1 Rashmi

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 54

Accepted Manuscript

CO2 sorption behavior of imidazole, benzimidazole and benzoic acid based co-
ordination polymers

Rashmi A. Agarwal, Neeraj K. Gupta

PII: S0010-8545(16)30369-1
DOI: http://dx.doi.org/10.1016/j.ccr.2016.11.002
Reference: CCR 112335

To appear in: Coordination Chemistry Reviews

Received Date: 1 September 2016


Revised Date: 2 November 2016
Accepted Date: 3 November 2016

Please cite this article as: R.A. Agarwal, N.K. Gupta, CO2 sorption behavior of imidazole, benzimidazole and benzoic
acid based coordination polymers, Coordination Chemistry Reviews (2016), doi: http://dx.doi.org/10.1016/j.ccr.
2016.11.002

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers
we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting proof before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
Review
CO2 sorption behavior of imidazole, benzimidazole and
benzoic acid based coordination polymers
Rashmi A. Agarwal*, Neeraj K. Gupta
Department of Chemistry, Indian Institute of Technology Kanpur,
208016, India *Corresponding Author’s email:
rashmi.a.agarwal@gmail.com

Contents
1. Introduction
2. Porous coordination polymers/metal organic frameworks
2.1 Synthesis
2.2 Characterization
2.3 Classification of imidazole, benzimidazole and benzoic acid based
coordination polymers
2.4 Benzene/imidazole and CO2 interactions in PCPs
2.5 Selection of best category PCPs
2.6 Adaptation to temperature and pressure
3. Post synthesis modifications of MOFs for CO2 adsorption
4. Conclusions; outlook and development needs

1
ABSTRACT:
Non corrosive gas, such as CO2, is an environment greenhouse pollutant, but at the same time it

is a reusable material with a highly attractive application profile and potential as a building block
to create added value for the chemical industry. Current technologies for carbon capture are
cumbersome, energy intensive and expensive. However, amongst new developments, porous
coordination polymers (PCPs) are being viewed as materials with great potential for CO2
capture. The ability to flexibly design and tune properties of such materials results in different
levels of affinity for CO2 sorption/desorption. Imidazole/benzimidazole and benzoic acid based
coordination polymers are of interest from the view point that various architectures such as rigid,
flexible, interpenetrating, etc., can be designed to influence cavity pore size and different types
of supramolecular interactions existing within the pores. CO2 sorption, in these structures, not

only depends on surface area and void volume but is also influenced by the electronic
environment, i.e., functionalization of the pores with either coordinatively unsaturated metal ions
or Lewis base groups. In this review, other important influencing factors including strong non-
covalent interactions (π-π bonding, anion-π bonding, CH-π bonding, and H-bonding) present on
the pore walls, facilitating selective adsorption of CO2 molecules in substantial quantities (even
if Langmuir surface area is very low) are discussed. These interaction potentials in the pores
decide quantitatively the sorption of CO2 as function of temperature and pressure.

1. Introduction
Carbon dioxide (CO2) is a gas vital to life on earth, and present in the earth's atmosphere as a

trace gas at a concentration of about 0.04 percent (400 ppm) by volume as well as in natural
sources. CO2 is an important greenhouse gas. Burning of carbon-based fuels has rapidly

increased its concentration in the atmosphere and represents over half of the global greenhouse
gas emissions. Sharply rising levels of atmospheric CO2 resulting from combustion of fossil
fuels is one of the greatest environmental concerns facing civilization today. It is expected that
the energy consumption will increase by 57% by 2030 according to the Energy Information
Administration (EIA) [1]. Approximately 60% of global warming effects are attributed to carbon
dioxide emission [2]. With rapid increase of global population and industrialization,
consumption of energy is growing exponentially. Our primary energy source has been the

2
inherent energy density abundance of fossil fuels and the economic dependence of modern
society on the acquisition and trade of these resources. It is also a major cause of ocean
acidification, attributed to the escalating level of atmospheric CO2, because it dissolves in water

to form carbonic acid.


There are four potential sources of carbon dioxide emission; fossil fueled power plants, industrial
processes, production of hydrogen from carbon rich feed stock, and transportation [3]. Fossil
fueled power plants rank foremost as CO2 emission source. It is predicted that at the global

scale, electricity generation from coal (since coal is the world’s cheapest and most widely used
electricity source) and gas is expected to increase to 21.9 trillion kWh in 2035 [4,5]. Mitigation
of CO2 from mixed component gas streams for example flue gas streams from power generation

has gained tremendous amount of attention due to growing concern of the environmental and
climatic impact of greenhouse gas emissions. Any serious effort to reduce anthropogenic CO2

emissions must keep in mind the geopolitical and economic reality that fossil fuels will most
likely continue to make a dominant contribution to the world's energy supply for decades to
come. There is, therefore, an intensive global drive to minimize or eliminate these emissions,
particularly through carbon capture and sequestration (CCS) [6-10]. This involves capturing
carbon dioxide that would otherwise be released to the atmosphere, compressing it, transporting
it to a suitable site and injecting it into deep geological formations where it can be safely trapped
or stored. CO2 is a reusable material with a highly attractive application profile and great
potential as a building block to create added value for the chemical industry. CO2 is used as a
precursor for production of various chemicals, as a solvent, in fire extinguishers, cleaning and
surface preparation, etc.; food industry mainly bakery products, beverages, winery, etc.; as inert
gas in pressure tools, welding, lasers, etc.; various mechanical and consumer goods, etc.;
agricultural fumigation, growth stimulation, etc.; medical gasses; oil exploration and recovery;
fuels via bio- transformation; industrial and consumer refrigerants; enhanced coal bed methane
recovery; coolant in nuclear industry [11-14]. Keeping in view the diverse applications, it makes
sense to capture, separate and store CO2. Also some research results reveal that the concentration
of CO2 in the atmosphere has increased from about 310 ppm to over 380 ppm during the last half

century [15].
Current technologies center on post combustion capture, generally using amine scrubbers [16,17]

3
(an inefficient energy-intensive process which generates toxic byproducts/corrosion, requiring
nearly 30% of the output of a power plant to operate and accounts for as much as 70% of the cost
of CCS); pre-combustion methods such as gasifying coal prior to burning it and oxy combustion
which burns coal in pure oxygen rather than air. Other existing CO2 capture technologies in the
context of performance considerations include porous solids, such as zeolites and activated
carbons. All are effective but are expensive and inefficient; and none will work for removing
carbon from the atmosphere. The separation of CO2 from gas streams is still a critical issue.
The development of more proficient technologies for CO2 capture from the flue streams of
power plants is considered key to reduction of greenhouse gas emissions and it’s elimination
from mixed component gas streams has gained a significant amount of attention due to growing
concerns of the environment. Several international climate change initiatives have identified the
urgent need for new concepts and materials to emerge that improve upon current technology and
capture materials offerings. With an ever increasing need for a more efficient, energy-saving and
environmentally benign process for gas separation, adsorbents with tailored structures and
tunable surface properties must be found. Intense efforts in these directions are ongoing. Apart
from establishing new techniques, the development of sorbent materials displaying high
separation performance and low capital cost are of critical importance. Amongst new
technologies for CO2 mitigation, the use of inorganic and nature inspired biomimetic structures
for example enzymes such as carbonic anhydrase, urease and Rubisco wherein quantum and
simulation techniques for development of CO2 fixation catalysts have been proposed [18].
Another class of materials which have seen considerable progress over the last two decades for
CO2 capture are PCPs/MOFs. Porous coordination polymers [19] or metal organic frameworks
[20-22] (PCPs or MOFs) present a potentially viable solution to this problem as a more
environmental friendly and efficient alternative, to sequester more CO2 with less energy required

to regenerate. The structural tuneability of PCPs could allow them to be optimized for specific
type of CO2 capture to be performed (oxy-fuel combustion, pre-combustion capture or post-

combustion capture) and potentially even for the specific power plant in which capture systems
are to be installed.
US Department of Energy has identified PCPs/MOFs as the most promising next-generation
technology for carbon capture. Pike Research has projected that the global market for carbon

4
capture and sequestration will be worth approximately $221 billion by 2030. PCPs/MOFs, a
class of crystalline porous materials constructed by metal-containing nodes bonded to organic
bridging ligands hold great potential as adsorbents or membrane materials in gas separation.
These represent an opportunity to create next-generation materials that are optimized for real-
world applications. Unique characteristics including large surface areas, tunaable pore sizes, high
thermal stabilities [23] and convenient functionalization processes can enable the design of
flexible and dynamic frameworks that reversibly change their structures and properties in
response to external stimuli [24-27]. Such architectures may find applications such as materials
displaying highly selective guest accommodation for gasses [28,29], sensors [30- 33], magnetism
[34,35], luminescence [36-41], anion exchange [42] and proton conduction [43,44] etc.
Interest in imidazole/benzimidazole and carboxylic acid based coordination polymers arises from
the fact that various architectures such as rigid, flexible, interpenetrating, etc., having different
levels of affinity for CO2 sorption/desorption can be designed. As an important family of

multidentate N-donor ligands, imidazole/benzimidazole based derivatives are excellent building


blocks as bridging ligands with excellent coordination ability, versatile conformations, as well as
outstanding stability for constructing 1D, 2D and 3D frameworks. Carboxylate ligands (benzoic
acid based derivative) provide different coordination modes and are potentially attractive for
creating porous structures. Tuning chemical functionalities of the pore walls should help to
increase CO2 uptake and adsorption selectivity.
Research on the subject has shown that CO2 sorption in PCPs/MOFs comprising of large surface
area as well as large pore diameter occurs mostly under high pressure conditions. It is attributed
to absence of any functionality in the structure and devoid of any interactions with CO2. But the
latest scientific research in this area has emphasized that PCPs/MOFs with multifunctional
groups consisting of unsaturated metal sites (UMSs) and Lewis basic sites (LBSs) have affinity
to capture CO2 from gaseous mixtures under STP conditions [45-49]. More recent studies have

also shown the importance of hydrogen bonding as a strategy to enhance the selective uptake of
CO2 probe molecules [50].

This review intends to focus mainly on other influencing factors which decide selective
separation and sorption of CO2 from gas mixtures. These include non-covalent interactions (π-π

bonding, anion-π bonding, CH-π bonding, and H-bonding) present on the pore walls which

5
facilitate adsorption in substantial quantities even if Langmuir surface area is very low. These
interaction potentials in the pores decide quantitatively the sorption of CO2 as function of
temperature and pressure. The review also discusses CO2 sorption/desorption at STP and the

influencing structural attributes of imidazole/benzimidazole and benzoic acid based coordination


polymers in which various architectures such as rigid, flexible, interpenetrating, etc., can be
designed to influence cavity pore size and different types of supramolecular interactions within
the pores.

2. Porous coordination polymers (PCPs)/metal organic frameworks (MOFs)


PCPs/MOFs are unique materials constructed by coordinate bonds between multidentate ligands
and metal atoms or small metal-containing clusters (also called secondary building units or
SBUs) [51–53]. Most of these materials are 3D structures incorporating uniform pores and a
network of channels. The integrity of these pores and channels can be retained after careful
removal of the guest moieties. The remaining voids within the 3D structures then can adsorb
other guest molecules [54,55].
2.1 Synthesis
Synthesis of PCPs is reasonably straightforward, using mostly water soluble metal salts with
organic linkers, namely carboxylic acids, N-donor moieties or a combination of N- and O-
donors in a polar organic solvent. Depending upon precursor and solvent, metal–organic
structures can be formed by self-assembly at temperatures ranging from room temperature to
solvo/hydrothermal conditions [56]; green synthesis and microwave synthesis [57].
2.2 Characterization
Being crystalline and highly porous structures, PCPs are mostly characterized by X-ray
diffraction (XRD) to determine crystallinity and phase purity. Porosity is characterized by
adsorption measurements using nitrogen sorption at 77 K or argon uptake at 87 K, and equivalent
surface areas are calculated according to, e.g., the Langmuir equation with determination of
sorption sites by neutron scattering [58]. It should be considered that the underlying model of
independent, equivalent and non-infringing sorption sites might be different on a molecular level.
Many reports have described localized rather than bulk volume adsorption phenomena on PCPs,
with some even clearly differentiating between various crystallographic sites and adsorption
strengths [58–61]. The adsorbates in the pores of PCPs can be investigated by UV-VIS, IR and

6
Raman spectroscopy [62]. Thermal stability of PCPs is studied by TGA.
2.3 Classification of imidazole, benzimidazole and benzoic acid based coordination
polymers Various architectures (rigid, flexible, interpenetrated, etc.) can be designed by
employing multidentate N-donor ligands based on imidazole/benzimidazole and carboxylate
ligand based on benzoic acid based coordination polymers. Imidazole/benzimidazole based
derivatives are excellent building blocks as bridging ligands with excellent coordination ability,
versatile conformations, as well as outstanding stability for constructing 1D, 2D and 3D
frameworks. Carboxylate ligands (benzoic acid derivative) provide different coordination modes
and are potentially attractive for creating porous structures. The ability to flexibly design and
tune chemical functionalities of such materials results in different levels of affinity for CO2
sorption/desorption.
Variations in structures of these coordination polymers may comprise of ligands in which
imidazole/benzimidazole and benzoic acid can be present either in the same moiety or as
different linkers; only imidazole/benzimidazole based or only carboxylate based. These
variations influence rigidity, flexibility, interpenetration, cavity pore size, and different types of
supramolecular interactions existing within the pores.
Often it has been observed that replacement of imidazol by benzimidazole reduces
interpenetration and results in permanent porosity and high thermal stability. The benzimidazoles
are predominantly basic compounds (pka 5.5) weaker than imidazole (pka 7.0), having the ability
to form salts with acids. The differences in framework architecture are due to that the imidazole
ring is susceptible to electrophilic attack on an annular carbon. Unless there is a strong electron
withdrawing substituent present in the ring, it is much less likely to become involved in
nucleophilic substitution reactions. Compared to imidazole, benzimidazole has a fused benzene
ring which provides sufficient electron withdrawal enabling a variety of nucleophilic substitution
reactions. The presence of basic nitrogen atoms in the heterocycle makes these structures suitable
candidates for CO2 sorption.
For simplicity of understanding, these coordination polymers are divided into the following four
categories:

Category 1: PCPs incorporating imidazole/benzimidazole and benzoic acid based moieties


in the same linker

7
Most porous structures incorporating imidazole and benzoic acid in the same linker do not
display any gas adsorption characteristics due to high fold of interpenetration yet are flexible in
nature after the removal of guest molecules. Replacing imidazole with benzimidazole results in
architectures with permanent porosity due to the presence of extensive aromatic benzene rings on
the pore surface. This enables CO2 uptake because of stabilization of delocalized π electrons of
the ring via dispersion-dominated interaction of the CO2 quadrupole [63]. Aijaz et al. [64]
reported a structure {[Cd(IBA)2].2DMF}n; [HIBA=4-(1H-imidazole-1-yl)benzoic acid] which

does not display gas adsorption characteristics, despite ~39.6% void volume. Inspite of 4- fold
interpenetration arising from spacious nature of a single network (Fig. 1), weak interactions
between imidazole C-H and lattice solvent DMF oxygen atoms stabilize the overall 3D structure.
At elevated temperatures, a densely packed framework is formed after removal of DMF which
restricts entry of CO2 at low pressures [65]. This phenomenon is completely reversible and the
dynamic nature of the framework is associated with the rotational motion of aromatic rings as
well as sliding of the interpenetrating frameworks. Absence of weak forces within the structure is
the reason for dynamic nature of the framework by which permanent porosity disappears.

Fig. 1. Perspective view of a single adamantane cage with HIBA connections between cadmium ions (left) and 4-
fold interpenetration of diamondoid networks (right).
Reproduced with permission from Ref. [64].

The 1H-imidazol-1-yl-containing moieties or its analogues (Scheme 1) can only act as neutral
ligands to utilize terminal N- atoms to ligate metal ions [66], therefore, showing rigid
coordination modes. Comparatively, the 1H-imidazol-4-yl group possesses more flexible
coordination modes [67] due to presence of two free N- atoms and one of these can be
deprotonated to form the imidazolate anion displaying intermolecular non-covalent bonding with

8
the structure to form a stable framework with permanent porosity. Because of these anion dipole-
quadrupole interactions between N- atom and probe molecules takes place [68- 70].

Scheme1 Schematic structures of HIBA and 4-H2IBA.

Replacing imidazole with benzimidazole in the following structure {[Co(bmzbc)2].2DMF}n;


JXNU-1 [Hbmzbc = 4-(benzimidazole-1-yl)benzoic acid], synthesized by Wang and coworkers
[71] which is a layered PCP (Fig. 4) stacked into 3D dense structure with permanent porosity.
Such dense structures result from π-π bonding between benzene rings of benzimidazole and other
intermolecular weak forces (H-bonding). The CO2 absorption exhibits saturated high adsorption
amount due to stabilization of the delocalized π electrons of benzene through π-π bonding
between benzene ring and CO2 quadrupole [72] leading to non-hysteretic behavior at low

temperature and pressure.

Fig. 2. Structure (left) and CO2 sorption isotherms (right) of {[Co(bmzbc)2].2DMF}n

Reproduced with permission from Ref. [71].

However, the ligand structure: 3,5-di(1H-imidazol-1-yl) benzoic acid [73,74] where a second
imidazole ring is incorporated leads again to an interpenetrated framework with dynamic

9
behavior devoid of porosity thus no sorption. Absence of any interaction sites within the
framework, to make it stable, is the reason for its dynamic nature due to which permanent
porosity does not exist.
Modifying such a structure to {Co2(BIIBA)3]·(NO3)·5H2O}n; [BIIBA = 3-(1H-
benzo[d]imidazol-1-yl)-5-(1H-imidazol-1-yl)benzoate] as reported by Agarwal et al. [75] results
in a rigid, porous 3D framework with asymmetrical ligand and two fold interpenetration (Fig. 3).
The presence of benzene rings of benzimidazole in the structure results in reduced
interpenetration due to π-π stacking interactions between two benzene rings as well as H-bonding
interactions between hydrogen of benzene rings and oxygen atoms of carboxylate groups of the
ligand. This provides stability to the framework and will lead to adsorption of CO2 due to π-π
interactions which includes a weak electron donation from the aromatic ring to the probe
molecule [76].
(a) (b)

(c)
(d)

Fig. 3. Single hexagonal network (a) space filled model showing large channels (b) 2-fold interpenetrated
framework (c) and space filled model (d) of {Co2(BIIBA)3]·(NO3)·5H2O}n along c axis. (hydrogen atoms and
lattice water molecules are omitted for clarity).
Reproduced with permission from Ref. [75].

Moving forward by replacing both imidazoles by benzimidazole rings, an increment in π-π


stacking interactions is obtained along with enhanced CO2 adsorption. Synthesis of the

10
following 3D isostructural series: {[Ni2(DBIBA)3]⋅Cl·18H2O}n, {[Co2(DBIBA)3]·Cl·9H2O}n

and {[Mn2(DBIBA)3]⋅Cl·3H2O}n (DBIBA = 5-di(1H-

benzo[d]imidazol-1-yl)benzoate), sorption/desorption characteristics without hysteresis [77,78].


2
The first two PCPs having surface area 782 and 794 m /g show selective and high uptake at low
2
temperature and pressure (Fig. 4 and 5) while third structure has largest surface area (823 m /g)
among all with no interpenetration, show high CO2 three and displays no selectivity for CO2
(Fig. 6). This is because of the slightly flexible nature as a result of weaker π-π bonding between
benzene rings of benzimidazole moieties, which are evenly distributed within the cavities and
intermolecular forces (anion-π, CH-π, H- bonding, etc.) in comparison to the former two
structures where these interactions are very strong. The ionic radii of Mn(II) metal node is the
largest or its density is very less compared to Ni(II) and Co(II). Due to this larger size, the
intensity of interactions decrease leading to slightly flexible nature, higher surface area as well as
large pore apertures which facilitate non-selective sorption of CO2, N2 and H2.

11
(a)

(b) (c)
Fig. 4. Strong π-π and anion-π bonding (a) CH-π and H-bonding (b) gas sorption isotherm (c) of
{[Ni2(DBIBA)3]⋅Cl·18H2O}n, green lines showing π-π bonding and purple lines for anion-π interactions while red

and cyan broken lines show CH-π and H-bonding respectively.


Reproduced with permission from Ref. [77].

12
(a) (b)

Fig. 5. Strong π-π and anion-π bonding (a) gas sorption isotherm (b) of {[Co2(DBIBA)3]·Cl·9H2O}n.

Reproduced with permission from Ref. [78].

(a) (b)

Fig. 6. Weak π-π and anion-π interactions shown by fragmented green lines (a) and gas sorption isotherm (b) of
{[Mn2(DBIBA)3]⋅Cl·3H2O}n

Reproduced with permission from Ref. [77].

13
Synthesis of {[Mn2(DBIBA)3]·Cl·DMF·8H2O}n PCP from a room temperature anion exchange
reaction of {[Mn2(DBIBA)3]·(NO3)·3DMF·4H2O}n [DBIBAH = 3,5-di(1H-benzo[d]imidazol-
1-yl)benzoic acid] [79] leads to a significant increase in gas uptake at low temperature and
pressure (Fig. 7). This is due to shifting of anions from the center to the corners of the cavities by
which surface area increases. Chloride anions creating strong H-bonding and interactions
facilitate dense packing of CO2 molecules at low temperature and pressure. In this case,
comparatively smaller voids containing high CO2 volume wherein strong London dispersive
forces (induced dipole-induced dipole) between CO2 molecules dominate, leading to hysteretic
desorption [80].
Increases of bulky and high capacity π-π stacking in organic ligands of the first category exert an
important influence on the formation of resulting diverse structures. In PCPs with benzimidazole
moiety, strong π-π interactions help to make the frameworks more stable thus preventing
interpenetration. At the same time, weak π electron donation from benzene ring to CO2

quadrupole promotes high gas sorption.

14
(a) (c)

(b) (d)

Fig. 7. Structure and gas adsorption/desorption isotherm of {[Mn2(DBIBA)3]·(NO3)·3DMF·4H2O}n (a), (b) and
{[Mn2(DBIBA)3]·Cl·DMF·8H2O}n (c), (d).

Reproduced with permission from Ref. [79].

Category 2: PCPs consisting of imidazole/benzimidazole based ligands


The second category comprises of architectures that include only imidazole/benzimidazole based
ligands. PCPs such as [Co(HL)]DMF·H2O and [Co(HL)]2H2O; [H3L=1,3,5-tri(1H-imidazol-4-

yl)benzene] reported by Chen et al. [78] are isomeric porous structures. The first structure
displays high selective uptake of CO2 at low temperature and pressure while the second structure
shows non-selectivity except at higher temperatures (Figs. 8 and 9). This is ascribed to the free
deprotonated imidazolate anions generating delocalized π electronic cloud leading to dipole-
quadrupole interactions with CO2 molecules [68-70] in the first PCP, while in second only

neutral N- atoms are present as Co(II) coordinates with different N- atoms of imidazole of the

15
ligand. The dipole-quadrupole interactions between imidazolate anions and probe molecules and
1D channels in the first PCP leads to high selective uptake of CO2 without hysteresis. But at
higher temperatures the kinetic energy of CO2 molecules is high which leads to lower uptake

and significant hysteresis with temperature. In the second structure a reduced electrostatic
environment and 3D channels leads to non-specific gas adsorption.

Fig. 8. The coordination environment and 3D porous structure of [Co(HL)]DMF·H2O (a), (c) and [Co(HL)]2H2O

(b), (d) isomers.


Reproduced with permission from Ref. [81].

16
CO2 at 195 K

(c) (d)

N 2 at 195 K

CO2 at 195 K

Fig. 9. CO2 adsorption isotherm at 195 K, temperature dependent stepwise and hysteretic sorption behavior and

selective adsorption of CO2 for [Co(HL)]DMF·H2O (a) and (b), the gas adsorption isotherm at 195 K and selective

uptake of CO2 at 273 K and 298 K for [Co(HL)]2H2O (c) and (d). Filled shapes: adsorption; open shape:
desorption.
Reproduced with permission from Ref. [81].

Fig. 10. ZIF crystal structures shown with heterolinks (left), the CO2 and CO adsorption isotherms for ZIF-69 at 273

K.
Reproduced with permission from Ref. [82].

17
ZIF 68-70 display selective CO2 sorption from CO2/CO gas mixtures [82] because of their

porous walls containing different types of ligands (Fig. 10) which contribute to high thermal and
chemical stability to the evacuated frameworks after solvent removal, leading to no
interpenetration. The highest CO2 separation is exhibited by ZIF-69 because of the presence of

π- π interactions between benzene rings of benzimidazole and CO2 quadrupole. At the same
time, due to the presence of chloride ion substituted groups in the benzimidazole ring, overall gas
separation increases due to Lewis acid- Lewis base interactions [80]. Comparatively, these
interactions are weaker in case of ZIF-68 and 70 because of different ligand moieties present. In
porous polymer networks (PPNs) reported by Zhou et al. [83], CO2 adsorption as well as
CO2/N2 selectivity (Fig. 11) is significantly enhanced because of exposed edges and faces of
phenyl and benzimidazole rings lead to high Langmuir surface area. Extensive π-π bonding
interactions between benzene rings stabilize the CO2 quadrupole through weak electron donation
from the aromatic rings to CO2 molecules as well as dipole-quadrupole interactions between free

N- atoms of imidazole groups and probe molecules [84]. At lower temperature, significant
hysteresis was observed which is due to dense packing of CO2 molecules in the cavities. At
higher temperature, greatly reduced CO2 uptake and hysteresis is obtained due to higher kinetic

energy of probe molecules which hinder uptake and compact packing.

18
(a)

(b) (c)

Fig. 11. Benzimidazole incorporated PPN-101 (a) CO2 isotherms at 273 K and 298 K (b) CO2 isotherms at 195 K

(c).
Reproduced with permission from Ref. [83].

From the above examples, it is evident that replacement of 1H-imidazol-1-yl moiety by 1H-
imidazol-4-yl groups, provides flexible coordination modes in the assembly process as well as
free deprotonated imidazolate anions which generate delocalized π electronic clouds leading to
dipole-quadrupole interactions between imidazolate anion and probe molecules. Benzimidazole
based structures show high CO2 uptake because of aromatic groups consisting of π-π interactions

which are largely exposed in the channels, where net charge is transferred from the aromatic
rings to the CO2 molecules.

Category 3: PCPs having architectures of imidazole/benzimidazole and benzoic acid as


different linkers

19
The third category comprises of architectures of imidazole/benzimidazole and benzoic acid
based PCPs. In these types of PCPs flexibility comes due to two separate linkers which can
adjust according to coordination modes of metal ion.
Beginning with a structure such as [Zn(L)(1,4-BDC)] [L = 1,4-di(1H-imidazol-4-yl)benzene],
investigated by Chen et al. [85], there is no gas uptake because of the high fold of
interpenetration due to a large vacancy of the diamond net [86]. Further, we look at the
following solvent dependent structures reported by Hua et al. [87] {[Zn2(tib)2(BDC-
Br)]2·2SO4·17H2O} and [Zn4(tib)2(BDC-Br)3(H2O)4SO4]·7.5H2O·2.5DMF} ;[tib = 1,3,5-
tris(1-imidazolyl)benzene; H2BDC-Br = 2-bromo-1,4-benzenedicarboxylate] with a non-

coordinated bromo-functional group for additional interaction sites [88]. Both structures have
some degree of flexibility after desolvation. The first structure having large void volume (29.5%)
but very less surface area shows moderate CO2 gas adsorption at low temperature and pressure

[Fig. 12] while the second exhibits low uptake. Due to absence of aromatic-aromatic interactions
surface area is very less for adsorption of probe molecules because of no weak contributing
bonding interactions. Electrostatic environment created by sulphate anions in the cavities of the
first structure leads to adsorption, where as in the second structure uptake is influenced by the
presence of unsaturated metal sites (generated by removal of coordinated water molecules at
higher temperature) even though the surface area is very low. Desorption in both PCPs is
accompanied by hysteresis which is due to high interaction potential between framework and
probe molecules. There can be covalent bonding through charge migration or non-covalent long
range bonding by electrostatic interaction between the dipole moment of anion and the
quadrupole moment of CO2. After saturation of covalent bonding, weak bonding starts between

these two moieties [89].


Agarwal et al. [90] reported the structures {[Cd(NPBI)(BDC)(H2O)]·2H2O}n and
{[Zn(NPBI)(BDC)]·H2O}n, [NPBI = 1,1'-(4-nitro-1,3-phenylene)bis(1H-benzo[d]imidazole),

BDC = 1,3-benzenedicarboxylate]. Both are 1D chain structures but due to strong hydrogen and
π-π bonding between chains, structure is constructed in to 3D supramolecular architecture with
1D dumbbell shaped small cavities. Both PCPs are highly flexible at room temperature, with the
first structure showing high selective CO2 uptake at low temperature and pressure with respect to
very small surface area (Fig. 13). This is because of the presence of -NO2 substituent at the pore

20
surface wherein dipole-dipole interactions are not primary but local electrostatic and polarization
interaction are dominant between the lone pairs, or the acidic hydrogen, and the CO2 quadrupole
[91]. Hysteresis on desorption is because of strong bonding between CO2 molecules due to

dense packing in very small surface area. Second PCP is near similar in structure to the first with
the same electronic environment within its cavities, but has lower void volume and surface area.
This structure will display selective sorption characteristics.
The structure {[Ni3(TBIB)2(BTC)2(H2O)6].5C2H5OH.9H2O}n; [TBIB = 1,3,5 tris(1H-

benzo[d]imidazole-1-yl)benzene; BTC = 1,3,5-benzene tri carboxylate] [92] is a layered porous


flexible architecture consisting of strong H-bonding between layers through coordinated water
molecules leading to a 3D framework with 1D channels. This structure has lower surface area
2
(297 m /g), higher thermal stability and shows comparatively higher selective CO2 sorption
characteristics and reduced hysteresis at low temperature and pressure (Fig. 14) compared to
afore discussed structures.
(a) (b)

(c)

Fig.12. 3D framework of {[Zn2(tib)2(BDC-Br)]2·2SO4·17H2O} (a) 2D network of [Zn4(tib)2(BDC-


Br)3(H2O)4SO4]·7.5H2O·2.5DMF} (b) Gas adsorption isotherms of CO2 (195 K) for 3D framework (a) and 2D

network (b).
Reproduced with permission from Ref. [87].

21
(a)
(b)

(c) (d)

Fig. 13. Packing, space filled diagram, π-π bonding between chains and gas sorption isotherm of
{[Cd(NPBI)(BDC)(H2O)]·2H2O}n (a), (b), (c) and (d).
Reproduced with permission from Ref. [90].

Due to aromatic π-π interactions between two benzene rings of benzimidazole within the layered
structure, polarization of the π – electronic cloud towards CO2 quadrupole and electrostatic

interactions generated through free oxygen atoms (acting as Lewis base) of the BTC ligand
where all carboxylate groups are present in a monodentate fashion is facilitated. This
environment makes this structure a potential candidate for post combustion CO2 separation. The

observed desorption with slight hysteresis is due to the lower kinetic energy of probe molecules
because of dense packing at lower temperature.

22
Fig. 14. 3D structure generating through H-bonding interactions between layers (left) CO2 adsorption isotherm of

{[Ni3(TBIB)2(BTC)2(H2O)6].5C2H5OH.9H2O}n.

Reproduced with permission from Ref. [92].

From the above discussion, it is clear that PCPs incorporating imidazole/benzimidazole and
benzoic acid based linkers can show interpenetration or none, rigid or flexible structures with
varying degree of CO2 sorption characteristics.

Category 4: PCPs incorporating only benzoic acid based ligands


The fourth category comprises of CPs having ligands incorporating benzoic acid derivatives.
Examples in this category include layered architecture such as
{[Cu2(HBTB)2(H2O)(EtOH)]H2O.EtOH}; [H3BTB = 1,3,5-tris(4-carboxy-phenyl)benzene)]

synthesized by Mu and coworkers [93]. Despite being a 2-fold interpenetrating network, this
shows preferential sorption of CO2 over CH4 at ambient pressures (Fig. 15) because of the

electrostatic environment created by uncoordinated carboxylic groups [91] and presence of


unsaturated metal sites (UMCs) [94]. Higher temperature and pressure is required to enhance gas
uptake within the 3D interconnected pores to reduce the influence of H-bonding between two
interpenetrated folds. If organic linker 1,3,5-tris(4-carboxy-phenyl)benzene is replaced by 1,3,5-
tris(4-carboxy-naphthyl)benzene (Scheme-2), in the following interpenetrated mesoporous
structure [Cu3(BTN)2(H2O)3] (NJTU-1) higher CO2 uptake is observed because of large

Langmuir surface area as evaluated by Duan and coworkers [95]. At lower temperature and
pressure the uptake is comparatively lower than at higher temperature. This is due to the π-π

23
stacking of aromatic rings of naphthalene between two interpenetrated units. The higher uptake
of CO2 probe molecules is effected by π-π interactions between aromatic moieties and non-
aromatic CO2 quadrupole which results in loosening of the interpenetrated units (due to

reduction of π-π stacking of aromatic rings) providing extra surface area at higher temperature
and pressure despite higher kinetic energy [96].

Scheme 2 Molecular structures of the ligands H3BTB and H3BTN.

(a) (c) (d)

(b) (e)

Fig. 15. The stack of different layers A (red), B (green) at [010] plane (a) and single-component isotherm of CO2,
CH4 in {[Cu2(HBTB)2(H2O)(EtOH)]H2O.EtOH} at 298K (b) packing view of three types of cages, (c) low

temperature and pressure CO2 isotherm (d) and high temperature and pressure CO2 isotherm for NJTU-1.

24
Reproduced with permission from Ref. [93 and 95].
Other examples include Cu3(BTC)2(H2O)9 (HKUST-1) [BTC = 1,3,5-benzentricarboxylate]

reported by Liang et al., [97]. From Figure 16 it is clearly seen that CO2 adsorption capacity
increases with increasing pressure and decreasing temperature. This is probably due to the low
density of open metal sites or the “blocked” open metal sites by the framework [93]. Higher gas
uptake is facilitated with reducing temperature due to lower kinetic energy of CO2 molecules as
the pores are narrow. The CO2 adsorption capacity of Cu-BTC declines after water sorption,

particularly at high relative humidity because pore volume occupied by H2O molecules will
reduce the free volume available for gas sorption. Absence of π-π aromatic interaction sites, not
significant CO2 adsorption is obtained despite the very high surface area.

Fig. 16. Gas sorption isotherms of Cu-BTC measured in the range 0 to 15 bar and at temperatures ranging from 25
to 105 °C.
Reproduced with permission from Ref. [97].
3+ 3+
Other architectures include flexible MIL-53 series M(OH)(O2C-C6H4-CO2), (M = Al , Cr )
reported by Bourrelly et al. [98] which normally display stepwise adsorption consisting of a 1500
2
m /g Langmuir surface area. At low pressures, adsorption is due to the specific interaction sites
produced by coordinated hydroxyl groups. Once these interaction sites are saturated, enhanced
CO2 uptake at high pressures is solely a physical phenomenon (Fig. 17). Some ultrahigh pore
MOFs, reported by Yaghi et al. [99,100] such as Zn4O(BDC)3 (MOF-5), [BDC=1,4-
benzenedicarboxylate], Zn4O(BTB)2 (MOF-177), [BTB=4,4′,44′′benzene-1,3,5- triyl-

25
tribenzoate], and Zn4O(BTE)14/9(BPDC)6/9 (MOF-210), [BTE=4,4′,4′′-[benzene- 1,3,5triyl-

tris(ethyne-2,1-diyl)]tribenzoate, BPDC=biphenyl-4,4’-dicarboxylate] display and high pressures


with increasing pore size (Fig. 18). In these structures there are no interaction sites present in the
pores thus high pressure is required to drive probe stepwise CO2 adsorption due to bulk

condensation at near room temperature molecules uptake.

(a)

(b) (c)

Fig. 17. Hydration and dehydration process occurring in MIL-53 (Al, Cr), (Left) MIL-53LT (hydrated). (Right)

MIL-53HT (dehydrated) (a) CO2 and CH4 adsorption isotherm of MIL-53 (Al) (b) and MIL-53 (Cr) (c).

Reproduced with permission from Ref. [98].

26
Fig. 18. CO2 adsorption isotherms of MOFs.
Reproduced with permission from Ref. [100].
SNU-77 reported by Park et al. [101] is a doubly interpenetrating Zn4O-type MOF which

exhibits no breathing effect and only its organic components perform rotational motions across a
range of temperatures. During activation this structure undergoes transformations to afford
different fine structures (SNU-77R, SNU-77S and SNU-77H) varying in pore size. As an
2
example SNU-77H (void volume 69%, Langmuir surface area 4180 m /g) shows a large uptake
of CO2 at low temperature and pressure, while having reduced adsorption at higher temperatures

and pressures (Fig. 19). This is because of edge-to-face π-π aromatic stacking interactions
present between two interpenetrating nets at low temperature and pressure, which facilitate
dispersion dominated π-electron delocalization towards CO2 molecules [102]. But at higher
temperatures there is a drastic reduction in uptake due to high rotational motion of organic
components. High pressures do not influence gas uptake as the kinetic energy of CO2 molecules
increases and at the same time there are high rotational motions within the pores.

27
(a)

(b) (c)

Fig. 19. X-ray crystal structure (a), gas sorption isotherm of SNU-77H and (c) CO2 sorption isotherm at high

temperature and pressure. Filled shapes: adsorption; open shapes: desorption.


Reproduced with permission from Ref. [101].
Lu et al. [103] synthesized NJU-Bai13 which shows large uptake of CO2 (Fig. 20) at high

temperature and pressure which is due to partial positive charge present on coordinatively
unsaturated metal sites as well as π electronic delocalization from alkyne units or phenyl rings to
probe molecules. Large pores and high kinetic energy of CO2 molecules enables this structure to

be an excellent candidate for high adsorption at high temperature. High pressure facilitates dense
packing of CO2 molecules within the cavities.

28
(a) (b)

Fig. 20. The connection mode between kagome layer (a) and gas adsorption isotherms of NJU-Bai13 at high
pressure range (0–20 bar) (b).
Reproduced with permission from Ref. [103].
Moving on to PCN-72 [104], this has a similar structure to MIL-53 except that it has a
coordinated DMSO molecule (Fig. 21) in place of a –OH group coordinated with metal in MIL-
53. At lower temperature activation, this structure is nonporous (free volume ~7.6%) but after
activation at higher temperatures (250 oC or higher) removal of the coordinated DMSO solvent
generates open metal sites. Partial positive charge on the open metal sites is the driving force for
the selective moderate uptake at low temperature and pressure.

Fig. 21. Single crystal X-ray structure (left) and gas adsorption isotherms of PCN-72 after being activated at 360 °C
(right).
Reproduced with permission from Ref. [104].
Although CO2 adsorption capacity of PCN-72 is much lower than the best performing Mg-

MOF- 74 [105] which has the highest reported adsorption capacity due to its larger pore volume
2
and Langmuir surface area (~1500 m /g). (Fig. 22). The driving force for higher CO2 adsorption

29
+2
is the presence of the highly charged open Mg sites with smaller ionic radius leading to strong
CO2–metal interactions. Consequently it could facilitate a greater degree of polarization on the
adsorbed CO2 molecules [106-108]. At higher temperatures, time for saturation of uptake within

the large pores is shortened. Uptake decreases with increasing temperature because of the high
kinetic energy of probe molecules and the large amount of heat released during strong
interactions.

Fig. 22. Crystal structure (left) and CO2 adsorption isotherms of Mg-MOF-74; open symbols, adsorption; filled

symbols, desorption; shot dash dot, Langmuir equation; solid line, Sips equation.
Reproduced with permission from Ref. [105].
Benzoic acid based PCPs normally display adsorption at a higher pressure because of larger pore
diameters and unsaturated metal sites present within the cavities. In systems comprising of
aromatic π-π interaction sites adsorption at lower temperatures and pressures is facilitated.

Table-1 CO2 adsorption capacities of PCPs/MOFs of discussed categories.


Surface
Void
area Pres
volu Qst Temperature Reason (dominant
PCP/ MOF (Langmu CO2 uptake sure Ref.
me (kJ/ (K) interactions)
ir), (bar)
(%) mol)
(m2g−1)

Dynamic nature of
{[Cd(IBA)2].2DMF} n ------ 39.6 ----- None ------ ------ framework after guest 64
removal
dispersion dominated
{[Co(bmzbc)2].2DMF}n ------ ----- 35.4 124/91/ 42 cm3 g-1 195/273/298 1 weak interactions 71
between aromatic rings

30
of benzimidazole and
CO2 quadrupole
Adsorption will be due
to delocalization of π
{Co2(BIIBA)3]·(NO3)·5H2O} ------ 18 % ----- ------ ------ ----- 75
electrons of benzene
towards CO2 molecule
135.4/50.7/ 44.8
{[Ni2(DBIBA)3]⋅Cl·18H2O}n 782 32.7 16.1 195/273/298 1 ibid 77
cm3 g-1
140.5/ 52.1/
{[Co2(DBIBA)3]·Cl·9H2O}n 794 32.8 16.0 195/273/298 1 ibid 78
44.2 cm3 g-1
160.4/52.7/ 44.1
{[Mn2(DBIBA)3]⋅Cl·3H2O}n 823 34 12.7 195/273/298 1 ibid 77
cm3 g-1
{[Mn2(DBIBA)3]·(NO3)·3DMF·4H2O}n 780 28.3 27.4 155/73/50 cm3 g-1 195/273/298 1 ibid
3 1
{[Mn2(DBIBA)3]·Cl·DMF·8H2O}n 800 33.6 33.2 260/100 cm g- 195/273 ----- ibid 79
3 1
{[Co(HL)]DMF·H2O} 1170 43 ----- 163.71 cm g- 195 1 ibid
Adsorption due to 3D
[Co(HL)]2H2O 32.7 89.85/49.33/36.57 81
------ 42.4 195/273/298 1 channels and reduced
2 cm3 g-1
interactions
No adsorption due to
[Zn(L)(1,4-BDC)] ------ ----- ----- ----- ------ ----- high fold of 85
interpenetration
Adsorption due to strong
{[Zn2(tib)2(BDC-Br)]2·2SO4·17H2O} ------- ----- ----- 66.484 cm3 g-1 195 0.99 anionic interactions and
large surface area
87
Adsorption due to open
{[Zn4(tib)2(BDC-
------- ----- ----- 36.084 cm3 g-1 195 1 metal sites (blocked)
Br)3(H2O)4SO4]·7.5H2O·2.5DMF}
despite low surface area
Adsorption due to
3 1
{[Cd(NPBI)(BDC)(H2O)]·2H2O}n ------- 13.6 40 80/30/23 cm g- 195/273/298 1 dispersion dominated
weak interactions
90
Adsorption will be due
{[Zn(NPBI)(BDC)]·H2O}n ------- 8.1 ----- ------ ------- ----- to dispersion dominated
weak interactions
Adsorption due to strong
{[Ni3(L1)2(L2)2(H2O)6].5C2H5OH.9H2O}n 297.63 36 ----- 100/55/40 cm3 g-1 195/273/298 1 92
electrostatic environment
Adsorption due to
dispersion dominated
[Cu3(BTN)2(H2O)3]
3030 ----- 25.1 350/298 cm3 g-1 273/298 20 weak interactions 95
or NJTU-1
because of large
aromatic rings

Adsorption due to
29.2- electrostatic interactions
Cu3(BTC)2(H2O)9 ------ ----- 12.7 mmol/g 298 15 97
35.0 by unsaturated metal
or HKUST-1
sites with adsorbate

M(OH)(O2C-C6H4-CO2) or MIL-53 ------ 10 mmol/g 304 30 Adsorption due to phase 98

31
M = Al/Cr transition from narrow
pore at low pressure to
larger pore at high
Pressure

Zn4O(BDC)3 [MOF-5] 4400 79 750 mg/g 298 20


Zn4O(BTB)2 [MOF-177] 5340 83 1470 mg/g 298 45
Adsorption due to large 100
MOF-200 10400 90 2400 mg/g 298 50
surface area
MOF-205 6170 85 1500 mg/g 298 37
Zn4O(BTE)14/9(BPDC)6/9 [MOF-210] 10400 89 2400 mg/g 298 50
Adsorption due to
4020/ 19.9
[Zn4O(TCBPA)2] ~900 cm3 g-1 dispersion dominated
4120/ to 195 42 101
SNU-77R/ SNU-77S/ SNU-77H (SNU-77H) weak interactions as well
4180 19.4
high surface area
Adsorption due to
electrostatic interactions
{Cu2(CPEIP)0.5·2H2O}∞ [NJU-Bai13] 4168 78.9 21.2 21.2 mmol/g 298 20 103
by unsaturated metal
sites with adsorbate
Adsorption due to
electrostatic interactions
PCN-72 75/ 35/ 30 cm3 g-1 195/ 273/ 295 1 104
by unsaturated metal
sites with adsorbate
Adsorption due to
7.22/ 7.059/6.949 electrostatic interactions
Mg-MOF-74 1733 73 278/298/318 1 105
mmol/g by unsaturated metal
sites with adsorbate

32
PCPs with high CO2 uptake and no hysteresis

Imidazole/ Imidazole/ Benzoic acid based


Imidazole/
benzimidazole and benzimidazole based benzimidazole and
benzoic acid in same benzoic acid Large pores/no
linker in different interactive sites
linker
Adsorption infludenced
by high temperature and
pressure
Narrow pores, Narrow/large pores, Narrow pores,
interactive sites interactive sites interactive sites Large pores/
(dispersion-dominated) (dispersion-dominated) /dispersion-dominated interactive sites

Adsorption infludenced
by high temperature and
pressure

Narrow pores/
Adsorption influenced by Adsorption influenced by interactive sites
Adsorption influenced by
low temperature/ low temperature/ low temperature/
low pressure low pressure low pressure Adsorption infludenced
by low temperature and
pressure

Scheme 3 Overview of PCPs displaying high CO2 uptake and no hysteresis.

2.4 Benzene/imidazole and CO2 interactions in PCPs


Benzene has a symmetrical structure, therefore its dipole moment is zero. In its structure, only
dispersion forces exist because the electrons around the nucleus are symmetrical. Having a larger
area for electrons to move in the electron cloud, dispersion forces are more likely to occur.
Further, it has a strong permanent quadrupole moment which is of the form wherein two dipoles
2
are simply aligned end-to-end. In the structure, because an sp C is more electronegative than H,
δ− δ+
it creates six C H bond dipoles which add up to a molecular quadrupole of electronegative
potential at the face and positive along the edge.
In benzene dimer, π-π stacking interaction occurs through stabilization of attractive electrostatic
quadrupole-quadrupole as in the case of T-shaped configurations [109], as the positive
quadrupole of one benzene ring interacts with the negative quadrupole of the second ring and

33
London forces are diminished due to decreased polarizability [110]. Parallel-displaced
configurations having near equivalent stability to T-shaped configurations also exist along with
London dispersive forces (dispersion interactions) [111].
In PCPs where pores are exposed with benzene rings, π-π interactions (repulsive electrostatic
forces) with adjacent benzene rings, are the primary driving forces that facilitate the entry of
CO2 molecules. The adsorption of CO2 molecules within the cavities creates more stable

electrostatic environment due to the easy displacement of negative quadrupole of benzene rings
towards the positive quadrupole of CO2 molecules (generated through considerable charge
separation in the C=O bonds). In cases where there are no π-π interactions, but benzene rings are
present at the pore surface, delocalization of π electrons of the benzene ring towards CO2
quadrupole is effected through weak dispersion-dominated interaction.
During adsorption, CO2 molecules are positioned on the top of benzene rings equidistant from

all the aromatic C atoms; this conformation is stabilized by the above modes of interaction
[63]. From binding energy calculations at the MP2/def2QZVPP level of theory, π-π interactions
significantly contribute to the total interactions between CO2 and aromatic molecules [112].
According to the π-π bonding theory [113,114] based on dispersive/repulsive interactions, weak
adsorption occurs through π-π interactions within the pores. This is due to stabilization of
delocalized π electrons of benzene through dispersion-dominated interaction of the CO2

quadrupole. However, Wheeler and coworkers [115,116] have reported that these interactions are
stronger between aromatic and non-aromatic moieties compared to that between two aromatic
moieties.
In the case of imidazole, which has a dipole moment because of its high polarity, the quadrupole
moment in CO2 molecules through dipole-induced dipole (Vander Waals forces) is created.
There are two different mechanisms by which imidazole and CO2 can interact: a) Lewis acid–
Lewis base interactions (the most important mechanism) with interaction energies correlating
with the shape and the intensity of the electrostatic potential surfaces and b) hydrogen bonding
interactions which provide further stabilization through strong hydrogen bonds with the -NH
group [84]. A major interpretation in this area is that non-coordinated excessive nitrogen sites as
Lewis-base centers incorporated in PCPs can significantly enhance CO2 uptake capacity and

selectivity on account of the dipole-quadrupole interactions.

34
If we combine benzene and imidazole to form benzimidazole, which has a higher dipole moment
than imidazole, one can take advantage of both benzene and imidazole functional sites for CO2

interactions [117,118]. This is of great interest in the synthesis of PCPs/MOFs for post
combustion CO2 sorption. Thus a structure with naphthimidazole should possess a larger surface

area and enhanced dispersion-dominated interaction potential because of enhanced aromaticity


due to two benzene rings and imidazole functional site.
Other structural features that influence CO2 sorption include:

Structural robustness: to provide robustness, aromatic moieties should be present in the


framework.
Coordination modes: metal nodes should have flexibility to freely coordinate with a number of
donor atoms.
Cationic frameworks: structures with anions exhibiting different modes of weak interactions
(H-bonding and anion-π).
Interpenetration: creation of confined narrow pores/channels for entrapment of small
molecules, due to strong electrostatic potential, which may find application in selective
separation of probe molecules from mixed gases.
Donor groups: presence of two linkers (one for O- and the other for N- donor atoms) in the
framework will lead to a varying degree of flexibility and adsorption.
Unsaturated metal sites (UMSs): the open metal sites with high charge density act as anchors
for probe molecules and are typically obtained by introducing coordinated solvent molecules
(e.g. H2O, DMF, etc.) as terminal ligands. These coordinated solvent molecules can be readily

removed by desolvation at elevated temperatures and/or under a vacuum.


Lewis basic sites (LBSs): Typically, the interaction between localized dipoles of uncoordinated
N-containing groups and the high quadrupole moment of CO2 induces dispersion and

electrostatic forces influencing adsorption and separation.


Functional sites: Besides the UMSs and LBSs, other polar functional groups (e.g. –F, –Br, – Cl,
–OH, –CN, –NO2, –SO3) may impose a similar effect and offer enhancement of CO2–PCP

interactions [119-121]. Understanding of the above interaction sites and interactions of CO2

molecules/host within the PCPs/MOFs allows for flexible design of optimal framework

35
chemistries for both adsorption and desorption applications. Infrared spectroscopy (indirect
information of molecular adsorption process) is a powerful tool for investigating CO2
molecules/host interactions. The symmetric stretch mode (ν1) for the linear CO2 molecule is

Raman active but not IR active. On adsorption at a given interaction site, antisymmetric modes
(ν2 and ν3) of CO2 become IR active [122]. As an example characterization of the type of

host/guest interaction, in MIL-53 (category 4) has been evidenced by IR spectroscopy through


the formation of an electron donor acceptor complex between CO2 and MOF. The adsorption of

−1
the gas molecules leads to a reasonable shift from 10 to 15 cm of the stretching mode and
splitting of the bending mode due to removal of degeneracy of bends [123]. In case of flexible
PCPs [124] Raman spectroscopy is more useful because the phonon modes of the PCPs do not
overlap as in case of IR wherein a large number of combination and overtone bands exist. By
Raman spectroscopy, changes in specific bonds within the PCP framework can be correlated to
structural changes and guest/host interactions. A technique for direct evidence of CO2/pore

surface is in situ crystallography of CO2@PCPs/MOFs [49, 125-127]. This study effectively

provides information about the location of gas molecules diffused within the pores. For example
a porous framework CaSDB (SDB: sulfonyldibenzoate) shows strong affinity for CO2 despite
not having any functional groups or open metal sites [128]. Single crystal X-ray diffraction along
with Raman and IR spectroscopy, theoretical calculations and DSC-XRD carried out for this
structure after CO2 adsorption showed that the SDB linkers were responsible for adsorption. The
main interaction between CO2 and the pore surface was between delocalized π aromatic system
of both aromatic system of both phenyl rings of the linker and the molecular quadrupole of CO2.
Carbon of CO2 molecules are situated between two centroids of the aromatic rings (Fig. 23). The
guest molecules are oriented near parallel to both phenyl rings with oxygen atoms positioned as
far as possible from the centroids and are relatively close to hydrogen atoms.

36
Fig. 23. X-ray structure of CO2 bound in CaSDB. a) The difference electron density map calculated before

assigning CO2 showing the carbon dioxide inside the channel (oxygen atoms are above and below the plane), the
white wire represents the superimposed structure of the CaSDB framework. b) Packing along [010] showing the
location of CO2 ; blue spheres: calcium, red: oxygen, black: carbon, and yellow: sulfur. Hydrogen atoms are omitted

for clarity. c) Local environment of the adsorbed CO2. Dashed lines represent phenyl.....CO2 π quadrupole
interactions. The CO2 molecule occupies two equivalent positions, with 32% occupancy on each.
Reproduced with permission from Ref. [128].

2.5 Selection of best category PCPs:


Table-2 Selected PCPs/MOFs from discussed categories
Void Surface area Pres
Catego Qst CO2 uptake Temperature
Best PCPs/MOFs volume (Langmuir), sure
ry 2 1
(kJ/ (cm3 g-1) (K)
(%) (m g− ) (bar)
mol)

{[Ni2(DBIBA)3]⋅Cl·18H2O}n 32.7 782 16.1 135.4/50.7 195/273 1


{[Co2(DBIBA)3]·Cl·9H2O}n 32.8 794 16.0 140.5/ 52.1 195/273 1
1
{[Mn2(DBIBA)3]·(NO3)·3DMF·4H2O}n 27 780 27.4 155/73 195/273 1
{[Mn2(DBIBA)3]·Cl·DMF·8H2O}n 33 800 ----- 260/100 195/273 1

37
2 {[Co(HL)]DMF·H2O} 43 1170 ----- 163.71 195 1
4 Mg-MOF-74 ----- 1733 73 161.73/ 158.12 278/298 1

Performance of adsorbents can be assessed by sorbents selection parameters including uptake


and selectivity under adsorption conditions, working capacity and regeneration of sorbents [129].
Table-2 shows, a group of six PCPs/MOFs identified from the discussed four categories that
display the highest CO2 sorption characteristics with no hysteresis at low pressures. Isomeric

structures from Category 1 (incorporating imidazole/benzimidazole and benzoic acid in same


ligand) such as {[Ni2(DBIBA)3]·Cl·18H2O}n and {[Co2(DBIBA)3]·Cl·9H2O}n (DBIBA = 5-

di(1H-benzo[d]imidazol-1-yl)benzoate), consisting of comparable void volume and Langmuir


surface areas have almost similar CO2 gas sorption properties with similar isosteric heat of
adsorption. It is noted from Table 2, that CO2 sorption not only depends on surface area and void

volume but is also influenced by the electronic environment, i.e., functionalization of the pores
with either of two groups coordinatively unsaturated metal ions, Lewis base groups and different
types of non-covalent interactions (π- π bonding, anion-π bonding, CH-π bonding, H-bonding,
etc.) existing within the pores. These interaction potentials in the pores decide quantitatively the
sorption of CO2 as function of temperature and pressure. Increased polarization effect due to

presence of dispersion dominated interactions, π electronic delocalization is favored towards


CO2 quadrupole. This effect becomes very prominent when there is presence of extensive

aromatic rings at the pore surface stabilizing through π-π interactions. When CO2 molecules
enter through pores these forces are shifted between aromatic and non-aromatic moieties (CO2)

with great stabilization. A dense packing is observed because there are many other weak forces
working simultaneously, helping CO2 to be adsorbed. These weak forces (quadrupole-

quadrupole) are deciding factors for selective entry as well as for removal of guest molecules.
From Figure 4c and 5b it is clear that sharp CO2 uptake is seen at ~0.19 bar or at a very low

pressure after which it gets saturated. So it can be concluded that the presence of these weak
interactions are driving forces for CO2 sorption in a PCP where surface area is not sufficient for
high uptake. Other best PCPs from Category 1 are {[Mn2(DBIBA)3]·(NO3)·3DMF·4H2O}n and
{[Mn2(DBIBA)3]·Cl·DMF·8H2O}n [DBIBAH = 3,5-di(1H-benzo[d]imidazol-1-yl)benzoic

38
2 −1
acid] which also have almost same surface areas of 780 and 800 m g respectively with high
3 −1
CO2 adsorption of 155 and 260 cm g at 195 K. Both structures are the same, having similar
type of weak forces (π-π bonding, anion-π bonding, CH-π bonding, H-bonding, etc.) at the pore
surface but interestingly show different properties. While the mechanistic reasons (electrostatic
environment gradient) for adsorption behavior are the same, the slightly increase in surface area
and dense CO2 adsorption of the latter structure is due to anion exchange (nitrate anion in the

first and chloride anion in the latter). Anion-tuned sorption [130] for selective adsorption
behavior is due to the host–guest interaction, which is dependent on the anions and the guest
size. This may be one potential route to enhance the surface area and improve sorption
characteristics should be explored further for these types of PCPs and others.
In Category 2 (consisting imidazole/benzimidazole based ligands) {[Co(HL)]DMF·H2O}
[H3L=1,3,5-tri(1H-imidazol-4-yl)benzene] is the best type of PCP regarding CO2 sorption

characteristics. It has a higher Langmuir surface area than the above mentioned PCPs. In this
case deprotonated imidazolate anions are playing a functional role through dipole-quadrupole
interactions where π electronic delocalization occur from imidazolate anions to probe molecules.
A sharp uptake is seen at the low pressure region (~0.19 bar).
Lastly Mg-MOF-74, belonging to Category 4 (incorporating only benzoic acid based ligands)
2 −1
displays the highest Langmuir surface area (1733 m g ), and comparable CO2 adsorption at
low temperature and pressure. This MOF has generally been considered as a benchmark for CO2

adsorption. It is established that the highly ionic characters of the Mg-O bond facilitate a greater
degree of polarization and higher adsorption. This is the structure from the table in which there
are exposed metal sites with high charge density which influence CO2 sorption at very low
pressure.
By finding the best PCPs/MOFs, it is clear that they all have different types of weak forces for
almost comparable gas adsorption. Only in Category 1 surface area is much lower compared to
the other PCPs/MOFs described. Conclusively it can be said that in that category dense packing
of CO2 molecules exist. It is important to take into consideration the synergistic electrostatic
effects from neighboring atoms or moieties which also influence CO2 adsorption due to the
increased polarization effect.

39
2.6 Adaptation to temperature and pressure: Suitable CO2 adsorbents for post combustion
processes should possess extraordinarily high CO2 uptake and selectivity at low temperature and
pressure. A number of weak forces present in the cavities are suitable to effect high selective
uptake of CO2 as well as desorption, irrespective of small or large surface area. Amongst
promising routes in designing a PCP with highly improved physisorption at low temperature and
pressure in a synergic fashion, is through the presence of multiple interaction sites, e.g.
hydrogen-bonding and electron-donor acceptor interactions which stabilize CO2 interaction
within the pores. In addition, to molecular confinement such as binding to more than one ligand,
restricts freedom of CO2 to diffuse out of the pores leading to higher affinity for CO2

adsorption. Mostly in PCPs, wherein open metal sites with high charge density are present for
anchoring probe molecules, dense packing of CO2 is observed at low temperature and pressure
because of large pore apertures in the framework followed by non-hysteretic desorption during
regeneration. In PCPs devoid of any interaction sites within the cavities, high uptake occurs at
high temperatures and pressures followed by non-hysteretic desorption.

3. Post synthesis modifications of PCP/MOFs for CO2 adsorption

For functionalizing PCPs/MOFs with a wide variety of different groups post synthetic
modification (PSM) is an important approach that can enhance the chemical and physical
properties of the framework. Pores in the structure can be precisely tailored through PSM, in
order to improve host guest interactions. By modifying the metal node or the organic ligand,
polar or basic substituent can be introduced. Long and coworkers [131] immobilized
ethylenediamine to unsaturated metal sites to increase CO2 uptake capacity (Fig. 24). The

modified MOF showed significant uptake of CO2 at extremely low pressure (up to 0.06 bar).
Due to crowding of ethylenediamine moieties beyond 0.1 bar uptake capacity was quite low
−1
within the pores. Heat of adsorption (Qst) was very high for the modified MOF (90 kJ mol )
−1
compared to the parent MOF (20 kJ mol ). PSM influences the breathing behavior of some
MOFs as reported by Wang and Cohen [132] where in different steps were observed in the
isotherms, attributed to structural transitions between a proposed np (narrow pore) and lp (large
pore) form of the framework.

40
Fig. 24. MOF with unsaturated metal sites (left) and saturated MOF after ethylenediamine immobilization (right).
Reproduced with permission from Ref. [131].

4. Conclusion, outlook and development needs


An important advantage over conventional adsorbents such as zeolites and activated carbons is
that pore size, structure and chemical nature of PCPs can be suitably modified to influence
adsorption. This flexibility widens opportunities to increase CO2 uptake capacity and selectivity

for PCPs in the immediate and near future. From the work by several research groups during the
period 2012 onwards, it is quite apparent that it is feasible to develop PCPs with excellent levels
of performance in the separation of CO2 from gas mixtures such as with N2, H2 and CH4. At the
same time, it is very evident from all reported approaches aiming to enhance affinity for CO2
that there is no single adsorption site approach that offers desired performances in terms of CO2
uptake and selectivity [133].
CO2 sorption and separation technologies using tunable PCP architectures have potential for
optimized use under both STP and PSA (pressure swing adsorption) operating conditions
prevalent in industry. Further, performance of PCPs under real gas mixture conditions and their
eventual use at the scale required for CO2 capture from power generation flue gases, still

presents a number of research and developmental challenges. These challenges may be broadly
divided into two categories. The first category entails the fundamental understanding of
important structure dependent physical and chemical properties affecting performance: uptake
and selectivity under adsorption conditions, working capacity, regeneration of sorbent, and
sorbent selection parameter [121]. The second category includes cost of manufacture and scales

41
of economy, PCP stability; stability toward water vapor and organic solvents, acid gases such as
SOx and NOx, H2S, and HCl, thermal regeneration, and cyclic processing, etc. These categories

require further research from a practical application point of view. While, computational gas
mixture studies can provide essential information in this regard, experimental investigations are
still considered the most reliable approach [134], even though the larger picture indicates that
rational design, development and synthesis of PCPs for target specific applications is still limited
[135–139].
As discussed earlier, PCPs represent materials with advantages such as versatility of tailoring
both architectures and properties for separation and capture. However, for real world
applications, several critical issues still need to be resolved [140]. During the last couple of
decades, though, numerous structurally differentiated PCPs have been synthesized and
characterized, only a small number have been evaluated for selective CO2 separation, sorption

and storage. While molecular simulation today plays a key role in scanning existing PCPs for
CO2 capture, by evaluating different criteria and also confirming experimental results, there are

a limited number of experimental evaluations of the effect of water on the separation


performance. The adsorption properties of MOFs can be strongly affected by water during
exposure to pre- and post-combustion flue gases due to the intrinsically weaker metal−ligand
bonds resulting in some cases in framework collapse. Water can play a relevant role in the CO2
adsorption properties of MOFs and in their capacity to discriminate CO2 from other gases.
Partial pore blockage by water upon CO2 adsorption has been reported for rigid MOFs.
Flue gas streams in real world CO2 separations always contain water and it does not make

economic sense to include additional steps to dry gas prior to separation. Thus, separation
materials must offer high tolerance to water and studies to resolve this challenge need to take
into consideration both physical co-adsorption of water in the pores of PCPs and chemisorption
of water to sites such as open metal centers. Another major challenge to overcome is co-
adsorption of CO2 and other gaseous species. In our opinion, calculations using simple mixing
theories will not work given the complex nature of flue gas streams and more rigorous
experimentation with actual gas streams, reproducibility of separation performance on long-term
exposure and response to repetitive cycles will be required. Such experimentation will help in the
evaluation of kinetic separation and advance the development of PCP-based membranes [141-

42
143] for CO2 separation which have great technological potential. Despite potential in different

CO2 capture scenarios, long-term studies focusing on the stability of membrane materials under
realistic operation conditions such as membrane activation, thermal cycles are required.
The ability to prepare PCPs in large quantities is imperative for their applications in CO2 capture

from large scale sources. As for any industrial production process, materials input costs are an
utmost critical factor for economic feasibility and viability for end use applications. PCPs are
generally prepared through hydro or solvo-thermal procedures and the total minimum cost for
preparation, include cost of reactors, chemicals, utilities, wash up and separation, and activation.
From the perspective of production, prevalent industrial process technologies or unit operations
can be adopted for their manufacture [144]. From raw materials input point of view, important
are organic linkers which are mostly aromatic compounds with structures comprising of several
aromatic rings and functional groups, including carboxylic acid, imidazole and derivatives,
hydroxyl, amine, etc. An analysis of raw material cost can provide an estimate of bare minimum
potential cost of a PCP/MOF as the raw material prices will form a major component of the total
product cost. For example, BASF has commercialized a few MOFs whose retail prices are
currently affordable for research purposes only. The raw materials cost to prepare following PCP
−1
{[M2(DBIBA)3]·Cl·9H2O}n (M = Ni, Co,) is in the range of approximately 11 US $ kg and is
a good potential for commercialization if it can be utilized in membranes. There are still many
more potential but costly organic linkers that have not been commercialized with their synthesis
limited to laboratory scale due to cost. To reduce cost of preparing such linkers at pilot and
subsequently commercial scales, it becomes imperative to use raw materials comprising of larger
fractions of aromatics and minimize use of pure reagents.
Further, the solvent exchange step which normally requires large quantities of organic solvents to
remove residual process solvents from within the pores and increase surface areas and pore
volumes involves energy intensive solvent recovery steps. The openness of the structures is
immensely useful for fast kinetics in the uptake of gases and removal of bound gases. An
alternative methodology to the solvent exchange step which may considerably reduce the use of
solvents is the supercritical drying technique followed by thermal regeneration which has been
reported to result in a larger increase in the surface areas of PCPs [145].

43
Abbreviation of
Full name of compound
compound
HIBA 4-(1H-imidazole-1-yl)benzoic acid
Hbmzbc 4-(benzimidazole-1-yl)benzoic acid
BIIBA 3-(1H-benzo[d]imidazol-1-yl)-5-(1H-imidazol-1-yl)benzoate
DBIBA 5-di(1H-benzo[d]imidazol-1-yl)benzoate
DBIBAH 3,5-di(1H-benzo[d]imidazol-1-yl)benzoic acid
H3L 1,3,5-tri(1H-imidazol-4-yl)benzene
L 1,4-di(1H-imidazol-4-yl)benzene
tib 1,3,5-tris(1-imidazolyl)benzene
H2BDC-Br 2-bromo-1,4-benzenedicarboxylate
NPBI 1,1'-(4-nitro-1,3-phenylene)bis(1H-benzo[d]imidazole)
BDC 1,3-benzenedicarboxylate
TBIB 1,3,5 tris(1H-benzo[d]imidazole-1-yl)benzene
BTC 1,3,5-benzene tri carboxylate
H3BTB 1,3,5-tris(4-carboxy-phenyl)benzene
BTN 1,3,5-tris(4-carboxy-naphthyl)benzene
BTB 4,4′,44″benzene-1,3,5-triyl-tribenzoate
BTE 4,4′,4″-[benzene-1,3,5triyl-tris(ethyne-2,1-diyl)]tribenzoate
BPDC biphenyl-4,4’-dicarboxylate

44
References
[1] Energy Information Administration, International Energy Outlook (2010). US DOE,
http://www.eia.doe.gov/oiaf/ieo/index.html.
[2] R.K. Pachauri, A. Reisinger, IPCC Fourth Assessment Report, Intergovernmental Panel on
Climate Change, (2007).
[3] S.A. Rackley, Carbon Capture and Storage, Elsevier, (2010).
[4] W.J.W. Botzen, J.M. Gowdy, J.C.J.M. Van Den Bergh, Climate Policy, 8 (2008) 569.
[5] International Energy Outlook: DOE/EIA-0484, U.S. Energy Information Administration.
http://www.eia.gov/forecasts/ieo/ index.cfm/, 2010
[6] B. Metz, O. Davidson, H. de Coninck, M. Loos, L. Meyer. IPCC, IPCC Special Report on
Carbon Dioxide Capture and Storage, Cambridge University Press, 2005.
[7] D.Y.C. Leung, G. Caramanna, M. M. Maroto-Valer, Renewable and Sustainable Energy
Reviews 39 (2014) 426.
[8] S. L'Orange Seigo, S. Dohle, M. Siegrist, Renewable and Sustainable Energy Reviews 38
(2014) 848.
[9] J. Ling, A. Ntiamoah, P. Xiao, D. Xu, P. A. Webley, Y. Zhai, Austin Chem. Eng. 1(2) (2014)
1009.
[10] B. Sreenivasulu, D.V. Gayatri, I. Sreedhar, K.V. Raghavan, Renewable and Sustainable
Energy Reviews 41 (2015) 1324.
[11] R. Pierantozzi, Carbon Dioxide, Kirk-Othmer Encyclopedia of Chemical Technology.
Wiley. doi:10.1002/0471238961.0301180216090518.a01/ 2001.
[12] IPCC Special Report on Carbon dioxide Capture and Storage,
https://www.ipcc.ch/pdf/special-reports/srccs/srccs_chapter7.pdf.
[13] N. Stafford, Nature 448 (2007) 526.
[14] J.M Austell, CO2 for Enhanced Oil Recovery Needs – Enhanced Fiscal Incentives,

Exploration & Production: the Oil & Gas Review (2005).


[15] R.G. Watts, Global Warming and the Future of the Earth, Morgan & Claypool Publishers,
Denver, (2007).
[16] J.D. Figueroa, T. Fout, S. Plasynski, H. McIlvried, R.D. Srivastava, Int. J. Greenhouse Gas

45
Control 2(1) (2008) 9.
[17] S. Freguia, G.T. Rochelle. AIChE J. 49 (2003) 1676.
[18] N.J. English, M.M. El-Hendawy, D.A. Mooney, J.M.D. MacElroy, Coord. Chem. Rev. 269
(2014) 85.
[19] S. Kitagawa, R. Kitaura, S. Noro, Angew. Chem., Int. Ed. 43 (2004) 2334.
[20] M. Eddaoudi, D.B. Moler, H. Li, B. Chen, T.M. Reineke, M. O’Keeffe, O.M. Yaghi, Acc.
Chem. Res. 34 (2001) 319.
[21] J.L.C. Rowsell, O.M. Yaghi, Microporous Mesoporous Mater. 73 (2004) 3.
[22] G. Ferey, Stud. Surf. Sci. Catal. 170 (2007) 66.
[23] R. Sabouni, H. Kazemian, S. Rohani, Environ. Sci. Pollut. Res. 21 (2014) 5427.
[24] R. Kitaura, K. Fujimoto, S. Noro, M. Kondo, S. Kitagawa, Angew. Chem. Int. Ed. 41 (2002)
133.
[25] E.J. Cussen, J.B. Claridge, M.J. Rosseinsky, C.J. Kepert, J. Am. Chem. Soc. 124 (2002)
9574.
[26] K. Uemura, S. Kitagawa, M. Kondo, K. Fukui, R. Kitaura, H.C. Chang, T. Mizutani,
Chem.-Eur. J. 8 (2002) 3586.
[27] S. Galli, N. Masciocchi, G. Tagliabue, A. Sironi, J.A.R. Navarro, J.M. Salas, L. Mendez-
Linan, M. Domingo, M. Perez-Mendoza, E. Barea, Chem.-Eur. J. 14 (2008) 9890.
[28] S.K. Makinen, N.J. Melcer, M. Parvez, G.K.H. Shimizu, Chem.-Eur. J. 7 (2001) 5176.
[29] K. Uemura, S. Kitagawa, M. Kondo, K. Fukai, R. Kitaura, H.C. Chang, T. Mizutani, Chem.-
Eur. J. 8 (2002) 3586.
[30] S. Kitagawa, M. Kondo, Bull. Chem. Soc. Jpn. 71 (1998) 1739.
[31] M. Albercht, M. Lutz, A.L. Spek, G.V. Koten, Nature 406 (2000) 970.
[32] J.A. Real, E. Andres, M.C. Munoz, M. Julve, T. Granier, A. Bousseksou, F. Varret, Science
268 (1995) 265.
[33] L.G. Beauvais, M.P. Shores, J.R. Long, J. Am. Chem. Soc. 122 (2000) 2763.
[34] M. Nihei, L. Han, H. Oshio, J. Am. Chem. Soc. 129 (2007) 5312.
[35] S.M. Neville, G.J. Halder, K.W. Chapman, M.B. Duriska, P.D. Southon, J.D. Cashion, J.-F.
Letard, B. Moubaraki, K.S. Murray, C.J. Kepert, J. Am. Chem. Soc. 130 (2008) 2869.

46
[36] D. Bradshaw, T.J. Prior, E.J. Cussen, J.B. Claridge, M.J. Rosseinsky, J. Am. Chem. Soc.
126 (2004) 6106.
[37] D.N. Dybtsev, H. Chun, S.H. Yoon, D. Kim, K. Kim, J. Am. Chem. Soc. 126 (2004) 32.
[38] R. Kitaura, K. Seki, G. Akiyama, S. Kitagawa, Angew. Chem. Int. Ed. 42 (2003) 428.
[39] J.-P. Zhang, S. Horike, S. Kitagawa, Angew. Chem. Int. Ed. 46 (2007) 889.
[40] R.A. Agarwal, P.K. Bharadwaj, Proc. Natl. Acad. Sci., India, Sect. A Phys. Sci. 84(2)
(2014) 251.
[41] R.A. Agarwal, Polyhedron 85 (2015) 740.
[42] R.A. Agarwal, P.K. Bharadwaj, Cryst. Growth Des. 14 (2014) 6115.
[43] R.A. Agarwal, Inorg. Chem. Commun. 70 (2016) 115.
[44] Y. Wang, Z. Tao, X. Yin, J. Shu, L. Chen, D. Sheng, Z. Chai, T.E. Albrecht-Schmitt, S.
Wang, Inorg. Chem. 54 (2015) 10023.
[45] D.M. D’Alessandro, T. McDonald, Pure Appl. Chem. 83(1) (2011) 57.
[46] G. Ferey, C. Serre, T. Devic, G. Maurin, H. Jobic, P.L. Llewellyn, G.D. Weireld, A.
Vimont, M. Daturif, J.-S. Chang, Chem. Soc. Rev. 40 (2011) 550.
[47] Y. Liu, Z.U. Wang, H.-C. Zhou, Greenhouse Gas Sci Technol. 2 (2012) 239.
[48] Z. Zhang, Y. Zhao, Q. Gong, Z. Li, Jing Li, Chem. Commun. 49 (2013) 653.
[49] E.J. Carrington, I.J. Vitorica-Yrezabal, L. Brammer, Acta Cryst. B70 (2014) 404.
[50] J. Liu, P.K. Thallapally, B.P. McGrail, D.R. Brown, J. Liu, Chem. Soc. Rev. 41 (2012)
2308.
[51] J.R. Long, O.M. Yaghi, Chem. Soc. Rev. 38 (2009) 1213.
[52] M. Eddaoudi, D.B. Moler, H. Li, Acc. Chem. Res. 34 (2001) 319.
[53] D.J. Tranchemontagne, J.L. Mendoza-Cortes, M. O’Keeffe, O.M. Yaghi, Chem. Soc. Rev.
38 (2009) 1257.
[54] H.L. Li, M. Eddaoudi, M. O’Keeffe, O.M. Yaghi, Nature 402 (1999) 276.
[55] J.L.C. Rowsell, E.C. Spencer, J. Eckert, J.A.K. Howard, O.M. Yaghi, Science 309 (2005)
1350.
[56] R.I. Walton, Chem. Soc. Rev. 31 (2002) 230.
[57] J. Klinowski, F.A. Almeida Paz, P. Silva, J. Rocha, Dalton Trans. 40 (2011) 321.

47
[58] N.L. Rosi, J. Eckert, M. Eddaoudi, D.T. Vodak, J. Kim, M. O’Keefe, O.M. Yaghi, Science
300 (2003) 1127.
[59] Y. Kubota, M. Takata, R. Matsuda, R. Kitaura, S. Kitagawa, K. Kato, M. Sakata, T.C.
Kobayashi, Angew. Chem. 117 (2005) 942.
[60] T. Yildirim, M.R. Hartman, Phys. Rev. Lett. 95 (2005) 215504.
[61] T. Mueller, G. Ceder, J. Phys. Chem. B, 109 (2005) 17974.
[62] H.K. Chae, D.Y. Siberio-Perez, J. Kim, Y.B. Go, M. Eddaoudi, A.J. Matzger, M. O’Keeffe,
O.M. Yaghi, Nature 427 (2004) 523.
[63] A. Torrisi, C. Mellot-Draznieks, R.G. Bell, J. Chem. Phys. 130 (2009) 194703.
[64] A. Aijaz, E. Barea, P.K. Bharadwaj Cryst. Growth Des. 9(10) (2009) 4480.
[65] S. Galli, N. Masciocchi, G. Tagliabue, A. Sironi, J.A.R. Navarro, J.M. Salas, L. Mendez-
Linan, M. Domingo, M. Perez-Mendoza, E. Barea, Chem.-Eur. J. 14 (2008) 9890.
[66] X.F. Wang, Y. Lv, T.A. Okamura, H. Kawaguchi, G. Wu, W.Y. Sun, N. Ueyama, Cryst.
Growth Des. 7 (2007) 1125.
[67] S.S. Chen, Y. Zhao, J. Fan, T.-a. Okamura, Z.S. Bai, Z.H. Chen, W.Y. Sun,
CrystEngComm. 14 (2012) 3564.
[68] W.-Y. Gao, W. Yan, R. Cai, K. Williams, A. Salas, L. Wojtas, X. Shi, S. Ma, Chem.
Commun. 48 (2012) 8898.
[69] Q. Lin, T. Wu, S.-T. Zheng, X. Bu, P. Feng, J. Am. Chem. Soc. 134 (2012) 784.
[70] R. Luebke, Ł.J. Weselin ́ski, Y. Belmabkhout, Z. Chen, L. Wojtas, M. Eddaoudi, Cryst.
Growth Des. 14 (2014) 414.
[71] Y.-L. Wang, L. Chen, C.-M. Liu, Y.-Q. Zhang, S.-G. Yin, Q.-Y. Liu. Inorg. Chem. 54
(2015) 11362.
[72] A.M. Plonka, D. Banerjee, W.R. Woerner, Z. Zhang, N. Nijem, Y.J. Chabal, J. Li, J.B.
Parise, Angew. Chem. Int. Ed. 51 (2012) 1.
[73] Z. Su, K. Cai, J. Fan, S.-S. Chen, M.-S. Chen, W.-Y. Sun, CrystEngComm. 12 (2010) 100.
[74] A. Aijaz, E.C. Sañudo, P.K. Bharadwaj, Cryst. Growth Des. 11 (2011) 1122.
[75] R.A. Agarwal, P.K. Bharadwaj, Polyhedron 85 (2015) 445.
[76] J. Li, C. Zhu, Z. Qiao, X. Chen, W. Wei, H. Ji, K. Sohlberg, Applied Surface Science

48
385 (2016) 578.
[77] R.A. Agarwal, A. Aijaz, E.C. Sañudo, Q. Xu, P.K. Bharadwaj, Cryst. Growth Des. 13
(2013) 1238.
[78] R.A. Agarwal, A. Aijaz, M. Ahmad, E.C. Sañudo, Q. Xu, P.K. Bharadwaj, Cryst. Growth
Des. 12 (2012) 2999.
[79] R.A. Agarwal, S. Mukherjee, E.C. Sañudo, S.K. Ghosh, P.K. Bharadwaj, Cryst. Growth
Des. 14 (2014) 5585.
[80] J.Y. Park,Y.S. Lee, Y. Jung, J. Nanopart. Res. 14 (2012) 793.
[81] S.-S. Chen, M. Chen, S. Takamizawa, P. Wang, G.-C. Lv, W.-Y. Sun. Chem. Commun. 47
(2011) 4902.
[82] R. Banerjee, A. Phan, B. Wang, C. Knobler, H. Furukawa, M. O’Keeffe, O.M. Yaghi,
Science 319 (2008) 939.
[83] M. Zhang, Z. Perry, J. Park, H.-C. Zhou, Polymer 55 (2014) 335.
[84] K.D. Vogiatzis, A. Mavrandonakis, W. Klopper, G.E. Froudakis, ChemPhysChem. 10
(2009) 374.
[85] S.-S. Chen, J. Fan, T.-a. Okamura, M.-S. Chen, Z. Su, W.-Y. Sun, N. Ueyama, Cryst.
Growth Des. 10(2) (2010) 812.
[86] J. Yang, J.F. Ma, Y.Y. Liu, S.R. Batten, CrystEngComm. 11 (2009) 151.
[87] J.-A. Hua, Y. Zhao, Y.-S. Kang, Y. Lu, W.-Y. Sun, Dalton Trans. 44 (2015) 11524.
[88] P. Kanoo, R. Matsuda, H. Sato, L.-C. Li, H. J. Jeon, S. Kitagawa, Inorg. Chem. 52 (2013)
10735.
[89] F. Grein, D.M. Chevrier, Theor. Chem. Acc. 131 (2012) 1110.
[90] R.A. Agarwal, S. Mukherjee, Polyhedron 106 (2016) 163.
[91] A. Torrisi, C. Mellot-Draznieks, R.G. Bell, J. Chem. Phys. 132 (2010) 044705.
[92] R.A. Agarwal, S. Mukherjee, Polyhedron 105 (2016) 228.
[93] B. Mu, F. Li, K.S. Walton, Chem. Commun. (2009) 2493.
[94] Z. Zhang, Z.-Z. Yao, S. Xiang, B. Chen, Energy Environ. Sci. 7 (2014) 2868.
[95] J. Duan, Q.Q. Li, Z. Lu, CrystEngComm. 17 (2015) 2087.
[96] Y. Cho, W.J. Cho, S. Youn, G. Lee, N.J. Singh, K.S. Kim, Acc. Chem. Res. 47 (2014)

49
3321.
[97] Z. Liang, M. Marshall, A.L. Chaffee, Energy & Fuels 23 (2009) 2785.
[98] S. Bourrelly, P.L. Llewellyn, C. Serre, F. Millange, T. Loiseau, G. Ferey, J. Am. Chem.
Soc. 127 (2005) 13519.
[99] N.L. Rosi, J. Eckert, M. Eddaoudi, D.T. Vodak, J. Kim, M. O’Keeffe, O.M. Yaghi Science
300 (2003) 1127.
[100] H. Furukawa, N. Ko, Y.B. Go, N. Aratani, S.B. Choi, E. Choi, A.Ö. Yazaydin, R.Q. Snurr,
M. O’Keeffe, J. Kim, O.M. Yaghi, Science 329 (2010) 424.
[101] H.J. Park, D.-W. Lim, W.S. Yang, T.-R. Oh, M.P. Suh. Chem. Eur. J. 17 (2011) 7251.
[102] N.J. Singh, S.K. Min, D.Y. Kim, K.S. Kim, J. Chem. Theory Comput. 5(3) (2009) 515.
[103] Z. Lu, L. Du, B. Zheng, J. Bai, M. Zhanga, R. Yun, CrystEngComm. 15 (2013) 9348.
[104] Y. Liu, Y.-P. Chen, T.-F. Liu, A.A. Yakovenko, A.M. Raiffa, H.-C. Zhou.
CrystEngComm. 15 (2013) 9688.
[105] Z. Bao, L. Yu, Q. Ren, X. Lu, S. Deng, Journal of Colloid and Interface Science 353
(2011) 549.
[106] Z. J. Zhang, Y.G. Zhao, Q.H. Gong, Z. Li, J. Li, Chem. Commun. 49 (2013) 653.
[107] K. Sumida, D.L. Rogow, J.A. Mason, T.M. McDonald, E.D. Bloch, Z.R. Herm, T.H. Bae,
J.R. Long, Chem. Rev. 112 (2012) 724.
[108] K. Poloni, K. Lee, R.F. Berger, B. Smit, J.B. Neaton, J. Phys. Chem. Lett. 5 (2014) 861.
[109] M.O. Sinnokrot, E.F. Valeev, C.D. Sherrill, J. Am. Chem. Soc. 124 (2002) 10887.
[110] J. Hwang, B.E. Dial, P. Li, M.E. Kozik, M.D. Smith, K.D. Shimizu, Chem. Sci. 6 (2015)
4358.
[111] T.C. Dinadayalane, J. Leszczynski, Struct. Chem. 20 (2009) 11.
[112] L. Chen, F. Cao, H. Sun, Int. J. Quantum Chem. 113 (2013) 2261.
[113] M. Franz, A.A. Hassan, N.G. Pinto, Carbon. 38 (2000) 1807.
[114] R.W. Coughlin, F.S. Ezra, R.N. Tan, J. Colloid Interface Sci. 28 (1968) 386.
[115] J.W.G. Bloom, S.E. Wheeler, Angew. Chem. Int. Ed. 50 (2011) 7847.
[116] S.E. Wheeler, J. Am. Chem. Soc. 133 (2011) 10262.
[117] S. Altarawneh, S. Behera, P. Jena, H.M. El-Kaderi, Chem. Commun. 50 (2014) 3571.
[118] T. Islamoglu, S. Behera, Z. Kahveci, T.-D. Tessema, P. Jena, H.M. El-Kaderi, ACS Appl.

50
Mater. Interfaces. 8 (2016) 14648.
[119] A. Torrisi, R.G. Bell, C. Mellot-Draznieks, Cryst. Growth Des. 10 (2010) 2839.
[120] H. Liu, Y. Zhao, Z. Zhang, N. Nijem, Y.J. Chabal, H. Zeng, J. Li, Adv. Funct. Mater. 21
(2011) 4754.
[121] S.J. Garibay, S.M. Cohen, Chem. Commun. 46 (2010) 7700.
[122] J.R. Ferraro, K. Nakamoto, C.W. Brown, Introductory Raman Spectroscopy, second ed.,
Academic Press, San Diego, California, 2003.
[123] A. Vimont, A. Travert, P. Bazin, J.-C. Lavalley, M. Daturi, C. Serre, G. Ferey, S.
Bourrelly, P.L. Llewellyn, Chem. Commun. (2007) 3291.
[124] N. Nijem, P. Thissen, Y. Yao, R.C. Longo, K. Roodenko, H. Wu, Y. Zhao, K. Cho, J. Li,
D.C. Langreth, Y.J. Chabal. J. Am. Chem. Soc. 133 (2011) 12849.
[125] W.L. Queen, M.R. Hudson, E.D. Bloch, J.A. Mason, M.I. Gonzalez, J.S. Lee, D. Gygi,
J.D. Howe, K. Lee, T.A. Darwish, M. James, V.K. Peterson, S.J. Teat, B. Smit, J.B. Neaton, J.R.
Long, C.M. Brown, Chem. Sci. 5 (2014) 4569.
[126] F. Gandara, T.D. Bennett, IUCrJ 1 (2014) 563.
[127] W.L. Queen, C.M. Brown, D.K. Britt, P. Zajdel, M.R. Hudson, O.M. Yaghi, J. Phys.
Chem. C 115 (2011) 24915.
[128] A.M. Plonka, D. Banerjee, W.R. Woerner, Z. Zhang, N. Nijem, Y.J. Chabal, J. Li, J.B.
Parise, Angew. Chem. Int. Ed. 52 (2013) 1692.
[129] Y.-S. Bae, R.Q. Snurr, Development and evaluation of porous materials for carbon dioxide
separation and capture. Angew. Chem., Int. Ed. 50 (2011) 11586.
[130] S. Wang, L. Li, J. Zhang, X. Yuana, C.-Y. Su, J. Mater. Chem. 21 (2011) 7098.
[131] M. Dinca, W.S. Han, Y. Liu, A. Dailly, C.M. Brown, J.R. Long, Angew. Chem., Int. Ed.
46 (2007) 1419.
[132] Z. Wang, S.M. Cohen, J. Am. Chem. Soc. 131 (2009) 16675.
[133] Y. Belmabkhout, V. Guillerm, M. Eddaoudi, Chem. Eng. J. 296 (2016) 386.
[134] M. Mohamedali, D. Nath, H. Ibrahim, A. Henni, Review of Recent Developments in CO2
Capture Using Solid Materials: Metal Organic Frameworks (MOFs) in Green House Gases.
http://dx.doi.org/10.5772/62275.

51
[135] V. Guillerm, Ł.J. Weselin ́ ski, Y. Belmabkhout, A.J. Cairns, V. D’Elia, Ł. Wojtas, K.
Adil, M. Eddaoudi, Nat. Chem. 6 (2014) 673.
[136] M. Eddaoudi, D.F. Sava, J.F. Eubank, K. Adil, V. Guillerm, Zeolite-like metalorganic
frameworks (ZMOFs): design, synthesis, and properties, Chem. Soc. Rev. 44 (2015) 228.
[137] Y. Liu, J.F. Eubank, A.J. Cairns, J. Eckert, V.C. Kravtsov, R. Luebke, M. Eddaoudi,
Angew. Chem. Int. Ed. 46 (2007) 3278.
[138] R. Luebke, J.F. Eubank, A.J. Cairns, Y. Belmabkhout, Ł. Wojtas, M. Eddaoudi, Chem.
Commun. 48 (2012) 1455.
[139] K. Sumida, D.L. Rogow, J.A. Mason, T.M. McDonald, E.D. Bloch, Z.R. Herm, T.-H. Bae,
J.R. Long, Chem. Rev. 112 (2012) 724.
[140] S. Keskin, T.M. van Heest, D.S. Sholl, ChemSusChem. 3 (2010) 879.
[141] J.-R. Li, Y. Ma, M.C. McCarthy, J. Sculley, J. Yub, H.-K. Jeong, P.B. Balbuena, H.-C.
Zhou. Coord. Chem. Rev. 255 (2011) 1791.
[142]. K. Sumida, D.L. Rogow, J.A. Mason, T.M. McDonald, E.D. Bloch, Z.R. Herm, T.-H.
Bae, and J.R. Long, Chem. Rev. 112 (2012) 724.
[143]. M. Pera-Titus Chem. Rev. 114 (2014) 1413.
[144] U. Mueller, M. Schubert, F. Teich, H. Puetter, K. Schierle-Arndt, J. Pastre, J. Mater.
Chem. 16 (2006) 626.
[145] A.P. Nelson, O.K. Farha, K.L. Mulfort, J.T. Hupp, J. Am. Chem. Soc. 131 (2009) 458.

52
CO2 sorption behaviour of imidazole, benzimidazole and benzoic
acid based coordination polymers
Rashmi A. Agarwal*, Neeraj K. Gupta
Department of Chemistry, Indian Institute of Technology Kanpur, 208016, India

Research Highlights
 PCPs designed by utilizing imidazole/benzimidazole and benzoic acid based linkers.
 CO2 sorption influenced by strong non-covalent interactions.
 These interactions include π-π, anion-π, CH-π, and H-bondings.
 Interaction potentials decide selective sorption as function of temperature & pressure.
 PCPs have good commercial potential if utilized in membranes for CO2 capture.

53

You might also like