temperature and 6 bar [199]. For membrane preparation, ZIF-8 nanoparticles up to 30
wt.% supplied by Aldrich were slurried into chloroform, sonicated for several periods, and
then added under strong stirring to a chloroform solution containing PPEES. As seen in SEM
micrographs (Figure 22), at 10 and 20 wt.% ZIF-8 loading, MMMs exhibited a rather good
dispersion in the polymer matrix and less interface void around the particles. However, an
aggregation of ZIF-8 crystals occurred for the sample with 30 wt.% loading.
Permeation results revealed that the permeability of CO2 for the ZIF-8/PPEES
MMMsdecreased as the experimental pressure increased. PFG NMR results (Table 12)
showed that the self-diffusion coefficient of CO2for the PPEES membranes is similar to the
apparent diffusion coefficients obtained by the Dual-Mode ModelDDM, and permeation
technique, Deff. Interestingly, the self-diffusion coefficient for the ZIF-8/PPEES MMMswith
different MOFs contents isnearly twice compared to Deff. This result proved that the
concentration of CO2 in Langmuir sites increased at higher filler loadings.

Figure 22. SEM micrographs showing transversal sections of (a) PPEES membrane, (b) 10 wt.%, (c) 20
wt.%, and (d) 30 wt.% ZIF-8/PPEES MMMs [199].
156 Hoang Vinh-Thang and Serge Kaliaguine

Table 12. Summary of values of CO2 diffusion coefficients expressed in cm2s-1,


at 303 K and 6 bar, in PPEES and MMMs, as well as in ZIF-8, determined
by the Dual-Mode Model, DDM, permeation and sorption results, Deff,
the PFG NMR technique, DNMR (Dapp), and the time lag method D[199]

Sample DNMRa 108 D 108 Deff 108 DDM 108


PPEES 2.1 1.87 2.87 2.84
MMM-10 5.2 1.65 1.91 1.89
MMM-20 7.7 2.80 3.68 3.65
MMM-30 9.3 4.72 4.09 4.14
ZIF-8 900b
a
Measurements performed at 298 K. b Experiments were carried out in ZIF-8 as received ( = 0.35 g
cm3) under gas pressure of 2.4 bar.

11. V-MIL-47 AND AL-MIL-53 BASED


MIXED-MATRIX-MEMBRANES
Al-MIL-53 crystals and the isostructural vanadium(4+) analogue, V-MIL-47 materials
are a type of MIL series materials developed by Freys group [218-219]. The structure of the
MIL-53/MIL-47 materials is built up from infinite chains of corner-sharing MO4(OH)2
(M=Al3+, Fe3+, Cr3+) or V4+O6 octahedra interconnected by the dicarboxylate groups. This
results in a 3D metal organic framework containing 1D diamond-shaped channels with pores
of free diameter close to 0.85 nm (Figure 23). Both Al-MIL-53 and V-MIL-47 were reported
as high potential adsorbents for H2 [220], CO2 and CH4 [221],olens, alkylnaphthalenes and
dichlorobenzenes [222],styrene and ethylbenzene [223], etc.

Figure 23. MIL-47 (left) and MIL-53 (right)structures [218].

To our knowledge, however, up to now, the separation performance of Al-MIL-53 and V-


MIL-47 based MMMs, was only studied by Basu et al. in 2009 [191]. In particular in the
separation of Rose Bengal (RB) from isopropanol, V-MIL-47 and Al-MIL-53 particles as
fillers were mixed with polydimethylsiloxane (PDMS). Al-MIL-53 crystals are commercial
Basolite A100TM products purchased from Sigma-Aldrich. The preparation of two MMMs is
similar to that of Cu3(BTC)2 based MMMssynthesized by the same authors (see section 7). As
an example, cross-section SEM images of unfilled and Al-MIL-53/PDMS MMMs showed a
fairly good dispersion (Figure 24).
MOF-Based Mixed-Matrix-Membranesfor Industrial Applications 157

Figure 24.(a) SEM cross-section view of an unfilled PDMS membrane with a thickness of 28.81.2m
and (b) top view of a 20 wt.% Al-MIL-53/PDMS MMMs. Cross-sections of MSTFA modified Al-MIL-
53/PDMS MMMs: (c) 5wt.%, (d) 10 wt.%, (e) 15 wt.%, and (f) 20 wt.%, taken at a magnification of
3500 [191].

Figure 25. (a, b) Permeance of V-MIL-47 and Al-MIL-53/PDMS MMMs loaded with different contents
of unmodified and MSTFA functionalized MOFs,respectively; and (c, d) RB permeance (, l/m2.h.bar)
and RB retention (, %) of MSTFA functionalized MOF-based MMMs, respectively [191].
158 Hoang Vinh-Thang and Serge Kaliaguine

As shown in Figure 25, the permeance of MMMs filled and MSTFA-unmodified V-MIL-
47 and Al-MIL-53 crystals were improved with the MOF loading. However, the RB retention
was decreased. This was attributed to the presence of a large number of non-selective voids
(see Figure 24). The modification with MSTFA enhanced both RB permeability and RB
retention over a period of filtration of 20 h.

12. ZIF-90BASED MIXED-MATRIX-MEMBRANES


Very recently, Koros and co-authors presented an interesting paper related to a high-
performance gas-separation MMMs containing submicrometer-sized ZIF-90 crystals [200].
ZIF-90 has a sodalite cagelike structure with 0.35 nm pore windows, through which size
exclusion of CH4 from CO2/CH4 mixtures is possible. Furthermore, the imidazole linker in
ZIF-90 contains a carbonyl group, which has a favorable chemical noncovalent interaction
with CO2. ZIF-90 is a Zn-based MOFs which was synthesized from Zn(NO3)2 and imidazole-
2-carboxaldehyde in DMF. By quickly adding nonsolvents such as methanol and water in the
synthesis medium, these authors obtained ZIF-90A and ZIF-90B crystals, with particle sizes
of 0.81 and 0.66-2 m, respectively, as shown in Figure 26.

Figure 26. SEM images of submicrometer-sized ZIF-90 particles; a) ZIF-90A synthesized using
methanol as a nonsolvent; b) ZIF-90B synthesized using deionized water as a nonsolvent [200].

Based on these ZIF-90 particles, three kind of 15% wt% ZIF-90-based MMMs were
synthesized using different poly(imide)s such as Ultem 1000 (SABIC), Matrimid 5218
(Vantico), and 6FDA-DAM (6FDA-DAM poly(imide) (6FDA: 2,2-bis (3,4-carboxyphenyl)
hexafluoropropane dianhydride and DAM: diaminomesitylene; molecular weight ~128,000,
polydispersity index = 5.97). ZIF-90 particles were dispersed in DCM (dichloromethane) and
the mixture was sonicated with an ultrasound horn followed by adding of the polymers.
As seen in Figures 27-28, ZIF-90-based Ultem and Matrimid MMMs showed
significantly enhanced CO2permeability without any loss in CO2/CH4 selectivity. The
invariable selectivity was attributed to the mismatch between the permeabilities of ZIF-90
(several thousand barrer) and those of Ultem and Matrimid polyimides (1-10 Barrer). On the
other hand, ZIF-90/6FDA-DAM MMMs (a highly permeable polymer) showed substantial
enhancements in both CO2 permeability and CO2/CH4 selectivity. The 6FDA-DAM MMMs
containing smaller particles (ZIF-90A, =720 Barrer, = 37) showed slightly better
performance than that with higher particle sizes (ZIF-90B, =590 Barrer, = 34),
MOF-Based Mixed-Matrix-Membranesfor Industrial Applications 159

whereas commercial 6FDA-DAM membranes achieved and values around of 400


Barrer and 15, respectively. Furthermore, the authors also reported a result of pure-
component permeation measurement of N2 on the ZIF-90A/6FDA-DAM MMMs, for
example, an ideal CO2/N2 selectivity of 22 as compared to 14 for pure 6FDA-DAM
polyimide.

Figure 27. Gas-permeation properties of mixed-matrix membranes containing 15 wt.% of ZIF-90


crystals measured with pure gases [200].

Figure 28. CO2/CH4 separation performance of the ZIF-90 based MMMs, compared with the compiled
data on MOF-containing MMMs [200].

From this result, the authors showed that the performances of ZIF-90 membranes with
Ultem and Matrimid polyimides are comparable with other MOF-based MMMs because of
the low permeability. In contrast, ZIF-90/6FDA-DAM MMMs showed much higher CO2
160 Hoang Vinh-Thang and Serge Kaliaguine

permeability (>700 barrer) than any other MOF-based MMMs, combined with a good
CO2/CH4 mixed-gas selectivity.

13. SUMMARY AND FUTURE OPPORTUNITIES


Based on the mentioned-above discussions, mixed-matrix membranes with MOF-based
particles embedded in a continuous polymer matrix have enormous potential in gas separation
applications. These membranes are relatively straightforward to prepare and suitable for all
kinds of MOFs particles dispersing in the polymer matrix. Although a very large number of
MOFs are known, only not more than ten kinds of MOFs have been used to synthesize
MMMs. Modification of fillers and polymer matrix to improve their interface interaction and
fillers dispersion for MMMs have become the object of an intense research attention.
However, no research on the introduction of functional groups into MOF particles, which
could also change chemical affinities of gas penetrants, i.e. solubility and diffusivity, has been
reported yet.
Then, more research and development are still required in order to explore MOF-based
MMMs for gas separation. The following topics might therefore be proposed for the future
research on MOF-based MMMs:

(i) Understanding the basic interactions between the polymers and the MOF particles,
(ii) Synthesizing nano-sized MOF particles without agglomeration;
(iii) Understanding the intrinsic separation performances of MOFs;
(iv) Synthesizing new MOFs with excellent separation and storage properties;
(v) Functionalizing MOFs with halogeno- or aminogroups to improve the adhesion and
compatibility between the surface of MOFs and the polymers;
(vi) Developing novel approaches to uniformly and easily disperse MOF particles in the
continuous polymer matrix;
(vii) Developing new prediction models to guide the selection of both MOFs and
polymers with good MMM separation performance.
(viii) Developing new applications of MMMs not only for gas separations, but also for
other industrial processes, i.e., dry bio-ethanol production for bio-fuels

BIBLIOGRAPHY
[1] Henis, J. M. S.; Tripodi, M. K. Sep. Sci. Technol. 1980, 15, 1059-1068.
[2] Baker, R. W. Membrane Technology and Applications; Wiley and Sons Ltd:
Chichester, 2004.
[3] Kulkarni, S. S.; Hasse, D. J. US Patent 2010, 7,776,137.
[4] Basu, S.; Khan, A.L.; Cano-Odena, A.; Liu, C.; Vankelecom, Ivo F.J. Chem. Soc. Rev..
2010, 39, 750-768.
[5] Cranford, R.; Roy, C. US Patent 2009, 7,556,677.
[6] Burns, R. L.; Koros, W. J. J. Membr. Sci. 2003, 211, 299-309.
[7] Koros, W. J.; Fleming, G. K. J. Membr. Sci. 1993, 83, 1-80.
MOF-Based Mixed-Matrix-Membranesfor Industrial Applications 161

[8] Matsuura, T. Synthetic Membranes and Membrane Separation Processes; CRC Press:
Boca Raton, FL, 1994.
[9] Stern, S. A. J. Membr. Sci. 1994, 94, 1-65.
[10] Singh, R. Chemtech., 1998, 28, 33-44.
[11] Koros, W. J.; Mahajan, R. J. Membr. Sci. 2000, 175, 181-196.
[12] Pandey, P.; Chauhan, R. S. Prog. Polym. Sci. 2001, 26, 853-893.
[13] Baker, R.W. Ind. Eng. Chem. Res. 2002, 41, 1393-1411.
[14] Bredesen, R.; Jordal, K.; Bolland, O. Chem. Eng. Process. 2004, 43, 1129-1158.
[15] Smitha, B.; Suhanya, D.; Sridhar, S.; Ramakrishna, M. J. Membr. Sci. 2004, 241, 1-21.
[16] Lin, H.; Freeman, B. D. J. Mol. Struct. 2005, 739, 57-74.
[17] Ulbricht, M. Polymer 2006, 47, 2217-2262.
[18] Sridhar, S.; Smitha, B.; Aminabhavi, T. M. Sep. Purif. Rev. 2007, 36, 113-174.
[19] Bernardo, P.; Drioli, E.; Golemme, G. Ind. Eng. Chem. Res. 2009, 48, 4638-4663.
[20] Yampolskii, Y.; Freeman, B. Membrane Gas Separation; John Wiley and Sons Ltd,
2010.
[21] Favre, E. In Comprehensive Membrane Science and Engineering; Eds.: Drioli, E.;
Giorno,L.; Elsevier, 2010; Vol. 2, pp 155-212.
[22] Robeson, L. M. J. Membr. Sci. 2008, 320, 390-400.
[23] Lin, Y.S. Sep. Purif. Technol. 2001, 25, 39-55.
[24] Bowen, T. C.; Noble, R. D.; Falconer, J. L. J. Membr. Sci. 2004, 245, 1-33.
[25] Caro, J.; Noack, M. Micropor. Mesopor. Mater. 2008, 115, 215-233.
[26] Caro, J.; Noack, M. In Advances in Nanoporous Materials; Elsevier, 2009, Vol. 1, pp 1-
96.
[27] van den Bergh, J.; Nishiyama, N.; Kapteijn, F. In Novel Concepts in Catalysis and
Chemical Reactors: Improving the Efciency for the Future; Eds.: Cybulski, A.;
Moulijn, J. A.; Stankiewicz, A.; Wiley-VCH Verlag GmbH and Co. KGaA, 2010, pp
211-238.
[28] Mahajan, R.; Zimmerman, C. M.; Koros, W. J. In Polymer Membranes for Gas and
Vapor Separation - Chemistry and Materials Science, Eds.: Freeman, B. D.; Pinnau, I.;
ACS Symposium Series, Vol. 733; American Chemical Society, 1999, pp 277-286.
[29] Chung, T. -S.; Jiang, L. Y.; Li, Y.; Kulprathipanja, S. Prog. Polym. Sci. 2007, 32, 483-
507.
[30] Liu, C.; Kulprathipanja, S.; Hillock, A. M. W.; Husain, S.; Koros, W. J. In Advanced
Membrane Technology and Applications; Eds.: Li, N. N.; Fane, A. G.; Winston Ho,
W.S.; Matsuura, T.; John Wiley and Sons Ltd, 2008, pp 789-819.
[31] Nunes, S. P. In Inorganic Membranes: Synthesis, Characterization and Applications;
Eds.: Mallada, R.; Menndez, M.; Membrane Science and Technology Series, Vol. 13;
Elsevier, 2008, pp 121-134.
[32] Liu, C.; Kulprathipanja, S. In Zeolites in Industrial Separation and Catalysis; Ed.:
Kulprathipania, S.; Wiley-VCH Verlag GmbH and Co. KGaA, Weinheim, 2010, pp
329-353.
[33] Aroon, M. A.; Ismail, A. F.; Matsuura, T.; Montazer-Rahmati, M. M. Sep. Purif.
Technol. 2010, 75, 229-242.
[34] Ismail, A. F.; Goh, P. S.; Sanip, S. M.; Aziz, M. Sep. Purif. Technol. 2009, 70, 12-26.
[35] Okumus, E.; Gurkan, T.; Yilmaz, L. Sep. Sci. Technol. 1994, 29, 2451-2473.
[36] Miyazawa, K.; Inagaki, S. Chem. Commun. 2000, 2121-2122.
162 Hoang Vinh-Thang and Serge Kaliaguine

[37] Kruk, M.; Jaroniec, M.; Ko, C. H.; Ryoo, R. Chem. Mater. 2000, 12, 1961-1968.
[38] Ravikovitch, P. I.; Neimark, A. V. J. Phys. Chem. B. 2001, 105, 6817-6823.
[39] Vinh-Thang, H.; Huang, Q.; Ei, M.; Trong-On, D.; Kaliaguine, S. Langmuir 2005, 21,
2051-2057.
[40] Liu, Y.; Zhang, W.; Pinnavaia, T.J. J. Am. Chem. Soc. 2000, 122, 8791-8792.
[41] Trong-On, D.; Kaliaguine, S. US Patent 2003, 6,669,924 B1.
[42] Trong-On, D.; Kaliaguine, S. Angew. Chem. Int. Ed. 2001, 40, 3248-3251.
[43] Vinh-Thang, H.; Huang, Q.; Ungureanu, A.; Ei, M.; Trong-On, D.; Kaliaguine, S.
Langmuir 2006, 22, 4777-4786.
[44] Nohair, B.; MacQuarrie, S.; Crudden, C. M.; Kaliaguine, S. J. Phys. Chem. C.2008,
112, 6065-6072.
[45] Ghaffari-Nik, O.; Nohair, B.; Kaliaguine, S. Micropor. Mesopor. Mater. 2011, 143,
221-223.
[46] Ghaffari-Nik, O.; Chen, X. Y.; Kaliaguine, S. J. Membr. Sci.,in press.
[47] Petropoulos, J. H. J. Polym. Sci., Polym. Phys. Ed. 1985, 23, 1309-1324.
[48] Bouma, R. H. B.; Checchetti, A.; Chidichimo, G.; Drioli, E. J. Membr. Sci., 1997, 128,
141-149.
[49] Maxwell, J. C. Treatise on Electricity and Magnetism; Oxford University Press,
London, 1873.
[50] Bruggeman, D. A. G. Ann. Phys. (Leipzig) 1935, 24, 636-679.
[51] Lewis, T.; Nielsen, L. J. Appl. Polym. Sci. 1970, 14, 1449-1471.
[52] Nielsen, L. J. Appl. Polym. Sci. 1973, 17, 3819-3820.
[53] Pal, R. J. Reinf. Plast. Compos. 2007, 26, 643-651.
[54] Pal, R. J. Colloid Inter. Sci. 2008, 317, 191-198.
[55] Moore, T. T.; Mahajan, R.; Vu, D. Q.; Koros, W. J. AIChE J. 2004, 50, 311-321.
[56] Felske, J. D. Int. J. Heat Mass Transfer. 2004, 47, 3453-3461.
[57] Bottcher, C. J. F. Recueil des Travaux Chimiques des Pays-Bas 1945, 64, 47-51.
[58] Higuchi, W. I.; Higuchi, T. J. Am. Pharm. Assoc. Sci. 1960, 49, 598-606.
[59] Cheng, S. C.; Vachon, R. I. Int. J. Heat Mass Transfer. 1969, 12, 249-264.
[60] Agari, Y.; Uno, T. J. Appl. Polym. Sci. 1986, 32, 5705-5712.
[61] Hashemifard, S. A.; Ismail, A. F.; Matsuura, T. J. Membr. Sci. 2010, 347, 53-61.
[62] Hashemifard, S. A.; Ismail, A. F.; Matsuura, T. J. Membr. Sci. 2010, 350, 259-268.
[63] Yaghi, O. M.; Li, H. L. J. Am. Chem. Soc. 1995, 117, 10401-10402.
[64] Yaghi, O. M.; Li, H. L.; Davis, C.; Richardson, D.; Groy, T. L. Acc. Chem. Res. 1998,
31, 474-484.
[65] Li, H. L.; Eddaoudi, M.; O'Keeffe, M.; Yaghi, O. M. Nature, 1999, 402, 276-279.
[66] Kaskel, S. In Handbook of Porous Solids, Eds.: Sch t h, F.; Sing, K. S. W.; Weitkamp,
J.; Wiley-VCH Verlag GmbH, Weinheirn, 2002, Vol. 2, pp 1190-1249.
[67] James, S. L. Chem. Soc. Rev. 2003, 32, 276-288.
[68] Yaghi, O. M.; OKeeffe, M.; Ockwig, N. W.; Chae, H. K.; Eddaoudi, M.; Kim, J.
Nature 2003, 423, 705-714.
[69] Rowsell, J. L. C.; Yaghi, O. M. Micropor. Mesopor. Mater. 2004, 73, 3-14.
[70] Rosseinsky, M. J. Micropor. Mesopor. Mater. 2004, 73, 15-30.
[71] Kitagawa, S.; Kitaura, R.; Noro, S. Angew. Chem. Int. Ed. 2004, 43, 2334-2375.
[72] Uemura, K.; Matsuda, R.; Kitagawa, S. J. Solid State Chem. 2005, 178, 2420-2429.
MOF-Based Mixed-Matrix-Membranesfor Industrial Applications 163

[73] Ct, A. P.; Benin, A.; Ockwig, N.; OKeeffe, M.; Matzger, A.; Yaghi, O. M. Science
2005, 310, 1166-1170.
[74] Park, K. S.; Ni, Z.; Ct, A. P.; Choi, J. Y.; Huang, R.; Uribe-Romo, F.; Chae, H.;
O'Keefe, M.; Yaghi, O. M. Proc. Nat. Acad. Sci. USA 2006, 103, 10186-10191.
[75] Hayashi, H.; Ct, A.P.; Furukawa, H.; O'Keefe, M.; Yaghi, O. M. Nat. Mater. 2007, 6,
501-506.
[76] Frey, G. Stud. Surf. Sci. Catal. 2007, 168, 327-374.
[77] Frey, G. Stud. Surf. Sci. Catal. 2007, 170, 66-84.
[78] Frey, G. Chem. Soc. Rev. 2008, 37, 191-241.
[79] Wright, P. A. In Microporous Framework Solids; Ed.: J.A. Connor; RSC Materials
Monographs Series; The Royal Society of Chemistry, 2008, pp 46-70.
[80] Tranchemontagne, D. J.; Ni, Z.; OKeeffe, M.; Yaghi, O. M. Angew. Chem. Int. Ed.
2008, 47, 5136 -5147.
[81] Frey, G. Dalton Trans. 2009, 4400-4415.
[82] Prakash, M. J.; Lah, M. S. Chem. Commun. 2009, 3326-3341.
[83] MacGillivray, L. R. Metal-organic frameworks: design and application; John Wiley
and Sons, Inc., Hoboken, New Jersey, 2010.
[84] Fang, Q.; Sculley, J.; Zhou, H. -C. J.; Zhu, G. In Comprehensive Nanoscience and
Technology;Eds: Andrews, D. L.; Scholes, G. D.; Wiederrecht, G.P.;Academic Press,
Elsevier, 2010; Vol. 5, pp 1-20.
[85] Janiak, C.; Vieth, J. K. New J. Chem. 2010, 34, 2366-2388.
[86] Natarajan, S.; Mahata, P. Chem. Soc. Rev. 2009, 38, 2304-2318.
[87] Alaerts, L.; de Vos, D. E. In Novel Concepts in Catalysis and Chemical Reactors; Eds.:
Cybulski, A.; Moulijn, J. A.; Stankiewicz, A.; Wiley-VCH Verlag GmbH and Co.
KGaA, Weinheim; 2010, pp 73-94.
[88] Meek, S. T.; Greathouse, J. A.; Allendorf, M. D. Adv. Mater. 2011, 23, 249-267.
[89] Carn, A.; Carbonell, C.; Imaz, I.; Maspoch, D. Chem. Soc. Rev. 2011, 40, 291-305.
[90] Park, Y.; Choi, S.; Kim, H.; Kim, K.; Won, B. -H.; Choi, K.; Choi, J. -S.; Ahn, W. -S.;
Won, N.; Kim, S.; Jung, D.; Choi, S. -H.; Kim, G. -H.; Cha, S. -S.; Jhon, Y.; Yang, J.;
Kim, J. Angew. Chem. Int. Ed. 2007, 46, 8230-8233.
[91] Qiu, L. -G.; Xu, T.; Li, Z. -Q.; Wang, W.; Wu, Y.; Jiang, X.; Tian, X. -Y.; Zhang, L. -
D. Angew. Chem. Int. Ed. 2008, 47, 9487-9491.
[92] Jiang, H. -L.; Tatsu, Y.; Lu, Z.-H.; Xu, Q. J. Am. Chem. Soc. 2010, 132, 5586-5587.
[93] Xuan-Dong, D.; Vinh-Thang, H.; Kaliaguine, S. Micropor. Mesopor. Mater. 2010, 141,
135-139.
[94] Tranchemontagne, D. J.; Mendoza-Corts, J. L.; OKeeffe, M.; Yaghi, O. M. Chem.
Soc. Rev. 2009, 38, 1257-1283.
[95] Wang, Z.; Cohen, S. M. Chem. Soc. Rev. 2009, 38, 1315-1329.
[96] Perry IV, J. J.; Perman, J. A.; Zaworotko, M. J. Chem. Soc. Rev. 2009, 38, 1400-1417.
[97] Zacher, D.; Schmid, R.; Wll, C.; Fischer, R. A. Angew. Chem. Int. Ed. 2011, 50, 176-
199.
[98] Klinowski, J.; Paz, F. A. A.; Silva, P.; Rocha, J. Dalton Trans. 2011, 40, 321-330.
[99] Cohen, S. M. Chem. Sci. 2010, 1, 32-36.
[100] Farha, O. K.; Hupp, J. T. Acc. Chem. Res.2010, 43, 1166-1175.
[101] Rowsell, J. L. C.; Yaghi, O. M. Angew. Chem. Int. Ed. 2005, 44, 4670-4679.
[102] Hu, Y. H.; Zhang, L. Adv. Mater. 2010, 22, E117-E130.
164 Hoang Vinh-Thang and Serge Kaliaguine

[103] Murray, L. J.; Dinca, M.; Long, J. R. Chem. Soc. Rev. 2009, 38, 1294-1314.
[104] Lin, X.; Jia, J.; Hubberstey, P.; Schrder, M.; Champness, N. R. Cryst. Eng. Comm.
2007, 9, 438-448.
[105] Zhao, D.; Yuan, D.; Zhou, H. -C. Energy Environ. Sci. 2008, 1, 222-235.
[106] Thomas, K. M. Dalton Trans. 2009, 38, 1487-1505.
[107] Collins, D. J.; Zhou, H. -C. J. Mater. Chem. 2007, 17, 3154-3160.
[108] Xiao, B.; Yuan, Q. Particuology 2009, 7, 129-140.
[109] Meilikhov, M.; Yusenko, K.; Esken, D.; Turner, S.; van Tendeloo, G.; Fischer, R. A.
Eur. J. Inorg. Chem. 2010, 3701-3714.
[110] Isaeva, V. I.; Kustov, L. M. Russ. J. Gen. Chem. 2007, 77, 721-739.
[111] Dinca, M.; Long, J. R. Angew. Chem. Int. Ed. 2008, 47, 6766-6779.
[112] Schrder, M. Functional Metal-Organic Frameworks: Gas Storage, Separation and
Catalysis; Topics in Current Chemistry, Vol. 293; Springer, 2010.
[113] Phan, A.; Doonan, C. J.; Uribe-Romo, F. J.; Knobler, C. B.; OKeeffe, M.; Yaghi, O.
M. Acc. Chem. Res.2010, 43, 58-67.
[114] Li, J. -R.; Kuppler, R. J.; Zhou, H. -C. Chem. Soc. Rev. 2009, 38, 1477-1504.
[115] Keskin, S.; Liu, J.; Rankin, R. B.; Johnson, J. K.; Sholl, D. S. Ind. Eng. Chem. Res.
2009, 48, 2355-2371.
[116] Liu, D.; Zhong, C. J. Mater. Chem. 2010, 20, 10308-10318.
[117] Zou, R.; Abdel-Fattah, A. I.; Xu, H.; Zhao, Y.; Hickmott, D. D.; Cryst. Eng. Comm.
2010, 12, 1337-1353.
[118] Fletcher, A. J.; Thomas, K. M.; Rosseinsky, M. J. J. Solid State Chem. 2005, 178, 2491-
2510.
[119] Keskin, S.; van Heest, T. M.; Sholl, D. S. Chem. Sus. Chem. 2010, 3, 879-891.
[120] Liu, Y.; Xuan, W.; Cui, Y. Adv. Mater. 2010, 22, 4112-4135.
[121] Farrusseng, D.; Aguado, S.; Pinel, C. Angew. Chem. Int. Ed. 2009, 48, 7502-7513.
[122] Corma, A.; Garcia, H.; Llabrs i Xamena, F. X. Chem. Rev. 2010, 110, 4606-4655.
[123] Ranocchiari, M.; van Bokhoven, J. A. Phys. Chem. Chem. Phys. 2011,13, 6388-6396.
[124] Mueller, U.; Schubert, M.; Teich, F.; Puetter, H.; Schierle-Arndt, K.; Pastr, J. J. Mater.
Chem. 2006, 16, 626-636.
[125] Custelcean, R.; Moyer, B. A. Eur. J. Inorg. Chem. 2007, 1321-1340.
[126] Uemura, T.; Yanai, N.; Kitagawa, S. Chem. Soc. Rev. 2009, 38, 1228-1236.
[127] Ma, L.; Abney, C.; Lin, W. Chem. Soc. Rev. 2009, 38, 1248-1256.
[128] Czaja, A. U.; Trukhan, N.; Mller, U. Chem. Soc. Rev. 2009, 38, 1284-1293.
[129] Lee, J. -Y.; Farha, O. K.; Roberts, J.; Scheidt, K. A.; Nguyen, S. B. T.; Hupp, J. T.
Chem. Soc. Rev. 2009, 38, 1450-1459.
[130] Isaeva, V. I.; Kustov, L. M. Petroleum Chemistry 2010, 50, 167-180.
[131] Jiang, H. -L.; Xu, Q. Chem. Commun. 2011, 47, 3351-3370.
[132] Silva, C. G.; Corma, A.; Garcia, H. J. Mater. Chem. 2010, 20, 3141-3156.
[133] Allendorf, M. D.; Bauer, C. A.; Bhakta, R. K.; Houk, R. J. T. Chem. Soc. Rev. 2009, 38,
1330-1352.
[134] Rocha, J.; Carlos, L. D.; Paz, F. A. A.; Ananias, D. Chem. Soc. Rev. 2011, 40, 926-940.
[135] Kurmoo, M. Chem. Soc. Rev. 2009, 38, 1353-1379.
[136] Maspoch, D.; Ruiz-Molina, D.; Veciana, J. J. Mater. Chem. 2004, 14, 2713-2723.
[137] Zacher, D.; Shekhah, O.; Wll, C.; Fischer, R. A. Chem. Soc. Rev. 2009, 38, 1418-
1429.
MOF-Based Mixed-Matrix-Membranesfor Industrial Applications 165

[138] Shekhah, O.; Liu, J.; Fischer, R. A.; Wll, C. Chem. Soc. Rev. 2011, 40, 1081-1106.
[139] McKinlay, A. C.; Morris, R. E., Horcajada, P.; Frey, G.; Gref, R.; Couvreur, P.; Serre,
C. Angew. Chem. Int. Ed. 2010, 49, 6260-6266.
[140] Keskin, S.; Kizilel, S. Ind. Eng. Chem. Res. 2011, 50, 1799-1812.
[141] Yoo, Y.; Jeong, H. -K. Chem. Commun. 2008, 2441-2443.
[142] Hermes, S.; Zacher, D.; Baunemann, A.; Wll, C.; Fischer, R. A. Chem. Mater. 2007,
19, 2168-2173.
[143] Hermes, S.; Schrder, F.; Chelmowski, R.; Wll, C.; Fischer, R. A. J. Am. Chem. Soc.
2005, 127, 13744-13745.
[144] Liu, Y.; Ng, Z.; Khan, E. A.; Jeong, H. -K.; Ching, C.; Lai, Z. Micropor.
Mesopor.Mater. 2009, 118, 296-301.
[145] Yoo, Y.; Lai, Z.; Jeong, H. -K. Micropor. Mesopor. Mater. 2009, 123, 100-106.
[146] Ameloot, R.; Stappers, L.; Fransaer, J.; Alaerts, L.; Sels, B. F.; de Vos, D. E. Chem.
Mater. 2009, 21, 2580-2582.
[147] Biemmi, E.; Scherb, C.; Bein, T. J. Am. Chem. Soc. 2007, 129, 8054-8055.
[148] Shekhah, O.; Wang, H.; Kowarik, S.; Schreiber, F.; Paulus, M.; Tolan, M.; Sternemann,
C.; Evers, F.; Zacher, D.; Fischer, R. A.; Wll, C. J. Am. Chem. Soc. 2007, 129, 15118-
15119.
[149] Guo, H.; Zhu, G.; Hewitt, I. J.; Qiu, S. J. Am. Chem. Soc. 2009, 131, 1646-1647.
[150] Zacher, D.; Baunemann, A.; Hermes, S.; Fischer, R. A. J. Mater. Chem. 2007, 17,
2785-2792.
[151] Biemmi, E.; Darga, A.; Stock, N.; Bein, T. Micropor. Mesopor. Mater. 2008, 114, 380-
386.
[152] Ameloot, R.; Gobechiya, E.; Uji-i, H.; Martens, J. A.; Hofkens, J.; Alaerts, L.; Sels, B.
F.; de Vos, D. E. Adv. Mater. 2010, 22, 2685-2688.
[153] Schoedel, A.; Scherb, C.; Bein, T. Angew. Chem. Int. Ed. 2010, 49, 7225-7228.
[154] Guerrero, V. V.; Yoo, Y.; McCarthy, M. C.; Jeong, H. -K. J. Mater. Chem. 2010, 20,
3938-3943.
[155] Gascon, J.; Aguado, S.; Kapteijn, F. Micropor. Mesopor. Mater. 2008, 113, 132-138.
[156] Allendorf, M. D.; Houk, R. J. T.; Andruszkiewicz, L.; Talin, A. A.; Pikarsky, J.;
Choudhury, A.; Gall, K. A.; Hesketh, P. J. J. Am. Chem. Soc. 2008, 130, 14404-14405.
[157] Zybaylo, O.; Shekhah, O.; Wang, H.; Tapolsky, M.; Schmid, R.; Johannsmann, D.;
Wll, C. Phys. Chem. Chem. Phys. 2010, 12, 8092-8097.
[158] Nan, J.; Dong, X.; Wang, W.; Jin, W.; Xu, N. Langmuir 2011, 27, 4309-4312.
[159] Ranjan, R.; Tsapatsis, M. Chem. Mater. 2009, 21, 4920-4924.
[160] Arnold, M.; Kortunov, P.; Jones, D. J.; Nedellec, Y.; Krger, J.; Caro, J. Eur. J. Inorg.
Chem. 2007, 60-64.
[161] Horcajada, P.; Serre, C.; Grosso, D.; Boissire, C.; Perruchas, S.; Sanchez, C.; Frey, G.
Adv. Mater. 2009, 21, 1931-1935.
[162] Aguado, S.; Nicolas, C. -H.; Moizan-Basl, V.; Nieto, C.; Amrouche, H.; Bats, N.;
Audebrand, N.; Farrusseng, D. New J. Chem. 2011, 35, 41-44.
[163] Hu, Y.; Dong, X.; Nan, J.; Jin, W.; Ren, X.; Xu, N.; Lee, Y. M. Chem. Commun. 2011,
47, 737-739.
[164] Zacher, D.; Yusenko, K.; Btard, A.; Henke, S.; Molon, M.; Ladnorg, T.; Shekhah, O.;
Schpbach, B.; de los Arcos, T.; Krasnopolski, M.; Meilikhov, M.; Winter, J.; Terfort,
A.; Wll, C.; Fischer, R. A. Chem. Eur. J. 2011, 17, 1448-1455.
166 Hoang Vinh-Thang and Serge Kaliaguine

[165] Shekhah, O. Materials 2010, 3, 1302-1315.


[166] Li, Y. -S.; Bux, H.; Feldhoff, A.; Li, G. -L.; Yang, W. -S.; Caro, J. Adv. Mater. 2010,
22, 3322-3326.
[167] Li, Y. -S.; Liang, F. -Y.; Bux, H.; Feldhoff, A.; Yang, W. -S.; Caro, J. Angew.
Chem.Int. Ed. 2010, 49, 548-551.
[168] Huang, A.; Bux, H.; Steinbach, F.; Caro, J. Angew. Chem. Int. Ed. 2010, 49, 4958-4961.
[169] Li, Y. -S.; Liang, F. -Y.; Bux, H.; Yang, W.; Caro, J. J. Membr. Sci. 2010, 354, 48-54.
[170] McCarthy, M. C.; Varela-Guerrero, V.; Barnett, G. V.; Jeong, H. -K. Langmuir 2010,
26, 14636-14641.
[171] Bux, H.; Chmelik, C.; van Baten, J. M.; Krishna, R.; Caro, J. Adv. Mater. 2010, 22,
4741-4743.
[172] Bux, H.; Liang, F. -Y.; Li, Y. -S.; Cravillon, J.; Wiebcke, M.; Caro, J. J. Am. Chem.
Soc. 2009, 131, 16000-16001.
[173] Venna, S. R.; Carreon, M. A. J. Am. Chem. Soc. 2010, 132, 76-78.
[174] Lu, G.; Hupp, J. T. J. Am. Chem. Soc. 2010, 132, 7832-7833.
[175] Bux, H.; Chmelik, C.; Krishna, R.; Caro, J. J. Membr. Sci. 2010, 369, 284-289.
[176] Liu, Y.; Hu, E.; Khan, E. A.; Lai, Z. J. Membr. Sci. 2010, 353, 36-40.
[177] Guo, H.; Zhu, Y.; Qiu, S.; Lercher, J. A.; Zhang, H. Adv. Mater. 2010, 22, 4190-4192.
[178] Keskin, S.; Liu, J.; Johnson, J. K.; Sholl, D. S. Micropor. Mesopor. Mater. 2009, 125,
101-106.
[179] Keskin, S.; Sholl, D. S. Langmuir 2009, 25, 11786-11795.
[180] Watanabe, T.; Keskin, S.; Nair, S.; Sholl, D. S. Phys. Chem. Chem. Phys. 2009, 11,
11389-11394.
[181] Keskin, S. Ind. Eng. Chem. Res. 2010, 49, 11689-11696.
[182] Krishna, R.; van Baten, J. M. J. Membr. Sci. 2010, 360, 323-333.
[183] Yehia, H.; Pisklak, T. J.; Ferraris, J. P.; Balkus Jr., K. J.; Musselman, I. H. Polymer
Preprints 2004, 45, 35-36.
[184] Liu, C.; Kulprathipanja, S.; Wilson, S. T. US patent 2007, US2007/0209505A1.
[185] Liu, C.; McCulloch, B.; Wilson, S. T.; Benin, A. I.; Schott, M. E. US patent 2009,
US7637983B1.
[186] Perez, E. V.; Balkus Jr., K.J.; Ferraris, J. P.; Musselman, I. H. J. Membr. Sci. 2009, 328,
165-173.
[187] Adams, R.; Carson, C.; Ward, J.; Tannenbaum, R.; Koros, W. Micropor. Mesopor.
Mater., 2010, 131, 13-20.
[188] Keskin, S.; Sholl, D. S. Energy Environ. Sci. 2010, 3, 343-351.
[189] Keskin, S. J. Phys. Chem. C. 2010, 114, 13047-13054.
[190] Car, A.; Stropnik, C.; Peinemann, K. -V. Desalination 2006, 200, 424-426.
[191] Basu, S.; Maes, M.; Cano-Odena, A.; Alaerts, L.; de Vos, D. E.; Vankelecom, I. F. J. J.
Membr. Sci. 2009, 344, 190-198.
[192] Basu, S.; Cano-Odena, A.; Vankelecom, I. F. J. J. Membr. Sci. 2010, 362, 478-487.
[193] Hu, J.; Cai, H.; Ren, H.; Wei, Y.; Xu, Z.; Liu, H.; Hu, Y. Ind. Eng. Chem. Res. 2010,
49, 12605-12612.
[194] Zhang, Y.; Musselman, I. H.; Ferraris, J. P.; Balkus Jr., K. J. J. Membr. Sci. 2008, 313,
170-181.
[195] Musselman, I. H.; Balkus Jr., K. J.; Ferraris, J. P. Report DE-FG26-04NT42173, 2008.
[196] Perez, E. V. Ph.D Thesis, University of Texas at Dallas, 2009.
MOF-Based Mixed-Matrix-Membranesfor Industrial Applications 167

[197] Ordonez, Ma J. C.; Balkus Jr., K. J.; Ferraris, J. P.; Musselman, I. H. J. Membr. Sci.
2010, 361, 28-37.
[198] Ordonez, Ma J. C. M.S Thesis, University of Texas at Dallas,2009.
[199] Diaz, K.; Garrido, L.; Lopez-Gonzalez, M.; del Castillo, L. F.; Riande, E.
Macromolecules 2010, 43, 316-325.
[200] Bae, T. -H.; Lee, J. S.; Qiu, W.; Koros, W. J.; Jones, C. W.; Nair, S. Angew. Chem.
Int.Ed. 2010, 49, 9863-9866.
[201] Huang, L.; Wang, H.; Chen, J.; Wang, Z.; Sun, J.; Zhao, D.; Yan, Y. Micropor.
Mesopor. Mater. 2003, 58, 105-114.
[202] Pan, L.; Sander, M. B.; Huang, X.; Li, J.; Smith, M.; Bittner, E.; Bockrath, B.; Johnson,
J. K. J. Am. Chem. Soc. 2004, 126, 1308-1309.
[203] Pan, L.; Olson, D. H.; Ciemnolonski, L. R.; Heddy, R.; Li, J. Angew. Chem. Int. Ed.
2006, 45, 616-619.
[204] Ranjan, R.; Tsapatsis, M. Chem. Mater. 2009, 21, 4920-4924.
[205] Dybtsev, D. N.; Chun, H.; Kim, K. Angew. Chem. Int. Ed. 2004, 43, 5033 -5036.
[206] Lee, J. Y.; Olson, D. H.; Pan, L.; Emge, T. J.; Li, J. Adv. Funct. Mater. 2007, 17, 1255-
1262.
[207] Liu, J.; Lee, J. Y.; Pan, L.; Obermyer, R. T.; Simizu, S.; Zande, B.; Li, J.; Sankar, S. G.;
Johnson, J. K. J. Phys. Chem. C. 2008, 112, 2911-2917.
[208] Millward, A. R.; Yaghi, O. M. J. Am. Chem. Soc. 2005, 127, 17998-17999.
[209] Yang, Q.; Zhong, C. J. Phys. Chem. B. 2006, 110, 17776-17783.
[210] Wang, Q. M.; Shen, D.; Blow, M.; Lau, M. L.; Deng, S.; Fitch, F. R.; Lemcoff, N. O.;
Semanscin, J. Micropor. Mesopor. Mater. 2002, 55, 217-230.
[211] Chui, S. S.; Lo, S. M.; Charmant, J. P. H.; Orpen, A. G.; Williams, I. D. Science 1999,
283, 1148-1150.
[212] Vishnyakov, A.; Ravikovitch, P. I.; Neimark, A. V.; Blow, M.; Wang, Q. M. Nano
Letters 2003, 3, 713-718.
[213] Hartmann, M.; Kunz, S.; Himsl, D.; Tangermann, O. Langmuir 2008, 24, 8634-8642.
[214] Xiang, Z.; Cao, D.; Shao, X.; Wang, W.; Zhang, J.; Wu, W. Chem. Eng. Sci. 2010, 65,
3140-3146.
[215] Noro, S. I.; Kitagawa, S.; Kondo, M.; Seki, K. Angew. Chem. Int. Ed. 2000, 39, 2081-
2084.
[216] Furukawa, H.; Kim, J.; Plass, K. E.; Yaghi, O. M. J. Am. Chem. Soc. 2006, 128, 8398-
8399.
[217] Venna, S. R.; Jasinski, J. B.; Carreon, M. A. J. Am. Chem. Soc. 2010, 132, 18030-
18033.
[218] Bauer, S.; Serre, C.; Devic, T.; Horcajada, P.; Marrot, J.; Frey, G.; Stock, N. Inorg.
Chem. 2008, 47, 7568-7576.
[219] Loiseau, T.; Serre, C.; Huguenard, C.; Fink, G.; Taulelle, F.; Henry, M.; Bataille, T.;
Frey, G. Chem. Eur. J. 2004, 10, 1373-1382.
[220] Salles, F.; Jobic, H.; Maurin, G.; Koza, M. M.; Llewellyn, P. L.; Devic, T.; Serre, C.;
Frey, G. Phys. Rev. Lett. 2008, 100, 245901.
[221] Bourrelly, S.; Llewellyn, P. L.; Serre, C.; Millange, F.; Loiseau, T.; Frey, G. J. Am.
Chem. Soc. 2005, 127, 13519-13521.
[222] Alaerts, L.; Maes, M.; van der Veen, M. A.; Jacobs, P. A.; de Vos, D. E. Phys. Chem.
Chem. Phys. 2009, 11, 2903-2911.
168 Hoang Vinh-Thang and Serge Kaliaguine

[223] Maes, M.; Vermoortele, F.; Alaerts, L.; Couck, S.; Kirschhock, C. E. A.; Denayer, J. F.
M.; de Vos, D. E. J. Am. Chem. Soc. 2010, 132, 15277-15285.
In: Coordination Polymers and Metal Organic Frameworks ISBN: 978-1-61470-899-5
Editors: O. L. Ortiz and L. D. Ramirez, pp. 169-195 2012 Nova Science Publishers, Inc.

Chapter 5

COORDINATION POLYMERS: OPPORTUNITIES


IN HETEROGENEOUS CATALYSIS

Francesc X. Llabrs i Xamena


Instituto de Tecnologa Qumica, Universidad Politcnica de Valencia,
Consejo Superior de Investigaciones Cientficas,
Avda. de los Naranjos s/n, Valencia, Spain.

ABSTRACT
Coordination Polymers (CPs) are an emerging class of materials that are attracting
considerable interest in recent years. Their unique properties make these materials very
promising for applications in a number of fields, including heterogeneous catalysis. In
this chapter, with 126 references, we will revise the main strategies that have been
specifically developed for introducing catalytic active sites in these materials. The
enormous possibilities of this class of materials will be outlined throughout selected
examples taken from the recent literature. I hope that this chapter will be useful either as
an introductory lecture to those who approach the field of CPs or heterogeneous catalysis
for the first time, as well as an updated state-of-the-art vision for all scientists working in
this field.

1. INTRODUCTION
Coordination Polymers (hereafter CPs) are an emerging class of hybrid solid materials
formed by a self-assembled network of metal ions coordinated to polydentate organic ligands.
This is a broad definition covering many different types of materials with very different
characteristics, including amorphous and crystalline solids, as well as porous and non-porous
materials. To delimit the scope of this chapter, only those CPs which are crystalline, porous
and sustained by strong metal-ligand coordination bonds will be considered here [1]. Note
that are outside the scope of this chapter all those hybrid organic-inorganic amorphous
materials, mono- and bidimensional crystalline solids, as well as three-dimensional materials
in which one or more dimensions are supported by weak interactions, such as hydrogen bonds
170 Francesc X. Llabrs i Xamena

and stacking.1 Thus, the materials that will be considered here are usually defined with
more restrictive terms, such as Metal-Organic Frameworks (MOFs), Zeolite-Imidazolate
Frameworks (ZIFs), crystalline porous coordination polymers (PCPs), and others, which
could be considered subclasses of the broader family of CPs. The reason for using the more
restrictive definition throughout this chapter is because only in these materials we can find the
simultaneous occurrence of three important characteristics: crystallinity, porosity, and
existence of strong metal-ligand interactions in three-dimensions imparting robustness. This
unique combination of properties renders CPs a very special class of materials with promising
features for technologic applications, including heterogeneous catalysis.
CPs contain three well differentiated structural parts: the metallic component, the organic
ligand, and the pore system. As it will be demonstrated throughout this chapter, it is possible
to develop CP-based catalysts in which the catalytic activity is located in any of these three
parts, and this introduces an enormous potential for the design of the active center. The
existing strategies for introducing catalytic functionalities at the different parts of the CP will
be revised throughout this chapter and illustrated by selected examples.

1.1. Why Do We Want to Use CPs in Heterogeneous Catalysis?

In the following, the main properties that make CPs so appealing for applications in
heterogeneous catalysis are summarized.
The porous structures of some CPs provides unprecedented enormous apparent surface
2
areas of up to 6150 m2g-1 [2] and specific pore volumes of up to 2 cm3g-1, together with an
almost endless variety of pore dimensions and topologies, ranging from ultramicroporous to
mesoporous and from monodimensional channels to tridimensional cages or cavities. The
elevated pore volumes attained in certain CPs have triggered an intensive research on the
application of these materials for gas separation and storage [3-6], while the simultaneous
occurrence of strong metal-ligand interactions can confer permanent porosity to the materials.
Thus, in many CPs it is possible to remove completely the solvent molecules occluded inside
the pore system, which is necessary to liberate the inner space, without causing the collapse
of the crystalline structure. This property has provided an additional criterion for classifying
CPs into first, second, and third generation materials [7]. First generation CPs are materials
having a pore system sustained by guest molecules that irreversibly collapses when they are
removed. Second generation materials have a robust pore system, featuring permanent
porosity upon evacuation of the guest molecules. Finally, third generation CPs feature a
flexible pore system, which may change reversibly depending on the presence of guest
molecules or by the action of certain external stimuli, such as light, temperature or the
application of an electric field. This last category is also referred to as either dynamic porous
CPs [8] or breathing materials [9, 10]. While first generation CPs would find very limited (or

1
Note that metal-ligand coordination bonds are stronger than hydrogen bonds, and they have also more
directionality tan other weak interaction
2
Note that, most of the CPs described up to now are microporous (pore diameters <2nm). Strictly speaking, the
mechanism of adsorption in these microporous materials is pore filling as opposed to mono/multilayer
adsorption occurring in mesoporous materials. Therefore, the values of specific surface areas for these
materials extracted by using the Langmuir or the BET methods cannot be considered as their tru surface area,
but they have to be taken as apparent or equivalent surface areas.
Coordination Polymers 171

none) use in heterogeneous catalysis, second and third generation CPs could show, and have
already demonstrated [11-13], a very high potential for this and other applications. Thus, this
chapter is mainly focused on second and third generation CPs and their use in catalysis.
Besides permanent porosity, another important feature of CPs is that their pore size,
shape, dimensionality and chemical environment (e.g., hydrophilicity/hydrophobicity) can be
finely tuned by the judicious selection of their components (metal and organic ligand) and the
way in which they are connected (see sections 2.2 and 2.3). This opens the possibility to
control which molecules can diffuse within the pores, since only those molecules smaller that
the pore openings will penetrate the inner space of the solid. Thus, CPs can show molecular
sieve and shape-selective properties, which are highly desired for heterogeneous catalysis.
Meanwhile, both the metallic and organic components can modulate the host-guest
interactions sensed by adsorbed molecules or by the transition states formed during a reaction
occurring within the pores, especially when the dimensions of the substrates are similar to the
pores in which they are confined (the so-called electronic confinement effect [14, 15]). These
interactions can thus activate or orient the substrates around a catalytic center and, in general,
can modify the reactivity in the ground and excited states of adsorbed substrates.
To these advantages, it must be also added the possibility to design CPs containing chiral
catalytic centers [16, 17], which can lead to asymmetric (enantioselective) catalysis.
Asymmetric catalysis is routinely practiced in the synthesis of drugs, usually relying on the
use of expensive homogeneous chiral catalysts. The use of chiral organic ligands in the
synthesis of CPs can lead to a vast number of new chiral heterogeneous catalysts to be tested
3
in this field.

Figure 1. Summary of the most relevant properties of Coordination Polymers with respect to their
potential application in catalysis. Note that, in order to limit the scope of this chapter, only those CPs
that are crystalline, porous and have strong metal-ligand coordination bonds in all three dimensions of
space will be considered.

3 Besides the use of chiral molecules, it is also possible to generate chiral crystals from achiral ligands. In these
cases, the chirality of the architecture comes from the spatial disposition of the building units. Thus, a
homochiral crystal can be generated from simple achiral building units that crystallize in a chiral space group,
or induced by the presence of chiral guests inside the achiral host.
172 Francesc X. Llabrs i Xamena

On the other hand, a rational control over the framework flexibility of the material can
also have a large impact on its final catalytic properties. Although framework flexibility has
been largely overlooked when designing catalytic applications of CPs, it can be expected that
these properties will be considered in future developments. Taking the enzymatic systems as
source of inspiration, the objective has to be the preparation of materials capable of adapting
the pore space by conformational changes of their building units.
Finally, the rational design of CPs affords a means for readily preparing multi-functional
catalysts, which allows performing one-pot procedures involving multiple catalytic events
with only one catalyst and avoiding unnecessary (and costly) steps of separation and
purification of intermediate products [18]. These transformations known as tandem, domino
or cascade reactions represent a process intensification, which allows decreasing the energy
consumption and generation of wastes. This improves the atom economy and lowers the E
factors (kg subproduct/kg product) of the overall process. Although the use of CPs as
enantioselective and multi-functional catalysts is still in its infancy, a fast evolution of the
field can be anticipated in the next years, which will surely demonstrate the brilliant potential
of these materials as advanced heterogeneous catalysts.
It is important to note that up to now research in CPs has been mainly focused on the
discovery of new materials with a variety of compositions and architectures. Concerning their
applications, the main promise so far has been in gas storage, separation and purification,
while application in heterogeneous catalysis is mainly limited to the last 6-8 years (with some
noticeable exceptions of reports appeared earlier). Heterogeneous catalysis based on CPs is
currently evolving from a first initial stage, in which the main objective was to show that
these materials could be used as catalysts. These first reports generally consisted in
demonstrating that a given CP contained the necessary catalytic centers to catalyze a certain
reaction. In many cases, the performance of the material was poor and many concerns existed
regarding the stability of certain materials under reaction conditions. The current challenge is
now to develop truly efficient catalytic process using CPs, ideally exploiting the many
advantages that these materials have: High surface area and pore volume, design and
flexibility of the pore system, potential for chiral (enantioselective) catalysis and development
of multifunctional catalysts. This apparently seems an intrusion into a field that nowadays is
governed by porous zeolites and zeotypes (which are today the most successful type of porous
solid catalysts). However, this chapter is not presented as a dichotomy between CPs and
zeolites, but will rather try to illustrate the possibilities of CPs to complement them. The
emphasis is made to present the different strategies to introduce catalytic functions into CPs.
Some relevant references to original papers are given throughout the chapter that can serve to
illustrate practical application of CPs. The reader should consult the recent reviews covering
the use of CPs as catalysts for organic reactions for a comprehensive coverage of the subject
[11-13].

1.2. Bridging the Gap between Homogeneous


and Heterogeneous Catalysis with CPs

Homogeneous and heterogeneous catalysis have traditionally been considered as different


disciplines, and indeed they have been usually studied separately. However, the tendency
nowadays is to try to design and develop new materials to bring together the best advantages
Coordination Polymers 173

of both disciplines in a sole catalytic system, while avoiding the limitations of conventional
homogeneous and heterogeneous catalysts.
Metals in solution, either as metal salts or transition metal complexes, are widely used to
effect chemical transformations in two completely distinct scenarios. One consists in the use
of stoichiometric amounts of the metal salts as oxidizing reagents (such as KMnO4, K2Cr2O7
or Pb(OAc)4), as well as Lewis or Brnsted acids or bases (e.g., AlCl3 or NaOH); while a
second approach consists in using more sophisticated, specifically designed transition metal
complexes as efficient and selective catalysts. The use of stoichiometric metal salts has many
associated problems, such as the generation of large amounts of waste, either directly
generated in the chemical process or formed during the necessary neutralization of acids and
bases. The current tendency is whenever possible to replace these compounds by more benign
heterogeneous catalysts, such as zeolites or clays. For instance, 4-methoxyacetophenone, an
important chemical intermediate, has been traditionally produced in the liquid phase, using
dichloromethane as solvent and AlCl3 as Lewis acid reagent. But this process produces up to
4.5 kg of inorganic salts per kg of final product. Nowadays, Rhodia produces this molecule in
a continuous process in fixed bed reactors and without any solvent, using zeolite beta as acid
catalyst, thus reducing considerably the amount of wastes generated [19].
Besides their use as stoichiometric reagents, metal coordination or organometallic
complexes have also been widely used as homogeneous catalysts, sometimes with
extraordinary success. They are able to catalyze a large number of organic reactions, in many
cases with extraordinary chemo, regio and enantioselectivities. The main advantage is the
high activity and selectivity that can be achieved with these compounds, which is primarily
due to the possibility to change systematically the organic ligands that coordinate to the
central metal ion. In this way, the electronic and steric properties of the molecule can be
finely modulated to obtain the desired catalytic properties. Additionally, the well defined
structure of these metal coordination complexes enables the study of their interaction with
substrates by means of powerful quantum chemical calculations, which allows a precise
knowledge on the reaction mechanism as well as to engage in predictive reactivity patterns.
Homogeneous catalysts, however, are most of the times difficult to be recovered from the
reaction medium and/or they decompose during the reaction. Replacement of soluble metal
coordination complexes by heterogeneous catalysts can overcome these limitations, and in
some cases, can even introduce additional valuable features, such as shape-selective
properties. For instance, well-defined active sites ranging from protons to Lewis acids, and
even redox sites, can be introduced in the frameworks of zeolites and mesoporous materials.
When combining these well-defined framework or extraframework active sites with the
regular pore dimensions and topologies typical of zeolites, that can be selective towards
different potential transition states (shape-selectivity), highly active, selective, stable and
recyclable catalysts can be achieved. Probably the most popular example of this class is
titanium silicalite-1 (TS-1) [20]. TS-1 in combination with hydrogen peroxide has allowed the
industrial implementation of environmentally friendly technologies, such as phenol
hydroxylation and cyclohexanone ammoximation [20]. Owing to the relatively small pore
diameter of TS-1 (5.5 ), this zeolite catalyst cannot be used for reacting bulky substrates.
Therefore, new titanium-containing large pore zeolites [21] and mesoporous materials [22]
have been prepared to replace TS-1 when dealing with large substrates. Another four-valent
metal, i.e., tin, was successfully incorporated into the framework of pure silica zeolite beta
and provided unique catalytic activity for the Baeyer-Villiger oxidation with hydrogen
174 Francesc X. Llabrs i Xamena

peroxide [23] and Meerwein-Ponndorf-Verley reductions [24] among others [25]. Later,
zirconium [26], nionium and tantalum [27] beta zeolites were also synthesized and
successfully used in catalysis.
But heterogeneous catalysts have also their own limitations, which limit their
performance and application. For instance, zeolites are suited solid catalysts for gas phase
reactions, but due to its limited available pore size, they generally undergo fast deactivation
for liquid phase reactions. In addition, most of the efforts to develop zeolites with larger pores
in the nanometer scale have met with failure [28-31]. Another important concern when using
zeolites and related materials as heterogeneous catalysts is their stability, and in particular,
leaching of the metal active species into solution upon continuous use of the catalyst. Finally,
another important limitation in heterogeneous catalysis is that fine modulation of the
electronic and coordination properties of the metal active sites of the solid is usually much
more difficult than in the case of the homogeneous complexes.
In an attempt to overcome those limitations of both homogeneous and heterogeneous
catalysts, researchers have developed preparative methods to combine the well-controlled
active sites of transition metal complexes with the adsorption and pore selectivity effects and
recyclability of solid catalysts. Heterogeneization has been achieved by grafting or
impregnation of the active metal coordination complexes on solid carriers, by intercalation
within layered compounds, by introducing them in zeolite cavities by s hip in a bottle
technique [32], or by forming structured or non-structured mesoporous organic-inorganic
hybrid systems [33, 34]. In some cases, the objective of heterogeneization was simply to
anchor the active species on a solid carrier to achieve well isolated, uniform single sites that
will not interact between them and decompose. However, in other cases the heterogeneization
went further in such a way that the solid also intervene in the catalytic process, either by
stabilizing transition states or by introducing additional active sites [35, 36]. Unfortunately,
these heterogeneization methods have no general application, since not all the homogeneous
catalysts can be easily and successfully anchored to suitable supports. Sometimes the
necessary steps for immobilization can severely increases the prize of the catalyst, making it
not competitive with respect to alternative non-catalytic processes.
In this context, porous crystalline CPs can come into scene, since these compounds can
be considered as complimenting an expanding the work of zeolites in heterogeneous catalysis,
owing to the outstanding properties described in the previous section. In particular, it is very
interesting the possibility to prepare a large variety of CPs with pore dimensions larger than
those of classical zeolites, with a range of compositions and chemical properties, and with
access to chiral catalysts. This could overcome the limitations of zeolites when used in liquid
phase reactions or with large organic substrates, since wide pore CPs could avoid diffusion
control of the reaction. On the other hand, the modular construction of CPs through the
assembly of metal and organic ligands by coordination bonds allows changing systematically
the electronic and steric properties of the central metal ions, much like in soluble metal
complexes. However, being solid materials CPs will be easily recovered from the reaction
medium by simple filtration, and this will allow recycling the material for further use. Thus,
when the catalysis is based on metal activity at the nodes or at the ligands forming the walls,
CPs appear directly as solid counterparts of homogeneous catalysts. Indeed, a CP can be
viewed as a spatial array of metal coordination complexes. Keeping this idea in mind can help
in designing new CP-based heterogeneous catalysts, as exemplified in Figure 2. Some real
Coordination Polymers 175

examples will be shown in section 3, when the different strategies for preparing CP catalysts
will be presented.

Figure 2. CPs can be viewed as a spatial array of metal coordination complexes, and this idea can be
used when designing a new catalyst. Imagine a soluble metal coordination complex, MX4, as shown in
a), with a well recognized catalytic activity. A CP can be designed containing the same metal
component, M, and a multidentate ligand containing the same (or related) functional groups, X, as
shown in b). The resulting structure of this CP could be represented as shown in c), where the MX 4
units are evidenced. If MX4 has certain catalytic properties, this activity could also be present in the CP.

It is evident that CPs will also present limitations that will prevent their use under certain
conditions, such as high temperatures or in the presence of certain solvents and reagents.
However, their rational design and the large versatility in the engineering of their structures
and compositions can certainly help bridging the gap between homogeneous and
heterogeneous catalysis [35].

2. SYNTHESIS OF CPS
A large variety of metal atoms, i.e., alkaline, alkaline-earth, transition metals, main group
metals and rare-earth elements, have been successfully used in the synthesis of CPs. As
organic components, rigid molecules (such as conjugated aromatic systems) are usually
preferred over flexible ones, since they generally favor the preparation of crystalline, porous,
stable materials. Common choices for the organic linkers are based on some of the
compounds shown in Figure 3, including among others, polycarboxylic aromatic molecules,
bipyridines, and polyazaheterocycles (imidazole, triazole, tetrazole, pyrimidine, pyrazine, tec)
and their derivatives. Both neutral and charged molecules can be used, although cationic
ligands are less common in the synthesis of CPs because of their lower affinity to coordinate
to metal cations.
176 Francesc X. Llabrs i Xamena

Figure 3. Some common organic linkers used for the preparation of CPs.

2.1. Synthesis Methods

Typical syntheses of CPs are usually carried out in the liquid phase, either in pure
solvents or in suitable mixtures of solvents. Formation of the crystalline framework takes
places by self-assembly of the structural units forming a network of metal-ligand coordination
bonds. The synthetic method generally consists in mixing solutions containing the metal and
the ligand, with or without the aid of additional auxiliary molecules. Variants of the syntheses
in liquid media include saturation, diffusion and solvothermal methods.
In the saturation method, crystals are formed from saturated solutions, which can be
achieved by slow evaporation of the mother liquor.
Diffusion methods are usually preferred when well formed single crystals are sought (eg,
single crystals suitable for X-ray characterization). The method consists in slowly bringing
into contact the components forming the CP, usually by carefully forming different layers of
liquids containing the reactants separated by an additional layer of pure solvent.
Finally, in solvothermal methods, the synthesis is carried out in closed recipients
(autoclaves) at temperatures usually comprised between 100 and 250C and under
autogenous pressure. This method, initially developed for the synthesis of zeolites, favors
crystal growth due to the reduced viscosity of the solvent under solvothermal conditions. This
method can be preferred when the different solubility of the organic and metallic components
in the mother liquor prevents the use of the two other methods.
Besides the synthesis methods described above, alternative preparation procedures have
recently appeared, including microwave synthesis, electrochemical, and mecanochemical
(solvent-free) methods. Although the use of these methods are generally less extended, in
many cases they may represent important improvements as compared to liquid-phase
conv entional methods, such as reduced synthesis time, improved quality of the materials
formed, easiness of large scale production, or reduced waste production.
Microwave heating reduces considerably the crystallization time as compared to
conventional heating, so that it is possible to prepare CPs in few hours. Usually CPs prepared
by this method are microcrystalline, which is probably related with a high nucleation rate.
Electrochemical synthesis was developed by BASF as an alternative means to produce
large scale amounts of CPs without using the enormous amounts of solvents required by
conventional liquid phase synthesis. Mueller et al. applied this method for the first time to the
preparation of a copper trimesate material [37]. The method consisted in using copper plates
as the anodes, immersed in a methanolic solution containing the linker (trimesic acid) and
copper cathodes. Upon applying an electric current, a greenish-blue solid precipitated, which
showed an X-ray diffraction pattern similar to that of another copper trimesate material
prepared by a solvothermal route and known as HKUST-1 [38]. Interestingly, the material
Coordination Polymers 177

prepared electrochemically presented a higher surface area and a more symmetric fourfold
coordination environment of the copper ions as compared to the material prepared by the
solvothermal method, which probably reflects the presence in the latter material of nitrate
impurities occluded inside the pores.
Mechanochemical synthesis methods do not require the use of solvents. Solid precursors
are directly mixed and ground together in a ball mill to form the desired material, followed by
a thermal treatment at a convenient temperature. In general, metal salt precursors that are
either hydrated or that release solvent upon reaction (such as acetic acid) are preferred for this
type of synthesis methods. Pichon et al. described for the first time the preparation of a copper
isonicotinate CP, [Cu(ina)2] (ina = isonicotinate), using a mechanochemical method and
copper acetate as the metallic precursor, [39]. A more detailed study by the same authors
appeared some years later [40].

2.2. Systematic Design of Pore Size

Probably one of the most striking features of CPs that has attracted much interest is the
possibility to construct materials with crystalline networks having channels and cavities with
tunable pore sizes [41, 42]. This has allowed preparing materials with pore sizes ranging from
ultramicropores to mesopores [43]. One possible strategy for tuning the pore dimensions is
the so-called isoreticular expansion. This conceptually consists in expanding the pore size of
a known structure by using expanded but geometrically analogous organic bridging ligands,
so that the new material will have the same framework topology as the original compound,
but with larger pores.

Figure 4. Representative examples of ex panded ligands used for the preparation of isoreticular
coordination polymers of increasing pore size. Within a given isoreticular series, the materials contain
the same inorganic building block and geometrically analogous organic bridging molecules of different
size. Some examples of the isoreticular materials prepared with the ligands are given in parentheses [42,
44-46]. ADC = acetylenedicarboxylate; BDC = benzenedicarboxylate; BPDC = biphenyldicarboxylate;
TPDC = triphenyldicarboxylate; BTC = benzenetricarboxylate; BTB = 4,44-benzene-1,3,5-triyl-
tribenzoate; BTE = 4,4,4-[benzene-1,3,5-triyl-tris(ethyne-2,1-diyl)]tribenzoate; BBC = 4,4,4-
[benzene-1,3,5-triyl-tris(benzene-4,1-diyl)]tribenzoate.
178 Francesc X. Llabrs i Xamena

The first example of the use of this strategy is the well known IRMOF isoreticular series
prepared by the group of Prof. Yaghi [42]. All the materials belonging to this series are
constructed by the same inorganic building block, Zn4O (analogous to that found in basic zinc
acetate), while the organic linker is selected among various linear dicarboxylic molecules of
different dimensions, such as in IRMOF-0, IRMOF-1, IRMOF-10 and IRMOF-16, which
contain alkyne, phenyl, biphenyl an triphenyl moieties, respectively. Similar series of
isoreticular compounds having increasing pore sizes were prepared by Prof. Lillerud et al.
containing hexameric Zr6O4(OH)4 clusters as the inorganic building block and linear
dicarboxylic acids, viz. UiO-66, UiO-67 and UiO-68 [44]. Besides linear dicarboxylic ligands,
other expand ed ligands has also been used in this context, as in the case of the tricarboxylic
molecules geometrically analogous to trimesic acid, leading to the isoreticular series of
materials knows as MOF-177, MOF-180 and MOF-200 [45]. Some representative examples
of expan ded ligands used to control systematically the pore size of the isoreticular CPs are
shown in Figure 4.

2.3. Structure Directing Effects

The large number of metal ions that can be incorporated in CPs, together with the
virtually infinite possibilities in the selection of the organic substrate, lead to a myriad of
possible chemical compositions for CPs. The number of possibilities increases considerably if
we take into account that it is also possible to prepare CPs containing two (or more) types of
metals as well as two (or more) different types of ligands, which greatly boosts the possible
attainable architectures and range of properties. Additionally, the number of different CP
structures that can be in principle prepared is dramatically increased by the extended
occurrence of polymorphism and other isomerism. Thus, a given metal-ligand combination
can lead to a number of different structures depending on the particular synthesis conditions
used for the preparation of the CP (temperature and time of synthesis, the nature of the
solvent, or the presence of substituents in the organic ligand, among other factors). A clear
example of the effect of these parameters can be found in the zinc-imidazole system. Simply
by changing the solvent used in the synthesis, Tian et al. were able to prepare up to seven
different zinc imidazolate frameworks [47], which joined the family of other already existing
zinc imidazolates [48, 49]. All these materials had the same general formula, [Zn(im)2xG]
(im = imidazolate, G = guest molecule, and x = 0.2-1), and the framework was constructed by
the same building units: Zn2+ ions coordinated to 4 N atoms from four different imidazolate
ring in a tetrahedral geometry (ZnN4), while each imidazolate molecules coordinates to two
Zn2+ ion through each of the N atoms (Zn-im-Zn). Thus, the solvent was found to act as a
structure-directing agent for the resulting zinc imidazolates.
The presence of substituents in the imidazole ring was also found to be important for
achieving the final crystalline structure, as evidenced by Chen and co-workers [47, 50-52].
The authors prepared [Zn(bzim)25/3 H2O], [Zn(2-mim)23H2O] and [Zn(2-eim)2H 2O] from
benzimidazole, 2-methylimidazole, and 2-ethylimidazole, respectively, as well as a mixed
ligand compound [Zn(2-eim/2-mim)2H 2O]. The presence of substituents in the imidazole
ring avoided the formation of dense phases and directed the topology of the resulting
material. Thus, [Zn(bzim)25/3 H2O] adopts a distorted zeolite-related sodalite (sod) topology
with only 18% accessible volume to guests. [Zn(2-mim)23H2O] features a regular sod
Coordination Polymers 179

topology, with 47% of free volume composed of spherical cavities with a diameter of 12.5
and accessible through hexagonal windows of ca. 3.3 , with an apparent specific BET
surface area of 1029 m2g-1. [Zn(2-eim)2H 2O] assumes an analcime (ana) zeolite topology,
while the structure of the mixed ligand compound [Zn(2-eim/2-mim)2H 2O] has a zeolitic rho
topology, with 57% free volume and truncated cuboctahedra of 18.1 with pore windows of
7.4 .
Several other examples are found in the literature in which a given binary metal-ligand
system can yield different structures, depending on the particular synthesis conditions. For
instance, the system zinc-terephthalate is known to produce a number of different structures,
including MOF-2 [53, 54], MOF-3 [53, 55], MOF-5 [53, 56], [Zn2(OH)2(BDC)22 DEF]
(DEF = diethyl formamide) [57], [Zn(H2O)2(-O,O-BDC] [58], zinc terephthalate hydrate
[59], sodium zinc terephthalate hydrate DMF solvate (DMF = dimethyl formamide) [60],
MOCP-H and MOCP-H [50].

3. DESIGNING CPS FOR CATALYTIC APPLICATIONS


Porous CPs contain three well differentiated parts: the metallic component, the organic
linker and the porous system. As we will see in this section, it is possible to prepare CPs in
which the catalytic function is contained in any of these three parts. In some cases and for
certain CPs, the as-prepared material can be used directly as a catalyst, since it already
contains the necessary active sites to catalyze the chemical reaction. This category includes
those materials in which the metal sites can directly coordinate to the substrates and catalyze
the reaction, as well as CPs in which the organic ligands contain functional groups that can
act as (organo) catalysts. But unfortunately, the CPs that can be directly used as catalysts in
the as-synthesized form represent only a tiny proportion of the whole family of CPs. In most
cases, it has been necessary to develop specific strategies to modify the material before they
can be used in a catalytic reaction. In the following, we will revise the main situations we can
find when facing the application of a CP in catalysis, depending on whether the catalytic
function is introduced at the metallic site, at the organic linker or inside the pore system.

3.1. Catalysis at the Metallic Site

3.1.1. As-Synthesized Active CPs


This is a limited group of materials that contain metal ions that can directly coordinate to
the substrates to catalyze the reaction. Coordination of the substrate to the metal requires
either an expansion of the coordination sphere of the metal ion, or a displacement of one of
the ligands originally forming the CP, as shown schematically in Figure 5. In either case, the
crystalline network of the CP has to be highly flexible to prevent the collapse of the structure
by the local distortions produced upon substrate coordination. We have recently found [61]
that copper imidazolate, [Cu(im)2] (im = imidazolate) [62], and copper pyrimidinolate, [Cu(2-
pymo)2] (2-pymo = 2-hydroxypyrimidinolate) [63], both feature highly flexible networks that
can readily accommodate changes in the coordination sphere of copper upon substrate
binding, passing from tetra- to pentacoordinated sites. This high flexibility allows
180 Francesc X. Llabrs i Xamena

successfully using these two materials as catalysts for various reactions while preserving the
structure integrity [64-66].

Figure 5. Interaction of a substrate molecule, S, with a metal site, M, through a) expansion of the
coordination sphere around the metal ion; or b) displacement of one of the ligands.

3.1.2. CPs Containing Metal Nodes with Semiconducting Properties


This group is formed by CPs featuring metal oxide nanoclusters at their nodes, and whose
electronic configuration correspond to that of a semiconductor, i.e., with a band structure
containing a completely filled valence band and an empty conduction band separated by a
relatively small energy gap, Eg. It has been demonstrated that the materials that fulfill these
requirements can actually have semiconducting properties [67, 68]. Upon convenient optical,
electronic or thermal excitation, electrons can be excited from the valence to the conduction
band, thus generating an electron-hole pair. Examples of these materials are CPs containing
tetranuclear Zn4O or hexanuclear Zn6O4(OH)4 clusters. We have recently shown that both, the
electrons in the conduction band and the holes in the valence band, are long lived species that
decay to the ground state in the microsecond time scale [67]. Therefore, the lifetime of the
charge separated state is long enough to allow it to interact with suitable electron donors or
acceptors that can be present in the medium. We and others have demonstrated the utility of
this type of compounds as photocatalysts [67-70], i.e., the energy needed to create charge-
separated states was provided by UV light absorption (photons with energy higher than the
band gap of the material).

3.1.3. CPs with Coordinatively Unsaturated Metal Sites


It is possible to prepare CPs in which one of the coordination positions of the metal
centers is occupied by a labile ligand, which can be removed without causing the collapse of
the crystalline structure. In most cases, the labile ligands are solvent molecules that, when
Coordination Polymers 181

thermically removed, leave a free coordination position in the metal, which become available
for adsorbed substrates. A relevant example of this type of CPs corresponds to the copper
trimesate, [Cu3(BTC)2] (BTC = benzene tricarboxylate) [38], in which the copper sites are
known to lose reversibly their coordinated apical water molecule upon thermal activation
under vacuum [71, 72], thus leaving an accessible coordination vacancy on the copper ion, as
shown in Figure 6.

Figure 6. Schematic representation of the Cu2 dimmers in the copper trimesate [Cu3(BTC)2]. Each
copper ion bears a water molecule, that can be removed under vacuum to leave a coordination vacancy.

3.1.4. Anchoring of Catalytic Active Species to the Metal Nodes


Coordination vacancies on metal ions generated by displacement of labile ligands can be
used as anchoring points for introducing additional functionalities.

Figure 7. Summary of the main types of CPs that can be prepared in which the catalytic active site is
located at the metallic component.
182 Francesc X. Llabrs i Xamena

The first example of the use of this strategy was recently reported by Hwang et al. [73]
They prepared the chromium terephthalate MIL-101, which containg Cr3+ sites in which one
coordination position is occupied by a water molecule. After removing this water molecule by
a thermal treatment under vacuum, the resulting vacancy was used for grafting
ethylenediamine, in such a way that one of the N atoms coordinated to the Cr3+ ion while the
second N remained free and pointing towards the center of the pores. The authors also
demonstrated that these amino groups can be used as catalytic basic sites [73]. Similarly,
Banerjee et al. used the Cr3+ ions of MIL-101 as anchoring sites for introducing a proline
derivative containing a pyridine group as anchoring point [74].

3.2. Catalysis at the Organic Linker

3.2.1. CPs with Organic Functional Groups


These are materials containing functional groups at the organic component that can
catalyze a chemical reaction; i.e., they are organocatalysts. The catalytic activity is located at
the organic linker and not at the metal. It is evident that the organic linkers used to prepare
these CPs must contain two different types or organic functional groups, as shown in Figure
8: coordinative groups, L1, which are coordinated to the metal ions and are required to
construct the crystalline network; and reactive groups, L2, which are free (not coordinated to
any metal ion) and will be responsible for the catalytic properties of the material.The most
well known example of this type of CPs is the zinc aminoterephthalate, IRMOF-3. The
organic ligand contains two carboxylate coordinative groups (L1) and one amino reactive
group (L2), which has been claimed as an efficient base catalytic site [75, 76]. Another
example of ligand containing two types of functional groups that has been used for preparing
a CP is 1,3,5-benzene tricarboxylic acid tris[N-(4-pyridyl)amide], 4-BTAPA [77]. This
molecule coordinates to the metal ions through the pyridyl N atoms, while the amide groups
remain free and can be used as a basic functionality.

Figure 8. General structure and selected examples of ligands containing coordinative and reactive
functional groups.
Coordination Polymers 183

3.2.2. CPs Containing Metal Coordination Complexes as Building Blocks


In the vast majority of the existing CPs, the metal sites have no available coordination
positions (as in 3.1.1), nor they can be created by removal of labile ligand molecules (as in
3.1.3). On the contrary, the metal sites are completely blocked by tightly coordinative linkers
forming the crystalline network. A possible alternative that can be used to introduce metal
active sites in a CP is by binding the metal ion to a suitable organic molecule to form a metal
coordination complex. Then, this metal coordination complex is used as a linker to form the
CP. Therefore, these CPs contain two types of metal ions, as shown in Figure 9: one of them
(M1), which belongs to the metal coordination complex, is responsible for the catalytic
activity of the CP, while a second type of metal ions, M2, has only a structural role and is not
directly involved in catalysis.

Figure 9. Some CPs contain two types of metal ions: catalytic active sites (M1) and metal ions with a
mere structural role, M2. Selected examples of metal coordination complexes used as building blocks
for constructing CPs are shown.

Kitagawa and co-workers [78, 79] prepared a CP containing Cu(2,4-pydca)2 complexes


(2,4-pydca = pyridine-2,4-dicarboxylate) as metalloligands, which were coordinated to Zn2+
cations through one of the carboxylate groups to form a 3D structure. In this material, Zn2+
acted as a mere structural element, while Cu2+ ions were accessible for guest coordination.
Similarly, the same group [80] has prepared a series of materials containing metal Schiff base
complexes, M(H2salphdc) (M = Cu2+, Ni2+ or Co2+, salphdc = N,N-
phenylenebis(salicylideneimine)dicarboxylate), with Zn2+ cations at the nodes. Lin and co-
workers [81] have prepared a homochiral CP containing Cd2+ ions and the chiral ligand (R)-
6,6-dichloro-2,2-dihydroxy-1,1-binaphthyl-4,4-bipyridine as the organic building unit.
The ligand coordinates to Cd2+ through chlorine and the pyridine nitrogen, while the two
hydroxyl groups of the binaphthyl moiety remain uncoordinated and pointing to the channels.
184 Francesc X. Llabrs i Xamena

Post-synthesis modification of this material by adding titanium isopropoxide yielded a


titanium containing material, with titanium di-isopropoxide grafted to the walls of the MOF
through the dihydroxy groups. Szeto et al. prepared bimetallic materials containing Gd [82] or
Yb [83] ions at the nodes, and Pt2+ ions four-coordinated by two Cl and by two N atoms of
2,2-bipyridine-4,4-dicarboxylate (bpydc), which could act as potential catalytic sites. These
are some of the examples of materials in which a metalloligand is used as linker to form the
crystalline network of the CP.

3.2.3. CPs with Chelating Linkers


When a CP is formed by an organic linker that contains some functional group that is not
coordinated to the metal site (such as those described in 2.42.a), the additional functional
group can be used to coordinate to a second metal ion, as shown in Figure 10. In this way, it
is possible to introduce an additional metal site, which might not be possible to put directly in
the as-synthesized material. This can be seen as a combination of the two situations shown in
Figure 8 and Figure 9. Thus, it is also possible to prepare a CP in which the organic linker
contains two functional groups: one of them (L1) coordinates to a metal ions having a
structural role (M1), while the second type of functional groups (L2) serves as a chelating
agent to introduce a second metal ion (M2) with catalytic properties. It is also possible to
modify these functional groups L2 through organic reactions, in order to increase their
chelating properties, as it will be shown in section 3.4.

Figure 10. General procedure for the preparation of CPs with chelating groups to introduce a second
metal ion, and representative illustration of this method applied to prepare the material known as
IRMOF-3-SI-Au.
Coordination Polymers 185

Following this strategy, we described the preparation of the material known as IRMOF-3-
SI-Au [84]. It consisted of a zinc aminoterephthalate IRMOF-3, in which the amino groups
were reacted with salicylaldehyde to form the corresponding Schiff base, which was used to
chelate Au3+ ions. Thus, this material contained Zn2+ and Au3+ ions as the structural and
catalytic metal ions, respectively. The 3D structure was formed by an extended network of
Zn2+-carboxylate coordination bonds, while the Schiff base salicylidene imine acted as a
chelating ligand to coordinate and stabilize the catalytic Au3+ ions. A completely analogous
strategy was used by Ingleson et al. to introduce vanadyl acetylacetonate [85].

Figure 11. Summary of the main types of CPs that can be prepared in which the catalytic active site is
located at the organic linker.

3.3. CPs as Host Matrices or Nanometric Reaction Cavities

In these materials, none of the components forming the CP is directly involved in


catalysis. The pore system of the material either provides the physical space where a chemical
reaction takes places (nanometric reaction cavity), or it serves as a cage for encapsulating the
catalytic species (host matrices).

3.3.1. Encapsulated Metal Nanoparticles


Different metal nanoparticles have been successfully prepared inside the pore structure of
a number of CPs. Among the preparation procedures used are chemical vapor deposition of
suitable precursors, incipient wetness impregnation, co-precipitation and even simple
grinding, usually followed by in situ reduction of the precursors with H2 or NaBH4 to form
the metallic nanoparticles. In this way, small metallic nanoparticles of Pd [86-89], Cu [86],
Ru [90], Ag [91] or Au [92-94] have been incorporated inside the pores of different CPs.
What is relevant for this chapter is that, the encapsulated nanoparticles have shown to be
186 Francesc X. Llabrs i Xamena

catalytically active in some cases and for certain reactions, including olefin hydrogenation,
alcohol oxidation or C-C coupling reactions.

3.3.2. Encapsulated Metal Oxide Nanoparticles


Metal oxide nanoparticles have also been prepared inside the pore system of CPs. Among
the different composite materials thus prepared are SiO2 [95], TiO2 [96] and ZnO [97].
Common preparative techniques consist in the adsorption of suitable precursors
(tetramethoxysylane, titanium isopropoxide or diethyl zinc), followed by thermal annealing to
produce hydrolysis and condensation. Considering the interest in the preparation of
encapsulated semiconducting metal oxide clusters as a way to increase their photocatalytic
activity [98-101], it can be anticipated that CPs could also be appropriate matrices to include
metal oxide nanoparticles inside the pores.

3.3.3. Encapsulated Catalytic Active Molecular Species


The micropore system of CPs holds a considerable promise since it can be used to
encapsulate guests that can exhibit some catalytic activity. This strategy to transform a
homogeneous catalyst into a heterogeneous system has been widely explored in the case of
zeolites [102]. Depending on the size of the guest to be encapsulated with respect to the pore
openings and cavities of the CP, different preparative methods can be used. These include the
incorporation of the guest during the synthesis of the CP, incorporation by
adsorbing/impregnating from the liquid or the gas phase, or by the so-called ship-in-a-bottle
synthesis. In these processes, small precursors that can diffuse through the micropore system
of the host material react to form species that can be accommodated inside the cages but they
are too large to diffuse through the small window.
Following different preparation strategies, a number of molecular guests have been
successfully incorporated inside the pores of CPs, such as metalloporphyrin derivatives [103],
polyoxometallates (POMs) [104-107], or metal phthalocyanine complexes [108].

3.3.4. CPs as Nanometric Reaction Cavities


In this case, the pore system of a CP is the physical place in which a chemical reaction
takes place. Although none of the components of the solid participates in the reaction, the
properties inside the channel system in terms of polarity, hydrogen bond interactions,
hydrophilicity/hydrophobicity, viscosity, etc. may have an influence on the final products
formed. This is more so when the molecular size of a substrate or product of the reaction has
similar dimensions than the host matrix in which it is contained, since confinement effects
can become important. One example of performing a reaction inside the MOFs pores is the
styrene polymerization [109]. Radical polymerization of styrene was carried out inside CPs of
general formula [M2(bdc)2(teda)] (M: Zn2+ or Cu2+, teda = triethylenediamine), and no
difference due to the nature of the Zn2+ or Cu2+ metal was observed. Quantitative recovery of
the resulting polystyrene accommodated inside the CP was performed by dissolving the host
matrix with 0.1 N NaOH. It was observed that the resulting polymers exhibit an average
molecular weight, M w , around 56000 with a remarkably low polydispersity (1.66).
Analogous polymerization procedure in the absence of CP leads to a polydispersity of 4.68.
Moreover, EPR spectroscopy during the polymerization of styrene inside the CP showed an
intense signal assigned to the propagating radical. The signal did not disappear even after
Coordination Polymers 187

storing the sample for one week at 70oC. This remarkable fact of having a living radical
inside the CP contrasts with the EPR spectra in homogeneous phase and was attributed to the
suppression of termination reaction and radical transfer inside the CP channels. A detailed
study by the same group has specifically addressed the chain conformation and dynamics of
single polystyrene chains hosted inside the channels of monodirectional [Zn2(bdc)2(teda)]
[110]. The channel dimensions of this CP (0.750.75 nm2) do not allow accommodation of
more than one polystyrene chain (0.440.68 nm2) and, therefore, it was assumed that the
polymer chains were independent and surrounded by the CP framework.

Figure 12. Product distribution in the photolysis of o-methyl dibenzyl ketone when the photolysis is
carried out inside the pores of [Co3(4,4-BPhDC)3(4,4-BPY)].

Figure 13. Summary of the main uses of porous CPs as host matrices or nanometric reaction cavities in
which a chemical reaction takes place.

Photochemical reactions have also been conducted inside the pores of a CP. One example
is the study of the photochemistry of o-methyl dibenzyl ketone inside the pores of the
188 Francesc X. Llabrs i Xamena

material [Co3(4,4-BPhDC)3(4,4-BPY)] (BPhDC = biphenyl dicarboxylate; 4,4-BPY = 4,4-


bipyridine) [111]. Asymmetric dibenzyl ketones can undergo homolytic cleavage in the -
position of the CO to give two differently substituted benzyl radicals that can recombine to
give three possible diarylethanes. In solution the product distribution arises from the random
coupling of the two benzyl radicals, giving a 25, 50, 25 % of the three possible A-A, A-B and
B-B diarylethanes. When diffusion is restricted the product distribution changes, favoring the
asymmetric A-B diarylethane arising from recombination of the geminate radical pair.
Following this, it was observed that when performing the photolysis of o-dibenzyl ketone
inside [Co3(4,4-BPhDC)3(4,4-BPY)], asymmetric o-tolyl phenyl ethane was formed with a
60 % yield and 100 % selectivity accompanied with a 40 % of intramolecular hydrogen
abstraction, as shown in Figure 12 [111].
The lack of formation of ditolyl (A-A) and diphenyl ethane (B-B) indicates that
[Co3(4,4-BPhDC)3(4,4-BPY)] as host exerts 100 % cage effect allowing exclusively
recombination of the geminate benzyl radicals. To put this value into context, photolysis of
the same compound in NaX gives 78 % cage effect [112].

3.4. Post-Synthesis Modification of CPs

We have seen that it is possible to use different strategies to design a CP to introduce


suitable catalytic active species. In some cases, the as-prepared material can be used directly,
or after simply removing a labile ligand to create a coordination vacancy. In other cases, the
preparation of CP catalysts involves a more complicated synthesis effort, sometimes requiring
4
the use of pre-formed metalloligands or exot ic organic molecules. When all these
strategies cannot be used, we have seen that the pores of CPs can still be used to encapsulate
active guests or as confinement spaces in which a chemical reaction takes place.
It is usually found in the preparation of heterogeneous catalysts in general, and of CPs in
particular, that the as-prepared material still does not contain the desired chemical
functionality that will be responsible for the catalytic activity. Thus, post-synthesis treatments
are required to modify the properties of the solid and to introduce (or activate) the catalytic
sites. These post-synthesis treatments can be of very different types, depending on the type of
material and the reaction that has to be catalyzed. In the case of zeolites, zeotypes, and other
metal or metal oxide catalysts, post-synthesis treatments consisting of thermal activation,
ionic exchange, steaming or hydrogenation are routinely performed to pre-activate the
catalytic centers. In the case of CPs, post-synthesis modifications are also applied, and we
have already described some of them in the previous section.
Post-synthesis modification of CPs has recently emerged as a highly versatile tool to
prepare tailored materials for applications in catalysis, gas adsorption, etc. Chemical
functionalization of the framework can be accomplished by either introducing covalent
attachment to the organic linker, or by grafting of organic molecules at metal sites
coordinative vacancies created after solvent elimination.

4 It is not uncommon to see CPs containing organic linkers that are not commercially available and need to be
synthesized, sometimes requiring complicated reaction schemes with many steps. This reduces considerably
the overall yield and, consequently, the amount of CP that can be produced is very small (many times only a
couple of single crystals are obtained). In these cases, catalytic studies become very complicated or even
impossible to carry out.
Coordination Polymers 189

Removal of labile ligand molecules upon thermal outgassing of the CP to create a


coordination vacancy can be viewed as the simplest post-synthesis modification, in the broad
sense of the word. This operation is needed to get access to the Lewis acid character of the
metal sites, i.e., its ability to coordinate to substrates acting as electron donors to form acid-
base adducts. In fact, there are a number of reports describing the use of CPs with
coordinative unsaturated metal sites as Lewis acid [71, 89, 113-115] or oxidation [116-119]
catalysts. But creation of coordination vacancies at the metal ions can be further used to
introduce a new functionality that was not present (nor latent) in the as-synthesized material,
as described in section 3.1.4. In this way, coordination vacancies were used by Hwang et al.
[73], and later by Demessence et al. [120] and by Banerjee et al. [74], to introduce amino or
proline groups anchored to the metal sites. This is another type of post-synthesis modification
which we could refer to as functionalization at the metal nodes.
A rare example that implies post-synthesis modification of the metallic nodes was
reported by Das et al. [121]. The treatment consisted in replacing the metal ions at the nodes
of the CP by other suitable elements (isomorphous substitution). These authors described a
metathesis exchange process leading to the complete and fully reversible substitution of
framework Cd2+ ions by Pb2+ in a MOF single crystal, without altering its crystalline
integrity. Interestingly, the authors were also able to exchange Cd2+ ions by trivalent
lanthanide cations Dy3+ or Nd3+, which resulted in positively charged frameworks. This
charge excess was compensated by extraframework NO3- anions. Although this particular
case of metathesis process has not an immediate application for catalysis, it demonstrates the
feasibility of tuning the reactivity of a CP by post-synthesis selecting the identity and
concentration of the metal ions forming the framework nodes. This could have interesting
implications for the preparation of acid and oxidation catalysts.
A highly versatile and unique property of CPs is the possibility to modify covalently the
organic linkers forming the material, which most noticeable can be done after the synthesis of
the material. This is the basis of the covalent post-synthesis modification (CPM), which has
attracted much interest in recent years [122]. This method relies on the presence in the
organic linkers of the CP of functional groups that do not participate in the metal coordination
bond. The idea of CPM comes from the realization that these groups of the solid can be
covalently modified through conventional reactions typical of Organic Chemistry, exactly in
the same way as we would do with soluble molecules bearing the same functional group. For
instance, when incorporated to a CP the amino groups of aminoterphthalate show a reactivity
pattern analogous to aniline in solution. Thus, they can be reacted with many different
functional groups, as shown in Figure 14. This provides a highly flexible means for
introducing new properties to the materials and for modifying the chemical environment
inside the pores. Some examples of the modifications reported so far include the reaction of
these amino groups with alkyl anhydrides, isocyanates or carboxylic acids, to name only a
few.
Post-synthesis modifications of CPs have been applied very recently in different
scenarios. Goto et al. [123] used an azide functionalized ligand that was post-synthesis
reacted with organic molecules bearing a terminal alkyne (through a click reaction using
CuBr as the catalyst). In this way, the authors have introduced hanging groups with various
functionalities: ester, alcohol and alkyl chain. Farrusseng and co-workers have also developed
a generic route for the preparation of cl ickable CPs [124]. The method started also from
compounds containing an amino group at the linker, which were in situ transformed into the
190 Francesc X. Llabrs i Xamena

corresponding azide and further reacted with a molecule bearing the desired functionality
together with a terminal alkyne group.

Figure 14. Some examples of the possible ways of covalent post-synthesis modification of a CP
containing accessible amino groups in the organic linkers.

Ingleson et al. [125] have described the preparation of a material containing Brnsted
acid sites by post-synthesis protonation of the carboxylate ligands of a CP with anhydrous
HCl. The same authors also introduced an interesting modification of the amino groups of
IRMOF-3 with salicylaldehyde to form the corresponding imine [85]. In this way, the poorly
coordinating amino groups were transformed into a very good Schiff base ligand for metal
coordination. This ligand was used to complex metal ions, as demonstrated by covalently
anchoring vanadyl acetylacetonate. Practically by the same time, we also used a similar post-
synthesis strategy to prepare a metal-coordination complex covalently anchored to the organic
linkers of a preexisting CP, as shown in Figure 10. First, a covalent functionalization of the
available amino groups of IRMOF-3 was carried out with salicylaldehyde to form the
salicylideneimine and, in a second step, Au(III) sites were coordinated to the Schiff base
complex [84]. As a last example to illustrate the enormous potential of CPM, Farrusseng and
co-workers have recently described a post-synthesis modification consisting in the
introduction of a long (C12) alkyl amine in a pre-existing CP containing aldehyde
functionalities [126]. In this way, the authors have been able to modify the chemical
environment inside the pores of the material, by increasing considerable the hydrophobic
character of the material.

CONCLUSION
Throughout this chapter, we have shown the enormous potential that Coordination
Polymers may have concerning their application as heterogeneous catalysts. Different ways
have been outlined for introducing catalytic active sites at the metallic nodes, at the organic
Coordination Polymers 191

linkers or inside the pores of the CPs, as well as the possibility to use CPs as the physical
space where a chemical reaction takes place. The potential of using post-synthesis
modification methods for modifying the properties of pre-formed CPs, and specially the
covalent modification of the organic linkers, has also been reviewed and illustrated through
selected examples.
I hope that this brief chapter could serve as an introductory reading for those who are
approaching the field of CP in heterogeneous catalysis for the first time, since the main
objective has been to provide a general and broad vision of the current state of the art.

REFERENCES
[1] Rowsell, J.L.C.; Yaghi, O.M. Microporous Mesoporous Mater., 2004, 73, 3-14.
[2] Farha, O.K.; Yazaydin, A.O.; Eryazici, I.; Malliakas, C.D.; Hauser, B.G.; Kanatzidis,
M.G.; Nguyen, S.T.; Snurr, R.Q.; Hupp, J.T. Nature Chem., 2010, 2, 944-948.
[3] Rosi, N.L.; Eckert, J.; Eddaoudi, M.; Vodak, D.T.; Kim, J.; O'Keeffe, M.; Yaghi, O.M.
Science, 2003, 300, 1127-1129.
[4] Ferey, G.; Mellot-Draznieks, C.; Serre, C.; Millange, F.; Dutour, J.; Surble, S.;
Margiolaki, I. Science, 2005, 309, 2040-2042.
[5] Rowsell, J.L.C.; Yaghi, O.M. Angew. Chem., Int. Ed., 2005, 44, 4670-4679.
[6] Latroche, M.; Surble, S.; Serre, C.; Mellot-Draznieks, C.; Llewellyn, P.L.; Lee, J.H.;
Chang, J.S.; Jhung, S.H.; Ferey, G. Angew. Chem., Int. Ed., 2006, 45, 8227-8231.
[7] Kitagawa, S.; Kondo, M. Bull. Chem. Soc. Jpn., 1998, 71, 1739-1753.
[8] Kitagawa, S.; Kitaura, R.; Noro, S. Angew. Chem., Int. Ed., 2004, 43, 2334-2375.
[9] Maspoch, D.; Ruiz-Molina, D.; Wurst, K.; Domingo, N.; Cavallini, M.; Biscarini, F.;
Tejada, J.; Rovira, C.; Veciana, J. Nature Mater., 2003, 2, 190-195.
[10] Mellot-Draznieks, C.; Serre, C.; Surble, S.; Audebrand, N.; Ferey, G. J. Am. Chem.
Soc., 2005, 127, 16273-16278.
[11] Lee, J.Y.; Farha, O.K.; Roberts, J.; Scheidt, K.A.; Nguyen, S.T.; Hupp, J.T. Chem. Soc.
Rev., 2009, 38, 1450-1459.
[12] Corma, A.; Garcia, H.; Llabrs i Xamena, F.X. Chem. Rev., 2010, 110, 4606-4655.
[13] Farrusseng, D.; Aguado, S.; Pinel, C. Angew. Chem., Int. Ed., 2009, 48, 7502-7513.
[14] Marquez, F.; Garcia, H.; Palomares, E.; Fernandez, L.; Corma, A. J. Am. Chem. Soc.,
2000, 122, 6520-6521.
[15] Marquez, F.; Zicovich-Wilson, C.M.; Corma, A.; Palomares, E.; Garcia, H. J. Phys.
Chem. B, 2001, 105, 9973-9979.
[16] Kesanli, B.; Lin, W. Coord. Chem. Rev., 2003, 246, 305-326.
[17] Lin, W. J. Solid State Chem., 2005, 178, 2486-2490.
[18] Climent, M.J.; Corma, A.; Iborra, S. Chem. Rev., 2011, 111, 1072-1133.
[19] Spagnol, M.; Gilbert, L.; Guillot, H.; Tirel, P.J. 1997. WO Patent.
[20] Perego, C.; Carati, A.; Ingallina, P.; Mantegazza, M.A.; Bellussi, G. Appl. Catal. A:
Gen., 2001, 221, 63-72.
[21] Sheldon, R.A.; Wallau, M.; Arends, I.W.C.E.; Schuchardt, U. Acc. Chem. Res., 1998,
31, 485-493.
192 Francesc X. Llabrs i Xamena

[22] Corma, A.; Navarro, M.T.; Perez Pariente, J. J. Chem. Soc., Chem. Commun., 1994,
147-148.
[23] Corma, A.; Nemeth, L.T.; Renz, M.; Valencia, S. Nature, 2001, 412, 423-425.
[24] Corma, A.; Domine, M.E.; Valencia, S. J. Catal., 2004, 215, 294-304.
[25] Corma, A.; Renz, M. Chem. Commun., 2004, 550-551.
[26] Zhu, Y.; Chuah, G.; Jaenicke, S. J. Catal., 2004, 227, 1-10.
[27] Corma, A.; Llabrs i Xamena, F.X.; Prestipino, C.; Renz, M.; Valencia, S. J. Phys.
Chem. C, 2009, 113, 1130611315.
[28] Corma, A.; Diaz-Cabanas, M.J.; Jorda, J.L.; Martinez, C.; Moliner, M. Nature, 2006,
443, 842-845.
[29] Estermann, M.; McCusker, L.B.; Baerlocher, C.; Merrouche, A.; Kessler, H. Nature,
1991, 352, 320-323.
[30] Strohmaier, K.G.; Vaughan, D.E.W. J. Am. Chem. Soc., 2003, 125, 16035-16039.
[31] Sun, J.L.; Bonneau, C.; Cantin, A.; Corma, A.; Diaz-Cabanas, M.J.; Moliner, M.;
Zhang, D.L.; Li, M.R.; Zou, X.D. Nature, 2009, 458, 1154-U1190.
[32] Alvaro, M.; Carbonell, E.; Espl, M.; Garcia, H. Appl. Catal. B: Environm., 2005, 57,
37-42.
[33] Diaz, U.; Garcia, T.; Velty, A.; Corma, A. J. Mater. Chem., 2009, 19, 5970-5979.
[34] Diaz, U.; Vidal-Moya, J.A.; Corma, A. Microporous Mesoporous Mater., 2006, 93,
180-189.
[35] Corma, A. Catal. Rev. Sci. Eng., 2004, 46, 369-417.
[36] Comas-Vives, A.; Gonzalez-Arellano, C.; Boronat, M.; Corma, A.; Iglesias, M.;
Sanchez, F.; Ujaque, G. J. Catal., 2008, 254, 226-237.
[37] Mueller, U.; Schubert, M.; Teich, F.; Puetter, H.; Schierle-Arndt, K.; Pastr, J. J. Mater.
Chem., 2006, 16, 626-636.
[38] Chui, S.S.Y.; Lo, S.M.F.; Charmant, J.P.H.; Orpen, A.G.; Williams, I.D. Science, 1999,
283, 1148-1150.
[39] Pichon, A.; Lazuen-Garay, A.; James, S.L. Cryst. Eng. Comm., 2006, 8, 211-214.
[40] Pichon, A.; James, S.L. Cryst. Eng. Comm., 2008, 10, 1839-1847.
[41] Rosi, N.L.; Eddaoudi, M.; Kim, J.; O'Keeffe, M.; Yaghi, O.M. Cryst. Eng. Comm.,
2002, 4, 401-404.
[42] Eddaoudi, M.; Kim, J.; Rosi, N.; Vodak, D.; Wachter, J.; O'Keeffe, M.; Yaghi, O.M.
Science, 2002, 295, 469-472.
[43] Wang, X.S.; Ma, S.Q.; Sun, D.F.; Parkin, S.; Zhou, H.C. J. Am. Chem. Soc., 2006, 128,
16474-16475.
[44] Cavka, J.H.; Jakobsen, S.; Olsbye, U.; Guillou, N.; Lamberti, C.; Bordiga, S.; Lillerud,
K.P. J. Am. Chem. Soc., 2008, 130, 13850-13851.
[45] Furukawa, H.; Ko, N.; Go, Y.B.; N., A.; Choi, S.B.; Choi, E.; Yazaydin, A.O.; Snurr,
R.Q.; O'Keeffe, M.; Kim, J.; Yaghi, O.M. Science, 2010, 329, 424-428.
[46] Tranchemontagne, D.J.; Hunt, J.R.; Yaghi, O.M. Tetrahedron, 2008, 64, 8553-8557.
[47] Tian, Y.Q.; Zhao, Y.M.; Chen, Z.X.; Zhang, G.N.; Weng, L.H.; Zhao, D.Y. Chem. Eur.
J., 2007, 13, 4146-4154.
[48] Lehnert, R.; Seel, F. Zeitschrift Fur Anorganische Und Allgemeine Chemie, 1980, 464,
187-194.
[49] Park, K.S.; Ni, Z.; Cote, A.P.; Choi, J.Y.; Huang, R.D.; Uribe-Romo, F.J.; Chae, H.K.;
O'Keeffe, M.; Yaghi, O.M. Proc. Natl. Acad. Sci. , 2006, 103, 10186-10191.
Coordination Polymers 193

[50] Huang, L.M.; Wang, H.T.; Chen, J.X.; Wang, Z.B.; Sun, J.Y.; Zhao, D.Y.; Yan, Y.S.
Microporous Mesoporous Mater., 2003, 58, 105-114.
[51] Huang, X.C.; Lin, Y.Y.; Zhang, J.P.; Chen, X.M. Angew. Chem., Int. Ed., 2006, 45,
1557-1559.
[52] Zhang, J.P.; Chen, X.M. Chem. Commun., 2006, 1689-1699.
[53] Eddaoudi, M.; Li, H.L.; Yaghi, O.M. J. Am. Chem. Soc., 2000, 122, 1391-1397.
[54] Li, H.; Eddaoudi, M.; Groy, T.L.; Yaghi, O.M. J. Am. Chem. Soc., 1998, 120, 8571-
8572.
[55] Li, H.L.; Davis, C.E.; Groy, T.L.; Kelley, D.G.; Yaghi, O.M. J. Am. Chem. Soc., 1998,
120, 2186-2187.
[56] Li, H.; Eddaoudi, M.; O'Keeffe, M.; Yaghi, O.M. Nature, 1999, 402, 276-279.
[57] Loiseau, T.; Muguerra, H.; Ferey, G.; Haouas, M.; Taulelle, F. J. Solid State Chem.,
2005, 178, 621-628.
[58] Guilera, G.; Steed, J.W. Chem. Commun., 1999, 1563-1564.
[59] Yang, S.Y.; Long, L.S.; Huang, R.B.; Zheng, L.S. Main Group Metal Chem., 2002, 25,
329.
[60] Yang, S.Y.; Sun, Z.G.; Long, L.S.; Huang, R.B.; Zheng, L.S. Main Group Metal
Chem., 2002, 25, 579-580.
[61] Luz, I.; Llabrs i Xamena, F.X.; Boronat, M.; Corma, A. submitted for publication.
[62] Masciocchi, N.; Bruni, S.; Cariati, E.; Cariati, F.; Galli, S.; Sironi, A. Inorg. Chem.,
2001, 40, 5897-5905.
[63] Tabares, L.C.; Navarro, J.A.R.; Salas, J.M. J. Am. Chem. Soc., 2001, 123, 383-387.
[64] Llabrs i Xamena, F.X.; Casanova, O.; Galiasso Tailleur, R.; Garcia, H.; Corma, A. J.
Catal., 2008, 255, 220-227.
[65] Luz, I.; Llabrs i Xamena, F.X.; Corma, A. submitted for publication.
[66] Luz, I.; Llabrs i Xamena, F.X.; Corma, A. J. Catal., 2010, 276, 134-140.
[67] Alvaro, M.; Carbonell, E.; Ferrer, B.; Llabrs i Xamena, F.X.; Garcia, H. Chem. Eur. J.,
2007, 13, 5106-5112.
[68] Tachikawa, T.; Choi, J.R.; Fujitsuka, M.; Majima, T. J. Phys. Chem. C, 2008, 112,
14090-14101.
[69] Gomes Silva, C.; Luz, I.; Llabrs i Xamena, F.X.; Corma, A.; Garcia, H. Chem. Eur. J.,
2010, 16, 11133-11138.
[70] Gascon, J.; Hernandez-Alonso, M.D.; Almeida, A.R.; van Klink, G.P.M.; Kapteijn, F.;
Mul, G. Chem. Sus. Chem., 2008, 1, 981-983.
[71] Alaerts, L.; Seguin, E.; Poelman, H.; Thibault-Starzyk, F.; Jacobs, P.A.; De Vos, D.E.
Chem. Eur. J., 2006, 12, 7353-7363.
[72] Prestipino, C.; Regli, L.; Vitillo, J.G.; Bonino, F.; Damin, A.; Lamberti, C.; Zecchina,
A.; Solari, P.L.; Kongshaug, K.O.; Bordiga, S. Chem. Mater., 2006, 18, 1337-1346.
[73] Hwang, Y.K.; Hong, D.Y.; Chang, J.S.; Jhung, S.H.; Seo, Y.K.; Kim, J.; Vimont, A.;
Daturi, M.; Serre, C.; Ferey, G. Angew. Chem., Int. Ed., 2008, 47, 4144-4148.
[74] Banerjee, M.; Das, S.; Yoon, M.; Choi, H.J.; Hyun, M.H.; Park, S.M.; Geo, G.; Kim, K.
J. Am. Chem. Soc., 2009, 131, 7524-7525.
[75] Gascon, J.; Aktay, U.; Hernandez-Alonso, M.D.; van Klink, G.P.M.; Kapteijn, F. J.
Catal., 2009, 261, 75-87.
[76] Vermoortele, F.; Ameloot, R.; Vimont, A.; Serre, C.; De Vos, D.E. Chem. Commun.,
2011, 47, 1521-1523.
194 Francesc X. Llabrs i Xamena

[77] Hasegawa, S.; Horike, S.; Matsuda, R.; Furukawa, S.; Mochizuki, K.; Kinoshita, Y.;
Kitagawa, S. J. Am. Chem. Soc., 2007, 129, 2607-2614.
[78] Kitagawa, S.; Noro, S.; Nakamura, T. Chem. Commun., 2006, 701-707.
[79] Noro, S.; Kitagawa, S.; Yamashita, M.; Wada, T. Chem. Commun., 2002, 222-223.
[80] Kitaura, R.; Onoyama, G.; Sakamoto, H.; Matsuda, R.; Noro, S.; Kitagawa, S. Angew.
Chem., Int. Ed., 2004, 43, 2684-2687.
[81] Wu, C.D.; Hu, A.; Zhang, L.; Lin, W.B. J. Am. Chem. Soc., 2005, 127, 8940-8941.
[82] Szeto, K.C.; Prestipino, C.; Lamberti, C.; Zecchina, A.; Bordiga, S.; Bjorgen, M.;
Tilset, M.; Lillerud, K.P. Chem. Mater., 2007, 19, 211-220.
[83] Szeto, K.C.; Lillerud, K.P.; Tilset, M.; Bjorgen, M.; Prestipino, C.; Zecchina, A.;
Lamberti, C.; Bordiga, S. J. Phys. Chem. B, 2006, 110, 21509-21520.
[84] Zhang, X.; Llabrs i Xamena, F.X.; Corma, A. J. Catal., 2009, 265, 155-160.
[85] Ingleson, M.J.; Barrio, J.P.; Guilbaud, J.B.; Khimyak, Y.Z.; Rosseinsky, M.J. Chem.
Commun., 2008, 2680-2682.
[86] Hermes, S.; Schroter, M.K.; Schmid, R.; Khodeir, L.; Muhler, M.; Tissler, A.; Fischer,
R.W.; Fischer, R.A. Angew. Chem., Int. Ed., 2005, 44, 6237-6241.
[87] Sabo, M.; Henschel, A.; Froede, H.; Klemm, E.; Kaskel, S. J. Mater. Chem., 2007, 17,
3827-3832.
[88] Opelt, S.; Turk, S.; Dietzsch, E.; Henschel, A.; Kaskel, S.; Klemm, E. Catalysis
Communications, 2008, 9, 1286-1290.
[89] Henschel, A.; Gedrich, K.; Kraehnert, R.; Kaskel, S. Chem. Commun., 2008, 4192-
4194.
[90] Schroeder, F.; Esken, D.; Cokoja, M.; van den Berg, M.W.E.; Lebedev, O.I.; van
Tendeloo, G.; Walaszek, B.; Buntkowsky, G.; Limbach, H.H.; Chaudret, B.; Fischer,
R.A. J. Am. Chem. Soc., 2008, 130, 6119-6130.
[91] Moon, H.R.; Kim, J.H.; Suh, M.P. Angew. Chem., Int. Ed., 2005, 44, 1261-1265.
[92] Esken, D.; Turner, S.; Lebedev, O.I.; Van Tendeloo, G.; Fischer, R.A. Chem. Mater.,
22, 6393-6401.
[93] Ishida, T.; Nagaoka, M.; Akita, T.; Haruta, M. Chem. Eur. J., 14, 8456-8460.
[94] Liu, H.; Liu, Y.; Li, Y.; Tang, Z.; Jiang, H. J. Phys. Chem. C, 2010, 114, 13362-13369.
[95] Uemura, T.; Hiramatsu, D.; Yoshida, K.; Isoda, S.; Kitagawa, S. J. Am. Chem. Soc.,
2008, 130, 92169217.
[96] Muller, M.; Zhang, X.N.; Wang, Y.M.; Fischer, R.A. Chem. Commun., 2009, 119-121.
[97] Hermes, S.; Schroder, F.; Amirjalayer, S.; Schmid, R.; Fischer, R.A. J. Mater. Chem.,
2006, 16, 2464-2472.
[98] Bossmann, S.H.; Jockusch, S.; Schwarz, P.; Baumeister, B.; Goeb, S.; Schnabel, C.;
Payawan, L., Jr.; Pokhrel, M.R.; Woerner, M.; Braun, A.M.; Turro, N.J. Photochem.
Photobiol. Sci., 2003, 2, 477.
[99] Bossmann, S.H.; Shahin, N.; Le Thanh, H.; Bonfill, A.; Worner, M.; Braun, A.M.
Chem. Phys. Chem., 2002, 3, 401.
[100] Cosa, G.; Chretien, M.N.; Galletero, M.S.; Fornes, V.; Garcia, H.; Scaiano, J.C. J. Phys.
Chem. B, 2002, 106, 2460-2467.
[101] Cosa, G.; Galletero, M.S.; Fernndez, L.; Mrquez, F.; Garca, H.; Scaiano, J.C. New J.
Chem., 2002, 26, 1448-1455.
[102] Corma, A.; Garcia, H. Eur. J. Inorg. Chem., 2004, 1143-1164.
Coordination Polymers 195

[103] Alkordi, M.H.; Liu, Y.L.; Larsen, R.W.; Eubank, J.F.; Eddaoudi, M. J. Am. Chem. Soc.,
2008, 130, 12639-+.
[104] Maksimchuk, N.V.; Timofeeva, M.N.; Melgunov, M.S.; Shmakov, A.N.; Chesalov,
Y.A.; Dybtsev, D.N.; Fedin, V.P.; Kholdeeva, O.A. J. Catal., 2008, 257, 315-323.
[105] Sun, C.-Y.; Liu, S.-X.; Liang, D.-D.; Shao, K.-Z.; Ren, Y.-H.; Su, Z.-M. J. Am. Chem.
Soc., 2009, 131, 1883-1888.
[106] Juan-Alcaniz, J.; Ramos-Fernandez, E.V.; Lafont, U.; Gascon, J.; Kapteijn, F. J. Catal.,
269, 229-241.
[107] Wee, L.H.; Bajpe, S.R.; Janssens, N.; Hermans, I.; Houthoofd, K.; Kirschhock, C.E.A.;
Martens, J.A. Chem. Commun., 46, 8186-8188.
[108] Kockrick, E.; Lescouet, T.; Kudrik, E.V.; Sorokin, A.B.; Farrusseng, D. Chem.
Commun., 47, 1562-1564.
[109] Uemura, T.; Kitagawa, K.; Horike, S.; Kawamura, T.; Kitagawa, S.; Mizuno, M.; Endo,
K. Chem. Commun., 2005, 5968-5970.
[110] Uemura, T.; Horike, S.; Kitagawa, K.; Mizuno, M.; Endo, K.; Bracco, S.; Comotti, A.;
Sozzani, P.; Nagaoka, M.; Kitagawa, S. J. Am. Chem. Soc., 2008, 130, 67816788.
[111] Pan, L.; Liu, H.; Lei, X.; Huang, X.; Olson, D.H.; Turro, N.J.; Li, J. Angew. Chem., Int.
Ed., 2003, 42, 542-546.
[112] Turro, N.J. Acc. Chem. Res., 2000, 33, 637.
[113] Schlichte, K.; Kratzke, T.; Kaskel, S. Microporous Mesoporous Mater., 2004, 73, 81-
88.
[114] Horcajada, P.; Surble, S.; Serre, C.; Hong, D.Y.; Seo, Y.K.; Chang, J.S.; Greneche,
J.M.; Margiolaki, I.; Ferey, G. Chem. Commun., 2007, 2820-2822.
[115] Horike, S.; Dinca, M.; Tamaki, K.; Long, J.R. J. Am. Chem. Soc., 2008, 130, 5854-
5855.
[116] Kato, C.N.; Hasegawa, M.; Sato, T.; Yoshizawa, A.; Inoue, T.; Mori, W. J. Catal.,
2005, 230, 226-236.
[117] Zou, R.Q.; Sakurai, H.; Han, S.; Zhong, R.Q.; Xu, Q. J. Am. Chem. Soc., 2007, 129,
8402-+.
[118] Kim, J.; Bhattacharjee, S.; Jeong, K.-E.; Jeong, S.-Y.; Ahn, W.-S. Chem. Commun.,
2009, 3904-3906.
[119] Hwang, Y.K.; Hong, D.Y.; Chang, J.S.; Seo, H.; Yoon, M.; Kim, J.; Jhung, S.H.; Serre,
C.; Ferey, G. Appl. Catal. A: Gen., 2009, 358, 249-253.
[120] Demessence, A.; D'Alessandro, D.M.; Foo, M.L.; Long, J.R. J. Am. Chem. Soc., 2009,
131, 8784-8786.
[121] Das, S.; Kim, H.; Kim, K. J. Am. Chem. Soc., 2009, 131, 3814-3815.
[122] Wang, Z.Q.; Cohen, S.M. Chem. Soc. Rev., 2009, 38, 1315-1329.
[123] Goto, Y.; Sato, H.; Shinkai, S.; Sada, K. J. Am. Chem. Soc., 2008, 130, 14354-14355.
[124] Savonnet, M.; Bazer-Bachi, D.; Bats, N.; Perez-Pellitero, J.; Jeanneau, E.; Lecocq, V.;
Pinel, C.; Farrusseng, D. J. Am. Chem. Soc., 2010, 132, 4518-4519.
[125] Ingleson, M.J.; Barrio, J.P.; Bacsa, J.; Dickinson, C.; Park, H.; Rosseinsky, M.J. Chem.
Commun., 2008, 1287-1289.
[126] Canivet, J.; Aguado, S.; Daniel, C.; Farrusseng, D. Chem. Cat. Chem., 2011, 3, 675-
678.
In: Coordination Polymers and Metal Organic Frameworks ISBN: 978-1-61470-899-5
Editors: O. L. Ortiz and L. D. Ramirez, pp. 197-223 2012 Nova Science Publishers, Inc.

Chapter 6

HIGH PRESSURE GAS STORAGE ON POROUS SOLIDS:


A COMPARATIVE STUDY OF MOFS AND
ACTIVATED CARBONS

A. Linares-Solano, D. Cazorla-Amors,
J. P. Marco-Lozar and F. Surez-Garca
Dpto. Qumica Inorgnica, Universidad de Alicante,
E-03080 Alicante, Spain.

ABSTRACT
Porous materials provide an alternative for satisfying gas storage demands for on-
board storage in transportation technologies (e.g. CH4 and H2) and for capture, storage
and transport (e.g. CO2). The principle of their storages is the use of a high pressure
adsorption process (or physisorption), as a supercritical gas (e.g. H2 and CH4) or as a
subcritical one (e.g. CO2). Such adsorption process has some advantages as: its high
storage capacity (very much depending on the surface area, porosity and pore size of the
material), its fast kinetic of storage and release (reversibility), its short refueling time, its
low heat evolution and its efficient cyclability. Additionally, the porous solid (the
adsorbent) presents advantages; different types are available (e.g., zeolites, porous
carbons, MOFs, all of them with a large variety of materials), its porosity, morphology,
size and shape are tunable. Among them, two types of porous solids stand out: 1) the
classical activated carbons and 2) the recent and new type of porous materials (i.e.
MOFs and COFs).
Most of the papers report the gas storage capacity of an adsorbent refereed per unit
of weight (i.e. gravimetric basis). However, for applications where the volume of the tank
is an important controlling factor (e.g. in transportation), the gas storage capacity should
also be reported per unit of volume (volumetric basis). And, the density of material used
should be consistently measured (i.e. tap or packing). Unfortunately, this is not always
the case and very frequently (as it happens with recent and new porous materials)
calculated density (e.g. crystal density) is used. Using such crystal density, impressive
volumetric storage capacities have been reported for MOFs (also COFs), claiming that

Corresponding Author. Fax; +34 965903454. E-mail address: linares@ua.es (A. Linares-Solano).
198 A. Linares-Solano, D. Cazorla-Amors, J. P. Marco-Lozar et al.

they can achieve higher storage capacities for H2, CH4 and CO2 than other porous
materials such as zeolites and porous carbons. In our opinion, such claimed superior gas
storage capacity of MOFs in relation to activated carbons needs further evaluations.
In this chapter, we comparatively analyses the adsorption capacity of two activated
carbons (ACs) and MOF-5 for storing gases (H2, CH4 and CO2) at different temperatures
(77K and RT) and pressures (from 0.1MPa to 20MPa) both on gravimetric and
volumetric basis paying attention to the data reported in the literature as well as on the
suitability of different densities employed.
We advance that, from the data presented and discussed in this chapter, the
outstanding adsorption capacities of MOFs in relation to ACs on volumetric basis,
frequently claimed in the literature, is mainly due to the use of an unrealistic high density
(crystal density) which, not including the inter-particle space of the adsorbents, gives an
apparently high volumetric gas storage capacity. Using a density measured similarly in
both types of adsorbents (e.g. tap density) MOF presents, on volumetric basis, and for all
gases and conditions studied, lower adsorption capacities than ACs, due to its lower
inherent density.

INTRODUCTION
Currently, there are different methods for storing gases, all of which still are under
research to improve their performances. They vary depending on the nature of the gas to be
stored. For Hydrogen, the most difficult one due to its especial physical characteristics, there
are mainly four methods [1-4]: liquid hydrogen, compressed gas, metal hydrides and high
pressure adsorption on highly porous adsorbents. Nowadays, despite the worldwide effort
carried out by scientists and engineers, and the massive number of published papers, the
research still has to continue in all of these storage systems. In fact, none of them can meet
the required gravimetric and volumetric targets for on board hydrogen storage in vehicles
which, for light vehicles and for the w hole storage system, the 2009 revised DOE report
gives a target for 2015 of 5.5 wt% and 40 g H2/L [5].
Among the four mentioned approaches, high pressure adsorption on highly porous
adsorbents appears to be an interesting alternative to store gases, not only for an on-board
storage of hydrogen in transportation technologies, but also for methane (natural gas; NG) [6-
8]. Additionally, other gases also require the use of a suitable adsorbent contained in a given
tank volume for other applications such as [9-15]: i) capture, storage and transport (for short-
term CO2 storage for CCS technologies), ii) gasoline evaporative control or VOC
transportation, iii) safe transportation of very dangerous compounds (e.g. arsine), or iv) for
medical applications. In all these cases, the basic principle of their use is the adsorption
process (physisorption) and, hence, the challenge for improving results is focused in the
design, optimization and development of new and more suitable porous adsorbents.
For some of these storage applications (e.g. H2, CH4 and CO2) the basic adsorption
principle has to be extended to high pressure adsorption (HPA) where these gases (except
CO2) are in a supercritical state. In such adsorption process the molecules of gases to be
stored (the adsorbate) are thereby bond by Van der Waals forces to adsorbent which has to
have a high porosity and surface area [16].
Although these gas-adsorbent interactions are weak, they are strong enough to allow the
amount of gas stored in the tank containing the adsorbent be enhanced (the adsorbate density
is higher than the density of the gaseous phase, under similar conditions). Depending on the
High Pressure Gas Storage on Porous Solids 199

nature of the gas to be stored, and the storage temperature, the maximum adsorption pressure
to be used varies considerably; from high pressures (< about 4MPa) to very high pressures
(from 10 to 70MPa). The former, the high pressure (HPA; < 4MPa) is used, for example, to
store methane and carbon dioxide at room temperature (RT) and to store hydrogen at
cryogenic temperatures (e.g. 77K) [7, 14, 15, 17-21]. For hydrogen at RT, the target required
for on-board applications become more difficult to be accomplished and, additionally to the
possible future development of new adsorbents, the use of very high pressures are required
(VHAP in the range of 10-70MPa) [17, 19, 22, 23].
Such use of adsorbents to store gases has some advantages due to the adsorption process
by itself and to the adsorbent characteristics. Thus, in relation to the adsorption process, it
should be noted that: i) the current reached storage capacity (very much depending on the
surface area, porosity and pore size of the material) is reasonably high enough, and ii) the
adsorption process has a short refueling time, an efficient cyclability, a fast kinetics of storage
and release as well as a low heat evolution. In relation to the adsorbent, it should be noted that
[4, 17, 19-21, 24-34]: i) there is a large variety of available porous materials having quite
different properties, such as zeolites, highly activated carbons, activated carbon nanotubes,
zeolite template carbons, carbide-derived carbons, metal organic frameworks (MOFs) and
covalent organic frameworks (COFs) and ii) in all of them, their porosity, surface area,
morphology, size and shape are tunable and, hence, able for further improvements.
In this way, the development of new porous materials has been, and continues to be,
intensively investigated [29, 35-45]. Hence, the number of publications reporting their
subsequent performance improvements for gas storage has increased considerably. Among
the highest reported values, two families of porous solids have shown to have better
performances [17-21, 23, 26, 28-35, 46-50]: 1) the more cl assical one (i.e. activated
carbons) and 2) the more r ecent and new type of porous materials (i.e. MOFs and COFs).
Among these two types of adsorbents (continuously been designed and investigated) the
superior gas storage performance of one in relation to the other is, in our opinion, up to date
still uncertain. This is so because gas storage values are not always reported in a similar way
on both families of adsorbents. Generally, in one of these families, the results are mostly
expressed on a gravimetric basis whereas for the other family both gravimetric and
volumetric basis are used. Additionally, those reporting volumetric storage do not always use
comparable densities. Therefore, a conclusive opinion about the best adsorbent for gas storage
application is still lacking and more attention is required performing suitable comparatives
analysis.
In general, as commented above, most of the papers dealing with high pressure gas
adsorption storages present their results expressed per unit of weight of the material (i.e. on
gravimetric basis). However, they should also be reported per unit of volume of adsorbent
(i.e. volumetric basis), as was pointed out long time ago (1994) by Chahine et al [51]. This
should be done considering: i) that the adsorbent has to be confined on the limited tank
volume (a vehicle for an on-board storage application in transportation) and ii) that the
density of the adsorbent depends very much on its porosity, as it will be commented next.
Unfortunately, the results on a volumetric basis are not frequently reported and/or the type of
density used is not well defined. Thus, those papers that convert their gravimetric results to
volumetric ones do not always give details of the experimental procedure used to measure the
density of the adsorbent investigated. This is especially important in the case of the above
mentioned r ecent and new porous materials (MOFs and COFs) that mostly use a
200 A. Linares-Solano, D. Cazorla-Amors, J. P. Marco-Lozar et al.

cal culated density (e.g. crystal density). Using such crystal density many papers, published
in very prestigious journals, have reported very high volumetric storage capacities.
Thus, in many of these papers it is claimed that MOFs can achieve higher storage
capacities, for H2, CH4 and CO2, than porous carbons and more recently, the same statement
has also been extended to COFs. Such better performance of MOFs can be confirmed in the
following three interesting and high quality papers and on their many cited references, as an
example of many others published papers. These, are: i) Except ional H2 Saturation Uptake in
Microporous Metal-Organic Frameworks ([31], JACS 2006), ii) St orage of Hydrogen,
Methane, and Carbon Dioxide in Highly Porous Covalent Organic Frameworks for Clean
Energy Applications ([29], JACS 2009) and iii) on U ltrahigh Porosity in Metal-Organic
Frameworks ([35], Science 2010). As in most MOFs-papers reporting volumetric capacity,
in these papers, the authors convert their gravimetric data (g H2/g) to volumetric one (g H2/L)
by mean of a calculated density. Some conclusive sentences can be extracted from these three
papers, such as: Volumetric H2 storage in MOFs are excellent and approach practical
utility [31], COFs, a new class of porous crystals, have high capacities for hydrogen,
methane, and carbon dioxide and they compare favorably with the most common carbon
materials, zeolites, mesoporous solids, and MOFs, an aspect that places them firmly among
such useful porous materials in terms of their gas storage capacities [29] and the
calculated gravimetric hydrogen density in MOF-210 (176 mg g1) exceeds that of typical
alternative fuels (methanol and ethanol) and hydrocarbons (pentane and hexane) [35]. It
should be recalled that these three statements were done using an estimated density of the
adsorbent, calculated from the pore volume and the density of hydrogen at 77K.
Consequently, the current feeling of most scientists dealing with high pressure adsorption
gas storage is that MOFs are the most outstanding adsorbents among the different existing
ones. As stated before, in our opinion, further analysis is needed for confirming such
conclusive opinion.

ADSORPTION CAPACITY AND GAS STORAGE


CAPACITY CORRELATIONS
At a first glance, the best and simplest way to know the gas adsorption capacity of a
given adsorbent is expressing its capacity per unit weight. This will allow comparing
adsorbents from any family type. However, for some applications, especially for gas storage
where the adsorbent has to fill a given volume, the adsorption capacity should also be
expressed per unit of volume. In that case, the inherent sample density is an important
parameter that, together with its porosity, will increase the storage capacity. However,
additionally to the inherent sample density, the density used to convert gravimetric data to
volumetric ones, as it will be shown later, can be the responsible of an over estimated high
performance. Hence, attention has to be paid to the adsorbent density used.
The capacity to store gases (e.g. hydrogen, methane and carbon dioxide) at different
temperatures and pressures has been intensively studied in all types of adsorbents. As
mentioned above, there are two families that present better results [17-21, 23, 26, 28-31, 33-
35, 46, 47, 49, 52]; 1) MOFs, a fascinating porous solid family with thousands of compounds
(macromolecular coordination compounds). Their porosity can be controlled varying the
High Pressure Gas Storage on Porous Solids 201

metal building units and/or the size of the organic linkers [40, 44, 45, 53] and 2) porous
carbons that also constitute a large family that includes, among other, highly activated
carbons, carbon nanotubes, template carbons, and carbide-derived carbons. Interesting
tunable properties converge on all these porous carbons, such as, morphology, size and shape,
chemical inertness, electric and thermal conductivities, density and porosity (micropore
volume, surface area and pore size distribution) [9-12, 33, 54-56].
The gas storage in adsorbents of these two families has been widely correlated with their
porosity (e.g. surface area, pore size distribution, micropore volume, narrow micropore
volume -pore size smaller than 0.7 nm, approximately-). The resulting correlation varies
depending on variables such as the nature of the gas and the temperature and pressure of the
storage system. However, these correlations are quite similar for all the adsorbents studied if
they are kept constant. Figure 1 shows, as an example of other similar reported plots [2, 13,
17-20, 57-61], the amount (excess) of hydrogen stored at 77K (up to 4 MPa) for both MOFs
and porous carbons families. From it, two important observations have to be noticed: i) both
families of adsorbents well define a similar trend correlating porosity (BET surface area as
well as micropore volume) with excess of H2 adsorption, and ii) the family of MOFs can
reach higher H2 uptakes than activated carbons.

Figure 1. Gravimetric excess H2 adsorption on two well characterized series of activated carbons and
MOFs at 77K up to 4MPa as a function of their surface areas and micropore volumes (redrawn from
[17, 18] and [31] respectively).

From these two observations, and from similar reported data [13, 19-21, 26, 29], three
conclusions can be extracted: 1) this type of relationship is independent of the type of
adsorbents studied which confirms that H2 storage at 77K (up to 4MPa) occurs through a
physical adsorption process, 2) the gravimetric uptake is controlled by the adsorption capacity
of the adsorbent, determined, for example, by N2 adsorption at 77K and 3) MOFs have a
better performance than activated carbons, giving superior amount of hydrogen stored per
unit of gram (about 20%). Their BET surface area reaches higher values and, hence, higher
hydrogen storage (about 20%) than those of the activated carbons [13]. Of course, further
porosity development on any adsorbents will give rise to more hydrogen uptake per unit of
202 A. Linares-Solano, D. Cazorla-Amors, J. P. Marco-Lozar et al.

mass. In the case of MOFs such development is opened to new metal building units, type of
linkers and synthesis process. In this way, much higher values than those of Figure 1, have
recently been reported for the MOF-200 series, with extremely and unique high BET surface
area (above 6000m2/g) [35]. In the case of the carbon family, the search for new materials is
also opened trying to reach the t heoretical maximum value (9.2 wt.%) calculated for
graphene sheets separated by idealized slip-shaped pores of a given size [62]. To date the
highest reported value for hydrogen storage at 77K is near 7 wt% and correspond to a zeolite-
templated carbon [49] and to a doubly activated carbon [63].
It is important to point out that, such type of correlations (obtained on a gravimetric
basis) should not be extended to a volumetric concept. They are only true when the hydrogen
uptake is expressed on a gravimetric basis but this is not correct at all when the hydrogen
uptake is expressed on a volumetric basis. This is so, because the density depends on its
assessment method and because the adsorbent changes with its porosity, as it is clearly shown
in Figure 2a that compiles results obtained on a series of activated carbons [18, 19]. The
Figure plots the changes of two types of densities (among many others possible) as a function
of the sample porosity (apparent BET surface area and micropore volume). These two
densities are: the tap density (determined using a standard procedure ASTM D7481-09) and
the packing density (sample subjected in a mould to a pressure of 415kg/cm2).

Figure 2. (a) Evolution of the tap and packing densities of two well characterized series of activated
carbons as a function of their surface areas and micropore volumes (redrawn from [18]) and (b) their
volumetric excess H2 adsorption (using packing density) as a function of their surface areas and
micropores volume (redrawn from [18]).

Three observations from this Figure 2a merit comments: i) tap density is lower than
packing density (the higher is the pressure used to pack the solid the higher will be the
resulting density, as it will be discussed latter on), ii) for a given adsorbent, different densities
can be used to convert gravimetric to volumetric storages; the higher the density selected, the
higher will be the resulting reported storage capacity. This is especially important when gas
storage capacities of different adsorbents are compared and their densities have not been
assessed similarly and iii) both the tap and the packing densities decrease as the adsorbent
porosity (micropore volume and apparent BET surface area) increases. Using one of the
densities of Figure 2a (packing density), the results of Figure 1 are converted to volumetric
High Pressure Gas Storage on Porous Solids 203

data. The resulting plots for porous carbons is shown in Figure 2b, from which some
comments have to be done: 1) The plot of Figure 2b is quite different to that of Figure 1,
confirming our concerns about the conclusions that could be reached from the analysis of the
gas storage performed only on a gravimetric basis, 2) our claim that the analysis has to be
done also on a volumetric basis, measuring the density of the adsorbent and 3) to improve the
storage capacity of a given adsorbent, it is better to try to maximize the combination of both,
the density and the porosity of the adsorbent than just trying to develop its porosity. In fact,
Figure 2b shows that activated carbons with apparent BET surface area of 2000-3000m2/g
maximize their hydrogen storage in relation to one that has 4000m2/g. Similar plots have been
obtained [59] with other series of activated carbons (activated carbon fibers).
As commented above, and because the adsorbent has to fill a given tank volume, the gas
storage capacity of adsorbents under study should also be expressed on a volumetric basis.
However, doing so, it has to be noted that the reliability of the results depend too much on the
density used to convert the gravimetric data to volumetric ones. This could be solved, as
shown in Figure 2b, measuring the density of the adsorbent under study but, unfortunately, a
similar study has never been reported for the MOF family. Hence, up to day, a reliable
comparative analysis done on a volumetric basis and using a large numbers of adsorbents
from both families, has not been performed yet. We can only find a few exceptions on recent
publications [46, 64] that limit their study to a short number of samples (two or three). The
main difficulty to carry such type of study in the case of the MOF family is the synthesis of
enough amount of sample to determine their densities.
In this chapter, we deep into the claimed superior capacity of MOFs in relation to
activated carbons (ACs), reviewing and extending our comparative works on the field of gas
storage (H2, CH4 and CO2) [46, 65], with new data. The work discussed on this chapter
covers: i) the characterization of selected adsorbents from each families (MOFs and ACs)
such as density (tap and packing), porosity, surface area and adsorption capacity changes over
pressure and time, ii) high pressure adsorption of gases (H2, CH4 and CO2) under different
temperatures (77K and RT) and pressures (0.1, 4 and 20MPa) and iii) a comparative analysis
of both gravimetric and volumetric results paying attention to the data reported in the
literature as well as to the suitability of different densities employed.
To carry out this work the selected samples (see below) have to meet: i) to be the best
samples of each family, ii) to be widely studied, to allow results comparison, iii) to be
available, to have enough sample amount to carry out a deep characterization and iv) to be
similar on their adsorption capacity to allow such comparative study. As a result of such
sample selection (similar adsorption capacities and being similarly characterized using the
same equipments and experimental procedure) our aim to compare MOFs and ACs can be
easily carried out.
In summary, the gas adsorption performances of different adsorbents well done in terms
of their gravimetric uptakes might not be a suitable way for gas storage applications because
the adsorbent has to be contained in a given tank volume. Hence being its density the key
parameter, we advance, from the data discussed in this chapter, our concerns about the use of
the crystal density for expressing adsorbent volumetric capacity and about the claimed
superior capacity of MOF in relation to ACs.
204 A. Linares-Solano, D. Cazorla-Amors, J. P. Marco-Lozar et al.

MATERIAL AND METHODS


Samples Used

Three adsorbents (one MOF and two activated carbons) are comparatively analyzed in
the present chapter in terms of their gravimetric and volumetric gas storage capacities. These
three adsorbents are selected because all of them have high and very similar adsorption
capacities (per unit of gram), enough amount of sample are available (two of them from
companies), and both are among the samples most studied having high gas storage capacities
[17, 18, 20, 21, 31, 51, 61, 66-74]. These samples are: 1) MOF-5 (also known as IRMOF-1)
supplied by BASF. The received sample showed different shapes and sizes (flakes and
powders). Therefore, in order to homogenise the powder size, the sample was crushed before
experiments, 2) a KOH activated carbon known as Maxsorb3000 (also as MaxsorbMSC30
[51, 73, 74]) supplied by Kansai Coke and Chemical Co [75]. This sample, named in this
work as Max3, is among the commercial activated carbons, one with the better performance)
and 3) a second KOH activated carbon (AC1) which was prepared in our labs from a Spanish
anthracite, following the experimental procedure previously described [76].
MOF-5 selection. Selecting MOF-5, among the thousands of synthesized MOFs, is
because its adsorption capacity is high and comparable to Maxsorb3000 and it is, without
doubt, the archetype of the MOFs family. Therefore, it has been the most widely studied and
its storage capacity per unit of volume is the highest one among the reported MOFs values.
This is so, although other MOFs with much higher BET surface areas have recently been
synthesized (e.g. MOF-177 [31, 52, 77] or the MOF-200 series [35]). Consequently, their
gravimetric storage capacity for H2, CH4 and CO2, are considerably higher than MOF-5. An
example, redrawn from literature data [31, 35], can be seen in Figure 3a, where the H2
uptakes for three most performing MOFs (MOF-5, MOF-177 and MOF-210) are plotted.
Similar results and conclusions are obtained plotting their CH4 and CO2 adsorption capacities
per unit of weight. However, analyzing their reported H2 storages at 77K per unit of volume
[31, 35, 52], MOF-5 has the better performance, as shown in Figure 3b (redrawn from
published work [31]). Similar conclusion can be reached analyzing, per unit of volume, their
storage capacities for CH4 and CO2 [35, 36, 68].

Figure 3. H2 uptakes at 77K for MOF-5, MOF-177 and MOF-210, redrawn from published work [31,
35], (a) on a gravimetric basis and (b) on a volumetric basis (using their crystal densities)
High Pressure Gas Storage on Porous Solids 205

Sample Characterization

Prior to gas storage measurements, a sample characterization has been done. It includes
the measurement of density, porosity and apparent surface area, obtained as explained next.
Different types of sample density were measured: tap (also known as bulk or apparent),
packing (will depend on the pressure used to pack the sample) and helium (also known as true
density). Additionally, although the ACs are not crystalline materials their crystal densities
has been calculated (for comparison purposes) from experimental data. The procedure used
for tap, packing and helium densities is the usual ones: (i) by filling and vibrating a container
with a known weight of sample, obtaining the occupied volume; (ii) by pressing a given
amount of sample in a mould at a pressure of 415kg/cm2 (other pressures could be used) [46,
78] and (iii) using a Micromeritics Accupyc 1330 pycnometer for helium density. The fourth
density (crystal density) was estimated as the sum of the volume occupied by the atoms of the
material (i.e. 1/He density) plus the total pore volume obtained from the N2 adsorption
isotherm at P/Po = 0.99.
The textural characterization was performed by physical adsorption of N2 at 77K (using
an ASAP 2020, Micromeritics, equipment) and CO2 at 273K (using an Autosorb-6,
Quantachrome, equipment). Samples were previously degassed at 423K under vacuum for 4h.
From N2 adsorption isotherms, the apparent BET surface areas (SBET) as well as the
micropore volumes using the DR (Dubinin-Radushkevich) equation (pore size smaller than
2nm; VDR(N2)) were calculated. In the same way, from CO2 adsorption isotherms and using
the same equation, the volumes of narrow micropores (pore smaller than 0.7nm [79];
VDR(CO2)) were assessed.

Gas Storage Experiments

The storage of the different gases studied (hydrogen, methane and carbon dioxide) has
been carried out in different equipments, depending on the pressure required (from 0.1 to
20MPa) and the temperature used (room temperature and 77K).
For hydrogen storage three equipments were used depending on the range of pressures
analyzed: i) for sub-atmospheric pressures at 77K, an automatic volumetric adsorption
apparatus, (ASAP 2020; Micromeritics). The sample was outgassed at 423K under vacuum
for at least 4h before measuring their isotherms, ii) for high pressures (up to 4MPa) at 77K, a
Sartorius 4406- DMT high-pressure microbalance using the procedure described elsewhere
[18, 80]. Approximately 200mg of sample was outgassed i n situ at 423K during 4h under
vacuum. The experimental results were corrected for buoyancy effects related to the
displacement of gas by the sample, sample holder and pan and iii) for very high pressures (up
to 20MPa) at room temperature (298K), a fully automated volumetric apparatus, designed and
built up at University of Alicante [17, 81]. About 700mg of sample were degassed at 423K
during 4h under vacuum and the weight of the degassed sample was measured. After that, the
sample was located in the sample holder, where it was degassed at 423K during another 4h in
vacuum, in order to prepare the sample for the measurement. The bulk gas densities were
calculated by the equation of state of Modified-Benedic-Webb-Rubin proposed by Younglove
[82].
206 A. Linares-Solano, D. Cazorla-Amors, J. P. Marco-Lozar et al.

For methane and carbon dioxide their storages were studied at room temperature up to 3
and 4MPa respectively, in the same Sartorius device. In both cases, about 150mg of sample
were placed in the sample holder and degassed i n situ at 423K under vacuum until a
constant weight was measured. The experimental results were corrected for buoyancy effects
related to the displacement of gas by the sample, sample holder and pan [83].

RESULTS AND DISCUSSION OF SAMPLE CHARACTERIZATION


Table 1 summarizes the different sample densities measured and used in this study. As
expected, the density differs according to the experimental procedure used to assess it. It
increases in the following sequence: tap density < packing density < crystal density < true
density. Consequently the type of density used to convert the gravimetric results to volumetric
ones, will affect very much, and presumably in an unrealistic manner, the volumetric storage
results. This highlights the importance of reporting the density of the material and the need to
well describe how it has been measured.

Table 1. Values for the different densities of the samples studied

Tap density Packing density True density Crystal density


Sample
(g/cm3) (g/cm3) * (g/cm3) (g/cm3)**
Max3 0.36 0.41 2.1 0.48
AC-1 0.38 0.50 2.1 0.52
MOF-5 0.30 0.57 2.0 0.58
*
Sample pressed in a mould at 415kg/cm2, ** Calculated as the volume occupied by the atoms of the
material (i.e. 1/He density) plus the total pore volume (obtained from N 2 adsorption at P/Po =
0.99).

In relation to the MOF-5 used in this study, it has to be pointed out that the obtained
densities of Table 2 agree very well with reported data; a tap density 0.3g cm-3 by Mueller et
al. [69] and a crystal density 0.59g cm-3 by Wong-Foy et al. [31].
Tap and packing densities include the particle volume (i.e. the volume occupied by atoms
and the internal pore volume of the material), and the inter-particle space volume. The
difference between them is that in the second case, by applying pressure, the particles are
forced to settle down and, therefore, the inter-particle space is reduced, increasing the density.
The higher the packing pressure used the higher will be resulting packing density (more data
and discussion will be presented later on, after analyzing hydrogen storage) [46, 78]. Crystal
density of the materials, calculated as explained before, includes the volume occupied by the
atoms and the internal pore volume but it does not include the inter-particle space [31, 36, 46,
84]. Hence, it assumes that the adsorbent is a monocrystal. Helium density, having no sense
to be used for expressing volumetric storage capacity, is only used for calculation of the
crystal density.
Figure 4a shows the N2 adsorption/desorption isotherms at 77K for the three studied
samples (AC1, Max3 and MOF-5). In the literature a large number of adsorption isotherms
has been published for well synthesized MOF-5, giving BET ranging from 2500-3800m2/g.
This is due to small variations in the synthesis process and to its sensitivity to ambient
High Pressure Gas Storage on Porous Solids 207

conditions (e.g. air exposure and degradation [70, 85]), as it will be discussed later on.
Among these well synthesized MOF-5 [21, 31, 39, 46, 53, 58, 61, 67, 69, 85, 86] we have
selected one having comparable adsorption isotherm [87] which is also plotted in Figure 4a.
The Figure 4a shows the shapes and type of isotherms (type I of the IUPAC
classification). From it, it can be concluded that all the selected samples have quite
comparable adsorption capacity and are essentially microporous. However they present some
differences: 1) MOF-5 adsorbs more N2 than the ACs at low relative pressures and less at
high relative pressures than both ACs and 2) the isotherms of the two ACs present much
wider knee than the MOF-5 isotherm. These two observations indicate that MOF-5 has a
lower but more homogeneous microporous network system, than the two ACs used.

Figure 4. (a) N2 adsorption/desorption isotherms at 77K for the three selected samples and a MOF-5 (*)
redrawn from published work [87] and (b) Sub-atmospheric CO2 adsorption isotherms at 273K for the
three selected samples and a MOF-5 [*] redrawn from published work [88].

Figure 4b presents the sub-atmospheric CO2 adsorption isotherms at 273K of the three
selected samples. It includes a MOF-5 isotherm redrawn from the very scarce literature
reporting such type of data [88]. Because sub-atmospheric CO2 adsorption at 273K is related
with narrow microporosity (i.e., pore width < 0.7nm [79, 89]), it can be stated that: i) the
three materials have narrow micropores and ii) the amount of narrow microporosity is lower
for MOF-5 than for the two ACs.
Table 2 compiles the data obtained from the analysis of the N2 and CO2 adsorption
isotherms (BET and DR equations). It can be seen that the three adsorbents studied have
comparable N2 adsorption capacities per unit of weight and hence BET surface area and
micropore volumes. However, MOF-5 (having BET in agreement with data published by
other authors [31, 46, 50, 53, 69, 70, 85, 86]) has much lower CO2 adsorption capacity than
the ACs, as commented above.

Table 2. Porosity characterization of the samples used, deduced


from N2 (77 K) and CO2 (273K) adsorption isotherms

Sample SBET (m2/g) VDR(N2) (cm3/g) VDR(CO2) (cm3/g)


Max3 3180 1.31 0.70
AC-1 3120 1.25 0.72
MOF-5 2800 1.13 0.44
208 A. Linares-Solano, D. Cazorla-Amors, J. P. Marco-Lozar et al.

At a first glance, according to the shape of the low relative pressures range of the N2
adsorption isotherm (Figure 4a) one would expect that the VDR(CO2) value of Table 2 for
MOF-5 were even higher than those for the ACs. To confirm such low VDR(CO2) value (that
will affect its storage capacity) a N2 adsorption isotherm at 77K, covering the same low
relative pressure range as CO2 (i.e. 10-3 to 10-5) was carried out (Micromeritics; ASAP 2020)
as shown in Figure 5. MOF-5 has an unusual sigmoidal shape (to the best of our knowledge,
not reported before for N2 at 77K at these low relative pressures) which is not observed on the
ACs. Such behaviour seems to be related to the lower VDR(CO2) of MOF-5 in relation to both
ACs. Additionally, such unexpected N2 adsorption isotherm could indicate that MOF-5 has
two pore size distributions in the narrow micropore range. Because similar results have not
been reported, more work is needed to understand its origin and its possible relationship with
its sensitivity to ambient conditions (e.g. air exposure).
The lower volume of narrow micropores of MOF-5, in relation to both ACs, has to have
an effect on its performance for gas storage (hydrogen, carbon dioxide and methane)
especially when they are stored at RT, as it will be commented later on.

Figure 5. N2 adsorption isotherms at 77K for the three selected materials at low relative pressures.

HYDROGEN ADSORPTION AT 77 K
At Sub-Atmospheric Pressures (< 0.1MPa)

Figure 6 plots, on a gravimetric basis, the H2 adsorption/desorption isotherms obtained


for the three studied adsorbents at 77K and sub-atmospheric pressures. The figure also
includes a MOF-5 isotherm taken from the literature [86]. It can be observed that: i) the
adsorption capacities of both MOF-5 are identical, ii) all these hydrogen adsorption isotherms
are clearly of Type I and iii) all are completely reversible. However, comparisons of H 2
adsorption capacities reveal substantial differences between the two ACs and both MOF-5.
The amount of hydrogen adsorbed changes between 1.3 wt% (MOF-5 samples) to 3.3 wt%
(ACs samples). These differences are consistent with the previous comments about the low
narrow micropore volume of MOF-5, and are the first example of this chapter that confirms
the relevant contribution that the narrow microporosity has on the adsorption of H2 at these
conditions of low pressures and temperature [17, 19].
High Pressure Gas Storage on Porous Solids 209

Figure 6. H2 adsorption isotherms at sub-atmospheric pressure on the selected adsorbents at 77K in


gravimetric basis (wt%) and a MOF-5 [*] redrawn from published work [86].

For gas storage application we have highlighted the importance of expressing the amount
of hydrogen adsorbed per unit of volume. We have also pointed out the importance of the
density and the need of assessing the density in a similar way to carry out a comparison
without misunderstandings. Figure 7a plots the above results but now expressed on a
volumetric basis using, for the three samples, the same type of density (tap) that is obtained
similarly. Of course, such comparison carried out mixing different types of density will give
different, and wrong, results. The Figure shows, without doubt, that for low pressure H2
adsorption at 77K both ACs behave much better than MOF-5.

Figure 7. (a) H2 adsorption isotherms at sub-atmospheric pressure at 77K on volumetric basis using the
tap density of the adsorbents (g/L) and (b) the use of different densities for a given adsorbent (Max3).

Figure 7b plots the volumetric results of a given adsorbent (e.g. Max3) calculated using
the three densities of Table 1. The uptake increases, as expected, depending on the value of
the density used; the higher the density used (to convert gravimetric data to volumetric ones)
the higher will be adsorption results (tap < packing < crystal). Similar trends are observed for
the other two samples (AC1 and MOF-5).
210 A. Linares-Solano, D. Cazorla-Amors, J. P. Marco-Lozar et al.

At Moderate Pressures (< 4MPa)

Figure 8a shows the excess H2 adsorption isotherms at 77K (up to 4 MPa) for these three
samples, including the isotherm of MOF-5 (IRMOF-1) reported by Wong-Foy et al [31]. Two
points should be remarked from this Figure: i) the H2 uptakes expressed on a gravimetric
basis (wt%) for the two ACs and MOF-5 are quite similar. This confirms, independently of
the type of adsorbent used (ACs or MOF), that the apparent BET surface area and/or the
micropore volume control the H2 uptake at 77K at high pressures [13, 17, 18, 20, 21, 29], as it
was commented in Figure 1 and ii) it is concluded that both MOF (IRMOF-1 from reference
[31] and MOF-5 used in this work) are very similar because their H2 adsorption isotherms are
very close. For the purpose of this work, to point out such similarity of both MOF-5 samples
is important. Among the different reported values for MOF-5 [21, 50, 70, 85] the similar
results excludes problems related to its sensitivity to the synthesis process, manipulation (e.g.
air exposure) and degradation [85].

Figure 8. (a) Excess adsorption isotherms for hydrogen at 77K on the different materials up to 4 MPa in
gravimetric basis (wt%) and a MOF-5 [*] redrawn from published work [31] and (b) H2 adsorption
isotherms for the selected samples measured at 77K in volumetric units (g/L) and a MOF-5 [*] redrawn
from published work [31].

H2 adsorption isotherms on volumetric basis have been plotted in Figure 8b by using the
tap density of the sorbents. These isotherms have been plotted together with the isotherm
reported by Wong-Foy et al [31] calculated using the crystal density of MOF-5.
Firstly, contrarily to the similar results observed in Figure 8a, the volumetric results
obtained using the tap density for the three samples are quite different. Their performances
increase as follows: MOF-5 < Max3 < AC1 because of their different inherent densities
(Table 1). Secondly, the data for the MOF-5, obtained from the literature, has almost twice
volumetric uptake of H2 than the MOF-5 used in this study. This is so because to convert their
similar gravimetric uptakes to volumetric ones, we use the tap density (0.30cm3/g) whereas
for the reported MOF-5 a density two times higher is used (crystal density of 0.59cm3/g [31]).
This is another example, showing the significant differences that can exist using different
types of densities and hence the wrong conclusion that can be extracted comparing
adsorbents. These results highlight that much more work is needed before being able to get a
conclusive statement about adsorbent with the best performance for gas storage applications.
High Pressure Gas Storage on Porous Solids 211

This is especially relevant for MOFs in which unrealistically high adsorption values are
obtained using crystal density, as it will be shown next.

At Moderate Pressures (< 4MPa) Using Different Packing Densities

Up to this moment, we were presenting and discussing the volumetric hydrogen storage
using only the tap density. In this section, using two samples (Max3 and MOF-5), the effect
of increasing the packing pressure will be analyzed in order to increase hydrogen storage at
77K up to 4MPa. This analysis can be extended to the activated carbon AC1, as well as to the
results at room temperature for the three gases (H2, CH4 and CO2) that will be discussed later
on.
As shown before (Table 1) the density value (and hence the volumetric value) differs
according to the experimental procedure used to assess it, increasing from tap density <
packing density < crystal density < true density. The first two, tap and packing densities,
include the particle volume and the inter-particle space volume. By applying pressure, the
particles are forced to settle down and, therefore, the inter-particle space is reduced. The
higher the packing pressure used, the higher will be the resulting packing density and its
corresponding gas storage value. Hence, up to a certain limit, this is a way to increase the
filling of a given thank volume with the adsorbent.
Table 3 summarises the packing densities for the samples Max3 and MOF-5 at different
applied pressures. As it is expected, the packing density of the samples increases with the
applied pressure.

Table 3. Packing density of the adsorbents as a function of the pressures

Pressure density Pressure density


Sample Sample
(kg/cm2) (g/cm3) (kg/cm2) (g/cm3)
Max3-P1 415 0.41 MOF5-P1 415 0.57
Max3-P2 830 0.42 MOF5-P2 830 1.17
Max3-P3 1452 0.46 MOF5-P3 1452 1.31
Max3-P4 2075 0.47 MOF5-P4 2075 1.48

The evolution of these packing densities (starting with tap density with no pressure
applied) with the applied pressure is shown in Figure 9a. Crystal densities of both adsorbents
(i.e. the theoretical maximum packing density that can be achieved for a zero inter-particle
space) are also included. Figure 9b compiles the corresponding volumetric H2 uptake
calculated from these packing densities.
Figure 9a shows that the evolution of the packing density with pressure is different for
both samples. The packing density of Max 3 increases slowly with pressure (from 0.41 to
0.47g/cm3) tending to reach its theoretical crystal density (0.48g/cm3) at the maximum
pressure studied. Contrarily, MOF-5 is more easily compacted than the activated carbon, its
packing density increases with pressure (sharply up to 830kg/cm2 and slower above),
reaching its crystal density value (0.57g/cm3) at pressure of 415kg/cm2 and attaining at the
highest pressure used, an unreasonable high value (1.48g/cm3) and hence an unreasonable
high volumetric H2 storage (Figure 9b).
212 A. Linares-Solano, D. Cazorla-Amors, J. P. Marco-Lozar et al.

Figure 9. (a) Evolution of the packing density of Max 3 and MOF-5 versus compression pressure
applied. Full symbols correspond to their crystal density and (b) calculated volumetric H 2 uptakes using
packing densities of Table 3 (black points for Max3 and blue points for MOF-5).

Figure 10. H2 adsorption isotherms at 77K on volumetric units (g/L), using densities obtained at
different packing pressures measurements for Max3 (a) and MOF-5 (b).

Figure 10 plots the volumetric excess H2 adsorption isotherms that would be obtained if
the densities of Table 3 are used to convert their gravimetric data. In the case of Max3 (Figure
10a) a reasonable, and expected, volumetric increase is observed with the packing pressure
used. In the case of MOF-5 the volumetric storage values are unrealistically and artificially
high for pressures above 415kg/cm2, even higher than the crystal density value (Figure 9b),
High Pressure Gas Storage on Porous Solids 213

suggesting that the textural properties of the adsorbent are modified. This observation agrees
with a recent work of Zacharia et al. [90] that studied the preparation of monoliths by
mechanical compression of bulk powdered MOF-177 and the hydrogen adsorption capacities,
both on gravimetric and volumetric basis, of these monoliths. They observed that the H2
excess adsorption capacity on gravimetric basis diminishes with increasing the density of the
monoliths which was attributed to the decreasing of their porosity and the progressive
collapse of MOF-177.
Therefore, when a sample is subjected to high axial pressure, it is first necessary to check
that its structure can support such pressure and that its porous texture does not change. To
check it, both adsorbents have been analyzed by N2 adsorption at 77K and their adsorption
isotherms are plotted in Figure 11.

Figure 11. Effect of the applied pressure on N2 adsorption isotherms. (a) Max3 and (b) MOF-5.

The isotherms of Max3 (Figure 11a) remain practically unchanged by increasing the
packing pressure which indicates that its porosity is not significantly modified (the surface
area decreases only from 3180m2/g (fresh sample) to 3000m2/g (sample subjected to a
packing pressure of 2075kg/cm2). Contrarily, the isotherms of MOF-5 are strongly affected
by the applied pressure; N2 adsorption decreases drastically above 415kg/m2 and hence its
BET surface area (from 2800m2/g to only 85m2/g) proving that its texture is completely
damaged (collapsed). Similar results were recently reported [78]. The effect of the pressure
on its porosity depends on the pressure range applied; at pressures lower than 415kg/cm2, the
inter-particle space decreases, hence the packing density approaches the crystal density and
214 A. Linares-Solano, D. Cazorla-Amors, J. P. Marco-Lozar et al.

the porosity remains almost unchanged (see in Figure 11b) at higher pressures, the pore
volume noticeably decreases (collapses).
From Figures 9 and 10, we conclude that, for both materials, it is possible to increase
their density by decreasing the inter-particle space and consequently their volumetric H2
storage. However, in the case of MOF-5, the packing pressure used has to be limited to
415kg/m2 because, above that pressure, there are irreversible textural damages (Figure 10b).

Gas Adsorption at Room Temperature

Before analyzing the storage capacities of the three samples at RT, it has to be
remembered that the storage of H2, CH4 and CO2 takes place by a high pressure adsorption
process. Hence, both the properties of the adsorbents (mainly surface area, micropore volume
and narrow micropore volume) and the r elative pressure reached (depending on the
physical properties of the gas analyzed) will control the storage capacity of these adsorbents.
Under the experimental conditions used in this study, hydrogen and methane are
adsorbed as supercritical gases but carbon dioxide is not. Hence, the concept of relative
pressure (P/Po, being Po the saturation pressure) can only be used for CO2 but not for H2 and
CH4. For them a r elative pressure calculated as P/Pcs being Pcs the saturation pressure
calculated using the critical pressure and the empirical equation proposed by Dubinin [83, 91,
92] has been used. The P/Pcs value reached with the different gases, pressures and
temperatures have to be considered when comparing results.
H2 at 77K has a Pcs value of 7MPa, hence the results above discussed going from sub-
atmospheric pressure to 4MPa, cover a relative P /Pcs range from 0.014 (at 0.1MPa) to 0.57
(at 4MPa). The new results that we are going to discuss cover lower P /Pcs ranges for H2 at
RT (its Pcs at RT is 106MPa, the pressure analyzed is up to 20MPa, hence its P/Pcs is <
0.19) and for CH4 at RT (its Pcs at RT is 11.25MPa, the pressure analyzed is up to 3MPa,
hence its P/Pcs is < 0.27 at 3MPa). For CO2 at RT (its Po is 6.41MPa, the pressure analyzed
is up to 4MPa, hence its P/Po is < 0.6) which is slightly higher than 0.57 for H2 at 77K and
4MPa.

Hydrogen at High Pressures (< 20MPa)

The excess hydrogen adsorption isotherms, expressed on a gravimetric basis, and


measured at 298K and up to 20MPa, are shown in Figure 12a. The amount of hydrogen
adsorbed on gravimetric basis increases in the following order MOF-5 < Max3 < AC-1, being
0.86, 1.07 and 1.17 wt%, respectively. The Figure also presents an adsorption isotherm
previously published in the literature [85]. Similarly to what happens in Figure 8, the amounts
of H2 adsorbed by both MOF-5 agree very well. The lower capacity of both MOF-5 in
relation to the ACs can again be explained considering its lower VDR(CO2). Additionally, this
lower performance should be expected, in conformity with reported data obtained under
similar H2 storage conditions [17] where it was shown that the narrow micropore volume (<
0.7nm) controls the H2 storage at room temperature and a pressures lower than 20MPa [17,
19, 23, 58, 93] which is also in agreement with the reported value for the optimal pore size for
hydrogen storage (about 0.6-0.7nm [23, 94, 95]).
High Pressure Gas Storage on Porous Solids 215

Figure 12. Excess adsorption isotherms for hydrogen at 298K on the different materials studied here
and a MOF-5 [*] redrawn from published work [85], (a) in gravimetric basis (wt%) and (b) on
volumetric basis (g/L) and a MOF-5 [*] redrawn from published work [85] using their crystal density.

Figure 12b presents the excess adsorption calculated on a volumetric basis using the tap
density of Table 1. It also includes the volumetric adsorption isotherm of a MOF-5 redrawn
from Kaye et al [85] using its crystal density. It can be seen that their performances maintain
the same order; MOF-5 < Max3 < AC1 which is slightly increased due to the effect of their
different inherent tap densities. However, if crystal density is used, the resulting volumetric
capacity is apparently twice, in relation to the use of tap density (compare both MOF-5 in
Figure 12b).
In summary these results point out the importance of preparing adsorbents having a good
balance between porosity development and density to maximize their volumetric storage
capacity.

Methane at Moderate Pressures (< 4MPa)

Because natural gas (methane) is an alternative fuel to conventional ones as gasoline and
diesel, its storage under pressure is of a great interest. Currently, natural gas for vehicles is
mainly stored in a compressed form, at 20MPa (CNG). To reduce weight, multistage
compression and to increase safety transport, adsorbed natural gas (ANG) on an adsorbent,
has intensively been studied [6-8, 13, 29, 35, 38, 45, 96, 97]. For such purpose, again two
families of adsorbents give the best results; ACs and MOFs.
Figure 13 presents the methane isotherms for the two activated carbon samples measured
up to 3MPa and a MOF-5 redrawn from [98], expressed on a gravimetric basis (Figure 13a)
and expressed on a volumetric basis (Figure 13b), using the tap density of Table 1. The Figure
13b also includes the methane isotherm on MOF-5 on a volumetric basis using its crystal
density.
Figure 13 (a and b) allows to see that ACs can store higher amounts of CH4 than MOF-5,
as it also happens for H2 storage at RT up to 20MPa. The low VDR(CO2) of MOF-5, in
relation to the ACs, can explain its lower performance to store H2 and CH4 at RT because of
the low P /Pcs reached in both adsorbates. Again, if crystal density is used (Figure 13b)
MOF-5 becomes the material with the best artificial performance.
216 A. Linares-Solano, D. Cazorla-Amors, J. P. Marco-Lozar et al.

Figure 13. CH4 adsorption isotherms at 298K for the three adsorbents (the MOF-5 [*] is redrawn from
published work [98]); (a) on gravimetric basis (wt%) and (b) on a volumetric basis using theirs tap
densities for AC samples and MOF-5 (a) and using its crystal density for sample MOF-5 (b).

Carbon Dioxide at Moderate Pressures (< 4MPa)

Nowadays, carbon capture, storage and transport (for short-term CO2 storage for CCS
technologies) is one of the technological options that are being studied in order to control CO2
emissions to the atmosphere and, hence, to reduce climatic change effects. Among them,
adsorption on porous materials has a special interest and therefore it is intensively studied.
ACs and more recently MOFs have been postulated as interesting candidates for such capture
application [37, 68, 99, 100]. In the following we comparatively analyze these two materials
for storing CO2 at RT.

Figure 14. CO2 adsorption isotherms at 298K for the three adsorbents. (a) on a gravimetric basis (wt%);
MOF-5 [*] redrawn from [68] and (b) on a volumetric basis; MOF-5 [*] redrawn from [68] by using
crystal density.

Figure 14a plots the CO2 adsorption of the three samples studied and it includes a MOF-5
extracted from the literature [68], for comparative purposes. It can be seen that: i) both ACs
have higher adsorption than both MOF-5, ii) on a gravimetric basis, the MOF-5 used in this
study is quite comparable to the one from the literature, iii) MOF-5 shows an unusual S-
shaped isotherm not found for both ACs.
High Pressure Gas Storage on Porous Solids 217

Figure 14b plots the CO2 adsorption isotherms per liter of adsorbent, using the tap density
of Table 1. The above lower performance of MOF-5 in relation to both ACs, also shown for
hydrogen and methane at RT (Figures 12b and 13b), is maintained when the tap densities are
used. However, the use of a crystal density, results in a considerable artificial increase in its
volumetric uptake, becoming now the material with the best performance.
CO2 storage of at RT up to 4MPa, adsorbed as a subcritical gas, reaches a relative
pressure (P/Po) of 0.63 which means that a considerable porosity of the adsorbent is used.
Hence, from Figure 14, it can be noted: i) the adsorbent porosity development, controls CO2
adsorption capacity per unit of weight of adsorbent and per unit of volume; ii) in the latter
case, the porosity is not the only parameter that affects adsorption, since adsorbent density is
also important, and it increases with material density, iii) using tap density, the three
adsorbents behave as expected according to their porosity (MOF-5 < Max3 < AC1, and iv)
the highest adsorption capacity of MOFs compared to ACs, claimed by different authors, is
only related to the use of a higher density value (crystal density).

FINAL REMARKS
Volumetric Gas Storage Comparison with Best MOFs

MOFs with better performance than MOF-5, such as MOF-177 or MOF-210, have been
developed [29, 35, 52] but they are not easily available. Hence, we had to limit our
comparative ACs/MOFs study to MOF-5. Such comparison can be extended to other more
recently synthesized MOFs using their published volumetric storages. Because these data are
calculated using the crystal density of the adsorbents, we have also used them (Table 1) for
our three samples. Nevertheless, as commented several times over the chapter, its use gives
over estimated values because it does not take into account the inter particle spaces of the
adsorbents. Figure 15, presents the H2, CH4 and CO2 storing results of these five samples.

Figure 15. Comparative volumetric storages of H2 (77K) and CH4 and CO2 (RT) using crystal density
for two ACs (AC1 and Max3) and three MOFs (MOF-5, MOF-177 and MOF-210; redraw from [35])

Using comparable densities for both families of adsorbents (tap, as shown before) or in
Figure 15 crystal density, the claimed outstanding behavior of MOFs for storing gases, in
relation to other adsorbents (e.g. activated carbons), is not at all evident.
218 A. Linares-Solano, D. Cazorla-Amors, J. P. Marco-Lozar et al.

Samples Stability

For gas store applications, the adsorbent used has to have a suitable mechanical and
thermal stability as well as an appropriate chemical inertness. These characteristics, well
performed by most of the members of the porous carbon family, present some constrains in
MOFs family, related to the synthesis step, to its handling (air exposure) and to the presence
of humidity [85] that would require special dry conditions or pre-dried gas streams.
In order to analyze the stability of the selected materials, two samples (AC-1 and MOF-5)
were exposed to an atmosphere of air with a 70% relative humidity (RH) for 72 hours. Their
textural properties were analyzed before and after the exposure using N2 adsorption at 77K as
shown in Figure 16a. It can be seen that the exposure of AC1 to humidity has no effect on its
porosity. However, in the case of MOF-5, humidity causes total destruction of its porosity
[85]. Similar results were published for other MOF, such as MOF-177 [101], despite its very
interesting adsorption capacity.
Knowing the negative effect of humidity, the MOF-5 received was kept in a closed bottle
and handled with care trying to avoid as much as possible the exposure to air. However, after
a period of two years, we observed that the sample presents an important structural change
(its XRD pattern is completely different to the initial one) and has a noticeable change on its
textural porosity loosing most of its adsorption capacity, as it is shown in Figure 16b, and
hence its capacity to storage gases.

Figure 16. (a) N2 adsorption/desorption isotherms at 77K on AC-1 and MOF-5 samples before and after
exposing them to air with a 70% RH for 72 h and (b) after two year of the sample reception maintained
in a closed bottle.

CONCLUSIONS
This Chapter comparatively analyses the adsorption capacity of two activated carbons
and MOF-5 for storing gases (H2, CH4 and CO2) at different temperatures (77K and RT) and
pressures (from 0.1MPa to 20MPa). The results show that:

1) Independently of the type of adsorbents used, the storage occurs via a physisorption
process which is controlled by both the adsorbents characteristics (porosity and
density), by the nature of the gas and the experimental conditions used to store them.
High Pressure Gas Storage on Porous Solids 219

2) The lower is the relative pressure achieved (P/Pcs or P/Po) the more important will
be the contribution of the narrow microporosity (< 0.7nm) of the samples and this
narrow microporosity is higher for both ACs than for MOFs.
3) At higher relative pressures, most of the sample porosity will be used for the storage;
hence the micropore volume or the sample surface area will control its extension, as
it happens with H2 at 77K up to 4MPa.
4) Because the three samples have comparable surface area, their capacities to store
hydrogen are similar, especially if they are expressed per unit of weight.
5) When gravimetric results are converted to volumetric ones the inherent density of the
adsorbents (higher for both ACs than for MOF-5) will give higher storing capacity
results for both CAs than for MOF-5.
6) Because, the higher the density is, the higher will be the storage results, there is a
need of using similar densities when comparing the storage capacity of adsorbents.
7) Using a density measured similarly in the three samples (e.g. tap density), MOF-5
presents, for all gases and conditions studied lower adsorption capacities on
volumetric basis than ACs, due to its lower tap density value.
8) The outstanding adsorption capacities of MOFs in relation to ACs on volumetric
basis, frequently claimed in the literature, is due to the use of a higher density; the
crystal density which does not include the inter-particle space of the adsorbents.
9) Although the use of crystal density for gas storage over estimates the storage
capacity giving unrealistic high values, its use to compare the volumetric storage
capacity of different MOFs shows that MOF-5 is still one with the best performance,
although it has lower volumetric capacity than the two the ACs used in this study.

ACKNOWLEDGEMENTS
The authors thank BASF Chemical Company and Kansai Coke and Chemicals for
supplying MOF-5 and Maxsorb3000, respectively. The authors also thank the Spanish
MICINN (Project CTQ2009-10813/PPQ) and Generalitat Valenciana and FEDER (project
PROMETEO/2009/047) for their financial support.

REFERENCES
[1] G. D. Berry and S. M. Aceves, Energy and Fuels 12, 49 (1998).
[2] L. Schlapbach and A. Zuttel, Nature 414, 353 (2001).
[3] A. M. Seayad and D. M. Antonelli, Advanced Materials 16, 765 (2004).
[4] K. L. Lim, H. Kazemian, Z. Yaakob and W. R. W. Daud, Chemical Engineering and
Technology 33, 213 (2010).
[5] DOE Targets Light-Duty vehicles. 2010 [cited; Available from:
http://www1.eere.energy.gov/hydrogenandfuelcells/storage/pdfs/targets_onboard_hydro
_storage_explanation.pdf.
[6] D. Lozano-Castell, D. Cazorla-Amors, A. Linares-Solano and D. F. Quinn, Carbon
40, 989 (2002).
220 A. Linares-Solano, D. Cazorla-Amors, J. P. Marco-Lozar et al.

[7] D. Lozano-Castell, J. Alcaiz-Monge, M. A. Casa-Lillo, D. Cazorla-Amors and A.


Linares-Solano, Fuel 81, 1777 (2002).
[8] D. Lozano-Castell, D. Cazorla-Amors and A. Linares-Solano, Energy and Fuels 16,
1321 (2002).
[9] T. D. Burchell, ed. Carbon Materials for Advanced Technologies, Pergamon
Amsterdam (1999).
[10] E. i. Yasuda, M. Inagaki, K. Kaneko, M. Endo, A. Oya and Y. tanabe, eds. Carbon
alloys. Novel concepts to develop carbon science and technology, Elsevier, Amsterdam
(2003).
[11] A. P. Terzyk, P. A. Gauden and P. Kowalczyk, eds. Carbon Materials: Theory and
Practice, Research Signpost, Kerala (2008).
[12] [E. J. Bottani and J. M. D. Tascon, eds. Adsorption by Carbons, Elsevier, Amsterdam
(2008).
[13] R. E. Morris and P. S. Wheatley, Angewandte Chemie - International Edition 47, 4966
(2008).
[14] IPCC, Intergovernmental Panel on Climate Change. CO2 Capture and Storage. 2005.
[15] IEA Greenhouse Gas RandD Programme. CO2 Capture and Storage. 2006.
[16] J. E. Lennard-Jones, Transactions of the Faraday Society 28, 0333 (1932).
[17] M. Jord-Beneyto, F. Suarez-Garcia, D. Lozano-Castell, D. Cazorla-Amors and A.
Linares-Solano, Carbon 45 293 (2007).
[18] M. Jord-Beneyto, D. Lozano-Castell, F. Suarez-Garcia, D. Cazorla-Amors and A.
Linares-Solano, Microporous and Mesoporous Materials 112, 235 (2008).
[19] M. Jord-Beneyto, M. Kunowsky, D. Lozano-Castell, F. Suarez-Garcia, D. Cazorla-
Amors and A. Linares-Solano. Hydrogen Storage in Carbon Materials. In: A. P.
Terzyk, P. A. Gauden and P. Kowalczyk, eds. Carbon Materials : Theory and Practice.
Research Signpost, Kerala, 245 (2008).
[20] K. M. Thomas, Catalysis Today 120, 389 (2007).
[21] K. M. Thomas, Dalton Transactions 1487 (2009).
[22] R. von Helmolt and U. Eberle, Journal of Power Sources 165, 833 (2007).
[23] M. A. Casa-Lillo, F. Lamari-Darkrim, D. Cazorla-Amors and A. Linares-Solano,
Journal of Physical Chemistry B 106, 10930 (2002).
[24] D. Fraenkel, Journal of the Chemical Society-Faraday Transactions I 77, 2041 (1981).
[25] H. W. Langmi, A. Walton, M. M. Al Mamouri, S. R. Johnson, D. Book, J. D. Speight,
P. P. Edwards, I. Gameson, P. A. Anderson and I. R. Harris, Journal of Alloys and
Compounds 356, 710 (2003).
[26] M. G. Nijkamp, J. E. M. J. Raaymakers, A. J. van Dillen and K. P. de Jong, Applied
Physics A-Materials Science and Processing 72, 619 (2001).
[27] J. Weitkamp, M. Fritz and S. Ernst, International Journal of Hydrogen Energy 20, 967
(1995).
[28] P. Benard and R. Chahine, Scripta Materialia 56, 803 (2007).
[29] H. Furukawa and O. M. Yaghi, Journal of the American Chemical Society 131, 8875
(2009).
[30] M. Latroche, S. Surble, C. Serre, C. Mellot-Draznieks, P. L. Llewellyn, J. H. Lee, J. S.
Chang, S. H. Jhung and G. Ferey, Angewandte Chemie-International Edition 45, 8227
(2006).
High Pressure Gas Storage on Porous Solids 221

[31] A. G. Wong-Foy, A. J. Matzger and O. M. Yaghi, Journal of the American Chemical


Society 128, 3494 (2006).
[32] Y. Yurum, A. Taralp and T. N. Veziroglu, International Journal of Hydrogen Energy
34, 3784 (2009).
[33] V. Presser, M. Heon and Y. Gogotsi, Advanced Functional Materials 21, 810 (2011).
[34] H. Nishihara, P. X. Hou, L. X. Li, M. Ito, M. Uchiyama, T. Kaburagi, A. Ikura, J.
Katamura, T. Kawarada, K. Mizuuchi and T. Kyotani, Journal of Physical Chemistry C
113, 3189 (2009).
[35] H. Furukawa, N. Ko, Y. B. Go, N. Aratani, S. B. Choi, E. Choi, A. O. Yazaydin, R. Q.
Snurr, M. O'Keeffe, J. Kim and O. M. Yaghi, Science 329, 424 (2010).
[36] M. Eddaoudi, J. Kim, N. Rosi, D. Vodak, J. Wachter, M. O'Keefe and O. M. Yaghi,
Science 295, 469 (2002).
[37] G. Ferey, C. Mellot-Draznieks, C. Serre, F. Millange, J. Dutour, S. Surble and I.
Margiolaki, Science 309, 2040 (2005).
[38] H. Hayashi, A. P. Ct, H. Furukawa, M. O'Keeffe and O. M. Yaghi, Nature Materials
6, 501 (2007).
[39] H. Li, M. Eddaoudi, M. O'Keeffe and O. M. Yaghi, Nature 402, 276 (1999).
[40] O. M. Yaghi, M. O'Keeffe, N. W. Ockwig, H. K. Chae, M. Eddaoudi and J. Kim,
Nature 423, 705 (2003).
[41] T. Kyotani, T. Nagai, S. Inoue and A. Tomita, Chemistry of Materials 9, 609 (1997).
[42] Z. X. Ma, T. Kyotani and A. Tomita, Carbon 40, 2367 (2002).
[43] R. Ryoo, S. H. Joo and S. Jun, Journal of Physical Chemistry B 103, 7743 (1999).
[44] D. Zhao, D. J. Timmons, D. Yuan and H.-C. Zhou, Accounts of Chemical Research 44,
123 (2010).
[45] S. T. Meek, J. A. Greathouse and M. D. Allendorf, Advanced Materials 23, 249 (2011).
[46] J. Juan-Juan, J. P. Marco-Lozar, F. Suarez-Garcia, D. Cazorla-Amors and A. Linares-
Solano, Carbon 48, 2906 (2010).
[47] S. H. Sang, H. Furukawa, O. M. Yaghi and W. A. Goddard Iii, Journal of the American
Chemical Society 130, 11580 (2008).
[48] B. Xiao and Q. Yuan, Particuology 7, 129 (2009).
[49] Z. X. Yang, Y. D. Xia and R. Mokaya, Journal of the American Chemical Society 129,
1673 (2007).
[50] Y. H. Hu and L. Zhang, Advanced Materials 22, E117 (2010).
[51] R. Chahine and T. K. Bose, International Journal of Hydrogen Energy 19, 161 (1994).
[52] H. Furukawa, M. A. Miller and O. M. Yaghi, Journal of Materials Chemistry 17, 3197
(2007).
[53] J. L. C. Rowsell and O. M. Yaghi, Microporous and Mesoporous Materials 73, 3
(2004).
[54] A. Linares-Solano, D. Lozano-Castell, M. A. Lillo-Rdenas and D. Cazorla-Amors,
Chemistry and Physics of Carbon 30, 1 (2007).
[55] T. Kyotani, Carbon 38, 269 (2000).
[56] J. Lee, J. Kim and T. Hyeon, Advanced Materials 18, 2073 (2006).
[57] M. Hirscher and B. Panella, Journal of Alloys and Compounds 404, 399 (2005).
[58] B. Panella, M. Hirscher and S. Roth, Carbon 43, 2209 (2005).
[59] M. Kunowsky, J. P. Marco-Lozar, D. Cazorla-Amors and A. Linares-Solano,
International Journal of Hydrogen Energy 35, 2393 (2010).
222 A. Linares-Solano, D. Cazorla-Amors, J. P. Marco-Lozar et al.

[60] V. Fierro, A. Szczurek, C. Zlotea, J. F. Mareche, M. T. Izquierdo, A. Albiniak, M.


Latroche, G. Furdin and A. Celzard, Carbon 48, 1902 (2010).
[61] M. Hirscher and B. Panella, Scripta Materialia 56, 809 (2007).
[62] S. K. Bhatia and A. L. Myers, Langmuir 22, 1688 (2006).
[63] H. Wang, Q. Gao and J. Hu, Journal of the American Chemical Society 131, 7016
(2009).
[64] M. Schlichtenmayer, B. Streppel and M. Hirscher, International Journal of Hydrogen
Energy 36, 586 (2011).
[65] J. P. Marco-Lozar, F. Suarez-Garcia, D. Cazorla-Amors and A. Linares-Solano,
Carbon (2011).
[66] W. Zhou, H. Wu, M. R. Hartman and T. Yildirim, Journal of Physical Chemistry C
111, 16131 (2007).
[67] A. Dailly, J. J. Vajo and C. C. Ahn, Journal of Physical Chemistry B 110, 1099 (2006).
[68] A. R. Millward and O. M. Yaghi, Journal of the American Chemical Society 127,
17998 (2005).
[69] U. Mueller, M. Schubert, F. Teich, H. Puetter, K. Schierle-Arndt and J. Pastre, Journal
of Materials Chemistry 16, 626 (2006).
[70] B. Panella, M. Hirscher, H. Putter and U. Muller, Advanced Functional Materials 16,
520 (2006).
[71] E. Poirier, R. Chahine, P. Benard, L. Lafi, G. Dorval-Douville and P. A. Chandonia,
Langmuir 22, 8784 (2006).
[72] B. Schmitz, U. Muller, N. Trukhan, M. Schubert, G. Ferey and M. Hirscher, Chem.
Phys. Chem. 9, 2181 (2008).
[73] L. Zhou, Y. P. Zhou and Y. Sun, International Journal of Hydrogen Energy 29, 319
(2004).
[74] M. A. Richard, D. Cossement, P. A. Chandonia, R. Chahine, D. Mori and K. Hirose,
Aiche Journal 55, 2985 (2009).
[75] T. Otowa, Y. Nojima and T. Miyazaki, Carbon 35, 1315 (1997).
[76] D. Lozano-Castell, M. A. Lillo-Rdenas, D. Cazorla-Amors and A. Linares-Solano,
Carbon 39, 741 (2001).
[77] M. Dinca, A. Dailly, Y. Liu, C. M. Brown, D. A. Neumann and J. R. Long, Journal of
the American Chemical Society 128, 16876 (2006).
[78] J. Alcaiz-Monge, G. Trautwein, M. Perez-Cadenas and M. C. Roman-Martinez,
Microporous and Mesoporous Materials 126, 291 (2009).
[79] D. Cazorla-Amors, J. Alcaiz-Monge, M. A. Casa-Lillo and A. Linares-Solano,
Langmuir 14, 4589 (1998).
[80] F. Suarez-Garcia, E. Vilaplana-Ortego, M. Kunowsky, M. Kimura, A. Oya and A.
Linares-Solano, International Journal of Hydrogen Energy 34, 9141 (2009).
[81] E. Gadea-Ramos, F. Suarez-Garcia, D. Cazorla-Amors, M. Jord-Beneyto and A.
Linares-Solano, inventors; Method for measuring high-pressure absorption isotherms of
supercritical fluids and gases, involves determining mass balance for fluid or gas in
distributor/cell unit to calculate quantity of fluid or gas absorbed by sample patent
WO2008107505-A1. 2008.
[82] B. A. Younglove, Journal of Physical and Chemical Reference Data 11, 1 (1982).
[83] R. K. Agarwal and J. A. Schwarz, Carbon 26, 873 (1988).
[84] H. Frost and R. Q. Snurr, Journal of Physical Chemistry C 111, 18794 (2007).
High Pressure Gas Storage on Porous Solids 223

[85] S. S. Kaye, A. Dailly, O. M. Yaghi and J. R. Long, Journal of the American Chemical
Society 129, 14176 (2007).
[86] J. L. C. Rowsell, A. R. Millward, K. S. Park and O. M. Yaghi, Journal of the American
Chemical Society 126, 5666 (2004).
[87] K. S. Walton and R. Q. Snurr, Journal of the American Chemical Society 129, 8552
(2007).
[88] K. S. Walton, A. R. Millward, D. Dubbeldam, H. Frost, J. J. Low, O. M. Yaghi and R.
Q. Snurr, Journal of the American Chemical Society 130, 406 (2008).
[89] D. Cazorla-Amors, J. Alcaiz-Monge and A. Linares-Solano, Langmuir 12, 2820
(1996).
[90] R. Zacharia, D. Cossement, L. Lafi and R. Chahine, Journal of Materials Chemistry 20,
2145 (2010).
[91] M. M. Dubinin. Physical Adsorption of Gases and Vapors in Micropores. In: D. A.
Cadenhead, J. F. Danielli and M. D. Rosenberg, eds. Progress in Surface and
Membrane Science. Academic Press, New York, 1 (1975).
[92] V. Sebastin, J. Bosque, I. Kumakiri, R. Bredesen, A. Ansn, J. A. Maci-Agull, A.
Linares-Solano, C. Tllez and J. Coronas, Microporous and Mesoporous Materials.
[93] N. Texier-Mandoki, J. Dentzer, T. Piquero, S. Saadallah, P. David and C. Vix-Guterl,
Carbon 42, 2744 (2004).
[94] I. Cabria, M. J. Lopez and J. A. Alonso, Carbon 45, 2649 (2007).
[95] M. Rzepka, P. Lamp and M. A. Casa-Lillo, Journal of Physical Chemistry B 102,
10894 (1998).
[96] D. Lozano-Castell, D. Cazorla-Amors, A. Linares-Solano and D. F. Quinn, Carbon
40, 2817 (2002).
[97] A. C. Sudik, A. R. Millward, N. W. Ockwig, A. P. Cote, J. Kim and O. M. Yaghi,
Journal of the American Chemical Society 127, 7110 (2005).
[98] T. Duren, L. Sarkisov, O. M. Yaghi and R. Q. Snurr, Langmuir 20, 2683 (2004).
[99] C. F. Martin, M. G. Plaza, J. J. Pis, F. Rubiera, C. Pevida and T. A. Centeno,
Separation and Purification Technology 74, 225 (2010).
[100] M. G. Plaza, S. Garcia, F. Rubiera, J. J. Pis and C. Pevida, Chemical Engineering
Journal 163, 41 (2010).
[101] D. Saha and S. G. Deng, Journal of Physical Chemistry Letters 1, 73 (2010).
In: Coordination Polymers and Metal Organic Frameworks ISBN: 978-1-61470-899-5
Editors: O. L. Ortiz and L. D. Ramirez, pp. 225-247 2012 Nova Science Publishers, Inc.

Chapter 7

METAL-ORGANIC FRAMEWORKS FOR CO2


CAPTURE: WHAT ARE LEARNED
FROM MOLECULAR SIMULATIONS

Jianwen Jiang
Department of Chemical and Biomolecular Engineering,
National University of Singapore, 117576, Singapore.

ABSTRACT
CO2 capture is currently a topical issue in environmental protection and sustainable
development. This chapter reviews the recent molecular simulation studies for CO2
capture in metal-organic frameworks (MOFs). Emerged as an intriguing class of
nanoporous materials, MOFs have been considered versatile candidates for storage,
separation, catalysis and other widespread potential applications. However, the number of
MOFs synthesized to date is extremely large, experimentally testing and screening of
ideal MOFs for high-performance CO2 capture is formidable and time-consuming. With
ever-growing computational resources and advance in mathematical techniques,
molecular simulations have become an indispensable tool for materials characterization,
screening and design. At a molecular level, simulations can provide microscopic insights
from the bottom-up and establish structure-function relationships. Here, representative
simulation studies are summarized for CO2 capture in MOF sorbents and membranes
respectively, strategies (catenation, functionalization, exposed metals, ionic frameworks
and metal doping) are discussed for improving capture performance, and the effects of
water on CO2 capture are also considered. The chapter is concluded with the key insights
learned from simulations and the outlook for future endeavors.

1. INTRODUCTION
Fossil fuels (coal, oil and natural gas) supply over 85% of energy sources currently in use
worldwide [1]. The combustion of fossil fuels produces a huge quantity of CO2 emissions into

Tel: +65-65165083, Fax: +65-67791936, Email: chejj@nus.edu.sg.


226 Jianwen Jiang

the atmosphere, which is estimated to be about 30 gigatons per year. Since the industrial
revolution, CO2 concentration in the atmosphere has increased from 280 to 385 ppm and the
annual increase is at a high level of 1 2 ppm [2]. Carbon capture and sequestration (CCS) is
now not only a scientific interest, but a societal issue for environmental protection and
sustainable development. As a key step in CCS, CO2 is required to be captured and three
conceptual approaches have been proposed namely post-combustion, pre-combustion and
oxyfuel processes [3]. In a post-combustion process, CO2 is captured from exhaust flue gas
that consists of primarily N2 and 10 15 vol% CO2 at about 1 atm. The challenge in post-
combustion CO2 capture is associated with the low pressure and high flowrate of flue gas. In
a pre-combustion process, fuel is gasified and reacted with water to form shifted syngas rich
in H2 and CO2, and CO2 is separated before combustion. A limitation in pre-combustion CO2
capture involves the high-temperature of effluent gas and pre-cooling is required prior to
separation. In oxyfuel process, fuel is combusted with nearly pure O2 ( 95% purity) instead
of air to produce gas mixture containing mostly CO2 and H2O. The bottleneck in oxyfuel
process is the large-scale production of pure O2. In this chapter, the review will be focused on
post- and pre-combustion CO2 capture. On the other hand, natural gas is being consumed
more rapidly than any other fossil fuels. It is expected that the consumption of natural gas will
rapidly grow up to 153 trillion ft3 in 2030 [4]. Natural gas usually contains impurities such as
CO2 that reduces calorie content and needs to be separated.
A handful of technologies have been proposed for CO2 capture as illustrated in Figure 1
[5]. Among them, cryogenic distillation is energetically intensive due to the occurrence of
phase transition. Conventional amine scrubbing has been commercialized in the chemical
industry; however, solvent regeneration at 100 150 C consumes a large amount of energy.
Alternatively, separation based on sorbents and membranes has considerably high energy
efficiency, low capital cost, large separation capability and ease for scaling-up. Numerous
laboratory experiments have been carried out for CO2 capture in a wide variety of sorbents [6,
7] and membranes [8, 9], e.g., based on zeolites and metal-organic frameworks (MOFs).

Copyright 2008, Elsevier.

Figure 1. CO2 capture technologies: cost reduction benefit versus time to commercialization.
Reproduced with permission from ref. [5].
Metal-Organic Frameworks for CO2 Capture 227

In the last decade, MOFs have emerged as a new family of nanoporous materials [10].
MOFs are assembled by directly bonding metal clusters with organic linkers, thus there is no
need to use structure-directing agents as for zeolites. In contrast to the tetrahedral building
blocks in zeolites, MOFs can be synthesized from a large selection of inorganic clusters (e.g.
square-shaped, trigonal, tetrahedral and octahedral) and organic linkers (e.g. carboxylates,
imidazolates and tetrazolates). Consequently, MOFs possess a wide range of surface areas
and pore sizes, and have been considered versatile materials for storage, separation, catalysis
and biomedical applications [11-14]. Figure 2 demonstrates several typical MOFs, in which
IRMOF-1 (also called MOF-5) is a prototype MOF with internal cage illustrated by the large
sphere. Each oxide-centered Zn4O tetrahedron in IRMOF-1 is edge-bridged by six
carboxylate linkers resulting in an octahedral building unit, which reticulates into a three-
dimensional framework [15]. Tens of thousands of MOFs have been produced to date in this
burgeoning field and several (ZIF-8, Cu-BTC and MIL-53) are commercially available.
Enormous studies have been reported on the synthesis, characterization and applications
of MOFs. Figure 3 indicates the number of publications for MOFs increases rapidly in the
past few years. Among these, a large number of experiments have been conducted for CO2
capture [23-25]. Yaghi and coworkers synthesized two porous zeolitic imidazolate
frameworks (ZIF-95 and ZIF-100) that possess complex cages and selectively capture CO2
from several gas mixtures [19]. An et al. reported a cobalt adeninate MOF (bio-MOF-11) with
the Lewis basic amino and pyrimidine groups [20]. The bio-MOF-11 has a high heat of
adsorption for CO2, high CO2 capacity and exceptional selectivity for CO2 over N2. Panda et
al. synthesized a NH2 functionalized zeolitic tetrazolate framework (ZTF-1) [22]. Due to the
presence of free NH2 groups and uncoordinated tetrazolate nitrogen atoms, high CO2 and H2
uptake was observed. Caskey et al. evaluated the effects of metal substitution in a MOF with
cylindrical pores on surface area and CO2 uptake, and found Mg\DOBDC significantly
outperforms all other physisorptive materials for CO2 adsorption at low pressures [26].

Figure 2. Typical MOFs: IRMOF-1 [15], Cu-BTC [16], MIL-101 [17], rho-ZMOF [18], ZIF-100 [19],
bio-MOF-11 [20], cyclodextrin-MOF [21] and ZTF-1 [22]. Reproduced with permission.
228 Jianwen Jiang

Choi and Suh synthesized two flexible MOFs with highly selective CO2 adsorption over
N2, H2 and CH4 [27]. Gate opening and closing phenomena as well as hysteretic desorption
were observed in CO2 adsorption isotherms, which would allow efficient CO2 capture, storage
and sensing. In a rod packing MOF with highly polar CF3 groups pointing towards pore
surface, high adsorption of CO2 was reported by Hou et al [28]. Then based on single-
component adsorption data, the selectivities of CO2/CH4 and CO2/N2 were estimated.
Through breakthrough experiments, Couck et al. demonstrated that functionalizing MIL-53
by amino increases the selectivity of CO2/CH4 by orders of magnitude while maintaining a
very high capacity for CO2 capture [29].
While most experimental studies in MOFs for CO2 capture are based on MOF sorbents,
MOF membranes have recently been produced and tested for CO2 capture. Guo et al. used a
twin copper source technique to synthesize a copper-net supported Cu-BTC membrane, which
shows large permeability and excellent selectivity for H2 recycling [30]. A continuous MOF-5
membrane was prepared by Liu et al. on alumina substrate using in-situ solvothermal method
and the permeation of simple gases through the membrane was observed to follow Knudsen
behavior [31]. With a thermal seeding technique, Guerrero et al. anchored HKUST-1 crystal
on alumina support and found moderation separation of H2 over other gases such as CO2, N2,
O2 and CH4 [32]. Several zeolitic imidazolate framework (ZIF-7, ZIF-8 and ZIF-22)
membranes were fabricated by Caro and coworkers for H2 purification [33-36]. A molecular
sieving effect was observed in ZIF-7 and ZIF-22 because of their small apertures, which leads
to a high selectivity for H2.
The number of MOF materials synthesized to date is extremely large, thus experimentally
testing and screening of ideal MOFs for high-performance CO2 capture is a difficult and time-
consuming task. With ever-growing computational resources and advance in mathematical
techniques, molecular simulations have become an indispensable tool for materials science
and engineering. Simulations at an atomistic/molecular level provide microscopic insights
that are experimentally intractable, if not impossible, thus elucidate underlying physics from
the bottom-up. In addition, simulations can be used to secure fundamental interpretation of
experimental observations and establish structure-function relationships to guide practical
applications [37].

(Source: ISI Web of Science, 01 June 2011).

Figure 3. Number of publications for MOFs .


Metal-Organic Frameworks for CO2 Capture 229

While several reviews have been reported on simulation studies in MOFs [38-41], the
focus of this chapter is different and specifically on CO2 capture. Following the Introduction,
simulation methods are briefly introduced in Section 2. The representative simulation studies
examining MOF sorbents for CO2 capture are presented in Section 3, along with different
strategies to improve adsorption-based capture. In Section 4, simulation studies for CO2
capture in MOF membranes are illustrated. Furthermore, the effects of water on CO2 capture
are considered in Section 5. Finally, the conclusion and outlook are summarized in Section 6.

2. SIMULATION METHODS
The underlying information required in molecular simulation is intermolecular potentials,
which exclusively determine the accuracy and reliability of simulation results. Intermolecular
potentials are commonly referred to as a force field that consists of a set of potential functions
and parameters. The potential functions have bonded and nonbonded terms, where the bonded
term is decomposed into bond stretching, bending and torsional interactions; and the
nonbonded term includes short-ranged van der Waals and long-ranged electrostatic
interactions. The commonly used force fields for MOFs include universal force field (UFF)
[42] and DREIDING [43]. The parameters in a force field were usually derived from
experimental data over a limited range of conditions. With this type of force field, useful
information can be obtained for the properties that have been fitted; however, its semi-
empirical nature may lead to inaccurate or incorrect predictions for other properties, other
systems or at other conditions.
Quantum mechanics (QM) has been increasingly used to develop force fields as a
consequence of robust algorithms and powerful computers [44]. In principle, QM calculations
and molecular simulations can be performed simultaneously, that is, at each step of
simulation, interactions are computed onthe fly from QM [45]. However, this approach is
computationally extremely expensive and currently practical only for small systems. More
commonly used is a multiscale approach spanning different scales. First, the interaction
energies are calculated by QM at various positions and orientations, and subsequently fitted
by an analytic potential function. Then this analytical force field is used in simulation to
compute macroscopic properties. To accurately calculate interaction energies, an appropriate
level of QM theory is crucial. The most cost-effective method is density functional theory
(DFT); however, which has deficiency to describe the nonlocal London dispersion forces
induced by electron correlations. In contrast, wave function-based ab initio theories such as
coupled-cluster and Mller-Plesset perturbation methods are more accurate. Nevertheless,
they are computationally impracticable for large systems. As a compromise between accuracy
and efficiency, hybrid methods are often used [46].
With an appropriate force field for molecular interactions, simulation can be performed
using Monte Carlo (MC), molecular dynamics (MD), or a hybrid of MC and MD [47, 48].
MC simulation is a stochastic method to generate representative configurations. One
attractive aspect of MC simulation is that only energies rather than forces are evaluated
during configurational sampling. MC simulation can be performed with physically unnatural
motions, e.g. a jump from one position to the other or random insertion/deletion, and thus
significantly increasing efficiency. In particular, MC simulation is well-suited to calculate
230 Jianwen Jiang

thermodynamic properties for adsorption in MOFs. For example, commonly used grand-
canonical MC (GCMC) simulation with fixed temperature, volume and chemical potential
allows adsorbate molecules to exchange between adsorbed phase and bulk fluid reservoir,
thus adsorption isotherm can be readily generated. In a canonical ensemble with fixed
temperature, volume and number of molecules, Henrys constant, isosteric heat and binding
energy can be estimated by MC method. MD simulation follows Newtons second law of
motion and mimics the natural pathway of molecular motion to sample successive
configurations. The initial velocities of molecules at a given temperature are usually assigned
by the Maxwell-Boltzmann distribution. At each time, the forces between molecules are
calculated, then the equations of motion are solved numerically with a time step, and finally
the velocities and positions are updated. The time step is chosen to be sufficiently small so
that the total energy of system is conserved. MD simulation can be performed in an ensemble
with a constant volume, temperature or pressure. In addition to thermodynamic properties that
can be simulated by MC method, transport properties such as diffusivity and conductivity can
be determined from MD trajectory.
In most simulation studies for MOFs, the coordinates of framework atoms are adopted
from experimental crystallographic data and assumed to be rigid. This has two advantages:
there is no need to evaluate the intra-framework interactions, and the guest-framework
interactions can be pre-tabulated and simulation can be accelerated. Such a simplified
treatment is acceptable, particularly for small guest molecules (e.g. CO2, CH4, N2, H2) in
MOFs. For large molecules (e.g. benzene, xylene) or breathing MOFs, however, the
framework flexibility might be important and should be taken into account. The ordered
crystal structures of MOFs are mimicked using periodic boundary conditions, in which
simulation box is surrounded by its replicated images. In this regard, the simulation system
can be considered to be infinitely large. Simulation should run for sufficiently long to assure
equilibration or steady state is reached and then ensemble averages can be evaluated.

3. MOF SORBENTS FOR CO2 CAPTURE


Table 1 summarizes representative simulation studies for adsorption-based CO2 capture
in MOF sorbents at room temperature. We first highlight a few early studies in prototype
MOFs namely IRMOFs and Cu-BTC, and then discuss strategies to enhance capture
performance in catenated, functionalized, metal-exposed, ionic and metal-doped MOFs.
Yang and Zhong reported a simulation study toward molecular understanding for the
separation of CO2/CH4/H2 mixture in IRMOF-1 and Cu-BTC [49]. The results show that both
geometry and pore size largely affect separation efficiency and electrostatic interactions may
enhance the separation efficiency of gas mixtures composed of components with different
chemistries. Yang et al. also performed simulation on the separation of flue gas in Cu-BTC
[52]. They found that the electrostatic interactions between framework and adsorbates have a
large influence on the separation of CO2 and N2 with different quadrupole moments. Both
adsorbates at low pressures are mainly located in tetrahedral side pockets. Due to a greater
quadrupole moment, however, CO2 molecules are bound closer to the Cu atoms than N2
molecules. The simulation results suggest that Cu-BTC might be useful for the separation of
CO2 from flue gas.
Metal-Organic Frameworks for CO2 Capture 231

Table 1. MOF sorbents for CO2 capture

Material Mixture (ratio) Pressure (bar) Selectivity Reference


IRMOF-1 CO2/CH4 (50/50) 0 50 24 [49, 50]
CO2/H2 (50/50) 5 40
CO2/H2 (50/50) 0 50 80 150
Cu-BTC CO2/N2 (50/50) 0 20 20 35 [49, 51-54]
CO2/N2 (15/85) 0 50 20 30
CO2/CH4 (50/50) 0 50 37
IRMOF-13, IRMOF-14 CO2/CH4 (50/50) 0 50 3 5, 2 4 [53]
PCN-6, PCN-6 2 5, 2 4
IRMOF-11, -12, -13, -14 CO2/N2 (50/50) 0 20 4 20 [54]
IRMOF-9, -10, -11, -12, -13, -14 CO2/H2 0 20 5 100 [55]
IRMOFs and COFs CO2/CH4 (50/50) 0 20 17 [56]
CO2/H2 (50/50) 1 120
MIL-101 CO2/CH4 (50/50) 0 50 46 [57]
Zn(BDC)(TED)0.5 CO2/CH4 (50/50) 0 30 47 [58]
ZIF-8 and Zn(BDC)(TED)0.5 CO2/CH4 (50/50) 0 1.6 45 [59]
CO2/N2 (15/85) 02 10 12
ZIF-68 and ZIF-69 CO2/CH4 (50/50) 0 30 10 3 [60]
CO2/N2 (15/85) 30 13
Mixed-ligand MOFs CO2/CH4 (50/50) 0 15 4 30 [61]
Carborane-MOFs CO2/CH4 (50/50) 0 15 5 17 [62]
Cavity-modified MOFs CO2/N2 (50/50) 04 15 40 [63]
CO2/CH4 (50/50) 08 46
Diimide MOFs CO2/CH4 (50/50) 05 10 48 [64]
Co(II) carborane MOFs CO2/N2 (50/50) 0 10 10 100 [65]
CO2/CH4 (50/50) 5 50
Functionalized IRMOFs CO2/CH4 (50/50) 0 20 27 [66]
Functionalized PAFs CO2/H2 (15/85) 1 100 10 300 [67]
CO2/N2 (15/85) 3 90
CO2/CH4 (50/50) 2 25
bio-MOF-11 CO2/H2 (15/85) 0 30 230 375 [68]
CO2/N2 (15/85) 30 77
soc-MOF CO2/CH4 (50/50) 0 50 20 36 [53]
CO2/H2 (15/85) 0 30 300 600 [69]
CO2/H2 (15/85) 0 30 1.6 105 200
rho-ZMOF CO2/N2 (15/85) 1.9 104 100 [70]
CO2/CH4 (50/50) 3800 30
rht-MOF CO2/H2 (15/85) 0 50 40 60 [71]
+
Li -exchanged MOFs CO2/H2 (15/85) 0 50 2000 400 [72]
CO2/N2 (15/85) 300 50
Li-doped IRMOF-1 CO2/CH4 (10/90) 0 50 180 20 [73]
Metal-doped IRMOF-1 CO2/CH4 (10/90) 0 30 100 5 [74]

Babarao et al. predicted the adsorption and separation of CO2 and CH4 in silicalite,
IRMOF-1 and C168 schwarzite, which represent three different types of nanostructures [50].
The simulated adsorption isotherms and isosteric heats of pure gases agree well with available
experimental data. The predictions of mixture adsorption from the ideal-adsorbed-solution
theory (IAST) [75] are in accord with simulation results. Though IRMOF-1 has a
232 Jianwen Jiang

significantly higher adsorption capacity than silicalite and C168 schwarzite, the adsorption
selectivity of CO2 over CH4 is found to be close in the three adsorbents. Similarly, Liu and
Smit compared MOFs and zeolites for the separation of CO2/N2 mixture [54]. They selected
MFI with intersecting channels, DDR and LTA with cages, and seven MOFs (Cu-BTC, MIL-
47, IRMOF-1, -11, -12, -13 and -14) with a wide range of chemical compositions, pore sizes
and topologies. The simulation results reveal that although MOFs outperform over zeolites
for gas storage, their separation performance is comparable to zeolites. In addition, they
suggested that the difference in the quadrupole moments of sorbates should be considered in
selecting a suitable material for separation.

3.1. Catenated MOFs

It was reported from simulation that H2 adsorption in catenated IRMOFs can be


beneficial for improving H2 storage at cryogenic temperatures and low pressures [76]. A few
groups have investigated the effect of framework catenation on CO2 capture in MOFs. Jiang
and coworkers performed simulation to investigate the adsorption and separation of CO2/CH4
mixture in MOFs with catenated (IRMOF-13 and PCN-6) and non-catenated frameworks
(IRMOF-14 and PCN-6) [53]. Compared to IRMOF-14 and PCN-6, the catenated IRMOF-
13 and PCN-6 exhibit a greater extent of adsorption for CO2 at low pressures. This is
attributed to the formation of constricted pores by catenation, which induces a larger potential
overlap and a stronger affinity with adsorbate. However, the opposite is true at high pressures
due to the smaller pore volume in the catenated framework. As demonstrated in Figure 4, the
catenated PCN-6 has a higher selectivity for CO2/CH4 than the non-catenated PCN-6. This is
consistent with the recognized knowledge that adsorption separation is more efficient in a
structure with narrower pores. The authors also examined the adsorption of alkane isomer (C4
and C5) mixtures in IRMOF-13, -14 and PCN-6, -6, and observed similar behavior [77].

Copyright 2009, American Chemical Society.

Figure 4. Adsorption selectivity of CO2/CH4 mixture in PCN-6 and PCN-6. Reproduced with
permission from ref. [53].
Metal-Organic Frameworks for CO2 Capture 233

The enhanced selectivity as a consequence of catenation was also predicted by Liu and
Smit for CO2/N2 mixture in catenated IRMOF-11 and -13 and non-catenated IRMOF-12 and -
14 [54]. The predicted selectivity increases from 5 in IRMOF-12 and -14 to 20 in IRMOF-11
and -13. In a separate study, Zhong and coworkers examined the effect of catenation on the
separation of CO2/H2 mixture in three pairs of IRMOFs with and without catenation [55]. The
results show that CO2 selectivity in the catenated IRMOFs with multi-porous frameworks is
higher than that in the non-catenated counterparts. In addition, the electrostatic interactions
between adsorbate and framework are more pronounced in the catenated IRMOFs.
Nevertheless, these simulation studies discussed above reveal that the enhancement of
selectivity in catenated MOFs is not significant and catenation may not be a promising
strategy to improve CO2 capture.

3.2. Functionalized MOFs

Functionalization is a commonly used method in chemistry to tailor material properties


and has also been applied to MOFs. Zhong and coworkers predicted CO2/CH4 separation in
functionalized IRMOFs by substituting the hydrogen atoms in phenyl rings by CH3, F,
NH2 and OH groups [66]. As shown in Figure 5a, IRMOFNH2 and IRMOFOH exhibit the
highest selectivity, while IRMOFF and IRMOFCH3 are comparable to IRMOF-1 with only
a slight higher selectivity. This indicates that the introduction of a functional group with
strong electro-donating ability can strengthen electrostatic field and thus enhance selectivity.
However, the electro-withdrawing group F has a very weak effect. In IRMOF(XX)2 with
double substitutions, Figure 5b demonstrates that the selectivity in IRMOF(NH2)2 is higher
than in IRMOF(OH)2, due to the stronger electro-donating ability of NH2. The authors
further simulated the separation of CO2/CH4 mixture in IRMOF(NH2)n with multi-
substitutions (n = 14). The results in Figure 5c indicate the selectivity increases with
increasing number of NH2. Nevertheless, IRMOF(NH2)3 and IRMOF(NH2)4 have nearly
identical selectivity. This might be attributed to the counter-balance between the electrostatic
enhancement of NH2 and the steric hindrance effect of additional pendant NH2.

Copyright 2010, Elsevier.

Figure 5. Selectivities for the adsorption of equimolar CO2/CH4 mixture at 298 K in (a) IRMOF-1 and
IRMOFXX with single substitutions (XX = NH2, OH, F and CH3). (b) IRMOF(XX)2 with double
substitutions (XX = NH2 and OH). (c) IRMOF(NH2)n with different substituent number of NH2 groups
(n = 14). Reproduced with permission from ref. [66].
234 Jianwen Jiang

Babarao et al. computationally designed new porous aromatic frameworks (PAFs) by


introducing polar organic groups and then examined their separation capability toward CO2
[67]. The DFT calculations indicate that CO2 interacts the most strongly with ether O atoms,
especially in tetrahydrofuran. Among the functionalized PAFs considered, tetrahydrofuran-
like ether-PAF exhibits the highest adsorption capacity for CO2 at ambient conditions and
also the highest selectivities for CO2/CH4, CO2/N2 and CO2/H2 mixtures. It was found
electrostatic interactions play a dominant role in the predicted high CO2 selectivities in
functionalized PAFs.

8 MIL-101-OH
Selectivity (CO2/CH4)

7 MIL-101-NH2

6 MIL-101-CH3

MIL-101
4

3
0 20 40 60 80 100
P (kPa)

Figure 6. Selectivities of equimolar CO2/CH4 mixture in MIL-101 functionalized with OH, NH2
and CH3 groups.

Chen and Jiang investigated the adsorption of CO2/CH4 mixture in MIL-101


functionalized with OH, NH2 and CH3 groups [78]. MIL-101 is a chromium
terephthalate-based mesoscopic MOF and one of the most porous materials reported to date.
It is stable in air and boiling water, and its structure is not altered in organic solvents or
solvothermal conditions. These remarkable properties and the highly porous structure have
promoted considerable interest in MIL-101. The atomistic structure of MIL-101 was
constructed on the basis of experimental crystallographic data, energy minimization and
quantum mechanical optimization [57]. Figure 6 shows the predicted selectivity of CO2/CH4
in MIL-101 is enhanced upon introducing functional groups. In particular, polar OH and
NH2 groups have a larger enhancement than CH3 in the selectivity.

3.3. Metal-Exposed MOFs

Snurr and coworkers reported the separation of CO2/CH4 in carborane-based MOFs with
and without coordinately exposed metals [62]. The single-component adsorption isotherms
were measured at 298 K and fitted by a dual-site Langmuir-Freundlich model. Then, the
adsorption and selectivity were predicted using the IAST. In the pressure range from 0.1 to 15
bar, the selectivity for equimolar CO2/CH4 mixture was estimated to 5 17. By comparing
Metal-Organic Frameworks for CO2 Capture 235

the selectivities between MOFs with and without exposed metals, the results suggest that the
exposed metals in carborane-based MOFs can facilitate the separation of
(quadru)polar/nonpolar gas mixtures such as CO2/CH4. Also, the authors pointed out that the
carborane-based MOFs can be regenerated under milder conditions compared to zeolites, thus
requiring less expenditure of energy. In a separate study, they screened a diverse set of 14
MOFs for low-pressure CO2 capture from flue gas combining experimental and simulation
approach [79]. The results show that M/DOBDC (M = Zn, Mg, Co or Ni) with higher density
of exposed metals than HKUST-1 and UMCM-150 interact more strongly with CO2.
Particularly, Mg/DOBDC performs better because the relatively higher ionic character of
MgO bond promotes CO2 uptake. As illustrated in Figure 7, the CO2 uptake at 0.1, 0.5 and 1
bar correlates well with the isosteric heats; thus MOFs having a high density of exposed
metals are promising for CO2 capture.

Copyright 2009, American Chemical Society.

Figure 7. CO2 uptake and isosteric heats in screened 14 MOFs at 0.1, 0.5 and 1 bar. Reproduced with
permission from ref. [79].

Copyright 2010, American Chemical Society.

Figure 8. Real space Fourier difference scattering length density (yellow regions) in Mg-MOF-74 (color
code: Mg, blue; C, gray; O, red; and H, white), clearly indicating that adsorbed CO2 is located on top of
the exposed Mg atoms. Reproduced with permission from ref. [80].
236 Jianwen Jiang

Wu et al. combined neutron diffraction and first-principles calculations to investigate


CO2 adsorption in Mg-MOF-74 and HKUST-1 [80]. In both MOFs, the exposed metal ions
were identified to be the strongest binding sites through enhanced electrostatic interactions
between metal charges and CO2 multipole moments. Figure 8 illustrates the scattering length
density of CO2 in Mg-MOF-74. The adsorbed CO2 molecule is strongly attached through one
oxygen atom, while the other oxygen atom is relatively free. This high orientational disorder
contributes to the large apparent O-C-O bond bending, which was derived from diffraction
measurements. Nevertheless, only a small degree of bond bending was found by calculations,
suggesting that CO2 adsorption on the exposed metals is largely physisorption. Interestingly,
the overall metal-CO2 binding strength is right in the range for both CO2 adsorption and
desorption under typical flue gas conditions.

3.4. Ionic MOFs

In the continuous quest for novel MOFs to achieve high-performance gas adsorption and
separation, unique ionic MOFs have been produced. The presence of charge-balancing
nonframework ions in a nano-confined space can enhance the interactions with guest
molecules, which in turn increase adsorption capacity and selectivity.
Babarao and Jiang simulated the separation of CO2/CH4, CO2/N2 and CO2/H2 mixtures in
a zeolite-like MOF namely Na-rho-ZMOF [70]. As the first example of a 4-connected MOF
based on rho-zeolite, rho-ZMOF possesses a widely open anionic framework and charge-
balancing doubly protonated 1,3,4,6,7,8-hexahydro-2H-pyrimido[1,2-a]pyrimidine (HPP)
[18]. The HPP ions can be exchanged with other cations, e.g., Na+ ions. The simulation
results in Na-rho-ZMOF reveal that CO2 is adsorbed predominantly over other gases due to
the strong electrostatic interactions with Na+ ions and ionic framework. Figure 9a
demonstrates the typical locations of CO2 molecules in the eight-member ring of framework
for CO2/CH4 mixture adsorption at a total pressure of 500 kPa. CO2 molecules bind
preferentially to Na+ ions, as demonstrated by the radial distribution functions in Figure 9b.
With increasing pressure (data not shown), Na+ ions are coordinated by more CO2 molecules,
i.e., increasingly solvated by surrounding CO2 molecules. While the distance between Na+
and CO2 remains more or less constant with increasing pressure, the distance between CO2
molecules becomes shorter. The predicted CO2/CH4 selectivity shown in Figure 9c is
significantly higher than that in most MOFs and other nanoporous materials. By switching off
the charges in Na-rho-ZMOF, however, the selectivity in neutral framework decreases
substantially. This suggests that the electrostatic interactions play a pivotal role in the
extremely high selectivity observed in Na-rho-ZMOF.
Jiang and coworkers further systematically investigated the separation of CO2/H2 mixture
in three ionic MOFs (rho-ZMOF, soc-MOF and rht-MOF) [69-71]. As for CO2/CH4 mixture
in rho-ZMOF, the predicted CO2/H2 selectivities are much higher than in non-ionic MOFs.
Particularly noteworthy is that the selectivity behaves differently in the three ionic MOFs
with different free volumes and charge densities. The fractional free volume increases in the
order of rho-ZMOF < soc-MOF < rht-MOF, but the charge density increases in the opposite
order. Consequently, rho-ZMOF has the smallest porosity and largest charge density, and
thus exhibits the highest selectivity, followed by soc-MOF and rht-MOF.
Metal-Organic Frameworks for CO2 Capture 237

Copyright 2009, American Chemical Society.

Figure 9. Adsorption of equimolar CO2/CH4 mixture in Na-rho-ZMOF. (a) Locations of CO2 molecules
in the eight-member ring at a total pressure of 500 kPa. Na+ ions and CO2 molecules are represented by
balls and sticks, respectively. (b) Radial distribution functions between Na+ ions and adsorbates. (c)
Selectivities of CO2/CH4 in ionic and neutral frameworks. Reproduced with permission from ref. [70].

In addition, Jiang and coworkers simulated the adsorption of CO2/H2 mixture in rho-
ZMOFs exchanged with various cations including Na+, K+, Rb+, Cs+, Mg2+, Ca2+ and Al3+
[81]. Figure 10 shows the selectivities of CO2/H2 in the seven cation-exchanged rho-ZMOFs
versus total pressure. At a given pressure, the selectivity increases as Cs+ < Rb+ < K+ < Na+ <
Ca2+ < Mg2+ Al3+. This is because the electric field of cation increases as the charge-to-
diameter ratio becomes larger and thus the interaction with CO2 is enhanced. With increasing
pressure, the selectivity in each rho-ZMOF decreases sharply as a consequence of two
factors. First, the adsorption sites in rho-ZMOF are heterogeneous and adsorbate molecules
start to occupy less favorable sites at high pressures. Second, H2 is smaller than CO2 and can
pack into the partially filled cages more easily with increasing pressure. The predicted
selectivities of CO2/H2 mixture in rho-ZMOFs with various cations are substantially higher
200 than in other porous materials. Although the ionic MOFs considered are predicted to be good
(a) candidates for CO2 capture, experimental
250 (b) confirmation is desired to ascertain their high-
performance in practice.
150
200

105
150
NCO2

NCO2

+
100 Na
+
K
+
Na
+
100 Rb
+
K +
Cs
50 Rb
+
2+
Cs
+ 104 Mg
Mg
2+ 50 2+
Ca
2
SCO /H

2+
Ca 3+
Al
2

3+
Al
0 0
0 20 40 60 80 100 0 500 1000 1500 2000 2500 3000
P (kPa) 103 P (kPa)

102
1 10 100 1000
Ptotal (kPa)

Figure 10. Selectivities of CO2/H2 mixture in Na+, K+, Rb+, Cs+, Mg2+, Ca2+ and Al3+-exchanged
rho-ZMOFs.
238 Jianwen Jiang

3.5. Metal-Doped MOFs

Metal-doped MOFs have also been proposed for CO2 capture. Zhong and coworkers
estimated the separation of CO2/CH4 mixtures in Li-modified MOF-5 [73]. They considered
both physical and chemical approaches of Li doping. In the former, Li atoms were adsorbed
onto organic linkers behaving as non-framework atoms; in the latter, Li atoms were
introduced as part of framework by modifying organic linkers. As shown in Figure 11a, the
adsorption selectivities of CO2/CH4 in both physically and chemically doped MOF-5 were
predicted to be much higher than in MOF-5. This indicates that metal doping can significantly
improve the separation performance of CO2/CH4 mixture due to the enhancement of
electrostatic potentials. Comparing the two doping approaches, physical doping appears to
perform better at low pressures, while the opposite is true at high pressures. Based on the
electrostatic potential contours, the authors found that the regions with enhanced electrostatic
potentials are different in the two approaches. More specifically, the enhancement in physical
doping occurs around the metal atoms, while it is around the organic linkers in chemical
doping. As a consequence, the preferential adsorption sites for CO2 differ in these two cases.
It was further predicted that an increase in the number of doped Li atoms leads to a larger
enhancement in electrostatic potentials, and hence a higher selectivity.
They further simulated the separation of CO2/CH4 in IRMOF-1 doped with nine metals
(Li, Na, K, Rb, Cs, Mg, Ca, Al and Ga) [74]. The periodic DFT calculations indicate the
optimized structures of these metal doped IRMOFs appear to be stable. As illustrated in
Figure 11b, the selectivities in IRMOF-1-Na, -Li, -K, -Rb and -Cs are enhanced compared to
IRMOF-1, and the enhancement becomes weaker with increasing metal radius. However, the
selectivities in IRMOF-1-Ca, -Al and -Ga are slightly lower than in IRMOF-1. This suggests
that the doping of alkali metals can significantly enhance selectivity, while the influence of
alkaline earth metals and boron group metals is marginal. The work also demonstrates that the
Lennard-Jones interactions of metals should be incorporated with electrostatic interactions,
while considering metal-doping strategy to improve the separation performance of MOFs.

(a) Copyright 2010, Royal Society of Chemistry. (b) Copyright 2011, Elsevier.

Figure 11. Selectivity of CO2/CH4 in (a) physically and chemically doped MOF-5. Reproduced with
permission from ref. [73]. (b) IRMOF-1 doped with various metals. Reproduced with permission from
ref. [74].
Metal-Organic Frameworks for CO2 Capture 239

Lan et al. combined QM calculations and GCMC simulations to examine CO2 capture in
COFs doped by alkali (Li, Na and K), alkaline-earth (Be, Mg and Ca) and transition (Sc and
Ti) metals [82]. The results indicate that Li, Sc and Ti can bind with COFs stably, while Be,
Mg and Ca cannot because the binding of Be, Mg and Ca with COFs is very weak.
Furthermore, Li, Sc and Ti can significantly improve CO2 adsorption in COFs. However, the
binding energy of CO2 molecule with Sc and Ti exceeds the lower limit of chemisorption and
thus suffers from the difficulty of desorption. They concluded that Li is the best surface
modifier of COFs for CO2 capture among all the metals studied. As illustrated in Figure 12,
the CO2 uptake in Li-doped COF-102 and COF-105 at 298 K and 1 bar reaches 409 and 344
mg/g, respectively, which are approximately eight and four times those in the non-doped
counterparts.

Copyright 2010, American Chemical Society.

Figure 12. (a) Adsorption of a single CO2 molecule in Li-doped COF. (b) Excess gravimetric uptake of
CO2 in Li-doped COFs at 298 K. Reproduced with permission from ref. [82].

4. MOF MEMBRANES FOR CO2 CAPTURE


The current research on MOF membranes for gas separation is in an infant stage but very
active [83]. Table 2 highlights the simulation studies for permeation-based CO2 capture in
MOF membranes. To calculate permeation in a membrane, both equilibrium and dynamic
properties are required. Therefore, it is more challenging and time-consuming to examine
permeation-based separation than adsorption-based, and fewer simulation studies have been
reported for CO2 capture in MOF membranes.
Keskin and Sholl tested the performance of IRMOF-1 as a membrane for CO2/CH4
separation [84]. The single-component results predict that IRMOF-1 membrane would show a
strong ideal selectivity for CH4 in CO2/CH4 mixture. However, the predictions for mixture
permeation suggest that IRMOF-1 gives at best a weak selectivity for CO2. This study
indicates that mixture effects play a crucial role in determining the performance of IRMOF-1
membrane under conditions relevant for natural-gas separation, and suggests that modeling or
experimental studies that examine only single-component gases are insufficient to understand
the properties of MOF membranes in practical applications. They further evaluated the
performance of chemical diverse MOFs (IRMOF-8, -9, -10, -14 and Zn(BDC)(TED)0.5) for
membrane separation of CO2/CH4 and CO2/H2 mixtures [87]. In all the MOF membranes
examined, the separation of CO2 from CH4 or H2 is not favorable due to low CO2 selectivities.
240 Jianwen Jiang

Table 2. MOF membranes for CO2 capture

Material Mixture (ratio) Pressure (bar) Selectivity Reference


IRMOF-1 CO2/CH4 (50/50) 0 50 13 [84, 85]
CO2/N2 (50/50) 0 100 29 [86]
CO2/H2 (50/50) 0 100 18
Cu-BTC CO2/CH4 (50/50) 0 100 16 [87]
CO2/H2 (10/90) 5 40
IRMOF-1, 8, -9, -10, -14, CO2/CH4 (50/50) 0 100 14 [87]
COF-102 CO2/H2 (10/90) 0.5 2
bio-MOF-11 CO2/CH4 (10/90) 0 50 35 [88]
CO2/H2 (1/99) 20 60
ZIF-3 and ZIF-10 CO2/CH4 (10/90) 0 50 12 [89]
CO2/H2 (1/99) 45
MgMOF-74, ZnMOF-74 CO2/CH4 (50/50) 1 20 1 2000 [90]
MOF-177, ZIF-8 CO2/N2 (15/85) 2 500
CO2/H2 (15/85) 0.6 150

Babarao and Jiang compared the diffusion and separation of CO2 and CH4 in silicalite,
C168 schwarzite and IRMOF-1 [85]. As loading increases, the self-diffusivities in the three
frameworks decrease due to steric hindrance; the corrected diffusivities remain nearly
constant or decrease approximately in a linear manner depending on the type of adsorbate and
framework; the transport diffusivities generally increase except for CO2 in IRMOF-1. The
correlation effects were found to reduce in the order of MFI > C168 > IRMOF-1, opposite to
the increasing hierarchy of porosity (MFI < C168 < IRMOF-1) in the three frameworks. The
predicted self-, corrected, and transport diffusivities of pure CO2 and CH4 from the Maxwell-
Stefan formulation match well with simulation results. The permselectivity of CO2/CH4 is
marginal in IRMOF-1, slightly enhanced in MFI, and greatest in C168 schwarzite.
Krishna and van Baten compared the permeation selectivities Sperm of CO2/CH4, CO2/N2
and CO2/H2 mixtures across various MOF and zeolites membranes [90]. At a total pore
concentration equal to that at the upstream membrane face, they first calculated the adsorption
selectivities Sadsp and diffusion selectivities Sdiff, and then multiplied them to estimate Sperm.
Therefore, the Sperm reflect the balance between adsorption and diffusion in different
structures. As shown in Figure 13, the highest Sperm for CO2/CH4 mixture with values > 100
are observed in zeolites with 8-ring windows such as ERI, DDR and CHA. In these cases,
Sadsp and Sdiff complement each other. For permeation across MgMOF-74 and ZnMOF-74,
Sperm is higher in the former despite a lower Sdiff. This indicates that the higher Sadsp in
MgMOF-74 compensates more on the lower Sdiff. Further discussed by the authors, the
situations are different for CO2/N2 and CO2/H2 mixtures.
Hu and Jiang investigated H2 purification by a ZIF-7 membrane using non-equilibrium
MD simulation [91]. Figure 14 illustrates the simulation system which contains two
chambers: a five-component gas mixture (CO2:H2:CO:CH4:H2O = 15:74:5:5:1) on the left and
a vacuum on the right. The gas mixture mimics a typical effluent gas in syngas production.
The two chambers are separated by a ZIF-7 membrane. Two graphene plates are added into
the two chambers: the left one is exerted by a driving pressure Pext and the right one is fixed.
The simulation study demonstrates that ZIF-7 membrane can act as an ultra-selective
membrane only allowing H2 to permeate, which is attributed to the molecular sieving effect of
small apertures in ZIF-7 and the repellence of hydrophobic pores to H2O. With increasing
Metal-Organic Frameworks for CO2 Capture 241

Pext, the flux and permeance of H2 increase and the activation energy of H2 permeation
decreases. They also pointed out a few limitations in the MD simulation used and suggested
possible improvements.

Copyright 2011, Royal Society of Chemistry.

Figure 13. Permeation selectivities for CO2/CH4 mixture permeation across various MOF and zeolite
membranes. Reproduced with permission from ref. [90].

Figure 14. H2 purification by a ZIF-7 membrane. There are two chambers: a five-component gas
mixture (CO2:H2:CO:CH4:H2O = 15:74:5:5:1) on the left and a vacuum on the right. The two chambers
are separated by a ZIF-7 membrane. Two graphene plates are added into the two chambers: the left one
is exerted by a driving pressure Pext and the right one is fixed. Color codes: Zn, magenta; N, blue; C,
cyan; H, white; O, red; H2, green; graphene, yellow.
242 Jianwen Jiang

5. EFFECTS OF H2O ON CO2 CAPTURE


Gas mixtures usually contain moisture that may adversely affect separation. For example,
the Henrys constant of CO2 in cationic zeolite-X was found to drop exponentially with
increasing amount of H2O [92]. Jiang and coworkers have systematically examined the
effects of H2O on adsorption and separation in various neutral and ionic MOFs. Four
intriguing cases were observed depending on the nature of framework.

(1) In hydrated MIL-101, terminal H2O molecules enhance the adsorption of CO2 and
CH4 at low pressures and the enhancement is greater for CO2. The reason is that the
terminal H2O molecules act as additional adsorption sites and CO2 interacts more
strongly with them than CH4. At high pressures, however, the terminal H2O
molecules reduce the free volume of MIL-101 and lead to a smaller adsorption
capacity compared to the dehydrated MIL-101. The adsorption selectivity of
CO2/CH4 is slightly higher in the hydrated MIL-101 [57]. Similar phenomena were
observed by Yazaydin et al. for the adsorption of CO2 and CH4 as well as the
separation of CO2/CH4 in Cu-BTC with H2O molecules coordinated to the exposed
Cu atoms [93]. They found that the interactions between CO2 and the electric field
created by H2O molecules are responsible for the enhanced CO2 uptake. Liu and Smit
also observed an increase in the selectivities of CO2/CH4 and CO2/N2 mixtures with
increasing H2O amount in ZIF-68 and -69 [60].
(2) In Zn(BDC)(TED)0.5 (BDC = benzenedicarboxylate and TED = triethylenediamine),
H2O adsorption is vanishingly small as observed from both experiment and
simulation [58]. This is attributed to the highly hydrophobic nature of
Zn(BDC)(TED)0.5. With the BDC and TED linkers surrounding metal oxides,
Zn(BDC)(TED)0.5 interacts very weakly with H2O. Consequently, the adsorption and
selectivity of CO2/CH4 mixture remain nearly the same in the absence or presence of
H2O. This implies that a prewater treatment is not required prior to CO2/CH4
separation in such a MOF.
(3) In ionic Na-rho-ZMOF, the effects of H2O on adsorption and selectivity of CO2/CH4
mixture are significant [94]. Figure 15a shows the locations of CO2 and H2O
molecules in the eight-member ring for CO2/CH4/H2O (composition 50/50/0.1)
mixture at 500 kPa. A large number of H2O molecules are observed to surround Na+
ions. Compared to Figure 9a, the number of CO2 molecules in Figure 15a is fewer
and the distance from CO2 to Na+ is longer, changing from 2.5 2.6 to 4.0 4.4 .
This suggests the interaction between CO2 and Na+ is reduced in the presence of
H2O, because H2O is highly polar and interacts with Na+ ions much more strongly
than CO2 and thus has a substantial influence on CO2 adsorption. A very high peak is
observed in the radial distribution function of Na+H2O (Figure 15b), indicating a
strong affinity between Na+ and H2O. However, no peak is observed here for Na+
CO2, which is in contrast to Figure 9b. These further reveal that H2O competes the
adsorption sites with CO2 and leads to a pronounced reduction in CO2 adsorption.
With 0.1% of H2O present, CO2 adsorption drops substantially, while CH4 adsorption
is not discernibly affected. Consequently, the selectivity of CO2/CH4 is reduced
approximately by one order of magnitude (Figure 15c). Similar effects of H2O on
Metal-Organic Frameworks for CO2 Capture 243

CO2/H2 adsorption and separation were also observed in rht-MOF [71]. It is therefore
important to remove H2O before CO2 capture in these ionic MOFs.
(4) Despite similar ionic framework, soc-MOF exhibits different behavior from rho-
ZMOF and rht-MOF [69]. The nonframework ions NO3 in soc-MOF are located in
carcerand-like capsules, which are connected via narrow windows with dimensions
of 7.651 5.946 , and adsorbate molecules can enter only at high pressures [95].
With a small amount H2O present, the selectivity of CO2/H2 in soc-MOF increases at
low pressures as a consequence of the promoted adsorption of CO2 by H2O bound
onto exposed indium atoms. This is because H2O interacts preferentially with the
readily accessible indium atoms rather than the nonframework NO3 ions in the
capsules, and thus the indium atoms act as additional sites for CO2 adsorption. With
increasing H2O at high pressures, however, H2O competitively replaces CO2 and the
selectivity decreases.

Copyright 2009, Royal Society of Chemistry.

Figure 15. Adsorption of CO2/CH4/H2O (composition 50:50:0.1) in Na-rho-ZMOF. (a) Locations of


CO2 and H2O molecules in the eight-member ring at total pressure 500 kPa. Na+ ions are represented by
balls, CO2 and H2O molecules are represented by sticks. (b) Radial distribution functions between Na+
ions and adsorbates. (c) Selectivity of CO2/CH4. Reproduced with permission from ref. [94].

CONCLUSION AND OUTLOOK


This chapter reviews representative, rather than exhaustive, molecular simulation studies
for CO2 capture in MOFs. Several strategies are discussed to improve adsorption-based CO2
capture in MOF sorbents, such as catenation, functionalization, exposed metals, ionic
frameworks and metal doping. Simulation results reveal that the first three strategies
(catenation, functionalization and exposed metals) can enhance the separation factor of CO2
capture slightly or moderately, whereas the latter two (ionic frameworks and metal doping)
appear to be more efficient. Compared to MOF sorbents, the simulation studies for
permeation-based CO2 capture in MOF membranes are fewer. However, it is expected that
more simulation studies will be reported along this direction, as evidenced by increasing
experimental efforts. Furthermore, different effects of H2O on CO2 capture are unraveled by
simulations depending on MOF structures. Molecular simulations provide atomic-resolution
and time-resolved insights, which are essential in securing quantitative interpretation of
experimental results. With microscopic understanding, MOFs can be screened and designed
244 Jianwen Jiang

from the bottom-up for CO2 capture and other practical applications without resort to time-
consuming experiment testing, and thus expediting material and process development.
The current state of knowledge for CO2 capture in MOFs has been developed
progressively. Nevertheless, several challenges exist for future simulation endeavors in this
rapidly evolving field. (1) The accuracy of simulation is exclusively based on the force field
used. In most simulation studies for MOFs, empirical force fields have been used. These force
fields were constructed by fitting to the limited experimental data with certain empirical rules
and their semi-empirical nature may lead to inaccurate predictions. A more rational way is to
calculate guest-MOF interactions from QM methods [96]. However, the main obstacle is that
high-level methods are usually required but computationally extremely expensive.
Consequently, a cost-effective method is required as a compromise between accuracy and
speed. (2) Most simulation studies have used rigid structures for MOFs. This allows the use
of grid-interpolation technique to compute guest-framework interactions very efficiently and
is usually acceptable, but cannot reproduce structural change that might occur upon
adsorption, e.g., in flexible MOFs [97]. A force field that accounts for the flexibility of MOFs
should be used instead. While simulation studies have been reported on specific flexible
MOFs, a general force field is currently unavailable for MOFs and needed to be developed.
(3) Although the effects of H2O on CO2 capture in MOFs have been analyzed by simulations,
it is indispensable to evaluate the chemical stability of MOFs under moisture. A large number
of MOFs are unstable in water or atmosphere, which will impede their use for CO2 capture
and other applications (e.g. water desalination and biofuel purification).
In addition, the thermal stability of MOFs should also be taken into consideration in
practical applications [98], e.g. in high-temperature pre-combustion CO2 capture. Molecular
simulations are useful to unravel the chemical and thermal properties of MOFs from a
microscopic point of view, and thus provide guidelines on the selection of suitable building
blocks to synthesize stable MOFs. (4)
Another important issue is the mechanical properties of MOFs. In CO2 capture, the
pressure exerted on a separation system (MOF sorbent or membrane) may distort MOF
structure and lead to a deteriorated performance. It is crucial to understand quantitatively how
pressure affects the pore geometries, framework dimensionalities and selectivities of MOFs
[99]. This topic has received limited attention, but simulations can play an important role to
elucidate the structure-mechanical property relationships of MOFs. The challenges outlined
above provide new opportunities for in-depth molecular simulation investigations towards the
practical applications of MOFs for high-performance CO2 capture.

ACKNOWLEDGMENTS
The author acknowledges his coworkers and collaborators for their contributions to some
of the studies summarized in this chapter, and the National University of Singapore, the
Ministry of Education of Singapore, and the Singapore National Research Foundation for
support (R-279-000-198-112/133, R-279-000-243-123, R-279-000-238-112, R-279-000-297-
112 and R-279-000-261-281).
Metal-Organic Frameworks for CO2 Capture 245

REFERENCES
[1] Facing the Hard Truths about Energy: A Comprehensive View to 2030 of Global Oil
and Natural Gas, U.S. Department of Energy, Washington D.C., 2007.
[2] Doney, S. C.; Fabry, V. J.; Feely, R. A.; Kleypas, J. A. Annu. Rev. Mar. Sci. 2009, 1,
169.
[3] IPCC Special Report on Carbon Dioxide Capture and Storage, Intergovernmental
Panel on Climate Change, 2005.
[4] International Energy Outlook Report, U.S. Department of Energy, Washington D.C.,
2010.
[5] Figueroa, J. D.; Fout, T.; Plasynski, S.; Mcllvried, H.; Srivastava, R. D. Int. J.
Greenhouse Gas Control 2008, 2, 9.
[6] Choi, S. H.; Drese, J. H.; Jones, C. W. Chem. Sus. Chem. 2009, 2, 796.
[7] D'alessandro, D. M.; Mcdonald, T. Pure Appl. Chem. 2011, 83, 57.
[8] Shelley, S. Chem. Eng. News 2009, 105, 42.
[9] Brunetti, A.; Scura, F.; Barbieri, G.; Drioli, E. J. Membr. Sci. 2010, 359, 115.
[10] Yaghi, O. M.; O'keefe, M.; Ockwig, N. W.; Chae, H. K.; Eddaoudi, M.; Kim, J. Nature
2003, 423, 705.
[11] Ferey, G. Chem. Soc. Rev. 2008, 37, 191.
[12] Long, J. R.; Yaghi, O. M. Chem. Soc. Rev. 2009, 38, 1213.
[13] Coma, A.; Garcia, H.; Xamena, F. X. L. Chem. Rev. 2010, 110, 4606.
[14] Meek, S. T.; Greathouse, J. A.; Allendorf, M. D. Adv. Mater. 2011, 23, 249.
[15] Li, H.; Eddaoudi, M.; O'keeffe, M.; Yaghi, O. M. Nature 1999, 402, 276.
[16] Chui, S. S.-Y.; Lo, S. M.-F.; Charmant, J. P. H.; Orpen, A. G.; Williams, I. D. Science
1999, 283, 1148.
[17] Ferey, G.; Mellot-Draznieks, C.; Serre, C.; Millange, F.; Dutour, J.; Surble, S.;
Margiolaki, I. Science 2005, 309, 2040.
[18] Liu, Y. L.; Kravtsov, V. C.; Larsen, R.; Eddaoudi, M. Chem. Commun. 2006, 1488.
[19] Wang, B.; Cote, A. P.; Furukawa, H.; O'keeffe, M.; Yaghi, O. M. Nature 2008, 453,
207.
[20] An, J.; Geib, S. J.; Rosi, N. L. J. Am. Chem. Soc. 2010, 132, 38.
[21] Smaldone, R. A.; Forgan, R. S.; Furukawa, H.; Gassensmith, J. J.; Slawin, A. M. Z.;
Yaghi, O. M.; Stoddart, J. F. Angew. Chem. Int. Ed. 2010, 46, 8630.
[22] Panda, T.; Pachfule, P.; Chen, Y. F.; Jiang, J. W.; Banerjee, R. Chem. Commun. 2011,
47, 2011.
[23] D'alessandro, D. M.; Smit, B.; Long, J. R. Angew. Chem. Int. Ed. 2010, 49, 6058.
[24] Lin, X.; Champness, N. R.; Schroder, M. Top. Curr. Chem. 2010, 293, 35.
[25] Keskin, S.; Van Heest, T. M.; Sholl, D. S. ChemSusChem. 2010, 3, 879.
[26] Caskey, S. R.; Wong-Foy, A. G.; Matzger, A. J. J. Am. Chem. Soc. 2008, 130, 10870.
[27] Choi, H. S.; Suh, M. P. Angew. Chem. Int. Ed. 2009, 48, 6865.
[28] Hou, L.; Shi, W. J.; Guo, Y.; Jin, C.; Shi, Q. Z. Chem. Commun. 2011, 47, 5464.
[29] Couck, S.; Denayer, J. F. M.; Baron, G. V.; Remy, T.; Gascon, J.; Kapteijn, F. J. Am.
Chem. Soc. 2009, 131, 6326.
[30] Guo, H. L.; Zhu, G. S.; Hewitt, I. J.; Qiu, S. L. J. Am. Chem. Soc. 2009, 131, 1646.
[31] Liu, Y. Y.; Ng, Z. F.; Khan, E. A.; Jeong, H. K.; Ching, C. B.; Lai, Z. P. Micro. Meso.
Mater. 2009, 118, 296.
246 Jianwen Jiang

[32] Guerrero, V. V.; Yoo, Y.; Mccarthy, M. C.; Jeong, H. K. J. Mater. Chem. 2010, 20,
3938.
[33] Bux, H.; Liang, F. Y.; Li, Y. S.; Cravillon, J.; Wiebcke, M.; Caro, J. J. Am. Chem. Soc.
2009, 131, 16000.
[34] Li, Y. S.; Liang, F. Y.; Bux, H.; Feldhoff, A.; Yang, W. S.; Caro, J. Angew. Chem. Int.
Ed. 2010, 49, 548.
[35] Li, Y. S.; Liang, F. Y.; Bux, H.; Yang, W. S.; Caro, J. J. Membr. Sci. 2010, 354, 48.
[36] Huang, A. S.; Bux, H.; Steinbach, F.; Caro, J. Angew. Chem. Int. Ed. 2010, 49, 4958.
[37] Hill, J. R.; Subramanian, L.; Maiti, A. Molecular Modeling Techniques in Material
Sciences; Taylor and Francis: London, 2005.
[38] Keskin, S.; Liu, J.; Rankin, R. B.; Johnson, J. K.; Sholl, D. S. Ind. Eng. Chem. Res.
2009, 48, 2355.
[39] Duren, T.; Bae, Y. S.; Snurr, R. Q. Chem. Soc. Rev. 2009, 38, 1203.
[40] Liu, D. H.; Zhong, C. L. J. Mater. Chem. 2010, 20, 10308.
[41] Jiang, J. W.; Babarao, R.; Hu, Z. Q. Chem. Soc. Rev. 2011, 40, 3599.
[42] Rappe, A. K.; Casewit, C. J.; Colwell, K. S.; Goddard, W. A.; Skiff, W. M. J. Am.
Chem. Soc. 1992, 114, 10024.
[43] Mayo, S. L.; Olafson, B. D.; Goddard, W. A. J. Phys. Chem. 1990, 94, 8897.
[44] Chalasinski, G.; Szczesniak, M. M. Chem. Rev. 2000, 100, 4227.
[45] Car, R.; Parrinello, M. Phys. Rev. Lett. 1985, 55, 2471.
[46] Ramachandran, K. I.; Deepa, G.; Namboori, K. Computational Chemistry and
Molecular Modeling: Principles and Applications; Springer: Coimbatore, India, 2010.
[47] Allen, M. P.; Tildesley, D. J. Computer Simulation of Liquids; Oxford University Press:
Oxford, 1987.
[48] Frenkel, D.; Smit, B. Understanding Molecular Simulation: From Algorithms to
Applications; Academic Press: San Diego, 2002.
[49] Yang, Q. Y.; Zhong, C. L. J. Phys. Chem. B 2006, 110, 17776.
[50] Babarao, R.; Hu, Z. Q.; Jiang, J. W.; Chempath, S.; Sandler, S. I. Langmuir 2007, 23,
659.
[51] Yang, Q. Y.; Zhong, C. L. Chem. Phys. Chem. 2006, 7, 1417.
[52] Yang, Q. Y.; Xue, C. Y.; Zhong, C. L.; Chen, J. F. AIChE J. 2007, 53, 2832.
[53] Babarao, R.; Jiang, J. W.; Sandler, S. I. Langmuir 2009, 25, 5239.
[54] Liu, B.; Smit, B. Langmuir 2009, 25, 5918.
[55] Yang, Q. Y.; Xu, Q.; Liu, B.; Zhong, C. L.; Smit, B. Chin. J. Chem. Eng. 2009, 17, 781.
[56] Liu, Y. H.; Liu, D. H.; Yang, Q. Y.; Zhong, C. L.; Mi, J. G. Ind. Eng. Chem. Res. 2010,
49, 2902.
[57] Chen, Y. F.; Babarao, R.; Sandler, S. I.; Jiang, J. W. Langmuir 2010, 26, 8743.
[58] Chen, Y. F.; Lee, J. Y.; Babarao, R.; Li, J.; Jiang, J. W. J. Phys. Chem. C 2010, 114,
6602.
[59] Liu, Y.; Liu, H. L.; Hu, Y.; Jiang, J. W. J. Phys. Chem. B 2010, 114, 2820.
[60] Liu, B.; Smit, B. J. Phys. Chem. C 2010, 114, 8515.
[61] Bae, Y. S.; Mulfort, K. L.; Frost, H.; Ryan, P.; Punnathanam, S.; Broadbelt, L. J.; Hupp,
J. T.; Snurr, R. Q. Langmuir 2008, 24, 8592.
[62] Bae, Y. S.; Farha, O. K.; Spokoyny, A. M.; Mirkin, C. A.; Hupp, J. T.; Snurr, R. Q.
Chem. Commun. 2008, 4135.
[63] Bae, Y. S.; Farha, O. K.; Hupp, J. T.; Snurr, R. Q. J. Mater. Chem. 2009, 19, 2131.
Metal-Organic Frameworks for CO2 Capture 247

[64] Farha, O. K.; Bae, Y. S.; Hauser, B. G.; Spokoyny, A. M.; Snurr, R. Q.; Mirkin, C. A.;
Hupp, J. T. Chem. Commun. 2010, 46, 1056.
[65] Bae, Y. S.; Spokoyny, A. M.; Farha, O. K.; Snurr, R. Q.; Hupp, J. T.; Mirkin, C. A.
Chem. Commun. 2010, 46, 3478.
[66] Mu, W.; Liu, D. H.; Yang, Q. Y.; Zhong, C. L. Micro. Meso. Mater. 2010, 130, 76.
[67] Babarao, R.; Dai, S.; Jiang, D. E. Langmuir 2011, 27, 3451.
[68] Chen, Y. F.; Jiang, J. W. ChemSusChem 2010, 3, 982.
[69] Jiang, J. W. AIChE J. 2009, 55, 2422.
[70] Babarao, R.; Jiang, J. W. J. Am. Chem. Soc 2009, 131, 11417.
[71] Babarao, R.; Eddaoudi, M.; Jiang, J. W. Langmuir 2010, 26, 11196.
[72] Babarao, R.; Jiang, J. W. Ind. Eng. Chem. Res. 2011, 50, 62.
[73] Xu, Q.; Liu, D. H.; Yang, Q. Y.; Zhong, C. L.; Mi, J. G. J. Mater. Chem. 2010, 20, 706.
[74] Mu, W.; Liu, D. H.; Zhong, C. L. Micro. Meso. Mater. 2011, 143, 66.
[75] Myers, A. L.; Prausnitz, J. M. AIChE J. 1965, 11, 121.
[76] Ryan, P.; Broadbelt, L. J.; Snurr, R. Q. Chem. Commun. 2008, 4132.
[77] Babarao, R.; Tong, Y. H.; Jiang, J. W. J. Phys. Chem. B 2009, 113, 9129.
[78] Chen, Y. F.; Jiang, J. W. in preparation.
[79] Yazaydin, A. O.; Snurr, R. Q.; Park, T. H.; Koh, K.; Liu, J.; Levan, M. D.; Benin, A. I.;
Jakubczak, P.; Lanuza, M.; Galloway, D. B.; Low, J. J.; Willis, R. R. J. Am. Chem. Soc.
2009, 131, 18198.
[80] Wu, H.; Simmons, J. M.; Srinivas, G.; Zhou, W.; Yildirim, T. J. Phys. Chem. Lett.
2010, 1, 1946.
[81] Chen, Y. F.; Nalaparaju, A.; Jiang, J. W. in preparation.
[82] Lan, J. H.; Cao, D. P.; Wang, W. C.; Smit, B. ACS Nano 2010, 4, 4225.
[83] Zacher, D.; Shekhah, O.; Woll, C.; Fischer, R. A. Chem. Soc. Rev. 2009, 38, 1418.
[84] Keskin, S.; Sholl, D. S. J. Phys. Chem. C 2007, 111, 14055.
[85] Babarao, R.; Jiang, J. W. Langmuir 2008, 24, 5474.
[86] Keskin, S.; Sholl, D. S. Ind. Eng. Chem. Res. 2009, 48, 914.
[87] Keskin, S.; Sholl, D. S. Langmuir 2009, 25, 11786.
[88] Atci, E.; Erucar, I.; Keskin, S. J. Phys. Chem. C 2011, 115, 6833.
[89] Keskin, S. J. Phys. Chem. C 2011, 115, 800.
[90] Krishna, R.; Van Baten, J. M. Phys. Chem. Chem. Phys. 2011, 13, 10593.
[91] Hu, Z. Q.; Jiang, J. W. in preparation.
[92] Brandani, F.; Ruthven, D. M. Ind. Eng. Chem. Res. 2004, 43, 8339.
[93] Yazaydin, A. O.; Benin, A. I.; Faheem, S. A.; Jakubczak, P.; Low, J. J.; Willis, R. R.;
Snurr, R. Q. Chem. Mater. 2009, 21, 1425.
[94] Babarao, R.; Jiang, J. W. Energy Environ. Sci. 2009, 2, 1088.
[95] Liu, Y. L.; Eubank, J. F.; Cairns, A. J.; Eckert, J.; Kravtsov, V. C.; Luebke, R.;
Eddaoudi, M. Angew. Chem. Int. Ed. 2007, 46, 3278.
[96] Tafipolsky, M.; Amirjalayer, S.; Schmid, R. Mirco. Meso. Mater. 2010, 129, 304.
[97] Horike, S.; Shimomura, S.; Kitagawa, S. Nature Chem. 2009, 1, 695.
[98] Kang, I. J.; Khan, N. A.; Haque, E.; Jhung, S. H. Chem. Eur. J. 2011, 17, 6437.
[99] Tan, J. C.; Cheetham, A. K. Chem. Soc. Rev. 2011, 40, 1059.
In: Coordination Polymers and Metal Organic Frameworks ISBN: 978-1-61470-899-5
Editors: O. L. Ortiz and L. D. Ramirez, pp. 249-276 2012 Nova Science Publishers, Inc.

Chapter 8

HALOGEN BONDING IN THE ASSEMBLY


OF HIGH-DIMENSIONAL SUPRAMOLECULAR
COORDINATION POLYMERS

Ya-Juan Fan, Shuang-Quan Zang, Yong-Li Wei,


Chen-Xia Du, and Hong-Wei Hou
Department of Chemistry, Zhengzhou University, Henan, P. R. China.

ABSTRACT
The investigation of supramolecular assemblies based on halogen bonding (XB) has
been a field with rapid growth because a large variety of novel architectures which were
constructed through halogen bonding have been reported to possess potential
applications. Halogen bonding as well as related halogenhalogen and halogen
intermolecular interactions found in a given crystalline is valuable to inorganic chemists
on their study and poses an interesting challenge. This chapter will give a concise
overview on recent developments in the syntheses and preparations of high-dimensional
supramolecular coordination architectures based on halogen-related interactions. The
interplay of coordination bonds and such intermolecular forces highlights the complexity
and challenge in supramolecular assembly of high-dimensional coordination polymers.

INTRODUCTION
The design and synthesis of high-dimensional supramolecular coordination architectures
accompanied by non-covalent interactions has been an active field in recent years [1-3]. Some
characteristics of the complexes are related to the non-covalent interactions which determine
their architectures [4-7]. Many intermolecular interactions, such as hydrogen bonds [8, 9] and
stacking [10, 11] are effective in the construction of supramolecular architectures and
have been investigated in detail. Recently, halogen-related interactions, including halogen

E-mail: zangsqzg@zzu.edu.cn.
250 Ya-Juan Fan, Shuang-Quan Zang, Yong-Li Wei et al.

bonding as well as related halogenhalongen and halogen intermolecular interactions have


been exploited as another fundamental interaction which may be utilized in the deliberate
design of supramolecular materials [12-14].
Halogen bonding (XB) is a strong, specific, and directional interaction that gives rise to
well-defined supramolecular synthons. The investigation of such non-covalent interactions
was put forward over one century ago [15]. In recent years, Metrangolo and Resnati [12, 13,
16, 17] and other researchers [18, 19] have demonstrated the effectiveness of such
interactions in crystal engineering, thereby drawing widespread interest in many fields, most
notably in connection with drug-receptor interactions [20], halide-anion receptors [21],
supramolecular organic conductors [14], liquid crystals [22], crystal plasticity [23], nonlinear
optics (NLO) [24], luminescence [25] and nanomaterials [26].
Lee Brammers group [27] have successfully shown that although the attention on
halogen bonds has been mainly focused on organic molecules or adducts of the dihalogens,
they are pervasive across the periodic table. Some metal halides can be halogen bond donors,
and a number of discrete coordination molecules can serve as halogen bond acceptors.
In this chapter, firstly, we will illustrate the halogen-related interactions in some organic
supramolecules, and then give a concise overview of recent developments in the halogen-
interaction based metal-organic coordination systems.

1. HALOGEN-RELATED INTERACTIONS
Halogen-related interaction is a short non-bonding contact involving electrophilic
halogen atom and another electron-dense donor atom. Systematic searches by the Cambridge
Structural Database (CSD) reveal that halogen-related interactions such as CHhalogen,
halogenO (N, S), halogenhalogen and halogen interactions are ubiquitous non-
covalent interactions in Crystal Engineering. The energy of the halogen bonding ranges from
5 to 180 kJ/mol, with the weak Cl Cl interaction between chlorocarbons and the very strong
I I2 interaction in I3 being the extremes [16].
It has been proven that halogen-related interaction is effective in the formation of large
numbers of supramolecular architectures with interesting topologies. In this section, we will
select some examples to show different patterns of XB.

1.1. HalogenN (Or O, S) Halogen Bonding

A variety of halogen bonding acceptors such as carboxylate, pyridine, imidazole, amine


and other groups which contain N, O or S atoms have been widely used in crystal engineering
in which halogenN (or O, S) bonding were found. They play impotant roles in the self-
assembly of a multitude of supramolecular architectures. Metrangolo, Resnati, and co-
workers [12, 13, 16, 17] have demonstrated how halogen bonding is a strong, specific, and
directional interaction that provides an additional opportunity for the crystal engineer to
design novel architectures. They can prevail over hydrogen bonding in competitive
recognition processes.
Halogen Bonding in the Assembly 251

The attractive nature of halogen bonding causes the distances between involved atoms to
be shorter than the sum of van der Waals radiis of corresponding atoms. The stronger the
interaction is, the shorter the distance is. The strength of the interactions decreases from I to
Cl due to the polarisability of the halogen being IBrCl. [16]
Facchetti and coworkers have reported an interesting layer structure constructed from 4-
(2,4,6-trifluoro-3,5-diiodostyryl)pyridine [26]. In the structure of 1, 4-(2,4,6-trifluoro-3,5-
diiodostyryl)pyridine forms a helical chain through CHF (2.425 ) hydrogen interactions,
which is further fused together through IN intermolecular interactions into a two-
dimensional supramolecular layer (Figure 1). The IN distance is 2.860 and the CIN
angle is 172.9.
Aakerys group [28] have used tetrafluorodiiodobenzene and iso-nicotinamide to
construct the supramolecular structure of desired connectivity through hydrogen bonds and
halogen bonds. Compound 2 exhibits two-dimensional structure formed by the strong classic
IN halogen bonding with IN distance of 2.859 and the CIN angle almost being
180 and NHO (2.197 and 2.133 ) hydrogen bonding, as shown in Figure 2.

Figure 1. The layer structure of 4-(2,4,6-trifluoro-3,5-diiodostyryl)pyridine (1) constructed from IN


and CHF intermolecular interaction.

Figure 2. Perspective view of the two-dimensional structure of binary co-crystals


tetrafluorodiiodobenzene iso-nicotinamide (1 : 2) (2).

Figure 3. Perspective view of the trimeric systems, complexes 3 and 4, which constructed from the I
N and BrN halogen bonding.
252 Ya-Juan Fan, Shuang-Quan Zang, Yong-Li Wei et al.

Bruce and coworkers have reported that halogen bonding could induce liquid crystal
behavior. In 2007, they obtained two complexes 3 and 4 by the use of 1,4-
diiodotetrafluorobenzene (DITFB) and 1,4-dibromotetrafluorobenzene (DBrTFB) as halogen
bond-donor (electron-acceptor, Lewis acid) with 4-alkoxy-4'-stilbazoles. Results show how
trimeric systems may be constructed by using a difunctional halogen electron-acceptor [29].
Such two complexes are almost isomorphous. The halobenzene molecule was caught in the
middle of two 4-alkoxy-4'-stilbazoles molecules through IN (2.812 ) and BrN (2.867
) halogen bonding with the angle of 174, 175 respectively, as shown in Figure 3. The
linearity of halogen bonding angle unequivocally demonstrated charge transfer between both
species.
William Jones and co-workers [30] have obtained seven isostructural halogen-bonded
cocrystals by using 1,4-dibromo-and 1,4-diiodotetrafluorobenzene (donors) and
thiomorpholine, thioxane, morpholine, and piperazine (acceptors). Each solid is 1:1 cocrystals
composed of anticipated chains that are held together through IN, IO or I S halogen
bonding. The structure of (tfib)(tox) (5) was shown in Figure 4. Each 1,4-
diiodotetrafluorobenzene molecule bridges neighboring acceptors with one iodine contacting
oxygen atom whereas the other linking sulfur atom from a different tox molecule. The
distances of IO and I S are ca. 3.11 and 3.23 , respectively. The angles of the
corresponding atoms are close to 180.

Figure 4. 1D supramolecule structure of (tfib)(tox) (5). Containing IO and IS interactions.

Figure 5. Chain structure of 6, consisting of the IS and NH S interactions.

Figure 6. Layered structure of 7, consisting of IS and NH S interactions.


Halogen Bonding in the Assembly 253

Figure 7. Three-dimension supramolecular architecture of 8, constructed from IS and N S


interactions.

Penningtons group [31] have reported three complexes 68, using thiourea and its
derivatives with three different organoiodines (mercaptobenz-imidazole and 1,2-
diiodotetrafluorobenzene (7), thiourea and 1,4-diiodotetrafluorobenzene (6), ethylenethiourea
and tetraiodoethylene (8) respectively), in which the halogen bonding and hydrogen bonding
play vital roles in consolidating the structures. In complex 6, only one of the two iodine atoms
of 1,2-diiodotetrafluorobenzene participates in IS halogen bonding, so there is no extension
of the dimensionality of the structure beyond the NHS ribbons (Figure 5). Two iodine
atoms of 1,4-diiodotetrafluorobenzene participate in IS halogen bonding, so complex 7
exhibits two-dimensional network (Figure 6). Complex 8 shows three-dimensional
supramolecular network. The IS interaction distances range from 3.281 to 3.404 and the
N S interactions are from 3.317 to 3.489 (Figure 7).

1.2. HalogenHalogen Interactions

Halogen halogen interactions X X (X = Cl, Br, I) can be used as design elements in


the assembly of supramolecular architectures. Two types of such interactions have been
summarized according to their geometrical features by Desiraju and Parthasarathy [23, 32]
(Scheme 1). The type-I motif represents the interaction between identical portions of the
halogen atoms. The type-II motif corresponds to the interaction originating from the
electrostatic forces and repulsion anisotropy between halogen atoms polarized positively in
the polar region and negatively in the equatorial region.

Scheme 1.
254 Ya-Juan Fan, Shuang-Quan Zang, Yong-Li Wei et al.

Figure 7. Perspective view of the infinite 1D chain in 8.

Figure 8. Layer structure of 9.

Compared to the other halogens, iodine is easier to form halogen bonding because it is
more readily polarized.
Aakerys group [29] have also used iodine with iso-nicotinamide to assemble the
supramolecular structures with hydrogen bonds and halogen bonds. Compound 8 shows
infinite chain through IN (2.439 ) and II halogen bonds, as shown in Figure 7. The
halogenhalogen interaction belongs to type I with the distance between two iodine atoms of
3.631 and the angle 1=2=170.
Facchetti and coworkers [26] have also reported another compound, which shows a layer
supramolecular structure. In the structure of 9, the (4-(2,4,6-trifluoro-3,5-
diiodostyryl)pyridine 1-oxide) ligands form a 2D layer through IN (2.860 ) and CHF
(2.425 ) intermolecular interactions, which is further associated together through
interactions with the distance of 3.36 to a three-dimensional supramolecular architecture
(Figure 8). The type II motif of halogenhalogen bonding was also detected in the layer, in
which II distance is 3.747 .

1.3. Halogen Interactions

Systematic survey in the Cambridge Structural Database shows that the halogen
contact is also a ubiquitous weak interaction but has been overlooked in most cases because
of the existence of other interactions. Shishkin and co-workers [33, 34] have summarized
three types of halogen interactions according to the location of halogen atom with respect
to the aromatic plane, as shown in Scheme 2.
Halogen Bonding in the Assembly 255

Scheme 2.

Chlorine, bromine and iodine substituted derivatives of 5,5'-bisdiazo-dipyrromethane


have also been introduced into the assembling reactions by Yin and coworkers giving rise to a
series of complexes containing Cl , Br and I interactions, respectively [35]. Here
we take the bromine substituted derivative (10) as an example. In the structure of 10, the
dimers which are formed by NHN hydrogen bonding are fused together into a double
chain structure via two semilocalized Br interactions with the BrC distances of 3.493,
3.367, 3.395, 3.537, respectively. The double chain is extended to a 2D structure through
one delocalized Br interaction with the corresponding distance and angle being 3.581
and 174 (Figure 9). Each Br atom is involved in one localized Br interaction which holds
the layers together. Their study has demonstrated that various halogen interactions are
capable of influencing the packing of organic crystals decisively.

Figure 9. (a) Perspective view of the dimers of 10 formed by the NHN hydrogen bonding. (b)
Br interactions associated the dimers into two-dimensional layer in 10.
256 Ya-Juan Fan, Shuang-Quan Zang, Yong-Li Wei et al.

2. HALOGEN RELATED INTERACTIONS


IN METAL-ORGANIC NETWORKS

Combination of metal-organic compounds with halogen bonding is an intriguing tool in


crystal engineering and material design [36]. A wide variety of halogen bonding have been
found in metal-containing crystal structures. Halogen molecules or anions, simple ligands
bound to transition metals or main group metals and multifunctional ligands containing
halogen atoms may be involved in metal containing halogen bonded coordination
architectures.

2.1. Halogen Molecules or Anions Systems

First, we will discuss the interactions involving halogen molecules or anions in the
assembly of metal-organic polymers. Jack Y. Lus group [37] have made a novel iodine-
inclusion coordination polymer [CuI(C5H3NI2)0.5I2] (11) through an interesting reaction of
Cu(NO3)22.5H2O with 3,5-pyridinedicarboxylic acid and iodine under hydrothermal
conditions, in which reduction of copper(II) to copper(I) by pyridinecarboxylate, the
substitution of carboxylate groups by iodine nucleophiles and a self-assembly process
simultaneously happened. The I2 molecule acts as a 2-bridge linking two coordinated iodine
atoms from different zigzag chains through II interactions with the distance between two
iodine atoms being 3.485 . Then the zigzag ladder is extended to a supramolecular layer
(Figure 10), and such layers are connected with each other through longer II contacts (II,
3.697 ) involving adjacent coordinated iodine atoms and substitutional iodine atoms to
result in a 3D supramolecular structure, as shown in Figure 10. The II interactions between
inclusion iodine molecules and network play an important role in the construction of the 3D
metal-organic polymer.
More interestingly, when the reactions absent of copper (II) ions were conducted under
the same reaction condition, a single iodine-substituted complex IC5H3NCOOH (12) was
obtained. The colorless crystal has no inclusion iodine molecules or hydrogen bonds
(O1H N1, 2.649 ) or IO interactions (I1 O2 3.042 ) to contribute to the 2D
supramolecular structure (Figure 11).

Figure 10. Left: The 2D layer in [CuI(C5H3NI2)0.5I2] (11).(Dotted lines represent II interactions.)
Right: Perspective view of the 3D structure in 11.
Halogen Bonding in the Assembly 257

Figure 11. A view of the layer in IC5H3NCOOH (12). (Dotted lines represent IO interactions.).

Figure 12. The supramolecular structure of [Co(H2bbim)3]Cl2 2H2O (13). (Dotted lines represent
hydrogen bonds and Cl interactions.).

In 2008 Bao-Hui Ye et al. [38] reported a series of metal complexes with 2,2'-
biimidazole-like ligand and chloride. The structures of these complexes have shown that a
variety of Cl-related non-covalent interactions such as XH Cl (X=N, O, and C) and Cl
interactions play important roles in the assembly of the metal-organic polymers with an
increase in dimensionality as well as the stabilization of supramolecular architecture. Here,
we take [Co(H2bbim)3]Cl2 2H2O (13) for example. Cl1 and Cl2 anions are hydrogen bonded
to two water molecules to form a coplanar [(H2O)2Cl2]2- unit, which linked adjacent
[Co(H2bbim)3]2+ cations to result in a zipper-like double chain via NH Cl hydrogen
bonding. The distance between Cl2 anion and the imidazolyl ring from adjacent chain is
3.301 , which indicates the existence of Cl (imidazolyl) charge-assisted interactions
because the coordination of a positively charged Co(II) ion greatly enhances the electron-
deficient character of the imidazolyl ring and provides sufficient polarization to produce an
anion- charge-assisted interaction. Such chain is expanded to supramolecular sheet through
Cl interactions and stacking, as shown in Figure 12.
258 Ya-Juan Fan, Shuang-Quan Zang, Yong-Li Wei et al.

Figure 13. (a): The layer in [Hg( 3-pz-3-py)Cl2]n (14), containing short ClCl contacts. (b): A view of
the BrBr and Br interactions in [Hg( -pz-3-py)Br2]n (15).

Coordinated halide anions can control the structure of the polymers in the self-assembly
process through halogen-related non-covalent interactions too. Recently, Nasser Safari and
co-workers [39] have investigated two novel halogen-containing polymers based on a new
flexible tritopic pyrazine-pyridine ligand (pz-3-py) with HgX2 (X = Cl, Br). Results showed
that coordinated chloride and bromide anions play different roles in the supramolecular
architectures. In 14, short ClCl contacts are found in the 2D covalent net with the distance
of two adjacent chloride anions being 3.286 , which stabilize non-covalent interactions in
the structure (Figure 13a). As to 15, when bromide anion was used instead of chloride anion,
a remarkably different architecture was found. Complex 15 features a zigzag chain structure
and such chains are connected to each other through NHN and CH hydrogen bonds
extending to a 2D supramolecular architecture. The shortest distance between adjacent
bromide anions from different layers is 3.593 which is shorter than the sum of the van der
Waals radiis of two Br atoms (ca. 3.7 ), indicating there is also a halogenhalogen
interaction. In contrast to 14, the short BrBr contact plays an important role in the increase
of dimensionality (2D3D). Further more, the key factor for the 2D3D architecture is the
Br interaction, which is highly directional and effective in crystal packing (Figure 13b). It
is evident that Br is more effective than Cl to function as an electron-rich site on forming a
lone-pair interaction at its equator region with electron-deficient rings. This result is
consistent with the halogen bonding theory that the polarizability and anisotropic character of
the halogen atom increases from F to I.

2.2. Metal-Halide and Metal-Pseudohalide Systems

Lee Brammers group [40] have provided an opportunity to examine the potential
competition between halometallate ions and halide ions for both hydrogen bond (NHX)
Halogen Bonding in the Assembly 259

and halogen bond (CXX') formation. They have studied four complexes, namely, (4-
ClpyH)3[PtCl6]Cl (16), (4-XpyH)3[FeCl4]2Cl [X = Cl (17) or Br (18)] and (3-
IpyH)2[AuBr3X]X (X = Cl/Br) (19). In complex 16, 4-chloropyridinium cations, PtCl62 and
Cl anions are linked together mostly via NHCl and CHCl hydrogen bonds. Only a
weak asymmetric CClClC halogenhalogen interaction (ClCl 3.309 ) was
observed. As for 17 and 18, pairs of chloride ions are surrounded by a set of six 4-
chloropyridinium cations through hydrogen bonds. One of each three independent
halopyridinium cations forms a short halogen bond FeClXC (ClCl 3.366 for 17,
ClBr 3.361 for 18) with a single chloride atom from a FeCl4 anion, while the others
form much longer FeClXC interactions (3.59 < ClCl < 3.79 for 17, 3.51 <
ClBr < 3.89 for 18) with two chloride ligands along one edge of a FeCl4 anion.
Compared with CCl, the CBr group provides a greater electrophilic nature, which may
contribute to the shorter contact of FeClBrC in 18. In complex 19, when the
iodopyridinium cation and [AuBr3.35Cl0.65] anion were used in the system, more
halogenhalogen interactions have been observed. As shown in Figure 14, pairs of halide
anions and iodopyridinium cations form a roughly rectangular 12-membered ring through
symmetry-related NHX hydrogen bonds and CIX halogen bonds (IX 3.331 ),
and such rings are extended to a tape running along the a-axis by sharing halide anions.
[AuBr3.35Cl0.65] anions are arranged orderly to form a row propagating parallel to the a-axis
via short AuBrBrAu contacts with the distance between two Br atoms of 3.520 . The
tapes and the rows are parallel to each other and interlinked to achieve a 2D structure through
AuXX halogen interactions (XX 3.468 ). The iodopyridinium cation provides a
more electrophilic CX group and the [AuBr3.35Cl0.65] anion should have the most diffusive
charge distribution of the halometallate ions and thus provide the least competition with the
halide ion for formation of electrostatically strong interactions.
In 2006, Lee Brammer studied a series of eight complexes (4-X'pyH)2[CoX4] (4-
X'pyH=4-halopyridinium, X=Cl, Br or I) which are isostructural with both the organic and
inorganic halogen species varied [41]. This series of complexes provides a remarkable
opportunity to assess the relative importance of the charge transfer and electrostatic
contributions to the MXX'C halogen bonds by changing halogen species. The
electrostatic interaction between the positive (C)X and negative (M)X atoms is the
determinant factor in directing and governing the strength of the MXX'C halogen bonds.

Figure 14. A view of the non-covalent interactions in (3-HpyI)2[AuBr3.35Cl0.65]Br0.30Cl0.70 (19).


260 Ya-Juan Fan, Shuang-Quan Zang, Yong-Li Wei et al.

Figure 15. The CINC halogen bonding and CHN hydrogen bonding in [N-Me-3-I-
C5H4N]2[Ru(bipy)(CN)4]0.5(MeCN) (20).

Figure 16. The CINC halogen bonding and CI (CN) interaction in [N-Me-3,5-I2-
C5H3N]2[Ru(bipy)(CN)4] (21).

CXNCM interaction can also be used as an effective tool in the process of


supramolecular assemblies due to the electrostatic correspondence between an externally
directed lone pair of electrons in a cyanometallate anion and an electrophilic halogen atom in
an organic molecule.
Michael D. Ward and co-workers [42] had a structural study on the combination of
[Ru(bipy)(CN)4]2 anions with N-methylpyridinium and its halogen substitutes. As shown in
Figure 15, in [N-Me-3-I-C5H4N]2[Ru(bipy)(CN)4] 0.5(MeCN) (20), each [Ru(bipy)(CN)4]2
anion is associated with two N-methyl-3-iodopyridinium cations through CINC halogen
bonding, in which the IN distances are 2.809, 2.820, 2.906, 2.963 respectively, much less
than the sum of the van der Waals radiis of I and N atoms (3.53 ). Symmetry-related [N-Me-
3-I-C5H4N]2[Ru(bipy)(CN)4] units are connected each other through CHN hydrogen
bonds and stacking to form a binuclear unit and such unit is extended to 3D
supramolecular architecture via hydrogen bonds. In [N-Me-3,5-I2-C5H3N]2[Ru(bipy)(CN)4]
(21), each [Ru(bipy)(CN)4]2 anion is surrounded by four N-Me-3,5-I2-C5H3N cations, in
which three iodine atoms are involved in I-related interactions. As shown in Figure 16, IN
distances are ranging from 2.793 to 2.974 . The interactions involving I1 and I3 atoms show
some deviation from linearity of the interaction at nitrogen, with CIN angles close to
180, whereas the other one has an almost orthogonal approach of the CI group to the
cyanide ligand, indicating the former two interactions are CIN halogen bonding while the
latter is CI (CN) interaction. Instead of I-related interaction, the anion is associated to
the last N-Me-3,5-I2-C5H3N cation by CH (CN) hydrogen bonds involving one methyl
hydrogen atom and one aromatic ring proton from the cation. The cationic organohalogen
Halogen Bonding in the Assembly 261

compounds act as bridges linking adjacent anions to form a 1D structure. In contrast, when
the analogous brominated cations (N-methyl-3-bromopyridinium and N-methyl-3,5-
dibromopyridinium) were used in this system, no CBrNC halogen bonds were found in
the solid state. The solid-state structures suggest that charge-assisted CI NC halogen-
bonding interactions play an important role in controlling the solid state structures and
cyanide-halogen interactions are a useful tool in the preparation of multicomponent
assemblies.

2.3. Metal-Tetrathiafulvalenium (TTF) Systems

Scheme 3.

Figure 17. (a) The 1D ribbon structure formed via CI NC halogen bonding in -(DIETS)2[Au(CN)4]
(22). (b)(c) The connectivity pattern of the 3D supramolecular network in 22.
262 Ya-Juan Fan, Shuang-Quan Zang, Yong-Li Wei et al.

Halogen bonding has been introduced in the field of molecular conductors by Imakubo
and Kato [43-45]. In recent years, XB has been fully investigated as the first choice of non-
covalent interaction to control the solid-state structures of molecular conductors which
influence their electronic properties. A variety of neutral halogenated tetrathiafulvalenes
(TTFs) and their radical cation salts (Scheme 3) were introduced to the system, which has
been summarized in the very exhaustive reviews [14, 46]. Here, we only take the -
(DIETS)2[Au(CN)4] (22) for example, which is a supramolecular superconductor prepared by
electrochemical oxidation of a CH2Cl2 solution of DIETS in the presence of
[Au(CN)4](NBu4) as a supporting electrolyte. As shown in Figure 17a, the DIETS cations and
square planar tetracyanoaurate anions are arranged alternatively and associated together to
form a 1D ribbon structure via CI NC halogen bonding with the I N distance of 3.018
and CI N angle of 174 . Each DIETS molecule functions as bidentate XB donor with each
iodine atom connected to one tetracyanoaurate anion and the anion functions as a tetradentate
XB acceptor with two of the four cyano groups functioning as bifurcated XB acceptors. The
ribbons are extended to a 3D supramolecular network through CH N hydrogen bonds
involving the methylene groups of the DIETS cations and cyanoaurate nitrogen atoms of
[Au(CN)4] anions. The connectivity pattern of the 3D supramolecular network is depicted in
Figure 17b and 17c. Interestingly, the complex exhibits a peculiar electronic behavior with
the superconducting transition temperature being 8.6K which is higher than the highest
known value for unsymmetrical -donors. The strong iodine-based halogen bonding in this
complex appeared to be particularly relevant in the inducing uniaxial strain superconductivity
by improving the anisotropic character of organic conducting crystals. The IN interaction is
much more directional and particularly suitable for this purpose than the flexible HB.

2.4. SilverEthynide Systems

Scheme 4.
Halogen Bonding in the Assembly 263

Figure 18. 3D supramolecular structure in [(AgL1)(AgCF3CO2)4(H 2O)(CH3CH2CN)]2 (23)(Left) and


[(AgL2)(AgCF3CO2)4(H2O)(CH3CN)]2 (24)(Right). All irrespective atoms are omitted for clarity.

In 2009, our group have reported a series of new silver(I) ethynide complexes based on 1-
iodo-4-prop-2-ynyloxy-benzene (HL1) and 1-bromo-4-prop-2-ynyloxy-benzene (HL2)
(Scheme 4), in which the halogen-related interactions play important roles [47].
In the asymmetric unit of [(AgL1)(AgCF3CO2)4(H 2O)(CH3CH2CN)]2 (23), there are
two anionic L1 ligands, ten crystallographically independent Ag(I) ions, two bridging aqua
ligands and two coordinated propionitrile molecules. The ethynide groups are bound to
butterfly-shaped Ag4 baskets in the 41112 mode. Such two independent Ag4 baskets
can be regarded as firstly united together through the 3-bridged trifluoroacetate groups to
give a chain structure along the b axis, and the system of parallel chains as building units are
further fused together through strong AgIaryl interactions (Ag5I2 2.661 and
Ag10I1 2.660 ) along the a axis to give a 2D metal-organic network which is stabilized
by donor hydrogen bonds from O1W and O2W: O1WO10 2.855 ,O1WO15 2.792
,O2WO7 2.776 and O2WO18 2.878 .
Interestingly, each layer is bonded to its neighbors on both sides via I
(I2center(C4C9), 3.717 ), O3I1E (3.473 ), and weak N1I1E (3.628 )
interactions, imparting non-centrosymmetry upon the entire 3D assembly (Figure 18).
[(AgL2)(AgCF3CO2)4(H 2O)(CH3CN)]2 (24) is allomeric with 23, and neighboring
layers in 24 interact on both sides via Br (Br2center(IB), 3.556 ), O5Br1B (3.303
), and weak CH (H38BAcenter(II), 3.175 ) interactions, imparting non-
centrosymmetry upon the entire 3D assembly (Figure 18).

2.5. Metal-Porphyrin Systems

Israel Goldberg [48, 49] has reported a series of crystals based on two pre-designed
porphyrins which bear mixed and complementary pyridyl and iodophenyl molecular
recognition functions, namely, 5-(4-pyridyl)-10,15,20-tris(4-iodophenyl)porphyrin (PyTIPP)
and 5,15-bis(4-pyridyl)-10,20-bis(4-iodophenyl)porphyrin (PyDIPP) (Scheme 3). Their
diverging iodophenyl groups may be involved in the NI halogen bonding, as well as related
II and I interactions which can serve as important tools in the assembly of the
supramolecular structures.
264 Ya-Juan Fan, Shuang-Quan Zang, Yong-Li Wei et al.

Scheme 5.

Figure 19. The layered supramolecular aggregation in PyTIPP (25) (a) and PyDIPP(26) (b). The dashed
lines represent NI halogen bonding (contacts of type a), I interactions (contacts of type b and c)
and intermolecular CH N contacts.
Halogen Bonding in the Assembly 265

Figure 20. Views of the layer structure in [Zn(py)-PyTIPP] (27) (a) and [Cu-PyTIPP] (28) (b). The
coordinated pyridyl rings in 27 are omitted for clarity. The dashed lines represent NI halogen
bonding (contacts of type a), II (contacts of type b) and I interactions (contacts of type c, d and
e).

As shown in Figure 19, in the free base PyTIPP crystal, adjacent porphyrin units are
connected to each other through NI halogen bonding with the distance of the corresponding
atoms of 2.931 to form a chain running along the trans-related pyridyl-iodophenyl axis.
The other two sideway CI bonds of the porphyrin unit are almost perpendicular to the
iodophenyl and pyridyl rings of adjacent arrays with the iodine-to-aryl ring distances being
ca. 3.448 and 3.408 respectively, which indicate the existence of I interaction. Then
the linear structure is extended to a flat supramolecular layer. It is clear that the self-assembly
of PyTIPP in the crystalline phase is directed by halogen bonds. I interaction is also
detected in the PyDIPP, which plays an important role in the self-assembly of the
architecture.
As to their core-metalated derivatives, the DMF solvate of [Zn(py)-PyTIPP] (27) and the
pyridine solvate of [Cu-PyTIPP] (28) can also self-assemble via NI halogen bonding as
well as related II and I interactions. Both the crystal structures can be described as
being composed of flat supramolecular layers of porphyrin units, in which the chains are
formed by head-to-tail NI bonds involving the trans-related pyridyl and iodophenyl groups
of the porphyrin moiety and neighboring chains further interact on both sides via secondary I
interactions.
266 Ya-Juan Fan, Shuang-Quan Zang, Yong-Li Wei et al.

In complex 27, Zn center is coordinated by four nitrogen atoms from one porphyrin unit
and a single nitrogen atom from one discrete pyridyl molecule to achieve a slightly distorted
tetragonal pyramidal coordination environment [ZnN5]. Adjacent porphyrin units are linked
each other by head-to-tail NI halogen bonding with the separation of 2.873 to result in a
1D perfectly aligned supramolecular structure (Figure 20a). Neighboring rows are parallel to
each other with all NI halogen bonds oriented in the same direction. The other two iodine
atoms of the porphyrin moiety are involved in II and I interactions with other
iodophenyl groups from adjacent rows, by which the 1D structure is extended to 2D
supramolecular layer. There are some differences between the flat supramolecular layer in 28
and 27. In complex 28, the Cu center is four-coordinated by four nitrogen atoms of one
porphyrin unit to form a slightly distorted square-planar geometry. The component building
blocks are arranged in ABAB mode. As depicted in Figure 20b, in the A rows, the distance
between adjacent N and I atoms is 3.060 and the other two iodine atoms of the same
porphyrin unit are connected to iodophenyl groups of different B rows through I
interaction. In the B rows, the NI distance is 3.031 and the other two iodine atoms are
involved in I interaction with two pyridyl rings of neighboring A arrays.
More interestingly, the supamolecular layers in 25, 27 and 28 have polar axises and are
chiral, and all the layers are aligned in the same direction, imparting chirality upon the entire
3D assembly. This study provides a unique example of the induction of supramolecular
chirality by halogen bonding and confirms that such bonding can serve as an important tool in
the patterning of supramolecular assemblies.

2.6. Cu2(3-Iodobenzoate)4 Systems

In 2008, Lee Brammers group have investigated a series of metal-organic complexes


based on Cu(II) center and 3-iodobenzoate in which secondary building units (SBUs)(Figure
21) existed in the linear structures [50]. They found that these groups can form halogen-
bonded architectures with simple short linkers such as 1,4-dioxane and
diazabicyclo(2.2.2)octane (dabco). Increasing the linker size to 4,4'-bipyridine (bipy) or 4,4'-
bipyridylethane (bpe), novel coordination polymers were obtained, in which linear chains of
SBU are associated through various iodine-related interactions.
The structure of complex 29 is composed of discrete paddlewheel units (SBU I). Each
SBU contains two copper atoms which are bridged by four carboxylate groups from different
3-iodobenzoate ligands.

Figure 21. Secondary building units in 2934.


Halogen Bonding in the Assembly 267

Figure 22. (a) A view of the halogen bonding in [Cu2(2-3-iodobenzoate)4(DMF)2] (29) (b) Schematic
view of the 3D supramolecular structure in 29.

The two copper axial sites are occupied by oxygen atoms from different DMF molecules.
As shown in Figure 22, two of the phenyl rings are parallel to the CuCu axis and coplanar
with the corresponding carboxylate groups, while the other two aryl rings are twisted to the
planes of the corresponding carboxylate groups with the dihedral angle of ca. 31. The iodine
atom is close to oxygen atom of DMF molecule from neighbouring paddlewheels due to the
twist of the benzoate ring. The distance of the adjacent iodine and oxygen atoms is 3.387
which is shorter than the sum of the van der Waals radiis of the two atoms (ca. 3.53 ) and
the CIO angle is 162.6, indicating the existence of CIO halogen bonding. Each SBU
donates and accepts two halogen bonds, resulting in a 3D supramolecular architecture. If the
SBU is clarified as a four-connected node (connecting to other four SBUs through halogen
bonding), the structure of 29 can be classified as a diamond-type (6, 4) net with a (66) Schlfli
symbol (Figure 22).

Figure 23. (a) A view of the halogen bonding in [Cu2(2-3-iodobenzoate)4(dioxane)2]2(dioxane) (30).


(b) The connected pattern between adjacent chains in 30. (c) Schematic view of the layer in 30.
268 Ya-Juan Fan, Shuang-Quan Zang, Yong-Li Wei et al.

The structure of complex 30 is also composed of discrete SBU I. The difference is that
the four benzoate rings are approximately coplanar with the corresponding carboxylate
groups and the two axial coordination sites around copper are occupied by oxygen donors of
dioxane solvent molecules (Figure 23a). The other oxygen atom of each dioxane is connected
to iodine atom from neighboring SBU through CIO halogen bonding ( IO 3.138 ,
CIO 170.9). At the same time, two carboxylate oxygen atoms of the SBU form other two
CIO halogen bonding with IO distance and CIO angle being 3.376 and 159.8
respectively. As shown in Figure 23b, each SBU is connected by other four SBUs through
four pairs of halogen bonding in which the SBU donates and accepts four halogen bonds. In
this way, the complex forms a 2D supramolecular structure. If the SBU is clarified as a four-
connected node, the structure of 30 can be classified as a (4, 4) net (Figure 23c).
When diazabicyclo(2.2.2)octane (dabco) replaced 1,4-dioxane, only 1D structure through
halogen bonding was obtained. The SBU I in complex 31 is similar to that in 30 in which the
aromatic rings are almost coplanar with the corresponding carboxylate groups. Each axial site
of the SBU is occupied by nitrogen atom of coordinated dabco molecule. The other nitrogen
atom of each dabco is connected to iodine atom from neighboring SBU through CIN
halogen bonding, in which the short IN distance is 3.075 and the CIN angle is
168.3. However, only two iodine atoms and nitrogen atoms are involved in halogen bonding,
the other two iodine atoms and carboxylate oxygen atoms do not take part in the non-covalent
interaction. As depicted in Figure 24, each SBU acts as a 2-bridge linking two adjacent
paddlewheels through pairs of CIN halogen bonding to propagate a 1D structure. The
other iodine atoms are involved in several weak CHI hydrogen bonds which may
compensate for the lack of halogen bonding.
The alternate permutation of the two kinds of SBUs from parallel adjacent chains
conduce to the iodine atom of the SBU II close to the benzoate ring of the SBU I from
neighbouring chain and the CI halogen bonding form. Then the chains are associated
together through the weak halogen bonding producing a 2D supramolecular structure (Figure
25b). The layers are further connected to each other through a second type of halogen
bonding between the electrophilic site of the paddlewheel iodine atom and the nucleophilic
site of the iodine atoms of SBU II to create a 3D network. Figure 16c and 16d show the views
of the 3D structure and the CI as well as II interactions in the complex, respectively.
Both the iodine substituents of SBU II form CI and II interactions with the
neighboring SBU I, while only a pair of iodobenzoate ligands in SBU I were involved in the
iodine-related interactions.

Figure 24. 1D structure in [Cu2(3-iodobenzoate)4(dabco)2] (31).


Halogen Bonding in the Assembly 269

Figure 25. (a) A view of the 1D structure in [Cu2(2-3-iodobenzoate)4](2-bipy)[Cu(2-3-


iodobenzoate)2]BnOH (32) (b) The layer formed through CI halogen bonding. (c) 3D network. (d)
The iodine-related interaction in 32.
270 Ya-Juan Fan, Shuang-Quan Zang, Yong-Li Wei et al.

Figure 26. 2D network formed through CI interaction in [Cu(2-3-iodobenzoate)2(2-bpe)]BnOH


(33).

When bpe was used as linkers for the coordination network, complex 33 and 34 were
obtained from a concomitant crystallization. Compound 33 consists of mononuclear copper
centers (SBU II) which are interlinked by 2-bpe producing a 1D coordination structure. The
antiparallel chains allow each iodine atom to be close to a pyridyl ring of the bpe ligand from
adjacent chain. Such chain is extended to a 2D network through CI interactions, as
shown in Figure 26.

Figure 27. (a)(b) the single chain and the double chain in {[Cu2(2-3-iodobenzoate)2(2-3-
iodobenzoate)2](2-bpe)2}{[Cu(2-3-iodobenzoate)2](2-bpe)} (34). (c) 3D supramolecular structure
composed of SBU II chains in 34. (d)(e) The iodine-related interactions between two kinds of the
chains.
Halogen Bonding in the Assembly 271

Beside the type of the single chain which is similar to that in 33, a distinct type of
coordination chain was also found in 34, which is a double chain comprised of the third
copper iodobenzoate motif (SBU III) and bpe ligand (Figure 27a, 27b). In the SBU III, two
copper centers are first bridged by a pair of iodobenzoate carboxylate groups to form an
octatomic ring. Each metal atom is further coordinated to a chelating carboxylate group of the
iodobenzoate ligand as well as two nitrogen atoms from different bpe ligands. As shown in
Figure 27c, each chain is connected to other four such chains through CI interactions
involving the iodine substituents of SBU II and the aromatic rings of SBU II in a
neighbouring chain resulting in a 3D halogen-bonded network comprising only SBU II
moieties with large channel. The double chains run parallel to the single chains along the a-
axis filling in the channel. Two kinds of halogen bonds are detected between adjacent
different kinds of chains. One is CI halogen bond between the chelating iodobenzoate
ligand of SBU III and the aromatic rings of the bpe linker, the other is CIIC
halogenhalogen interaction between the bridging iodobenzoate ligands of SBU III and the
iodine atom of SBU II, As depicted in Figure 27d and 27e.

2.7. Metal-5-iipa Systems

Recently, a series of divalent coordination polymers based on 5-iodo-isophthalic acid (5-


iipa) and ancillary nitrogen ligands have been synthesized by our group [51]. Structural
determinations of these complexes have demonstrated that 5-iipa can act as an effective
bridging ligand and its tethered iodine atom plays an important role in the assembly of the
coordination polymers with increasing in dimensionality as well as the stability of the n-fold
interpenetrated structure. Here, we take [Mn4(5-iipa)4(bpe)45H2O]n (35) for instance.
In complex 35, Pairs of symmetry-related 5-iipa2 ligands act as bridges joining adjacent
Mn(II) bimetal pairs to result in a [Mn2(5-iipa)2]n double chain and such chains are further
linked by bpe ligands running along the b axis to furnish a 2D pillared layer in the ab plane,
as illustrate in Figure 28a. More interestingly, the distance between the adjacent iodine atoms
of 3.627 are considerably shorter than the sum of the van der Waals radiis of the two atoms
(ca. 4.0) which indicate there are II interactions between the iodine atoms. When the I I
halogen bonding between layers are taken into account, the resulting net of 35 becomes a 3D
architecture with large channels (ca. 13.88517.7502).

Figure 28. (a) The pillared layer structure of 35. (b) A view of the interpenetrating 3D supramolecular
structure of 35. (Dotted lines represent II interactions.).
272 Ya-Juan Fan, Shuang-Quan Zang, Yong-Li Wei et al.

It is a rare example in the metal-organic framworks. A further investigation reveals a


more striking feature of 35, i.e. two sets of the 3D architectures are interlaced each other in a
parallel fashion to give rise to a two-fold interpenetrating structure (Figure 28b) to minimize
the open channels and stabilize the architecture. In recent years, interpenetrating nets in which
independent motifs entangled together in different ways have been widely reported for metal-
organic frameworks through covalent bonds or other noncovalent interactions in the
literatures [52, 53]. The interpenetrating structures supported by halogen bonding and related
II and I intermolecular interactions are still rare. As far as we know, only very limited
examples with interpenetrating systems based on halogen bonds have been reported [54-57],
while such interdigitating 3D frameworks formed by metal-organic layers via II
interactions have not been found up to now.

CONCLUSION AND OUTLOOK


Different interesting architectures based on the cooperation of halogen bonding as well as
related II and I intermolecular interactions and coordinated bonds have been shown in
this chapter. Obviously, the introduction of the halogen-related interactions in the design of
metal-organic complexes is a very useful tool, which makes it possible for prediction and
synthesis of novel coordination polymers and functional materials. However, the co-existence
of these interactions in combination with metal coordination within a given crystalline
compound and the effect of halogen-related interactions in the assembly of high-dimensional
coordination polymers remain largely unexplored. As could be expected, this fascinating
research field will continue to expand within the near future for their interesting structures
and potential applications as functional materials.

ACKNOWLEDGMENTS
We gratefully acknowledge financial support by the National Natural Science Foundation
of China (No. 20901070), the China Postdoctoral Science Foundation (Grant 20090460859
and 201003399) and Zhengzhou University (P. R. China).

REFERENCES
[1] Jeffrey, G. A. An Introduction to Hydrogen Bonding, Oxford University Press: Oxford,
U. K., 1997.
[2] Lindoy, L. F.; Atkinson, I. Self-Assembly in Supramolecular Systems, Royal Society of
Chemistry: Cambridge, U. K., 2000.
[3] Wu, X.-T. Controlled Assembly and Modification of Inorganic Systems, Springer:
Heidelberg, Germany, 2009.
[4] Schottel, B. L.; Chifotides, H. T.; Dunbar, K. R. Anion- interactions, Chem. Soc. Rev.
2008, 37, 6883.
Halogen Bonding in the Assembly 273

[5] Zhao, L.; Chen, X.-D.; Mak, T.C.W. Silver(I)-ethynide supramolecular synthons in the
assembly of - stacked infinite columns, Organometallics 2008, 27, 24832489.
[6] Nishio, M.; Hirota, M.; Umezawa, Y. The CH/ Interaction: Evidence, Nature and
Consequences, Wiley-VCH, Weinheim, 1998.
[7] Espallargas, G. M.; Zordan, F.; Marn, L. A.; Adams, H.; Shankland, K.; van de Streek,
J.; Brammer, L. Rational Modification of the Hierarchy of Intermolecular Interactions
in Molecular Crystal Structures by Using Tunable Halogen Bonds, Chem. Eur. J. 2009,
15, 75547568.
[8] Goldberg, I. Metalloporphyrin molecular sieves, Chem. Eur. J. 2000, 6, 38633870.
[9] Telfer, S. G.; Wuest, J. D. Metallotectons: Comparison of Molecular Networks Built
from Racemic and Enantiomerically Pure Tris(dipyrrinato)cobalt(III) Complexes,
Cryst. Growth Des. 2009, 9, 19231931.
[10] Khavasi, H. R.; Fard, M. A. - Interactions Affect Coordination Geometries, Cryst.
Growth Des. 2010, 10, 18921896.
[11] Zheng, Y.-Z.; Speldrich, M.; Kgerler, P.; Chen, X.-M. The role of - stacking in
stabilizing a,a-trans-cyclohexane-1,4-dicarboxylatein a 2D Co(II) network, Cryst. Eng.
Comm. 2010, 12, 10571059.
[12] Metrangolo, P.; Resnati, G. Halogen Bonding, in Encyclopedia of Supramolecular
Chemistry, Marcel Dekker, New York, 2004, 628635.
[13] Metrangolo, P.; Resnati, G.; Pilati, T.; Biella, S.
Halogen Bonding: Fundamentals and
Applications in Structure and Bonding, Springer, Berlin, 2008, 126, 105136.
[14] Fourmigu, M.; Batail, P. Activation of Hydrogen- and Halogen-Bonding Interactions
in Tetrathiafulvalene-Based Crystalline Molecular Conductors, Chem. Rev. 2004, 104,
53795418.
[15] Guthrie, F. On the Iodide of Iodammonium. J. Chem. Soc. 1863, 16, 239-244.
[16] Metrangolo, P.; Neukirch, H.; Pilati, T.; Resnati, G. Halogen bonding based recognition
processes: A world parallel to hydrogen bonding , Acc. Chem. Res., 2005, 38, 386395.
[17] Metrangolo, P.; Carcenac, Y.; Lahtinen, M.; Pilati, T.; Rissanen, K.; Vij, A.; Resnati, G.
Nonporous Organic Solids Capable of Dynamically Resolving Mixtures of
Diiodoperfluoroalkanes, Science 2009, 323, 14611464.
[18] Awwadi, F. F.; Willett, R. D.; Peterson, K. A.; Twamley, B. The nature of
HalogenHalide synthons: Theoretical and crystallographic studies, J. Phys. Chem. A
2007, 111, 23192328.
[19] George, S.; Nangia, A.; Lam, C.-K.; Mak, T. C. W.; Nicoud, J.-F. Crystal engineering
of urea a-network via IO2N synthon and design of SHG active crystal N-4-
iodopheny-N '-4 '-N '-nitrophenylurea, Chem. Commun. 2004, 12021203.
[20] Auffinger, P.; Hays, F. A.; Westhof, E.; Ho, P. S. Halogen bonds in biological
molecules, Proc. Natl. Acad. Sci. USA 2004, 101, 1678916794.
[21] Sarwar, M. G.; Dragisic, B.; Sagoo, S.; Taylor, M. S. A Tridentate Halogen-Bonding
Receptor for Tight Binding of Halide Anions, Angew. Chem. Int. Ed. 2010, 49,
16741677.
[22] Prsang, C.; Loc Nguyen, H.; Horton, P. N.; Whitwood, A. C.; Bruce, D. W. Trimeric
liquid crystals assembled using both hydrogen and halogen bonding, Chem. Commun.
2008, 61646166.
[23] Reddy, C. M.; Kirchner, M. T.; Gundakaram, R. C.; Padmanabhan, K. A.; Desiraju, G.
R. Isostructurality, polymorphism and mechanical properties of some hexahalogenated
274 Ya-Juan Fan, Shuang-Quan Zang, Yong-Li Wei et al.

benzenes: The nature of halogenhalogen interactions, Chem. Eur. J. 2006, 12,


22222234.
[24] Cariati, E.; Forni, A.; Biella, S.; Metrangolo, P.; Meyer, F.; Resnati, G.; Righetto, S.;
Tordin, E.; Ugo, R. Tuning second-order NLO responses through halogen bonding,
Chem. Commun. 2007, 25902592.
[25] Laguna, A.; Lasanta, T.; Lpez-de-Luzuriaga, J. M.; Monge, M.; Naumov, P.; Olmos,
M. E. Combining Aurophilic Interactions and Halogen Bonding To Control the
Luminescence from Bimetallic Gold-Silver Clusters, J. Am. Chem. Soc. 2010, 132,
456457.
[26] Shirman, T.; Freeman, D.; Posner,Y. D.; Feldman, I.; Facchetti, A.; van der Boom, M.
E. Assembly of crystalline halogen-bonded materials by physical vapor deposition, J.
Am. Chem. Soc. 2008, 130, 81628163.
[27] Brammer, L.; Espallargas, G. M.; Libri S. Combining metals with halogen bonds,
Cryst. Eng. Comm. 2008, 10, 17121727.
[28] Aakery, C. B.; Desper, J.; Helfrich, B. A.; Metrangolo, P.; Pilati, T.; Resnati, G.;
Stevenazzib, A. Combining halogen bonds and hydrogen bonds in the modular
assembly of heteromeric infinite 1-D chains, Chem. Commun. 2007, 42364238.
[29] Bruce, D. W.; Metrangolo, P.; Meyer, F.; Prsang, C.; Resnati, G.; Terraneo, G.;
Whitwood, A. C. Mesogenic, trimeric, halogen-bonded complexes from
alkoxystilbazoles and 1,4-diiodotetrafluorobenzene, New J. Chem. 2008, 32, 477482.
[30] Cini, D.; Frii, T.; Jones W. Isostructural Materials Achieved by Using Structurally
Equivalent Donors and Acceptors in Halogen-Bonded Cocrystals, Chem. Eur. J. 2008,
14, 747 753.
[31] Arman, H. D.; Gieseking, R. L.; Hanks, T. W.; Pennington, W. T. Complementary
halogen and hydrogen bonding: sulfuriodine interactions and thioamide ribbonsw,
Chem. Commun. 2010, 46, 18541856.
[32] Desiraju, G. R.; Parthasarathy, R. The Nature of HalogenHalogen Interactions: Are
Short Halogen Contacts Due to Specific Attractive Forces or Due to Close Packing of
Nonspherical Atoms? J. Am. Chem. Soc. 1989, 111, 87258726.
[33] Schollmeyer, D.; Shishkin, O. V.; Ruhl, T.; Vysotsky, M. O. OH- and halogen-
interactions as driving forces in the crystal organisations of tri-bromo and tri-iodo trityl
alcohols, Cryst. Eng. Comm. 2008, 10, 715723.
[34] Shishkin, O. V. Evaluation of true energy of halogen bonding in the crystals of halogen
derivatives of trityl alcohol, Chem. Phys. Lett. 2008, 458, 96100.
[35] Yin, Z.-M.; Wang, W.-Y.; Du, M.; Wang, X.-G.; Guo, J.-H. Crystal engineering of 5,5'-
bisdiazo-dipyrromethane with halogen synthons, Cryst. Eng. Comm. 2009, 11,
24412446.
[36] Corradi, E.; Meille, S.V.; Messina, M.T.; Metrangolo, P.; Resnati, G. Halogen Bonding
versus Hydrogen Bonding in Driving Self-Assembly Processes, Angew. Chem. Int. Ed.
2000, 39, 17821786.
[37] Lu, J. Y.; Babb, A. M. A Simultaneous Reduction, Substitution, and Self-Assembly
Reaction under Hydrothermal Conditions Afforded the First Diiodopyridine Copper(I)
Coordination Polymer, Inorg. Chem. 2002, 41, 13391341.
[38] Zhong, Y.-R.; Cao, M.-L.; Mo, H.-J.; Ye, B.-H. Syntheses and Crystal Structures of
Metal Complexes with 2,2-Biimidazole-like Ligand and Chloride: Investigation of X-
Halogen Bonding in the Assembly 275

HCl (X=N, O, and C) Hydrogen Bonding and Cl- (imidazolyl) Interactions, Cryst.
Growth Des. 2008, 8, 22822290.
[39] Notash, B.; Safari, N.; Khavasi, H. R. Anion-Directed Self-Assembly in Coordination
Networks: Architectural Control via Cooperative Noncovalent Interactions, Inorg.
Chem. 2010, 49, 1141511420.
[40] Zordan, F.; Purver, S. L.; Adams, H.; Brammer, L. Halometallate and halide ions:
nucleophiles in competition for hydrogen bond and halogen bond formation in
halopyridinium salts of mixed halidehalometallate anions, Cryst. Eng. Comm. 2005, 7,
350354.
[41] G. M. Espallargas,;L. Brammer,; P. Sherwood, Designing Intermolecular Interactions
between Halogenated Peripheries of Inorganic and Organic Molecules: Electrostatically
Directed M_XX_C Halogen Bonds, Angew. Chem. Int. Ed. 2006, 45, 435440.
[42] Derossi, S.; Brammer, L.; Hunter, C. A.; Ward, M. D. Halogen Bonded Supramolecular
Assemblies of [Ru(bipy)(CN)4]2- Anions and N-Methyl-Halopyridinium Cations in the
Solid State and in Solution, Inorg. Chem. 2009, 48, 16661677.
[43] Imakubo, T.; Tajima, N.; Tamura, M.; Kato, R.; Nishio, Y.; Kajita, K. A
supramolecular superconductor -(DIETS)2[Au(CN)4], J. Mater. Chem. 2002, 12,
159161.
[44] Thoyon, D.; Okabe, K.; Imakubo, T.; Golhen, S.; Miyazaki, A.; Enoki, T.; Ouahab, L.
Conducting Materials Containing Paramagnetic Hexacyanometallate [Cr(CN)6]3- and
Iodine Substituted Organic Donor [DIETS], Mol. Cryst. Liq. Cryst. 2002, 376, 2532.
[45] Imakubo, T.; Miyake, A.; Sawa, H.; Kato, R. Synthesis and Physical Properties of
(DIETS)2[Au(CN)4]: A New -salt with a Unique DonorAnion Network, Synth. Met.
2001, 120, 927928.
[46] Bertani, R.; Sgarbossa, P.; Venzo, A.; Lelj, F.; Amati, M.; Resnati, G.; Pilati, T.;
Metrangolo, P.; Terraneo, G. Halogen bonding in metalorganicsupramolecular
networks, Coord. Chem. Rev. 2010, 254, 677695.
[47] Zang, S.-Q.; Cheng, P.-S.; Mak, T. C. W. SilverXaryl (X= I and Br) interaction in a
network assembly with a flexible polynuclear silverethynide supramolecular synthon,
Cryst. Eng. Comm. 2009, 11, 10611067.
[48] Muniappan, S.; Lipstman, S.; Goldberg, I. Rational design of supramolecular chirality
in porphyrin assemblies: the halogen bond case, Chem. Commun. 2008, 1777-1779.
[49] Lipstman, S.; Muniappan, S.; Goldberg, I. Supramolecular Reactivity of Porphyrins
with Mixed Iodophenyl and Pyridyl meso-Substituents, Cryst. Growth Des. 2008, 8,
16821688.
[50] Smart, P.; Espallargas, G. M.; Brammer, L. Competition between coordination network
and halogen bond network formation: towards halogen-bond functionalised network
materials using copper-iodobenzoate units, Cryst. Eng. Comm. 2008, 10, 13351344.
[51] S.-Q. Zang,; Y.-J. Fan,; J.-B. Li,; H.-W. Hou,; and T. C. W. Mak, Halogen Bonding in
the Assembly of Metal-Organic Networks Based on 5-Iodo-Isophthalic Acid, submitted
to Cryst. Growth Des.
[52] Batten, S. R.; Robson, R. Interpenetrating Nets: Ordered, Periodic Entanglement,
Angew. Chem., Int. Ed., 1998, 37, 14601494.
[53] Batten, S. R. "Interpenetration" in Encyclopedia of Supramolecular Chemistry (Eds.
Atwood, J.L.; Steed, J.W.), Marcel Dekker, New York, USA, 2004, 735741.
276 Ya-Juan Fan, Shuang-Quan Zang, Yong-Li Wei et al.

[54] Thaimattam, R.; Sharma, C. V. K.; Clearfield, A.; Desiraju, G. R. Diamondoid and
Square Grid Networks in the Same Structure. Crystal Engineering with the IodoNitro
Supramolecular Synthon, Cryst. Growth Des. 2001, 1, 103106.
[55] Liantonio, R.; Metrangolo, P.; Meyer, F.; Pilati, T.; Navarrini, W.; Resnati, G. Metric
engineering of supramolecular Borromean rings, Chem. Commun., 2006, 18191821.
[56] Metrangolo, P.; Meyer, F.; Pilati, T.; Proserpio, D. M.; Resnati, G. Highly
interpenetrated supramolecular networks supported by NI halogen bonding, Chem.
Eur. J., 2007, 13, 57655772.
[57] Garca, M. D.; Blanco, V.; Platas-Iglesias, C.; Peinador, C.; Quintela, J. M. Interplay
between Halogen/Hydrogen Bonding and Electrostatic Interactions in 1,1-Bis(4-
iodobenzyl)-4,4-bipyridine-1,1-diium Salts, Cryst. Growth Des. 2009, 9, 50095013.
In: Coordination Polymers and Metal Organic Frameworks ISBN: 978-1-61470-899-5
Editors: O. L. Ortiz and L. D. Ramirez, pp. 277-290 2012 Nova Science Publishers, Inc.

Chapter 9

SUBTRACTIVE APPROACH FOR INTRODUCING


FUNCTIONAL GROUPS ONTO
METALORGANIC FRAMEWORK

Teppei Yamada and Hiroshi Kitagawa


Division of Chemistry, Graduate School of Science, Kyoto University
Kitashirakawa-Oiwakecho, Sakyo-ku, Kyoto 606-8502, JAPAN.

1. INTRODUCTION
The aim of this work is to apply a new procedure in the synthesis of metalorganic
frameworks (MOFs), where the organic moiety bears functional groups that can be involved
in the proton conduction mechanism. Several method can be applied for introducing
functional groups onto MOFs, including a naive method which simply uses a ligand having
acidic functional groups for constructing a MOF, and several papers succeeded in introducing
it. However, acidic functional groups have ability to coordinate to metal site as well as other
functional groups for coordinating, and the obtained MOF structure is a sort of creation of a
chance.
Afterwards, use of a post-synthetic method in functionalizing MOFs has been developed
by research groups of Cohen, Cronin, Gamez, and Yaghi in parallel while various functional
groups were introduced using cl ick chemistry by Sada and co-workers. [1-6] However there
are still considerable difficulties associated with post-synthetic methods. For example, pore
volume is decreased by the functionalization reaction, and the gas absorption capability can
be reduced. A functionalization reaction commences at the peripheral part of the crystals, and
transport of a reactant to the central part tends to be hindered by a decrease in the pore
diameter, which sometimes results in loss of crystallinity.
Here we proposed a third way, a novel three-step procedure for introducing non-
coordinating hydroxyl groups into MOFs instead of using post-synthetic method, as
schematically represented in Figure 1. The three steps include the following: (a) a pre-
reaction, i.e. protection of the functional groups by introduction of protecting groups
(abbreviated as P reaction), (b) a complexation reaction of a MOF (C reaction), and (c) a
278 Teppei Yamada and Hiroshi Kitagawa

deprotection reaction as a post-synthetic process (D reaction). In this article we reviewed


the method, which we call a protectioncomplexationdeprotection (PCD) method.
We focused on a MOF structure consisting of the terephthalate bears hydroxyl groups, for
example 2,5-dihydroxy-terephthalate. In the literature there are several reports of stable
frameworks using 2,5-dihydroxy-terephthalate [7-9], but in these cases all the oxygen atoms
of this ligand are involved in coordination. Our aim is to have the two hydroxyl groups
remaining dangling in the framework and having hydrogen-bonding interactions with the
solvent to generate effective proton conducting pathways.
We protect hydroxyl groups of 2,5-dihydroxyterephthalic acid (H2dhybdc) by various
acyl groups and protected dicarboxylic acids were applied for complexation reactions. In the
former part of this article, IRMOFs are synthesized, and Rietveld refinement and spectral
analysis revealed that protected dicarboxylate is useful to construct IRMOF while non-
protected dhybdc resulted in amorphous compounds.
In the later part, the same protected ligand are applied to complexation reaction with zinc
and cadmium ions and single-crystal X-ray diffraction measurements of the obtained crystals
revealed that the non-coordinating hydroxyl groups were successfully introduced into the
frameworks of [Zn(dhybdc)(bpy)]4DMF (bpy = 4,4-bipyridine, DMF = N,N-dimethyl
formamide) and [Cd(dhybdc)(bpy)]. These results will provide the novel MOFs consisting
hydrogen bond networks within them.

Figure 1. Schematic illustrations of the steps in the protectioncomplexation deprotection (PCD)


method for preparing MOFs. Gray, red and blue balls represent carbon, oxygen and metal atoms,
respectively. Yellow balls represent acetyl groups computationally put into the void space of the
crystals.

2. RESULTS AND DISCUSSIONS


Synthesis and Structure of [Zn4O(Dacobdc)3]

First, both hydroxyl groups were protected with acetyl or pivaloyl groups. Given their
stability, it is suitable for protection and deprotection reactions. H2dacobdc was prepared by
Subtractive Approach for Introducing Functional Groups 279

the acetylation of hydroxyl groups with acetic anhydride, and H2dpivobdc with pivaloyl
chloride as shown in the experimental section.
The reaction of H2dhybdc, H2dacobdc with zinc ion and triethylamine results in the
precipitation of the IRMOF-1 derivatives in a similar way as the reaction of terephthalic acid
(H2bdc), zinc and triethylamine. XRD patterns of the MOFs obtained from H2bdc, H2dhybdc
and H2dacobdc are shown in Figure 2. These novel MOF decompose by exposure to air, thus
XRD patterns were measured when the compounds were still wet with DMF. Elemental
analyses were executed on dried samples after exposing to air, and the results indicate that the
specimens absorb water molecules from the moist air. Difference of elemental analyses is
observed with increase of the reaction time. The results suggest that prolonged reaction time
results in lower organic content, which may be associated with the formation of zinc oxide.
The reaction of H2dpivobdc results in no precipitates, which suggest that the pivaloyl
group prevents the constructing of MOF structure due to its steric hindrance.

(a)

(b)

(c)

(d)

5 10 15 20 25 30
2 / degree
Figure 2. (a) XRD pattern simulated from the reported crystal of IRMOF-1, obtained patterns of
IRMOF-1 (b), precipitate from the reaction of H2dhybdc and zinc (c) and that of H2dacobdc and zinc
(d).

OH
O OH O OH
HO O HO O
HO

Zn
O
Zn O O Zn Zn O O Zn
Zn O O Zn Zn O O Zn
O
Zn
Figure 3. Difference between the arrangements of coordination in bdc and dhybdc.
280 Teppei Yamada and Hiroshi Kitagawa

As shown in Figure 2, only broad peaks at 2 = 8 were observed for the complex
consisting of dhybdc, indicating that the sample is amorphous. Dhybdc ligand has hydroxyl
groups as well as carboxylate, thus some hydroxyl groups also coordinate to the zinc ions in
amorphous fashion, as shown in Figure 3. XRD pattern of [Zn4O(dacobdc)3] was quite similar
to that of IRMOF-1 indicating to have the same [Zn4O(dicarboxylate)3] framework. Space
group of [Zn4O(dacobdc)3] was proposed to Fm-3m, and lattice parameter was well fitted to a
= 25.832 by Pawley method. From these results, by protecting these hydroxyl groups in
dhybdc with acetyl groups, IRMOF structure was formed with prevention of the coordination
of hydroxyl groups.
The vibrational mode of C=O in acetyl group was observed at 1770 cm1 in the infrared
spectrum, showing that acetyl group was introduced (Figure 4e). From these results, acyl
groups in dacobdc ligand are not deprotected under the reaction conditions.

(a)

(b)

(c)

(d)

(e)

(f)

2000 1500 1000 500


Wavenumber / cm 1
Figure 4. IR spectra of IRMOFs and corresponding ligands; (a) IRMOF-1, (b) H2bdc, (c) precipitate
from the reaction of H2dhybdc and zinc, (d) H2dhybdc, (e) precipitate of H2dacobdc and zinc, and (f)
H2dacobdc.

Figure 5. (Right) Results of Rietveld refinement. Red line, black dots, green bars and blue lines
represent simulated and experimental XRD pattern, peak indices and residue, respectively. (Left)
Obtained crystal structure of [Zn4O(dacobdc)3]. Gray, red and pale blue balls represents carbon, oxygen
and zinc atoms, respectively. Solvents are omitted for clarity.
Subtractive Approach for Introducing Functional Groups 281

Figure 6. Summarized scheme of PCD reaction applied for IRMOF.

Rietveld refinement was executed for the XRPD pattern of [Zn4O(dhybdc)3] for analysis
of the structure. The result is shown in Figure 5. The simulated pattern from the Rietveld
refinement is in good agreement with the experimental one.
From these results, unprotected dhybdc and zinc resulted in the amorphous compound
due to the hydroxyl groups; however protected dacobdc prevented hydroxyl group from
complexation. The deprotection reaction was tried; however IRMOFs suffered decomposed
by nucleophilic agents and the reaction was not achieved with IRMOFs (Figure 6).

Syntheses and Crystal Structures of [Zn(Dhybdc)(Bpy)]4DMF and


[Cd(Dhybdc)(Bpy)]

Because IRMOFs are found to be decomposed to the aqueous condition, and pillared-
layer type MOF was applied for the PCD reaction. The reaction of H2dacobdc,
Zn(NO3)6H2O, and bpy in DMF afforded good block crystals of [Zn(dhybdc)(bpy)]4DMF.
Cadmium derivative ([Cd(dhybdc)(bpy)]) was obtained under similar reaction conditions.
The structures of [Zn(dhybdc)(bpy)]4DMF and [Cd(dhybdc)(bpy)] were determined by
single-crystal X-ray diffraction analyses, the details of which are given in Table 1.
[Zn(dhybdc)(bpy)]4DMF has a pillared-layer structure (Figure 7), in which each two-
dimensional (2D) layer consists of a square grid of dimeric Zn bridged by the dhybdc. One
carboxylate of each dhybdc coordinates in a monodentate fashion to one zinc ion at a distance
of 1.989 , and the other carboxylate coordinates to two separate zinc ions at distances of
2.004 and 2.009 (Figure 8). Thus, each zinc ion has a trigonal bipyramidal coordination
geometry, with one monodentate carboxylate oxygen atom, two bridging carboxylate oxygen
atoms, and two nitrogen atoms from the bpy pillars (Figure 8). One oxygen atom of
carboxylate ligand is located closely to zinc ion (2.787 ), indicating the weak bonding to it,
and 2D sheet of Zn(dhybdc) was distorted by the partial bonds. The coordination geometry in
[Zn(dhybdc)(bpy)]4DMF is different from those in the usual coordination compounds of
zinc ions with dhybdc [10-12] and in pillared-layer MOFs previously reported for the
terephthalate derivatives. [13-16] Four DMF molecules are captured in each crystal lattice.
282 Teppei Yamada and Hiroshi Kitagawa

Table 1. Crystallographic parameters of [Zn(dhybdc)(bpy)]4DMF [17] and


[Cd(dhybdc)(bpy)]

Empirical Formula ZnC30H40O10N6 CdC18H12O6N2


Formula weight 710.08 464.71
T, K 100(2) 100(2)
Crystal system tetragonal orthorhombic
Space group P-421c (#114) P21212 (#18)
a, 17.476(2) 13.137(5)
b, 17.476(2) 21.272(5)
c, 22.687(2) 11.640(5)
V, 3 6928.8(13) 3253(2)
Z 8 4
Crystal size, mm3 0.2 0.1 0.1 0.3 0.1 0.1
Unique reflections 4709 4690
Parameters 465 245
3
Dcalc 1.36 1.546
F(000) 3012 1560
(Mo K), cm1 11.106 7.37
R1 [I > 2 0.0546 0.0531
wR2 [all reflections] 0.1538 0.1227
Goodness of fit 1.030 1.073
Max Shift/Error 0.003 0.000
Max peak, e 3 1.20 3.070
Min peak, e 3 0.32 0.624

Figure 7. Crystal structure of [Zn(dhybdc)(bpy)]4DMF: (top) view along the c-axis; (bottom)
perspective view of 2D layers and pillars. Solvent molecules have been omitted for clarity. Gray, red,
and blue and pale blue balls represent carbon, oxygen, nitrogen and zinc atoms respectively. [17].
Subtractive Approach for Introducing Functional Groups 283

The structure of [Cd(dhybdc)(bpy)] is shown in Figures 9 and 10. The structure is


roughly the same as [Zn(dhybdc)(bpy)]4DMF, consisting of 2D square grid of Cd(dhybdc)
pillared by bpy. One carboxylate group of each dhybdc coordinates in a bidentate fashion to
one cadmium ion at distances of 2.290 and 2.467 . The other carboxylate coordinates to two
separate cadmium ions at distances of 2.359 and 2.384 (Figure 9). Thus, each cadmium ion
has a distorted octahedral coordination geometry, with one bidentate carboxylate oxygen
atom, two bridging carboxylate oxygen atoms, and two nitrogen atoms from the bpy pillars.
The difference in coordination between Zn and Cd is probably due to the ionic radii. 2D
sheets of Cd(dhybdc) is flat, unlike the Zn compound (Figure 10). DMF molecules are
disordered and SQUEEZE technique was applied.

Figure 8. (top) Crystal structure of [Zn(dhybdc)(bpy)]4DMF in the ab plane. (bottom) Coordination


geometry of zinc ions. Solvent molecules have been omitted for clarity. [17]

Figure 9. Crystal structure of [Cd(dhybdc)(bpy)] along the c-axis. Gray, red and yellow balls represent
carbon, oxygen and cadmium atoms, respectively.
284 Teppei Yamada and Hiroshi Kitagawa

Figure 10. Perspective view of 2D layers and pillars of [Cd(dhybdc)(bpy)]. Solvent molecules have
been omitted for clarity. Gray, red, blue and yellow balls represent carbon, oxygen, nitrogen and
cadmium atoms, respectively.

Unexpectedly and fortuitously, the acetoxyl groups were totally removed, and the dhybdc
ligands were observed in both [Zn(dhybdc)(bpy)]4DMF and [Cd(dhybdc)(bpy)], indicating
that the C and D reactions were one-pot reactions and that the non-coordinating hydroxyl
groups were successfully introduced into the cavities. Each hydroxyl group of the ligand is
intramolecularly and intermolecularly hydrogen-bonded to an adjacent carboxylate group and
contributes to the stabilization of the 2D layer structure of [Zn(dhybdc)(bpy)]4DMF or
[Cd(dhybdc)(bpy)].
Figure 11 shows the result of TG analysis of [Zn(dhybdc)(bpy)]4DMF. As shown in the
figure, 45% loss of weight can be observed stepwise until 300 C corresponding to the four
equivalents of DMF.

Figure 11. TG plot of [Zn(dhybdc)(bpy)]4DMF. [17].

Spectroscopic Analyses

Electronic and infrared spectra of [Zn(dhybdc)(bpy)]4DMF are shown in Figures 12 and


13. Absorption peak can be observed at 360 nm, corresponding to -* transition of dhybdc
(Figure 12). In Figure 13, the stretching mode of the hydroxyl group was observed at 3450
Subtractive Approach for Introducing Functional Groups 285

cm1 and one of carbonyl groups cannot be observed around 1800 cm1, confirming the
presence of non-coordinated hydroxyl groups and the elimination of acetyl groups.

Figure 12. Electronic spectrum of [Zn(dhybdc)(bpy)]4DMF. [17].

Figure 13. FT-IR spectrum of [Zn(dhybdc)(bpy)]4DMF. [17].

Elucidation of Deprotection Condition

Although the MOFs obtained here suffered deprotection reaction during the complexation
reaction, we concluded that deprotection reaction does not occur before the complexation
reaction but occurs simultaneously or after the crystal growth of [Zn(dhybdc)(bpy)] or
[Cd(dhybdc)(bpy)], the deprotected MOFs. The first reason is the NMR analysis of
deprotection conditions of H2dacobdc itself (Figure 14). As shown in the figure, rate of the
deprotection reaction is very slow under the same conditions as the MOF preparation, and at
least one of the hydroxyl groups is protected before the complexation reaction occurs. Thus,
the deprotection reaction of the ligands should proceed simultaneously with crystal growth or
286 Teppei Yamada and Hiroshi Kitagawa

might occur after the crystal growth. In addition, X-ray powder diffraction peaks
corresponding to [Zn(dhybdc)(bpy)] were not observed in a sample obtained from a mixture
of H2dhybdc, bpy, and Zn(NO3)6H2O.

Figure 14. Longitudinal changes of proportions of H2dacobdc (circle), H2acohybdc (diamond) and
H2dhybdc (square). [17].

The another collateral reason is the shape of the voids in the MOFs. Conventional MOFs
consisting of Zn, terephthalate, and bpy form interpenetrated structures, and consequently
only a small pore (ca. 1 1 ) can be constructed, [14] while [Zn(dhybdc)(bpy)] and
[Cd(dhybdc)(bpy)] synthesized here does not display an interpenetrated structure (Figure 7).
MOFs show very high tendency to construct interpenetrating structure when the connector
ligands are long, as is the case here, and it is difficult to avoid the interpenetration by the
reported method before. If the acetyl group is present during the complexation as discussed
here, then its bulkiness would prevent the interpenetration. This offers further support to the
conclusion that the D reaction process occurs during or after the coordination reaction.
Consequently, the PCD method can be used to prevent interpenetration and to obtain
materials with larger pores and also wider apertures (Figure 15).

Figure 15. Schematic representation of PCD method applied to [Zn(dhybdc)(bpy)].


Subtractive Approach for Introducing Functional Groups 287

3. SUMMARY
A novel reaction strategy has been executed for IRMOF and pillared-layer type MOFs.
Protected dacobdc ligand offers the novel IRMOF; however deprotection reaction was not
achieved. Pillared-layer type [Zn(dhybdc)(bpy)]4DMF and [Cd(dhybdc)(bpy)] were
synthesized by a protectioncomplexationdeprotection (PCD) method. This procedure
prevents interpenetration and coordination of hydroxyl groups to zinc ions, and uncoordinated
hydroxyl groups were successfully introduced in the pores.
This procedure for the modification of MOFs offers several advantages. Acidic functional
groups can be protected from coordination reactions. The modification of acidic groups in
MOFs is of significance for tuning the ionic conductivity [18-22] or cation exchange
capability. [23,24] Bulky protective groups also prevent the interpenetration reaction.
Avoidance of the interpenetration reaction is vital for the establishment of pores with
controlled aperture diameters, which critically affect the gas selection or molecular motion in
the frameworks. After the D reaction, the pore size increases. Enlarged pores may improve
transport of reactants and byproducts. The reaction in the MOFs is then expected to run to
completion.
This PCD method provides us a sophisticated technique to obtain functionalized MOFs
with large pore aperture. The most promising future is a cation exchange MOF which requires
acidic functional groups and large pore. It will open the door for a solid acid catalyst and a
proton conductor which are analogous to ion exchange polymer resign.

4. EXPERIMENTAL
2,5-Dihydroxyterephthalic acid was purchased from Aldrich. Sulfuric acid, N,N-
dimethylformamide (DMF) and acetic anhydride were purchased from Wako pure chemicals.
4,4-Bipyridine was purchased from Tokyo Chemical Industry. Zinc nitrate hexahydrate was
purchased from Kishida Reagents Chemicals. All chemicals were used without further
purification.

Preparation of 2,5-Diacetoxyterephthalic Acid (H2dacobdc)

Figure 16. Synthetic scheme of H2dacobdc.[17].


288 Teppei Yamada and Hiroshi Kitagawa

A suspension of 2,5-dihydroxyterephthalic acid (0.99 g, 5.0 mmol) and excess acetic


anhydride (2.50 g, 24 mmol) were warmed to 50 C to which were added several drops of
conc. sulfuric acid. The precipitate was filtered, washed with ethanol and dried in vacuo. A
faint yellow powder was obtained (800 mg, 56.7%). Molecular ratio of un-reacted compound
was estimated to be lower than 1% by 1H NMR spectroscopy, and the compound was used to
following reaction without further purification. Elemental analyses calculated (found) % for
C8H6O6: C 51.07 (50.87); H 3.57 (3.63). 1H-NMR / ppm; 13.63 (br, 2H), 7.68(s, 2H),
2.25(s, 6H).

Preparation of 2,5-Bis(Pivaloyloxy)Terephthalic Acid (H2dpivobdc)

Figure 17. Synthetic scheme of H2dpivobdc.

2,5-dihydroxyterephthalic acid (198 mg, 0.10 mmol) was suspended to DMF (1 mL) and
warmed at 50 C, and the suspension was dissolved. Excess pivaloyl chloride (250 mg, 2.1
mmol) was added to the solution. After 30 minutes of stirring, excess pivaloyl chloride and
DMF was eliminated under reduced pressure, and the remaining suspension was filtered off,
washed carefully with ethanol and ether, and H2dpivobdc was obtained as white powder.
Elemental analyses calculated (found) % for C8H6O6: C 51.07 (50.87); H 3.57 (3.63). 1H-
NMR / ppm; 13.63 (br, 2H), 7.68(s, 2H), 2.25(s, 6H).

Preparation of [Zn(Dhybdc)(Bpy)]4DMF

Zn(NO3)26H2O (119 mg, 0.40 mmol), H2dacobdc (89 mg, 0.32 mmol) and 4,4-
bipyridine (31 mg, 0.20 mmol) were dissolved in 5 mL of N,N-dimethylformamide and the
solution was heated to 55 C for a week and then cooled slowly to room temperature.
Yellowish block crystals were obtained.

Preparation of [Cd(Dhybdc)(Bpy)]

Cd(NO3)24H2O (123 mg, 0.40 mmol), H2dacobdc (89mg, 0.32 mmol) and 4,4-
bipyridine (31 mg, 0.20 mmol) were dissolved in 5 mL of N,N-dimethylformamide and the
solution was heated to 55 C for two week and then cooled slowly to room temperature.
Yellowish block crystals were obtained.
Subtractive Approach for Introducing Functional Groups 289

Preparation of [Zn4O(Dacobdc)3]

Triethylamine (1 mL) was added dropwise to the stirred solution of Zn(NO3)26H2O (116
mg, 0.40 mmol) and H2dacobdc (85 mg, 0.30 mmol) in DMF (5 mL). White precipitate was
formed instantly. After 30 seconds, the suspension was filtered off, washed with DMF
thoroughly, and dried. Elemental analyses calculated (found) % for
[Zn4O(dacobdc)3](DMF)2(H2O)4: C 37.75 (37.54); H 3.47 (3.58); N 2.10 (2.18).

Evaluation of Deprotection Condition

H2dacobdc (5.64 mg, 0.02 mmol) and bpy (1.56 mg, 0.01 mmol) were dissolved into
DMF-d7 (0.75 mL) in an NMR tube and 1H-NMR spectrum was measured. The tube was kept
at 55 C and 1H-NMR measurements were executed several times. Molecular ratios of
H2dacobdc, 2-acetoxy-5-hydroxyterephthalic acid (H2acohybdc) and H2dhybdc were
evaluated from the areas of their peaks of phenyl protons.

REFERENCES
[1] Song, Y.-F.; Cronin, L. Angew. Chem. Int. Ed. 2008, 47, 4635-4637.
[2] Ingleson, M. J.; Barrio, J. P.; Bacsa, J.; Dickinson, C.; Park, H.; Rosseinsky, M. J.
Chem. Commun. (Cambridge, U. K.) 2008, 1287-1289.
[3] Costa, J. S.; Gamez, P.; Black, C. A.; Roubeau, O.; Teat, S. J.; Reedijk, J. Eur. J. Inorg.
Chem. 2008, 1551.
[4] Haneda, T.; Kawano, M.; Kawamichi, T.; Fujita, M. J. Am. Chem. Soc. 2008, 130,
1578-1579.
[5] Wang, Z.; Cohen, S. M. J. Am. Chem. Soc. 2007, 129, 12368-12369.
[6] Goto, Y.; Sato, H.; Shinkai, S.; Sada, K. J. Am. Chem. Soc. 2008, 130, 14354-14355.
[7] Rosi, N. L.; Kim, J.; Eddaoudi, M.; Chen, B.; O'Keeffe, M.; Yaghi, O. M. J. Am. Chem.
Soc. 2005, 127, 1504-1518.
[8] Dietzel, P. D. C.; Morita, Y.; Blom, R.; Fjellvaag, H. Angew. Chem. Int. Ed. 2005, 44,
6354.
[9] Vitillo, J. G.; Regli, L.; Chavan, S.; Ricchiardi, G.; Spoto, G.; Dietzel, P. D. C.;
Bordiga, S.; Zecchina, A. J. Am. Chem. Soc. 2008, 130, 8386.
[10] Hasegawa, S.; Horike, S.; Matsuda, R.; Furukawa, S.; Mochizuki, K.; Kinoshita, Y.;
Kitagawa, S. J. Am. Chem. Soc. 2007, 129, 2607.
[11] Babarao, R.; Jiang, J. Langmuir 2008, 24, 6270.
[12] Hwang, Y. K.; Hong, D.-Y.; Chang, J.-S.; Jhung, S. H.; Seo, Y.-K.; Kim, J.; Vimont,
A.; Daturi, M.; Serre, C.; Ferey, G. Angew. Chem., Int. Ed. 2008, 47, 4144-4148.
[13] Qin, C.; Wang, X.; Carlucci, L.; Tong, M.; Wang, E.; Hu, C.; Xu, L. Chem. Commun.
2004, 1876.
[14] Tao, J.; Tong, M. L.; Chen, X. M. J. Chem. Soc., Dalton Trans. 2000, 3669.
[15] Ma, B.-Q.; Mulfort, K. L.; Hupp, J. T. Inorg. Chem. 2005, 44, 4912-4914.
[16] Dinca, M.; Long, J. R. J. Am. Chem. Soc. 2005, 127, 9376.
290 Teppei Yamada and Hiroshi Kitagawa

[17] Yamada, T.; Kitagawa, H. J. Am. Chem. Soc. 2009, 131, 6312-6313.
[18] Nagao, Y.; Fujishima, M.; Ikeda, R.; Kanda, S.; Kitagawa, H. Synth. Met. 2003, 133-
134, 431-432.
[19] Nagao, Y.; Ikeda, R.; Kanda, S.; Kubozono, Y.; Kitagawa, H. Mol. Cryst. Liq. Cryst.
2002, 379, 89-89.
[20] Nagao, Y.; Kubo, T.; Nakasuji, K.; Ikeda, R.; Kojima, T.; Kitagawa, H. Synth. Met.
2005, 154, 89-92.
[21] Fujishima, M.; Enyo, M.; Kanda, S.; Ikeda, R.; Kitagawa, H. Chem. Lett. 2006, 35,
546-547.
[22] Fujishima, M.; Ikeda, R.; Kanda, S.; Kitagawa, H. Mol. Cryst. Liq. Cryst. 2002, 379,
223-223.
[23] Klontzas, E.; Mavrandonakis, A.; Tylianakis, E.; Froudakis, G. E. Nano Lett. 2008, 8,
1572-1576.
[24] Blomqvist, A.; Arajo, C. M.; Srepusharawoot, P.; Ahuja, R. Proc. Natl. Acad. Sci. U.
S. A. 2007, 104, 20173-20176.
In: Coordination Polymers and Metal Organic Frameworks ISBN: 978-1-61470-899-5
Editors: O. L. Ortiz and L. D. Ramirez, pp. 291-295 2012 Nova Science Publishers, Inc.

Chapter 10

PERFORMANCE OF METAL-ORGANIC FRAMEWORK


MIL-101 IN THE LIQUID PHASE ADSORPTION
OF HETEROCYCLIC NITROGEN COMPOUNDS

Alexey L. Nuzhdin,a Konstantin A. Kovalenko,b


Vladimir P. Fedin,b and Galina A. Bukhtiyarovaa
a
Boreskov Institute of Catalysis, SB RAS, Novosibirsk, Russian Federation
b
Nikolaev Institute of Inorganic Chemistry, SB RAS, Novosibirsk,
Russian Federation.

ABSTRACT
The heterocyclic nitrogen compounds containing in liquid hydrocarbons streams
poison many industrial catalysts. The selective removal of the nitrogen species from
refinery streams by adsorption at ambient temperature is a promising approach.
Remarkable adsorption capacity and selectivity towards heterocyclic nitrogen compounds
were observed for metal-organic framework [Cr3O(C8H4O4)3F(H2O)2] (MIL-101) under
the sorption from isooctane and the hydrotreated gas oil. The adsorption capacity of MIL-
101 towards heterocyclic nitrogen compounds is significantly higher than the capacity of
conventional adsorbents such as activated carbons, activated alumina, silica-based
adsorbents and zeolite-based materials especially under the sorption from hydrocarbon
with low nitrogen content. The spent MIL-101 can be regenerated and reused in the next
adsorption cycle without loss of its adsorption capability. The very high adsorption
capacity and selectivity of metal-organic framework MIL-101 for the nitrogen
compounds, along with its good regenerability, indicate that MIL-101 may be the
promising adsorbent for deep denitrogenation of liquid hydrocarbon streams.

The elimination of nitrogen compounds from various refinery streams is of great practical
importance because the nitrogen compounds poison many industrial catalysts due to the
strong adsorption at the acidic sites [1]. The nitrogen compounds found in liquid

Tel.: (+7) 383 3269410 fax: (+7) 383 3308056; E-mail: anuzhdin@catalysis.ru.
292 Alexey L. Nuzhdin, Konstantin A. Kovalenko, Vladimir P. Fedin et al.

hydrocarbons streams are generally divided into two groups: basic aromatic heterocycles,
such as aniline, pyridine, quinoline, acridine, and their alkyl substituent derivatives, and
nonbasic aromatic heterocycles, such as pyrrole, indole, carbazole, and their alkyl substituent
derivatives. The selective removal of the nitrogen species from refinery streams by adsorption
at ambient temperature is a promising approach [2-5]. The selectivity means that the
prospective adsorbent should have high adsorption capacity towards the nitrogen compounds
rather than to the large fraction of aromatic compounds in the hydrocarbons streams.
The adsorption of heterocyclic nitrogen compounds by well-known classes of porous
materials such as activated carbons [2-5], activated alumina [2, 3], silica-based adsorbents [6,
7] and zeolite-based materials [8, 9] was extensively studied. The best results among these
materials were achieved for activated carbons. However, the adsorption capacities of
activated carbons are considerably decreased in the presence of a large amount of aromatic
compounds. In the past decade, metal-organic frameworks (MOFs) have attracted
considerable attention due to their unique combination of properties such as high surface area,
crystalline open structures, tunable pore size and functionality [10, 11]. Mesoporous
chromium(III) coordination polymer MIL-101 (Figure 1) possesses a rigid zeotype crystal
structure and extremely large surface area [12]. This material is built from Cr3O triangular
cluster complexes bridged by linear terephthalate linkers. Importantly, MIL-101 is resistant to
air, water, common solvents and thermal treatment (up to 320oC). All these unprecedented
features of MIL-101 justify its using as good adsorbent. Here we present data on superior
adsorption capacity of metal-organic framework MIL-101 towards heterocyclic nitrogen
compounds exceeding the capacity of conventional adsorbents.

Figure 1. Schematic representation of the crystal structure of [Cr3O(C8H4O4)3F(H2O)2] (MIL-101),


showing two major and two smaller pores of framework.

We investigated the adsorption of heterocyclic nitrogen compounds (quinoline, indole


and carbazole) over MIL-101 using two different type of solvent: pure isooctane and the
hydrotreated gas oil, containing about 8 ppmw S, 0.5 ppmw N; 21.7 wt% mono-, 1.9 wt% di-
and 0.8 wt% poli-aromatics. The adsorption capabilities of MIL-101 increase in the order
carbazole < indole < quinoline irrespective of solvent used (Figure 2). Interestingly, the
adsorption capability changes weakly when adsorption occurs from hydrotreated gas oil
instead of isooctane in spite of the presence of substantial amounts of aromatic compounds.
Performance of Metal-Organic Framework MIL-101 293

The result obtained let us to conclude that MIL-101 reveals very high adsorption selectivity
towards nitrogen compounds.

Figure 2. Adsorption isotherms for heterocyclic nitrogen compounds at 25 C on MIL-101.

We have evaluated the adsorption equilibrium constants, K (g/g) and the maximum
adsorption capacities, qm (mg-N/g-A) of MIL-101 for quinoline and indole under the sorption
from isooctane on the basis of Langmuir isotherms:

Kqm Ce
q
1 KCe

where q is the equilibrium adsorption capacity (milligrams of nitrogen per gram of adsorbent,
mg-N/g-A), Ce is the equilibrium concentration of nitrogen in the liquid phase (ppmw). To
obtain the adsorption parameters, the linear regression for a plot of Ce/q versus Ce was
constructed, as shown in Figure 3.

Figure 3. The plots of Ce/q vs Ce for quinoline and indole adsorption on MIL-101 from isooctane.
294 Alexey L. Nuzhdin, Konstantin A. Kovalenko, Vladimir P. Fedin et al.

The excellent linear correlation indicates that the adsorption follows the Langmuir
isotherm. The obtained qm and K values for MIL-101 and corresponding parameters for
activated carbons [13] are listed in Table 1.

Table 1. Parameters for adsorption nitrogen compounds over MIL-101


and activated carbons (AC) from alkanes on the bases of Langmuir isotherms

Adsorbent N- qm, K, g/g Equilibrium capacity,


compound mg-N/g-A mg-N/g-A
at 25 at 150
ppmw ppmw
MIL-101 indole 23.31.1 0.0480.007 12.61.4 20.41.3
MIL-101 quinoline 26.00.5 0.1150.014 19.31.0 24.60.7
(a)
AC1 indole 23.2 0.020 7.72 17.39
AC1(a) quinoline 15.1 0.029 6.36 12.26
AC2(a) indole 19.3 0.028 7.88 15.55
AC2(a) quinoline 17.6 0.033 7.90 14.62
(a)
- Data from Ref. 13.

The adsorption equilibrium constants of MIL-101 are significantly higher than for
activated carbons. The maximum adsorption capacities of MIL-101 also exceed the
corresponding parameters of activated carbons. As consequence, the adsorption capacity of
metal-organic framework MIL-101 towards heterocyclic nitrogen compounds is significantly
higher than the capacity of activated carbons especially under the sorption from hydrocarbon
with low nitrogen content (Table 1).
The adsorption of nitrogen compounds from the commercial middle-distillate
hydrocarbon streams using metal-organic framework MIL-101 was also studied [14]. The
detailed analysis of the nitrogen compounds in the light cycle oil before and after adsorption
by means of chromatography with atomic emission detector (JAS) allowed us to conclude
that the sorption treatment leads to a decrease in the concentration of all N-containing
components [15] with the preferential adsorption of the small or less sterically bulky
molecule. The spent metal-organic framework MIL-101 could be regenerated by methanol
extracting of the N-containing compounds in reflux conditions, dried and reused in the next
adsorption cycle, without loss of its adsorption capability.
In summary, metal-organic framework MIL-101 exhibits a remarkable adsorption
capacity for heterocyclic nitrogen compounds exceeding all reported adsorbents. The very
high adsorption capacity and selectivity of MIL-101 for the nitrogen compounds, along with
its good regenerability, indicate that MIL-101 may be the promising adsorbent for deep
denitrogenation of liquid hydrocarbon streams.

EXPERIMENTAL DETAILS
Materials The quinoline and indole were purchased from Aldrich. The metal-organic
framework MIL-101 was prepared as described by Frey et al. [12].
Performance of Metal-Organic Framework MIL-101 295

Adsorption experiment The metal-organic framework MIL-101 (10 mg) was added to the
solution of individual nitrogen compound in isooctane or hydrotreated gas oil (2 mL). After
stirring the mixture for 4 h at room temperature, the adsorbent was collected by filtration. The
filtrate was analyzed by chemiluminescence using ANTEK 9000 NS Instrument.
Regeneration procedure The extraction of the N-containing compounds from MIL-101
(50 mg) was carried out by methanol (35 mL) in reflux conditions for 3 hours. Metal-organic
framework was filtrated, dried at 140 C for 3 hours and reused in the next cycles of sorption
experiments.

REFERENCES
[1] Furimsky, E.; Massoth, F. E. Catal. Rev. Sci. Eng. 2005, 47, 297-489.
[2] Kim, J. H.; Ma, X.; Zhou, A.; Song, C. Catal. Today 2006, 111, 74-83.
[3] Almarri, M.; Ma, X.; Song, C. Ind. Eng. Chem. Res. 2009, 48, 951-960.
[4] Sano, Y.; Choi, K.-H.; Korai, Y.; Mochida, I. Appl. Catal. B 2004, 49, 219-225.
[5] Sano, Y.; Choi, K.-H.; Korai, Y.; Mochida, I. Energ. Fuel. 2004, 18, 644-651.
[6] Kwon, J.-M.; Moon, J.-H.; Bae, Y.-S.; Lee, D.-G.; Sohn, H.-C.; Lee, C.-H. Chem. Sus.
Chem. 2008, 1, 307-309.
[7] Bae, Y.-S.; Kim, M.-B.; Lee, H.-J.; Lee, C.-H.; Ryu, J. W. AICHE J. 2006, 52, 510-
521.
[8] Hernandez-Maldonado, A. J.; Yang, R. T. Angew. Chem. Int. Ed. 2004, 43, 1004-1006.
[9] Shiraishi, Y.; Yamada, A.; Hirai, T. Energ. Fuel. 2004, 18, 1400-1404.
[10] Frey, G. Dalton Trans. 2009, 44004415.
[11] Cychosz, K. A.; Ahmad, R.; Matzger, A. J. Chem. Sci. 2010, 1, 293-302.
[12] Frey, G.; Mellot-Draznieks, C.; Serre, C.; Millange, F.; Dutour, J.; Surbl, S.;
Margiolak, I. Science 2005, 309, 2040-2042.
[13] Almarri, M.; Ma, X.; Song, C. Energ. Fuel. 2009, 23, 3940-3947.
[14] Nuzhdin, A.L.; Kovalenko, K.A.; Dybtsev, D.N.; Bukhtiyarova, G.A. Mendeleev
Commun. 2010, 20, 57-58.
[15] Wiwel, P.; Knudsen, K.; Zeuthen, P.; Whitehurst, D. Ind. Eng. Chem. Res. 2000, 39,
533-540.
INDEX

annealing, 186
A antiferromagnet, 67
aqueous solutions, 116
access, 52, 113, 117, 174, 189
argon, 112
accommodation, 187
Aristotle, 107
acetic acid, 9, 177
aromatic compounds, 292
acetonitrile, 61, 150
aromatic rings, 268, 271
acetylation, 279
aromatics, 292
acid, 62, 78, 108, 115, 137, 138, 140, 142, 143, 147,
Arrhenius equation, 116
148, 152, 173, 176, 178, 189, 190, 252, 256, 278,
arsenates, 2, 23
279, 287, 288, 289
assessment, 202
acidic, iv, 9, 11, 12, 13, 277, 287, 291
assets, ii, 129
activated carbon, ii, iii, iv, 99, 100, 131, 197, 198,
asymmetry, 81
199, 201, 202, 203, 204, 211, 215, 217, 218, 291,
atmosphere, 34, 61, 109, 113, 216, 218, 226, 244
292, 293, 294
atmospheric pressure, 205, 208, 209, 214
activation energy, 116, 120, 241
attachment, 188
active site, ii, iii, 52, 99, 115, 117, 120, 169, 173,
automobiles, 125
174, 179, 181, 183, 185, 190
Avogadro number, 63
adaptability, 6
ADC, 177
adhesion, 145, 153, 160 B
adsorption isotherms, 102, 112, 205, 206, 207, 208,
209, 210, 212, 213, 214, 215, 216, 217, 228, 231, band gap, 180
234 barriers, 92
aggregation, 132, 148, 152, 153, 155, 264 base, 153, 182, 183, 185, 189, 190, 265
alcohols, 274 behaviors, i, 2, 33, 77, 78, 103
algae, 33 Beijing, 99
alkaline earth metals, 238 bending, 229, 236
alkane, 232 benign, 173
alternative energy, 100 benzene, 19, 22, 107, 138, 140, 144, 177, 181, 182,
alumino-phosphates, 2 230, 263
alumino-silicates, 2 binding energy, 230, 239
amine, 74, 126, 140, 190, 226, 250 biofuel, 244
amino, 16, 182, 185, 189, 190, 227, 228 biogas, 130
amino groups, 182, 185, 189, 190 biomedical applications, 227
ammonia, 113, 125, 126, 129 biotechnology, ii, 129
anchoring, 181, 182, 190 blends, 146, 147
aniline, 189, 292 Boltzmann constant, 63
anisotropy, 65, 70, 76, 77, 78, 79, 80, 93, 253 Boltzmann distribution, 230
298 Index

bonding, i, iv, 2, 68, 227, 249, 250, 251, 252, 253, 173, 174, 178, 179, 182, 185, 186, 187, 188, 189,
254, 255, 256, 257, 258, 260, 261, 262, 263, 264, 190, 191, 201, 218, 226, 230, 232, 238, 239, 244
265, 266, 267, 268, 269, 271, 272, 273, 274, 275, chemical bonds, 45
276, 278, 281 chemical industry, 226
bonds, iv, 3, 4, 5, 35, 36, 51, 60, 81, 118, 126, 169, chemical inertness, 201, 218
170, 171, 174, 176, 185, 249, 250, 251, 254, 259, chemical properties, 174
260, 265, 266, 267, 268, 271, 272, 273, 274, 281 chemical reactions, ii, 99, 113
bounds, 143 chemical stability, 130, 133, 244
breathing, 36, 170, 230 chemical vapor deposition, 133, 185
bromine, 255 chemicals, 287
BTC, 108, 134, 135, 136, 137, 144, 145, 146, 147, chemiluminescence, 294
148, 153, 156, 177, 181, 227, 228, 230, 231, 232, chemisorption, 100, 239
240, 242 China, 99, 249, 272
building blocks, ii, 2, 59, 60, 93, 183, 227, 244, 266 chiral catalyst, 171, 174
butadiene, 130 chiral molecules, 171
by-products, 114 chirality, 171, 266, 275
chloride anion, 258
chlorine, 88, 183
C chloroform, 136, 137, 140, 148, 150, 153, 155
chromatography, 294
Ca2+, 237
chromium, 123, 182, 234, 292
cadmium, 12, 21, 278, 283, 284
clarity, 150, 263, 265, 280, 282, 283, 284
calorie, 226
classes, 17, 19, 113, 292
candidates, iv, 113, 133, 216, 225, 237
classification, i, 1, 5, 17, 19, 20, 207
carbon, 100, 130, 131, 133, 141, 199, 200, 202, 203,
clean energy, 122
204, 205, 206, 208, 214, 216, 218, 220, 278, 280,
cleavage, 188
282, 283, 284
clusters, i, 1, 2, 5, 32, 53, 93, 117, 120, 121, 178,
carbon dioxide, 130, 141, 199, 200, 205, 206, 208,
180, 227
214
coal, 225
carbon materials, 200
cobalt, 28, 50, 54, 80, 85, 227, 273
carbon nanotubes, 199, 201
collateral, 286
carbonyl groups, 285
color, 11, 43, 235
carboxylic acid, 189
combined effect, 64
carboxylic acids, 189
combustion, 225, 244
catalysis, i, ii, iii, iv, 2, 60, 99, 100, 113, 117, 121,
commercial, ii, 129, 130, 153, 156, 159, 204, 294
122, 123, 126, 169, 171, 172, 174, 179, 183, 185,
comparative analysis, 203
188, 189, 225, 227, 291
compatibility, 145, 160
catalyst, ii, 99, 114, 115, 116, 117, 118, 121, 122,
compensation, 82
172, 173, 174, 175, 179, 189, 287
competition, 258, 275
catalytic activity, 32, 33, 116, 117, 118, 121, 170,
complement, 32, 172, 240
173, 175, 182, 183, 186, 188
complexity, i, iv, 1, 18, 19, 249
catalytic effect, 122
composites, 132
catalytic properties, 51, 122, 172, 173, 175, 182, 184
composition, 131, 132, 147, 242, 243
catalytic system, 173
compression, 38, 40, 212, 213, 215
cation, 11, 17, 68, 72, 73, 74, 237, 259, 260, 262,
condensation, 8, 186
287
conduction, iv, 180, 277
C-C, 186
conductivity, 230, 287
cellulose, 130
conductor, 287
challenges, 100, 244
conductors, 250, 262
charge density, 236
configuration, 60, 65, 79, 180
chelates, 29
confinement, 117, 171, 186, 188
chemical, ii, 2, 8, 9, 11, 14, 17, 45, 99, 100, 101,
conformity, 214
107, 113, 125, 126, 129, 130, 133, 158, 160, 171,
Index 299

connectivity, i, 2, 4, 5, 6, 7, 18, 19, 24, 28, 29, 31, density functional theory, 229
32, 35, 47, 48, 49, 54, 251, 261, 262 density values, 12
construction, 5, 16, 124, 133, 174, 249, 256 Department of Energy, 100, 245
consumption, 226 deposition, 133, 134, 135, 274
containers, 10 depth, 244
contaminant, 130 derivatives, 73, 175, 186, 253, 255, 265, 274, 279,
conversion rate, 52 281, 292
cooling, 64, 69, 71, 74, 226 desorption, 110, 112, 118, 206, 207, 208, 218, 228,
cooperation, 272 236, 239
copper, 24, 30, 54, 75, 78, 88, 124, 133, 134, 135, destruction, 218
140, 148, 176, 177, 179, 181, 228, 256, 266, 267, detectable, 115
268, 270, 271, 275 deviation, 260
correlation, 131, 201, 240, 293 DFT, 68, 229, 234, 238
correlations, 49, 201, 202, 229 dichotomy, 172
cost, ii, 113, 129, 130, 131, 226, 229, 244 differential scanning, 118
coupling constants, 76 differential scanning calorimetry, 118
covalency, 93 diffraction, 34, 39, 40, 44, 62, 67, 68, 102, 108, 110,
covalent bond, 272 236, 286
covering, 169, 172, 208 diffusion, 8, 134, 135, 155, 156, 174, 176, 188, 240
crown, 62 diffusion process, 8
crystal architecture, i, 1, 3, 5, 7, 9, 17, 19, 23, 24, 31, diffusivities, 141, 143, 155, 240
33, 43, 54 diffusivity, 140, 160, 230
crystal growth, 8, 176, 285 dimensionality, i, 1, 17, 19, 23, 27, 28, 29, 30, 31,
crystalline, iv, 9, 11, 14, 42, 132, 169, 170, 171, 174, 171, 253, 257, 258, 271
175, 176, 177, 178, 179, 180, 182, 183, 184, 189, dimethylformamide, 78, 101, 108, 111, 138, 140,
205, 249, 265, 272, 274, 292 287, 288
crystalline solids, 169 dimethylsulfoxide, 62
crystallinity, i, 2, 12, 34, 35, 37, 39, 42, 108, 118, directionality, 170
170, 277 disorder, 236
crystallization, 7, 8, 9, 10, 11, 12, 13, 14, 15, 16, 24, dispersion, 117, 140, 148, 150, 155, 156, 160, 229
29, 32, 33, 34, 36, 37, 38, 39, 40, 41, 42, 43, 102, displacement, 39, 179, 180, 181, 205, 206
103, 176, 270 disposition, 8, 24, 42, 171
crystals, 11, 12, 13, 14, 115, 117, 123, 140, 142, 148, distillation, 130, 226
152, 153, 155, 156, 158, 159, 171, 176, 200, 251, distilled water, 40
255, 262, 263, 274, 277, 278, 281, 288 distortions, 179
cyanide, 46, 83, 88, 104, 105, 124, 260 distribution, 5, 132, 187, 188, 201, 237, 243, 259
cycles, 3, 5, 6, 7, 20, 21, 22, 23, 25, 27, 28, 30, 31, distribution function, 237, 243
32, 42, 100, 115, 117, 295 diversity, 3, 6, 7, 54, 113, 121, 122
cyclohexanone, 173 DMF, 92, 108, 109, 111, 140, 143, 150, 152, 158,
179, 265, 267, 278, 279, 281, 283, 284, 287, 288,
289
D DOI, 58
donors, 133, 180, 189, 250, 252, 262, 268
damages, 214
doping, iv, 225, 238, 243
database, 4, 5
drawing, 84, 85, 250
decay, 180
drug delivery, 2
decomposition, ii, 2, 32, 53, 54, 109, 113, 114, 115,
drugs, 171
118, 119, 120, 121, 122
DSC, 33, 37, 118, 119, 120
defects, 141
dyes, 53
deficiency, 229
deformation, 140
degenerate, 65, 85 E
degradation, 207, 210
dehydration, 42, 102 effluent, 226, 240
300 Index

effluents, 53 fibers, 147, 148, 203


electric current, 176 filler particles, 150
electric field, 3, 170, 237, 242 fillers, 131, 145, 148, 153, 156, 160
electrolyte, 262 filtration, 158, 174, 294
electron, 3, 63, 68, 79, 82, 88, 110, 180, 189, 229, financial, 219, 272
250, 252, 257, 258 financial support, 219, 272
electron paramagnetic resonance, 88 first generation, 2, 170
electronic structure, 80 fixed bed reactors, 173
electrons, 60, 68, 80, 82, 180, 260 flexibility, 3, 6, 10, 16, 24, 33, 40, 54, 133, 153, 172,
electrostatic host-guest interactions, 2 179, 230, 244
elongation, 146 flue gas, 226, 230, 235, 236
e-mail, 1 fluid, 222, 230
emission, 114, 294 food, ii, 129, 130
endothermic, 119 force, 61, 229, 244
energy, ii, 2, 65, 80, 86, 92, 100, 114, 116, 120, 129, force constants, 61
130, 172, 180, 219, 225, 226, 234, 235, 250, 274 formamide, 179, 278
energy consumption, ii, 129, 172 formation, 2, 54, 60, 61, 88, 113, 114, 119, 126, 131,
energy density, 114 140, 178, 188, 232, 250, 259, 275, 279
energy efficiency, 130, 226 formula, 13, 14, 15, 16, 44, 51, 82, 91, 178, 186
engineering, ii, 16, 17, 99, 113, 121, 123, 175, 228, free volume, 153, 179, 236, 242
250, 256, 273, 274, 276 freedom, 43
environment, i, 1, 4, 5, 6, 9, 17, 28, 33, 42, 63, 65, functionalization, iv, 133, 188, 189, 190, 225, 243,
79, 80, 83, 133, 171, 177, 189, 190, 266 277
environmental protection, iii, 225, 226 funding, 54
enzymes, 33 fungi, 33
EPR, 68, 79, 186
equilibrium, 106, 239, 240, 293, 294
equipment, 205 G
erosion, 130
gas diffusion, 131
ester, 189
gas sorption, 101, 123, 125
ethanol, 130, 160, 200, 288
gel, 133
ethyl acetate, 150
geometry, i, 1, 4, 5, 8, 16, 17, 26, 64, 77, 82, 83, 178,
ethylene, 9, 17, 19, 21, 130
230, 266, 281, 283
evacuation, 101, 170
germanates, 2, 23
evaporation, 134, 135, 136, 137, 176
Germany, 272
evidence, 78, 120, 140
glassy polymers, 130
evolution, iii, 15, 37, 38, 39, 40, 44, 47, 48, 49, 50,
graphene sheet, 202
120, 172, 197, 199, 211
graphite, 134, 135
exchange rate, 61
growth, iv, 133, 249, 285
excitation, 180
guidelines, 244
exclusion, 153, 158
experimental condition, 214, 218
exposure, 207, 208, 210, 218, 279 H
extraction, 2, 8, 295
extrusion, 147 halogen, i, iv, 61, 68, 78, 249, 250, 251, 252, 253,
254, 255, 256, 258, 259, 260, 261, 262, 263, 264,
265, 266, 267, 268, 269, 271, 272, 273, 274, 275,
F
276
halogens, 254
fabrication, 131
Hamiltonian, 65, 66, 71, 72, 75, 76, 80, 82, 83, 84,
families, 32, 199, 200, 201, 203, 215, 217
85, 86, 88, 89, 90, 91
ferromagnetism, 70
heat capacity, 67
ferromagnets, 79
heating rate, 118, 120
fiber, 11, 147, 148
Index 301

helium, 205 infancy, 135, 172


heterogeneous catalysis, i, iii, 133, 169, 170, 171, inorganic fillers, 131
172, 174, 175, 191 inorganic rigid scaffolds, 2
hexane, 136, 137, 146, 200 INS, 110
histogram, 5 insertion, 229
homogeneous catalyst, 173, 174, 186 integrity, 2, 180, 189
homolytic, 188 interface, 131, 133, 144, 155, 160
host, 2, 3, 17, 39, 171, 185, 186, 187, 188 interfacial layer, 132
humidity, 218 intermolecular interactions, iv, 67, 72, 86, 249, 251,
Hunter, 275 254, 272
hybrid, i, 1, 3, 5, 6, 7, 8, 9, 11, 12, 17, 18, 19, 20, 23, interphase, 132
24, 26, 28, 29, 31, 32, 33, 40, 42, 43, 47, 49, 52, inventors, 222
53, 54, 132, 169, 174, 229 inversion, 70, 136, 137
hydrides, 100, 113, 122, 198 iodine, 252, 253, 254, 255, 256, 260, 262, 265, 266,
hydrocarbons, iv, 130, 200, 291, 292 267, 268, 269, 270, 271, 274
hydrogen abstraction, 188 ions, ii, 35, 51, 52, 59, 60, 63, 67, 71, 74, 77, 80, 82,
hydrogen atoms, 104, 150, 233 83, 91, 99, 101, 105, 106, 108, 112, 113, 115,
hydrogen bonds, 3, 29, 51, 169, 170, 249, 251, 254, 117, 121, 122, 148, 177, 178, 182, 183, 185, 189,
256, 257, 258, 259, 260, 262, 263, 268, 274 236, 237, 242, 243, 256, 258, 263, 275, 278, 280,
hydrogen gas, 130 281, 283, 287
hydrogen peroxide, 173 IR spectra, 14, 36, 280
hydrogenation, 186, 188 IR spectroscopy, 9
hydrolysis, 8, 113, 114, 115, 116, 117, 118, 186 iron, 80, 83, 92
hydrophilicity, 171, 186 irradiation, 53
hydrophobicity, 171, 186 isophthalic acid, 271
hydrothermal synthesis, 8, 9, 10, 16, 19 isotherms, 103, 109, 110, 111, 112, 205, 206, 207,
hydrothermal system, 9 208, 210, 213, 215, 218, 222, 293, 294
hydroxyl, 183, 277, 278, 280, 281, 284, 285, 287 Israel, 263
hydroxyl groups, 183, 277, 278, 280, 281, 284, 285,
287
hypothesis, 103, 120 K
hysteresis, 70, 71, 79, 92, 109, 124
K+, 237
hysteresis loop, 70, 71, 79, 124
ketones, 188
kinetics, 100, 113, 114, 115, 118, 120, 121, 199
I KOH, 204

ideal, iv, 132, 138, 140, 141, 142, 149, 152, 159,
225, 228, 231, 239 L
identification, 9, 15
lanthanide, 189
identity, 68, 189
lattices, 73, 77
image, 115, 151
layer-by-layer growth, 133
images, 139, 140, 144, 146, 147, 148, 151, 152, 154,
leaching, 174
156, 158, 230
lead, 78, 80, 119, 120, 171, 178, 229, 242, 244
immobilization, 174
Lewis acids, 173
impregnation, 174, 185
lifetime, 100, 180
improvements, 114, 130, 176, 199, 241
light, 3, 9, 53, 54, 60, 143, 170, 198, 294
impurities, 49, 102, 177, 226
light cycle, 294
India, 246
liquid crystals, 250, 273
indium, 243
liquid phase, 173, 174, 176, 293
induction, 52, 114, 119, 120, 266
liquids, 126, 176
induction period, 52, 114, 119, 120
low temperatures, 8, 48, 66, 74, 78, 79, 80, 88, 91,
industrial revolution, 226
92, 100, 110
industries, ii, 129, 130
302 Index

LTA, 232 metal-organic compounds, 256


luminescence, 133, 135, 250 metals, i, iv, 60, 67, 88, 133, 175, 178, 225, 234,
Luo, 55, 58 236, 238, 239, 243, 256, 274
metavanadate chains, 6, 7, 26, 30, 46, 51
methanol, 101, 115, 116, 117, 150, 158, 200, 294,
M 295
methylene blue, 53, 54
magnesium, 100
MFI, 232, 240
magnet, 76, 87, 88, 91, 92, 93
Mg2+, 8, 237
magnetic field, 65, 66, 70, 74, 79, 87
microcrystalline, 66, 176
magnetic moment, 64, 67, 79
microporous materials, ii, 99, 170
magnetic properties, i, ii, 2, 32, 47, 54, 59, 60, 63,
microscope, 9
68, 69, 71, 72, 73, 75, 76, 77, 79, 80, 82, 83, 86,
milligrams, 293
93
Ministry of Education, 244
magnetic structure, 79
mixing, 138, 147, 152, 176, 209
magnetism, 133
MOCVD, 133, 134
magnetization, 70, 71, 74, 76, 78, 83, 88, 91, 92
modelling, 135
magnets, 60
models, 131, 136, 141
magnitude, 77, 82, 91, 228, 242
modifications, 188, 189
majority, 60, 183
modules, 2
manganese, 40, 54, 80, 90
modulus, 146
manipulation, 210
moisture, 242, 244
marches, 132
molar ratios, 14, 115, 116
MAS, 126
mole, 9, 14
mass, 202, 222
molecular dynamics, 229
materials science, 228
molecular weight, 113, 158, 186
matrix, i, ii, 129, 131, 133, 136, 138, 147, 148, 151,
momentum, 63
159, 160, 186
monolayer, 103, 106
matter, 132
Moon, 194, 295
measurement, 39, 40, 43, 106, 159, 205
morphology, iii, 8, 132, 133, 146, 197, 199, 201
measurements, 14, 32, 33, 34, 36, 44, 49, 66, 67, 68,
motif, 124, 253, 254, 271
70, 77, 88, 92, 105, 112, 119, 205, 212, 236, 278,
multidimensional, 59
289
mechanical properties, 146, 244, 273
media, 77, 176 N
medical, 133, 198
melting, 119 Na+, 8, 236, 237, 242, 243
membrane separation processes, 129 nanocrystals, 139
membranes, i, ii, iv, 129, 130, 131, 133, 134, 135, nanomaterials, 250
142, 145, 147, 148, 151, 152, 153, 155, 159, 160, nanometer, 153, 174
225, 226, 228, 229, 239, 240, 241, 243 nanometer scale, 174
Mendeleev, 295 nanoparticles, 116, 117, 131, 149, 153, 155, 185, 186
mesoporous materials, 131, 170, 173 nanostructures, 231
metal complexes, 173, 174, 257 natural gas, 130, 141, 198, 215, 225
metal ion, ii, 59, 60, 61, 68, 76, 80, 82, 84, 85, 86, NCS, 61, 77, 78
99, 100, 105, 106, 112, 113, 118, 121, 122, 169, neutral, 11, 12, 13, 52, 61, 68, 73, 78, 82, 84, 88,
173, 174, 178, 179, 180, 181, 182, 183, 184, 185, 108, 175, 236, 237, 242, 262
189, 190, 236 NH2, 227, 233, 234
metal ions, ii, 59, 60, 68, 76, 80, 82, 84, 86, 99, 100, nickel, 11, 24, 28, 35, 42, 43, 46, 48, 54, 76, 80, 115,
105, 106, 112, 113, 118, 121, 122, 169, 174, 178, 117
179, 181, 182, 183, 184, 185, 189, 190, 236 nicotinamide, 251, 254
metal nanoparticles, 185 nitrates, 8
metal oxides, 242 nitrogen, iv, 16, 17, 24, 40, 51, 83, 101, 103, 105,
metal salts, 173 106, 107, 110, 112, 126, 129, 130, 183, 227, 260,
Index 303

262, 266, 268, 271, 281, 282, 283, 284, 291, 292, permission, 5, 10, 11, 12, 13, 33, 34, 35, 36, 37, 41,
293, 294 42, 44, 45, 48, 50, 52, 226, 227, 232, 233, 235,
nitrogen compounds, iv, 291, 292, 293, 294 237, 238, 239, 241, 243
nitroxide, 72 peroxide, 174
NMR, 8, 126, 137, 155, 156, 285, 288, 289 Petroleum, 164
noble metals, 126 pH, i, 1, 8, 9, 11, 12, 13, 14
nodes, 22, 32, 174, 180, 183, 189, 190 pharmaceutical, ii, 129, 130
nonlinear optics, 250 phase diagram, 9
nucleation, 176 phase inversion, 146
nucleophiles, 256, 275 phenol, 173
phophites, 2
phosphates, i, 1, 2, 23
O photocatalysts, 180
photoelectron spectroscopy, 115
octane, 111, 266, 268
photolysis, 187, 188
OH, 8, 61, 156, 178, 179, 180, 233, 234, 274
photons, 180
oil, iv, 146, 225, 291, 292, 294
physical characteristics, 198
oligomers, 5
physical properties, 2, 214
opportunities, 133, 244
physicochemical properties, 2, 3, 32
optimization, 2, 9, 198, 234
physics, 228
orbit, 50, 60, 63, 64, 65, 67, 76, 80, 85, 93
pku, 99
organ, 173
plants, 129, 130
organic ligand, 2, 8, 11, 14, 16, 17, 18, 19, 20, 24,
plasticity, 250
27, 28, 29, 31, 32, 33, 36, 39, 44, 46, 47, 49, 52,
platinum, 116
54, 107, 110, 113, 117, 118, 121, 122, 132, 133,
PMDA, 137, 147
134, 169, 170, 171, 173, 174, 178, 179, 182
POFs, 133
organic polymers, i, 1, 256, 257
poison, iv, 291
organic solvents, 114, 150, 234
polar, 228, 234, 242, 253, 266
orthogonality, 82, 83
polarity, 186
overlap, 81, 84, 232
polarizability, 258
ox, ii, 59, 68, 77, 78, 79, 81, 82, 83, 84, 86, 87, 88,
polarization, 257
89, 92
pollutants, ii, 2, 54
oxidation, i, 1, 3, 16, 52, 54, 60, 61, 173, 186, 189,
pollution, 130
262
poly(vinyl chloride), 130
oxide clusters, 186
polyamides, 130
oxide nanoparticles, 186
polychloroprene, 130
oxygen, 4, 8, 19, 24, 35, 46, 51, 82, 83, 88, 89, 130,
polydimethylsiloxane, 144, 145, 153, 156
236, 252, 267, 268, 278, 280, 281, 282, 283, 284
polydispersity, 158, 186
polyimide, 130, 136, 137, 138, 142, 145, 146, 147,
P 148, 159
polyimides, 130, 140, 144, 153, 158, 159
parallel, 20, 21, 22, 23, 24, 25, 26, 27, 28, 51, 65, 66, polymer, i, ii, 1, 2, 129, 130, 131, 132, 133, 135,
89, 259, 263, 266, 267, 268, 271, 272, 273, 277 138, 140, 145, 146, 148, 150, 152, 153, 155, 158,
patents, 138 160, 187, 256, 287, 292
pathways, 9, 126, 278 polymer chain, 131, 133, 153, 187
pedal, 10, 47 polymer chains, 131, 133, 153, 187
permeability, 130, 131, 132, 135, 138, 140, 141, 142, polymer matrix, 131, 132, 133, 135, 138, 140, 141,
143, 144, 145, 147, 148, 149, 152, 153, 154, 155, 146, 155, 160
158, 159, 228 polymer properties, 132
permeation, 131, 139, 145, 148, 152, 155, 156, 159, polymer swelling, 153
228, 239, 240, 241, 243 polymerization, 2, 8, 9, 186
304 Index

polymers, i, iv, 2, 18, 32, 102, 122, 123, 124, 135, reaction time, 8, 10, 12, 279
138, 143, 144, 145, 146, 147, 153, 158, 160, 170, reactions, ii, 2, 9, 10, 12, 14, 15, 52, 54, 61, 172,
177, 186, 249, 258, 266, 271, 272 173, 174, 180, 184, 186, 187, 189, 255, 256, 278,
polymorphism, 178, 273 284, 287
polynuclear complexes, i, ii, 59, 60 reactive groups, 182
polypropylene, 130 reactivity, 126, 171, 173, 189
polystyrene, 186 reading, 191
pore openings, 171, 186 reagents, 2, 8, 11, 173, 175
porosity, ii, iii, 2, 99, 100, 101, 105, 113, 123, 170, reception, 218
171, 197, 198, 199, 200, 201, 202, 203, 205, 213, receptors, 250
215, 217, 218, 219, 236, 240 recognition, 60, 250, 263, 273
porous materials, ii, iii, 2, 99, 100, 106, 107, 113, recombination, 188
123, 131, 132, 133, 169, 197, 199, 200, 216, 234, recovery, 129, 130, 186
237, 292 recycling, 174, 228
porphyrins, 263 regeneration, 114, 226
potassium, 106 regression, 293
precipitation, 185, 279 relaxation, 88, 92
prediction models, 141, 160 relaxation times, 92
preparation, ii, 52, 59, 60, 77, 79, 100, 115, 117, 130, reliability, 203, 229
135, 145, 147, 155, 156, 172, 175, 176, 177, 178, reproduction, 89
184, 185, 186, 188, 189, 190, 213, 247, 261, 285 repulsion, 60, 253
prevention, 280 requirements, 77, 180
principles, 236 researchers, 33, 107, 132, 174, 250
prismatic single crystal, 11 resistance, 130
project, 94, 219 resolution, 243
proline, 182, 189 resources, iv, 225, 228
propagation, 17, 43 response, i, 2, 3, 17, 32, 33, 34, 39, 40, 41, 43, 46
propane, 19, 22, 130 rhenium, 60, 61, 63, 68, 74, 76, 77, 78, 86, 90
propylene, 130 rings, i, 1, 22, 36, 52, 53, 71, 123, 233, 258, 259,
protection, 277, 278, 287 265, 266, 267, 268, 271, 276
protons, 173, 289 room temperature, 8, 10, 39, 43, 44, 63, 67, 69, 71,
prototype, 227, 230 74, 79, 91, 107, 110, 113, 114, 115, 116, 124,
purification, 130, 172, 228, 240, 241, 244, 287, 288 126, 155, 199, 205, 206, 211, 214, 230, 288, 294
purity, 103, 226 routes, 61
PVAc, 140, 141 Royal Society, 163, 238, 241, 243, 272
pyrimidine, 78, 175, 227, 236 rubber, 130
pyromellitic dianhydride, 147 rubbers, 130
rules, 244

Q
S
quantum chemical calculations, 173
safety, 215
salts, 8, 62, 63, 66, 67, 68, 72, 73, 77, 116, 173, 262,
R 275
Sartorius, 205, 206
radial distribution, 236, 242
saturation, 70, 74, 91, 110, 111, 115, 122, 124, 176,
radicals, 188
214
radius, 238
scaling, 131, 226
reactant, 117, 277
scattering, 110, 235, 236
reactants, 11, 12, 115, 176, 287
science, 9, 220
reaction mechanism, 173
scope, 169, 171
reaction medium, 173, 174
second generation, 2, 34
reaction rate, 115, 120
seed, 134
Index 305

seeding, 134, 228 stability, 8, 11, 12, 44, 61, 101, 102, 105, 113, 117,
selectivity, iv, 52, 130, 131, 135, 140, 141, 142, 143, 130, 131, 172, 174, 218, 271, 278
144, 145, 147, 148, 149, 150, 153, 154, 158, 160, stabilization, 257, 284
173, 174, 188, 227, 228, 232, 233, 234, 236, 237, stars, 86, 143
238, 239, 242, 243, 291, 292, 294 state, iii, 3, 60, 61, 63, 65, 68, 79, 83, 87, 88, 114,
self-assembly, 132, 176, 250, 256, 258, 265 126, 169, 180, 191, 198, 205, 230, 244, 261, 262
SEM micrographs, 155 states, i, 1, 3, 4, 16, 60, 65, 171, 173, 174, 180
semiconductor, ii, 129, 130, 133, 180 stoichiometry, 8, 9, 11, 12, 13, 14, 16
sensing, 2, 228 stress, 86
sensitivity, 206, 208, 210 stretching, 36, 61, 229, 284
sensors, 134, 135 strong interaction, 110, 259
shape, iii, 3, 12, 67, 69, 74, 132, 133, 142, 171, 173, structural changes, i, 2, 39, 40
197, 199, 201, 208, 286 structural characteristics, i, 1, 17
showing, 155, 210, 280, 292 structural knowledge, 86
signals, 36, 88 structural transformations, i, 2, 43
silica, iv, 114, 131, 135, 173, 291, 292 structural transitions, 32
silicon, 135 styrene, 156, 186
silver, 52, 53, 54, 263, 275 styrene polymerization, 186
simulation, iii, 108, 135, 141, 225, 229, 230, 231, subgroups, 7
232, 233, 235, 236, 239, 240, 242, 243, 244 substitutes, 260
simulations, iv, 225, 228, 229, 239, 243, 244 substitution, ii, 59, 60, 61, 77, 189, 227, 256
Singapore, 225, 244 substitution reaction, 61
single chain, 21, 93, 270, 271 substitutions, 233
single crystals, 11, 12, 13, 14, 88, 176, 188 substrate, 52, 120, 121, 178, 179, 180, 186, 228
SiO2, 186 substrates, ii, 52, 99, 171, 173, 174, 179, 181, 189
skeleton, 2, 3, 23 sulfate, 21
sodium, 179 sulfur, 252, 274
solid state, 9, 77, 126, 261 sulfuric acid, 288
solubility, 141, 144, 149, 152, 160, 176 Sun, 56, 124, 125, 167, 192, 193, 195, 222
solution, 8, 9, 11, 13, 15, 16, 53, 100, 102, 105, 108, superconductivity, 262
109, 113, 114, 115, 116, 117, 140, 146, 147, 148, superconductor, 262, 275
153, 155, 173, 174, 176, 188, 189, 231, 262, 288, suppression, 187
289, 294 surface area, ii, iii, 99, 101, 103, 105, 106, 107, 109,
solvent molecules, 76, 170, 180, 268 110, 111, 112, 117, 123, 133, 138, 139, 144, 147,
solvents, 61, 175, 176, 177, 292 148, 153, 170, 172, 177, 179, 197, 198, 199, 201,
solvothermal reactions, 9 202, 203, 204, 205, 207, 210, 213, 214, 219, 227,
solvothermal synthesis, 133 292
sorption, iv, 101, 102, 103, 104, 109, 110, 113, 122, surface modification, 133
123, 125, 144, 156, 291, 293, 294, 295 surface properties, 133
sorption experiments, 295 susceptibility, 48, 49, 50, 51, 63, 66, 67, 68, 70, 71,
sorption isotherms, 101, 109 73, 75, 76, 77, 78, 86, 88, 89, 90, 91, 92, 93
sorption process, 102 sustainable development, iii, 225, 226
Spain, 1, 169, 197 symmetry, 64, 65, 80, 81, 82, 83, 86, 93, 259, 271
spatial array, 174, 175 synergistic effect, 52
species, ii, iv, 3, 8, 32, 59, 61, 93, 174, 180, 185, synthesis, iv, 2, 3, 8, 60, 61, 115, 116, 117, 118, 120,
186, 188, 252, 259, 291, 292 123, 133, 135, 138, 139, 141, 144, 145, 147, 148,
specific surface, ii, 99, 100, 101, 105, 106, 107, 113, 152, 153, 158, 171, 175, 176, 177, 178, 179, 184,
170 186, 188, 189, 190, 191, 202, 203, 206, 210, 218,
spectroscopy, 110, 112, 186, 288 227, 249, 272, 277
spin, ii, 50, 51, 59, 60, 63, 64, 65, 67, 68, 70, 72, 73, synthetic methods, 277
75, 76, 77, 79, 80, 82, 83, 85, 88, 90, 91, 93, 101,
124, 133, 148
sponge, 147
306 Index

transmission electron microscopy, 115


T transport, iii, 130, 131, 132, 140, 141, 152, 197, 198,
215, 216, 230, 240, 277, 287
tantalum, 174
transport processes, 132
target, 198, 199
transportation, iii, 100, 197, 198, 199
techniques, iv, 130, 133, 155, 186, 225, 228
treatment, 101, 189, 230, 242, 294
technologies, iii, 197, 226
tricarboxylic acid, 182
technology, ii, 129, 130, 220
trifluoroacetate, 263
TEM, 115
triphenylphosphine, 62
temperature dependence, 48, 64, 87
tunneling, 88
template molecules, 2
twist, 267
tensile strength, 146
testing, iv, 225, 228, 244
tetrahydrofuran, 150, 234 U
tetrahydrofurane, 61
textural character, 205 ultrasound, 158
texture, 213 uniform, 100, 113, 117, 174
TGA, 101, 102, 105, 109, 146 universal gas constant, 109
thermal activation, 181, 188 urea, 273
thermal analysis, 114 Uruguay, 59, 94
thermal decomposition, 113, 114, 117, 119, 120, USA, 163, 273, 275
121, 125, 126, 127 UV, 53, 180
thermal expansion, 32, 39, 40, 44, 45, 46, 47, 60, 124 UV light, 53, 180
thermal properties, 32, 54, 244
thermal stability, 34, 44, 101, 102, 130, 138, 139,
153, 218, 244 V
thermal treatment, 177, 182, 292
thermodiffractometric studies, 32 vacancies, 8, 102, 181, 188, 189
thermodynamic properties, 230 vacuum, 44, 45, 101, 107, 181, 182, 205, 206, 240,
thermodynamics, 119, 120 241
thermogravimetry, 44 valence, 180
thermolysis, 114 Valencia, 169, 192
thin films, 133, 135 vanadium, i, 1, 3, 4, 5, 6, 7, 8, 9, 11, 12, 13, 15, 16,
tin, 173 17, 19, 20, 21, 22, 23, 24, 25, 26, 27, 28, 29, 31,
titania, 135 32, 33, 36, 47, 49, 50, 54, 156
titanium, 173, 184, 186 vanadium oxides, 3, 5, 6, 8, 17, 20, 21, 22, 23, 32,
titanium isopropoxide, 184, 186 47, 54
toluene, 136, 140, 146, 150 vapor, 274
topology, 12, 17, 18, 32, 177, 178 variables, 9, 201
total energy, 230 variations, 206
toxicity, 113 vector, 43
TPA, 140, 141 vehicles, 198, 215, 219
trade, 130, 131 versatility, 175
trade-off, 130, 131 vibration, 36
trajectory, 230 viscosity, 8, 176, 186
transformation, 34, 35, 36, 39, 40, 42, 43, 125 vision, iii, 169, 191
transformations, 172, 173
transition metal, i, ii, 1, 2, 3, 7, 17, 19, 23, 59, 60, 67, W
68, 93, 111, 114, 132, 173, 174, 175, 256
transition metal ions, ii, 59, 67, 68, 111 Washington, 245
transition metal phosphates, i, 1, 2, 23 waste, 53, 173, 176
transition temperature, 74, 262 water, i, iv, 2, 8, 9, 11, 12, 24, 32, 33, 34, 35, 36, 37,
translation, 6 38, 39, 40, 41, 42, 43, 60, 71, 102, 104, 105, 106,
transmission, 115
Index 307

115, 130, 133, 148, 158, 181, 182, 225, 226, 229,
234, 244, 257, 279, 292
Y
weak interaction, 27, 29, 100, 169, 170, 254
yield, 179, 188
weight loss, 109
windows, 111, 144, 158, 179, 240, 243
workers, 114, 178, 183, 189, 190, 250, 252, 254, Z
258, 260, 277
worldwide, 100, 130, 198, 225 zeolites, i, ii, iii, 2, 3, 46, 99, 100, 131, 135, 140,
172, 173, 174, 176, 186, 188, 197, 198, 199, 200,
226, 227, 232, 235, 240
X zeotype materials, 2
zinc, 122, 125, 138, 178, 179, 182, 185, 186, 278,
XPS, 115, 117, 118
279, 280, 281, 282, 283, 287
X-ray diffraction, 9, 12, 14, 34, 46, 102, 176, 278,
zinc oxide, 279
281
zirconium, 174
XRD, 102, 105, 108, 118, 218, 279, 280
ZnO, 186

Das könnte Ihnen auch gefallen