You are on page 1of 195

MOLECULAR SIMULATIONS FOR CO

2
CAPTURE IN
METAL-ORGANIC FRAMEWORKS





CHEN YIFEI









NATIONAL UNIVERSITY OF SINGAPORE
2012


MOLECULAR SIMULATIONS FOR CO
2
CAPTURE IN
METAL-ORGANIC FRAMEWORKS




CHEN YIFEI
(B.Eng., Hebei University of Technology,
M. Eng., Tianjin University, Tianjin, China)





A THESIS SUBMITTED
FOR THE DEGREE OF DOCTOR OF ENGINEERING

DEPARTMENT OF CHEMICAL AND BIOMOLECULAR
ENGINEERING
NATIONAL UNIVERSITY OF SINGAPORE
2012
i
Acknowledgements
First of all, I would like to express my sincere gratitude to my supervisor Professor
Jiang Jianwen. His close guidance, suggestions, and discussions have helped me all
the time during my study and research at NUS. His patience and encouragement have
been of central importance for me to complete my PhD program. His immense
knowledge and enthusiasm in research have motivated me and will have substantial
impact on my future professional career.

I would like to thank my group members: Babarao Ravichandar, Hu Zhongqiao,
Anjaiah Nalaparaju, Luo Zhonglin, Fang Weijie, Zhang Liling, Liang Jianchao, Xu
Ying, Li Jianguo, Krishan Mohan Gupta, Naresh Thota, Zhang Kang and Huang
Zongjun for their interactions during my personal and professional time at NUS. I am
grateful for their suggestions, discussions, and comments on my research.

I would like to thank my family and friends for their support and encouragement. I am
also grateful to NUS for granting me the scholarship.

ii
Table of Contents
Acknowledgements ....................................................................................................... i
Table of Contents ......................................................................................................... ii
Summary ...................................................................................................................... vi
List of Tables ................................................................................................................ ix
List of Figures ............................................................................................................... x
List of Abbreviations .................................................................................................. xv
Chapter 1. Introduction ............................................................................................... 1
1.1 MOF Structures ................................................................................................ 3
1.2 MOF Synthesis ................................................................................................ 6
1.3 MOF Applications ............................................................................................ 7
1.4 Objective ........................................................................................................ 11
1.5 Thesis Outline ................................................................................................ 12
Chapter 2. Literature Review ................................................................................... 13
2.1 Experimental Studies ..................................................................................... 13
2.1.1 H
2
, CH
4
and CO
2
Storage .................................................................... 13
2.1.2 Water Adsorption ................................................................................ 18
2.1.3 Gas Separation .................................................................................... 19
2.1.4 Adsorption and Separation of Alkanes ................................................ 22
2.2 Simulation Studies ......................................................................................... 23
2.2.1 H
2
, CH
4
, CO
2
Adsorption .................................................................... 23

iii
2.2.2 Water Adsorption ................................................................................ 30
2.2.3 Gas Separation .................................................................................... 30
2.2.4 Adsorption and Separation of Alkanes ................................................ 33
Chapter 3. Models and Methods ............................................................................... 35
3.1 Atomic Models ............................................................................................... 35
3.2 Computational Methods ................................................................................. 40
3.2.1 Density Functional Theory ................................................................. 40
3.2.2 Interaction Potential ............................................................................ 41
3.2.2 Molecular Dynamics Simulation ........................................................ 43
3.2.3 Monte Carlo Simulation ...................................................................... 43
3.3 Analysis Methods ........................................................................................... 45
3.3.1 Radial Distribution Functions ............................................................. 45
3.3.2 Adsorption Selectivity ......................................................................... 46
3.3.3 Mean-Squared Displacement .............................................................. 46
Chapter 4. Adsorption of CO
2
and CH
4
in MIL-101 .............................................. 47
4.1 Models and Methods ...................................................................................... 47
4.2 Results and Discussion .................................................................................. 53
4.2.1 Sensitivity of Framework Charges ...................................................... 53
4.2.2 United-Atom and Five-site Models of CH
4
........................................ 54
4.2.3 Adsorption of Pure CO
2
and CH
4
........................................................ 55
4.2.4 Adsorption of CO
2
/CH
4
Mixture ......................................................... 63
4.3 Summary ........................................................................................................ 64

iv
Chapter 5. Adsorption and Separation in Hydrophobic Zn(BDC)(TED)
0.5
......... 66
5.1 Models and Methods ...................................................................................... 66
5.2 Results and Discussion .................................................................................. 71
5.2.1 CH
3
OH/H
2
O ........................................................................................ 71
5.2.2 CO
2
/CH
4
.............................................................................................. 75
5.2.3 Hexane ................................................................................................ 79
5.3 Summary ........................................................................................................ 81
Chapter 6. CO
2
Capture in Bio-MOF-11 ................................................................. 83
6.1 Models and Methods ...................................................................................... 83
6.2 Results and Discussion .................................................................................. 87
6.2.1 Pure Gases ........................................................................................... 87
6.2.2 CO
2
/H
2
Mixture .................................................................................. 91
6.2.3 CO
2
/N
2
Mixture .................................................................................. 94
6.3 Summary ........................................................................................................ 97
Chapter 7. CO
2
Adsorption in Cation-Exchanged MOFs ...................................... 99
7.1 Models............................................................................................................ 99
7.2 Methods........................................................................................................ 102
7.3 Results and Discussion ................................................................................ 104
7.3.1 Characterization of cations ............................................................... 105
7.3.2 Isosteric Heat and Henrys Constant ................................................. 107
7.3.3 CO
2
/H
2
Mixture ................................................................................ 112
7.4 Conclusions .................................................................................................. 115

v
Chapter 8. Ionic Liquid/MOF Composite for CO
2
Capture ................................ 116
8.1 Models and Methods .................................................................................... 116
8.1.1 [BMIM][PF
6
] .................................................................................... 116
8.1.2 IRMOF-1........................................................................................... 118
8.1.3 IL/IRMOF-1 Composite and Adsorption of CO
2
/N
2
Mixture .......... 120
8.2 Results and Discussion ................................................................................ 122
8.2.1 Structure and Dynamics of IL in IL/IRMOF-1 ................................. 122
8.2.2 Separation of CO
2
/N
2
Mixture in IL/IRMOF-1 ................................ 125
8.3 Conclusions .................................................................................................. 129
Chapter 9. Conclusions and Recommendation ..................................................... 131
9.1 Conclusions .................................................................................................. 131
9.2 Recommendation ......................................................................................... 135
References ................................................................................................................. 138
Appendix ................................................................................................................... 151
Publications .............................................................................................................. 162

vi
Summary
As a special class of hybrid nanoporous materials, metal-organic frameworks
(MOFs) have received considerable interest in the past decade. The achievable large
surface areas, high porosities, and tunable structures place them at the frontier for a
wide range of potential applications such as gas storage, separation, catalysis and drug
delivery. Since a vast variety of MOFs with different pore shapes and dimensions
have been synthesized and many more are possible, experimental screening of
appropriate MOFs for specific application is a formidable task. As an alternative,
molecular simulations can provide microscopic insights and quantitative guidelines
that otherwise are experimentally inaccessible or difficult to obtain, and thus assist in
the rational screening and design of novel MOFs. In this thesis, molecular simulations
have been performed primarily for CO
2
capture in different MOFs with diverse
structures and functionalities.
Firstly, CO
2
adsorption is investigated in a mesoporous MOF namely MIL-101,
which is one of the most porous materials reported to date. The simulation results
agree well with experimental data and the terminal water molecules play an
interesting role in adsorption. At low pressures, the terminal water molecules act as
additional adsorption site and enhance gas adsorption; however, they decrease the
available free volume and reduce adsorption at high pressures. The hydrated MIL-101
has a higher adsorption selectivity for CO
2
/CH
4
mixture.
Secondly, the adsorption and separation of CO
2
/CH
4
, as well as methanol/water,

vii
in highly hydrophobic Zn(BDC)(TED)
0.5
are examined. Good agreement is found
between simulation and experimental results and a large separation factor for
methanol/water is predicted. This reveals that Zn(BDC)(TED)
0.5
could be a good
candidate for the purification of liquid fuel. The simulation results also imply that
water has a marginal effect on CO
2
/CH
4
separation, thus pre-water treatment is not
required prior to separation.
Thirdly, CO
2
capture is investigated in bio-MOF-11 consisting of biological
ligands. The simulation results are in accordance with experimental data. The
predicted adsorption selectivities of CO
2
/H
2
and CO
2
/N
2
mixtures in bio-MOF-11 are
higher than in many porous materials, which suggests bio-MOF-11 might be
interesting for pre- and post-combustion CO
2
capture. In addition, water has a
negligible effect on the separation of these two CO
2
-containing mixtures.
Fourthly, CO
2
adsorption is simulated in rho zeolite-like MOFs (rho-ZMOFs)
exchanged with a series of cations (Na
+
, K
+
, Rb
+
, Cs
+
, Mg
2+
, Ca
2+
and Al
3+
). The
isosteric heat

and Henrys constant at infinite dilution increase monotonically with
increasing charge-to-diameter ratio of cation (Cs
+
< Rb
+
< K
+
< Na
+
< Ca
2+
< Mg
2+
<
Al
3+
). The adsorption selectivity of CO
2
/H
2
mixture increases as Cs
+
< Rb
+
< K
+
<
Na
+
< Ca
2+
< Mg
2+
~ Al
3+
. The simulation study provides microscopic insight into the
important role of cations in governing gas adsorption and separation, and suggests
that the performance of ionic rho-ZMOFs can be tailored by cations.
Finally, a new composite of ionic liquid (IL) [BMIM][PF
6
] supported on
IRMOF-1 is proposed for CO
2
capture. The confinement effects of IRMOF-1 on the

viii
structure and mobility of cation and anion are examined. Ions in the composite
interact strongly with CO
2
, particularly [PF
6
]

anion is the most favorable site for CO


2

adsorption. The composite selectively adsorbs CO
2
from CO
2
/N
2
mixture, with
selectivity significantly higher than polymer-supported ILs. In addition, the selectivity
increases with increasing IL ratio in the composite.

ix
List of Tables
Table 3.1 The structure parameters of the five MOFs. ................................................ 35
Table 4.1. Potential parameters and atomic charges of CO
2
, CH
4
, and terminal H
2
O.52
Table 5.1. Potential parameters and atomic charges.................................................... 70
Table 6.1. Atomic charges in the fragmental cluster of bio-MOF-11. ......................... 85
Table 6.2. LJ Potential Parameters and Charges for CO
2
, H
2
, N
2
, and H
2
O. .............. 86
Table 6.3. Parameters in the dual-site Langmuir-Freundlich equation fitted to the
adsorption of pure CO
2
, H
2
, and N
2
. ............................................................................ 88
Table 6.4. Selectivities and capacities for the adsorption of CO
2
/H
2
mixture in porous
materials. The capacities are for CO
2
at a total pressure of 1 bar for mixture. ............ 93
Table 6.5. Selectivities and capacities for the adsorption of CO
2
/N
2
mixture in porous
materials. The capacities are for CO
2
at a total pressure of 1 bar for mixture. ............ 96
Table 7.1. Charges Z, well depths c /k
B
and collision diameters of cations. .......... 101
Table 7.2. Lennard-Jones parameters of framework atoms in rho-ZMOF. ............... 102
Table 7.3. Porosity, isosteric heat and Henrys constant of CO
2
adsorption in
rho-ZMOFs. ............................................................................................................... 106
Table 8.1. Atomic charges in [BMIM]
+
and [PF
6
]

. .................................................. 117
Table 8.2. Simulated and experimental densities of [BMIM][PF
6
] at 1 atm. ............ 118
Table 9.1. CO
2
selectivities at ambient conditions in different MOFs. ..................... 135



x
List of Figures
Figure 1.1 Number of publications for MOFs. (Data from Scopus using metal
organic frameworks as the topic on 10 November 2011) ............................................ 2
Figure 1.2 Examples of SBUs from carboxylate MOFs. Color scheme: O, red; N,
green; C, black. In inorganic units, metal-oxygen polyhedra are blue, and the polygon
or polyhedron defined by carboxylate carbon atoms (SBUs) are red. In organic SBUs,
the polygons or polyhedrons to which linkers (all C
6
H
4
units in these examples) are
attached are shown in green.
4
......................................................................................... 3
Figure 1.3 Single crystal structures of isoreticular MOFs (IRMOF-n, n = 1 to 16).
Color code: Zn (blue polyhedra), O (red spheres), C (black spheres), Br (green
spheres in 2), amino-groups (blue spheres in 3). The large yellow spheres represent
the largest van der Waals spheres that would fit in the cavities without touching the
frameworks. All hydrogen atoms have been omitted for clarity.
5
.................................. 4
Figure 4.1. A unit cell of dehydrated MIL-101 constructed from experimental
crystallographic data, energy minimization and density functional theory calculation
(see the text for details). The pentagonal and hexagonal windows are enlarged for
clarity. Color code: Cr, orange polyhedra; F, cyan; C, blue; O, red; H, white. ............ 48
Figure 4.2. Merz-Kollman charges of Cr
3
O trimer with terminal fluorine and water
molecules in (a) dehydrated and (b) hydrated MIL-101. The cleaved bonds of Cr
3
O
(indicated by the circles) were saturated by methyl group. Color code: Cr, orange; F,
cyan; C, blue; O, red; H, white. ................................................................................... 49
Figure 4.3. Mulliken charges of the Cr
3
O trimer with terminal fluorine and water
molecules in (a) dehydrated and (b) hydrated MIL-101. The cleaved bonds of Cr
3
O
(indicated by the circles) were saturated by methyl group. Color code: Cr, orange; F,
cyan; C, blue; O, red; H, white. ................................................................................... 49
Figure 4.4. Electrostatic potential maps around the Cr
3
O trimer in (a) dehydrated and
(b) hydrated MIL-101. ................................................................................................. 51
Figure 4.5. CO
2
adsorption in dehydrated MIL-101. The squares, diamonds and
circles are experimental data
133
in MIL-101a, MIL-101b and MIL-101c, respectively.
...................................................................................................................................... 54
Figure 4.6. CH
4
adsorption in dehydrated MIL-101. The circles are the experimental
data in MIL-101c.
133
.................................................................................................... 55
Figure 4.7. Adsorption isotherms of CO
2
and CH
4
on a gravimetric basis. The squares,
diamonds, and circles are the experimental data in MIL-101a (as-synthesized),

xi
MIL-101b (activated by hot ethanol) and MIL-101c (activated by hot ethanol and KF),
respectively.
133
............................................................................................................. 56
Figure 4.8. Adsorption isotherms of CO
2
and CH
4
in MIL-101 (a) at low pressure and
(b) high pressure based on the number of molecules per unit cell. .............................. 57
Figure 4.9. Snapshots of CO
2
and CH
4
in a pentagonal window in dehydrated
MIL-101 at 10, 100, and 1000 kPa. Color code: Cr, orange; F, cyan; C, blue; O, red; H,
white; CO
2
, green; CH
4
, pink. ...................................................................................... 58
Figure 4.10. Radial distribution functions of CO
2
and CH
4
around Cr
1
and Cr
2
atoms
in dehydrated MIL-101 at 10, 100, 1000, and 5000 kPa. ............................................ 59
Figure 4.11. Schematic locations of CO
2
and CH
4
near Cr
3
O trimer in dehydrated
MIL-101. Color code: Cr, orange; F, cyan; C, blue; O, red; H, white. ........................ 60
Figure 4.12. Radial distribution functions of CO
2
and CH
4
around Cr
1
and Cr
2
atoms
in hydrated MIL-101 at 10, 100, 1000, and 5000 kPa. ................................................ 62
Figure 4.13. Radial distribution functions of CO
2
and CH
4
around the oxygen atoms
of terminal water molecules in hydrated MIL-101 at 10, 100, 1000, and 5000 kPa. .. 63
Figure 4.14. Adsorption of equimolar CO
2
/CH
4
mixture. (a) isotherm and (b)
selectivity of CO
2
over CH
4
. ........................................................................................ 64
Figure 5.1. A unit cell of Zn(BDC)(TED)
0.5
constructed from the experimental
crystallographic data and first-principles optimization. Color code: Zn, pink polyhedra;
N, green; C, blue; O, red; H, white. ............................................................................. 67
Figure 5.2. Channels along the Z, X, and Y (from top to bottom) axes in
Zn(BDC)(TED)
0.5
. The green regions denote the small windows. .............................. 68
Figure 5.3. Atomic charges in a fragmental cluster of Zn(BDC)(TED)
0.5
. The
dangling bonds (indicated by circles) were terminated by hydrogen. Color code: Zn,
pink polyhedra; N, green; C, blue; O, red; H, white. ................................................... 69
Figure 5.4. Isotherms of pure CH
3
OH and H
2
O at 303 K. The filled circles are
experimental data. The upper and lower triangles are adsorption and desorption data
from simulation. The insets show the isotherms a function of reduced pressure. The
saturation pressure P
o
is 21.7 kPa for CH
3
OH and 4.2 kPa for H
2
O. .......................... 71
Figure 5.5. Density contours of CH
3
OH at 1 kPa (top) and 10 kPa (bottom). ............ 72
Figure 5.6. Radial distribution functions of CH
3
OH around Zn, N, C2, and C4 atoms
of Zn(BDC)(TED)
0.5
at 1 and 10 kPa. ......................................................................... 74
Figure 5.7. (a) Adsorption and (b) selectivity of CH
3
OH/H
2
O mixture at 303 K. ...... 74

xii
Figure 5.8. Adsorption of pure CO
2
and CH
4
at 298 K. The open symbols are
simulation results and the filled symbols are experimental data.
327,328
........................ 76
Figure 5.9. Density contours of CO
2
at 10, 100, and 3000 kPa (from top to bottom).76
Figure 5.10. Radial distribution functions of CO
2
around Zn, N, C2, and C4 atoms of
Zn(BDC)(TED)
0.5
at 10, 100, and 3000 kPa. ............................................................... 77
Figure 5.11. (a) Adsorption and (b) selectivity of CO
2
/CH
4
equimolar mixture at 298
K. The filled symbols refer to the CO
2
/CH
4
mixture with 0.1% H
2
O. ........................ 78
Figure 5.12. Adsorption of hexane at 313 K. The open symbols are simulation results
and the filled symbols are experimental data.
150
.......................................................... 80
Figure 5.13. Density contours of hexane

at 0.001 kPa (top) and 10 kPa (bottom). .... 81
Figure 6.1. (a) Cobalt-adeninate-acetate cluster. N1 and N6 are the Lewis basic
pyrimidine and amino groups, while N3, N7, and N9 are bonded with cobalt. (b) A
unit cell of bio-MOF-11. The cavities are indicated by the green circles. Co: pink, O:
red, C: grey, H: white, N1: green, N6: blue, N3, N7, and N9: cyan. ........................... 84
Figure 6.2. A fragmental cluster of bio-MOF-11 used to calculate atomic charges. The
dangling bonds (indicated by circles) were terminated by hydrogen atoms. Color code:
Co, pink; O, red; N, cyan; C, grey; H, white. .............................................................. 85
Figure 6.3. Adsorption isotherms of pure CO
2
and N
2
at 298 K and of H
2
at 77 K,
respectively. The open symbols are from simulation and the filled symbols are from
experiment.

The lines are fits of the dual-site Langmuir-Freundlich equation to the
simulation data. ............................................................................................................ 87
Figure 6.4. Radial distribution functions of CO
2
around N1, N6, and Co atoms in
bio-MOF-11 at 298 K and 10 kPa. N1 and N6 are in the pyrimidine and amino groups,
respectively. ................................................................................................................. 89
Figure 6.5. (a) Simulation snapshot and (b) density contour of CO
2
in bio-MOF-11 at
298 K and 10 kPa. CO
2
molecules are represented by sticks. The density has a unit of
1/
3
and brighter color indicates a higher density. Co: pink, O: red, C: grey, H: white,
N1: green, N6: blue, N3, N7, and N9: cyan. ................................................................ 90
Figure 6.6. Density contours of CO
2
and H
2
for CO
2
/H
2
mixture (15:85) in
bio-MOF-11 at 298 K and 100 kPa. The density has a unit of 1/
3
. The density
distributions are largely similar to Figure 5.5b for pure CO
2
. ..................................... 91
Figure 6.7. (a) Adsorption isotherm and (b) selectivity of CO
2
/H
2
mixture (15:85) in
bio-MOF-11 as a function of total pressure in the absence and presence of 0.1 % H
2
O.
...................................................................................................................................... 92

xiii
Figure 6.8. (a) Adsorption isotherm and (b) selectivity of CO
2
/N
2
mixture (15:85) in
bio-MOF-11 as a function of total pressure. The open symbols are from simulation
and the filled symbols are from IAST. ......................................................................... 94
Figure 6.9. (a) Adsorption isotherm and (b) selectivity of CO
2
/N
2
mixture (15:85) in
bio-MOF-11 as a function of total pressure in the absence and presence of 0.1 % H
2
O.
...................................................................................................................................... 95
Figure 6.10. (a) Adsorption isotherm and (b) selectivity of CO
2
/N
2
mixture (15:85) in
bio-MOF-11 as a function of total pressure at 298 K. The open symbols are from
simulation and the filled symbols are from IAST. ....................................................... 97
Figure 7.1. Crystal structure of rho-ZMOF. Color code: In, cyan; N, blue; C, grey; O,
red; and H, white. The a-cage, double eight-membered ring (D8MR), 6-membered
ring (6MR) and 4-membered ring (4MR) are indicated. The yellow spheres in the
D8MR represent inaccessible cages. .......................................................................... 100
Figure 7.2. Atomic charges in a fragmental cluster of rho-ZMOF. Color code: In,
cyan; N, blue; C, grey; O, red; and H, white. ............................................................ 101
Figure 7.3. Equilibrium and initial locations of cations (a) K
+
(b) Ca
2+
(c) Al
3+
. The
initial locations are indicated in pink. ........................................................................ 105
Figure 7.4. Porosity | versus the packing fraction of cation q in rho-ZMOFs. The
solid line is a linear correlation between | and q. ..................................................... 107
Figure 7.5. (a) Isosteric heats and (b) Henrys constants for CO
2
adsorption in
rho-ZMOFs versus the charge-to-diameter ratio of cation. The dotted lines are to
guide the eye. ............................................................................................................. 108
Figure 7.6. Adsorption isotherms of CO
2
in rho-ZMOFs (a) low-pressure regime and
(b) high pressure regime. ........................................................................................... 109
Figure 7.7. Density contours of CO
2
in Na-rho-ZMOF at 10, 100 and 1000 kPa. The
locations of Na
+
ions are indicted by the large spheres. The density scale is the
number of CO
2
molecules per
3
. .............................................................................. 110
Figure 7.8. Radial distribution functions (a) CO
2
around Na
+
ions, N and In atoms in
Na-rho-ZMOF at 10 kPa (b) CO
2
around Na
+
ions in Na-rho-ZMOF at 10, 100 and
1000 kPa..................................................................................................................... 111
Figure 7.9. Selectivities of CO
2
/H
2
mixture in rho-ZMOFs. The composition of
CO
2
/H
2
mixture is 15/85. ........................................................................................... 113
Figure 7.10. Selectivity of CO
2
/H
2
/H
2
O mixture in rho-ZMOFs. The mole fraction of
H
2
O in the mixture is 0.1%. ....................................................................................... 114

xiv
Figure 8.1. Atomic types in [BMIM]
+
and [PF
6
]

. .................................................... 116
Figure 8.2. IRMOF-1 structure. Color code: Zn, orange; O, red; C, grey; H, white. 119
Figure 8.3. Atomic charges in a fragmental cluster of IRMOF-1. The dangling bonds
indicated by dashed circles are terminated by methyl groups. .................................. 119
Figure 8.4. Adsorption isotherm of CO
2
in IRMOF-1 at 300 K.
39
The filled symbols
are from simulation and the open symbols are from experiment............................... 120
Figure 8.5. [BMIM][PF
6
]/IRMOF-1 composite at a weight ratio W
IL/IRMOF-1
= 0.4. N:
blue, C in [BMIM]
+
: green, P: pink, F: cyan; Zn: orange, O: red, C in IRMOF-1: grey,
H: white. ..................................................................................................................... 121
Figure 8.6. Radial distribution functions of [BMIM]
+
and [PF
6
]

in IL/IRMOF-1 at
W
IL/IRMOF-1
= 0.4 and in bulk phase, respectively. The solid lines are in IL/IRMOF-1
and the dash lines are in bulk phase. .......................................................................... 122
Figure 8.7. Radial distribution functions of (a) [BMIM]
+
and (b) [PF
6
]

around O
1
, O
2
,
Zn, and C
3
atoms of IRMOF-1 at W
IL/IRMOF-1
= 0.4. .................................................. 123
Figure 8.8. Mean-squared displacements of [BMIM]
+
and [PF
6
]

in IL/IRMOF-1at
W
IL/IRMOF-1
= 0.4, 0.86 and 1.27. ................................................................................ 124
Figure 8.9. Reduced velocity correlation functions of [BMIM]
+
and [PF
6
]

in
IL/IRMOF-1 at W
IL/IRMOF-1
= 0.4, 0.86, 1.27 and 1.5. ............................................... 125
Figure 8.10. Simulation snapshot of CO
2
/N
2
mixture (P
total
= 1000 kPa) in
IL/IRMOF-1 at W
IL/IRMOF-1
= 0.4. .............................................................................. 126
Figure 8.11. Radial distribution functions of CO
2
(P
total
= 10 kPa) around Zn, N
1
, N
2
,
and P atoms in IL/IRMOF-1 at W
IL/IRMOF-1
= 0.4. ...................................................... 127
Figure 8.12. Radial distribution functions of CO
2
around P atom in IL/IRMOF-1 (a)
P
total
= 10, 100, 1000 kPa and W
IL/IRMOF-1
= 0.4, (b) P
total
= 100 kPa and W
IL/IRMOF-1
=
0.4, 0.86, 1.27 and 1.5. ............................................................................................... 128
Figure 8.13. Selectivity of CO
2
/N
2
mixture in IL/IRMOF-1 at W
IL/IRMOF-1
= 0, 0.4,
0.86, 1.27 and 1.5. ...................................................................................................... 129


xv
List of Abbreviations
5-AT 5-aminottrazole
B3LYP Becke, three-parameter, Lee-Yang-Parr
BDC benzenedicarboxylate
bpy bipyridine
bipy 4,4-bipyridine
BSSE basis set superposition error
BTC 1,3,5-benzenetricarboxylate
CCS carbon capture and sequestration
COF covalent organic frameworks
DFT density functional theory
dhtp 2,5-dihydroxyterephthalate
DMAz N,Ndimethylformamide-azine-dihydrochloride
DMF N,N-dimethylformamide
DNP double- numerical polarization
DOE Department of Energy
ESP Electrostatic Potentials
ETS-10 Engelhard Titano Silicate-10
FAU Faujasite
F-pymo 5-fluoropyrimidin-2-olate
GCMC Grand Canonical Monte Carlo

xvi
HPP 1,3,4,6,7,8-hexahydro-2H- pyrimido[1,2-a] pyrimidine
IAST ideal adsorbed solution theory
IDC imidazole-4,5-dicarboxylate
ImDC 4,5-imidazoledicarboxylate
IL ionic liquid
IRMOF isoreticular metal-organic framework
LJ Lennard-Jones
MC Monte Carlo
MD Molecular Dynamics
MFI Mobil Five
MIL Material Institute Lavoisier
MK Mera-Kollman
MAMS mesh-adjustable molecular sieves
MOF Metal-organic Frame work
MP2 second order MllerPlesset
MSD Mean Squared Displacement
MTV Multivariate
NDC 2,6-naphthalenedicarboxylate
PCN Porous Coordination Network
POM polyoxometalate-based
prz piperazine
PVDF polyvinylidene fluoride

xvii
pyz pyrazine
pzdc 2,3-pyrazinedicarboxylate
RCSR Reticular Chemistry Structure Resource
RTILs room temperature ionic liquids
SBU secondary building block
ScD supercritical drying
SILMs supported ionic liquid membranes
tbip 5-tert-butyl isophthalate
TED triethylenediamine
TIP3P Three point transferable interaction potential
TraPPE Transferrable Potentials for Phase Equlibria
UFF universal force field
ZIF zeolitic imidazolate frameworks
ZMOF zeolitic-like MOF
[BMIM] 1-n-butyl-3-methylimidazolium
[PF
6
] hexafluorophosphate
[HMIM] 1-n-hexyl-3-methylimidazolium
[TF
2
N] bis(trifluoromethylsulfony)-imide
[SCN] thiocyanate
Chapter 1. Introduction
1
Chapter 1. Introduction
Human being has been using porous materials for centuries.
1
Based on various
criteria (pore size, shape, arrangement, and chemical composition), porous materials
can be classified into different types. For example, they can be classified into three
types based on the pore size: microporous (pore size < 2 nm), mesoporous (pore size
between 2 and 50 nm), and macroporous (pore size > 50 nm). Porous materials with
pore size < 100 nm are usually termed as nanoporous materials. Traditional
nanoporous materials include inorganic (zeolites) and organic (activated carbons and
polymers). Although these materials have been utilized in many industrial processes
such as water purification, gas separation, and catalysis, they have certain limitations.
For example, highly porous activated carbons are not well ordered. On the other hand,
highly ordered zeolites lack diversity because only limited number of elements can be
used in tuning the tetrahedral building blocks.
Hybrid nanoporous materials consisting of both organic and inorganic moieties
possess unique features. They can have both highly porous and highly ordered
structures. Recently, a newly emerged class of hybrid materials named as
metal-organic frameworks (MOFs)
2
or also called porous coordination polymers
(PCPs) have attracted a great deal of attention. MOFs are crystalline structures
assembled from organic linkers and metal oxides. Compared with traditional
nanoporous materials, almost all cations can participate in MOF frameworks. In
addition, the wide variety of organic linkers and linker functionalities leads to a vast
Chapter 1. Introduction
2
diversity of MOFs. In principle, MOFs could have infinitely number of different
structures. The controllable organic linkers allow MOFs with designed functionality
and tunable pore size, surface area, and porosity. Both hydrophobic and hydrophilic
groups can be present in the frameworks, and the pores can range from microporous
to mesoporous. Therefore, MOFs are considered versatile materials for many potential
applications.
3
Over the past decade, a large number of MOFs with various topologies
and functionalities have been synthesized, and their applications in gas storage,
separation, catalysis and drug delivery have been explored. Figure 1.1 demonstrates
that the number of publications for MOFs increase rapidly in the recent years.

0
200
400
600
800
1000
1200
P
u
b
l
i
c
a
t
i
o
n

n
u
m
b
e
r
s
1998 1999 2000 2001 2002 2003 2004 2005 2006 2007 2008 2009 2010
Years

Figure 1.1 Number of publications for MOFs. (Data from Scopus using metal
organic frameworks as the topic on 10 November 2011)

In this thesis, CO
2
capture by adsorption in different MOFs is investigated. The
subsequent sections provide an overview for the structures, synthesis and typical
applications of MOFs. A more detailed literature review for the specific applications
in gas adsorption and separation will be presented in Chapter 2.
Chapter 1. Introduction
3
1.1 MOF Structures
Crystalline MOFs can be conceptually designed and constructed directly from
molecular building blocks. This route was coined as reticular synthesis by Yaghi.
4
The
molecular building blocks are linked by strong bonds and retain their structures
throughout synthesis process. The design strategy of reticular chemistry is based on
the direct expansion of secondary building units (SBUs). As shown in Figure 1.2 for
carboxylate MOFs, SBUs are the geometric units defined by the points of extension.
4




Figure 1.2 Examples of SBUs from carboxylate MOFs. Color scheme: O, red; N,
green; C, black. In inorganic units, metal-oxygen polyhedra are blue, and the polygon
or polyhedron defined by carboxylate carbon atoms (SBUs) are red. In organic SBUs,
the polygons or polyhedrons to which linkers (all C
6
H
4
units in these examples) are
attached are shown in green.
4

Chapter 1. Introduction
4
By extending different SBUs with a wide variety of organic linkers, various MOFs
can be produced. For example, Eddaoudi et al. developed a series of MOFs from the
prototype MOF-5 by using various functional organic linkers.
5
Sixteen highly
crystalline isoreticular MOFs (IRMOFs) produced are shown in Figure 1.3.



Figure 1.3 Single crystal structures of isoreticular MOFs (IRMOF-n, n = 1 to 16).
Color code: Zn (blue polyhedra), O (red spheres), C (black spheres), Br (green
spheres in 2), amino-groups (blue spheres in 3). The large yellow spheres represent
the largest van der Waals spheres that would fit in the cavities without touching the
frameworks. All hydrogen atoms have been omitted for clarity.
5


To date, tens of thousands of MOFs have been synthesized and characterized.
Based on the framework flexibility, MOFs can be categorized into rigid and flexible.
The former have rigid frameworks, largely similar to inorganic counterparts (e.g.
zeolites). In contrast, the latter can change frameworks at external stimuli like
pressure, temperature and accommodating of guest molecules.
6-8
The change may
include stretching, rotational, breathing and scissoring, and induce various effects
Chapter 1. Introduction
5
crystalline structures. Based on the framework properties, there are chiral MOFs,
9,10

magnetic MOFs,
11
luminescence MOFs.
12
The complex structures of MOFs can be
reduced to underlying nets
13
and the important nets are collected in Reticular
Chemistry Structure Resource (RCSR) database. As there are a tremendously large
number of MOFs, here we specifically introduce typical examples of MOFs, such as
IRMOFs, zeolitic imidazolate frameworks (ZIFs), covalent organic frameworks
(COFs), zeolite-like MOFs (ZMOFs), and MILs (Materials of Institut Lavoisier).
Yaghis group pioneered the development of a series IRMOFs.
5
The reticular
frameworks have a large open space up to 91.9% of crystal volume, and the free pore
diameter varies from 3.8 to 19.1 . They also synthesized ZIFs
14
and COFs.
15-17
ZIF
structures are based on the nets of aluminosilicate zeolites, in which oxygen atoms
and tetrahedral Si or Al atoms are substituted by imidazolate linkers and transition
metals, respectively. Two prototypical ZIFs (ZIF-8 and ZIF-11) exhibit permanent
porosity and high thermal/chemical stability.
14
COFs consist of light elements (B, C,
N, O) via strong covalent bonds. The pores in COFs can run in 2D and 3D with a size
ranging from 6.4 to 34.1 . Because of the unique structures, COFs exhibit high
thermal stability, permanent porosity, low density, and high surface area. ZMOFs have
similar topologies and structural properties to inorganic zeolites.
18
However, the
difference is oxygen atoms in zeolites are substituted by organic linkers, leading to
extra-large cavities and pores in ZMOFs. This edge expansion approach offers a great
potential towards the design and synthesis of widely open materials. In addition, some
ZMOFs possess ionic frameworks and contain charge-balancing nonframework ions.
Chapter 1. Introduction
6
For example, rho-ZMOF contains 48 extraframework ions per unit cell to neutralize
the anionic framework.
18

Ferey and co-workers synthesized a series of 3D rare earth diphosphonates named
as MIL-n.
19-21
They also extended to compounds containing transition metals (V, Fe,
Ti) and metallic dicarboxylates.
22-24
The first synthesized MIL-53(Cr) can exist in two
forms, one is filled with water molecules at low temperatures and the other is
dehydrated at high temperatures.
25
The transition between the hydrated and
dehydrated crystals is fully reversible and considered as breathing effect. Similar
breathing effect occurs when Cr metal is replaced by Al, Fe and Ga. This is due to the
presence of OH groups in one-dimensional channels that have strong interactions with
water molecules.
26-28
In MIL-47(V), however, no breathing occurs because of the
absence of OH groups in the skeleton.
29
Ferey et al. also synthesized chromium
terephthalate-based mesoscopic MIL-101,
30
which is one of the most porous materials.
It is stable in air or boiling water and its structure is not altered in various organic
solvents or solvothermal conditions.
1.2 MOF Synthesis
MOFs are usually synthesized by self-assembly at a low temperature (below
300 ) using organic or inorganic solvent without additional template. The
traditional synthesis methods include classical coordination chemistry and
solvothermal syntheses. In the traditional synthesis, temperature is crucial because it
can change the properties of solution and hence the dimension and structure of a MOF.
In addition, pH value, solution concentration and the chemical nature of cations can
Chapter 1. Introduction
7
also influence crystal structure. Furthermore, the nature of initial metallic salts and
precursors also play an important role in synthesis.
3

In recent years, new methods have been proposed including (1) hydrothermal
synthesis using immiscible solvents (2) electrochemical synthesis (3) microwave
synthesis
3
(4) sonication synthesis
31,32
(5) mechanochemical method
33,34
and (6)
high-throughput method. These new methods are regarded as environmentally
friendly and can significantly reduce reaction time. Their combination can also been
used for MOF synthesis.
35
In addition, the way to activate MOF samples is crucial to
determine the final crystal structure. Hupps group developed supercritical drying
(ScD) method, in which supercritical carbon dioxide is used to increase the accessible
surface area of MOF samples.
36

1.3 MOF Applications
MOF are potential candidates in many important areas such as gas storage,
5,37-39

separation,
40-43
catalysis,
44,45
sensing,
46
drug storage and delivery,
47-49
templates for
new materials synthesis,
50
luminescent and fluorescent materials,
51
magnetic
materials,
52
proton conductors.
53,54
Several reviews have summarized the potential
applications of MOFs.
3,55,56
A brief introduction is presented here for the applications
of MOFs in storage, separation, and catalysis. More detailed discussions in these areas
are described in Chapter 2.
Storage
Gas storage in porous materials has become increasingly important as the
growing concerns for energy and environment. The most extensively studied gases in
Chapter 1. Introduction
8
storage are H
2
and CH
4
for clean energy as well as CO
2
for environmental protection.
Particularly, H
2
is regarded as an ideal energy carrier due to the absence of carbon and
zero emission. The development of safe, efficient and high-capacity storage system is
a key step for the practical utilization of H
2
. The U.S. Department of Energy (DOE)
has set the targets for H
2
storage as of 7.5 wt% or 70 g/L for the ultimate Full Fleet
target.
57

With high porosity and large surface area, MOFs have attracted considerable
attention for gas storage. Rosi et al. first measured H
2
adsorption in MOF-5,
IRMOF-6 and IRMOF-8 and showed that MOFs have much larger capacity for H
2

storage than traditional zeolites and active carbons.
37
After this first experimental
measurement, numerous studies have been reported for H
2
storage in MOFs over the
past few years, as summarized in several reviews.
55,58-61
CH
4
is the major component
of nature gas and an alternative fuel to fossil fuels. The storage target set by the U.S.
DOE is 180 v/v at 35 bar (the volume of gas adsorbed at standard temperature and
pressure per volume of the storage vessel). Kitagawa and coworkers reported the first
CH
4
adsorption in M
2
(4,4-bpy)
3
(NO
3
)
4
](H
2
O)
x
(M = Co, Ni, and Zn).
62
CH
4
capacity
in this MOF is about 71 v/v anhydrous sample at 30 atm. Dren et al.
63
and Wang
64

used simulations to examine CH
4
storage in various MOFs. Furukawa et al.
investigated the adsorption of H
2
, CH
4
and CO
2
in seven COFs.
65
Among many
reported studies, MOF-200 and MOF-210 exhibit the highest adsorption capacity for
CO
2
(64.32 and 65.23 mmol/g, respectively, at 298 K and 50 bar).
66

As a biologically important gas, NO storage been examined in MOFs. Morriss
Chapter 1. Introduction
9
group measured the adsorption, storage and delivery of NO in MOFs with accessible
open metal sites.
67,68
The porous MOFs showed exceptionally good performance for
the adsorption and water-triggered delivery of NO. In addition, drug storage and
delivery in MOFs have also been reported. Fereys group determined the adsorption
and delivery of ibuprofen in MIL-101
47
and MIL-53.
48
The loading was found to be
1.38 g/g in MIL-101 and 0.22 g/g in MIL-53. An et al. reported cation-triggered
procainamide release in a bio-MOF,
69
which was constructed from biocompatible
linkers. They found that the drug was complete released after 72 hours and the
framework maintained crystalline structure in the whole process. The biomedical
applications in MOFs were recently summarized by Keskin et al.
70

Separation
Porous materials are commonly used in industry adsorbents for gas separation. A
suitable material should have large capacity and high selectivity. MOFs have the
potential for gas separation due to the tunable pore sizes and surface properties.
Numerous studies have been reported for gas separation in different MOFs. As
demonstrated in a breakthrough experiment, CO
2
can be separated completely from
CO
2
/CH
4
mixture in Mg-MOF-74.
71
ZIF-68, ZIF-69, and ZIF-70 exhibit large
capacity for CO
2
and unusual selective adsorption for CO
2
/CO mixture.
72
In a
breakthrough experiment for CO
2
/CO mixture, the complete retention of CO
2
and
passage of CO were observed. Such a high selectivity is based on the difference
quadruple moments of CO
2
and CO. ZIF-95 and ZIF-100 have the capability to
efficiently separate CO
2
from CH
4
, CO and N
2
.
73
This is attributed to the combined
Chapter 1. Introduction
10
effects of the appropriate aperture size and the strong quadrupolar interaction of CO
2

with the N atoms on the framework surface.
Important factors influencing separation efficiency include the size/shape of gas,
and the interaction of gas with framework. With regard to the second factor, different
strategies have been proposed to improve separation by tailoring MOF structures,
such as the addition of open metal sites and functionalized groups. Recently, Zhou
and coworkers reviewed the selective adsorption and separation in MOFs.
74,75

Catalysis
With high metal contents, well-defined pores and narrow pore distributions,
MOFs have potential application in heterogeneous catalysis. Fujita et al. first reported
MOF-based catalyst for the cyanosilylation of aldehydes and imines.
76
Several studies
investigated the catalytic properties of MOFs with chiral porous.
10,77-80
The catalytic
properties of common MOFs such as MOF-5,
81,82
HKUST-1,
44,83
MIL-101
45,84,85
were
examined. It was found that the catalytic activity of MOF-5 is attributed to the
encapsulated zinc-hydroxide clusters or to the hydrolytically degraded form of the
parent framework. MIL-101 has a stronger catalytic activity than HKUST-1 for the
cyanosilylation of benzaldehyde because of the greater Lewis acidity of Cr(III) vs.
Cu(II). The modified MIL-101(Cr) was tested for its catalytic activity for
Knoevenagel condensation of benzaldehyde with nitriles.
85

Hasegawa et al.
86
synthesized a MOF and found it was able to catalyze the
Knoevenagel condensation reaction due to its selective heterogeneous base catalytic
properties. Alkordi and Eddaoudi et al.
87
reported the catalytic properties of
Chapter 1. Introduction
11
rho-ZMOF with cationic porphyrins encapsulated in the framework. The encapsulated
porphyrin with Mn metallated showed catalytic activity towards the oxidation of
cyclohexane. Jiang et al.
88
synthesized a Cu-MOF that has comparable catalytic
activity for the ring-opening reactions of epoxides in the presence of alcohols and
aniline under ambient, solvent-free conditions. A comprehensive review for MOFs
used in catalysis was recently presented by Corma et al.
89

1.4 Objective
As a new class of nanoporous materials, MOFs are considered versatile
candidates for a wide variety of important applications. However, a very large number
of MOFs have been synthesized and many more are possible, experimental exploring
the performance of MOFs for specific application is not economically feasible. With
ever-increasing computational power, molecular simulations have become
increasingly important in materials science. In particular, microscopic insights gained
from simulations are indispensable in quantitatively elucidating the underlying
physics and subsequently provide guidelines for the rational design of new materials.
The objective of this thesis is to investigate CO
2
capture using molecular
simulations in five unique MOFs with different pore sizes and functionalities. Firstly,
adsorption of CO
2
and CH
4
is simulated in mesoporous MIL-101 and compared with
experiment. The effect of the coordinated water molecules on adsorption is carefully
examined. Secondly, adsorption and separation of CO
2
/CH
4
as well as CH
3
OH/H
2
O
are examined in highly hydrophobic Zn(BDC)(TED)
0.5
. Thirdly, CO
2
capture from
syngas and flue gas is studied in a bio-metal organic framework. The effect of
Chapter 1. Introduction
12
humidity on separation is estimated. Fourthly, CO
2
adsorption and CO
2
/H
2
separation
in cations exchanged rho-ZMOFs are explored to study the influence of cations.
Finally, CO
2
separation from flue gas in an ionic-liquid/MOF composite is
investigated. The effect of ionic liquid loading on separation is examined in detail.
1.5 Thesis Outline
There are eight chapters in this thesis and the outline is as follows. In Chapter 1,
the general background of MOFs is introduced, such as the structures, synthesis, and
applications of MOFs. Chapter 2 reviews the current state of both experimental and
simulation studies for adsorption and separation in MOFs, including gas storage and
separation, water adsorption, and adsorption of alkanes. In Chapter 3, the adsorption
of CO
2
and CH
4
in MIL-101 is examined. In Chapter 4, the separation of CH
3
OH/H
2
O
and CO
2
/CH
4
in highly hydrophobic MOF Zn(BDC)(TED)
0.5
are investigated.
Chapter 5 describes the CO
2
capture in bio-MOF-11. Chapter 6 discusses CO
2

adsorption in different cation-exchanged rho-ZMOFs. CO
2
capture in an ionic
liquid/metal-organic framework composite is proposed in Chapter 7. Finally, the
conclusions and recommendation for future work are presented in Chapter 8.

Chapter 2. Literature Review
13
Chapter 2. Literature Review
MOFs are considered versatile materials for a wide rang of applications.
Nevertheless, most current studies have been focused on gas storage and separation.
In this Chapter, a literature review is provided for the corresponding experimental and
simulation studies.
2.1 Experimental Studies
2.1.1 H
2
, CH
4
and CO
2
Storage
H
2
Storage
H
2
is regarded as an ideal energy carrier and its storage in MOFs has been
extensively studied. The first experimental measurement of H
2
storage was conducted
in IRMOF-1, IRMOF-6, and IRMOF-8.
37
In this study, the length of organic link was
found to play a dominate role in tuning H
2
capacity. Therefore, H
2
uptake in
IRMOF-6 and -8 are much higher than IRMOF-1. Compared with inorganic cluster,
organic linker has a lower binding energy. However, it governs the free volume and
surface area, thus determines the capacity for H
2
storage. The predominant effect of
organic linkers on H
2
storage capacity was also demonstrated by Rowsell et al.
90,91
In
MIL-53(Al and Cr), Ferey et al. found H
2
storage capacities are 3.8 wt% and 3.1 wt%
respectively at 77 K and 1.6 MPa.
38
Zhao et al. prepared three MOFs with flexible
linkers and compared their H
2
capacities with active carbons at a supercritical
temperature of H
2
.
92
Surprisingly, a hysteresis of H
2
adsorption was observed,
Chapter 2. Literature Review
14
implying the windows opening of the frameworks upon adsorption. Chen et al.
synthesized MOF-505 with open metal sites and found H
2
adsorption in the fully
activated MOF-505 is 2.47 wt% at 77 K and 750 torr.
93
In a MOF that has physical
characteristic similar to single walled carbon nanotubes but a stronger interaction with
H
2
, Pan et al. concluded that H
2
capacity is affected not only by pore volume but also
by pore quality, and the ideal storage materials should have the largest possible pore
volume and the proper pore which can fit with H
2
molecules.
94

Rowsell et al. reported several strategies to enhance H
2
storage in MOFs, such as
linker modification to optimize pore size and adsorption energy, impregnation,
catenation, and including of open metal sites and lighter metals.
95
Following this, the
effects of catenation and open metal sites on H
2
uptake have been extensively studied.
For example, Yaghi and coworkers
96,97
measured H
2
storage capacities in various
MOFs and investigated the effects of functionalization, catenation, metal oxide cluster,
and organic linker on H
2
adsorption at low pressures. The saturation capacities in
MOF-177 and IRMOF-20 at 77 K and below 80 bar were determined to be 7.5 wt%
and 6.7 wt%, respectively, which had achieved the U.S. DOE target for H
2
storage in
2010 (6wt%). Dinca et al. synthesized MOFs with exposed coordination sites (Mn
2+
,
Li
+
, Cu
+
, Fe
+
, Co
2+
, Ni
2+
, Cu
2+
, Zn
2+
) leading to high H
2
storage capacity ranging from
2.00 to 2.29 wt% at 77 K and 900 torr.
98,99
Particularly, the MOF with exposed Co
2+

exhibits a high isosteric heat (10.5 kJ/mol) because of the strong interaction between
H
2
and unsaturated Co
2+
. Ma et al. showed that both catenation and unsaturated
metals in PCN-6 can enhance H
2
uptake.
100
In PCN-10 and PCN-11 containing open
Chapter 2. Literature Review
15
metal sites, the measured H
2
uptakes at 77 K and 760 torr were 2.34 wt% and 2.55
wt%, respectively.
101
With a high porosity and metal sites, PCN-12 exhibits high H
2
uptake of 3.05 wt% at 77 K and 1 bar.
102
Schroders group synthesized a series of
MOFs, which show high H
2
storage capacity because of the exposed metal sites and
high surface areas.
103-106
CPO-27-Ni with open metal sites was found to possess the
highest isosteric heat for H
2
(-13.5 kJ/mol).
107
Dinca and Long have reviewed the
experimental studies for H
2
adsorption in various MOFs with open metal sites.
59

In a separate study, a microporous copper MOF with polar network and narrow
pores was observed to exhibit the highest H
2
capacity (3.07 wt%) at 77 K and ambient
pressure.
108
H
2
capacity can be also enhanced in MOFs doped with alkali metals (Li
+
,
Na
+
, and K
+
)
109,110
and modified by ligands.
111
New technique such as spillover has
been proposed to enhance H
2
capacity in MOFs.
112-114
Using this technique, the
capacities in MOF-5, IRMOF-8, and MOF-177 were enhanced by 3.3, 3.1, and 2.5
times, respectively. The capacity in IRMOF-8 was further increased to eight times
higher than that in pure IRMOF-8 using hydrogen spillover with bridges. The effect
of MOF structures on H
2
storage by spillover has been investigated in detail.
115
In
addition, MOF composites incorporated with carbon nanotube were examined for H
2

storage
116,117
More comprehensive review on H
2
uptake in MOFs can refer to a recent
critical report by Long and co-workers.
118

CH
4
Storage
CH
4
is also considered a clean energy carrier like H
2
. Although CH
4
storage in
MOFs is far less studied, it is still an interesting research field. After the first
Chapter 2. Literature Review
16
experimental measurement,
62
Eddaoudi et al. synthesized a series of IRMOFs and
tested for CH
4
storage.
5
Based on the prototype IRMOF-1, they functionalized or
modified the organic linkers and synthesized 16 highly crystalline MOFs with open
space up to 91.1% and pore size from 3.8 to 28.8 . One member of this series
exhibits a high capacity for CH
4
storage of 155 v/ at 36 atm and ambient temperature.
Kim measured CH
4
adsorption in Zn(BDC)(TED)
0.5
up to 35 bar over temperature
range from 198 to 296 K. The adsorption capacity was 137 v/v at 35 bar and 296 K,
and three adsorption sites were indentified.
119
Among CuBTC, Zn(BDC)(TED)
0.5
and
MIL-101(Cr),
120
Kaskel and coworkers found CuBTC has the highest excess CH
4

adsorption of 228 v/v at 150 bar and 303 K. Wu et al. studied CH
4
storage in five
MOFs (M
2
(dhtp), M = Mg, Mn, Co, Ni, Zn) with open metal sites.
121
These MOFs
have large capacity ranging from 149 v/v to 190v/v at 298 K and 35 bar. Particularly,
Ni
2
(dhtp) shows the largest capacity of 200 v/v. They also found that the primary
adsorption sites are the open metal sites. In another study for CH
4
adsorption in
MOF-5 and ZIF-8, they found the primary adsorption sites in ZIF-8 are associated
with the organic linkers, in contrast to the metal oxide clusters in MOF-5.
122

Zhou and coworkers synthesized a series of PCNs and tested for CH
4
storage. In
PCN-10 and PCN-11 with unsaturated metal sites, PCN-11 exhibits an excess CH
4

uptake of 171 v/v at 298 K and 35 bar, approaching the U.S. DOE target.
101
In
microporous PCN-14 based on anthracene derivative consisting of nanoscopic cages,
the excess CH
4
capacity was reported to be 230 v/v (excess 220 v/v).
123
This is the
highest CH
4
capacity reported to date, higher than the U.S. DOE target at 290 K and
Chapter 2. Literature Review
17
35 bar, and also higher than in MOF-200, -205 and -210 with ultrahigh porosity.
66

CO
2
Storage
CO
2
emissions have caused detrimental effects on environment such as global
warming, sea-level rise, and an irreversible increase in the acidity level of oceans. The
removal of CO
2
is thus practically important and MOFs have been extensively studied
for their application for CO
2
storage.
Yaghis group first reported CO
2
adsorption in MOF-2 to characterize the
microporosity.
124
Later, they examined CO
2
storage capacity in nine MOFs at ambient
temperature up to 42 bar.
125
These selected MOFs represented different characteristics
such as square channels (MOF-2), open metal sites (MOF-505 and Cu
3
(BTC)
2
),
hexagonally packed cylindrical channels (MOF-74), interpenetrated (IRMOF-11),
amino- and alkyl-functionalized (IRMOF-3 and -6), and highly porous (IRMOF-1 and
MOF-177). Because of the high porosity, MOF-177 shows a higher CO
2
capacity than
other MOFs studied. At 35 bar, CO
2
loading in MOF-177 is 9 times higher than the
pressurized CO
2
. They also synthesized a series of ZIFs with high thermal and
chemical stability, and selective adsorption for CO
2
.
72
Furthermore, they measured
CO
2
adsorption in various 1D, 2D and 3D COFs and found 3D COFs have a better
performance.
16,17,65
In MOFs of ultrahigh porosity they synthesized, MOF-210
exhibits the highest CO
2
storage capacity of 2870 mg/g.
66

CO
2
capacity can be enhanced in functionalized MOFs. It was suggested that
varying amine substituent in the frameworks would affect CO
2
adsorption.
126-131

Recently, multivariate (MTV) MOF-5 frameworks decorated with different functional
Chapter 2. Literature Review
18
groups were synthesized.
132
The properties of MTV-MOFs with mixed functional
groups in one pore were found to outperform the simple combination of their
constituents. In MTV-MOF-5-EIH, CO
2
capacity was determined to be approximately
4 times higher than MOF-5.
In a number of MILs produced by Ferey and coworkers, CO
2
adsorption has
been examined. Specifically, Llewellyn et al. measured MIL-101 has a high CO
2

capacity of 40 mmol/g (i.e. 390 v/v) at 303 K and 5 MPa.
133
Similarly, Chowdhury et
al. determined CO
2
, CH
4
, C
3
H
8
, SF
6
and Ar adsorption in MIL-101 at 283, 319 and
351 K.
134
The preferential adsorption sites were found to be the bare metal sites inside
supertetrahedra cages. Strong interaction was also found between CO
2
and the open
metal sites in MOF-74 that has an exceptionally high CO
2
capacity.
135,136
In addition,
CO
2
adsorption in MOFs has also been investigated in the presence of water.
137,138

2.1.2 Water Adsorption
Studies have shown that a number of MOFs are not stable in water.
114,139-142

Understanding the properties of water in MOFs is crucial to identify and design
water-resistant MOFs for technological applications, e.g., waste water treatment and
biofuel purification. Wang et al. showed that Cu-BTC exhibits a high H
2
O adsorption
capacity and a reversible color change upon H
2
O adsorption.
143
In three MOFs,
Kondo found the adsorption isotherms of H
2
O are all type .
144
This reveals H
2
O
adsorbs strongly on the hydrophilic sites of the three MOFs, which correspond to
crystalline water. Kitagawa and coworkers reported H
2
O and methanol adsorption in a
dynamic microporous MOF with 1D hydrophilic channels.
145
In a 3D porous MOF
Chapter 2. Literature Review
19
namely [Zn
6
(IDC)
4
(OH)
2
(Hprz)
2
]
n
, Gu et al. observed the selective adsorption of H
2
O
over organic solvents and the reversible formation of MOF channels upon H
2
O
adsorption/desorption.
146
Similarly, Barcia et al. found the selective adsorption of H
2
O
over methanol in Cu(R-GLA-Me)(4,4-bipy)
0.5
due to size/shape discrimination.
147

However, the selective adsorption of alcohols over H
2
O was observed in some MOFs
Cu(hfipbb)(H
2
hfipbb)
0.5
,
148
Zn(tbip),
149
Zn(BDC)(TED)
0.5
,
150
and ZIF-71.
151

2.1.3 Gas Separation
Gas separation in MOFs has been widely studied, particularly for the mixtures of
light gases (CO
2
, H
2
, N
2
, CH
4
, etc). For example, size- or shape-selective adsorption
has been observed in several MOFs. Dybtsev et al. synthesized a MOF with high
thermal stability and permanent porosity, which selectively adsorbs H
2
and CO
2
over
N
2
and other gases of large kinetic diameters.
152
This is attributed to the small
apertures that block the adsorption of large molecules. Similarly, selective uptake of
H
2
and O
2
over N
2
and CO was reported in the first magnesium-based MOF
Mg
3
(NDC)
3
.
153
It was suggested that Mg
3
(NDC)
3
could be potentially used for N
2

separation from air, H
2
/CO separation for fuel cell, and enrichment of H
2
from
ammonia synthesis. Selective adsorption of H
2
and O
2
over N
2
and CO due to size
exclusion was also reported in PCN-13
154
and PCN-17.
155
In addition, Cu(F-pymo)
2
156

and pillar-layer MOFs
157
were found to selectively adsorb H
2
over N
2
. The
mesh-adjustable molecular sieves (MAMS) possessing infinite numbers of mesh sizes
were proposed to have the ability to separate any two gases with different kinetic
diameters.
158,159
This is because the mesh size increases linearly with temperature;
Chapter 2. Literature Review
20
therefore, different gases can be separated by tuning temperature. For example, H
2

exhibits a higher uptake over CO, N
2
and O
2
at 77 K; however, O
2
has a much higher
uptake than CO and N
2
at 87 K. Further increasing temperature, N
2
is selectively
adsorbed over CO and CH
4
at 113 K. CH
4
can also be separated from C
2
H
4
at 143 K.
In MIL-96, CO
2
was found to has a shorter adsorption equilibrium time than CH
4

because of the small apertures in MIL-96.
158
CO
2
selective adsorption over CH
4
was
also found in PCN-5.
159
Recently, two MOFs with the same static aperture size but
different effective aperture sizes were reported to exhibit different separation
properties for N
2
/Ar.
160

Besides the size-/shape-based selective adsorption, interaction between gas and
MOF is also an important factor to determine separation selectivity. The interaction is
mainly governed by the nature of adsorbate and adsorbent. Both the polarity and
quadrupole moment of adsorbate and the functional group in MOF can affect the
interaction. In addition, the proper pore size may increase the interaction in the pore.
It was found Cu
2
(pzdc)
2
(pyz) selectively adsorbs C
2
H
2
over CO
2
.
161
The reason is that
the H atoms of C
2
H
2
and the non-coordinated O atoms in the framework form
hydrogen bonds, thus C
2
H
2
binds more strongly than CO
2
. Because of the different
gas-framework interactions, CO
2
and H
2
were found to be selectively adsorbed over
N
2
in an interdigitated 3D MOF.
162

MOFs with open metal sites can enhance gas separation. Britt et al. performed
dynamic separation experiment to measure the dynamic capacity of CO
2
in
Mg-MOF-74 replete with open metal sites.
71
Their breakthrough curves demonstrated
Chapter 2. Literature Review
21
that CO
2
can be separated completely from CH
4
. In addition, the isosteric heat of CO
2

in Mg-MOF-74 is moderate, suggesting the energy for regeneration is not high.
Mn(NDC) with open metal sites exhibits a larger capacity for CO
2
than CH
4
at
ambient temperature.
163
Snurr and coworkers reported the separation of CO
2
/CH
4
in a
carborane-based MOF with and without open metal sites, and observed a higher
selectivity in the former.
164
A 2D interpenetrating MOF with unsaturated metal sites
and uncoordinated carboxylic group was found to have high CO
2
/CH
4
selectivity.
165

Furthermore, Li-doped MOFs also show enhanced CO
2
/CH
4
selectivity.
166

Another strategy to improve gas separation is to functionalize MOFs. At 298 K
and 1 bar, the selectivity of CO
2
/N
2
in Cu-BTTri increases from 21 to 25 by
functionalizing the framework with ethylenediamine.
129
A 3D porous MOF with
tetrazole functionalized aromatic carboxylic acid was found to exhibit high selectivity
for CO
2
/CH
4
at 195, 273, and 298 K.
167
With both amino and pyrimidine groups
presented in the framework, bio-MOF-11 exhibits high CO
2
/N
2
selectivity of 81 at
273 K and 75 at 298 K.
168
Banerjee et al. synthesized a series of ZIFs and the
predicted selectivity of CO
2
/N
2
was in the range between 17 and 50.
169
Particularly,
ZIF-78 was found to have a higher selectivity because the presence of NO
2
enhances
interaction between CO
2
and framework. Post-modified MOFs with polar group CF
3

were also observed to increase the selectivity of CO
2
/N
2
.
170
In a rht-type MOF
decorated with acylamide (CONH), Zheng et al. determined the selectivity of
CO
2
/N
2
is 22 at 1 bar and 33 at 20 bar.
171
The selectivity is enhanced upon
comparison with the non-decorated framework PCN-61. A MOF functionalized with
Chapter 2. Literature Review
22
flexible alkyl ether chains was found to have high selective sorption of CO
2
over N
2

due to the gate effect of the side chains.
172

2.1.4 Adsorption and Separation of Alkanes
Despite numerous studies reported for the adsorption and separation of light
gases in MOFs, there are few studies for alkanes. Pan et al. first experimentally
reported the separation of hydrocarbons in microporous MOFs built from
paddle-wheel metal clusters and dicarboxylate ligands.
148
These MOFs possess
irregular-shaped microchannels with alternating large cages connected by small
windows, and the internal surfaces are highly hydrophobic. Selective adsorption in
these MOFs was observed, specifically nC
2
, C
3
and C
4
olefins and alkanes can be
separated from branched alkanes, and n-C
4
from higher alkanes and olefins. Chen et
al. examined the chromatographic separation of linear and branched isomers of
pentane and hexane in microporous MOF-508 based on size- and shape-selectivity.
173

The pores in MOF-508 may be tuned to match the sizes of alkanes. Later, their group
presented that Zn(BDC)(TED)
0.5
with two types of intersecting pores can kinetically
separate hexane isomers by fixed-bed adsorption.
174

Adsorption of a series of alkanes, along with alcohols and aromatics, was
determined in a multifunctional microporous MOF and the results indicated that the
MOF can be used for the separation of components with similar boiling point.
175

Adsorption and diffusion of alkanes in Cu-BTC were studied by infrared microscopy,
and strong inflection in isotherms was observed due to the preferential location close
to the mouth of octahedral pockets.
176
Luebbers et al. reported the adsorption of more
Chapter 2. Literature Review
23
than 30 volatile organic compounds in IRMOF-1 to examine the effect of structural
degradation on adsorption.
177
Recently, gas phase separation of ethane and ethene in
ZIF-7 was conducted and the results showed that gate-opening effect controls the
separation.
178
Adsorption equilibria (as well as diffusivities) of ethane, ethylene,
propane, and propylene in Mg-MOF-74 were experimentally determined. The
relatively high adsorption selectivities suggested that this MOF could be potentially
used to separate C
2
H
4
/C
2
H
6
, C
3
H
6
/C
3
H
8
, and C
3
H
6
/C
2
H
4
.
179

Trung et al. studied adsorption of alkanes in flexible MIL-53(Al, Cr).
180
They
showed that the metal type and length of n-alkanes govern the shape of adsorption
curve. Llewellyn et al. investigated the adsorption of short linear alkanes in flexible
MIL-53(Fe) and observed multisteps in adsorption isotherms.
181
This is because there
are four discrete pore openings in the whole adsorption process. They also performed
simulation and found various structural states of flexible MIL-53(Fe) in the presence
of short linear alkanes.
181
By combining experiment and simulation, very recently
they found the adsorption process of n-alkanes in functionalized MIL-53(Fe)-X (X =
CH
3
, Cl, Br, NH
2
) is different from that in the original MIL-53(Fe).
182

2.2 Simulation Studies
2.2.1 H
2
, CH
4
, CO
2
Adsorption
H
2
Adsorption
Numerous simulation studies have been reported for H
2
adsorption in various
MOFs. Kawakami et al. reported in 2001 the first simulation study of H
2
in
Zn(BDC)(H
2
O) and showed H
2
has higher adsorption than Ar and N
2
at their boiling

Chapter 2. Literature Review
24
temperatures.
183
Combining quantum chemical calculation and molecular simulation,
Sagara et al. examined H
2
adsorption in IRMOFs.
184-186
The zinc-oxide corners were
identified to have a higher binding energy compare with the organic linkers. The
binding energy increases with increasing size of organic linker or by adding NH
2
or
CH
3
group. In addition, they proposed new IRMOFs with a high H
2
storage capacity
of 6.5 wt% at room temperature. Yang and Zhong simulated H
2
adsorption in
IRMOF-1, IRMOF-8 and IRMOF-18 and concluded that metal-oxide clusters are the
preferential adsorption sites.
187
In IRMOF-1, Zhang et al. suggested that the
adsorption sites are mostly determined by adsorbate-MOF interaction at a low
temperature; as temperature increases, however, the loading and diffusion ability of
adsorbed molecules would influence.
188
Zhou et al. investigated H
2
adsorption in
ZIF-8 at 77 K and pressure range from 10 to 8000 kPa.
189
They identified the first
adsorption site is located at the both sides of imidazolate ring and close to imidazolate
C=C bond, and the secondary adsorption site is the pore channel. With ab initio
derived force field, Han et al. simulated H
2
adsorption in 10 ZIFs and found the
adsorption sites are diverse with ZIF structures.
190

At cryogenic temperatures, quantum effects might play a substantially important
role in determining H
2
adsorption. To address this, Garberoglio et al. predicted H
2

adsorption isotherms in various MOFs using classical and modified potential.
191
From
both experiment and simulation, Liu et al. examined H
2
adsorption in
Zn(BDC)(TED)
0.5
at 77 K and 298 K up to 50 bar.
192
The simulation results agree
well with experimental data when quantum effects were included. Zhong and
Chapter 2. Literature Review
25
coworkers reported the influences of pore size, pressure and temperature on quantum
effects.
193
Their results suggested that quantum effects increase with decreasing pore
size and are mainly determined by gas-gas and gas-MOFs interactions.
H
2
adsorption in MOFs is governed by many factors. Frost et al. discussed the
effects of isosteric heat, surface area, and free volume.
194,195
The required isosteric
heat for H
2
storage to meet the U.S. DOE target was predicted to be 10~15 kJ/mol
with a free volume of 1.6~2.4 cm
3
/g and larger than 20 kJ/mol with a lower free
volume. They also proposed correlations for H
2
uptake with isosteric heat, surface
area and free volume at 0.1, 30 and 120 bar, respectively. Assfour and Seifert
concluded that H
2
uptake at 77 K is mainly determined by isosteric heat at low
pressures and by free volume at high pressures, while H
2
uptake at 300 K is only
related to free volume.
196
The effect of catenation on H
2
adsorption was investigated
by Jung et al.
197
and catenated MOFs were found to exhibit a higher capacity because
of the stronger potential overlap generated by catenation. Similar catenation effect
was also reported for improving H
2
storage at cryogenic temperatures.
198
H
2
uptake
was found to increase at low pressures by the inclusion of C
60
into MOF-177 at 77
and 300 K; however, the uptake at high pressures decreases due to the reduced pore
volume.
199

The binding energy of H
2
in MOFs can be enhanced by open metal sites. As
several theoretical studies indicated, it can be tuned up to 10-50 kJ/mol by using
different transition metals in MOFs.
200-202
A new set of potential parameters used to
accurate describe H
2
adsorption in MOFs with open metal site were derived from ab
Chapter 2. Literature Review
26
initio calculations.
203
On the other hand, doping alkali elements on the organic linkers
of MOFs is another strategy to improve H
2
uptake.
204-213
IRMOF-14 composed of
-SO
3
-1
group and Li cation was predicted to have a higher H
2
binding energy and
hence a higher H
2
capacity at ambient conditions.
214
A simulation study of H
2

adsorption in metal alkoxide (Li, Mg, Mn, Ni, Cu) modified MOFs (IRMOF-1, -10,
-16, UiO-68, and UMCM-150) revealed that magnesium alkoxide might be promising
to enhance H
2
storage at room temperature.
215
In ionic rht-MOF balanced by NO
3
-

ions, high H
2
adsorption capacity was predicted (2.4 wt% at 1 bar and 6.2 wt% at 50
bar).
216

CH
4
Adsorption
In 2004, Dren et al. simulated CH
4
adsorption of in several IRMOFs and
compared with traditional zeolites, MCM-41, carbon nanotubes, and molecular
squares.
63
A complex interplay was found for CH
4
adsorption with surface area, free
volume, interaction strength, and pore size. Later, they determined the effects of
organic linker on CH
4
adsorption.
217
Based on simulated CH
4
isotherms in nine
IRMOFs, they correlated the adsorption amount to isosteric heat, surface area and free
volume at low, medium, and high loadings, respectively.
218
Similar correlations were
also proposed by Wang from the simulation of CH
4
in 10 MOFs and it was suggested
that the desired MOF for CH
4
storage should have high isosteric heat, surface area,
free volume, and low density of framework.
64

Different strategies such as MOFs with functionalized groups, open metal sites
and catenation have been proposed to enhance CH
4
storage. Jhon et al. simulated CH
4

Chapter 2. Literature Review
27
adsorption as well as diffusion in alkoxy-functionalized IRMOFs.
219
They found that
functionalization increases adsorption at low and moderate pressures (< 50 bar at 298
K), but reduces saturation capacity and diffusion. Based on IRMOF-1, Zeng and
Zhang designed 10 MOFs with different functional groups for CH
4
storage.
220
NO
2

functionalized MOFs were demonstrated to have the highest adsorption at 298 K and
3.5 MPa. Specifically, the capacity of CH
4
in a MOF substituted by four NO
2
groups
was predicted to be 228 v/v. From simulation on CH
4
adsorption in six heterometallic
MOFs (MOF-Fe/AgBF
4
-1, MOF-Fe/AgPF
6
-1, MOF-Fe/AgSb
6
-1, MOF-Co/AgBF
4
-1,
MOF-Co/AgPF
6
-1, MOF-Co/AgSb
6
-1), Liu et al. suggested that anions are the
preferential adsorption sites for CH
4
at low pressure.
221
In five MOFs with large
surface area (MOF-5 and MOF-177), catenation (IRMOF-11 and MOF-14), and open
metal sites (MOF-74), Gallo et al. showed that CH
4
storage capacity in MOF-74 is
170 v/v at 35 atm and 298 K, close to the U.S. DOE target.
222
Thornton predicted CH
4

adsorption in a MOF with magnesium-decorated fullerenes would reach 265 v/v,
higher than the DOE target.
223
Zhong and coworkers designed a new MOF by
modifying the organic linkers of PCN-14 for CH
4
storage.
224
The new MOF was
predicted to have a very large CH
4
capacity of 241 v/v at 298 K and 3.5 MPa, which
is 34% higher than the U.S. DOE target. They concluded that the rational
modification of organic linkers is a feasible way to improve CH
4
storage capacity.
They also suggested that catenation
225
and pore topology
226
are important factors to
enhance CH
4
adsorption capacity.
CO
2
Adsorption
Chapter 2. Literature Review
28
CO
2
adsorption in MOFs has been also studied using molecular simulations.
Babarao et al. compared CO
2
adsorption in IRMOF-1, silicalite, and C
168

schwarzite.
41
The simulated adsorption isotherms match well with experimental data,
and CO
2
has a higher adsorption capacity in IRMOF-1 than the other two materials.
Later, they simulated CO
2
storage in series of IRMOFs, UMCM-1, and COFs.
39,227

CO
2
affinity was found to be enhanced by functional groups or catenation. The
saturation CO
2
capacities were correlated with free volume, porosity, surface area and
framework density. CO
2
adsorption in Zn(BDC)(H
2
O) was theoretically studied by
Kawakami et al.
183
Although the simulation overestimated experiments, the tendency
was correct. They attributed the deviations to the possible unsaturated adsorption in
experimental samples.
Good agreement was found by Snurr and coworkers between simulation and
experiment for CO
2
adsorption in IRMOF-1 at different temperatures.
228
It was
suggested that electrostatic interactions between CO
2
and frameworks are responsible
for the observed stepped isotherms. They also simulated CO
2
adsorption in MOF-177
and IRMOF-3 and the results agreed well with experimental data. Later, they
examined the effect of coordinated water molecules on CO
2
adsorption in Cu-BTC,
and the presence of coordinated water molecules was found to increase CO
2
uptake
and the selectivity over N
2
and CH
4
.
137
As electrostatic interactions are important to
mimic CO
2
adsorption in MOFs, the calculation of atomic charges for MOF atoms is
always needed. However, quantum chemical calculation is time consuming. In order
to improve the efficiency, they introduced a rapid charge equilibration method and
Chapter 2. Literature Review
29
demonstrated its accuracy for CO
2
adsorption in 14 different MOFs.
229

Zhong and coworkers predicted CO
2
adsorption in various MOFs to evaluate the
effects of organic linker, pore size, and topology.
230
They correlated CO
2
uptake with
isosteric heat, free volume, and surface area; suggested that CO
2
-framework
electrostatic interactions would enhance adsorption amount. They also performed
simulation to investigate CO
2
, CH
4
and H
2
adsorption in 2D COFs.
231
As attributed to
CO
2
-CO
2
electrostatic interactions, stepped isotherms were observed and affected by
temperature, pore size, and CO
2
-COF interactions. From simulation in ZIF-68 and
ZIF-69, they found CO
2
is preferred to adsorb in the small pores formed by
2-nitroimidazole linkers and the chlorine atoms in ZIF-69 increase the binding energy
between CO
2
and framework.
232
Later, they computationally examined the influences
of framework charges on CO
2
adsorption in 20 MOFs with different topologies, pore
sizes, and chemical characteristics.
233
The relationships between framework charges
with pressure and pore size were identified.
Ramsahye et al. explored the breathing effect of MIL-53 upon CO
2
adsorption
from simulation.
234,235
They also used first-principles calculation to probe different
adsorption sites for CO
2
in MIL-53(Al, Cr) and MIL-47(V).
236
By functionalizing
MIL-53(Al) with different groups -OH, -COOH, -NH
2
, and -CH
3
, Torrise et al.
examined the impact of these groups on CO
2
adsorption at low pressures and room
temperature.
237
It was suggested that -(OH)
2
functionalized MIL-53 was an optimal
material for CO
2
capture at low pressures.
Chapter 2. Literature Review
30
2.2.2 Water Adsorption
Although the studies of water adsorption in MOFs are few, it is important for the
rational design of water-resistant MOFs and their applications in aqueous media.
Greathouse and Allendorf reported a molecular dynamics simulation study for water
in MOF-5.
238
It was found MOF-5 structure collapsed upon the water content larger
than 3.9%. By fitting the charges of framework atoms to reproduce the experimental
data of water adsorption in Cu-BTC, Castillo et al. showed that water has a strong
affinity with the open metal sites and water adsorption is sensitive to the atomic
charges.
239
Water adsorption on the open metal sites in Cu-BTC was further explored
by DFT calculations and the predicted adsorption enthalpy agreed well with
experimental data using DFT/CC correction scheme.
240,241
Nalaparaju et al.
investigated water adsorption, diffusion and vibration in cation-exchanged
rho-ZMOFs and revealed the important interplay between water and nonframework
ions.
242
Combining simulation and experimental methods, Maurin and coworkers
examined water adsorption and diffusion in MIL-53(Cr).
243
The results suggested that
MIL-53(Cr) is mildly hydrophobic in the large pore form and hydrophilic in the
narrow pore form. Thus MIL-53(Cr) can be tuned to hydrophobic or hydrophilic
without functionalization.
2.2.3 Gas Separation
Gas separation based on adsorption in MOFs has been extensively studied using
molecular simulations. Dren and Snurr simulated the adsorption of CH
4
/nC
4
H
10

mixture in IRMOFs to determined the influence of organic linkers on adsorption and
Chapter 2. Literature Review
31
separation.
217
Furthermore, they proposed a hypothetical IRMOF with a larger organic
linker, which exhibits a higher selectivity for C
4
H
10
. In a mixed-ligand MOF, Bae et al.
predicted the adsorption isotherms of CO
2
/CH
4
using ideal-adsorbed solution theory
(IAST), which compared well with simulation results.
244
Martin-Calvo et al.
examined the adsorption and separation of CO
2
, CH
4
and N
2
in IRMOF-1 and
Cu-BTC.
245
They found the adsorption selectivity of CO
2
over CH
4
and N
2
in
Cu-BTC was higher than in IRMOF-1, suggesting the efficiency largely depends on
the shape, composition and linker. Combining experiment and simulation, Yazaydin et
al. observed that CO
2
selectivity over N
2
and CH
4
in CuBTC increases in the presence
of 4 wt% of H
2
O.
137

Babarao et al. simulated the adsorption and separation of CO
2
and CH
4
in
IRMOF-1, silicalite, and C
168
schwarzite.
41
The predicted adsorption isotherms and
isosteric heats of pure gases agree well with available experimental data, and the
predictions of mixture adsorption from the IAST are in accordance with simulation
results. As the functionality of MOFs can be readily modified by tuning metal-oxide
and organic linker, Babarao et al. further explored a variety of MOFs with unique
characteristics such as open metal sites, interpenetration, and ionic framework for
CO
2
/CH
4
separation.
246
The selectivity was found to increase slightly in the presence
of open metal sites and interpenetration, but significantly in ionic framework. Jiang
and coworkers systematically investigated CO
2
/H
2
separation in ionic MOFs
(soc-MOF, rht-MOF and rho-ZMOF).
42,216,247
The predicted CO
2
/H
2
selectivity is
significantly higher than in non-ionic MOFs. By switching off the charges on the
Chapter 2. Literature Review
32
framework and ions, the selectivity was found to decrease substantially. This indicates
the important role of electrostatic interactions in the observed high selectivity in ionic
MOFs. Particularly noteworthy is 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.
Zhong and coworkers showed the separation selectivity of CO
2
/CH
4
/C
2
H
6
mixture in manganese-formate MOFs is higher than in IRMOF-1, Cu-BTC, and most
carbon and zeolite materials
248
and the selectivity can be enhanced when electrostatic
interactions are included.
249
They also simulated in Cu-BTC for the separation of flue
gas (CO
2
/N
2
/O
2
),
250
CH
4
/H
2
, CH
4
/N
2
, N
2
/H
2
, CO
2
/H
2
, CO
2
/N
2
, CO
2
/CH
4
,
251
and
CO
2
/CO, C
2
H
4
/CO
2
, and C
2
H
4
/CO mixtures.
252
The side pockets of Cu-BTC were
found to have positive contributions to mixture separation and Cu-BTC was suggested
to be useful for the purification and capture of CO
2
. By comparing the separation of
CH
4
/CO
2
/H
2
mixture in COFs and MOFs, they concluded COFs and MOFs have
similar separation performance.
253
They predicted a higher CO
2
/CH
4
selectivity in
metal-modified
254,255
and organic-group functionalized
256
MOF-5 from the prototype
MOF-5. Combining experimental and simulation methods, they examined the effect
of temperature on the separation of CO
2
, CH
4
, CO, and N
2
binary mixtures in
ZIF-8.
257
Furthermore, they computationally screened 105 MOFs for CH
4
/H
2

Chapter 2. Literature Review
33
separation in pressure swing adsorption based on high working capacity and
selectivity.
258

Liu et al. investigated the effect of interpenetration on CH
4
/H
2
separation in
MOFs and observed a higher selectivity in interpenetrated frameworks than the
non-interpenetrated counterparts due to the enhanced potential overlap in the
former.
259
Liu and Smit simulated the separation of CO
2
/N
2
and CH
4
/N
2
mixtures in
three zeolites (MFI, LTA and DDR) and seven MOFs (Cu-BTC, MIL-47, IRMOF-1,
-12, -14, -11 and -13). The results showed that MOFs are better for gas storage but
their separation performance is comparable to zeolites.
260
By matching experimental
data with adjustable force field parameters, they examined the separation of CO
2
/N
2
,
CO
2
/CH
4
, and CH
4
/N
2
in ZIF-68 and -69. In the presence of Cl atoms, ZIF-69 was
found to exhibit a higher selectivity.
261
. Recently, they investigated the effects of
interpenetration and mixed-ligand on CH
4
/H
2
separation. Similar with their previous
study, interpenetration in mixed-ligand MOFs was found to enhance adsorption
selectivity.
262

2.2.4 Adsorption and Separation of Alkanes
Jiang and Sandler conducted a simulation study for pure and mixed linear and
branched alkanes in IRMOF-1.
263
A linear alkane is preferentially adsorbed over a
shorter one at low pressures, whereas the reverse is found at high pressures. In
addition, a linear isomer adsorbs more than its branched analogue as a result of
configurational entropy. This study also indicates that the extent of alkane adsorption
in IRMOF-1 is larger than in silicalite and carbon nanotube. Later, Jiang and
Chapter 2. Literature Review
34
coworkers simulated the adsorption and diffusion of C
4
and C
5
isomer mixtures in
catenated and noncatenated MOFs.
43
The catenated MOFs showed a larger separation
factor compared with the the non-catenated counterparts. From the adsorption of a
mixture of 13 alkanes in Zn(BDC)(TED)
0.5
,
264
Dubbeldam et al. observed the
segregation of alkanes according to the degree of branching and suggested a new
possibility for the design MOFs. For propane, n-butane, n-pentane and n-hexane in
cobalt formate frameworks, Krishna and van Baten revealed the possibility for
separating C
3
/nC
6
, C
3
/nC
4
, nC
4
/nC
6
and nC
4
/nC
5
mixtures because the adsorbed phase
contains primarily short alkanes.
265
Jorge et al. presented simulation for the adsorption
of pure and mixed propane and propylene in Cu-BTC.
266
The predicted selectivity for
propylene over propane is approximately 4. Maurin and coworkers examined the
adsorption and diffusion of long n-alkanes (C
5
-C
9
) in MIL-47(V) using UFF-Dreiding
force fields, and the simulated isotherms and enthalpies agree fairly well with
experimental data.
267



Chapter 3. Models and Methods
35
Chapter 3. Models and Methods
3.1 Atomic Models
In this thesis, five MOFs with unique structures were selected to study CO
2

capture, namely MIL-101, Zn(BDC)(TED)
0.5
, bio-MOF-11, cations exchanged
rho-ZMOF, and IL/IRMOF-1 composite. The objective is to determine the factors that
influence CO
2
capture and assist in future materials design. Table 3.1 listed the
structure parameters of the five MOFs.

Table 3.1 The structure parameters of the five MOFs.

MOF Unit Cell () Cell Angle
(degree)
Pore Volume
(cm
3
/g)
Pore Size

()
Surface
Area (m
2
/g)
MIL-101
133
a=b=c=
88.869
===90 2.15 12, 16 (29,
34)
4230
Zn(BDC)(TED)
0.5
150,268
a=b=10.9288,
c=9.6084
===90 0.65 3.2 (7.5) 1450
Bio-MOF-11
168
a=b=15.4355,
c=22.775
===90 0.45 5.2 (5.8) 1040
rho-ZMOF
18
a=b=c=
31.062
===90 0.48 5.6 (18.2) 1067
IRMOF-1
5
a=b=c=
25.832
===90 1.55 11.2 (18.6) 3800

MIL-101
MIL-101 is a chromium terephthalate-based mesoscopic MOF and one of the
most porous materials reported to date. It is stable under air atmosphere or boiling
water and its structure is not altered in organic solvents or solvothermal conditions.
These remarkable properties and its highly porous structure have resulted in
Chapter 3. Models and Methods
36
considerable interest in MIL-101 for gas storage, catalysis, drug delivery, etc. Despite
a number of experimental studies in this unique MOF,
133,134,269-273
we are not aware of
molecular simulation study reported for gas adsorption in MIL-101. However, such
molecular information is important to understand the microscopic adsorption behavior
and facilitate the development of new super-adsorbents.
Zn(BDC)(TED)
0.5

A novel MOF named Zn(BDC)(TED)
0.5
(BDC = benzenedicarboxylate, TED =
triethylenediamine) was recently synthesized and characterized.
150,268
It is thermally
stable up to 282 C, readily activated and highly porous. H
2
sorption capacity in
Zn(BDC)(TED)
0.5
is 2.1 wt% at 78 K and 1 bar. In addition, Zn(BDC)(TED)
0.5
is
highly hydrophobic and adsorbs a large amount of hydrocarbons giving the highest
capacities among all MOFs reported to date.
150
Attributed to these salient features,
several studies have been carried out in Zn(BDC)(TED)
0.5
. Adsorption and diffusion
of H
2
were examined using both experimental and simulation methods.
192
Segregation
of complex alkane mixtures were observed from simulation, which suggested a new
possibility for the design and creation of highly selective adsorption sites in MOFs.
264

Adsorption of hexane isomers was studied and compared with experimental results.
265

Diffusion of binary mixtures was studied in Zn(BDC)(TED)
0.5
, other MOFs and
zeolites by a unified Maxwell-Stefan description.
274

Bio-MOF-11
A unique set of bio-MOFs based on adenine have been developed recently by
Rosi and coworkers.
79,195
As a natural nitrogen-based heterocycle, adenine consists of
Chapter 3. Models and Methods
37
four imino nitrogen atoms and one exocyclic amino nitrogen atom. Interestingly, all
the five nitrogen atoms can coordinate with metals. Adenine possesses multiple
binding modes and is an ideal biomolecular building block for bio-MOFs. Among
several synthesized bio-MOFs, bio-MOF-1 can serve as a host for adsorbing a
cationic drug,
69
bio-MOF-11 shows exceptional ability to adsorb CO
2
,
168
a
zinc-adeninate macrocycle has large cavities for gas sorption.
275
For further
development of new bio-MOFs as ideal sorbents for CO
2
capture, a fundamental
understanding of their adsorption properties from a molecular level is crucial.
rho-ZMOF
In the continuous quest for novel MOFs to achieve high-performance gas
adsorption and separation, ionic MOFs with charge-balancing nonframework ions
have been produced. The presence of nonframework ions in a nano-confined space
can enhance the interactions between adsorbate and framework, which in turn
increases adsorption capacities. Recently, Eddaoudi and coworkers synthesized rho
zeolite-like MOF (ZMOF),
18
which is topologically relevant to rho-zeolite. The
substitution of oxygen atom in rho-zeolite by 4,5-imidazoledicarboxylic acid
(H3ImDC) leads to extra-large cavities. Our group has performed systematic
simulation studies on gas adsorption and separation in several ionic MOFs (soc-MOF,
rho-ZMOF, rht-MOF and Li-MOF).
42,216,246,247,276
Good agreement was observed
between simulated and available experimental data. The nonframework ions in an
ionic MOF can be exchanged. For example, as-synthesized rho-ZMOF contains
doubly protonated 1,3,4,6,7,8-hexahydro-2H- pyrimido[1,2-a]pyrimidine (HPP). The
Chapter 3. Models and Methods
38
HPP cations in parent rho-ZMOF can be fully exchanged with other cations such as
Na
+
ions.
18
However, it remains unknown how different nonframework ions would
affect gas adsorption and separation.
IL/IRMOF composite
There has been considerable interest in using ionic liquids (ILs) for CO
2

capture.
277
As a unique class of green solvents, ILs are nonvolatile and nonflammable
with high thermal stability, and thus might potentially substitute traditional
solvents.
278,279
Since the first observation of high CO
2
dissolution in ILs,
280
a large
number of experimental and theoretical studies have examined CO
2
behavior in
various ILs.
281-284
For example, Shiflett and Yokozeki measured the solubilities and
diffusivities of CO
2
in 1-n-butyl-3-methylimidazolium hexafluorophosphate
[BMIM][PF
6
] and 1-n-butyl-3-methylimidazolium tetrafluoroborate [BMIM][BF
4
].
285

Bara et al. synthesized imidazolium-based ILs with one, two, or three oligo(ethylene
glycol) substituents and determined the solubilities of CO
2
, N
2
, and CH
4
.
286
Compared
to corresponding alkyl analogues, these ILs were observed to have 30 75% higher
ideal solubility selectivities for CO
2
/N
2
and CO
2
/CH
4
mixtures. By measuring CO
2

solubility in a range of ILs, Muldoon et al. found ILs containing more fluoroalkyl
chains on either cation or anion improve CO
2
solubility when compared to less
fluorinated ILs.
287
Shi and Maginn simulated the sorption of CO
2
and gas mixtures in
1-n-hexyl-3-methylimidazolium bis(trifluoromethylsulfonyl)-imide [HMIM]
[TF
2
N].
288,289
Quantum chemical COSMO-RS method was applied by
Gonzalez-Miquel et al. to predict CO
2
/N
2
selectivity in 224 ILs and they found
Chapter 3. Models and Methods
39
[SCN]

-based ILs can enhance selective separation.


290
On the other hand,
task-specific ILs have been investigated for CO
2
capture, e.g. by functionalizing
ILs
291
or tuning the basicity of ILs.
292

There are two major issues in the use of ILs for CO
2
capture, i.e., the cost and
high viscosity of ILs. As a new strategy, supported ILs (SILs) have been recently
proposed in which ILs impregnate the pores of a solid porous support.
293
In addition
to enhanced mechanical strength, SILs reduce the amount needed and viscosity of ILs,
and thus increase separation efficiency. Scovazzo et al. first tested polyethersulfone
supported non-hexafluorophosphate ILs for gas separation and found the
ideal-selectivity versus permeability ratios in these SILs were above the Robeson
upper-bound for polymers.
294
Brennecke and coworkers demonstrated the
high-temperature separation of CO
2
/H
2
mixture using an amine-functionalized IL
encapsulated on a cross-linked nylon support.
295
Using both hydrophilic and
hydrophobic polyvinylidene fluoride (PVDF) as supports, Neves et al. measured gas
permeation in 1-n-alkyl-3-methylimidazolium based ILs.
296
Their results show that
the hydrophobic PVDF-SILs are more stable and have a higher affinity for CO
2
.
Most studies to date have used either organic polymers or inorganic zeolites as
supports to prepare SILs.
297,298
Although numerous experimental and simulation
studies have been reported separately for ILs
281-284
and MOFs,
74,299,300
we are not
aware of any study for MOF-supported ILs.
Chapter 3. Models and Methods
40
3.2 Computational Methods
3.2.1 Density Functional Theory
In quantum chemistry, many approximate methods have been developed to solve
the Schrodinger equation with a low computational cost. Density functional theory
(DFT) is a robust method that is much more efficient than ab initio wave functional
methods. DFT is based on the theorems by Hohenberg and Kohn (1964) who proved
the ground state properties are directly related with electron density. For a system with
N electrons, the total ground state energy can be expressed as

ee ee
E T V U T V U = u + + u = u u + u u + u u (3.1)
where u is the electron wavefunction,

T is the kinetic energy,

V is the energy
from external fields, and

ee
U is electron-electron interaction energy. For a
many-body electronic system, U
ee
can be written as
1 ( ) ( ')
'
2 '
ee xc
r r
U drdr E
r r

= +

}
(3.2)
where E
xc
is the exchange-correlation energy and (r) is the electronic density.
E
xc
remains unknown and must be approximated. Local-density approximation
(LDA) is a simple widely used approximation based on exact exchange energy for a
uniform electron gas. In contrast, generalized gradient approximation (GGA)
considers the inhomogeneous electron gas in a natural molecular system. Commonly
used GGA functionals include BP, PW91, BLYP, and PBE. Hybrid functionals (e.g.
B3LYP) which combines the gradient-corrected functionals with exact HF-type
exchange is also a commonly used functional. The electron density (r) is calculated
Chapter 3. Models and Methods
41
from molecular orbitals mathematically expressed by basis set. A basis set determines
the number and types of atomic orbitals used in the expansion of molecular orbitals.
There are numerous basis sets, for example, the minimal basis set STO-3G,
split-valence basis set 6-31G(p), and correlation-consistent basis set cc-pVTZ.
3.2.2 Interaction Potential
In classical simulation, interaction potential of a system is usually determined by
empirical force field without solving the Schrodinger equation. The interaction
potential can be subdivided into two contributions

total bonded non bonded
U U U

= + (3.3)
where
bonded
U is the intramolecular energy and
non bonded
U

is the intermolecular energy.
The intramolecular energy
bonded
U consists of

stretching bonded bend torsion
U U U U = + + (3.4)
where
stretching
U is the stretching energy arising from the change of bond lengths,
bend
U is the bending energy resulting from the change of angles between two
successive chemical bonds, and
torsion
U is the torsional energy arising from the change
of dihedral angles.
The non-bonded intermolecular energy has

non bonded vdW Coulombic
U U U

= + (3.5)
where
vdW
U is the van der Waals interaction energy and
Coulombic
U is the coulombic
electrostatic potential energy. The van der Waals interaction is usually mimicked by
Lennard-Jones (LJ) potential. For a system composed of different types of atoms, the
total LJ potential is
Chapter 3. Models and Methods
42

12 6
.
( ) 4
ij ij
LJ ij
i j
ij ij
i j
u r
r r
o o
c
<
(
| | | |
(
=
| |
| |
(
\ . \ .

(3.6)
where
ij
and
ij
are the collision diameter and well depth, respectively; The cross
parameters
ij
and
ij
were evaluated by combining rule, for example, Lorentz-
Berthelot rules

ij i j
= c c c (3.7)

ij
= (
i
+
j
)/2. (3.8)
The electrostatic interaction is modeled by the Coulombs law

0
,
( )
4
i j
Coulomb
i j
ij
i j
q q
u r
r tc
<
=

(3.9)
where q
i
and q
j
are the charges on particle i and j,
0
= 8.8542 10
-12
C
2
N
-1
m
-2
is the
permittivity of vacuum. The charges are usually estimated by fitting electrostatic
potential (ESP),
301
which can be calculated by DFT.
Different force fields have been developed for simulations. For example,
Universal Force Field (UFF)
302
is a general force field can be used for all the elements
in the periodic table. Dreiding force field
303
can be used for organic, biological and
main-group inorganic systems. PCFF
304
is developed for polymers and organic
materials. COMPASS
305
is a force filed to predict both gas- and condensed-phase
properties for most common organics, polymers and small inorganic molecules.
AMBER
306
is a force field widely used for nucleic acids and proteins. CVFF
307
,
CHARMM
308
, OPLS-AA
309
and GROMOS
310
are also common force fields for
organic and biomolecular systems.
Chapter 3. Models and Methods
43
3.2.2 Molecular Dynamics Simulation
Molecular dynamics (MD) simulation is a technique to compute equilibrium and
transport properties in a system in which the motion of particles obeys the classical
mechanics laws. The transport properties such as mass transfer (diffusion coefficients),
heat transfer (thermal conductivity) or momentum transfer (viscosity) can be
calculated from MD simulation performed at equilibrium ensembles (NVE, NVT,
NPT) or non-equilibrium ensembles (when perturbations or the specific boundary
conditions are imposed).
Classical MD simulation is based on the integration of Newtons equation of
motion,

2
2
, =1, 2, ,
i
i
i
r
m F i N
t
c
=
c

(3.10)
where m
i
is the mass of particle i, i r

is the position of the particle, i F

is the force
acting on the particle from the system. For a system with N interacting particles, the
force is derived from potential energy
( ( )) i
i
F U r = V

(3.11)
These equations are solved numerically using a small time step, usually in
femtosecond scale.
3.2.3 Monte Carlo Simulation
Monte Carlo (MC) simulation is a statistical method to generate the collection of
configurations with desired probability distribution at given conditions (statistical
ensembles) such as temperature, pressure, volume, or chemical potential. MC
Chapter 3. Models and Methods
44
simulation is efficient because only potential energy rather than force is evaluated in
configuration sampling. Moreover, MC can perform motions without physical mean,
for example, a jump from one position to another, insertion or deletion of a new
molecule. In its simplest form, MC method is an algorithm for numerical integration
in mathematics. A set of parameters are randomly selected or randomly perturbed, and
a function of these parameters is evaluated. Depending on the system of interest,
various types of trial moves can be attempted to reach equilibrium, for example,
translation, rotation, displacement, regrowth. After that, statistically averaged
properties can be obtained. The ensembles for MC simulation generally include
canonical (NVT, in which the number of particles, volume, and temperature T are
constant), microcanonical (NVE, in which number of particles, volume, energy are
constant), isobaric-isothermal (NPT, in which number of particles, pressure,
temperature are constant), grand-canonical (VT, in which the chemical potential,
volume and temperature are constant).
In this thesis, adsorption in MOFs was simulated by the grand-canonical Monte
Carlo (GCMC) method. As the chemical potentials of adsorbate in adsorbed and bulk
phases are identical at thermodynamic equilibrium, GCMC allows one to directly
relate the chemical potentials of adsorbate in both phases and has been widely used to
simulate adsorption. The framework atoms in MOFs were assumed to be rigid and the
potential energies between adsorbate atoms and framework were pre-tabulated. This is
because the low-energy equilibrium configurations are involved in adsorption and the
flexibility of framework has only a marginal effect. A recent simulation study on the
Chapter 3. Models and Methods
45
adsorption of noble gases in IRMOF-1 demonstrated that rigid and semi-flexible
frameworks gave close results at both low and room temperatures.
311
For most cases,
the LJ interactions were evaluated with a spherical cutoff of 15 (13 for
Zn(BDC)(TED)
0.5
and bio-MOF-11) with long-range corrections added. The
coulombic interactions were calculated using the Ewald sum method because it is the
best technique for electrostatic interactions in a periodic system. The real/reciprocal
space partition parameter and the cutoff for reciprocal lattice vectors were chosen to
be 0.2
-1
and 8 (10 for MIL-101), respectively, to ensure the convergence of Ewald
sum. The number of trial moves in a typical GCMC simulation was 2 10
7
, though
additional trial moves were used at high pressures. The first 10
7
moves were used for
equilibration and the second 10
7
moves for ensemble averages. Five types of trial
moves were randomly attempted in the GCMC simulations: displacement, rotation,
and partial regrowth at a neighboring position; complete regrowth at a new position;
and swap with reservoir including creation and deletion with equal probability. For
mixtures, another type of trial move, the exchange of molecular identity, was also
included. Unless otherwise mentioned, the simulation uncertainties were smaller than
the symbol sizes presented below.
3.3 Analysis Methods
3.3.1 Radial Distribution Functions
Radial distribution function represents the probability of finding a particle at a
certain distance from another particle. To examine the structure of a system, g(r) is
calculated by
Chapter 3. Models and Methods
46

2
( )
4
A
=
A
ij
ij
i j
N V
g r
r r N N t
(3.12)
where r is the distance between the geometric centers-of-mass of species i and j, N
ij

is the number of species j around i within a shell from r to r + r, V is the volume, N
i

and N
j
are the numbers of species i and j. Essentially, g(r) gives the ratio of local
density at position r to averaged overall density in the system.
3.3.2 Adsorption Selectivity
The separation factor of a binary mixture is usually quantified by adsorption
selectivity

/
( / )( / ) =
i j i j j i
S x x y y
(3.13)

where
i
x and
i
y are the mole fractions of component i in adsorbed and bulk phases,
respectively.
3.3.3 Mean-Squared Displacement
The dynamic properties of a system is evaluated by mean-squared displacement

2
1
1
MSD( ) ( )
N
i
i
t t
N
=
= A

r (3.14)
where N is the number of ions and r
i
(t) is the displacement of i
th
ion at time t.


Chapter 4. Adsorption of CO
2
and CH
4
in MIL-101
47
Chapter 4. Adsorption of CO
2
and CH
4
in MIL-101
In this chapter, adsorption of CO
2
and CH
4
in MIL-101 is investigated at 303 K
combining quantum mechanics and molecular simulation. Both dehydrated and
hydrated MIL-101 are considered to examine the effect of terminal water molecules.
4.1 Models and Methods
MIL-101 is assembled by corner-sharing supertetrahedra, which consist of Cr
3
O
trimers and 1,4-benzenedicarboxylic acids.
30
Connected by 8.6 apertures, the
supertetrahedra include 4 vertices and 6 edges that are occupied by the trimers and
organic linkers, respectively. MIL-10 has an augmented three-dimensional MTN
zeotype structure with a giant cell volume (~ 702,000
3
). Two types of mesoporous
quasi-spherical cages exist in MIL-101: a small cage of 20 supertetrahedra and a free
diameter of 29 accessible through a pentagonal window with 12 aperture, and a
large one of 28 supertetrahedra and a free diameter of 34 accessible through both
hexagonal and pentagonal windows with 14.7 16 aperture. MIL-101 is highly
porous with BET and Langmuir surface areas of about 4100 200 and 5900 300
m
2
/g, respectively. Each octahedral Cr is bonded with four oxygen atoms from
carboxylates, one
3
O oxygen atom, and one terminal site. The terminal site could be
occupied by a fluorine atom or a terminal water molecule. As a consequence, there are
three terminal sites in each Cr
3
O trimer and the ratio of fluorine to water is 1:2. The
terminal water molecules can be removed after dehydration, thus providing exposed
metallic Lewis acid sites in MIL-101.
30

Chapter 4. Adsorption of CO
2
and CH
4
in MIL-101
48



Figure 4.1. A unit cell of dehydrated MIL-101 constructed from experimental
crystallographic data, energy minimization and density functional theory calculation
(see the text for details). The pentagonal and hexagonal windows are enlarged for
clarity. Color code: Cr, orange polyhedra; F, cyan; C, blue; O, red; H, white.

Figure 4.1 shows a unit cell of dehydrated MIL-101, in which the pentagonal and
hexagonal windows are enlarged for clarity. The unit cell structure was constructed by
three steps based on experimental crystallographic data and computational approaches.
In the first step, all the disordered and terminal water molecules in the
crystallographic data were removed. Indeed, only the oxygen atoms of water
molecules were removed because the hydrogen atoms were not in the crystallographic
data. To construct hydrated MIL-101, however, only one terminal water molecule in

hexagonal
pentagonal
Chapter 4. Adsorption of CO
2
and CH
4
in MIL-101
49

Figure 4.2. Merz-Kollman charges of Cr
3
O trimer with terminal fluorine and water
molecules in (a) dehydrated and (b) hydrated MIL-101. The cleaved bonds of Cr
3
O
(indicated by the circles) were saturated by methyl group. Color code: Cr, orange; F,
cyan; C, blue; O, red; H, white.


Figure 4.3. Mulliken charges of the Cr
3
O trimer with terminal fluorine and water
molecules in (a) dehydrated and (b) hydrated MIL-101. The cleaved bonds of Cr
3
O
(indicated by the circles) were saturated by methyl group. Color code: Cr, orange; F,
cyan; C, blue; O, red; H, white.

each Cr
3
O trimer was removed. In the second step, hydrogen and fluorine atoms were
added respectively to the phenyl rings and the exposed Cr sites (one in each Cr
3
O
trimer). For hydrated MIL-101, hydrogen atoms were also added to the oxygen atoms
of terminal water molecules. The unit cell was then energetically minimized using
Materials Studio.
312
The framework atoms were represented by the Universal Force
(b) (a)
C
3
(-0.094)
H (0.154)
Cr
1
(1.557)
O
1
(-0.889)
C
2
(-0.093)
C
1
(0.853)
Cr
2
(1.886)
O
2
(-0.766)
Cr
2
(1.886)
F (-0.523)
Cr
1
(1.448)
C
3
(-0.120)
H (0.148)
C
2
(-0.037)
C
1
(0.689)
Cr
2
(1.747)
O
2
(-0.654)
Cr
2
(1.747)
F (-0.526)
OW (-0.758)
Cr
1
(1.153)
C
3
(-0.043)
H (0.098)
O
1
(-0.822)
C
2
(-0.069)
C
1
(0.631)
Cr
2
(1.155)
O
2
(-0.537)
Cr
2
(1.155)
F (-0.512)
OW (-0.437)
(b)
C
3
(-0.041)
H (0.105)
Cr
1
(1.159)
O
1
(-0.764)
C
2
(-0.070)
C
1
(0.641)
Cr
2
(1.152)
O
2
(-0.533)
Cr
2
(1.152)
F (-0.505)
(a)
Chapter 4. Adsorption of CO
2
and CH
4
in MIL-101
50
Field (UFF).
302
The phenyl rings in the experimental crystallographic data were
slightly non-planar and transformed to planar after the energy minimization. In the
third step, geometry optimizations were conducted by density-functional theory (DFT)
to precisely determine the positions of the added hydrogen and fluorine atoms.
Because of the tremendously large number of atoms in a unit cell, the DFT
optimizations were performed on a Cr
3
O trimer as shown in Figure 4.2. To maintain
the original hybridization, the cleaved bonds of the trimer were saturated by methyl
groups. The Becke exchange plus Lee-Yang-Parr correlation functional and the
effective core potentials were adopted in the DFT calculations. In the latter,
pseudo-potentials are used to represent the potential of the nucleus and core electrons
experienced by the valence electrons. This allows only the softer valence electron
wave functions to be explicitly treated, which is usually the portion that controls the
chemistry and can significantly reduce computational cost. Double- numerical
polarization (DNP) basis set was used in the DFT calculations, which is comparable
to 6-31G(d,p) Gaussian-type basis set. DNP basis set incorporates p-type polarization
into hydrogen atoms and d-type polarization into heavier atoms. In both dehydrated
and hydrated MIL-101, the optimized distance from F to Cr is 1.86 . In hydrated
MIL-101, the oxygen atom (OW) of terminal water is about 2.26 from Cr.
The atomic charges of MIL-101 framework atoms were fitted to the electrostatic
potentials using the Merz-Kollman (MK) scheme.
301,313
It is worthwhile to note that
the concept of atomic charges is solely an approximation, and no unique
straightforward method is currently available to determine them on a rigorous level.
Chapter 4. Adsorption of CO
2
and CH
4
in MIL-101
51
To evaluate the sensitivity of framework charges, the Mulliken population analysis
314

was also used to estimate the charges. The MK and Mulliken charges are listed in
Figure 4.2 and Figure 4.3, respectively.
The DFT calculations also gave the electrostatic potential maps around the trimer.
Figure 4.4 illustrates that the fluorine atom possesses large electronegativity in both
dehydrated and hydrated MIL-101. The significant difference is the electric fields
generated by the terminal H
2
O molecules in the hydrated MIL-101. As we shall see,
this has an intriguing effect on adsorption.



Figure 4.4. Electrostatic potential maps around the Cr
3
O trimer in (a) dehydrated and
(b) hydrated MIL-101.


The free volumes of the dehydrated and hydrated MIL-101 structures were
estimated by Monte Carlo methods.
315
A single helium atom was attempted to insert
into the host structure and
He
ad B
exp[ ( ) / ] d r r
V
u k T -

gave the free volume, in which


He
ad
u is the interaction between helium and adsorbent. It should be noted that the free
volume estimated by helium adsorption is slightly greater than the actual geometrical
(b)
(a)
Chapter 4. Adsorption of CO
2
and CH
4
in MIL-101
52
volume.
316
The calculated free volumes in the dehydrated and hydrated MIL-101 were
1.89 and 1.85 cm
3
/g, respectively; and the porosities were 0.83 and 0.81.

Table 4.1. Potential parameters and atomic charges of CO
2
, CH
4
, and terminal H
2
O.

Site (Adsorbate) o () c /k
B
(K) Z(e)
C (CO
2
) 2.789 29.66 +0.576
O (CO
2
)

3.011 82.96 0.288

CH
4
3.73 148.0 0
OW (H
2
O) 3.151

76.42

0.401

HW (H
2
O) 0 0 +0.758

For the adsorption of CO
2
and CH
4
in MIL-101, CO
2
was represented as a
three-site rigid molecule and its intrinsic quadrupole moment was described by a
partial charge model.
317
The CO bond length was 1.18 and the bond angle ZOCO
was 180. CO
2
CO
2
interactions were modeled as a combination of Lennard-Jones
(LJ) and coulombic potentials. A united-atom model was used for CH
4
with the LJ
potential parameters from the TraPPE force field that was developed to reproduce the
critical parameters and saturated liquid densities of alkanes.
318
The LJ potential
parameters of the terminal water molecules were adopted from TIP3P (three-point
transferable interaction potential) model.
319
Table 4.1 lists the potential parameters
and atomic charges of CO
2
, CH
4
and terminal water. The dispersion interactions of the
framework atoms in MIL-101 were modeled using the Universal Force Field
(UFF).
302
The Lorentz-Berthelot combining rules were used to calculate the cross LJ
interaction parameters. A number of simulation studies have shown that UFF can
Chapter 4. Adsorption of CO
2
and CH
4
in MIL-101
53
accurately predict gas adsorption in various MOFs.
39,41,191,320
The adsorption of CO
2

and CH
4
in MIL-101 was simulated by grand canonical Monte Carlo (GCMC) method
as described in Chapter 3.
4.2 Results and Discussion
4.2.1 Sensitivity of Framework Charges
CO
2
adsorption in the dehydrated MIL-101 was simulated with the MK and
Mulliken charges on the framework atoms and the results are shown in Figure 4.5. For
comparison, the experimental data are included from the study by Llewellyn et al.
133

They measured the adsorption of CO
2
and CH
4
in three MIL-101 samples, namely,
MIL-101a as-synthesized, MIL-101b activated by ethanol, and MIL-101c activated by
ethanol and KF.
133
These samples differ primarily in the amount of residual
terephthalic acid and in the density of Lewis acid Cr sites (500, 700, and 1000 mol/g,
respectively). CH
4
adsorption was found to be essentially the same in the three
samples. However, CO
2
adsorption was affected by the activation method, and the
extent of adsorption increased in the order of MIL-101a < MIL-101b < MIL-101c.
133

Among the three samples, MIL-101b and MIL-101c were activated more thoroughly
and contained less amount of residual terephthalic acid and more open Cr sites;
consequently, they had the stronger interactions with CO
2
and the larger free volume
for adsorption than MIL-101a. The dehydrated MIL-101 model in our study is similar
to experimental MIL-101b and MIL-101c rather than MIL-101a. As seen in Figure 4.5,
the predicted CO
2
adsorption from both the MK and Mulliken charges match well
with the experimental data of MIL-101b and MIL-101c. The simulation results from
Chapter 4. Adsorption of CO
2
and CH
4
in MIL-101
54
the Mulliken charges are in good agreement with MIL-101b in the pressure range of 0
to 2000 kPa and the deviation is less than 6%. For the ESP charges, the deviation is
less than 19% at < 1000 kPa and is less than 7% at higher pressure. This validates the
models used in our study and suggests that CO
2
adsorption is relatively insensitive to
the method, either MK or Mulliken, for the framework charges. Unless otherwise
stated, the results for CO
4
presented below are based on the MK charges.
P (kPa)
0 1000 2000 3000 4000 5000
N

(
m
m
o
l
/
g
)
0
10
20
30
40


Figure 4.5. CO
2
adsorption in dehydrated MIL-101. The squares, diamonds and
circles are experimental data
133
in MIL-101a, MIL-101b and MIL-101c, respectively.

4.2.2 United-Atom and Five-site Models of CH
4

To examine the performance of united-atom and five-site models of CH
4
, Figure
4.6 shows the adsorption isotherms of CH
4
in the dehydrated MIL-101 predicted from
both models.
133
Because CH
4
adsorption was found experimentally to be insensitive to
the activation method, only the experimental data in MIL-101c are included in Figure
4.6. The predictions of both united-atom and five-site model are in fairly good
agreement with the experimental data, especially at low pressures. Nevertheless, the
A : ESP charges
: Mulliken charges
: exp. (MIL-101a)
: exp. (MIL-101b)
: exp. (MIL-101c)

Chapter 4. Adsorption of CO
2
and CH
4
in MIL-101
55
united-atom model overestimates, whereas the five-site model underestimates. Such a
trend was also observed in CH
4
adsorption on graphite and in graphitic slit pores, and
attributed to the better packing of the united-atom model in confined space.
321
Unless
otherwise stated, the results for CH
4
are based on the united-atom model.

P (kPa)
0 1000 2000 3000 4000 5000
N

(
m
m
o
l
/
g
)
0
5
10
15
20

Figure 4.6. CH
4
adsorption in dehydrated MIL-101. The circles are the experimental
data in MIL-101c.
133


4.2.3 Adsorption of Pure CO
2
and CH
4

Figure 4.7 shows the adsorption isotherms of CO
2
and CH
4
in both dehydrated and
hydrated MIL-101 on a gravimetric basis. CH
4
uptake increases almost linearly with
increasing pressure and is still not saturated at 50 bar because of the very porous
structure of MIL-101. This implies that the interaction between CH
4
and the MIL-101
is weak and a higher pressure is needed to attain saturation. Despite a slight
overestimation at high pressures, the simulated isotherms of CH
4
in both dehydrated
and hydrated MIL-101 are in good agreement with the experimental data. The
dehydrated MIL-101 exhibits greater adsorption than its hydrated counterpart at high
A : united-atom model
: five-site model
: exp. (MIL-101c)


Chapter 4. Adsorption of CO
2
and CH
4
in MIL-101
56
pressures. As will be seen, however, the opposite is observed at low pressures due to
the presence of terminal water molecules.

P (kPa)
0 1000 2000 3000 4000 5000
N

(
m
m
o
l
/
g
)
0
10
20
30
40
CO
2
CH
4

Figure 4.7. Adsorption isotherms of CO
2
and CH
4
on a gravimetric basis. The squares,
diamonds, and circles are the experimental data in MIL-101a (as-synthesized),
MIL-101b (activated by hot ethanol) and MIL-101c (activated by hot ethanol and KF),
respectively.
133


CO
2
adsorption is apparently greater than CH
4
because CO
2
has a quadrupole
moment and thus interacts more strongly with the adsorbent. The simulated isotherms
of CO
2
in dehydrated and hydrated MIL-101 compare fairly well with the
experimental data in MIL-101b and MIL-101c samples. The deviations can be
attributed primarily to difference between theoretical model and real sample. In our
study, the model does not contain any residual terephthalic acid, though some
probably remained in the sample. CO
2
uptake in MIL-101 at 50 bar ranges from 30 to
40 mmol/g depending on the activation method, which is higher than that in IRMOF-1
(22 mmol/g)
125
and Cu-BTC (11 mmol/g),
125
and comparable to that in MOF-177
(33.5 mmol/g at 42 bar)
125
and UMCM-1, IRMOF-10 and IRMOF-14 (40 ~ 45
mmol/g).
39
Similar to CH
4
adsorption, the dehydrated MIL-101 has greater adsorption
A : dehydrated
V : hydrated
: exp. (MIL-101a)
: exp. (MIL-101b)
: exp. (MIL-101c)

Chapter 4. Adsorption of CO
2
and CH
4
in MIL-101
57
for CO
2
at high pressures but the reverse is true at low pressures, which will be
discussed below.

P (kPa)
0 50 100 150 200 250 300
0
200
400
600
800
1000
CO
2
CH
4
CO
2
CH
4
P (kPa)
3000 3500 4000 4500 5000
2000
3000
6000
7000
CO
2
CH
4
N

(
m
o
l
e
c
u
l
e
s
/
U
C
)
(a) (b)

Figure 4.8. Adsorption isotherms of CO
2
and CH
4
in MIL-101 (a) at low pressure and
(b) high pressure based on the number of molecules per unit cell.

Figure 4.8 shows the adsorption isotherms of CO
2
and CH
4
on the basis of
molecules per unit cell. At low pressures, the adsorption particularly of CO
2
in
hydrated MIL-101 is greater than in dehydrated framework. Interestingly, the opposite
is observed at high pressures. Such as phenomenon was recently found of CO
2
and
CH
4
in microporous Cu-BTC
137
and of CO
2
/H
2
mixture in charged soc-MOF.
247
At
low pressures, the enhanced adsorption in hydrated MIL-101 is attributed to the
terminal water molecules that act as additional adsorption sites; thus, adsorbates
experience stronger interactions. As seen in Figure 4.8a, the enhancement is more
pronounced for CO
2
than CH
4
. This is because CO
2
is a quadrapolar molecule and
interacts strongly with the charged hydrogen and oxygen atoms of the terminal water
molecules. However, the united-atom model of CH
4
can only interact with water
molecules through weak disperse interaction. At high pressures, the pores in MIL-101
are largely filled by adsorbate, and the free volume is a dominant factor to determine
V : hydrated
A : dehydrated

V : hydrated
A : dehydrated
Chapter 4. Adsorption of CO
2
and CH
4
in MIL-101
58
adsorption capacity. Apparently, the terminal water molecules occupy part of the
volume and prevent the framework from accommodating more adsorbate molecules.
Consequently, hydrated MIL-101 exhibits lower adsorption at high pressures.




Figure 4.9. Snapshots of CO
2
and CH
4
in a pentagonal window in dehydrated
MIL-101 at 10, 100, and 1000 kPa. Color code: Cr, orange; F, cyan; C, blue; O, red;
H, white; CO
2
, green; CH
4
, pink.

To identify the favorable adsorption sites in MIL-101, Figure 4.9 illustrates the
simulation snapshots of CO
2
and CH
4
in a pentagonal window in the dehydrated
MIL-101. At 10 kPa, the adsorbates are located exclusively in the supertetrahedra.
Attributed to the strong overlap of surface potentials, the microporous
supertetrahedral are the most favorable adsorption sites. At 100 kPa with increased
amount of adsorption, adsorbates appear in the pentagonal window in addition to
supertetrahedra. Compared to CH
4
, CO
2
molecules also occupy the edge of the
pentagonal window close to the exposed Cr
2
sites. At 1000 kPa, more adsorbates
CO2
CH4
10 kPa 100 kPa 1000 kPa
Chapter 4. Adsorption of CO
2
and CH
4
in MIL-101
59
populate the pentagonal window. Though not shown here, at very high pressures the
mesoporous cages are gradually filled.

CO
2
-Cr
1
r ()
2 4 6 8 10 12 14 16
g
(
r
)
0
1
2
3
4
10 kPa
100 kPa
1000 kPa
5000 kPa
CO
2
-Cr
2
r ()
2 4 6 8 10 12 14 16
0
1
2
3
4
10 kPa
100 kPa
1000 kPa
5000 kPa

CH
4
-Cr
1
r ()
2 4 6 8 10 12 14 16
g
(
r
)
0
1
2
3
4
10 kPa
100 kPa
1000 kPa
5000 kPa
CH
4
-Cr
2
r ()
2 4 6 8 10 12 14 16
0
1
2
3
4
10 kPa
100 kPa
1000 kPa
5000 kPa

Figure 4.10. Radial distribution functions of CO
2
and CH
4
around Cr
1
and Cr
2
atoms
in dehydrated MIL-101 at 10, 100, 1000, and 5000 kPa.

The simulation snapshots in Figure 4.9 provide only qualitative information. To
quantitatively characterize the adsorption sites, the radial distribution functions of
adsorbates around the framework atoms were calculated. Figure 4.10 shows the g(r)
of CO
2
and CH
4
around Cr
1
and Cr
2
atoms in dehydrated MIL-101 at various
pressures. For all the four cases shown, pronounced peaks are observed at r = 5 ~ 9 ,
revealing that the supertetrahedra are preferential adsorption sites. Furthermore, two
additional peaks are observed in the g(r) of CO
2
-Cr
1
. This is due to the attractions
Chapter 4. Adsorption of CO
2
and CH
4
in MIL-101
60
between fluorine atoms (bonded to Cr
1
sites) and CO
2
molecules in two neighboring
supertetrahedra. As illustrated by the electrostatic potential maps in Figure 4.4, the
fluorine atoms possess large electronegativity and thus interact strongly with CO
2

molecules.
There is a distinct difference in the short region of g(r) around Cr
1
and Cr
2
. In the
dehydrated MIL-101, Cr
1
sites are bonded with fluorine atoms, whereas Cr
2
sites are
exposed. Small peaks are observed in the g(r) of CO
2
-Cr
2
at r = 3.5 ; though there
are no peaks in the g(r) of CH
4
-Cr
2
, shoulders exist in the region r = 3 ~ 4 . In
remarkable contrast, such peaks and shoulders are not seen around Cr
1
atoms. That is,
adsorbate molecules are closer to Cr
2
rather than Cr
1
sites. Because of the quadrapole
moment, CO
2
interacts more strongly with the Cr
2
sites than CH
4
, leading to the peaks
in g(r) of CO
2
-Cr
2
. Figure 4.11 schematically demonstrates the locations of CO
2
and
CH
4
molecules around Cr
2
sites. We should note that the region near Cr
2
sites is
substantially smaller than the size of supertetrahedra; therefore, the peaks in g(r) at
3.5 are much lower and narrower than those at 5 ~ 9 .



Figure 4.11. Schematic locations of CO
2
and CH
4
near Cr
3
O trimer in dehydrated
MIL-101. Color code: Cr, orange; F, cyan; C, blue; O, red; H, white.

CH4
CO2
Chapter 4. Adsorption of CO
2
and CH
4
in MIL-101
61
With rising pressure, the number of adsorbed molecules increases. As defined by
eq. (3.1), g(r) is the local density relative to the overall density. In Figure 4.10, the
peaks of g(r) at 5 ~ 9 decrease when pressure rises, as also observed in our recent
studies for the adsorption of gas mixture and water in charged rho zeolite-like
MOF.
56,242
Interestingly, different behavior is found for the peaks at 3.5 closed to
the exposed Cr
2
sites. Specifically, the peak of CO
2
-Cr
2
increases with rising pressure
from 10 to 100 and then to 1000 kPa, but decreases from 1000 to 5000 kPa. This
implies that CO
2
molecules are increasingly adsorbed near the exposed Cr
2
sites until
1000 kPa; however, these regions are small and get saturated beyond 1000 kPa.
Unlike CO
2
-Cr
2
, the shoulder of CH
4
-Cr
2
still increases marginally when pressure
rises from 1000 to 5000 kPa. This is because the interaction of CH
4
-Cr
2
is weaker
than that of CO
2
-Cr
2
, thus the saturation of Cr
2
sites by CH
4
appears at a higher
pressure. The radial distribution functions discussed here provide useful insights into
the favorable interaction sites of adsorbates in the dehydrated MIL-101.
We next consider the hydrated MIL-101, in which both Cr
1
and Cr
2
sites are
coordinated thus exhibit similar g(r). As shown in Figure 4.12, the g(r) are almost the
same for CO
2
-Cr
1
and CO
2
-Cr
2
, also for CH
4
-Cr
1
and CH
4
-Cr
2
. While there are peaks
at 5 ~ 9 , no peaks or shoulders are observed at 3.5 , which are significantly
different from Figure 4.10. The pressure dependence of the g(r) is similar to the
dehydrated framework, that is, the peak decreases with rising pressure.

Chapter 4. Adsorption of CO
2
and CH
4
in MIL-101
62
CO
2
-Cr
2
r ()
2 4 6 8 10 12 14 16
0
1
2
3
4
CO
2
-Cr
1
r ()
2 4 6 8 10 12 14 16
g
(
r
)
0
1
2
3
4

CH
4
-Cr
1
r ()
2 4 6 8 10 12 14 16
g
(
r
)
0
1
2
3
4
CH
4
-Cr
2
r ()
2 4 6 8 10 12 14 16
0
1
2
3
4

Figure 4.12. Radial distribution functions of CO
2
and CH
4
around Cr
1
and Cr
2
atoms
in hydrated MIL-101 at 10, 100, 1000, and 5000 kPa.

Figure 4.13 shows the radial distribution functions of CO
2
and CH
4
around the
oxygen atoms (OW) of the terminal water molecules in the hydrated MIL-101 at 10,
100, 1000 and 5000 kPa, respectively. Similar to the g(r) of CO
2
-Cr
2
and CH
4
-Cr
2
in
Figure 3.10, peaks and shoulders are observed in the g(r) of CO
2
-OW and CH
4
-OW at
r = 3.8 . The reason is that the terminal water molecules act as additional adsorption
sites, especially for CO
2
. Comparing Figures 4.10 and 4.13, we find that the peaks and
shoulders are broader in the latter, suggesting that the terminal water molecules result
in stronger interactions with adsorbates than the exposed Cr
2
sites. As a consequence,
adsorption at low pressures is greater for both CO
2
and CH
4
in the hydrated MIL-101.

10 kPa
100 kPa
1000 kPa
5000 kPa
10 kPa
100 kPa
1000 kPa
5000 kPa
10 kPa
100 kPa
1000 kPa
5000 kPa
10 kPa
100 kPa
1000 kPa
5000 kPa
Chapter 4. Adsorption of CO
2
and CH
4
in MIL-101
63
r ()
2 4 6 8 10 12 14 16
g
(
r
)
0
1
2
3
4
10 kPa
100 kPa
1000 kPa
5000 kPa
CO
2
-OW

r ()
2 4 6 8 10 12 14 16
0
1
2
3
4
10 kPa
100 kPa
1000 kPa
5000 kPa
CH
4
-OW


Figure 4.13. Radial distribution functions of CO
2
and CH
4
around the oxygen atoms
of terminal water molecules in hydrated MIL-101 at 10, 100, 1000, and 5000 kPa.

4.2.4 Adsorption of CO
2
/CH
4
Mixture
Adsorption of equimolar mixture of CO
2
and CH
4
was simulated in the dehydrated
and hydrated MIL-101. As shown in Figure 4.14a, the extent of adsorption of each
component is not substantially affected by hydration. With a closer look, however, we
find that CO
2
adsorption at low pressures is marginally higher in the hydrated
MIL-101 than in the dehydrated counterpart. At high pressures, there is no discernable
different in the two structures. In contrast, CH
4
adsorption in dehydrated MIL-101 is
slightly higher at moderate and high pressures. As shown in Figure 4.14b, with
increasing pressure the selectivity of CO
2
/CH
4
first increases, then drops rapidly, and
finally increases slightly. The initial increase is due to the preferential adsorption of
CO
2
in the multiple supertetrahedra. Upon adsorption occurring in the less attractive
windows and mesoporous cage, the selectivity drops. At high pressures, the selectivity
increases is because of the cooperative attractions between adsorbed CO
2
molecules.
Interestingly, the selectivity of CO
2
/CH
4
is higher in the hydrated MIL-101, implying
Chapter 4. Adsorption of CO
2
and CH
4
in MIL-101
64
that the separation efficacy could be enhanced by introducing terminal H
2
O molecules.
This was also observed in the adsorption of CO
2
/CH
4
in Cu-BTC
137
and of CO
2
/H
2
in
charged soc-MOF.
247


P (kPa)
0 1000 2000 3000 4000 5000
0
1000
2000
3000
4000
CH
4
P (kPa)
0 1000 2000 3000 4000 5000
3
4
5
6
7
(a) (b)
N

(
m
o
l
e
c
u
l
e
s
/
U
C
)
S
e
l
e
c
t
i
v
i
t
y

C
O
2
/
C
H
4
CO
2

Figure 4.14. Adsorption of equimolar CO
2
/CH
4
mixture. (a) isotherm and (b)
selectivity of CO
2
over CH
4
.

4.3 Summary
We have investigated the adsorption of CO
2
and CH
4
in mesoporous MIL-101.
The dehydrated and hydrated MIL-101 structures were constructed on the basis of
experimental crystallographic data, molecular mechanics minimization, and
first-principles calculations. Adsorption occurs exclusively in the microporous
supertetrahedra at low pressures, and then in the mesoporous cages at high pressures.
The simulated isotherms of pure CO
2
and CH
4
match well with measurements, despite
variations in experimental data depending on the activation method used. The
terminal water molecules in hydrated MIL-101 have an interesting effect on
adsorption by acting as additional adsorption sites that enhance adsorption at low
pressures. The enhancement is greater for CO
2
as it is a quadrapolar molecule and
V : hydrated
A : dehydrated

V : hydrated
A : dehydrated
Chapter 4. Adsorption of CO
2
and CH
4
in MIL-101
65
interacts strongly with charged hydrogen and oxygen atoms of the water molecules.
At high pressures, however, the terminal water molecules reduce free volume and lead
to less adsorption compared to the dehydrated MIL101.
Through structural analysis, adsorbates in dehydrated MIL-101 were found to
locate preferentially near the exposed Cr
2
sites as evidenced by the peaks and
shoulders in the radial distribution functions. However, such peaks and shoulders
were not seen around the fluorine saturated Cr
1
. In hydrated MIL-101, the radial
distribution functions behave similar for all Cr sites. Compared with those in
dehydrated MIL-101, the peaks and shoulders around the terminal water molecules
are higher and broader. This reveals that the terminal water molecules interact more
strongly than the exposed Cr
2
sites for adsorbates. Therefore, the hydrated MIL-101
exhibits greater adsorption at low pressures. The adsorption selectivity of CO
2
over
CH
4
is slightly higher in the hydrated MIL-101. This study provides useful insights
into the microscopic adsorption behavior in a unique MOF and underlines the
interesting effect of the terminal water molecules on adsorption.
Chapter 5. Adsorption and Separation in Zn(BDC)(TED)
0.5

66
Chapter 5. Adsorption and Separation in Hydrophobic
Zn(BDC)(TED)
0.5

In this Chapter, adsorption and separation of CH
3
OH/H
2
O and CO
2
/CH
4
in
Zn(BDC)(TED)
0.5
are investigated using molecular simulations. The objective is to
provide the microscopic understanding of adsorption behavior in this novel MOF at a
molecular level. The study is technologically important toward the capability of
Zn(BDC)(TED)
0.5
in the purification of (methanol-based) liquid fuel and natural gas.
The removal of H
2
O from liquid fuel and of CO
2
from natural gas is crucial to their
efficacy in combustion. We compare the experimental and simulated adsorption
isotherms of single component CH
3
OH, H
2
O, CO
2
, and CH
4
, and then predict the
adsorption of mixtures. The effect of H
2
O on the separation of CO
2
/CH
4
mixture is
also examined. As mentioned that among all MOFs reported, Zn(BDC)(TED)
0.5

exhibits the highest adsorption for hydrocarbons. Therefore, adsorption of hexane
(n-C
6
) was further simulated in Zn(BDC)(TED)
0.5
. To provide the microscopic
understanding of adsorption, the density distributions and structural properties of
adsorbates in Zn(BDC)(TED)
0.5
are presented from simulations.
5.1 Models and Methods
Zn(BDC)(TED)
0.5
possesses a paddle-wheel structure, in which the metal-oxide
building units Zn
2
(COO)
4
are bridged by BDC linkers to form a 2D square-grid net
[Zn
2
(1,4-bdc)
2
]. The TED pillars occupy the axial sites to extend the 2D layers into a
3D framework and are disordered along the crystallographic 4-fold axis. The structure
Chapter 5. Adsorption and Separation in Zn(BDC)(TED)
0.5

67
has a space group of P
4
/mmm, and lattice constants of a = 10.929 and c = 9.608 .
Figure 5.1 shows the crystal structure (3 3 3 unit cells) constructed from the
experimental crystallographic data and first-principles optimization. The optimization
was conducted on a unit cell of Zn(BDC)(TED)
0.5
by the periodical density functional
theory (DFT) using DMol3.
312
The same basis set was used as described in Chapter 4.
The effective core-potentials were adopted which represent the interactions of nucleus
and core electrons on the valence electrons. In such a way only the wave functions of
the softer valence electrons are explicitly treated, which usually controls the chemistry,
and can significantly reduce computational cost.

XY plane XZ plane YZ plane


Figure 5.1. A unit cell of Zn(BDC)(TED)
0.5
constructed from the experimental
crystallographic data and first-principles optimization. Color code: Zn, pink
polyhedra; N, green; C, blue; O, red; H, white.

There exist interlacing channels in the framework of Zn(BDC)(TED)
0.5
. Figure
5.2 shows the morphologies and diameters of the channels calculated using the HOLE
program.
322
The wide open channels are along the Z axis with a diameter ranging
from 7.3 to 9.2 , which are connected by the small windows along the X and Y axes
with a diameter of 3.6 . Therefore, small molecules may enter into the windows. The
free volume
free
V of Zn(BDC)(TED)
0.5
was evaluated from Monte Carlo
Chapter 5. Adsorption and Separation in Zn(BDC)(TED)
0.5

68
simulation
315
with helium as a probe. The
free
V was estimated to be 0.79 cm
3
/g; the
porosity was 64.9%, in good agreement with the experimental accessible volume
(61.3%).
150




Figure 5.2. Channels along the Z, X, and Y (from top to bottom) axes in
Zn(BDC)(TED)
0.5
. The green regions denote the small windows.

The atomic charges of framework atoms in Zn(BDC)(TED)
0.5
were also
calculated from DFT. Because of the large number of atoms in a unit cell, a
fragmental cluster shown in Figure 5.3 was used. The dangling bonds on the
fragmental cluster were terminated by hydrogen atoms. The DFT calculation used the
Becke exchange plus Lee-Yang-Parr functional and was carried out with Gaussian
03.
323
The 6-31G(d) basis set was used for all atoms except Zn atoms, for which
Chapter 5. Adsorption and Separation in Zn(BDC)(TED)
0.5

69
LANL2DZ basis set was used. LANL2DZ is a double-zeta basis set and contains
effective pseudo-potentials to represent the potentials of nucleus and core electrons.
The atomic charges were estimated by fitting to the electrostatic potentials from DFT
calculation.
313
The dispersion interactions of framework atoms in Zn(BDC)(TED)
0.5

were modeled by the Dreiding force field
303
which has been commonly adopted in the
simulation of MOFs.


Figure 5.3. Atomic charges in a fragmental cluster of Zn(BDC)(TED)
0.5
. The
dangling bonds (indicated by circles) were terminated by hydrogen. Color code: Zn,
pink polyhedra; N, green; C, blue; O, red; H, white.

H
2
O was mimicked by the three-point transferable interaction potentials (TIP3P)
model,
319
in which the O-H bond length was 0.9572 and the ZHOH angle was
104.52. It has been shown that the TIP3P model gives a reasonably good interaction
potential compared to the experimental value.
324
CO
2
was represented as a three-site
rigid molecule as described in Chapter 4. The intermolecular interactions of H
2
O and
CO
2
were modeled by the additive LJ and Coulombic potentials. CH
3
OH, CH
4
and
n-C
6
were represented by united-atom models with CH
x
as a single interaction site.
The potential parameters were adopted from the transferable potentials for phase
C1 (0.855)
C2 (-0.105)
C3 (-0.142)
H1 (0.131)
O (0.743)
C4 (-0.032)
H2 (0.051)
Zn (1.205) N (0.098)
Chapter 5. Adsorption and Separation in Zn(BDC)(TED)
0.5

70
equilibria (TraPPE) force field, developed to reproduce the critical parameters and
saturated liquid densities of alkanes and alcohols.
318,325
The cross LJ parameters were
evaluated by the Lorentz-Berthelot combining rules. In addition to the LJ interaction
(and the Coulombic interaction for CH
3
OH), there were bending interaction in
CH
3
OH and n-C
6

2
0
( ) 0.5 ( )
bending
u k
u
u u u = (5.1)
Furthermore, the torsion interaction was also taken into account in n-C
6

0 1 2 3
( ) [1 cos ] [1 cos(2 )] [1 cos(3 )]
torsion
u c c c c | | | | = + + + + + (5.2)


Table 5.1. Potential parameters and atomic charges.

Species
LJ and Coulombic Potential
Bond Length
Bending Angle and Force
Constant
site o () c/k
B
(K) q (e)
CH
4
CH
4
3.73

148.0

0


CO
2

C 2.789

29.66

+0.576

r
C-O
= 1.18

ZOCO
= 180
o

O 3.011

82.96

0.288

H
2
O
O 3.151 76.47 0.834
r
H-O
= 0.9527

ZHOH
= 104.52
o

H 0 0 +0.417

CH
3
OH
CH
3
3.75 98.0 +0.265 r
CH3-O
= 1.43
r
O-H
= 0.945

ZCH3OH
= 108.5
o

k

/k
B
= 55400 K

O 3.02 93.0 0.700

H 0 0 +0.435
n-C
6

CH
3
3.75 98.0
0

r
CHx-CHy
= 1.54
x y 2
CH CH CH 113.0 =

u

k

/k
B
= 62500 K
CH
2
3.95 46.0 0


The force field parameters used are listed in Table 5.1. GCMC simulations were
carried out for the adsorption of pure components and mixtures in Zn(BDC)(TED)
0.5
.
Chapter 5. Adsorption and Separation in Zn(BDC)(TED)
0.5

71
5.2 Results and Discussion
5.2.1 CH
3
OH/H
2
O
Figure 5.4 shows the adsorption of pure CH
3
OH and H
2
O in Zn(BDC)(TED)
0.5
at
303 K. The simulated isotherms are in fairly good agreement with experimental
results, particularly for CH
3
OH at low pressures (deviation ~ 10.8%) and for H
2
O
(deviation ~ 13.3%) over the entire pressure range under study. Note that the
saturation pressure is 21.7 kPa for CH
3
OH and 4.2 kPa for H
2
O.
326
With increasing
pressure, the uptake of CH
3
OH increases sharply until P/P
0
= 0.2 and finally
approaches a plateau. The simulation underestimates the saturation loading of CH
3
OH.
Specifically, the predicted saturation loading is approximately is 460 mg/g and lower
than the experimental value of 510 mg/g.
150
This might imply that the force field used
to model the interactions of CH
3
OH-CH
3
OH and CH
3
OH-adsorbent might not be very
accurate, and the agreement could be improved by more sophisticated modeling.

P (kPa)
0 5 10 15 20 25
N

(
m
g
/
g
)
0
100
200
300
400
500
P/P
o
0.0 0.1 0.2 0.3 0.4 0.5
0
100
200
300
400
500
N

(
m
g
/
g
)
P (kPa)
0 4 8 12 16
N

(
m
g
/
g
)
0
100
200
300
400
500
P/P
o
0 1 2 3 4
N

(
m
g
/
g
)
0
100
200
300
400
500
(a) CH
3
OH (b) H
2
O


Figure 5.4. Isotherms of pure CH
3
OH and H
2
O at 303 K. The filled circles are
experimental data. The upper and lower triangles are adsorption and desorption data
from simulation. The insets show the isotherms a function of reduced pressure. The
saturation pressure P
o
is 21.7 kPa for CH
3
OH and 4.2 kPa for H
2
O.

Chapter 5. Adsorption and Separation in Zn(BDC)(TED)
0.5

72
H
2
O uptake from simulation is negligible at P/P
o
< 3.0, which is consistent with
the experimental data up to P/P
o
= 0.4. Interestingly, a hysteresis is observed in the
simulated adsorption and desorption isotherms of H
2
O at P/P
o
between 2.6 and 3.1,
indicating the occurrence of capillary phase transition. Zn(BDC)(TED)
0.5
possesses a
highly hydrophobic framework with the BDC and TED linkers surrounding the metal
oxides; therefore, the interaction with H
2
O is very weak. It is instructive to note that
the behavior of H
2
O in Zn(BDC)(TED)
0.5
is significantly different from that in a
hydrophilic MOF. Our recent study revealed that H
2
O is strongly adsorbed in
cation-exchanged rho-ZMOFs because of the high affinity of ionic framework and
extraframework ions.
242
Specifically, H
2
O in rho-ZMOFs exhibits a three-step
adsorption mechanism. At low pressures, H
2
O is preferentially adsorbed on the
extraframework ions; with increasing pressure, adsorption occurs near the framework
and finally in the large cage.
XY plane XZ plane YZ plane


Figure 5.5. Density contours of CH
3
OH at 1 kPa (top) and 10 kPa (bottom).
I
I
II
III
III
III
III
I, II
I
II
Chapter 5. Adsorption and Separation in Zn(BDC)(TED)
0.5

73
Figure 5.5 shows the density contours of CH
3
OH in Zn(BDC)(TED)
0.5
at 1 and
10 kPa, respectively. On the XY plane, CH
3
OH molecules exclusively occupy the
open square channels. Nevertheless, the channels are not homogeneous due to the
asymmetrical locations of dimethylenes (CH
2
CH
2
). Based on the adsorbate density,
three adsorption sites (I, II, and III) are roughly identified in each channel. Site I is at
the corner and proximal to the metal oxide, phenyl ring, and dimethylene. As a polar
molecule, CH
3
OH experiences strong interaction at site I with the metal oxides and
the overlap of surface potentials. Compared with site I, sites II and III are less
favorable. This is because the steric hindrance of dimethylenes prevents CH
3
OH at
sites II and III from being near the metal oxides. It is clearly seen on the XZ and YZ
planes that the distribution of CH
3
OH is extended along the Z axis. CH
3
OH is also
located in the mouth of small windows. With increasing pressure from 1 to 10 kPa,
the density distribution in the channels is less inhomogeneous and the small windows
are partially occupied. However, the density in the small windows is very low.
The observed change in the density contours with increasing pressure can be
further elucidated from the radial distribution functions g(r) of CH
3
OH around typical
framework atoms. As shown in Figure 5.6, the peak heights in the g(r) around Zn and
N atom rise when pressure increases from 1 to 10 kPa. This indicates that CH
3
OH
molecules are located closer to the metal oxides at 10 kPa, particularly for molecules
at sites II and III. Consequently, the density difference among the three types of
adsorption sites reduces and the distribution tends to be homogeneous, which is
observed in Figure 5.5. CH
3
OH molecules also stay near the phenyl rings at 10 kPa,
Chapter 5. Adsorption and Separation in Zn(BDC)(TED)
0.5

74
evidenced from the rising peak height in the g(r) around C2 atom. Nevertheless, the
peak height in the g(r) around C4 atom remains essentially unchanged.
C2
r ()
2 4 6 8 10 12
Zn
r ()
2 4 6 8 10 12
g

(
r
)
0.0
0.5
1.0
1.5
2.0
2.5
1 kPa
10 kPa
N
r ()
2 4 6 8 10 12
C4
r ()
2 4 6 8 10 12
1 kPa
10 kPa
1kPa
10kPa
1 kPa
10 kPa

Figure 5.6. Radial distribution functions of CH
3
OH around Zn, N, C2, and C4 atoms
of Zn(BDC)(TED)
0.5
at 1 and 10 kPa.

P (kPa)
1 2 3 4 5 6 7 8 9 10
N

(
m
g
/
g
)
0
100
200
300
400
P (kPa)
1 2 3 4 5 6 7 8 9 10
S
C
H
3
O
H
/
H
2
O
0
5
10
15
20
25
CH
3
OH
H
2
O
(a)
(b)


Figure 5.7. (a) Adsorption and (b) selectivity of CH
3
OH/H
2
O mixture at 303 K.

The adsorption of equimolar CH
3
OH/H
2
O mixture in Zn(BDC)(TED)
0.5
at 303 K
is shown in Figure 5.7a. The loading of CH
3
OH is much higher than H
2
O, as
discussed above, which is attributed to the stronger interaction of CH
3
OH with the
hydrophobic framework. In contrast to the negligible adsorption of pure H
2
O,
appreciable amount of H
2
O is adsorbed from the CH
3
OH/H
2
O mixture. This is due to
the co-adsorption between H
2
O and the already adsorbed CH
3
OH. As seen in Figure
5.7b, the selectivity CH
3
OH over H
2
O is approximately 20 at 1 kPa. With increasing
pressure, the selectivity decreases gradually as a consequence of entropy (packing)
Chapter 5. Adsorption and Separation in Zn(BDC)(TED)
0.5

75
effect. H
2
O molecule is smaller compared to CH
3
OH and thus can fit into the
channels more effectively at high pressures, leading to the decrease in selectivity. The
high selectivity at low pressures suggests that Zn(BDC)(TED)
0.5
is a promising
candidate for separation of CH
3
OH/H
2
O. Experimental study has shown that in a
pressure swing adsorption process, the purity of produced CH
3
OH reaches 94.7%
after the first cycle.
150

5.2.2 CO
2
/CH
4

Figure 5.8 shows the adsorption of pure CO
2
and CH
4
in Zn(BDC)(TED)
0.5
at
298 K. CO
2
adsorption is greater than CH
4
due to two reasons. First, CO
2
has stronger
dispersion and electrostatic interactions with the framework compared to CH
4
. Second,
the temperature 298 K considered is subcritical for CO
2
(T
c
= 304.4 K), but
supercritical for CH
4
(T
c
= 190.5 K); that is, CO
2
is substantially more condensable
than CH
4
at 298 K. Fairly good agreement is found between simulated and
experimental isotherms, particularly for CH
4
. The overestimation by simulation may
attribute to the fact that real samples usually contain impurities (solvents, like DMF),
which block the channels and reduce the adsorption. At 30 bar the amount of CO
2

adsorption is about 600 mg/g (13.6 mmol/g) in Zn(BDC)(TED)
0.5
, larger than that in
silicalite, carbon nanotube, F-MOF-1, Mn-MOF, and Cu-BTC.
48,147
This is because
Zn(BDC)(TED)
0.5
has a greater free volume, which governs adsorption capacity at a
high pressure.
Chapter 5. Adsorption and Separation in Zn(BDC)(TED)
0.5

76
P (kPa)
0 500 1000 1500 2000 2500 3000

N
e
x

(
m
g
/
g
)
0
100
200
300
400
500
600
CO
2
CH
4


Figure 5.8. Adsorption of pure CO
2
and CH
4
at 298 K. The open symbols are
simulation results and the filled symbols are experimental data.
327,328


XY plane XZ plane YZ plane





Figure 5.9. Density contours of CO
2
at 10, 100, and 3000 kPa (from top to bottom).

Chapter 5. Adsorption and Separation in Zn(BDC)(TED)
0.5

77
To identify the adsorption sites for CO
2
in Zn(BDC)(TED)0.5, the density
contours of CO
2
at 10, 100, and 3000 kPa are shown in Figure 4.9. The contours of
CO
2
differ remarkably from those of CH
3
OH in Figure 4.5. As discussed above,
CH
3
OH is a polar molecule and has strong interaction with the metal oxides. However,
CO
2
is nonpolar and located preferentially near the phenyl rings as seen on the XZ
and YZ planes at 10 kPa. With increasing pressure, further adsorbed CO
2
molecules
tend to be close to the metal oxides and dimethylenes. At 3000 kPa, the regions
between dimethylenes are densely occupied because of the large available space for
adsorption; furthermore, CO
2
molecules also extend into the small windows. It is
worthwhile to note that compared with CH
3
OH, CO
2
is more likely to enter the small
windows because of its linear shape.

Zn
r ()
2 4 6 8 10 12
g

(
r
)
0.0
0.5
1.0
1.5
2.0
2.5
10 kPa
100 kPa
3000 kPa
N
r ()
2 4 6 8 10 12
C2
r ()
2 4 6 8 10 12
C4
r ()
2 4 6 8 10 12
10 kPa
100 kPa
3000 kPa
10 kPa
100 kPa
3000 kPa
10 kPa
100 kPa
3000 kPa

Figure 5.10. Radial distribution functions of CO
2
around Zn, N, C2, and C4 atoms of
Zn(BDC)(TED)
0.5
at 10, 100, and 3000 kPa.

As an additional support for the observed shift in the adsorption sites, Figure
4.10 shows the radial distribution functions of CO
2
around typical framework atoms
at different pressure. With increasing pressure, the peak heights in the g(r) around Zn
and N atoms rise and demonstrate that the locations of CO
2
molecules move toward
the metal oxides at high pressures. The (first) peak height reduces in the g(r) around
C2 atom, but rises in the g(r) around C4 atom. This indicates more CO
2
molecules are
Chapter 5. Adsorption and Separation in Zn(BDC)(TED)
0.5

78
located near the dimethylenes when pressure increases, as observed above.
The adsorption of equimolar CO
2
/CH
4
mixture is shown in Figure 5.11. CO
2

isotherm increases rapidly at low pressures and then slowly with increasing pressure.
As a comparison, the extent of CH
4
uptake is much smaller and approaches saturation
at a low pressure. The selectivity of CO
2
over CH
4
exhibits similar behavior to CO
2

isotherm and increases from 4 to 7 with increasing pressure. The increase in
selectivity is attributed to the increased cooperative interactions between adsorbed
CO
2
molecules. The selectivity in Zn(BDC)(TED)
0.5
is similar to that in neutral MOFs
such as IRMOF-1, IRMOF-13, IRMOF-14, Cu-BTC, PCN-6, and PCN-6. Compared
with charged soc-MOF and rho-ZMOF, however, the selectivity in Zn(BDC)(TED)
0.5

is several orders of magnitude smaller.
42,246


P (kPa)
0 500 1000 1500 2000 2500 3000
N

(
m
g
/
g
)
0
100
200
300
400
500
600
CO
2
CH
4
CO
2
(with 0.1% H
2
O)
CH
4
(with 0.1% H
2
O)
P (kPa)
0 500 1000 1500 2000 2500 3000
S
C
O
2
/
C
H
4
3
4
5
6
7
8
CO
2
/CH
4
CO
2
/CH
4
(with 0.1% H
2
O)
(a)
(b)


Figure 5.11. (a) Adsorption and (b) selectivity of CO
2
/CH
4
equimolar mixture at 298
K. The filled symbols refer to the CO
2
/CH
4
mixture with 0.1% H
2
O.

In practice, gas mixture usually contains a small amount of moisture. The
presence of H
2
O may be adverse to the separation of gas mixture. For instance, upon
the addition of H
2
O into CO
2
/CH
4
mixture, the interaction between CO
2
and Na
+
in
Na-rho-ZMOF is substantially reduced. Consequently, CO
2
adsorption drops and the
Chapter 5. Adsorption and Separation in Zn(BDC)(TED)
0.5

79
selectivity decreases by one order of magnitude.
329
In other cases, however, the
addition of H
2
O is beneficial, e.g., the presence of pre-adsorbed H
2
O in BaX zeolite
increases the selectivity toward p-xylene in p-xylene/m-xylene mixture.
330
The
adsorption of CO
2
and CH
4
in Cu-BTC is enhanced by H
2
O molecules coordinated to
the open metal sites.
137
With a trace amount of H
2
O, the selectivity of CO
2
over H
2
in
soc-MOF increases at low pressures due to the promoted adsorption of CO
2
by H
2
O
bound onto the indium atoms, but decreases at high pressures as a result of the
competitive adsorption of H
2
O over CO
2
.
247

To examine the effect of H
2
O here, the adsorption of CO
2
/CH
4
mixture in the
presence of 0.1% H
2
O was simulated in Zn(BDC)(TED)
0.5
. As seen in Figure 4.11, the
addition of H
2
O has a marginal effect on the uptake of CO
2
or CH
4
and subsequently
the selectivity. The reason is that the hydrophobic Zn(BDC)(TED)
0.5
has very weak
affinity for H
2
O and the adsorption of H
2
O is negligible. This implies that a pre-water
treatment is probably not required prior to separation process. Interestingly, the effect
of H
2
O on the separation of gas mixture in Zn(BDC)(TED)
0.5
is significantly different
from that in other MOFs mentioned above, in which a pre-water treatment is crucial.
5.2.3 Hexane
Adsorption of n-C
6
was simulated in Zn(BDC)(TED)
0.5
at 313 K. Figure 5.12
shows that the simulated and experimental isotherms exhibit similar trend. As a
function of pressure, the adsorption increases rapidly at low pressures and then
reaches saturation. The saturation capacity is approximately 450 mg/g from
simulation, larger than the experimental value 400 mg/g. The deviation (< 15%) may
Chapter 5. Adsorption and Separation in Zn(BDC)(TED)
0.5

80
attribute to the difference of the crystal structures used. In the simulation, a perfect
crystal was adopted; however, experimental sample may contain impurities and
defects. As a consequence, the free volume and capacity in experiment are reduced.
Our results are in good agreement with the simulation results by Krishna and van
Baten.
265
A recent experimental study reported the kinetic separation of C
6
isomers
using fixed-bed adsorption.
174
The measured loading of n-C
6
is one order of
magnitude lower than the simulation results. This is probably because of the phase
transition or contamination of Zn(BDC)(TED)
0.5
sample during adsorption process.

P (torr)
0 20 40 60 80 100
N

(
m
g
/
g
)
0
100
200
300
400
500


Figure 5.12. Adsorption of hexane at 313 K. The open symbols are simulation results
and the filled symbols are experimental data.
150


The density contours of n-C
6
at 0.001 and 10 kPa are shown in Figure 5.13.
Similar to the contours of CO
2
, n-C
6
interacts strongly with the phenyl rings rather
than the metal oxides. Consequently, at 0.001 kPa the preferential site is near the
phenyl rings. At 10 kPa, more adsorbed n-C
6
molecules are closer to the metal oxides
and dimethylenes, as observed in CO
2
adsorption at a high pressure. Compared to
CO
2
, however, n-C
6
is larger in size and cannot extend into the small windows.
Chapter 5. Adsorption and Separation in Zn(BDC)(TED)
0.5

81
XY plane YZ plane


Figure 5.13. Density contours of hexane

at 0.001 kPa (top) and 10 kPa (bottom).

5.3 Summary
We have investigated the adsorption and separation of CH
3
OH/H
2
O and CO
2
/CH
4

mixtures in highly hydrophobic Zn(BDC)(TED)
0.5
. The framework contains wide
open channels along the Z axis, which are connected by small windows along the X
and Y axes. In general, the simulated isotherms of all pure components (CH
3
OH, H
2
O,
CO
2
, CH
4
, and n-C
6
) under study are in fairly good agreement with experimental
results. Because of the hydrophobic BDC and TED linkers, the affinity for H
2
O is
very weak and thus H
2
O adsorption is negligible. The selectivity CH
3
OH over H
2
O is
approximately 20 at 1 kPa, and decreases gradually with increasing pressure due to
entropy effect. The high selectivity at low pressures suggests that Zn(BDC)(TED)
0.5

could be used for purification of CH
3
OH (or other alcohols) from H
2
O. Three
favorable adsorption sites are identified in Zn(BDC)(TED)
0.5
for CH
3
OH. At site I
Chapter 5. Adsorption and Separation in Zn(BDC)(TED)
0.5

82
near the corner, CH
3
OH interacts strongly with the metal oxides and experiences the
strong surface potentials of the phenyl rings and dimethylenes. Sites II and III are less
favorable because of the steric hindrance of bulky TED linkers. With increasing
pressure, the density distribution in the square channels is less inhomogeneous and the
small windows are partially occupied. From the analysis of radial distribution
functions, CH
3
OH molecules are located near the metal oxides at a high pressure.
In contrast to polar CH
3
OH that has strong interaction with the metal oxides,
CO
2
is nonpolar and the region near the phenyl rings is the preferential adsorption site.
With increasing pressure, CO
2
molecules are proximal to the metal oxides and
dimethylenes, and also extend into the small windows. The selectivity of CO
2
over
CH
4
increases with increasing pressure, attributed to the increased cooperative
interactions between CO
2
molecules. In the presence of 0.1% H
2
O in CO
2
/CH
4

mixture, the adsorption and selectivity remain essentially the same. This implies that
H
2
O has a marginal effect on the separation of CO
2
/CH
4
in Zn(BDC)(TED)
0.5
and a
pre-water treatment is perhaps not required.


Chapter 6. CO
2
Capture in Bio-MOF-11
83
Chapter 6. CO
2
Capture in Bio-MOF-11
In this Chapter, we report the first molecular simulation study for mixture
adsorption in a bio-MOF; more specifically, the adsorption of CO
2
/H
2
and CO
2
/N
2

mixtures in bio-MOF-11. While it is relatively straightforward to experimentally
determine the adsorption of pure gases, quantitative measurement on the adsorption of
gas mixtures is challenging. First, the models and force fields are validated by
comparing simulated and experimental adsorption of pure CO
2
, H
2
, and N
2
; then the
favorable adsorption sites are identified in bio-MOF-11. Finally, the adsorption and
separation of CO
2
/H
2
and CO
2
/N
2
mixtures are examined and compared with those in
other porous materials. A gas mixture usually contains moisture; therefore, the effect
of H
2
O on the separation is also investigated.
6.1 Models and Methods
Rosi and coworkers
168
synthesized bio-MOF-11 by a solvothermal reaction. It has
a formula of Co
2
(ad)
2
(CO
2
CH
3
)
2
and is thermally stable up to 200 C. The framework
consists of paddle-wheel cobalt-adeninate-acetate cluster (Figure 6.1 a) as the
secondary building unit (SBU). There are five types of nitrogens, among which N1
and N6 are Lewis basic sites; N3, N7, and N9 are covalently bonded to cobalt atoms.
A three-dimensional structure with augmented lvt network topology is generated by
the SBUs. The structure is tetragonal with lattice length of a = b = 15.4355 and c =
22.775 . The cavities with diameter of 5.8 are periodically distributed throughout
the structure (Figure 6.1 b). These cavities form interlacing narrow channels along the
Chapter 6. CO
2
Capture in Bio-MOF-11
84
crystallographic a and b dimensions.
168




Figure 6.1. (a) Cobalt-adeninate-acetate cluster. N1 and N6 are the Lewis basic
pyrimidine and amino groups, while N3, N7, and N9 are bonded with cobalt. (b) A
unit cell of bio-MOF-11. The cavities are indicated by the green circles. Co: pink, O:
red, C: grey, H: white, N1: green, N6: blue, N3, N7, and N9: cyan.

To estimate the charges of bio-MOF-11 framework atoms, a fragmental cluster
(Figure 6.2) was cleaved and saturated by hydrogen atoms. The electrostatic
potentials around the cluster were calculated by density-functional theory (DFT). It
has been widely recognized that first-principles derived charges fluctuate appreciably
when small basis set is used; however, they tend to converge beyond 6-31G(d) basis
set.
331
Consequently, 6-31G(d) basis set was used in the DFT calculation for all the
atoms except Co metals, for which LANL2DZ basis set was used with effective
pseudopotentials. The DFT calculation used the Lee-Yang-Parr correlation functional
(B3LYP) and was carried out with Gaussian 03.
323
The concept of atomic charges is
solely an approximation and no unique straightforward method is currently available
to determine atomic charges rigorously. In this study, the atomic charges (Table 6.1)
were estimated by fitting to the electrostatic potential using the CHelpG scheme.
332
In
addition to the Coulombic interactions, the dispersion interactions of the framework
cavity
N6
N1
N3
N7
N9
(b)

cavity
(a)
Chapter 6. CO
2
Capture in Bio-MOF-11
85
atoms were presented by Lennard-Jones (LJ) potential with parameters adopted from
the Universal Force Field (UFF).
302
A number of simulation studies have
demonstrated that UFF can accurately predict gas adsorption in various
MOFs.
39,191,192,320,333



Figure 6.2. A fragmental cluster of bio-MOF-11 used to calculate atomic charges.
The dangling bonds (indicated by circles) were terminated by hydrogen atoms. Color
code: Co, pink; O, red; N, cyan; C, grey; H, white.

Table 6.1. Atomic charges in the fragmental cluster of bio-MOF-11.

No. 1 2 3 4 5 6 7 8 9
Atom Co O N1 N3 N6 N7 N9 C1 C2
q (e) 0.728 -0.749 -0.713 -0.613 -0.952 -0.096 -0.393 0.104 0.568
No. 10 11 12 13 14 15 16 17
18
Atom C3 C4 C5 C6 C7 H1 H2 H3 H4
q (e) 0.563 0.760 -0.310 0.964 -0.462 0.068 0.008 0.442 0.130

Adsorbate CO
2
was represented by the elementary physical model (EPM), which
was fitted to the experimental vapor-liquid equilibrium data of bulk CO
2
.
334
The CO
bond was assumed to be rigid at 1.161 , while the ZOCO bond was flexible and
1
2
4
3
6
5
8
9 10
1112
13
14
15
16
17
18
7
Chapter 6. CO
2
Capture in Bio-MOF-11
86
governed by a harmonic potential k
u
(uu
0
)
2
with force constant k
u
/k
B
= 153355.79
(K/rad
2
) and equilibrium angle u
0
= 180. The intermolecular CO
2
CO
2
interactions
were modeled by the additive pair-wise site-site LJ and Coulombic potentials as stated
in Chapter 3. Both H
2
and N
2
were mimicked by two-site models. The bond length
was 0.74 for H-H and 1.10 for N-N. The LJ potential parameters for H
2
and N
2

were fitted respectively to their experimental bulk properties.
335,336
H
2
O was
represented by the three-point transferable interaction potentials (TIP3P) model, in
which the O-H bond length was 0.9572 and the ZHOH angle was 104.52.
319
Table
6.2 lists the LJ potential parameters and charges for CO
2
, H
2
, N
2
, and H
2
O. The
adsorption of pure CO
2
, H
2
, and N
2
as well as CO
2
/H
2
and CO
2
/N
2
mixtures were
simulated by GCMC method.

Table 6.2. LJ Potential Parameters and Charges for CO
2
, H
2
, N
2
, and H
2
O.

Adsorbate Site () / k
B
(K) q (e)
CO
2

C 2.785 28.999 +0.6645
O 3.064 82.997 -0.33225
H
2
H 2.50 14.5 0
N
2
N 3.32 36.4 0
H
2
O
H 0 0 +0.417
O 3.151 76.47 -0.834

Chapter 6. CO
2
Capture in Bio-MOF-11
87
6.2 Results and Discussion
6.2.1 Pure Gases
Figure 6.3 shows the adsorption isotherms of pure CO
2
, H
2
, and N
2
in
bio-MOF-11. All the isotherms belong to type I, which is the characteristic feature of
adsorption in a microporous adsorbent. Perfect agreement is found between the
simulated and experiment results for CO
2
with deviation of less than 4.3%, indicating
the accuracy of the models and force fields used in the study. The deviations for H
2

and N
2
are a little bit larger, about 13% and 40%, respectively. On this basis, we
envision that the simulated selectivities of mixtures shown below are reliable, though
no experimental data are available for comparison.

P (kPa)
0 20 40 60 80 100
N

(
m
m
o
l
/
g
)
0
1
2
3
4
w
t
%
0
1
2
3
4
CO
2
H
2
N
2


Figure 6.3. Adsorption isotherms of pure CO
2
and N
2
at 298 K and of H
2
at 77 K,
respectively. The open symbols are from simulation and the filled symbols are from
experiment.

The lines are fits of the dual-site Langmuir-Freundlich equation to the
simulation data.

A closer look at Figure 6.3 reveals that simulation slightly underestimates
experiment particularly for H
2
at low pressures, but overestimates at high pressures.
Chapter 6. CO
2
Capture in Bio-MOF-11
88
The deviations might attribute to the factor that the structure used in simulation is a
perfect crystal, while experimental samples usually contain impurities. The impurities
in the channels cause stronger interaction with adsorbate and consequently enhance
adsorption at low pressures. Another effect of the impurities is that they block the
channels and decrease free volume, then reduce adsorption capacity at high pressures.
Furthermore, the simulated isotherms were fitted to the dual-site
Langmuir-Freundlich equation,
337

1 1 2 2
2 1
1 2
2 1
( )
1 1
n n
n n
N k P N k P
N P
k P k P
= +
+ +
(6.1)
where P is the pressure, N
i
is maximum loading in site i ( = 1 and 2), k
i
is the affinity
constant, and n
i
is used to characterize the deviation from the simple Langmuir
equation. The fitted parameters are listed in Table 6.3 and will be used to predict the
adsorption of binary mixtures by the ideal-adsorbed-solution theory (IAST).
338

Table 6.3. Parameters in the dual-site Langmuir-Freundlich equation fitted to the
adsorption of pure CO
2
, H
2
, and N
2
.

Adsorbate N
1
k
1
n
1
N
2
k
2
n
2

CO
2
3.809 8.76 10
-3

1.354 4.207

1.18 10
-2

0.665

H
2
8.002 3.67 10
-5

1.008 5.97 10
-4

9.59 10
-7

3.581

N
2
3.666 5.04 10
-4

1.059 2.865

9.59 10
-7

0.702


CO
2
capacity in bio-MOF-11 is 4.1 mmol/g at 100 kPa, about 1-3 times higher
than that in most MOFs (e.g. MOF-2, MOF-177, MOF-505, IRMOF-1, -3, -6 and
-11).
125
Bio-MOF-11 also outperforms many zeolites and activated carbons,
339

amine-functionalized and imidazole-based MOFs in terms of CO
2
capacity.
168
H
2

Chapter 6. CO
2
Capture in Bio-MOF-11
89
adsorption in bio-MOF-11 is notably high with a capacity of 1.5 wt% at 77 K and 1
bar, which is higher than that in numerous MOFs (e.g. MOF-177, ZIF-8, MIL-100,
IRMOF-2, -3, -9, -18, and -20).
58
The high capacities observed in bio-MOF-11 are
attributed to the narrow channels and the Lewis basic sites present in the framework.
The channels have narrow diameter of 4.2 ~ 5.2 , leading to a substantial overlap of
the potential energy fields for adsorbate molecules. The linker adenine consists of
two Lewis basic sites, one amino group and the other pyrimidine nitrogen. A total of
four amino groups and four pyrimidine nitrogens are directly exposed to each
cavity.
168
These electron-rich sites interact strongly with adsorbate, particularly CO
2

that has a quadrupole moment.
To identify the favorable adsorption sites of CO
2
in bio-MOF-11, Figure 6.4
shows the radial distribution functions g(r) of CO
2
around N1, N6, and Co atoms at
298 K and 10 kPa. N1 and N6 are in the pyrimidine and amino groups, respectively.
Co
r
2 4 6 8 10 12
g

(
r
)
0.0
0.5
1.0
1.5
2.0
2.5
()
N1 N6


Figure 6.4. Radial distribution functions of CO
2
around N1, N6, and Co atoms in
bio-MOF-11 at 298 K and 10 kPa. N1 and N6 are in the pyrimidine and amino groups,
respectively.

Chapter 6. CO
2
Capture in Bio-MOF-11
90
A pronounced peak is centered at r ~ 4.0 in the g(r) around both N1 and N6. The
g(r) around Co is essentially zero at r < 4.0 and exhibits a lower peak at r ~ 8.0 .
This structural analysis reveals that CO
2
molecules are preferentially adsorbed onto
the Lewis basic sites in adenine linkers with the shortest distance approximately 2.7 ,
rather than proximal to the Co atoms. Such a behavior is remarkably different from
common MOFs, for example, the preferentially sites in IRMOF-1 were found to be
close to metal clusters.
340
Nevertheless, the behavior is similar to ZIF-8 in which the
adsorption sites are near 2-methylimidazolate linkers.
341
We therefore infer that the
biomolecular linkers in bio-MOF-11 should be tuned in order to further enhance
adsorption capacity of CO
2
.


Figure 6.5. (a) Simulation snapshot and (b) density contour of CO
2
in bio-MOF-11 at
298 K and 10 kPa. CO
2
molecules are represented by sticks. The density has a unit of
1/
3
and brighter color indicates a higher density. Co: pink, O: red, C: grey, H: white,
N1: green, N6: blue, N3, N7, and N9: cyan.

Figure 6.5 shows the simulation snapshot and density contour of CO
2
adsorption
in bio-MOF-11 at 298 K and 10 kPa. For clarity, only one cavity in the framework is
illustrated. Apparently, CO
2
molecules are located dominantly in the cavity and
adsorbed onto the amino and pyrimidine groups. A few CO
2
molecules seem to be
(a)
(b)
Chapter 6. CO
2
Capture in Bio-MOF-11
91
located in the diagonal regions of the cavity. However, they are indeed in a cavity
along the perpendicular dimension, as can been seen by rotating Figure 5.5 by 90.
6.2.2 CO
2
/H
2
Mixture
The adsorption of CO
2
/H
2
mixture with a composition of 15:85 was simulated in
bio-MOF-11 to represent pre-combustion CO
2
capture. The composition is typically
encountered for CO
2
/H
2
mixture in H
2
production. As shown in Figure 5.6, both CO
2

and H
2
in the mixture are preferentially located in the cavities. However, there are two
major differences between CO
2
and H
2
. First, H
2
has a much lower extent of
adsorption and therefore a lower density in the cavities. Second, H
2
is smaller and can
approach the adenine linkers more closely than CO
2
, as observed in the broader
distribution of H
2
.



Figure 6.6. Density contours of CO
2
and H
2
for CO
2
/H
2
mixture (15:85) in
bio-MOF-11 at 298 K and 100 kPa. The density has a unit of 1/
3
. The density
distributions are largely similar to Figure 5.5b for pure CO
2
.

Figure 6.7a shows the adsorption isotherm at 298 K as a function of total pressure.
CO
2
uptake increases sharply with increasing pressure and then approaches to a
plateau, while H
2
uptake is vanishingly small. Over the entire range of pressure, CO
2

H2 CO2
Chapter 6. CO
2
Capture in Bio-MOF-11
92
is predominantly adsorbed than H
2
due to three reasons. First, CO
2
is a three-site
molecule and has a stronger interaction with the framework than two-site H
2
. Second,
the temperature 298 K considered is subcritical for CO
2
(T
c
= 304.4 K), but
supercritical for H
2
(T
c
= 33.2 K); that is, CO
2
is more condensable than H
2
at 298 K.
Third, the presence of Lewis basic sites in adenines significantly enhance the
interactions with CO
2
.
P
total
(kPa)
0 500 1000 1500 2000 2500 3000
N

(
m
m
o
l
/
g
)
0
1
2
3
4
5
6
CO
2

H
2
CO
2
(0.1% H
2
O)

H
2
(0.1% H
2
O)

CO
2
(a)
H
2

P
total
(kPa)
0 500 1000 1500 2000 2500 3000
S
C
O
2
/
H
2
200
250
300
350
400
dry CO
2
/H
2
wet CO
2
/H
2
(b)


Figure 6.7. (a) Adsorption isotherm and (b) selectivity of CO
2
/H
2
mixture (15:85) in
bio-MOF-11 as a function of total pressure in the absence and presence of 0.1 % H
2
O.

Figure 6.7b shows the selectivity of CO
2
over H
2
as a function of total pressure.
With increasing pressure, the selectivity increases sharply, passes a maximum of 375
at 400 kPa, and then decreases. It is expected that the selectivity will reach a constant
upon further increasing pressure. The initial increase at low pressures is caused by the
strong interactions between CO
2
molecules and the multiple adsorption sites in
bio-MOF-11, and further promoted by the cooperative intermolecular interactions of
adsorbed CO
2
molecules. The decrease in selectivity is attributed primarily to the
entropy (packing) effect at high pressures. H
2
molecule is smaller in size and can fit
Chapter 6. CO
2
Capture in Bio-MOF-11
93
into the channels more effectively. In the pressure range under study, the selectivity of
CO
2
/H
2
mixture in bio-MOF-11 is between 230 and 375. Numerous experimental and
simulation have examined the separation of CO
2
/H
2
mixture in other porous materials,
as listed in Table 6.4 for comparison. While bio-MOF-11 has a lower CO
2
capacity
than other materials, it exhibits a higher selectivity compared with zeolites and
non-ionic MOFs.

Table 6.4. Selectivities and capacities for the adsorption of CO
2
/H
2
mixture in porous
materials. The capacities are for CO
2
at a total pressure of 1 bar for mixture.

Material
Temperature,
Pressure
Composition
(CO
2
:H
2
)
Selectivity
Capacity
(mmol/g)
ETS-10
230
298 K, 10 bar 50:50 3.5 2.45
MFI
215
298 K, 10 bar 50:50 5.0 2.95
Na-4A
342
298K, low pressure 1.4:98.6 70.65 1.3
IRMOF-n
(n = 9 ~ 14)
343

298 K, 0 ~ 2 MPa
5:95, 30:70,
50:50, 70:30,
95:5
10 ~ 100
Cu-BTC
344
298 K, 0 ~6 MPa 50:50 70 ~ 150
soc-MOF
247
298 K, 0 ~ 3000 kPa 15:85 300 ~ 600 2.7
rho-ZMOF
42
298 K, 0 ~ 3000 kPa 15:85
1.6 10
5
~
200
4.3
bio-MOF-11 298 K, 0 ~ 3000 kPa 15:85 230 ~ 375 1.3

A gas mixture usually contains moisture which would affect adsorption and
separation. For example, in the presence of H
2
O, the selectivity of CO
2
over H
2
in
soc-MOF increases at low pressures due to the promoted adsorption of CO
2
by H
2
O
bound onto metal atoms, but decreases at high pressures as a result of the competitive
adsorption of H
2
O over CO
2
.
247
In Na-rho-ZMOF, the interaction between CO
2
and
Na
+
is substantially reduced by a trace amount of H
2
O added into CO
2
/CH
4
mixture;
consequently, CO
2
adsorption drops and the selectivity decreases by one order of
Chapter 6. CO
2
Capture in Bio-MOF-11
94
magnitude.
329
To examine the effect of H
2
O in this study, the adsorption of CO
2
/H
2

mixture was simulated in the presence of 0.1% H
2
O (mole fraction). As seen in Figure
6.7, H
2
O has a negligible effect on the adsorption of both CO
2
and H
2
; and the CO
2
/H
2

selectivity is slightly enhanced due to co-adsorption between CO
2
and adsorbed H
2
O.
The predictions from IAST are shown in Figure 5.8 for CO
2
/H
2
mixture using
the parameters given in Table 6.3. While the adsorption isotherm is well predicted
upon comparison with simulation, the selectivity is underestimated by IAST. In IAST
prediction, even a small inaccuracy in the mole fractions would lead to a large
deviation in the selectivity. The deviation of 23% between IAST and GCMC
simulation may be attributed to the non-ideal gas mixture resulted from the different
polarities of CO
2
and H
2
.
345


P
total
(kPa)
0 500 1000 1500 2000 2500 3000
S
C
O
2
/
H
2
200
250
300
350
400
Sim.
IAST
P
total
(kPa)
0 500 1000 1500 2000 2500 3000
N

(
m
m
o
l
/
g
)
0
1
2
3
4
5
6
CO
2
(Sim.)
H
2
(Sim.)
CO
2
(IAST)
H
2
(IAST)
CO
2
(a) (b)
H
2

Figure 6.8. (a) Adsorption isotherm and (b) selectivity of CO
2
/N
2
mixture (15:85) in
bio-MOF-11 as a function of total pressure. The open symbols are from simulation
and the filled symbols are from IAST.

6.2.3 CO
2
/N
2
Mixture
To examine post-combustion CO
2
capture by bio-MOF-11, the adsorption of
CO
2
/N
2
mixture with a composition of 15:85 was simulated. Figure 6.9a shows the
Chapter 6. CO
2
Capture in Bio-MOF-11
95
adsorption isotherm of the mixture at 298 K as a function of the total pressure. Similar
to Figure 6.7a, CO
2
uptake increases sharply with increasing pressures and reaches
saturation at high pressures. In contrast, N
2
uptake only increases slightly at low
pressures. CO
2
has a substantially greater adsorption than N
2
, because of the three
reasons mentioned above for CO
2
/H
2
mixture. Nevertheless, the selectivity of CO
2
/N
2

versus pressure is qualitatively different from that of CO
2
/H
2
. As shown in Figure
6.9b, the selectivity increases monotonically with increasing pressure and approaches
a constant at high pressures. The increase of selectivity is due to the stronger
CO
2
-CO
2
intermolecular interactions as pressure increases. The decrease of selectivity
seen in Figure 6.7b for CO
2
/H
2
is not observed for CO
2
/N
2
. This is because the
molecular sizes of CO
2
and N
2
are not significantly different, unlike CO
2
and H
2
in
which H
2
is much smaller than CO
2
; therefore, the entropy effect is not dominant for
CO
2
/N
2
mixture at high pressures.
P
total
(kPa)
0 500 1000 1500 2000 2500 3000
N

(
m
m
o
l
/
g
)
0
1
2
3
4
5
CO
2
N
2
CO
2
(0.1% H
2
O)

N
2
(0.1% H
2
O)

(a)
CO
2
N
2

P
total
(kPa)
0 500 1000 1500 2000 2500 3000
S
C
O
2
/
N
2
20
40
60
80
100
dry CO
2
/N
2
wet CO
2
/N
2
(b)

Figure 6.9. (a) Adsorption isotherm and (b) selectivity of CO
2
/N
2
mixture (15:85) in
bio-MOF-11 as a function of total pressure in the absence and presence of 0.1 % H
2
O.

Chapter 6. CO
2
Capture in Bio-MOF-11
96
CO
2
/N
2
separation has been investigated in other nanoporous materials as listed
in Table 6.5. Upon comparison, bio-MOF-11 has a selectivity of 30 ~ 77, exhibiting a
better separation capability than zeolites and non-ionic MOFs. We note that the
selectivities of both CO
2
/H
2
and CO
2
/N
2
mixtures in bio- MOF-11 are smaller than in
rho-ZMOF.
42
This is not unexpected because rho-ZMOF consists of a highly ionic
framework and extraframework cations, which induce substantially strong
interactions with CO
2
molecules and exceptionally high selectivities.

Table 6.5. Selectivities and capacities for the adsorption of CO
2
/N
2
mixture in porous
materials. The capacities are for CO
2
at a total pressure of 1 bar for mixture.

Material
Temperature,
Pressure
Composition
(CO
2
:N
2
)
Selectivity
Capacity
(mmol/g)
Na-4A
342
298 K, low pressure 5.1:94.9 18.75 2.1
silicalite, ITQ-3,
ITQ-7
346

308 K, P > 10 bar 50:50, 10:90 30, 100, 10 0.5, 2.4, 1.6
MFI
347
300 K, 420 kPa 50:50 13.7
FAU
348
298 K, 1 bar 50: 50 20
MOF-508b
349
303 K, 1 ~ 4 bar 50:50 3 ~ 6 1.58
zinc-paddle
wheel MOFs
170

298 K, 0 ~ 8 bar 15:85 25~45
Cu-BTC
250
298 K, 1 ~ 5 MPa 15.6:84.4 20 ~ 38
rho-ZMOF
42
298 K, 3000 kPa 15:85 1.9 10
4
~ 100 4.4
bio-MOF-11 298 K, 3000 kPa 15:85 30 ~ 77 1.2

The effect of H
2
O on the adsorption of CO
2
/N
2
mixture is also simulated. As
seen in Figure 6.9, the addition of H
2
O slightly increases the selectivity of CO
2
/N
2
,
more obviously at high pressures. The performance of IAST for CO
2
/N
2
mixture
shown in Figure 6.10 is similar to that for CO
2
/H
2
mixture. While the adsorption
isotherm is well predicted, the selectivity is only in qualitative agreement with
simulation.
Chapter 6. CO
2
Capture in Bio-MOF-11
97

P
total
(kPa)
0 500 1000 1500 2000 2500 3000
S
C
O
2
/
N
2
20
30
40
50
60
70
80
Sim.
IAST
P
total
(kPa)
0 500 1000 1500 2000 2500 3000
N

(
m
m
o
l
/
g
)
0
1
2
3
4
5
CO
2
(Sim.)
N
2
(Sim.)
CO
2
(IAST)
N
2
(IAST)
(a) (b)
CO
2
N
2

Figure 6.10. (a) Adsorption isotherm and (b) selectivity of CO
2
/N
2
mixture (15:85) in
bio-MOF-11 as a function of total pressure at 298 K. The open symbols are from
simulation and the filled symbols are from IAST.

6.3 Summary
We have reported the first molecular simulation study for CO
2
capture in a
bio-metal organic framework (bio-MOF-11). The bio-MOF-11 consists of
paddle-wheel cobalt-adeninate-acetate clusters and interlacing narrow channels. The
amino and pyrimidine groups are the Lewis basic sites, preferentially for adsorption.
This contrasts remarkably to most MOFs in which the metal clusters are the favorable
sites. To further facilitate CO
2
adsorption in bio-MOF-11, we suggest that the adenine
linkers rather than the metals should be functionalized. Because of the Lewis basic
sites and the narrow channels in the framework, bio-MOF-11 exhibits large capacities
of CO
2
and H
2
, and indeed larger than many zeolites, activated carbons, and MOFs.
The simulated adsorption isotherms of CO
2
, H
2
, and N
2
agree very well with
experimental data, which demonstrates the accuracy of the models and potential used.
In the attempt to evaluate the capability of bio-MOF-11 for pre- and
post-combustion CO
2
capture, it was found that CO
2
is more favourably adsorbed
Chapter 6. CO
2
Capture in Bio-MOF-11
98
than H
2
and N
2
. As a function of pressure, the selectivities of CO
2
/N
2
and CO
2
/H
2

mixtures behave quantitatively different. Due to the strong interactions between CO
2

and framework, the selectivity of CO
2
/H
2
at low pressures increases with increasing
pressure. At high pressures, however, it decreases because H
2
has a smaller molecular
size and the entropy effect comes into play. The selectivity of CO
2
/N
2
increases
monotonically at low pressures and gradually approaches a constant. The selectivities
of both mixtures in bio-MOF-11 are higher than in most zeolites and non-ionic MOFs.
In the presence of H
2
O (0.1% in mole fraction), the selectivities are slightly enhanced,
particularly for CO
2
/N
2
mixture at high pressures. This study suggests that
adenine-based bio-MOFs might be useful for CO
2
capture. The microscopic insight
from molecular simulation is important for the quantitative understanding of
adsorption mechanism and the rational design of new bio-MOFs for emerging
applications.




Chapter 7. CO
2
Adsorption in Cation-Exchanged MOFs
99
Chapter 7. CO
2
Adsorption in Cation-Exchanged MOFs
In this Chapter, a simulation study is reported to investigate the adsorption of CO
2
,
as well as CO
2
/H
2
mixture, in cation-exchanged rho-ZMOFs. The cations considered
include mono-, di- and trivalent Na
+
, K
+
, Rb
+
, Cs
+
, Mg
2+
, Ca
2+
and Al
3+
. Following
this section, the models used for rho-ZMOF framework, cations and adsorbates are
briefly described in 7.1. The simulation methods are introduced in 7.2. In section 7.3,
the cations in rho-MOFs are first characterized, and then the isosteric heat and
Henrys constant of CO
2
adsorption at infinite dilution are presented. In addition,
adsorption isotherm and adsorbate structure are examined. Finally, the adsorption of
CO
2
/H
2
mixture is discussed, along with the effect of H
2
O.
7.1 Models
As the first example of 4-connected MOF with rho topology, rho-ZMOF was
synthesized by the directed assembly of In atoms and H
3
ImDC.
18
It has a space group
of Im-3m and a lattice constant of 31.062. In rho-ZMOF, each In atom is
coordinated to four N atoms and four O atoms of four separate doubly deprotonated
H
3
ImDC (HImDC) ligands to form an eight-coordinated dodecahedron. Each
independent HImDC is coordinated to two In atoms via N-, O-hetero-chelation
resulting in two rigid five-membered rings. As illustrated in Figure 7.1, rho-ZMOF
structure contains truncated cuboctahedron (o-cage) linked via double
eight-membered ring (D8MR). The substitution of oxygen in rho-zeolite by HImDC
generates a very open structure with extra-large cavity of 18.2 in diameter. Unlike
Chapter 7. CO
2
Adsorption in Cation-Exchanged MOFs
100
rho-zeolite and other rho-aluminosilicate or aluminophosphate, rho-ZMOF possesses
twice as many positive charges (48 vs. 24) in a unit cell to neutralize the anionic
framework. As mentioned, the HPP cations in parent rho-ZMOF can be exchanged
with other cations. In addition, experimental thermogravimetric analysis showed that
all residential water molecules in rho-ZMOF can be completely evacuated.
18



Figure 7.1. Crystal structure of rho-ZMOF. Color code: In, cyan; N, blue; C, grey; O,
red; and H, white. The a-cage, double eight-membered ring (D8MR), 6-membered
ring (6MR) and 4-membered ring (4MR) are indicated. The yellow spheres in the
D8MR represent inaccessible cages.

In this study, seven cation-exchanged rho-ZMOFs are considered with
monovalent Na
+
, K
+
, Rb
+
, Cs
+
, divalent Mg
2+
, Ca
2+
and trivalent Al
3+
. The cations
were modeled as charged Lennard-Jones (LJ) particles with parameters listed in Table
7.1. The LJ potential parameters were adopted from the Universal Force Field
(UFF).
302



o-cage
6MR
4MR
Chapter 7. CO
2
Adsorption in Cation-Exchanged MOFs
101
Table 7.1. Charges Z, well depths c /k
B
and collision diameters of cations.

Cation Z(e)
*
c/k
B
(K) o ()
Na
+
+1 15.083 2.658
K
+
+1 17.597 3.396
Rb
+
+1 20.111 3.665
Cs
+
+1 22.624 4.024
Mg
2+
+2 55.807 2.691
Ca
2+
+2 119.658 3.028
Al
3+
+3 253.897 4.008


Figure 7.2. Atomic charges in a fragmental cluster of rho-ZMOF. Color code: In,
cyan; N, blue; C, grey; O, red; and H, white.

The atomic charges of rho-ZMOF framework atoms were calculated from
density-functional theory (DFT) using a fragmental cluster (in Figure 7.2) and same
method as stated in Chapter 6. The atomic charges were estimated by fitting to the
Electrostatic Potentials (ESP). In addition, the van der Waals interactions of the
framework atoms were mimicked by LJ potential with parameters from the UFF (in
Table 7.2).


In (2.840)
N (-0.381)
C1 (0.950)
O2 (-0.813)
O1 (-0.751)
C2 (-0.257)
C3 (-0.525)
H2 (0.683)
H1 (0.426)
Chapter 7. CO
2
Adsorption in Cation-Exchanged MOFs
102
Table 7.2. Lennard-Jones parameters of framework atoms in rho-ZMOF.

Atom () c /k
B
(K)
In 3.976 301.157
N 3.260 34.690
O 3.118 30.166
C 3.431 52.790
H 2.571 22.122

CO
2
was modeled by a three-site rigid molecule and its intrinsic quadrupole
moment was described by a partial charge model.
317
The partial charges on C and O
atoms were q
C
= 0.576e and q
O
= 0.288e (e = 1.6022 10
-19
), respectively. The CO
bond length was 1.18 and the bond angle ZOCO was 180. The intermolecular
interactions of CO
2
were represented by a combination of LJ and Coulombic
potentials. In addition, the adsorption of CO
2
/H
2
mixture was also examined. H
2
was
mimicked as a three-site model with explicit charge.
350
The H-H bond length was 0.74
and the charge on H atom were +0.468e. The center-of-mass between the two H
atoms was a LJ core with a charge of 0.936e. This model gave a quadrupole moment
of 2.05 10
-40
Cm
2
for H
2
.
7.2 Methods
In each cation-exchanged rho-ZMOF with Na
+
, K
+
, Rb
+
, Cs
+
, Mg
2+
, Ca
2+
or Al
3+
,
the locations of cations were determined by simulated annealing in a canonical
ensemble. The initial temperature was 600 K and decreased to 300 K with an interval
of 20 K. The simulation box contained one unit cell of rho-ZMOF with appropriate
Chapter 7. CO
2
Adsorption in Cation-Exchanged MOFs
103
number of cations (48 Na
+
, K
+
, Rb
+
, Cs
+
, 24 Mg
2+
, Ca
2+
or 16 Al
3+
). The framework
was assumed to be rigid during simulation and periodic boundary conditions were
applied in three dimensions. The unit cell was divided into fine grids with energy
landscape tabulated in advance and then used by interpolation during simulation. In
such a way, the simulation was accelerated by two orders of magnitude. A spherical
cutoff of 15.0 was used to evaluate the LJ interactions with long-range corrections
added. For the Coulombic interactions, a simple spherical truncation could result in
significant errors; consequently, the Ewald sum with a tin-foil boundary condition was
used. The real/reciprocal space partition parameter and the cutoff for reciprocal lattice
vectors were chosen to be 0.2
-1
and 8, respectively, to ensure the convergence of the
Ewald sum. These methods to calculate the LJ and Coulombic interactions were also
used in the simulation described below. The cations were introduced into the box
randomly and followed by 10
7
trial moves. Specifically, two types of trial moves
including translation and regrowth were used with equal probability. After equilibrium,
the free volume in each cation-exchanged rho-ZMOF was estimated using helium as a
probe

He
ad B free
exp[ ( ) / ] d r r
V
V u k T = -

(7.1)
where
He
ad
u is the interaction between helium and framework,
He
2.58 s = and
B He
/ 10.22 K k e = .
351

The isosteric heat and Henrys constant of CO
2
adsorption at infinite dilution and
298 K were calculated by Monte Carlo (MC) simulation in a canonical ensemble. A
single gas molecule was added into rho-ZMOF and subject to three types of trial
Chapter 7. CO
2
Adsorption in Cation-Exchanged MOFs
104
moves, namely translation, rotation and regrowth. As shown in Figure 7.1, there exist
quasi-spherical inaccessible cages within the D8MR. They are isolated from the
-cage and inaccessible by gas molecules, thus were blocked during simulation.
Nevertheless, it should be point out that these cages are very small and have
negligible effect on the adsorption. It was found that a simulation without blocking
these cages gave, within statistical uncertainty, nearly identical results. Specifically,
the isosteric heat was evaluated by
o o
st
a
Q RT U =
(7.2)
where
o
a
U is the ensemble averaged adsorption energy of a gas molecule with
framework. The Henrys constant
H
K was evaluated by
H a
exp[ ( , )] d d u
K
= =
| |

=
}
r r
(7.3)
where
a
( , ) u = r is the adsorption energy for a gas molecules at position r and
orientation
=
.
The adsorption isotherms of CO
2
or CO
2
/H
2
mixture at 298 K were simulated
using GCMC method stated in Chapter 3. CO
2
/H
2
mixture was assumed to have a
bulk composition of 15/85, which is practically found in H
2
production. A gas mixture
usually contains moisture or the separation process might be operated under humid
condition. To examine the effect of H
2
O, the adsorption of CO
2
/H
2
/H
2
O mixture was
also simulated with 0.1% of H
2
O (mole fraction). The cations were allowed to move
during the GCMC simulation.
7.3 Results and Discussion
First, the cations in rho-ZMOFs are characterized, particularly in terms of their
Chapter 7. CO
2
Adsorption in Cation-Exchanged MOFs
105
equilibrium locations. Then, the isosteric heat and Henrys constant of CO
2
adsorption
are presented and related with the charge-to-diameter ratio of cation. The adsorption
isotherm, density contour and radial distribution function of CO
2
molecules are shown
as a function of pressure. Finally, the adsorption selectivities of CO
2
/H
2
and
CO
2
/H
2
/H
2
O mixtures are examined.
7.3.1 Characterization of cations
The equilibrium locations of cations are tabulated in appendix (Tables A1-A7),
along with the simulation snapshots. For cations with identical valence (e.g. Na
+
, K
+
,
Rb
+
and Cs
+
; Mg
2+
and Ca
2+
), the locations are largely similar. As identified in our
previous study,
42
there are two types of favorable sites for Na
+
ions in Na-rho-ZMOF.
Site I is in the single eight-membered ring (S8MR) and at the entrance to the o-cage.
Site II is in the o-cage and proximal to the moiety of organic link. Compared with site
II, site I has a larger coordination number with neighboring atoms in the S8MR and
thus a stronger interaction with the framework. It was also revealed that the mobility
of cations is very small and can be regarded as local vibration at the favorable sites.
42




Figure 7.3. Equilibrium and initial locations of cations (a) K
+
(b) Ca
2+
(c) Al
3+
. The
initial locations are indicated in pink.

(a) (b) (c)
Chapter 7. CO
2
Adsorption in Cation-Exchanged MOFs
106

Table 7.3. Porosity, isosteric heat and Henrys constant of CO
2
adsorption in
rho-ZMOFs.

MOF Porosity

o
st
(kJ/mol) Q
3
H
(mmol/cm kPa) K
Na-rho-ZMOF

0.548 58.25 50.37
K-rho-ZMOF
0.512 53.99 9.52
Rb-rho-ZMOF
0.491 51.73 5.98
Cs-rho-ZMOF
0.466 40.27 0.82
Mg-rho-ZMOF
0.572 84.42 1.66 10
5

Ca-rho-ZMOF
0.566 73.71 2.64 10
4

Al-rho-ZMOF

0.559 89.91 2.78 10
6


It is interesting to examine how the equilibrium locations shift from the initial
locations at which cations are randomly introduced into simulation cell. Figure 7.3
illustrates typically the equilibrium and initial locations of three cations (K
+
, Ca
2+
and
Al
3+
). Apparently, there are deviations between the two locations. The initial positions
are in a random pattern, whereas the equilibrium locations are more ordered at the
favorable sites.
For each cation-exchanged rho-ZMOF, the free volume was estimated and
porosity was calculated. By definition, the porosity | is equal to the ratio of free
volume to specific volume,
free
/ V V | = . As indicated in Table 7.3, | decreases with
increasing cation diameter for cations with identical valence. To further quantify,
Figure 7.4 shows | versus the packing fraction of cation q. The latter is defined as
3
cation cation
( / ) N V q o = , where
cation
N is the number of cations. Clearly, | decreases
monotonically with increasing q. A good correlation is found between | and q
0.59 1.23q | = (7.4)
Chapter 7. CO
2
Adsorption in Cation-Exchanged MOFs
107
With this correlation, the porosity can be predicted for other cation-exchanged
rho-ZMOFs.

| = 0.59 1.23q
0.40
0.45
0.50
0.55
0.60
0.00 0.02 0.04 0.06 0.08 0.10 0.12
q
cation
P
o
r
o
s
i
t
y

|

Figure 7.4. Porosity | versus the packing fraction of cation q in rho-ZMOFs. The
solid line is a linear correlation between | and q.

7.3.2 Isosteric Heat and Henrys Constant
The isosteric heats
o
st
Q

and Henrys constants K
H
for CO
2
adsorption at infinite
dilution are listed in Table 7.3. Due to the strong electrostatic interactions between
CO
2
and cation/ionic framework, the
o
st
Q and K
H
are substantially larger than those in
neutral MOFs.
39
The
o
st
Q

in Na, K, Rb and Cs-rho-ZMOFs are comparable to
experimentally measured
o
st
Q in Na-ZSM5 (50 kJ/mol)
352
and Na-MOR (65
kJ/mol).
353
Intriguingly, the
o
st
Q

and K
H
vary drastically with different cations. For
monovalent cations (Na
+
, K
+
, Rb
+
and Cs
+
), the
o
st
Q

and K
H
decrease in the order of
Na
+
> K
+
> Rb
+
> Cs
+
, i.e., with increasing cation diameter. This is also observed for
divalent cations Mg
2+
> Ca
2+
. In addition, the
o
st
Q

and K
H
in Mg-rho-ZMOF and
Chapter 7. CO
2
Adsorption in Cation-Exchanged MOFs
108
Ca-rho-ZMOF are significantly larger than in monovalent cation-rho-ZMOFs. Among
the seven rho-ZMOFs, Al-rho-ZMOF exhibits the highest
o
st
Q

and K
H
. Such a trend
was previously found for adsorption in cation-exchanged zeolites. The enthalpy of
CO
2
adsorption in alkali cation-exchanged zeolite-Y follows Li
+
> Na
+
> K
+
> Cs
+
.
354

In monovalent and divalent zeolite-X, the enthalpy of N
2
decreases as Li
+
> Na
+
> K
+

and Mn
2+
> Ca
2+
> Sr
2+
> Ba
2+
.
355
In zeolite-X and MOR, the enthalpy and Henrys
constant of N
2
and Ar also decreases with increasing cation diameter.
356-358

Figure 7.5 shows the behavior of
o
st
Q

and K
H
versus the charge-to-diameter ratio
Z/o
cation
of cation, which can be used to quantify the electric field generated by cation.
Both
o
st
Q

and K
H
increase monotonically with increasing Z/o
cation
as attributed to the
strong electrostatic interactions between CO
2
and cations as well as ionic framework.

Z/o
cation
0.2 0.3 0.4 0.5 0.6 0.7 0.8
Q
s
t

(
k
J
/
m
o
l
)
10
30
50
70
90
Z/o
cation
0.2 0.3 0.4 0.5 0.6 0.7 0.8
l
o
g

(
K
H
)
0
2
4
6
o
Cs
+
Rb
+
K
+
Na
+
Ca
2+
Mg
2+
Al
3+
Cs
+
Rb
+
K
+
Na
+
Ca
2+
Mg
2+
Al
3+
(a) (b)

Figure 7.5. (a) Isosteric heats and (b) Henrys constants for CO
2
adsorption in
rho-ZMOFs versus the charge-to-diameter ratio of cation. The dotted lines are to
guide the eye.

Figure 7.6 shows the adsorption isotherms of CO
2
in cation-exchanged
rho-ZMOFs at 298 K. At low pressures (< 100 kPa), CO
2
uptake increases in the
sequence Cs
+
< Rb
+
< K
+
< Na
+
< Ca
2+
< Mg
2+
< Al
3+
, despite small difference
Chapter 7. CO
2
Adsorption in Cation-Exchanged MOFs
109
between Mg- and Al-rho-ZMOFs. This is analogous to the behavior of
o
st
Q

and K
H
,
following the increasing order of Z/
cation
. It implies that electrostatic interactions
predominantly govern CO
2
adsorption at low pressures. Similar trend was reported by
a number of studies in cation-exchanged zeolites. For example, in zeolite-X
exchanged with monovalent and divalent cations, N
2
and Ar exhibit increasing
capacity K
+
< Na
+
< Li
+
and Ba
2+
< Sr
2+
< Ca
2+
< Mn
2+
.
355,358
The capacities of N
2
, O
2

and CO in cation-exchanged zeolite-X and MOR increase in the sequence of Cs
+
< K
+

< Na
+
< Li
+
.
356,357
CO
2
capacity in zeolite-X increases as Cs
+
< Rb
+
< K
+
< Na
+
< Li
+
,
and also in zeolite-Y.
359
In alkali cation-exchanged chabazite, CO
2
capacity in
K-CHA is 78.7% and 70.1% of those in Na-CHA and Li-CHA, respectively.
360
At
high pressures, CO
2
uptake largely follows the trend at low pressures, i.e., increases

P (kPa)
0 20 40 60 80 100
C
O
2

(
m
m
o
l
/
g
)
0
2
4
6
8
10
Na
+
K
+
Rb
+
Cs
+
Mg
2+
Ca
2+
Al
3+
P (kPa)
0 200 400 600 800 1000
C
O
2

(
m
m
o
l
/
g
)
0
2
4
6
8
10
12
Na
+
K
+
Rb
+
Cs
+
Mg
2+
Ca
2+
Al
3+
(a) (b)

Figure 7.6. Adsorption isotherms of CO
2
in rho-ZMOFs (a) low-pressure regime and
(b) high pressure regime.

with increasing Z/
cation
. However, Mg-rho-ZMOF exhibits a marginally higher
saturation capacity than Al-rho-ZMOF. This is attributed to the free volume of
framework. As listed in Table 7.3, the porosity in Mg-rho-ZMOF is 0.572, slightly
Chapter 7. CO
2
Adsorption in Cation-Exchanged MOFs
110
larger than that in Al-rho-ZMOF (0.559). Consequently, Mg-rho-ZMOF has a larger
free volume available for adsorption and a higher saturation capacity. The effect of
free volume for CO
2
adsorption is not obvious at low pressures because electrostatic
interactions play a major role, but becomes increasingly more important at high
pressures, particularly near saturation.
To examine how the locations of CO
2
molecules adsorbed in the framework shift
with pressure, Figure 7.7 shows the density contours of CO
2
in Na-rho-ZMOF at 10,
100 and 1000 kPa. At a low pressure (10 kPa), CO
2
molecules are mostly located in
the 6-membered ring (6MR), but also in the -cage and 4-membered ring (4MR). In
particularly, it is observed that CO
2
molecules are adsorbed proximately to Na
+
ions.
Therefore, Na
+
ions can be regarded as preferential sites for CO
2
adsorption due to
strong electrostatic interactions. In addition, the cations are located slightly different
at 10, 100 and 1000 kPa. This implies that the locations of cations are shifted upon
CO
2
adsorption.

10 kPa 100 kPa 1000 kPa


Figure 7.7. Density contours of CO
2
in Na-rho-ZMOF at 10, 100 and 1000 kPa. The
locations of Na
+
ions are indicted by the large spheres. The density scale is the
number of CO
2
molecules per
3
.



Chapter 7. CO
2
Adsorption in Cation-Exchanged MOFs
111
Figure 7.8a shows the g(r) for CO
2
around Na
+
ions, N and In atoms in
Na-rho-ZMOF at 10 kPa. The peak position in the g(r) of CO
2
-Na
+
is at r = 3.6 ,
shorter than those of CO
2
-N and CO
2
-In. In addition, the peak height of CO
2
-Na
+
is
more pronounced than the other two. This confirms that Na
+
ions are preferential sites
for CO
2
adsorption. With increasing pressure to 100 and then to 1000 kPa, Figure 7.7
shows that more CO
2
molecules are adsorbed in the a-cage and Na
+
ions appear to be
solvated by CO
2
. Thus, it can be expected the electrostatic interactions between CO
2

and Na
+
ions are largely reduced. Figure 6.8b compares the g(r) of CO
2
-Na
+
at 10,
100 and 1000 kPa. With increasing pressure, the peak height in g(r) drops as a
consequence of the fact that more CO
2
molecules are adsorbed further away from the
Na
+
ions. Although not shown here, it was observed in our previous study that the
distance between CO
2
molecules becomes shorter with increasing pressure.
42
This
indicates an enhancement in cooperative interactions between CO
2
molecules.

r (A)
2 3 4 5 6 7 8
g
(
r
)
0
1
2
3
4
5
6
In-CO
2
N-CO
2
Na
+
-CO
2
(a) 10 kPa

r (A)
2 3 4 5 6 7 8
g
(
r
)
0
1
2
3
4
5
6
1000 kPa
100 kPa
10 kPa
(b) Na
+
-CO
2


Figure 7.8. Radial distribution functions (a) CO
2
around Na
+
ions, N and In atoms in
Na-rho-ZMOF at 10 kPa (b) CO
2
around Na
+
ions in Na-rho-ZMOF at 10, 100 and
1000 kPa.

Chapter 7. CO
2
Adsorption in Cation-Exchanged MOFs
112
7.3.3 CO
2
/H
2
Mixture
Figure 7.9 shows the selectivity of CO
2
/H
2
in cation-exchanged rho-ZMOFs as a
functional of total pressure. At a given pressure, the selectivity increases as Cs
+
< Rb
+

< K
+
< Na
+
< Ca
2+
< Mg
2+
~ Al
3+
, which largely follows the increasing order of the
charge-to-diameter ratio of cation. 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. Particularly as shown in Figure 7.7, more CO
2

molecules are adsorbed in the a-cage when pressure increases and the electrostatic
interactions between CO
2
and Na
+
ions reduce substantially. Second, H
2
is smaller
than CO
2
and can pack into the partially filled cages more easily with increasing
pressure. At ambient conditions (298 K and 100 kPa), the selectivity ranges from 800
in Cs-rho-ZMOF to 3000 in Al-rho-ZMOF. The simulated selectivity of CO
2
/H
2

mixture in rho-ZMOFs is significantly higher than that predicted in other porous
materials. For example, the selectivity ranges from 300 to 600 in soc-MOF,
247

approximately 5 in MFI and 3.5 in ETS-10 for an equimolar CO
2
/H
2
mixture.
361
In an
activated carbon, the selectivity is between 60 and 90 for CO
2
/H
2
mixtures with
different mole fractions.
362
In zeolite Na-4A, the selectivity is 70 for a mixture with
98.6% H
2
and 1.4% CO
2
.
342
The selectivity is 40 in IRMOF-1 and 150 in Cu-BTC at
298 K and 1 atm.
344


Chapter 7. CO
2
Adsorption in Cation-Exchanged MOFs
113
P (kPa)
1 10 100 1000
S
C
O
2
/
H
2
1e+2
1e+3
1e+4
1e+5
Na
K
Rb
Cs
Mg
Ca
Al


Figure 7.9. Selectivities of CO
2
/H
2
mixture in rho-ZMOFs. The composition of
CO
2
/H
2
mixture is 15/85.

A gas mixture usually contains moisture or the separation process might be
carried out under humid condition. Consequence, the separation performance could be
affected by H
2
O and this has indeed been observed in experimental and simulation
studies. A very small amount of H
2
O in Li-LSX significantly affects the adsorption
capacity of N
2
, O
2
and Ar due to the shielding of cations by H
2
O.
363
In cationic
zeolite-X, the Henrys constant of CO
2
declines exponentially with the loading of
H
2
O.
364
For CO
2
/CH
4
mixture in rho-ZMOF, the interaction between CO
2
and Na
+
is
substantially reduced upon adding H
2
O; the selectivity is thus decreased by one order
of magnitude.
329
In other cases, however, the presence of H
2
O is beneficial for
adsorption and separation. For example, the presence of pre-adsorbed H
2
O in BaX
increases the selectivity toward p-xylene in an equimolar p-xylene/m-xylene
mixture.
330
The adsorption of CO
2
and CH
4
is enhanced in Cu-BTC via the occupation
of open metal sites by coordinated water molecules.
137
In the presence of H
2
O, the
selectivity of CO
2
over H
2
in soc-MOF increases at low pressures due to the promoted
adsorption of CO
2
by H
2
O bound onto metal atoms, but decreases at high pressures as
Chapter 7. CO
2
Adsorption in Cation-Exchanged MOFs
114
a result of the competitive adsorption of H
2
O over CO
2
.
247

To examine the effect of H
2
O on the separation of CO
2
/H
2
mixture in rho-ZMOFs,
a CO
2
/H
2
/H
2
O mixture was considered with 0.1% of H
2
O (mole fraction). As shown
in Figure 7.10, the selectivities in all the seven rho-ZMOFs decrease substantially
compared with Figure 7.9. This is attributed to the competitive adsorption between
H
2
O and CO
2
. H
2
O is highly polar and interacts with cations more strongly than CO
2
,
thus has an adverse effect on CO
2
adsorption and CO
2
/H
2
separation. Therefore, it is
important to remove H
2
O before separation process. Furthermore, it is observed that
the degree of decrease in selectivity is about 1 order of magnitude in rho-ZMOFs
exchanged with monovalent cations, and 1 ~ 2 orders of magnitude in rho-ZMOFs
exchanged with di- and trivalent cations. Apparently, H
2
O has a greater effect on the
selectivities in rho-ZMOFs exchanged with higher valent cations. Interestingly, in the
presence of H
2
O, the selectivity is close in the seven cation-exchanged rho-ZMOFs.
This is because the electrostatic interactions by cations are largely shielded by H
2
O
and the difference is small among various cation-exchanged rho-ZMOFs.
P (kPa)
1 10 100 1000
S
C
O
2
/
H
2
10
100
1000
Na
K
Rb
Cs
Mg
Ca
Al

Figure 7.10. Selectivity of CO
2
/H
2
/H
2
O mixture in rho-ZMOFs. The mole fraction of
H
2
O in the mixture is 0.1%.
Chapter 7. CO
2
Adsorption in Cation-Exchanged MOFs
115
7.4 Conclusions
We have investigated CO
2
adsorption in cation-exchanged rho-ZMOFs. The
cations include monovalent Na
+
, K
+
, Rb
+
, Cs
+
, divalent Mg
2+
, Ca
2+
and trivalent Al
3+
.
In all the seven cation-exchanged rho-ZMOFs studied, the isosteric heat and Henrys
constant at infinite dilution increase following Cs
+
< Rb
+
< K
+
< Na
+
< Ca
2+
< Mg
2+
<
Al
3+
, in accord with the increasing order of the charge-to-diameter ratio of cation. At
low pressures, cations act as preferential sites for CO
2
adsorption and the electrostatic
interactions of CO
2
-cations govern capacity, which increases in the order of Cs
+
< Rb
+

< K
+
< Na
+
< Ca
2+
< Mg
2+
< Al
3+
. At high pressures, the electrostatic interactions are
largely reduced because CO
2
molecules are located more in the a-cage. Consequently,
the free volume plays a more important role and Mg-rho-ZMOF exhibits a higher
capacity than Al-rho-ZMOF. For the adsorption of CO
2
/H
2
mixture, the selectivity at a
given pressure increases as Cs
+
< Rb
+
< K
+
< Na
+
< Ca
2+
< Mg
2+
~ Al
3+
, and largely
follows the increasing order of the charge-to-diameter ratio. With increasing pressure,
the selectivity decreases sharply due to the occupation of adsorbate molecules in less
favorable sites and the smaller size of H
2
. At 298 K and 1 bar, the selectivity ranges
from 800 to 3000, significantly higher than in other nanoporous materials such as
silicalite, activated carbon and non-ionic MOFs. Upon adding 0.1% H
2
O, the
selectivity decreases substantially because H
2
O is more competitively adsorbed
against CO
2
. Particularly, H
2
O induces a strong shielding on cations and the
selectivity exhibits approximately the same value in all the seven cation-exchanged
rho-ZMOFs.
Chapter 8. Ionic Liquid/MOF Composite for CO
2
Capture
116
Chapter 8. Ionic Liquid/MOF Composite for CO
2
Capture
In this Chapter, a composite of ionic liquid (IL) [BMIM][PF
6
] supported on
metal-organic framework IRMOF-1 is investigated for CO
2
capture by molecular
computation. The microscopic properties of IL confined in the MOF are also provided;
this fundamental understating is indispensable for the synthesis of novel MOFs using
ILs as templates
365
or bridging ligands.
366
The structural and dynamic properties of IL
in MOF are presented and the adsorption behavior of CO
2
/N
2
mixture in IL/MOF
composite is examined.

8.1 Models and Methods
8.1.1 [BMIM][PF
6
]
The IL considered is 1-butyl-3-methylimidazolium hexafluorophosphate
[BMIM][PF
6
], which is one of the commonly studied imidazolium-based ILs. Figure
8.1 illustrates the molecular structure of [BMIM][PF
6
].



Figure 8.1. Atomic types in [BMIM]
+
and [PF
6
]

.

The interactions of IL molecules include bonded stretching, bending, torsional
potentials and non-bonded Lennard-Jones (LJ), electrostatic potentials. The potential
F
H
1

C
1

C
2

C
3

C
4

C
5

C
6

C
7

C
8

H
3

H
4

H
5

H
2

H
6

H
7

H
8

H
9

H
10

H
11

H
13

H
14

H
12

[BMIM]
Chapter 8. Ionic Liquid/MOF Composite for CO
2
Capture
117
parameters of anion [PF
6
]

were adopted from Lopes et al.


367
For cation [BMIM]
+
, the
bonded interactions and non-bonded LJ interactions were represented by the AMBER
force field.
306
To estimate the atomic charges of [BMIM]
+
, density functional theory
(DFT) calculations with the Lee-Yang-Parr correlation (B3LYP) functional were
carried out using GAUSSIAN 03 package.
323
Initially, [BMIM]
+
was geometrically
optimized at 6-31G(d) basis set and the electrostatic potentials were calculated at
6-311+G(d,p) basis set. Thereafter, the atomic charges were estimated by fitting the
electrostatic potentials with Merz-Lollman (MK) scheme.
301
Table 8.1 lists the atomic
charges of [BMIM]
+
and [PF
6
]

.

Table 8.1. Atomic charges in [BMIM]
+
and [PF
6
]

.

Atom C
1
C
2
C
3
C
4
C
5
C
6
C
7
C
8
N
1
N
2

Charge -0.158 -0.151 -0.236 -0.316 -0.342 0.136 0.112 -0.315 0.218 0.276

Atom H
1
H
2
H
3
H
4
H
5
H
6
H
7
H
8
H
9
H
10

Charge 0.205 0.251 0.162 0.164 0.170 0.143 0.145 0.009 0.009 0.009

Atom H
11
H
12
H
13
H
14
H
15

Charge 0.008 0.082 0.111 0.079 0.230

Atom P F
Charge 1.34 -0.39

To validate the force field, the density of [BMIM][PF
6
] was estimated from NPT
(constant temperature and pressure) MD simulation using GROMACS package
v.4.5.3.
368
The initial velocities in MD simulation were assigned according to the
Maxwell-Boltzmann distribution at 300 K. The equations of motion were integrated
with a time step of 2 fs by Leapfrog algorithm.
369
Temperature was controlled by
velocity-rescaled Berendsen thermostat with a relaxation time of 0.1 ps and pressure
Chapter 8. Ionic Liquid/MOF Composite for CO
2
Capture
118
was maintained at 1 atm using Berendsen algorithm.
370
The LJ interactions were
evaluated with a cutoff value of 14 and the electrostatic interactions were calculated
using particle-mesh Ewald method with a grid spacing of 1.2 and a fourth-order
interpolation. The simulation trajectories were saved with a time interval of 2 ps.
Table 8.2 gives the simulated densities of [BMIM][PF
6
] at five different temperatures
(25, 40, 50, 60 and 70 C). Good agreement is observed between the simulated and
experimental data. Though the simulations slightly underestimate experimental data,
the deviations are less than 0.5%. To a certain extent, this suggests that the force field
used for [BMIM][PF
6
] is accurate.

Table 8.2. Simulated and experimental densities of [BMIM][PF
6
] at 1 atm.

Temperature
(C)
Simulated Density
(kg/m
3
)
Experimental
Density
371
(kg/m
3
)
25 1354 1360
40 1344 1348
50 1337 1340
60 1328 1332
70 1320 1323

8.1.2 IRMOF-1
The MOF support in this study for IL/MOF composite is IRMOF-1, which is a
prototype MOF. As shown in Figure 8.2, IRMOF-1 has a cubic structure and a lattice
constant of 25.832 .
5
The formula of IRMOF-1 is Zn
4
O(BDC)
3
, where BDC is
1,4-benzenedicarboxylate. Each oxide-centered Zn
4
O tetrahedron is edge-bridged by
six carboxylate linkers resulting in an octahedral Zn
4
O(O
2
C)
6
building unit, which
reticulates into a three-dimensional structure. Straight channels exist in the framework
Chapter 8. Ionic Liquid/MOF Composite for CO
2
Capture
119
with alternating size of 15 and 12 . The atomic charges of IRMOF-1 framework
atoms were calculated from DFT based on a fragmental cluster shown in Figure 8.3.


Figure 8.2. IRMOF-1 structure. Color code: Zn, orange; O, red; C, grey; H, white.


Figure 8.3. Atomic charges in a fragmental cluster of IRMOF-1. The dangling bonds
indicated by dashed circles are terminated by methyl groups.

The dangling bonds of the cluster were terminated by methyl groups. The 6-31G(d)
basis set was used for all atoms except metal atoms, for which LANL2DZ basis set
was used. After the DFT calculation, the atomic charges were fitted to electrostatic
potentials, as listed in Figure 8.3. The dispersion interactions of IRMOF-1 framework
atoms were represented by LJ potential with parameters from the UFF.
302
The
Zn (1.51))
O
2
(-0.72)) O
1
(-1.79))
H (0.16))
C
3
(-0.25))
C
2
(0.20))
Chapter 8. Ionic Liquid/MOF Composite for CO
2
Capture
120
Lorentz-Berthelot combining rules were used to evaluate cross interaction parameters.
As shown in Figure 8.4 for CO
2
adsorption in IRMOF-1,
39
the agreement between
simulation and experiment is fairly good.

P (kPa)
1 10 100 1000
N

(
m
m
o
l
/
g
)
0
5
10
15
20
25
30
0 1500 3000 4500
0
5
10
15
20
25
30
5000


Figure 8.4. Adsorption isotherm of CO
2
in IRMOF-1 at 300 K.
39
The filled symbols
are from simulation and the open symbols are from experiment.

8.1.3 IL/IRMOF-1 Composite and Adsorption of CO
2
/N
2
Mixture
To examine the properties of IL/IRMOF-1 composite, we consider four different
weight ratios of IL over IRMOF-1 (W
IL/IRMOF-1
= 0, 0.4, 0.86, 1.27, 1.5). At each ratio,
the desired number of [BMIM][PF
6
] were added into IRMOF-1 and subject to energy
minimization using the steepest descent method. Then, MD simulation was conducted
for [BMIM][PF
6
] in IRMOF-1 framework at 300 K. The LJ and electrostatic
interactions were evaluated using the same manner as described above for
[BMIM][PF
6
]. The time step was 2 fs and the trajectory was saved with a time
interval of 2 ps. The total MD simulation duration was 20 ns, with the final 10 ns used
for analysis. Figure 8.5 schematically illustrates an equilibrium snapshot of the
Chapter 8. Ionic Liquid/MOF Composite for CO
2
Capture
121
IL/IRMOF-1 composite at W
IL/IRMOF-1
= 0.4.



Figure 8.5. [BMIM][PF
6
]/IRMOF-1 composite at a weight ratio W
IL/IRMOF-1
= 0.4. N:
blue, C in [BMIM]
+
: green, P: pink, F: cyan; Zn: orange, O: red, C in IRMOF-1: grey,
H: white.

The capability of IL/IRMOF-1 for CO
2
capture was evaluated by simulating
adsorptive separation of CO
2
/N
2
mixture in the composite. GCMC method was used
in the simulation. CO
2
was represented as a three-site molecule and its intrinsic
quadrupole moment was described by a partial charge model. The partial charges on C
and O atoms were q
C
= 0.576e and q
O
= 0.288e (e = 1.6022 10
-19
), respectively. The
CO bond length used was 1.18 and the bond angle ZOCO was 180. CO
2
CO
2

interactions were modeled as a combination of LJ and electrostatic potentials.
317
N
2

was mimicked as a two-site model with the LJ potential parameters fitted to the
experimental bulk properties.
336
The bulk composition of CO
2
/N
2
mixture was
assumed to be 0.15:0.85, representing a typical flue gas. The IL/IRMOF-1 composite
was treated as rigid during adsorption simulation.
Chapter 8. Ionic Liquid/MOF Composite for CO
2
Capture
122
8.2 Results and Discussion
8.2.1 Structure and Dynamics of IL in IL/IRMOF-1
Figure 8.6 shows the g(r) of [BMIM]
+
and [PF
6
]

in IL/IRMOF-1 composite at
weight ratio W
IL/IRMOF-1
= 0.4. For comparison, the g(r) of IL in bulk phase are also
plotted. Due to the confinement effect of IRMOF-1 framework, IL molecules are
packed more ordered in the composite. Consequently, the peaks of all ion pairs in the
composite are observed to be higher than in bulk phase. In particular, the attractive
[BMIM]
+
-[PF
6
]

pair exhibits a pronounced peak at r = 4.4 . Nevertheless, the


[BMIM]
+
-[BMIM]
+
or [PF
6
]

-[PF
6
]

pair has a lower peak because of unfavorable
repulsive interaction. The peak position for [BMIM]
+
-[BMIM]
+
or [PF
6
]

-[PF
6
]

pair
in the composite is slightly shifted to a greater value compared to that in bulk phase,
indicating the distance of cation-cation or anion-anion increases. This is because the
confinement effect in the composite leads to a sharper g(r) for cation-anion pair, and
thus each cation (or anion) is surrounded by a larger number of anions (or cations).
() r
2 4 6 8 10 12
g

(
r
)
0
2
4
6
8
10
[BMIM]
+
-[BMIM]
+
[BMIM]
+
-[PF
6
]
-
[PF
6
]
-
-[PF
6
]
-
[BMIM]
+
-[BMIM]
+
[BMIM]
+
-[PF
6
]
-
[PF
6
]
-
-[PF
6
]
-


Figure 8.6. Radial distribution functions of [BMIM]
+
and [PF
6
]

in IL/IRMOF-1 at
W
IL/IRMOF-1
= 0.4 and in bulk phase, respectively. The solid lines are in IL/IRMOF-1
and the dash lines are in bulk phase.
Chapter 8. Ionic Liquid/MOF Composite for CO
2
Capture
123
r
2 4 6 8 10 12
g

(
r
)
0
1
2
3
4
[BMIM]
+
- O
1
[BMIM]
+
- O
2
[BMIM]
+
- Zn
[BMIM]
+
- C
3
() ()
(a) (b)
r
2 4 6 8 10 12
g

(
r
)
0
1
2
3
4
5
[PF
6
]
-
- O
1
[PF
6
]
-
- O
2
[PF
6
]
-
- Zn
[PF
6
]
-
- C
3
(b)


Figure 8.7. Radial distribution functions of (a) [BMIM]
+
and (b) [PF
6
]

around O
1
, O
2
,
Zn, and C
3
atoms of IRMOF-1 at W
IL/IRMOF-1
= 0.4.

Figure 8.7 further shows the g(r) of [BMIM]
+
and [PF
6
]

around IRMOF-1
framework atoms (O
1
, O
2
, Zn, and C
3
atoms) at W
IL/IRMOF-1
= 0.4. Non-zero g(r) is
observed at r < 5 for [BMIM]
+
-C
3
and [BMIM]
+
-O
2
, but at r > 5 for [BMIM]
+
-Zn
and [BMIM]
+
-O
1
. This indicates [BMIM]
+
is proximal to the benzene ring and
carboxylate group in IRMOF-1 rather than the metal cluster. With a chain-like
structure, [BMIM]
+
prefers to reside in the open pore of IRMOF-1 due to
configuration entropy effect. The small corner near the metal cluster cannot
accommodate [BMIM]
+
. However, pronounced peaks are observed for [PF
6
]

around
all the four atoms at r < 5 . Particularly, the peaks around O
1
and Zn atoms are
higher, suggesting [PF
6
]

is preferentially located in the corner near the metal cluster.


Compared to [BMIM]
+
, [PF
6
]

exhibits higher peaks around the four atoms; therefore,


[PF
6
]

possesses a stronger interaction with the framework. Note that the peaks at r =
10 ~ 11 are due to the presence of framework atoms on the other side of the
structure.
Chapter 8. Ionic Liquid/MOF Composite for CO
2
Capture
124

t (ps)
0 1000 2000 3000 4000 5000 6000
M
S
D

(



)
0
5
10
15
20
25
30
[BMIM]
+
0.4
[BMIM]
+
0.86
[BMIM]
+
1.27
[PF
6
]
-
0.4
[PF
6
]
-
0.86
[PF
6
]
-
1.27

2



Figure 8.8. Mean-squared displacements of [BMIM]
+
and [PF
6
]

in IL/IRMOF-1at
W
IL/IRMOF-1
= 0.4, 0.86 and 1.27.

The dynamics of IL is evaluated by mean-squared displacement (MSD). Figure
8.8 shows the MSDs of [BMIM]
+
and [PF
6
]

in IL/IRMOF-1 at W
IL/IRMOF-1
= 0.4, 0.86
and 1.27. Though not shown, the mobility of IL in bulk phase is greater than in the
composite. The mobility of both [BMIM]
+
and [PF
6
]

reduces with increasing


W
IL/IRMOF-1
, as attributed to enhanced confinement effect. At a given W
IL/IRMOF-1
, the
mobility of bulky [BMIM]
+
is greater than [PF
6
]

. Such an interesting phenomenon


was also observed in simulations
372,373
and experiments
374,375
for imidazolium-based
bulk ILs, and interpreted as a result of the less hindered displacement of [BMIM]
+

ring along the direction of C
1
atom (see Figure 8.1).
376
In the IL/IRMOF-1 composite
under study, [PF
6
]

interacts more strongly with the framework as discussed above;


this also contributes to the observed smaller mobility of [PF
6
]

.

Chapter 8. Ionic Liquid/MOF Composite for CO
2
Capture
125
t (ps)
0 1 2 3 4
C
(
t
)
/
C
(
0
)
-0.4
-0.2
0.0
0.2
0.4
0.6
0.8
1.0
[BMIM]
+
0.4
[BMIM]
+
0.86
[BMIM]
+
1.27
[BMIM]
+
1.5
(a)

t (ps)
0 1 2 3 4
C
(
t
)
/
C
(
0
)
-0.4
-0.2
0.0
0.2
0.4
0.6
0.8
1.0
(b)
[PF
6
]
-
0.4
[PF
6
]
-
0.86
[PF
6
]
-
1.27
[PF
6
]
-
1.5


Figure 8.9. Reduced velocity correlation functions of [BMIM]
+
and [PF
6
]

in
IL/IRMOF-1 at W
IL/IRMOF-1
= 0.4, 0.86, 1.27 and 1.5.

To provide further insight into the dynamics of IL in the IL/IRMOF-1 composite,
velocity correlation functions were calculated by
( ) (0) ( ) C t t = v v (8.1)
where v(t) is the velocity of center-of-mass of cation or anion of at time t. To calculate
C(t), a 100 ps trajectory was generated with a small time interval of 0.1 ps after 20 ns
MD simulation. Figure 8.9 shows the reduced velocity correlation functions C(t)/C(0)
of [BMIM]
+
and [PF
6
]

in the composite. The trend at different W


IL/IRMOF-1
is largely
identical. For either cation or anion, the C(t)/C(0) approaches zero in less than 2 ps
and then exhibits marginal oscillation. Therefore, the velocity of [BMIM]
+
or [PF
6
]


becomes uncorrelated rapidly within a picosecond scale. This is similar to the time
scale observed for IL in bulk phase,
376
and reveals that confinement has an
insignificant effect on the velocity correlation of IL.
8.2.2 Separation of CO
2
/N
2
Mixture in IL/IRMOF-1
Adsorption of CO
2
/N
2
mixture in IL/IRMOF-1 was simulated to evaluate the
Chapter 8. Ionic Liquid/MOF Composite for CO
2
Capture
126
performance of IL/IRMOF-1 composite for CO
2
capture. The CO
2
/N
2
mixture was
assumed to possess a composition of 15:85 to mimic a typical flue gas. Figure 8.10
shows the simulation snapshot of CO
2
/N
2
mixture with a total pressure of 1000 kPa in
the composite at W
IL/IRMOF-1
= 0.4. [PF
6
]

anions are observed to preferentially locate


in the corner of IRMOF-1, which was evidenced by the g(r) in Figure 8.7. Compared
to N
2
, more CO
2
molecules are adsorbed in the composite and proximal to the ions.
This is attributed to the strong interaction between CO
2
and ions.



Figure 8.10. Simulation snapshot of CO
2
/N
2
mixture (P
total
= 1000 kPa) in
IL/IRMOF-1 at W
IL/IRMOF-1
= 0.4.

To identify the favorable sites in the composite for CO
2
adsorption, Figure 8.11
shows the g(r) of CO
2
around the framework (Zn) and IL (N
1
, N
2
and P) atoms. A
sharp peak is observed in the g(r) of CO
2
-P at r ~ 4.1 , indicating [PF
6
]

anion is the
most favorable site. The g(r) of CO
2
-N
1
and CO
2
-N
2
are less pronounced with broad
peaks at r ~ 5.0 and 5.2 , respectively. As a comparison, the g(r) of CO
2
-Zn has the
lowest peak at r ~ 4.5 . The structural analysis reveals that CO
2
is preferentially
adsorbed onto IL, particularly [PF
6
]

anion, in the IL/IRMOF-1 composite. Such


CO
2

N
2

Chapter 8. Ionic Liquid/MOF Composite for CO
2
Capture
127
behavior is remarkably different from the adsorption in IRMOF-1, in which the
favorable site was found to be near the metal cluster.
39
In the IL/IRMOF-1 composite,
the cations and anions (especially [PF
6
]

) occupy the corner of the metal cluster and


act as adsorption sites for CO
2
adsorption.

r
2 4 6 8 10 12
g

(
r
)
0.0
0.5
1.0
1.5
2.0
2.5
3.0
Zn in IRMOF-1
N
1
in IL
N
2
in IL
P in IL
()


Figure 8.11. Radial distribution functions of CO
2
(P
total
= 10 kPa) around Zn, N
1
, N
2
,
and P atoms in IL/IRMOF-1 at W
IL/IRMOF-1
= 0.4.

Figure 8.12a shows the g(r) of CO
2
around P atom of [PF
6
]

at 10, 100 and 1000


kPa in the IL/IRMOF-1 composite at W
IL/IRMOF-1
= 0.4. The peak in g(r) of CO
2
-P
drops with increasing pressure, whereas the coordination number of CO
2
molecules
around P atoms increases (data not shown). This is because the local density of CO
2

increases less rapidly than overall density. Similar behavior was also found in our
previous studies for CO
2
capture in rho zeolite-like MOF
42
. As shown in Figure 8.12b,
the peak in g(r) of CO
2
-P also drops with increasing W
IL/IRMOF-1
in the IL/IRMOF-1
composite. At a low W
IL/IRMOF-1
, only a few IL molecules exist and reside closely to
the IRMOF-1 framework; consequently, CO
2
molecules are largely localized because
Chapter 8. Ionic Liquid/MOF Composite for CO
2
Capture
128
the ions act as favorable adsorption sites. With increasing W
IL/IRMOF-1
, however, the
ions start to occupy the open pores and distribute homogenously in the IRMOF-1
framework; therefore, CO
2
molecules are adsorbed uniformly and the local density
relative to average density drops.

r
2 4 6 8 10 12
g

(
r
)
0.0
0.5
1.0
1.5
2.0
2.5
3.0
10 kPa
100 kPa
1000 kPa
()
(a)

r
2 4 6 8 10 12
g

(
r
)
0.0
0.5
1.0
1.5
2.0
2.5
3.0
0.4
0.86
1.27
1.5
()
(b)

Figure 8.12. Radial distribution functions of CO
2
around P atom in IL/IRMOF-1 (a)
P
total
= 10, 100, 1000 kPa and W
IL/IRMOF-1
= 0.4, (b) P
total
= 100 kPa and W
IL/IRMOF-1
=
0.4, 0.86, 1.27 and 1.5.

Figure 8.13 shows the selectivity
2 2
CO / N
S in the IL/IRMOF-1 composite at
different W
IL/IRMOF-1
. In the absence of IL,
2 2
CO / N
S remains nearly 5 over the entire
pressure range; however, it increases with increasing W
IL/IRMOF-1
. This is because IL
acts as favorable sites for CO
2
adsorption and more such sites are available when
W
IL/IRMOF-1
increases. Nevertheless,
2 2
CO / N
S

decreases with increasing pressure,
particularly at high W
IL/IRMOF-1
because of the number of adsorption sites decreases at
high pressures. At the highest W
IL/IRMOF-1
= 1.5 in this study,
2 2
CO / N
S reaches
approximately 90 at infinite dilution and 70 at 1 bar. This value is higher than the
selectivities reported for CO
2
/N
2
mixture in [BMIM][PF
6
] (22.6),
290
in PVDF
supported [BMIM][PF
6
] (20),
296
and in many other supported ILs (10 56).
297

Chapter 8. Ionic Liquid/MOF Composite for CO
2
Capture
129

P
total
(kPa)
0 200 400 600 800 1000
S
C
O
2
/
N
2
0
20
40
60
80
100
1.5
1.27
0.86
0.4
0

Figure 8.13. Selectivity of CO
2
/N
2
mixture in IL/IRMOF-1 at W
IL/IRMOF-1
= 0, 0.4,
0.86, 1.27 and 1.5.

8.3 Conclusions
We have examined the microscopic properties of [BMIM][PF
6
] in IL/IRMOF-1
composite and the capability of the composite for CO
2
capture. IL molecules in the
composite are packed more ordered than in bulk phase, particularly for the
electrostatic attractive [BMIM]
+
-[PF
6
]

pair. This is attributed to the confinement


effect of nanoporous IRMOF-1 framework. The bulky chain-like [BMIM]
+
cation
tends to reside in the open pore of IRMOF-1. The small [PF
6
]

has a stronger
interaction with the framework than [BMIM]
+
and stays preferentially in the corner
near the metal cluster. As also observed in bulk phase, the mobility of [BMIM]
+
in the
IL/IRMOF-1 composite is greater than [PF
6
]

. Due to the enhanced confinement


effect when the ratio of IL in the composite increase, the mobility of both [BMIM]
+

and [PF
6
]

reduces. Nevertheless, confinement has no significant effect on the


velocity correlation. It is found that the velocity of [BMIM]
+
or [PF
6
]

is essentially
uncorrelated after 2 picoseconds, which is close to the time scale for bulk ILs.
Chapter 8. Ionic Liquid/MOF Composite for CO
2
Capture
130
In the IL/IRMOF-1 composite, CO
2
is more strongly adsorbed than N
2
from
CO
2
/N
2
mixture. From structural analysis, [PF
6
]

anion is identified to be the most


favorable site for CO
2
adsorption in the composite. In IRMOF-1, however, the
favorable site is in the corner of the metal cluster. Upon adding IL into IRMOF-1, ions
especially [PF
6
]

occupy the corner and act as strong ionic adsorption sites for CO
2
.
With increasing ratio of IL in the composite, ions start to occupy the open pores and
distribute more homogenously, thus CO
2
molecules tend to locate uniformly in the
composite. Compared with IRMOF-1, the IL/IRMOF-1 composite has a substantially
higher selectivity for CO
2
/N
2
separation. With increasing ratio of IL, the selectivity in
the composite increases as the number of ionic adsorption sites becomes larger. At a
weight ratio IL/IRMOF-1 = 1.5, the selectivity is about 70 at ambient conditions (300
K and 1 bar) and higher than in most supported ILs. The simulation results suggest
that the IL/IRMOF-1 composite is potentially a good material for CO
2
capture. In the
current study, the common [BMIM][PF
6
] and prototype IRMOF-1 are considered.
Nevertheless, thousands of ILs and MOFs have been synthesized with their readily
tunable structures; therefore, new candidates will be further explored.




Chapter 9. Conclusions and Recommendation
131
Chapter 9. Conclusions and Recommendation
9.1 Conclusions
This thesis aims to investigate the adsorption and separation of gas mixtures
(particularly CO
2
-containing mixtures) in various MOFs with diverse structures and
functionalities. Molecular simulations as well as first-principles calculations have
been employed to evaluate adsorption mechanisms and microscopic insights. The
main contents of the thesis are summarized below.
1. Adsorption of CO
2
and CH
4
in dehydrated and hydrated mesoporous MIL-101
is examined. Adsorption occurs exclusively in the microporous supertetrahedra at low
pressures and also in the mesoporous cages at high pressures. The simulated
isotherms match well with experimentally measured data that depend on activation
method. The terminal water molecules act as additional adsorption sites and enhance
adsorption at low pressures. This is because the terminal water molecules interact
more strongly than the exposed Cr sites for adsorbates. The enhancement is greater
for CO
2
as it is a quadrapolar molecule and interacts strongly with the charged
hydrogen and oxygen atoms of the water molecules. The opposite occurs at high
pressures because the terminal water molecules reduce free volume and lead to less
adsorption. This study provides useful insights into the microscopic adsorption
behavior in a unique MOF and underlines the interesting effect of the terminal water
molecules on adsorption.
2. Adsorption and separation of CH
3
OH/H
2
O and CO
2
/CH
4
in Zn(BDC)(TED)
0.5

Chapter 9. Conclusions and Recommendation
132
are studied. Zn(BDC)(TED)
0.5
possesses a highly hydrophobic framework because of
the existence of BDC and TED linkers. The simulated isotherms of CH
3
OH and H
2
O
are in fairly good agreement with experimental data. While H
2
O adsorption is
vanishingly small, CH
3
OH has a much strong adsorption. The selectivity is
approximately 20 at low pressures and decreases with increasing pressure. CH
3
OH
interacts strongly with the metal oxides, particularly at low pressures. The isotherms
of CO
2
and CH
4
from simulation also match well with experimental data. As a
nonpolar molecule, CO
2
exhibits different favorable sites from polar CH
3
OH. At low
pressures, CO
2
is located preferentially near the phenyl rings, then the metal oxides
and dimethylenes with increasing pressure. The selectivity of CO
2
over CH
4
increases
as a function of pressure, with a magnitude similar to other neutral MOFs. H
2
O has a
marginal effect on the selectivity of CO
2
/CH
4
. This study provides a quantitative
understanding of adsorption and separation in Zn(BDC)(TED)
0.5
and suggests that
Zn(BDC)(TED)
0.5
is a good candidate for the purification of liquid fuels.
3. Biological metal-organic framework (bio-MOF-11) for CO
2
capture is
explored. The biomolecular linkers (adenines) in bio-MOF-11 contain Lewis basic
amino and pyrimidine groups as the preferential adsorption sites. The simulated and
experimental adsorption isotherms of CO
2
, H
2
, and N
2
are in perfect agreement. As
attributed to the presence of multiple Lewis basic sites and nano-sized channels,
bio-MOF-11 exhibits larger adsorption capacities compared with zeolites, activated
carbons, and many other MOFs. For the adsorption of CO
2
/H
2
and CO
2
/N
2
mixtures
in bio-MOF-11, CO
2
is more dominantly adsorbed than H
2
and N
2
. With increasing
Chapter 9. Conclusions and Recommendation
133
pressure, the selectivity of CO
2
/H
2
initially increases due to the strong interactions
between CO
2
and framework, and then decreased as a consequence of entropy effect.
However, the selectivity of CO
2
/N
2
increases monotonically with increasing pressure
and finally reaches a constant. The selectivities in bio-MOF-11 are higher than in
many nanoporous materials. In addition, the simulation results indicate that a small
amount of H
2
O has an insignificant effect on the separation. The simulation study
provides microscopic insights into adsorption mechanism in bio-MOF-11 and
suggests that bio-MOFs might be interesting for CO
2
capture.
4. CO
2
adsorption is investigated in rho-ZMOFs exchanged with a series of
cations (Na
+
, K
+
, Rb
+
, Cs
+
, Mg
2+
, Ca
2+
and Al
3+
). The isosteric heat

and Henrys
constant at infinite dilution increase monotonically with increasing charge-to-diameter
ratio of cation (Cs
+
< Rb
+
< K
+
< Na
+
< Ca
2+
< Mg
2+
< Al
3+
). At low pressures, cations
act as preferential adsorption sites and the capacity follows the charge-to-diameter
ratio. However, the free volume of framework becomes predominant with increasing
pressure and Mg-rho-ZMOF exhibits the highest saturation capacity. For CO
2
/H
2

mixture, the selectivity increases as Cs
+
< Rb
+
< K
+
< Na
+
< Ca
2+
< Mg
2+
~ Al
3+
. At
ambient conditions, the selectivity is in the range of 800 ~ 3000 and significantly
higher than in other MOFs. In the presence of 0.1% H
2
O, the selectivity decreases
drastically because of the competitive adsorption between H
2
O and CO
2
, and shows
similar value in all the cation-exchanged rho-ZMOFs. This simulation study reveals
the important role of cations in governing gas adsorption and separation, and suggests
that the performance of ionic rho-ZMOFs can be tailored by cations.
Chapter 9. Conclusions and Recommendation
134
5. A composite of ionic liquid (IL) [BMIM][PF
6
] supported on IRMOF-1 is
proposed for CO
2
capture. Due to the confinement effect, IL in the composite exhibits
ordered structure. The bulky [BMIM]
+
cation resides in the open pore of IRMOF-1,
whereas the small [PF
6
] anion prefers to locate in the metal cluster corner and
possesses a strong interaction with the framework. [BMIM]
+
exhibits a greater
mobility than [PF
6
], which is also observed in simulation and experimental studies of
imidazolium-based ILs in bulk phase. With increasing IL ratio in the composite and
thus enhancing confinement effect, the mobility of [BMIM]
+
and [PF
6
] reduces. Ions
in the composite interact strongly with CO
2
, particularly [PF
6
] is the most favorable
site for CO
2
. The composite selectively adsorbs CO
2
from CO
2
/N
2
mixture, with
selectivity significantly higher than polymer-supported ILs. Furthermore, the
selectivity increases with increasing IL ratio in the composite. This computational
study demonstrates MOF-supported IL might be potentially useful for CO
2
capture.
Table 9.1 summarizes CO
2
selectivity at ambient conditions over other gases in
the five MOFs. It can be concluded that MOFs contain functional groups
(bio-MOF-11), ionic framework (rho-ZMOF) and more adsorption sites (IL/MOF
composite) possess better performance for CO
2
capture. Overall, the
recommendations for MOFs well-suited for CO
2
capture should have the following
characteristics: proper pore size, high density of open metal site, high polar functional
group, ionic framework, and more adsorption sites. In addition, H
2
O has different
effects on CO
2
capture in various MOFs. In MIL-101, terminal H
2
O enhances
selectivity; in rho-ZMOF the presence of H
2
O largely decreases the selectivity; in
Chapter 9. Conclusions and Recommendation
135
Zn(BDC)(TED)
0.5
and bio-MOF-11, H
2
O has a negligible effect on CO
2
selectivity.

Table 9.1. CO
2
selectivities at ambient conditions in different MOFs.

Gas mixture MOF CO
2
Selectivity at ambient conditions
CO
2
/CH
4

MIL-101 5
Zn(BDC)(TED)
0.5
4.5
CO
2
/H
2

bio-MOF-11 339
rho-ZMOF 800-3000 (Cs
+
-Mg
2+
)
CO
2
/N
2

bio-MOF-11 37
IL/IRMOF-1 4.8-70 (IL:0-1.5)
9.2 Recommendation
In this thesis, we have focused on the potential application of various MOFs with
unique structures and functionalities for CO
2
capture. However, MOFs are versatile
materials for a wide range of applications such as storage, separation, catalysis, and
biosensing. In addition, tens of thousands of MOFs have been synthesized and a few
have been commercially available. Several interesting and practically important topics
are recommended for future simulation studies.
1. The removal of harmful NO
x
in exhaust gas is of central importance for air
quality. Currently, Ba-based materials are often used for NO
x
storage. Although the
nature of storage mechanism is not well recognized, it is generally recognized that the
basicity of Ba-based materials and alkali/alkaline earth metals (K, Mg, and Ca) play a
key role in NO
x
storage.
377
MOFs with Lewis basic sites, alkali/alkaline earth metal
Chapter 9. Conclusions and Recommendation
136
ions could be potentially used for NO
x
storage. Studies have shown that NO can be
adsorbed in MOFs due to the coordination between NO and open metal sites.
68,378

Toward the screening and design of high-performance MOFs for NO
x
storage, it is
important to elucidate the interactions and microscopic properties of NO
x
in MOFs,
particularly those with open metal sites.
2. Most current experimental and simulations for MOFs have been concentrated
on gas storage and separation, and thus endeavors for liquid separation are lagged
well behind. Nevertheless, a recent trend has been to explore the use of MOFs for
liquid separation.
379,380
Molecular simulation study for liquid separation in MOFs is
scarce due to the significant amount of computational time required to sample liquid
phase. Consequently, the microscopic understanding of liquid separation in MOFs is
far from complete. To the best of our knowledge, only two simulation studies have
been reported in this area, one for water desalination
381
and the other for biofuel
purification.
382
With increasing demands for clean water, liquid fuels and other
liquid-based applications, more efforts are expected in order to provide deep
molecular insights.
3. MOFs may find application in the removal of toxic compounds. For instance,
paracresol is a protein-bound uremic toxin and associated with cardiovascular disease.
It is difficult to remove paracresol by conventional dialysis method.
383
As an
alternative, uremic toxins can be separated by adsorption in porous adsorbents for
blood purification. Wernert
384,385
and Boulet et al.
385-387
have studied the adsorption of
prarcresol in pure silica and cation-compensated alumino-silicalite. It was found that
Chapter 9. Conclusions and Recommendation
137
silicalite has a high affinity for parecresol and can eliminate 60% paracresol, which is
2 times higher than the conventional dialysis method.
384
MOFs with high porosities
and large surface areas could be better used for this purpose. Currently, there is no
study on this topic and useful microscopic information can be obtained from
simulations.
4. Porous materials are potential candidates for drug loading/delivery. There are
several factors determining the loading capacity and release rate of drug in a porous
carrier, such as pore size, shape, and host affinity. In traditional porous materials
including inorganic silica and polymeric matrixes, drug loading capacity is usually not
sufficiently high and encapsulated drug is difficult to be released specifically. To
achieve a high loading and a controlled release, porous carriers with large volumes
and well-defined structures are desired. In this regard, mesoporous MOFs provide a
wealth of opportunities for drug delivery. However, very few simulation studies are
available on this topic
48,49
and the fundamental understanding of drug-MOF
interactions remains largely elusive. Therefore, molecular-level study is crucial for the
development of new MOFs as novel drug devices.


References
138
References

(1) Guido, K. Hybrid materials:synthesis, characterization, and applications, 2007.
(2) Li, H.; Eddaoudi, M.; O'Keeffe, M.; Yaghi, O. M. Nature 1999, 402, 276.
(3) Ferey, G. Chem. Soc. Rev. 2008, 37, 191.
(4) Yaghi, O. M.; O'Keeffe, M.; Ockwig, N. W.; Chae, H. K.; Eddaoudi, M.; Kim, J. Nature 2003, 423,
705.
(5) Eddaoudi, M.; Kim, J.; Rosi, N.; Vodak, D.; Wachter, J.; O'Keeffe, M.; Yaghi, O. M. Science 2002,
295, 469.
(6) Yaghi, O. M.; Li, H. L.; Davis, C.; Richardson, D.; Groy, T. L. Acc. Chem. Res. 1998, 31, 474.
(7) Kitagawa, S.; Uemura, K. Chem. Soc. Rev. 2005, 34, 109.
(8) Fletcher, A. J.; Thomas, K. M.; Rosseinsky, M. J. Journal of Solid State Chemistry
Reticular Chemistry: Design, Synthesis, Properties and Applications of Metal-Organic Polyhedra and
Frameworks 2005, 178, 2491.
(9) Lin, W. B.; Wang, Z. Y.; Ma, L. J. Am. Chem. Soc. 1999, 121, 11249.
(10) Seo, J. S.; Whang, D.; Lee, H.; Jun, S. I.; Oh, J.; Jeon, Y. J.; Kim, K. Nature 2000, 404, 982.
(11) Dechambenoit, P.; Long, J. R. Chem. Soc. Rev. 2011, 40, 3249.
(12) Cui, Y.; Yue, Y.; Qian, G.; Chen, B. Chem. Rev. 2011.
(13) OKeeffe, M.; Yaghi, O. M. Chem. Rev. 2011.
(14) 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. U. S. A. 2006, 103, 10186.
(15) Cote, A. P.; Benin, A. I.; Ockwig, N. W.; O'Keeffe, M.; Matzger, A. J.; Yaghi, O. M. Science 2005,
310, 1166.
(16) El-Kaderi, H. M.; Hunt, J. R.; Mendoza-Cortes, J. L.; Cote, A. P.; Taylor, R. E.; O'Keeffe, M.;
Yaghi, O. M. Science 2007, 316, 268.
(17) Cote, A. P.; El-Kaderi, H. M.; Furukawa, H.; Hunt, J. R.; Yaghi, O. M. J. Am. Chem. Soc. 2007,
129, 12914.
(18) Liu, Y. L.; Kravtsov, V. C.; Larsen, R.; Eddaoudi, M. Chem. Commun. 2006, 1488.
(19) Riou, D.; Roubeau, O.; Ferey, G. Micro. Meso. Mater. 1998, 23, 23.
(20) Riou, D.; Serre, C.; Ferey, G. J. Solid State Chem. 1998, 141, 89.
(21) Riou, D.; Ferey, G. J. Mater. Chem. 1998, 8, 2733.
(22) Serpaggi, F.; Ferey, G. J. Mater. Chem. 1998, 8, 2737.
(23) Serpaggi, F.; Ferey, G. J. Mater. Chem. 1998, 8, 2749.
(24) Serpaggi, F.; Ferey, G. Micro. Meso. Mater. 1999, 32, 311.
(25) Serre, C.; Millange, F.; Thouvenot, C.; Nogues, M.; Marsolier, G.; Louer, D.; Ferey, G. J. Am.
Chem. Soc. 2002, 124, 13519.
(26) Loiseau, T.; Serre, C.; Huguenard, C.; Fink, G.; Taulelle, F.; Henry, M.; Bataille, T.; Ferey, G.
Chem.-Eur. J. 2004, 10, 1373.
(27) Millange, F.; Guillou, N.; Walton, R. I.; Greneche, J. M.; Margiolaki, I.; Fereya, G. Chem.
Commun. 2008, 4732.
References
139
(28) Volkringer, C.; Loiseau, T.; Guillou, N.; Ferey, G.; Elkaim, E.; Vimont, A. Dalton Trans. 2009,
2241.
(29) Barthelet, K.; Marrot, J.; Riou, D.; Ferey, G. Angew. Chem. Int. Edit. 2002, 41, 281.
(30) Ferey, G.; Mellot-Draznieks, C.; Serre, C.; Millange, F.; Dutour, J.; Surble, S.; Margiolaki, I.
Science 2005, 309, 2040.
(31) Son, W.-J.; Kim, J.; Kim, J.; Ahn, W.-S. Chem. Commun. 2008, 6336.
(32) Jung, D.-W.; Yang, D.-A.; Kim, J.; Kim, J.; Ahn, W.-S. Dalton Trans. 2010, 39, 2883.
(33) Friscic, T. J. Mater. Chem. 2010, 20, 7599.
(34) Bennett, T. D.; Cao, S.; Tan, J. C.; Keen, D. A.; Bithell, E. G.; Beldon, P. J.; Friscic, T.; Cheetham,
A. K. J. Am. Chem. Soc. 2011, 133, 14546.
(35) Sabouni, R.; Kazemian, H.; Rohani, S. Chem. Eng. J. 2010, 165, 966.
(36) Nelson, A. P.; Farha, O. K.; Mulfort, K. L.; Hupp, J. T. J. Am. Chem. Soc. 2009, 131, 458.
(37) Rosi, N. L.; Eckert, J.; Eddaoudi, M.; Vodak, D. T.; Kim, J.; O'Keeffe, M.; Yaghi, O. M. Science
2003, 300, 1127.
(38) Ferey, G.; Latroche, M.; Serre, C.; Millange, F.; Loiseau, T.; Percheron-Guegan, A. Chem.
Commun. 2003, 2976.
(39) Babarao, R.; Jiang, J. W. Langmuir 2008, 24, 6270.
(40) Llewellyn, P. L.; Bourrelly, S.; Serre, C.; Filinchuk, Y.; Ferey, G. Angew. Chem. Int. Edit. 2006, 45,
7751.
(41) Babarao, R.; Hu, Z. Q.; Jiang, J. W.; Chempath, S.; Sandler, S. I. Langmuir 2007, 23, 659.
(42) Babarao, R.; Jiang, J. W. J. Am. Chem. Soc. 2009, 131, 11417.
(43) Babarao, R.; Tong, Y. H.; Jiang, J. W. J. Phys. Chem. B. 2009, 113, 9129.
(44) Alaerts, L.; Seguin, E.; Poelman, H.; Thibault-Starzyk, F.; Jacobs, P. A.; De Vos, D. E. Chem.-Eur.
J. 2006, 12, 7353.
(45) 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.
(46) Achmann, S.; Hagen, G.; Kita, J.; Malkowsky, I. M.; Kiener, C.; Moos, R. Sensors 2009, 9, 1574.
(47) Horcajada, P.; Serre, C.; Vallet-Regi, M.; Sebban, M.; Taulelle, F.; Ferey, G. Angew. Chem. Int.
Edit. 2006, 45, 5974.
(48) Horcajada, P.; Serre, C.; Maurin, G.; Ramsahye, N. A.; Balas, F.; Vallet-Regi, M.; Sebban, M.;
Taulelle, F.; Ferey, G. J. Am. Chem. Soc. 2008, 130, 6774.
(49) Babarao, R.; Jiang, J. W. J. Phys. Chem. C. 2009, 113, 18287.
(50) Bhakta, R. K.; Herberg, J. L.; Jacobs, B.; Highley, A.; Behrens, R.; Ockwig, N. W.; Greathouse, J.
A.; Allendorf, M. D. J. Am. Chem. Soc. 2009, 131, 13198.
(51) Allendorf, M. D.; Bauer, C. A.; Bhakta, R. K.; Houk, R. J. T. Chem. Soc. Rev. 2009, 38, 1330.
(52) Kurmoo, M. Chem. Soc. Rev. 2009, 38, 1353.
(53) Sadakiyo, M.; Yamada, T.; Kitagawa, H. J. Am. Chem. Soc. 2009, 131, 9906.
(54) Wiers, B. M.; Foo, M.-L.; Balsara, N. P.; Long, J. R. J. Am. Chem. Soc. 2011, 133, 14522.
(55) Kuppler, R. J.; Timmons, D. J.; Fang, Q.-R.; Li, J.-R.; Makal, T. A.; Young, M. D.; Yuan, D.; Zhao,
D.; Zhuang, W.; Zhou, H.-C. Coordination Chemistry Reviews
Functional Hybrid Nanomaterials: - Design, Synthesis, Structure, Properties and Applications 2009,
253, 3042.
(56) Mueller, U.; Schubert, M.; Teich, F.; Puetter, H.; Schierle-Arndt, K.; Pastre, J. J. Mater. Chem.
2006, 16, 626.
References
140
(57) U.S. Department of Energy, DOE Targets for Onboard Hydrogen Storage Systems for Light-Duty
Vehicles. In
http://www1.eere.energy.gov/hydrogenandfuelcells/storage/pdfs/targets_onboard_hydro_storage.pdf,
2009.
(58) Collins, D. J.; Zhou, H.-C. J. Mater. Chem. 2007, 17, 3154.
(59) Dinca, M.; Long, J. R. Angew. Chem. Int. Edit. 2008, 47, 6766.
(60) Thomas, K. M. Dalton Trans. 2009, 1487.
(61) Ma, S.; Zhou, H.-C. Chem. Commun. 2010, 46, 44.
(62) Kondo, M.; Yoshitomi, T.; Matsuzaka, H.; Kitagawa, S.; Seki, K. Angew. Chem. Int. Edit. 1997,
36, 1725.
(63) Duren, T.; Sarkisov, L.; Yaghi, O. M.; Snurr, R. Q. Langmuir 2004, 20, 2683.
(64) Wang, S. Y. Energy Fuels 2007, 21, 953.
(65) Furukawa, H.; Yaghi, O. M. J. Am. Chem. Soc. 2009, 131, 8875.
(66) Furukawa, H.; Ko, N.; Go, Y. B.; Aratani, N.; Choi, S. B.; Choi, E.; Yazaydin, A. .; Snurr, R. Q.;
OKeeffe, M.; Kim, J.; Yaghi, O. M. Science 2010, 329, 424.
(67) McKinlay, A. C.; Xiao, B.; Wragg, D. S.; Wheatley, P. S.; Megson, I. L.; Morris, R. E. J. Am.
Chem. Soc. 2008, 130, 10440.
(68) Xiao, B.; Wheatley, P. S.; Zhao, X. B.; Fletcher, A. J.; Fox, S.; Rossi, A. G.; Megson, I. L.;
Bordiga, S.; Regli, L.; Thomas, K. M.; Morris, R. E. J. Am. Chem. Soc. 2007, 129, 1203.
(69) An, J. Y.; Geib, S. J.; Rosi, N. L. J. Am. Chem. Soc. 2009, 131, 8376.
(70) Keskin, S.; Kzlel, S. Ind. Eng. Chem. Res. 2011, 50, 1799.
(71) Britt, D.; Furukawa, H.; Wang, B.; Glover, T. G.; Yaghi, O. M. Proc. Natl. Acad. Sci. 2009, 106,
20637.
(72) Banerjee, R.; Phan, A.; Wang, B.; Knobler, C.; Furukawa, H.; O'Keeffe, M.; Yaghi, O. M. Science
2008, 319, 939.
(73) Wang, B.; Cote, A. P.; Furukawa, H.; O'Keeffe, M.; Yaghi, O. M. Nature 2008, 453, 207.
(74) Li, J. R.; Kuppler, R. J.; Zhou, H. C. Chem. Soc. Rev 2009, 38, 1477.
(75) Li, J. R.; Sculley, J.; Zhou, H. C. Chem. Rev. 2011, ASAP.
(76) Fujita, M.; Kwon, Y. J.; Washizu, S.; Ogura, K. J. Am. Chem. Soc. 1994, 116, 1151.
(77) Evans, O. R.; Ngo, H. L.; Lin, W. B. J. Am. Chem. Soc. 2001, 123, 10395.
(78) Wu, C. D.; Hu, A.; Zhang, L.; Lin, W. B. J. Am. Chem. Soc. 2005, 127, 8940.
(79) Lin, W. B. Mrs Bulletin 2007, 32, 544.
(80) Mantion, A.; Massuger, L.; Rabu, P.; Palivan, C.; McCusker, L. B.; Taubert, A. J. Am. Chem. Soc.
2008, 130, 2517.
(81) Ravon, U.; Domine, M. E.; Gaudillere, C.; Desmartin-Chomel, A.; Farrusseng, D. New J. Chem.
2008, 32, 937.
(82) Sabo, M.; Henschel, A.; Froede, H.; Klemm, E.; Kaskel, S. J. Mater. Chem. 2007, 17, 3827.
(83) Schlichte, K.; Kratzke, T.; Kaskel, S. Micro. Meso. Mater. 2004, 73, 81.
(84) Henschel, A.; Gedrich, K.; Kraehnert, R.; Kaskel, S. Chem. Commu. 2008, 4192.
(85) Hwang, Y. K.; Hong, D. Y.; Chang, J. S.; Jhung, S. H.; Seo, Y. K.; Kim, J.; Vimont, A.; Daturi, M.;
Serre, C.; Frey, G. Angew. Chem. Int. Edit. 2008, 47, 4144.
(86) Hasegawa, S.; Horike, S.; Matsuda, R.; Furukawa, S.; Mochizuki, K.; Kinoshita, Y.; Kitagawa, S.
J. Am. Chem. Soc. 2007, 129, 2607.
(87) Alkordi, M. H.; Liu, Y. L.; Larsen, R. W.; Eubank, J. F.; Eddaoudi, M. J. Am. Chem. Soc. 2008,
References
141
130, 12639.
(88) Jiang, D.; Mallat, T.; Krumeich, F.; Baiker, A. J. Catal. 2008, 257, 390.
(89) Corma, A.; Garca, H.; Llabrs i Xamena, F. X. Chem. Rev. 2010, 110, 4606.
(90) Rowsell, J. L. C.; Millward, A. R.; Park, K. S.; Yaghi, O. M. J. Am. Chem. Soc. 2004, 126, 5666.
(91) Rowsell, J. L. C.; Eckert, J.; Yaghi, O. M. J. Am. Chem. Soc. 2005, 127, 14904.
(92) Zhao, X. B.; Xiao, B.; Fletcher, A. J.; Thomas, K. M.; Bradshaw, D.; Rosseinsky, M. J. Science
2004, 306, 1012.
(93) Chen, B. L.; Ockwig, N. W.; Millward, A. R.; Contreras, D. S.; Yaghi, O. M. Angew. Chem. Int.
Edit. 2005, 44, 4745.
(94) Pan, L.; Sander, M. B.; Huang, X. Y.; Li, J.; Smith, M.; Bittner, E.; Bockrath, B.; Johnson, J. K. J.
Am. Chem. Soc. 2004, 126, 1308.
(95) Rowsell, J. L. C.; Yaghi, O. M. Angew. Chem. Int. Edit. 2005, 44, 4670.
(96) Rowsell, J. L. C.; Yaghi, O. M. J. Am. Chem. Soc. 2006, 128, 1304.
(97) Wong-Foy, A. G.; Matzger, A. J.; Yaghi, O. M. J. Am. Chem. Soc. 2006, 128, 3494.
(98) Dinca, M.; Dailly, A.; Liu, Y.; Brown, C. M.; Neumann, D. A.; Long, J. R. J. Am. Chem. Soc.
2006, 128, 16876.
(99) Dinca, M.; Long, J. R. J. Am. Chem. Soc. 2007, 129, 11172.
(100) Ma, S.; Sun, D.; Ambrogio, M.; Fillinger, J. A.; Parkin, S.; Zhou, H.-C. J. Am. Chem. Soc. 2007,
129, 1858.
(101) Wang, X. S.; Ma, S. Q.; Rauch, K.; Simmons, J. M.; Yuan, D. Q.; Wang, X. P.; Yildirim, T.; Cole,
W. C.; Lopez, J. J.; de Meijere, A.; Zhou, H. C. Chem. Mater. 2008, 20, 3145.
(102) Wang, X. S.; Ma, S. Q.; Forster, P. M.; Yuan, D. Q.; Eckert, J.; Lopez, J. J.; Murphy, B. J.; Parise,
J. B.; Zhou, H. C. Angew. Chem. Int. Edit. 2008, 47, 7263.
(103) Lin, X.; Telepeni, I.; Blake, A. J.; Dailly, A.; Brown, C. M.; Simmons, J. M.; Zoppi, M.; Walker,
G. S.; Thomas, K. M.; Mays, T. J.; Hubberstey, P.; Champness, N. R.; Schroder, M. J. Am. Chem. Soc.
2009, 131, 2159.
(104) Yan, Y.; Lin, X.; Yang, S. H.; Blake, A. J.; Dailly, A.; Champness, N. R.; Hubberstey, P.;
Schroder, M. Chem. Commu. 2009, 1025.
(105) Ibarra, I. A.; Yang, S.; Lin, X.; Blake, A. J.; Rizkallah, P. J.; Nowell, H.; Allan, D. R.;
Champness, N. R.; Hubberstey, P.; Schroder, M. Chem. Commun. 2011, 47, 8304.
(106) Yang, S. H.; Lin, X.; Dailly, A.; Blake, A. J.; Hubberstey, P.; Champness, N. R.; Schroder, M.
Chem.-Eur. J. 2009, 15, 4829.
(107) 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.
(108) Lssig, D.; Lincke, J.; Moellmer, J.; Reichenbach, C.; Moeller, A.; Glser, R.; Kalies, G.;
Cychosz, K. A.; Thommes, M.; Staudt, R.; Krautscheid, H. Angew. Chem. Int. Edit. 2011, 50, 10344.
(109) Mulfort, K. L.; Wilson, T. M.; Wasielewski, M. R.; Hupp, J. T. Langmuir 2009, 25, 503.
(110) Xiang, Z.; Hu, Z.; Yang, W.; Cao, D. Int. J. Hydrogen Energy.
(111) Zhao, Y. G.; Wu, H. H.; Emge, T. J.; Gong, Q. H.; Nijem, N.; Chabal, Y. J.; Kong, L. Z.;
Langreth, D. C.; Liu, H.; Zeng, H. P.; Li, J. Chem.-Eur. J. 2011, 17, 5101.
(112) Li, Y. W.; Yang, R. T. J. Am. Chem. Soc. 2006, 128, 726.
(113) Li, Y. W.; Yang, R. T. J. Am. Chem. Soc. 2006, 128, 8136.
(114) Li, Y.; Yang, R. T. Langmuir 2007, 23, 12937.
(115) Cao, W.; Li, Y.; Wang, L.; Liao, S. J. Phys. Chem. C 2011, 115, 13829.
References
142
(116) Yang, S. J.; Cho, J. H.; Nahm, K. S.; Park, C. R. Int. J. Hydrogen Energy 2010, 35, 13062.
(117) Prasanth, K. P.; Rallapalli, P.; Raj, M. C.; Bajaj, H. C.; Jasra, R. V. Int. J. Hydrogen Energy 2011,
36, 7594.
(118) Murray, L. J.; Dinca, M.; Long, J. R. Chem. Soc. Rev 2009, 38, 1294.
(119) Kim, H.; Samsonenko, D. G.; Das, S.; Kim, G. H.; Lee, H. S.; Dybtsev, D. N.; Berdonosova, E.
A.; Kim, K. Chem. Asian J. 2009, 4, 886.
(120) Senkovska, I.; Kaskel, S. Micro. Meso. Mater. 2008, 112, 108.
(121) Wu, H.; Zhou, W.; Yildirim, T. J. Am. Chem. Soc. 2009, 131, 4995.
(122) Wang, M.; Xie, M. H.; Wu, C. D.; Wang, Y. G. Chem. Commun. 2009, 2396.
(123) Ma, S. Q.; Sun, D. F.; Simmons, J. M.; Collier, C. D.; Yuan, D. Q.; Zhou, H. C. J. Am. Chem.
Soc. 2008, 130, 1012.
(124) Li, H.; Eddaoudi, M.; Groy, T. L.; Yaghi, O. M. J. Am. Chem. Soc. 1998, 120, 8571.
(125) Millward, A. R.; Yaghi, O. M. J. Am. Chem. Soc. 2005, 127, 17998.
(126) Arstad, B.; Fjellvag, H.; Kongshaug, K. O.; Swang, O.; Blom, R. Adsorption 2008, 14, 755.
(127) Vaidhyanathan, R.; Iremonger, S. S.; Dawson, K. W.; Shimizu, G. K. H. Chem. Commun. 2009,
5230.
(128) Couck, S.; Denayer, J. F. M.; Baron, G. V.; Remy, T.; Gascon, J.; Kapteijn, F. J. Am. Chem. Soc.
2009, 131, 6326.
(129) Demessence, A.; D'Alessandro, D. M.; Foo, M. L.; Long, J. R. J. Am. Chem. Soc. 2009, 131,
8784.
(130) Serra-Crespo, P.; Ramos-Fernandez, E. V.; Gascon, J.; Kapteijn, F. Chem. Mater. 2011, 23, 2565.
(131) Vaidhyanathan, R.; Liang, J. M.; Iremonger, S. S.; Shimizu, G. K. H. Supramol. Chem. 2011, 23,
278.
(132) Deng, H.; Doonan, C. J.; Furukawa, H.; Ferreira, R. B.; Towne, J.; Knobler, C. B.; Wang, B.;
Yaghi, O. M. Science 2010, 327, 846.
(133) Llewellyn, P. L.; Bourrelly, S.; Serre, C.; Vimont, A.; Daturi, M.; Hamon, L.; De Weireld, G.;
Chang, J. S.; Hong, D. Y.; Hwang, Y. K.; Jhung, S. H.; Ferey, G. Langmuir 2008, 24, 7245.
(134) Chowdhury, P.; Bikkina, C.; Gumma, S. J. Phys. Chem. C. 2009, 113, 6616.
(135) Caskey, S. R.; Wong-Foy, A. G.; Matzger, A. J. J. Am. Chem. Soc. 2008, 130, 10870.
(136) Dietzel, P. D. C.; Johnsen, R. E.; Fjellvag, H.; Bordiga, S.; Groppo, E.; Chavan, S.; Blom, R.
Chem. Commu. 2008, 5125.
(137) 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.
(138) Liu, J.; Wang, Y.; Benin, A. I.; Jakubczak, P.; Willis, R. R.; LeVan, M. D. Langmuir 2010, 26,
14301.
(139) Schrock, K.; Schroder, F.; Heyden, M.; Fischer, R. A.; Havenith, M. Phys. Chem. Chem. Phys.
2008, 10, 4732.
(140) Kusgens, P.; Rose, M.; Senkovska, I.; Frode, H.; Henschel, A.; Siegle, S.; Kaskel, S. Micro.
Meso. Mater. 2009, 120, 325.
(141) Cychosz, K. A.; Matzger, A. J. Langmuir 2010, 26, 17198.
(142) Ehrenmann, J.; Henninger, S. K.; Janiak, C. Eur. J. Inorg. Chem. 2011, 2011, 471.
(143) Wang, Q.; Shen, D.; Blow, M.; Lau, M.; Deng, S.; Fitch, F. R.; Lemcoff, N. O.; Semanscin, J.
Micro. Meso. Mater. 2002, 55, 217.
(144) Kondo, A.; Daimaru, T.; Noguchi, H.; Ohba, T.; Kaneko, K.; Kanob, H. J. Colloid Interface Sci.
References
143
2007, 314, 422.
(145) Hu, S.; Zhang, J. P.; Li, H. X.; Tong, M. L.; Chen, X. M.; Kitagawat, S. Cryst. Growth Des. 2007,
7, 2286.
(146) Gu, J. Z.; Lu, W. G.; Jiang, L.; Zhou, H. C.; Lu, T. B. Inorg. Chem. 2007, 46, 5835.
(147) Chen, B. L.; Ji, Y. Y.; Xue, M.; Fronczek, F. R.; Hurtado, E. J.; Mondal, J. U.; Liang, C. D.; Dai,
S. Inorg. Chem. 2008, 47, 5543.
(148) Pan, L.; Olson, D. H.; Ciemnolonski, L. R.; Heddy, R.; Li, J. Angew. Chem. Int. Edit. 2006, 45,
616.
(149) Pan, L.; Parker, B.; Huang, X. Y.; Olson, D. H.; Lee, J.; Li, J. J. Am. Chem. Soc. 2006, 128,
4180.
(150) Lee, J. Y.; Olson, D. H.; Pan, L.; Emge, T. J.; Li, J. Adv. Funct. Mater. 2007, 17, 1255.
(151) Lively, R. P.; Dose, M. E.; Thompson, J. A.; McCool, B. A.; Chance, R. R.; Koros, W. J. Chem.
Commun. 2011, 47, 8667.
(152) Dybtsev, D. N.; Chun, H.; Yoon, S. H.; Kim, D.; Kim, K. J. Am. Chem. Soc. 2004, 126, 32.
(153) Dinca, M.; Long, J. R. J. Am. Chem. Soc. 2005, 127, 9376.
(154) Ma, S. Q.; Wang, X. S.; Collier, C. D.; Manis, E. S.; Zhou, H. C. Inorg. Chem. 2007, 46, 8499.
(155) Ma, S. Q.; Wang, X. S.; Yuan, D. Q.; Zhou, H. C. Angew. Chem. Int. Edit. 2008, 47, 4130.
(156) Navarro, J. A. R.; Barea, E.; Rodriguez-Dieguez, A.; Salas, J. M.; Ania, C. O.; Parra, J. B.;
Masciocchi, N.; Galli, S.; Sironi, A. J. Am. Chem. Soc. 2008, 130, 3978.
(157) Chang, Z.; Zhang, D. S.; Chen, Q.; Li, R. F.; Hu, T. L.; Bu, X. H. Inorg. Chem. 2011, 50, 7555.
(158) Loiseau, T.; Lecroq, L.; Volkringer, C.; Marrot, J.; Ferey, G.; Haouas, M.; Taulelle, F.; Bourrelly,
S.; Llewellyn, P. L.; Latroche, M. J. Am. Chem. Soc. 2006, 128, 10223.
(159) Ma, S.; Wang, X. S.; Manis, E. S.; Collier, C. D.; Zhou, H. C. Inorg. Chem. 2007, 46, 3432.
(160) Liu, X.; Oh, M.; Lah, M. S. Cryst. Growth Des. 2011, ASAP.
(161) Matsuda, R.; Kitaura, R.; Kitagawa, S.; Kubota, Y.; Belosludov, R. V.; Kobayashi, T. C.;
Sakamoto, H.; Chiba, T.; Takata, M.; Kawazoe, Y.; Mita, Y. Nature 2005, 436, 238.
(162) Zou, Y.; Hong, S.; Park, M.; Chun, H.; Lah, M. S. Chem. Commun. 2007, 5182.
(163) Moon, H. R.; Kobayashi, N.; Suh, M. P. Inorg. Chem. 2006, 45, 8672.
(164) Bae, Y. S.; Farha, O. K.; Spokoyny, A. M.; Mirkin, C. A.; Hupp, J. T.; Snurr, R. Q. Chem.
Commun. 2008, 4135.
(165) Mu, B.; Li, F.; Walton, K. S. Chem. Commu. 2009, 2493.
(166) Bae, Y. S.; Hauser, B. G.; Farha, O. K.; Hupp, J. T.; Snurr, R. Q. Micro. Meso. Mater. 2011, 141,
231.
(167) Zhang, S.-M.; Chang, Z.; Hu, T.-L.; Bu, X.-H. Inorg. Chem. 2010, 49, 11581.
(168) An, J.; Geib, S. J.; Rosi, N. L. J. Am. Chem. Soc. 2010, 132, 38.
(169) Banerjee, R.; Furukawa, H.; Britt, D.; Knobler, C.; OKeeffe, M.; Yaghi, O. M. J. Am. Chem.
Soc. 2009, 131, 3875.
(170) Bae, Y. S.; Farha, O. K.; Hupp, J. T.; Snurr, R. Q. J. Mater. Chem. 2009, 19, 2131.
(171) Zheng, B.; Bai, J.; Duan, J.; Wojtas, L.; Zaworotko, M. J. J. Am. Chem. Soc. 2010, 133, 748.
(172) Henke, S.; Fischer, R. A. J. Am. Chem. Soc. 2011, null.
(173) Chen, B. L.; Liang, C. D.; Yang, J.; Contreras, D. S.; Clancy, Y. L.; Lobkovsky, E. B.; Yaghi, O.
M.; Dai, S. Angew. Chem. Int. Edit. 2006, 45, 1390.
(174) Barcia, P. S.; Zapata, F.; Silva, J. A. C.; Rodrigues, A. E.; Chen, B. L. J. Phys. Chem. B. 2007,
111, 6101.
References
144
(175) Li, K.; Olson, D. H.; Lee, J. Y.; Bi, W.; Wu, K.; Yuen, T.; Xu, Q.; Li, J. Adv. Funct. Mater. 2008,
18, 2205.
(176) Chmelik, C.; Karger, J.; Wiebckw, M.; Caro, J.; van Baten, J. M.; Krishna, R. Micro. Meso.
Mater. 2009, 117, 22.
(177) Luebbers, M. T.; Wu, T.; Shen, L.; Masel, R. I. Langmuir 2010, 26, 11319.
(178) Gucuyener, C.; van den Bergh, J.; Gascon, J.; Kapteijn, F. J. Am. Chem. Soc. 2010, 132, 17704.
(179) Bao, Z.; Alnemrat, S.; Yu, L.; Vasiliev, I.; Ren, Q.; Lu, X.; Deng, S. Langmuir 2011, 27, 13554.
(180) Trung, T. K.; Trens, P.; Tanchoux, N.; Bourrelly, S.; Llewellyn, P. L.; Loera-Serna, S.; Serre, C.;
Loiseau, T.; Fajula, F.; Ferey, G. J. Am. Chem. Soc. 2008, 130, 16926.
(181) Llewellyn, P. L.; Horcajada, P.; Maurin, G.; Devic, T.; Rosenbach, N.; Bourrelly, S.; Serre, C.;
Vincent, D.; Loera-Serna, S.; Filinchuk, Y.; Ferey, G. J. Am. Chem. Soc. 2009, 131, 13002.
(182) Ramsahye, N. A.; Trung, T. K.; Bourrelly, S.; Yang, Q.; Devic, T.; Maurin, G.; Horcajada, P.;
Llewellyn, P. L.; Yot, P.; Serre, C.; Filinchuk, Y.; Fajula, F. o.; Frey, G. r.; Trens, P. J. Phys. Chem. C
2011, 115, 18683.
(183) Kawakami, T.; Takamizawa, S.; Kitagawa, Y.; Maruta, T.; Mori, W.; Yamaguchi, K. Polyhedron
2001, 20, 1197.
(184) Sagara, T.; Klassen, J.; Ganz, E. J. Chem. Phys. 2004, 121, 12543.
(185) Sagara, T.; Klassen, J.; Ortony, J.; Ganz, E. J. Chem. Phys. 2005, 123, 014701.
(186) Sagara, T.; Ortony, J.; Ganz, E. J. Chem. Phys. 2005, 123, 214707.
(187) Yang, Q. Y.; Zhong, C. L. J. Phys. Chem. B. 2005, 109, 11862.
(188) Zhang, L.; Wang, Q.; Liu, Y. C.; Wu, T.; Chen, D.; Wang, X. P. Chem. Phys. Lett. 2009, 469,
261.
(189) Zhou, M.; Wang, Q.; Zhang, L.; Liu, Y. C.; Kang, Y. J. Phys. Chem. B. 2009, 113, 11049.
(190) Han, S. S.; Choi, S. H.; Goddard, W. A. J. Phys. Chem. C. 2010, 114, 12039.
(191) Garberoglio, G.; Skoulidas, A. I.; Johnson, J. K. J. Phys. Chem. B. 2005, 109, 13094.
(192) 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.
(193) Xu, Q.; Liu, D.; Yang, Q.; Zhong, C. Mol. Sim. 2009, 35, 748.
(194) Frost, H.; Duren, T.; Snurr, R. Q. J. Phys. Chem. B. 2006, 110, 9565.
(195) Frost, H.; Snurr, R. Q. J. Phys. Chem. C. 2007, 111, 18794.
(196) Assfour, B.; Seifert, G. Int. J. Hydrogen Energy 2009, 34, 8135.
(197) Jung, D. H.; Kim, D.; Lee, T. B.; Choi, S. B.; Yoon, J. H.; Kim, J.; Choi, K.; Choi, S. H. J. Phys.
Chem. B. 2006, 110, 22987.
(198) Ryan, P.; Broadbelt, L. J.; Snurr, R. Q. Chem. Commun. 2008, 4132.
(199) Han, S. S.; Mendoza-Cortes, J. L.; Goddard, W. A. Chem. Soc. Rev. 2009, 38, 1460.
(200) Kosa, M.; Krack, M.; Cheetham, A. K.; Parrinello, M. J. Phys. Chem. C 2008, 112, 16171.
(201) Sun, Y. Y.; Kim, Y. H.; Zhang, S. B. J. Am. Chem. Soc. 2007, 129, 12606.
(202) Yang, Q. Y.; Zhong, C. L. J. Phys. Chem. B. 2006, 110, 655.
(203) Fischer, M.; Kuchta, B.; Firlej, L.; Hoffmann, F.; Frba, M. J. Phys. Chem. C. 2010, 114, 19116.
(204) Han, S. S.; Goddard, W. A. J. Am. Chem. Soc. 2007, 129, 8422.
(205) Blomqvist, A.; Araujo, C. M.; Srepusharawoot, P.; Ahuja, R. Proc. Natl. Acad. Sci. U.S.A 2007,
104, 20173.
(206) Klontzas, E.; Mavrandonakis, A.; Tylianakis, E.; Froudakis, G. E. Nano. Lett. 2008, 8, 1572.
(207) Mavrandonakis, A.; Tylianakis, E.; Stubos, A. K.; Froudakis, G. E. J. Phys. Chem. C. 2008, 112,
References
145
7290.
(208) Dalach, P.; Frost, H.; Snurr, R. Q.; Ellis, D. E. J. Phys. Chem. C. 2008, 112, 9278.
(209) Choi, Y. J.; Lee, J. W.; Choi, J. H.; Kang, J. K. Appl. Phys. Lett. 2008, 92, 173102.
(210) Cao, D.; Lan, J.; Wang, W.; Smit, B. Angew. Chem. Int. Edit. 2009, 48, 4730.
(211) Stergiannakos, T.; Tylianakis, E.; Klontzas, E.; Froudakis, G. E. J. Phys. Chem. C. 2010, 114,
16855.
(212) Maark, T. A.; Pal, S. Int. J. Hydrogen Energy 2010, 35, 12846.
(213) Han, S. S.; Choi, S. H.; Goddard, W. A. J. Phys. Chem. C. 2011, 115, 3507.
(214) Mavrandonakis, A.; Klontzas, E.; Tylianakis, E.; Froudakis, G. E. J. Am. Chem. Soc. 2009, 131,
13410.
(215) Getman, R. B.; Miller, J. H.; Wang, K.; Snurr, R. Q. J. Phys. Chem. C. 2011, 115, 2066.
(216) Babarao, R.; Eddaoudi, M.; Jiang, J. W. Langmuir 2010, 26, 11196.
(217) Duren, T.; Snurr, R. Q. J. Phys. Chem. B. 2004, 108, 15703.
(218) Dren, T.; Snurr, R. Q. Using molecular simulation to characterise metal-organic frameworks
and judge their performance as adsorbents. In Stud. Surf. Sci. Catal.; P.L. Llewellyn, F. R.-R. J. R.,
Seaton, N., Eds.; Elsevier, 2006; Vol. Volume 160; pp 161.
(219) Jhon, Y. H.; Cho, M.; Jeon, H. R.; Park, I.; Chang, R.; Rowsell, J. L. C.; Kim, J. J. Phys. Chem.
C 2007, 111, 16618.
(220) Zeng, Y. Y.; Zhang, B. J. Acta Phys. Chim. Sin. 2008, 24, 1493.
(221) Liu, D.; Yang, Q.; Zhong, C. Mol. Sim. 2009, 35, 213.
(222) Gallo, M.; Glossman-Mitnik, D. J. Phys. Chem. C. 2009, 113, 6634.
(223) Thornton, A. W.; Nairn, K. M.; Hill, J. M.; Hill, A. J.; Hill, M. R. J. Am. Chem. Soc. 2009, 131,
10662.
(224) Zhou, Z.; Xue, C. Y.; Yang, Q. Y.; Zhong, C. L. Acta Chim. Sin. 2009, 67, 477.
(225) Xue, C.; Zhou, Z. e.; Yang, Q.; Zhong, C. Chin. J. Chem. Eng. 2009, 17, 580.
(226) Ye, Q.; Yan, S.; Liu, D.; Yang, Q.; Zhong, C. Mol. Sim. 2010, 36, 682.
(227) Babarao, R.; Jiang, J. W. Energy Environ. Sci. 2008, 1, 139.
(228) Walton, K. S.; Millward, A. R.; Dubbeldam, D.; Frost, H.; Low, J. J.; Yaghi, O. M.; Snurr, R. Q.
J. Am. Chem. Soc. 2008, 130, 406.
(229) Wilmer, C. E.; Snurr, R. Q. Chem. Eng. J. 2011, 171, 775.
(230) Yang, Q. Y.; Zhong, C. L.; Chen, J. F. J. Phys. Chem. C 2008, 112, 1562.
(231) Yang, Q.; Zhong, C. Langmuir 2009, 25, 2302.
(232) Liu, D.; Zheng, C.; Yang, Q.; Zhong, C. J. Phys. Chem. C. 2009, 113, 5004.
(233) Zheng, C.; Liu, D.; Yang, Q.; Zhong, C.; Mi, J. Ind. Eng. Chem. Res. 2009, 48, 10479.
(234) Ramsahye, N. A.; Maurin, G.; Bourrelly, S.; Llewellyn, P. L.; Loiseau, T.; Serre, C.; Ferey, G.
Chem. Commun. 2007, 3261.
(235) Ramsahye, N. A.; Maurin, G.; Bourrelly, S.; Llewellyn, P. L.; Devic, T.; Serre, C.; Loiseau, T.;
Ferey, G. Adsorption 2007, 13, 461.
(236) Ramsahye, N. A.; Maurin, G.; Bourrelly, S.; Llewellyn, P. L.; Serre, C.; Loiseau, T.; Devic, T.;
Ferey, G. J. Phys. Chem. C. 2008, 112, 514.
(237) Torrisi, A.; Bell, R. G.; Mellot-Draznieks, C. Cryst. Growth Des. 2010, 10, 2839.
(238) Greathouse, J. A.; Allendorf, M. D. J. Am. Chem. Soc. 2006, 128, 10678.
(239) Castillo, J. M.; Vlugt, T. J. H.; Calero, S. J. Phys. Chem.C 2008, 112, 15934.
(240) Watanabe, T.; Sholl, D. S. J. Chem. Phys. 2010, 133, 094509.
References
146
(241) Grajciar, L. s.; Bludsk, O.; Nachtigall, P. J. Phys. Chem. Lett. 2010, 1, 3354.
(242) Nalaparaju, A.; Babarao, R.; Zhao, X. S.; Jiang, J. W. Acs Nano 2009, 3, 2563.
(243) Salles, F.; Bourrelly, S.; Jobic, H.; Devic, T.; Guillerm, V.; Llewellyn, P.; Serre, C.; Ferey, G.;
Maurin, G. J. Phys. Chem. C. 2011, 115, 10764.
(244) 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.
(245) Martin-Calvo, A.; Garcia-Perez, E.; Castillo, J. M.; Calero, S. Phys. Chem. Chem. Phys. 2008,
10, 7085.
(246) Babarao, R.; Jiang, J. W.; Sandler, S. I. Langmuir 2009, 25, 5239.
(247) Jiang, J. AIChE J. 2009, 55, 2422.
(248) Wang, S. Y.; Yang, Q. Y.; Zhong, C. L. J. Phys. Chem. B. 2006, 110, 20526.
(249) Yang, Q. Y.; Zhong, C. L. Chemphyschem 2006, 7, 1417.
(250) Yang, Q. Y.; Xue, C. Y.; Zhong, C. L.; Chen, J. F. AIChE J. 2007, 53, 2832.
(251) Xue, C.; Yang, Q.; Zhong, C. Mol. Sim. 2009, 35, 1249.
(252) Wang, S. Y.; Yang, Q. Y.; Zhong, C. L. Sep. Purif. Tech. 2008, 60, 30.
(253) Liu, Y.; Liu, D.; Yang, Q.; Zhong, C.; Mi, J. Ind. Eng. Chem. Res. 2010, 49, 2902.
(254) Xu, Q.; Liu, D.; Yang, Q.; Zhong, C.; Mi, J. J. Mater. Chem. 2010, 20, 706.
(255) Mu, W.; Liu, D.; Zhong, C. Micro. Meso. Mater. 2011, 143, 66.
(256) Mu, W.; Liu, D.; Yang, Q.; Zhong, C. Micro. Meso. Mater. 2010, 130, 76.
(257) Huang, H.; Zhang, W.; Liu, D.; Liu, B.; Chen, G.; Zhong, C. Chem. Eng. Sci. 2011, 66, 6297.
(258) Wu, D.; Wang, C.; Liu, B.; Liu, D.; Yang, Q.; Zhong, C. AIChE J. 2011, in press.
(259) Liu, B.; Yang, Q.; Xue, C.; Zhong, C.; Chen, B.; Smit, B. J. Phys. Chem. C 2008, 112, 9854.
(260) Liu, B.; Smit, B. Langmuir 2009, 25, 5918.
(261) Liu, B.; Smit, B. J. Phys. Chem. C 2010, 114, 8515.
(262) Liu, B.; Sun, C. Y.; Chen, G. J. Chem. Eng. Sci. 2011, 66, 3012.
(263) Jiang, J. W.; Sandler, S. I. Langmuir 2006, 22, 5702.
(264) Dubbeldam, D.; Galvin, C. J.; Walton, K. S.; Ellis, D. E.; Snurr, R. Q. J. Am. Chem. Soc. 2008,
130, 10884.
(265) Krishna, R.; van Baten, J. M. Mol. Sim. 2009, 35, 1098.
(266) Jorge, M.; Lamia, N.; Rodrigues, A. E. Colloids. Surf. A 2010, 357, 27.
(267) Droche, I.; Rives, S.; Trung, T.; Yang, Q.; Ghoufi, A.; Ramsahye, N. A.; Trens, P.; Fajula, F.;
Devic, T.; Serre, C.; Frey, G.; Jobic, H.; Maurin, G. J. Phys. Chem. C 2011, 115, 13868.
(268) Dybtsev, D. N.; Chun, H.; Kim, K. Angew. Chem. Int. Edit. 2004, 43, 5033.
(269) Hong, D. Y.; Hwang, Y. K.; Serre, C.; Ferey, G.; Chang, J. S. Adv. Funct. Mater. 2009, 19, 1537.
(270) Jhung, S. H.; Lee, J. H.; Yoon, J. W.; Serre, C.; Ferey, G.; Chang, J. S. Adv. Mater. 2007, 19, 121.
(271) 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. Edit. 2006, 45, 8227.
(272) Hamon, L.; Serre, C.; Devic, T.; Loiseau, T.; Millange, F.; Ferey, G.; De Weireld, G. J. Am. Chem.
Soc. 2009, 131, 8775.
(273) Gu, Z.-Y.; Yan, X.-P. Angew. Chem. 2010, 122, 1519.
(274) Krishna, R.; van Baten, J. M. Chem. Eng. Sci. 2009, 64, 3159.
(275) An, J.; Fiorella, R. P.; Geib, S. J.; Rosi, N. L. J. Am. Chem. Soc. 2009, 131, 8401.
(276) Babarao, R.; Jiang, J. W. Ind. Eng. Chem. Res. 2011, 50, 62.
(277) Shiflett, M. B.; Drew, D. W.; Cantini, R. A.; Yokozeki, A. Energy Fuels 2010, 24, 5781.
References
147
(278) Zhu, S. D.; Wu, Y. X.; Chen, Q. M.; Yu, Z. N.; Wang, C. W.; Jin, S. W.; Ding, Y. G.; Wu, G.
Green Chem. 2006, 8, 325.
(279) Plechkova, N. V.; Seddon, K. R. Chem. Soc. Rev. 2008, 37, 123.
(280) Blanchard, L. A.; Hancu, D.; Beckman, E. J.; Brennecke, J. F. Nature 1999, 399, 28.
(281) Bara, J. E.; Camper, D. E.; Gin, D. L.; Noble, R. D. Acc. Chem. Res. 2010, 43, 152.
(282) Karadas, F.; Atilhan, M.; Aparicio, S. Energy Fuels 2010, 24, 5817.
(283) Hu, Y. F.; Liu, Z. C.; Xu, C. M.; Zhang, X. M. Chem. Soc. Rev. 2011, 40, 3802.
(284) Hasib-ur-Rahman, M.; Siaj, M.; Larachi, F. Chem. Eng. Process. 2010, 49, 313.
(285) Shiflett, M. B.; Yokozeki, A. Ind. Eng. Chem. Res. 2005, 44, 4453.
(286) Bara, J. E.; Gabriel, C. J.; Lessmann, S.; Carlisle, T. K.; FInotello, A.; Gin, D. L.; Noble, R. D.
Ind. Eng. Chem. Res. 2007, 46, 5380.
(287) Muldoon, M. J.; Aki, S. N. V. K.; Anderson, J. L.; Dixon, J. K.; Brennecke, J. F. J. Phys. Chem.
B. 2007, 111, 9001.
(288) Shi, W.; Maginn, E. J. J. Phys. Chem. B. 2008, 112, 2045.
(289) Shi, W.; Maginn, E. J. J. Phys. Chem. B 2008, 112, 16710.
(290) Gonzalez-Miquel, M.; Palomar, J.; Omar, S.; Rodriguez, F. Ind. Eng. Chem. Res. 2011, 50, 5739.
(291) Bates, E. D.; Mayton, R. D.; Ntai, I.; Davis Jr, J. H. J. Am. Chem. Soc. 2002, 124, 926.
(292) Wang, C.; Luo, X.; Luo, H.; Jiang, D. E.; Li, H. R.; Dai, S. Angew. Chem. Int. Ed. 2011, 50,
4918.
(293) Lozano, L. J.; Godinez, C.; de los Rios, A. P.; Hernandez-Fernandez, F. J.; Sanchez-Segado, S.;
Alguacil, F. J. J. Membr. Sci. 2011, 376, 1.
(294) Scovazzo, P.; Kieft, J.; Finan, D. A.; Koval, C.; DuBois, D.; Noble, R. D. J. Membr. Sci. 2004,
238, 57.
(295) Myers, C.; Pennline, H.; Luebke, D.; Ilconich, J.; Dixon, J. K.; Maginn, E. J.; Brennecke, J. F. J.
Membr. Sci. 2008, 322, 28.
(296) Neves, L. A.; Crespo, J. G.; Coelhoso, I. M. J. Membr. Sci. 2010, 357, 160.
(297) Scovazzo, P. J. Membr. Sci. 2009, 343, 199.
(298) Le Bideau, J.; Viau, L.; Vioux, A. Chem. Soc. Rev. 2011, 40, 907.
(299) Duren, T.; Bae, Y. S.; Snurr, R. Q. Chem. Soc. Rev. 2009, 38, 1203.
(300) Jiang, J. W.; Babarao, R.; Hu, Z. Q. Chem. Soc. Rev. 2011, 40, 3599.
(301) Singh, U. C.; Kollman, P. A. J. Comput. Chem. 1984, 5, 129.
(302) Rappe, A. K.; Casewit, C. J.; Colwell, K. S.; Goddard, W. A.; Skiff, W. M. J. Am. Chem. Soc.
1992, 114, 10024.
(303) Mayo, S. L.; Olafson, B. D.; Goddard, W. A. J. Phys. Chem. 1990, 94, 8897.
(304) Sun, H.; Mumby, S. J.; Maple, J. R.; Hagler, A. T. J. Am. Chem. Soc. 1994, 116, 2978.
(305) Sun, H. J. Phys. Chem. B 1998, 102, 7338.
(306) Cornell, W. D.; Cieplak, P.; Bayly, C. I.; Gould, I. R.; Merz, K. M.; Ferguson, D. M.; Spellmeyer,
D. C.; Fox, T.; Caldwell, J. W.; Kollman, P. A. J. Am. Chem. Soc. 1995, 117, 5179.
(307) Dauber-Osguthorpe, P.; Roberts, V. A.; Osguthorpe, D. J.; Wolff, J.; Genest, M.; Hagler, A. T.
Proteins: Structure, Function, and Bioinformatics 1988, 4, 31.
(308) Brooks, B. R.; Bruccoleri, R. E.; Olafson, B. D.; States, D. J.; Swaminathan, S.; Karplus, M. J.
Comput. Chem. 1983, 4, 187.
(309) Jorgensen, W. L.; Maxwell, D. S.; Tirado-Rives, J. J. Am. Chem. Soc. 1996, 118, 11225.
(310) Christen, M.; Hnenberger, P. H.; Bakowies, D.; Baron, R.; Brgi, R.; Geerke, D. P.; Heinz, T.
References
148
N.; Kastenholz, M. A.; Krutler, V.; Oostenbrink, C.; Peter, C.; Trzesniak, D.; van Gunsteren, W. F. J.
Comput. Chem. 2005, 26, 1719.
(311) Greathouse, J. A.; Kinnibrugh, T. L.; Allendorf, M. D. Ind. Eng. Chem. Res. 2009, 48, 3425.
(312) Materials Studio. 4.3 ed.; Accelrys, San Diego, 2008.
(313) Besler, B. H.; Merz, K. M.; Kollman, P. A. J. Comp. Chem 1990, 11, 431.
(314) Mulliken, R. S. J. Chem. Phys. 1955, 23, 1833.
(315) Myers, A. L.; Monson, P. A. Langmuir 2002, 18, 10261.
(316) Do, D. D.; Herrera, L. F.; Do, H. D. J. Colloid Interface Sci. 2008, 328, 110.
(317) Hirotani, A.; Mizukami, K.; Miura, R.; Takaba, H.; Miya, T.; Fahmi, A.; Stirling, A.; Kubo, M.;
Miyamoto, A. Appl. Surf. Sci. 1997, 120, 81.
(318) Martin, M. G.; Siepmann, J. I. J. Phys. Chem. B. 1998, 102, 2569.
(319) Jorgensen, W. L.; Chandrasekhar, J.; Madura, J. D.; Impey, R. W.; Klein, M. L. J. Chem. Phys.
1983, 79, 926.
(320) Skoulidas, A. I.; Sholl, D. S. J. Phys. Chem. B. 2005, 109, 15760.
(321) Do, D. D.; Do, H. D. J. Phys. Chem. B 2005, 109, 19288.
(322) Smart, O. S.; Neduvelil, J. G.; Wang, X.; Wallace, B. A.; Sansom, M. S. P. J. Mol. Graphics 1996,
14, 354.
(323) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.;
Zakrzewski, V. G.; Montgomery, J. A.; Stratmann, R. E.; Burant, J. C.; Dapprich, S.; Millam, J. M.;
Daniels, A. D.; Kudin, K. N.; Strain, M. C.; Farkas, O.; Tomasi, J.; Barone, V.; Cossi, M.; Cammi, R.;
Mennucci, B.; Pomelli, C.; Adamo, C.; Clifford, S.; Ochterski, J.; Petersson, G. A.; Ayala, P. Y.; Cui, Q.;
Morokuma, K.; Malick, D. K.; Rabuck, A. D.; Raghavachari, K.; Foresman, J. B.; Cioslowski, J.; Ortiz,
J. V.; Stefanov, B. B.; Liu, G.; Liashenko, A.; Piskorz, P.; Komaromi, I.; Gomperts, R.; Martin, R. L.;
Fox, D. J.; Keith, T.; Al-Laham, M. A.; Peng, C. Y.; Nanayakkara, A.; Gonzalez, C.; Challacombe, M.;
Gill, P. M. W.; Johnson, B. G.; Chen, W.; Wong, M. W.; Andres, J. L.; Head-Gordon, M.; Replogle, E.
S.; Pople, J. A. Gaussian 03, Revision D.01 ed.; Gaussian, Inc.: Wallingford CT, 2004.
(324) Mark, P.; Nilsson, L. J. Phys. Chem. A 2001, 105, 9954.
(325) Chen, B.; Potoff, J. J.; Siepmann, J. I. J. Phys. Chem. B. 2001, 105, 3093.
(326) P.J. Linstrom, W. G. M., Eds. NIST Chemistry WebBook, NIST Standard Reference database
Number 69, June 2005, National Institute of Standards and Technology, Gaithersburg MD, 20899
(327) Lee, J. Y. Synthesis and Gas Sorption Study of Microporous Metal Organic Frameworks for
Hydrogen and Methane Storage. PhD, The State University of New Jersey, 2007.
(328) Liang, Z.; Marshall, M.; Chaffee, A. L. Micro. Meso. Mater. 2010, 132, 305.
(329) Babarao, R.; Jiang, J. W. Energy Environ. Sci. 2009, 2, 1088.
(330) Moise, J. C.; Bellat, J. P. J. Phys. Chem. B. 2005, 109, 17239.
(331) Hariharan, P. C.; Pople, J. A. Chem. Phys. Lett. 1972, 16, 217.
(332) Breneman, C. M.; Wiberg, K. B. J. Comp. Chem 1990, 11, 361.
(333) Vishnyakov, A.; Ravikovitch, P. I.; Neimark, A. V.; Bulow, M.; Wang, Q. M. Nano. Lett. 2003, 3,
713.
(334) Harris, J. G.; Yung, K. H. J. Phys. Chem. 1995, 99, 12021.
(335) Pantatosaki, E.; Papadopoulos, G. K.; Jobic, H.; Theodorou, D. N. J. Phys. Chem. B. 2008, 112,
11708.
(336) Murthy, C. S.; Singer, K.; Klein, M. L.; McDonald, I. R. Mol. Phys. 1980, 41, 1387.
(337) Ruthven, D. M. Principles of Adsorption and Adsorption Processes; Wiley, Newyork 1984.
References
149
(338) Myers, A. L., and J. M. Prausnitz. AIChE J. 1965, 11, 121.
(339) Choi, S. H.; Drese, J. H.; Jones, C. W. ChemSusChem 2009, 2, 796.
(340) Dubbeldam, D.; Frost, H.; Walton, K. S.; Snurr, R. Q. Fluid Phase Equilib. 2007, 261, 152.
(341) Wu, H.; Zhou, W.; Yildirim, T. J. Am. Chem. Soc. 2007, 129, 5314.
(342) Akten, E. D.; Siriwardane, R.; Sholl, D. S. Energy Fuels 2003, 17, 977.
(343) Yang, Q. Y.; Xu, Q.; Liu, B.; Zhong, C. L.; Smit, B. Chin. J. Chem. Eng. 2009, 17, 781.
(344) Yang, Q. Y.; Zhong, C. L. J. Phys. Chem. B. 2006, 110, 17776.
(345) Cessford, N. F.; Seaton, N. A.; Dren, T. Ind. Eng. Chem. Res. 2012, 51, 4911.
(346) Goj, A.; Sholl, D. S.; Akten, E. D.; Kohen, D. J. Phys. Chem. B. 2002, 106, 8367.
(347) Bernal, M. P.; Coronas, J.; Menendez, M.; Santamaria, J. AIChE J. 2004, 50, 127.
(348) Seike, T.; Matsuda, M.; Miyake, M. J. Mater. Chem. 2002, 12, 366.
(349) Bastin, L.; Barcia, P. S.; Hurtado, E. J.; Silva, J. A. C.; Rodrigues, A. E.; Chen, B. J. Phys. Chem.
C. 2008, 112, 1575.
(350) Darkrim, F.; Levesque, D. J. Chem. Phys. 1998, 109, 4981.
(351) Hirschfelder, J. O.; Curtiss, C. F.; Bird, R. B. Molecular Theory of Gases and liquids; John
Wiley: New York 1964.
(352) Dunne, J. A.; Rao, M.; Sircar, S.; Gorte, R. J.; Myers, A. L. Langmuir 1996, 12, 5896.
(353) Delgado, J. A.; Uguina, M. A.; Gomez, J. M.; Ortega, L. Sep. Purif. Tech. 2006, 48, 223.
(354) Plant, D. F.; Maurin, G.; Deroche, I.; Gaberova, L.; Llewellyn, P. L. Chem. Phys. Lett. 2006, 426,
387.
(355) Maurin, G.; Llewellyn, P.; Poyet, T.; Kuchta, B. J. Phys. Chem. B. 2005, 109, 125.
(356) Pillai, R. S.; Sethia, G.; Jasra, R. V. Ind. Eng. Chem. Res. 2010, 49, 5816.
(357) Peter, S. A.; Sebastian, J.; Jasra, R. V. Ind. Eng. Chem. Res. 2005, 44, 6856.
(358) Maurin, G.; Llewellyn, P. L.; Poyet, T.; Kuchta, B. Micro. Meso. Mater. 2005, 79, 53.
(359) Walton, K. S.; Abney, M. B.; LeVan, M. D. Micro. Meso. Mater. 2006, 91, 78.
(360) Ridha, F. N.; Webley, P. A. Separ Sci. Technol. 2009, 67, 336.
(361) Gallo, M.; Nenoff, T. M.; Mitchell, M. C. Fluid Phase Equilib. 2006, 247, 135.
(362) Cao, D. P.; Wu, J. Z. Carbon 2005, 43, 1364.
(363) Hutson, N. D.; Zajic, S. C.; Yang, R. T. Ind. Eng. Chem. Res. 2000, 39, 1775.
(364) Brandani, F.; Ruthven, D. M. Ind. Eng. Chem. Res. 2004, 43, 8339.
(365) Parnham, E. R.; Morris, R. E. Acc. Chem. Res. 2007, 40, 1005.
(366) Han, L.; Zhang, S.; Wang, Y.; Yan, X.; Lu, X. Inorg. Chem. 2009, 48, 786.
(367) Lopes, J. N. C.; Deschamps, J.; Padua, A. A. H. J. Phys. Chem. B. 2004, 108, 2038.
(368) Van Der Spoel, D.; Lindahl, E.; Hess, B.; Groenhof, G.; Mark, A. E.; Berendsen, H. J. C. J.
Comput. Chem. 2005, 26, 1701.
(369) Hockney, R. W. Methods Comput. Phys 1970, 9, 136.
(370) Berendsen, H. J. C.; Postma, J. P. M.; Vangunsteren, W. F.; Dinola, A.; Haak, J. R. J. Chem. Phys.
1984, 81, 3684.
(371) Gu, Z.; Brennecke, J. F. J. Chem. Eng. Data 2002, 47, 339.
(372) Morrow, T.; Maginn, E. J. Phys. Chem. B. 2002, 106, 12807.
(373) Lee, S. U.; Jung, J.; Han, Y. K. Chem. Phys. Lett. 2005, 406, 332.
(374) Kanakubo, M.; Harri, s. K. R.; Tsuchihashi, N.; Ibuki, K.; Ueno, M. J. Phys. Chem. B. 2007, 111,
2062.
(375) Umecky, T.; Kanakubo, M.; Ikushima, Y. Fluid Phase Equilib. 2005, 228, 329.
References
150
(376) Urahata, S. M.; Ribeiro, M. C. C. J. Chem. Phys. 2005, 122, 024511.
(377) Roy, S.; Baiker, A. Chem. Rev. 2009, 109, 4054.
(378) Bonino, F.; Chavan, S.; Vitillo, J. G.; Groppo, E.; Agostini, G.; Lamberti, C.; Dietzel, P. D. C.;
Prestipino, C.; Bordiga, S. Chem. Mater. 2008, 20, 4957.
(379) Alaerts, L.; Maes, M.; Giebeler, L.; Jacobs, P. A.; Martens, J. A.; Denayer, J. F. M.; Kirschhock,
C. E. A.; De Vos, D. E. J. Am. Chem. Soc. 2008, 130, 14170.
(380) Hu, Y.; Dong, X.; Nan, J.; Jin, W.; Ren, X.; Xu, N.; Lee, Y. M. Chem. Commun. 2011, 47, 737.
(381) Hu, Z.; Chen, Y.; Jiang, J. J. Chem. Phys. 2011, 134, 134705.
(382) Nalaparaju, A.; Zhao, X. S.; Jiang, J. W. Energy Environ. Sci. 2011, 4, 2107.
(383) Dhondt, A.; Vanholder, R.; Biesen, W. V.; Lameire, N. Kidney Int. 2000, 58, S47.
(384) Wernert, V.; Schaf, O.; Ghobarkar, H.; Denoyel, R. Micro. Meso. Mater. 2005, 83, 101.
(385) Wernert, V.; Schaf, O.; Faure, V.; Brunet, P.; Dou, L.; Berland, Y.; Boulet, P.; Kuchta, B.;
Denoyel, R. J. Biotechnol. 2006, 123, 164.
(386) Berge-Lefranc, D.; Pizzala, H.; Paillaud, J. L.; Schaf, O.; Vagner, C.; Boulet, P.; Kuchta, B.;
Denoyel, R. Adsorption 2008, 14, 377.
(387) Narasimhan, L.; Boulet, P.; Kuchta, B.; Schaef, O.; Denoyel, R.; Brunet, P. Langmuir 2009, 25,
11598.


Appendix
151
Appendix
Table A1. Equilibrium locations of Na
+
ions.
Ion index X Y Z
1 23.193 17.440 9.239
2 18.263 27.145 31.033
3 25.030 22.456 30.241
4 0.017 27.937 13.034
5 0.170 18.057 3.263
6 15.571 11.903 28.243
7 19.426 15.517 2.802
8 15.563 11.732 3.061
9 22.459 6.035 30.143
10 28.440 15.546 12.133
11 29.614 6.138 23.137
12 8.136 8.829 17.770
13 8.828 23.616 16.960
14 3.356 12.867 30.972
15 15.623 28.516 18.896
16 6.219 22.471 30.305
17 27.532 12.943 0.062
18 6.208 29.366 23.624
19 16.453 7.012 21.668
20 30.998 18.241 27.321
21 31.052 3.541 12.800
22 29.899 24.873 8.444
23 3.201 0.094 18.232
24 27.650 31.057 18.132
25 6.800 30.515 8.832
26 23.981 9.246 16.387
27 18.623 30.982 27.159
28 28.296 15.460 19.183
29 0.920 8.644 5.920
30 12.760 0.003 27.403
31 8.033 6.441 29.606
32 18.340 30.975 3.800
33 18.761 2.478 15.560
34 11.691 15.597 2.853
35 12.160 2.529 15.530
36 21.951 23.590 17.014
37 7.226 14.493 10.093
Appendix
152
38 9.129 16.618 23.813
39 16.615 7.063 9.619
40 15.491 19.145 28.366
41 15.538 28.308 11.547
42 29.458 24.708 23.021
43 24.734 1.642 7.863
44 13.803 21.990 7.479
45 2.843 19.263 15.497
46 12.848 31.022 3.394
47 21.593 16.508 23.932
48 2.407 12.314 15.673

















Appendix
153
Table A2. Equilibrium locations of K
+
ions.
Ion index X Y Z
1 21.638 14.857 7.271
2 18.565 27.698 31.050
3 24.001 23.876 2.209
4 3.852 31.041 12.147
5 27.906 18.462 31.026
6 19.072 15.554 27.988
7 15.540 18.515 2.885
8 15.520 11.897 3.137
9 24.821 8.753 0.250
10 22.319 0.069 24.766
11 15.538 3.055 18.973
12 17.138 22.732 23.045
13 0.027 15.592 7.852
14 8.509 22.745 17.714
15 3.841 12.237 0.009
16 15.377 9.398 24.309
17 6.887 15.177 21.931
18 3.439 18.686 0.007
19 30.979 3.804 12.158
20 8.745 0.296 6.367
21 3.548 30.965 18.751
22 28.689 28.633 12.402
23 22.429 22.681 13.135
24 21.599 6.760 15.171
25 0.349 8.575 5.853
26 12.847 28.283 31.043
27 8.537 0.389 24.887
28 18.538 3.335 0.062
29 15.644 3.247 11.817
30 12.445 28.182 15.516
31 18.665 28.206 15.580
32 3.111 18.883 15.555
33 9.360 7.230 14.814
34 9.394 15.049 6.948
35 12.232 15.455 27.997
36 22.552 0.500 6.620
37 7.215 24.101 28.829
38 2.066 23.995 7.137
39 14.887 23.970 9.238
40 27.888 11.989 15.541
41 12.291 3.648 0.039
Appendix
154
42 24.423 15.372 21.646
43 3.210 12.008 15.581
44 29.168 23.709 23.900
45 30.968 12.368 27.476
46 28.011 18.907 15.534
47 27.610 0.010 18.605
48 30.988 6.204 22.323



















Appendix
155
Table A3. Equilibrium locations of Rb
+
ions.
Ion index X Y Z
1 24.239 15.534 9.508
2 18.569 0.046 27.918
3 22.364 24.721 0.094
4 3.339 0.048 18.724
5 0.234 22.397 6.320
6 15.555 12.213 27.842
7 18.948 15.628 3.252
8 19.077 3.340 15.528
9 23.833 6.999 1.774
10 23.974 1.869 23.753
11 6.915 9.356 15.756
12 17.711 22.556 22.664
13 31.027 24.624 22.588
14 0.002 12.354 3.333
15 15.449 6.854 21.735
16 3.228 18.676 31.058
17 30.985 3.198 12.433
18 15.375 21.752 6.871
19 8.624 31.047 24.584
20 8.398 30.789 6.802
21 27.704 12.050 15.616
22 0.022 6.457 22.390
23 12.462 27.868 0.001
24 6.047 8.700 30.874
25 18.780 0.041 3.460
26 15.441 11.983 3.369
27 11.959 3.272 15.520
28 22.070 24.148 15.356
29 8.999 24.045 15.373
30 7.018 23.656 29.394
31 9.448 15.751 6.814
32 15.537 18.54 28.066
33 24.763 31.008 8.623
34 0.041 27.628 12.278
35 15.560 27.723 11.964
36 3.167 17.715 17.661
37 12.262 3.401 0.052
38 22.545 13.441 22.651
39 3.373 15.483 11.965
40 31.055 15.517 23.300
41 28.821 12.346 28.702
Appendix
156
42 27.746 18.999 15.583
43 27.928 0.032 18.562
44 9.391 15.516 24.136
45 15.583 6.742 9.383
46 15.445 28.151 18.377
47 27.600 18.871 0.041
48 1.867 7.228 7.288



















Appendix
157
Table A4. Equilibrium locations of Cs
+
ions.
Ion index X Y Z
1 22.746 15.561 7.146
2 18.899 27.914 31.050
3 24.650 22.543 0.007
4 3.106 0.042 18.798
5 22.686 30.936 6.253
6 18.766 15.575 27.654
7 15.510 18.945 3.563
8 15.512 13.334 2.411
9 24.535 8.524 30.928
10 31.036 6.556 22.576
11 15.582 3.411 18.962
12 15.525 22.406 23.901
13 0.073 24.575 22.494
14 31.058 12.069 3.264
15 6.946 15.548 21.890
16 0.053 18.938 27.807
17 3.229 31.049 12.214
18 8.484 31.019 6.304
19 8.444 30.997 24.600
20 0.033 24.909 8.559
21 27.543 12.093 15.414
22 12.085 27.732 31.047
23 6.402 8.494 0.070
24 18.995 3.308 31.040
25 21.941 7.173 15.568
26 12.683 27.920 15.525
27 23.940 22.267 15.526
28 8.740 15.545 7.251
29 9.223 7.039 15.593
30 31.035 6.526 8.473
31 12.383 15.618 27.680
32 22.579 30.926 24.697
33 15.605 24.051 9.085
34 3.420 17.811 13.252
35 12.114 3.206 31.048
36 3.430 12.275 15.538
37 23.927 15.621 21.978
38 31.042 12.117 27.816
39 28.715 17.686 15.554
40 27.825 31.058 18.942
41 15.639 8.866 7.168
Appendix
158
42 8.365 22.736 16.973
43 31.036 18.878 3.255
44 27.725 30.963 12.041
45 15.618 8.867 24.006
46 15.478 3.294 12.387
47 6.115 22.573 0.093
48 18.909 27.589 15.603



















Appendix
159
Table A5. Equilibrium locations of Mg
2+
ions.
Ion index X Y Z
1 19.241 15.574 2.786
2 18.069 27.787 31.024
3 31.026 27.329 18.332
4 15.508 2.755 19.404
5 1.881 24.480 7.809
6 15.490 11.841 28.244
7 11.506 15.523 2.892
8 15.474 2.898 11.547
9 22.648 5.758 1.055
10 31.030 12.709 27.087
11 2.587 12.172 15.534
12 7.078 16.639 21.404
13 28.581 18.671 15.52
14 16.997 21.736 23.575
15 6.315 23.075 29.520
16 0.082 12.703 3.938
17 19.113 28.342 15.547
18 23.935 9.609 16.679
19 27.365 18.378 30.998
20 2.847 2.774 12.694
21 14.113 23.597 9.370
22 6.233 1.385 22.923
23 27.216 31.053 12.767
24 13.044 3.200 0.041









Appendix
160
Table A6. Equilibrium locations of Ca
2+
ions.
Ion index X Y Z
1 21.297 16.223 6.999
2 18.585 28.481 2.532
3 27.800 19.464 15.480
4 27.293 31.032 12.365
5 0.613 25.269 22.498
6 15.549 11.754 27.881
7 12.014 15.485 3.101
8 15.496 3.184 11.703
9 25.297 8.673 0.640
10 0.080 3.684 18.656
11 12.850 3.165 18.290
12 9.786 24.000 14.826
13 3.206 15.514 11.604
14 3.184 15.515 19.495
15 3.169 12.682 31.039
16 12.222 2.896 2.935
17 15.575 19.389 27.945
18 27.677 18.432 31.045
19 3.717 0.023 12.411
20 19.013 27.977 15.519
21 12.405 31.019 27.330
22 5.852 22.664 0.759
23 27.866 11.705 15.531
24 22.396 0.571 25.083









Appendix
161
Table A7. Equilibrium locations of Al
3+
ions.









Ion index X Y Z
1 15.490 24.069 9.470
2 12.431 28.390 31.001
3 28.544 28.503 11.995
4 28.059 18.891 0.122
5 18.804 15.524 27.553
6 17.685 8.605 8.705
7 18.967 3.122 30.985
8 31.013 5.283 22.277
9 13.169 3.513 17.795
10 17.990 27.601 17.956
11 5.781 23.049 22.649
12 2.987 12.232 0.035
13 2.619 2.632 11.893
14 13.169 15.545 2.724
15 3.478 15.528 12.302
16 27.907 15.547 18.260
Publications
162
Publications
1. Y. F. Chen, Z. Q. Hu, K. M. Gupta. J. W. Jiang. Ionic Liquid/Metal-Organic
Framework Composite for CO
2
Capture: A Computational Investigation. Journal
of Physical Chemistry C. 2011, 115, 21736-21742.
2. Y. F. Chen, J. W. Jiang. A Bio-Metal-Organic Framework for Highly Selective
CO
2
Capture: A Molecular Simulation Study. ChemSusChem. 2010, 3, 982-988.
3. Y. F. Chen, J. Y. Lee, R. Babarao, J. Li, and J. W. Jiang. A Highly Hydrophobic
Metal-Organic Framework Zn(BDC)(TED)
0.5
for Adsorption and Separation of
CH
3
OH/H
2
O and CO
2
/CH
4
. Journal of Physical Chemistry C. 2010, 114, 6602 -
6609.
4. Y. F. Chen, R. Babarao, S. I. Sandler, and J. W. Jiang. Metal-Organic Framework
MIL-101 for Adsorption and Effect of Terminal Water Molecules: From Quantum
Mechanics to Molecular Simulation. Langmuir 2010, 26, 8743-8750.
5. Z. Q. Hu, Y. F. Chen, and J. W. Jiang. Zeolitic Imidazolate Framework-8 as a
Reverse Osmosis Membrane for Water Desalination: Insight from Molecular
Simulation. Journal of Chemical Physics. 2011, 134, 134705.
6. P. Pachfule, Y. F. Chen, J. W. Jiang, R. Banerjee. Fluorinated Metal Organic
Frameworks (F-MOFs): Advantageous for Higher H
2
and CO
2
Adsorption or not?
Chemistry-A European Journal. 2012, 18, 688-694.
7. P. Pachfule, Y. F. Chen, S. C. Sahoo, J. W. Jiang, R. Banerjee. Structural
Isomerism and effect of Fluorination of Gas Adsorption in Copper-Tetrazolate
Based Metal Organic Frameworks. Chemistry of Materials. 2011, 23, 2908-2916.
8. P. Pachfule, Y. F. Chen, J. W. Jiang, R. Banerjee. Experimental and Computational
Approach of Understanding the Gas Adsorption in Amino Functionalized
Interpenetrated Metal Organic Frameworks. Journal of Materials Chemistry. 2011,
21, 17737-17745.
Publications
163
9. T. Panda, P. Pachfule, Y. F. Chen, J. W. Jiang. And R. Banerjee. Amino
Functionalized Zeolitic Tetrazolate Framework (ZTF) with High Capacity for
Storage of Carbon Dioxide. Chemical Communications. 2011, 47, 2011-2013.

Presentations
1. Y. F. Chen, A. Nalaparaju, J. W. Jiang, Adsorption and Separation of CO
2
/H
2
in
Mono-, Di- and Trivalent Cation-Exchanged Zeolite-like Metal-Organic
Frameworks: Atomistic Simulation Study, 14
th
Asia Pacific Confederation of
Chemical Engineering Congress, Singapore (Feb. 2012).
2. R. Babarao, Y. F. Chen, J. W. Jiang, Effect of Water on Adsorption in
Metal-Organic Frameworks: Insight from Molecular Simulations, AIChE Annual
Conference, Salt Lake City, Utah, USA (Nov. 2010).
3. Y. F. Chen, J. Lee, J. Li, J. W. Jiang, Adsorption and Separation of CO
2
/CH
4
and
CH
3
OH/H
2
O in Highly Hydrophobic Zn(BDC)(TED)
0.5
. 2
nd
International
Conference on Metal-Organic Frameworks and Open Framework Compounds,
Marseille, France (Sep. 2010).
4. Y. F. Chen, R. Babarao, J. W. Jiang, Metal-Organic Framework MIL-101 for
Adsorption and Effect of Terminal Water Molecules: A Computational Study. 2
nd

International Conference on Metal-Organic Frameworks and Open Framework
Compounds, Marseille, France (Sep. 2010).

NATURE PUBLI SHI NG GROUP LI CENSE
TERMS AND CONDI TI ONS
May 10, 2012

This is a License Agreement between Yifei Chen ("You") and Nature
Publishing Group ("Nature Publishing Group") provided by Copyright
Clearance Center ("CCC"). The license consists of your order details, the terms
and conditions provided by Nature Publishing Group, and the payment terms
and conditions.
Al l payment s must be made i n f ul l t o CCC. For pay ment i nst r uct i ons, pl ease see
i nf or mat i on l i st ed at t he bot t om of t hi s f or m.
License Number 2905190533550
License dat e May 10, 2012
Licensed cont ent publisher Nat ure Publishing Group
Licensed cont ent publicat ion Nat ure
Licensed cont ent t it le Ret icular synt hesis and t he design of new mat erials
Licensed cont ent aut hor Omar M. Yaghi, Michael O' Keeffe, Nat han W. Ockwig, Hee K. Chae,
Mohamed Eddaoudi, Jaheon Kim
Licensed cont ent dat e Jun 12, 2003
Type of Use reuse in a t hesis/ dissert at ion
Request or t ype academic/ educat ional
Format print and elect ronic
Port ion figures/ t ables/ illust rat ions
Number of
figures/ t ables/ illust rat ions
1
High- res required no
Figures Figure 2 on page 708
Aut hor of t his NPG art icle no
Your reference number
Tit le of your t hesis /
dissert at ion
Molecular simulat ions for CO2 capt ure in met al- organic frameworks
Expect ed complet ion dat e May 2012
Est imat ed size ( number of
pages)
176
Tot al 0. 00 USD
Terms and Condit ions
Terms and Conditions for Permissions
Nature Publishing Group hereby grants you a non-exclusive license to
reproduce this material for this purpose, and for no other use, subject to the
conditions below:
Page 1 of 4 Rightslink Printable License
5/10/2012 https://s100.copyright.com/App/PrintableLicenseFrame.jsp?publisherID...
1. NPG warrants that it has, to the best of its knowledge, the rights to license
reuse of this material. However, you should ensure that the material you
are requesting is original to Nature Publishing Group and does not carry
the copyright of another entity (as credited in the published version). If
the credit line on any part of the material you have requested indicates
that it was reprinted or adapted by NPG with permission from another
source, then you should also seek permission from that source to reuse the
material.

2. Permission granted free of charge for material in print is also usually
granted for any electronic version of that work, provided that the material
is incidental to the work as a whole and that the electronic version is
essentially equivalent to, or substitutes for, the print version. Where print
permission has been granted for a fee, separate permission must be
obtained for any additional, electronic re-use (unless, as in the case of a
full paper, this has already been accounted for during your initial request
in the calculation of a print run). NB: In all cases, web-based use of full-
text articles must be authorized separately through the 'Use on a Web Site'
option when requesting permission.

3. Permission granted for a first edition does not apply to second and
subsequent editions and for editions in other languages (except for
signatories to the STM Permissions Guidelines, or where the first edition
permission was granted for free).

4. Nature Publishing Group's permission must be acknowledged next to the
figure, table or abstract in print. In electronic form, this acknowledgement
must be visible at the same time as the figure/table/abstract, and must be
hyperlinked to the journal's homepage.

5. The credit line should read:
Reprinted by permission from Macmillan Publishers Ltd: [JOURNAL
NAME] (reference citation), copyright (year of publication)
For AOP papers, the credit line should read:
Reprinted by permission from Macmillan Publishers Ltd: [JOURNAL
NAME], advance online publication, day month year (doi: 10.1038/sj.
[JOURNAL ACRONYM].XXXXX)

Note: For republication from the British Journal of Cancer, the
following credit lines apply.
Reprinted by permission from Macmillan Publishers Ltd on behalf of
Cancer Research UK: [JOURNAL NAME] (reference citation), copyright
(year of publication) For AOP papers, the credit line should read:
Reprinted by permission from Macmillan Publishers Ltd on behalf of
Cancer Research UK: [JOURNAL NAME], advance online publication,
day month year (doi: 10.1038/sj.[JOURNAL ACRONYM].XXXXX)
Page 2 of 4 Rightslink Printable License
5/10/2012 https://s100.copyright.com/App/PrintableLicenseFrame.jsp?publisherID...


6. Adaptations of single figures do not require NPG approval. However, the
adaptation should be credited as follows:

Adapted by permission from Macmillan Publishers Ltd: [JOURNAL
NAME] (reference citation), copyright (year of publication)

Note: For adaptation from the British Journal of Cancer, the
following credit line applies.
Adapted by permission from Macmillan Publishers Ltd on behalf of
Cancer Research UK: [JOURNAL NAME] (reference citation), copyright
(year of publication)

7. Translations of 401 words up to a whole article require NPG approval.
Please visit http://www.macmillanmedicalcommunications.com for more
information. Translations of up to a 400 words do not require NPG
approval. The translation should be credited as follows:

Translated by permission from Macmillan Publishers Ltd: [JOURNAL
NAME] (reference citation), copyright (year of publication).

Note: For translation from the British Journal of Cancer, the
following credit line applies.
Translated by permission from Macmillan Publishers Ltd on behalf of
Cancer Research UK: [JOURNAL NAME] (reference citation), copyright
(year of publication)
We are certain that all parties will benefit from this agreement and wish you
the best in the use of this material. Thank you.
Special Terms:
v1.1
I f y ou w oul d l i k e t o pay f or t hi s l i cense now , pl ease r emi t t hi s l i cense al ong w i t h y our
pay ment made pay abl e t o " COPYRI GHT CLEARANCE CENTER" ot her w i se y ou w i l l be
i nv oi ced w i t hi n 48 hour s of t he l i cense dat e. Pay ment shoul d be i n t he f or m of a check
or money or der r ef er enci ng y our account number and t hi s i nv oi ce number
RLNK500776570.
Once y ou r ecei ve y our i nv oi ce f or t hi s or der , y ou may pay y our i nv oi ce by cr edi t car d.
Pl ease f ol l ow i nst r uct i ons pr ov i ded at t hat t i me.

Mak e Pay ment To:
Copy r i ght Cl ear ance Cent er
Dept 001
P. O. Box 843006
Bost on, MA 02284- 3006

For suggest i ons or comment s r egar di ng t hi s or der , cont act Ri ght sLi nk Cust omer
Suppor t : cust omer car e@copy r i ght .com or + 1- 877- 622- 5543 ( t ol l f r ee i n t he US) or + 1-
978- 646- 2777.
Page 3 of 4 Rightslink Printable License
5/10/2012 https://s100.copyright.com/App/PrintableLicenseFrame.jsp?publisherID...
l i cense f or y our r ef er ence. No pay ment i s r equi r ed.
Page 4 of 4 Rightslink Printable License
5/10/2012 https://s100.copyright.com/App/PrintableLicenseFrame.jsp?publisherID...
THE AMERI CAN ASSOCI ATI ON FOR THE ADVANCEMENT OF SCI ENCE LI CENSE
TERMS AND CONDI TI ONS
May 10, 2012

This is a License Agreement between Yifei Chen ("You") and The American
Association for the Advancement of Science ("The American Association for
the Advancement of Science") provided by Copyright Clearance Center
("CCC"). The license consists of your order details, the terms and conditions
provided by The American Association for the Advancement of Science, and
the payment terms and conditions.
Al l payment s must be made i n f ul l t o CCC. For pay ment i nst r uct i ons, pl ease see
i nf or mat i on l i st ed at t he bot t om of t hi s f or m.
License Number 2905191314867
License dat e May 10, 2012
Licensed cont ent publisher The American Associat ion for t he Advancement of Science
Licensed cont ent publicat ion Science
Licensed cont ent t it le Syst emat ic Design of Pore Size and Funct ionalit y in I soret icular
MOFs and Their Applicat ion in Met hane St orage
Licensed cont ent aut hor Mohamed Eddaoudi, Jaheon Kim, Nat haniel Rosi, David Vodak,
Joseph Wacht er, Michael O' Keeffe, Omar M. Yaghi
Licensed cont ent dat e Jan 18, 2002
Volume number 295
I ssue number 5554
Type of Use Thesis / Dissert at ion
Request or t ype Scient ist / individual at a r esearch inst it ut ion
Format Print and elect ronic
Port ion Figure
Number of figures/ t ables 1
Order reference number
Tit le of your t hesis /
dissert at ion
Molecular simulat ions for CO2 capt ure in met al- organic frameworks
Expect ed complet ion dat e May 2012
Est imat ed size( pages) 176
Tot al 0. 00 USD
Terms and Condit ions
American Association for the Advancement of Science TERMS AND
CONDITIONS
Regarding your request, we are pleased to grant you non-exclusive, non-
transferable permission, to republish the AAAS material identified above in
your work identified above, subject to the terms and conditions herein. We
must be contacted for permission for any uses other than those specifically
Page 1 of 9 Rightslink Printable License
5/10/2012 https://s100.copyright.com/App/PrintableLicenseFrame.jsp?publisherID...
identified in your request above.
The following credit line must be printed along with the AAAS material:
"From [Full Reference Citation]. Reprinted with permission from AAAS."
All required credit lines and notices must be visible any time a user accesses
any part of the AAAS material and must appear on any printed copies and
authorized user might make.
This permission does not apply to figures / photos / artwork or any other
content or materials included in your work that are credited to non-AAAS
sources. If the requested material is sourced to or references non-AAAS
sources, you must obtain authorization from that source as well before using
that material. You agree to hold harmless and indemnify AAAS against any
claims arising from your use of any content in your work that is credited to
non-AAAS sources.
If the AAAS material covered by this permission was published in Science
during the years 1974 - 1994, you must also obtain permission from the author,
who may grant or withhold permission, and who may or may not charge a fee
if permission is granted. See original article for author's address. This condition
does not apply to news articles.
The AAAS material may not be modified or altered except that figures and
tables may be modified with permission from the author. Author permission for
any such changes must be secured prior to your use.
Whenever possible, we ask that electronic uses of the AAAS material
permitted herein include a hyperlink to the original work on AAAS's website
(hyperlink may be embedded in the reference citation).
AAAS material reproduced in your work identified herein must not account for
more than 30% of the total contents of that work.
AAAS must publish the full paper prior to use of any text.
AAAS material must not imply any endorsement by the American Association
for the Advancement of Science.
This permission is not valid for the use of the AAAS and/or Science logos.
AAAS makes no representations or warranties as to the accuracy of any
information contained in the AAAS material covered by this permission,
including any warranties of
merchantability or fitness for a particular purpose.
If permission fees for this use are waived, please note that AAAS reserves the
Page 2 of 9 Rightslink Printable License
5/10/2012 https://s100.copyright.com/App/PrintableLicenseFrame.jsp?publisherID...
right to charge for reproduction of this material in the future.
Permission is not valid unless payment is received within sixty (60) days of the
issuance of this permission. If payment is not received within this time period
then all rights granted herein shall be revoked and this permission will be
considered null and void.
In the event of breach of any of the terms and conditions herein or any of
CCC's Billing and Payment terms and conditions, all rights granted herein shall
be revoked and this permission will be considered null and void.
AAAS reserves the right to terminate this permission and all rights granted
herein at its discretion, for any purpose, at any time. In the event that AAAS
elects to terminate this permission, you will have no further right to publish,
publicly perform, publicly display, distribute or otherwise use any matter in
which the AAAS content had been included, and all fees paid hereunder shall
be fully refunded to you. Notification of termination will be sent to the contact
information as supplied by you during the request process and termination shall
be immediate upon sending the notice. Neither AAAS nor CCC shall be liable
for any costs, expenses, or damages you may incur as a result of the
termination of this permission, beyond the refund noted above.
This Permission may not be amended except by written document signed by
both parties.
The terms above are applicable to all permissions granted for the use of AAAS
material. Below you will find additional conditions that apply to your particular
type of use.
FOR A THESIS OR DISSERTATION
If you are using figure(s)/table(s), permission is granted for use in print and
electronic versions of your dissertation or thesis. A full text article may be used
in print versions only of a dissertation or thesis.
Permission covers the distribution of your dissertation or thesis on demand by
ProQuest / UMI, provided the AAAS material covered by this permission
remains in situ.
If you are an Original Author on the AAAS article being reproduced, please
refer to your License to Publish for rules on reproducing your paper in a
dissertation or thesis.
FOR JOURNALS:
Permission covers both print and electronic versions of your journal article,
however the AAAS material may not be used in any manner other than within
the context of your article.
Page 3 of 9 Rightslink Printable License
5/10/2012 https://s100.copyright.com/App/PrintableLicenseFrame.jsp?publisherID...
FOR BOOKS/TEXTBOOKS:
If this license is to reuse figures/tables, then permission is granted for non-
exclusive world rights in all languages in both print and electronic formats
(electronic formats are defined below).
If this license is to reuse a text excerpt or a full text article, then permission is
granted for non-exclusive world rights in English only. You have the option of
securing either print or electronic rights or both, but electronic rights are not
automatically granted and do garner additional fees. Permission for translations
of text excerpts or full text articles into other languages must be obtained
separately.
Licenses granted for use of AAAS material in electronic format
books/textbooks are valid only in cases where the electronic version is
equivalent to or substitutes for the print version of the book/textbook. The
AAAS material reproduced as permitted herein must remain in situ and must
not be exploited separately (for example, if permission covers the use of a full
text article, the article may not be offered for access or for purchase as a stand-
alone unit), except in the case of permitted textbook companions as noted
below.
You must include the following notice in any electronic versions, either
adjacent to the reprinted AAAS material or in the terms and conditions for use
of your electronic products: "Readers may view, browse, and/or download
material for temporary copying purposes only, provided these uses are for
noncommercial personal purposes. Except as provided by law, this material
may not be further reproduced, distributed, transmitted, modified, adapted,
performed, displayed, published, or sold in whole or in part, without prior
written permission from the publisher."
If your book is an academic textbook, permission covers the following
companions to your textbook, provided such companions are distributed only
in conjunction with your textbook at no additional cost to the user:

- Password-protected website
- Instructor's image CD/DVD and/or PowerPoint resource
- Student CD/DVD
All companions must contain instructions to users that the AAAS material may
be used for non-commercial, classroom purposes only. Any other uses require
the prior written permission from AAAS.
If your license is for the use of AAAS Figures/Tables, then the electronic rights
granted herein permit use of the Licensed Material in any Custom Databases
that you distribute the electronic versions of your textbook through, so long as
the Licensed Material remains within the context of a chapter of the title
Page 4 of 9 Rightslink Printable License
5/10/2012 https://s100.copyright.com/App/PrintableLicenseFrame.jsp?publisherID...
identified in your request and cannot be downloaded by a user as an
independent image file.
Rights also extend to copies/files of your Work (as described above) that you
are required to provide for use by the visually and/or print disabled in
compliance with state and federal laws.
This permission only covers a single edition of your work as identified in your
request.
FOR NEWSLETTERS:
Permission covers print and/or electronic versions, provided the AAAS
material reproduced as permitted herein remains in situ and is not exploited
separately (for example, if permission covers the use of a full text article, the
article may not be offered for access or for purchase as a stand-alone unit)
FOR ANNUAL REPORTS:
Permission covers print and electronic versions provided the AAAS material
reproduced as permitted herein remains in situ and is not exploited separately
(for example, if permission covers the use of a full text article, the article may
not be offered for access or for purchase as a stand-alone unit)
FOR PROMOTIONAL/MARKETING USES:
Permission covers the use of AAAS material in promotional or marketing
pieces such as information packets, media kits, product slide kits, brochures, or
flyers limited to a single print run. The AAAS Material may not be used in any
manner which implies endorsement or promotion by the American Association
for the Advancement of Science (AAAS) or Science of any product or service.
AAAS does not permit the reproduction of its name, logo or text on
promotional literature.
If permission to use a full text article is permitted, The Science article covered
by this permission must not be altered in any way. No additional printing may
be set onto an article copy other than the copyright credit line required above.
Any alterations must be approved in advance and in writing by AAAS. This
includes, but is not limited to, the placement of sponsorship identifiers,
trademarks, logos, rubber stamping or self-adhesive stickers onto the article
copies.
Additionally, article copies must be a freestanding part of any information
package (i.e. media kit) into which they are inserted. They may not be
physically attached to anything, such as an advertising insert, or have anything
attached to them, such as a sample product. Article copies must be easily
removable from any kits or informational packages in which they are used. The
only exception is that article copies may be inserted into three-ring binders.
FOR CORPORATE INTERNAL USE:
Page 5 of 9 Rightslink Printable License
5/10/2012 https://s100.copyright.com/App/PrintableLicenseFrame.jsp?publisherID...
The AAAS material covered by this permission may not be altered in any way.
No additional printing may be set onto an article copy other than the required
credit line. Any alterations must be approved in advance and in writing by
AAAS. This includes, but is not limited to the placement of sponsorship
identifiers, trademarks, logos, rubber stamping or self-adhesive stickers onto
article copies.
If you are making article copies, copies are restricted to the number indicated
in your request and must be distributed only to internal employees for internal
use.
If you are using AAAS Material in Presentation Slides, the required credit line
must be visible on the slide where the AAAS material will be reprinted
If you are using AAAS Material on a CD, DVD, Flash Drive, or the World
Wide Web, you must include the following notice in any electronic versions,
either adjacent to the reprinted AAAS material or in the terms and conditions
for use of your electronic products: "Readers may view, browse, and/or
download material for temporary copying purposes only, provided these uses
are for noncommercial personal purposes. Except as provided by law, this
material may not be further reproduced, distributed, transmitted, modified,
adapted, performed, displayed, published, or sold in whole or in part, without
prior written permission from the publisher." Access to any such CD, DVD,
Flash Drive or Web page must be restricted to your organization's employees
only.
FOR CME COURSE and SCIENTIFIC SOCIETY MEETINGS:
Permission is restricted to the particular Course, Seminar, Conference, or
Meeting indicated in your request. If this license covers a text excerpt or a Full
Text Article, access to the reprinted AAAS material must be restricted to
attendees of your event only (if you have been granted electronic rights for use
of a full text article on your website, your website must be password protected,
or access restricted so that only attendees can access the content on your site).
If you are using AAAS Material on a CD, DVD, Flash Drive, or the World
Wide Web, you must include the following notice in any electronic versions,
either adjacent to the reprinted AAAS material or in the terms and conditions
for use of your electronic products: "Readers may view, browse, and/or
download material for temporary copying purposes only, provided these uses
are for noncommercial personal purposes. Except as provided by law, this
material may not be further reproduced, distributed, transmitted, modified,
adapted, performed, displayed, published, or sold in whole or in part, without
prior written permission from the publisher."
FOR POLICY REPORTS:
These rights are granted only to non-profit organizations and/or government
agencies. Permission covers print and electronic versions of a report, provided
Page 6 of 9 Rightslink Printable License
5/10/2012 https://s100.copyright.com/App/PrintableLicenseFrame.jsp?publisherID...
the required credit line appears in both versions and provided the AAAS
material reproduced as permitted herein remains in situ and is not exploited
separately.
FOR CLASSROOM PHOTOCOPIES:
Permission covers distribution in print copy format only. Article copies must
be freestanding and not part of a course pack. They may not be physically
attached to anything or have anything attached to them.
FOR COURSEPACKS OR COURSE WEBSITES:
These rights cover use of the AAAS material in one class at one institution.
Permission is valid only for a single semester after which the AAAS material
must be removed from the Electronic Course website, unless new permission is
obtained for an additional semester. If the material is to be distributed online,
access must be restricted to students and instructors enrolled in that particular
course by some means of password or access control.
FOR WEBSITES:
You must include the following notice in any electronic versions, either
adjacent to the reprinted AAAS material or in the terms and conditions for use
of your electronic products: "Readers may view, browse, and/or download
material for temporary copying purposes only, provided these uses are for
noncommercial personal purposes. Except as provided by law, this material
may not be further reproduced, distributed, transmitted, modified, adapted,
performed, displayed, published, or sold in whole or in part, without prior
written permission from the publisher."
Permissions for the use of Full Text articles on third party websites are granted
on a case by case basis and only in cases where access to the AAAS Material is
restricted by some means of password or access control. Alternately, an E-Print
may be purchased through our reprints department
(brocheleau@rockwaterinc.com).
REGARDING FULL TEXT ARTICLE USE ON THE WORLD WIDE WEB
IF YOU ARE AN ORIGINAL AUTHOR OF A SCIENCE PAPER
If you chose "Original Author" as the Requestor Type, you are warranting that
you are one of authors listed on the License Agreement as a "Licensed content
author" or that you are acting on that author's behalf to use the Licensed
content in a new work that one of the authors listed on the License Agreement
as a "Licensed content author" has written.
Original Authors may post the Accepted Version of their full text article on
their personal or on their University website and not on any other website. The
Accepted Version is the version of the paper accepted for publication by
AAAS including changes resulting from peer review but prior to AAASs copy
editing and production (in other words not the AAAS published version).
Page 7 of 9 Rightslink Printable License
5/10/2012 https://s100.copyright.com/App/PrintableLicenseFrame.jsp?publisherID...
FOR MOVIES / FILM / TELEVISION:
Permission is granted to use, record, film, photograph, and/or tape the AAAS
material in connection with your program/film and in any medium your
program/film may be shown or heard, including but not limited to broadcast
and cable television, radio, print, world wide web, and videocassette.
The required credit line should run in the program/film's end credits.
FOR MUSEUM EXHIBITIONS:
Permission is granted to use the AAAS material as part of a single exhibition
for the duration of that exhibit. Permission for use of the material in
promotional materials for the exhibit must be cleared separately with AAAS
(please contact us at permissions@aaas.org).
FOR TRANSLATIONS:
Translation rights apply only to the language identified in your request
summary above.
The following disclaimer must appear with your translation, on the first page of
the article, after the credit line: "This translation is not an official translation by
AAAS staff, nor is it endorsed by AAAS as accurate. In crucial matters, please
refer to the official English-language version originally published by AAAS."
FOR USE ON A COVER:
Permission is granted to use the AAAS material on the cover of a journal issue,
newsletter issue, book, textbook, or annual report in print and electronic
formats provided the AAAS material reproduced as permitted herein remains
in situ and is not exploited separately
By using the AAAS Material identified in your request, you agree to abide by
all the terms and conditions herein.
Questions about these terms can be directed to the AAAS Permissions
department permissions@aaas.org.
Other Terms and Conditions:
v 2
I f y ou w oul d l i k e t o pay f or t hi s l i cense now , pl ease r emi t t hi s l i cense al ong w i t h y our
pay ment made pay abl e t o " COPYRI GHT CLEARANCE CENTER" ot her w i se y ou w i l l be
i nv oi ced w i t hi n 48 hour s of t he l i cense dat e. Pay ment shoul d be i n t he f or m of a check
or money or der r ef er enci ng y our account number and t hi s i nv oi ce number
RLNK500776578.
Once y ou r ecei ve y our i nv oi ce f or t hi s or der , y ou may pay y our i nv oi ce by cr edi t car d.
Pl ease f ol l ow i nst r uct i ons pr ov i ded at t hat t i me.

Mak e Pay ment To:
Copy r i ght Cl ear ance Cent er
Dept 001
P. O. Box 843006
Page 8 of 9 Rightslink Printable License
5/10/2012 https://s100.copyright.com/App/PrintableLicenseFrame.jsp?publisherID...
Bost on, MA 02284- 3006

For suggest i ons or comment s r egar di ng t hi s or der , cont act Ri ght sLi nk Cust omer
Suppor t : cust omer car e@copy r i ght .com or + 1- 877- 622- 5543 ( t ol l f r ee i n t he US) or + 1-
978- 646- 2777.
Gr at i s l i censes ( r ef er enci ng $0 i n t he Tot al f i el d) ar e f r ee. Pl ease r et ai n t hi s pr i nt abl e
l i cense f or y our r ef er ence. No pay ment i s r equi r ed.
Page 9 of 9 Rightslink Printable License
5/10/2012 https://s100.copyright.com/App/PrintableLicenseFrame.jsp?publisherID...

You might also like