You are on page 1of 9

Tetrahedron Letters xxx (2016) xxxxxx

Contents lists available at ScienceDirect

Tetrahedron Letters
journal homepage: www.elsevier.com/locate/tetlet

Digest paper

Recent developments in porous materials for H2 and CH4 storage


Jia Liu a,b, Ruqiang Zou b,, Yanli Zhao a,c,
a

Division of Chemistry and Biological Chemistry, School of Physical and Mathematical Sciences, Nanyang Technological University, 21 Nanyang Link, Singapore 637371, Singapore
Beijing Key Laboratory for Theory and Technology of Advanced Battery Materials, Department of Materials Science and Engineering, College of Engineering, Peking University, Beijing
100871, China
c
School of Materials Science and Engineering, Nanyang Technological University, 50 Nanyang Avenue, Singapore 639798, Singapore
b

a r t i c l e

i n f o

a b s t r a c t

Article history:
Received 10 August 2016
Revised 15 September 2016
Accepted 26 September 2016
Available online xxxx

Emerging classes of porous crystalline materials such as metalorganic frameworks and covalent-organic
frameworks have received significant attention for selective gas storage. However, lack of coincident
standard and capacity calculation methods brings challenges to compare with existing materials.
Herein, we briefly discussed the H2 and CH4 storage capacity of some representative porous materials
at high pressure and proposed the conception of adsorptive density of H2 and CH4 that can be used as
a standard to evaluate the average intensity of the potential field in the pores. Important physical properties of these porous materials such as surface area, pore volume, and crystal density were illustrated for
evaluating the gas storage capacity. High pressure isotherm data were used to calculate the gravimetric
and volumetric uptake. Other important factors such as mechanical property, packing density, and impurity were also considered during the discussions. Promising potential of these porous materials for
improving gas storage capacity was highlighted.
2016 Elsevier Ltd. All rights reserved.

Keywords:
Covalent-organic frameworks
Gas storage
Metalorganic frameworks
Physical adsorption
Porous materials

Contents
Introduction. . . . . . . . . .
H2 storage . . . . . . . . . . .
Methane storage . . . . . .
Summary . . . . . . . . . . . .
Acknowledgments . . .
References and notes .

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

Introduction
A lot of gases are important species in our daily life. They normally exist as gas mixtures such as air and natural gas. Some useful
gases including H2, O2, and CH4 are highly flammable and volatile,
while high concentrations of gases like CO, CO2, and NOx are harmful to the environment and human being. Therefore, developing
approaches and technologies to selectively adsorb and store gases
is highly required. Using porous materials to selectively adsorb certain gases is one of the strategies developed so far. Among these
porous materials, metalorganic frameworks (MOFs) and covalent-organic frameworks (COFs) have attracted increasing atten Corresponding authors.
E-mail addresses: rzou@pku.edu.cn (R. Zou), zhaoyanli@ntu.edu.sg (Y. Zhao).

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

00
00
00
00
00
00

tion for their unique capabilities of selective gas capture and


storage. On the other hand, coincident standard and capacity calculation methods employed to compare their capabilities with existing materials have not been well standardized. In this Digest, we
highlighted the H2 and CH4 storage capacity of some representative
MOF and COF materials. The conception of adsorptive density of H2
and CH4 was proposed in order to serve as a standard to evaluate
the average intensity of the potential field in the pores of these
gas adsorbents.
H2 storage
We are now facing two critical problems of energy shortage and
air pollution because of large demand and over-exploitation of fossil fuel. Pursuing substitutable clean energy has become an urgent

http://dx.doi.org/10.1016/j.tetlet.2016.09.085
0040-4039/ 2016 Elsevier Ltd. All rights reserved.

Please cite this article in press as: Liu, J.; et al. Tetrahedron Lett. (2016), http://dx.doi.org/10.1016/j.tetlet.2016.09.085

J. Liu et al. / Tetrahedron Letters xxx (2016) xxxxxx

task. H2 with a heat value three times higher than oil and completely free of pollution during use is considered one of most
promising alternative energy resources. One of the bottlenecks in
H2 applications is its onboard storage. H2 can only be liquidized
by compressing it below its critical temperature (33.2 K), which
is a great challenge and results in inconceivable energy consumption. H2 liquefaction is not possible for commercialization at this
stage. While compressing H2 could be easily carried out, the storage capacity is as low as 2 wt % even at 20 MPa. Even by using purpose made cylinder with a weight of 110 kg, its H2 storage capacity
just reaches 6% (density of 30 kg m3) at 70 MPa, which is less than
half density of liquid H2 at 20 K.1
The U.S. Department of Energy published a revised onboard H2
storage target (5.5 wt %/40 g L1 by 2017; 7.5 wt %/70 g L1 ultimate) for vehicles.2 Aiming at this target, some new H2 storage
techniques have been developed, including solid H2 storage,3 fluid
H2 storage,4 and adsorptive storage.5 Among these approaches, the
storage of H2 using porous materials is one of the promising techniques. Various materials with high porosity and surface area were
purposely synthesized and extensively used for H2 storage based
on physical adsorptive interaction between H2 molecule and
exposed atoms on the surface of the materials.68 H2 is chemically
stable at room temperature due to its high HH bond energy
(436 kJ/mol). Its electron energy level of Sigma orbital is very low
(10.36 eV), and the gap of HOMOLUMO is as large as about
12.6 eV.9 Therefore, the chemisorption of H2 is hard to occur under
mild conditions. Only at high temperature, electrons of H2 can
jump above Fermi energy level and then react with some atoms
such as Ni for chemical adsorption. However, these chemically
adsorbed H2 cannot be released at room temperature. Generally,
chemical adsorption requires high temperature for both adsorption
and desorption. The chemisorption of H2 is not a discussion focus
in this Digest. With regard to physical adsorption, on one hand,
H2 molecule has only two electrons tightly focused on the core,
leading to a small instantaneous dipole and thus a very weak dispersion force. On the other hand, the Coulombic interaction
depended on the third item of Taylor expansion for the point
charge system is also very weak, because of a weak quadrupole
moment of H2 (2.12 e-40 cm2).10 The weak physical adsorptive
interaction between H2 and adsorbents triggers an obviously
experimental result that the H2 gravimetric uptake has an approximately linear relation with the surface area (SBET) of the adsorbents (Fig. 1).37,1034
Zhou has proposed a single layer adsorption mechanism for gas
adsorption above its critical temperature.1,35 Gases such as H2 with

Figure 1. Relationship of H2 uptake with the surface area (SBET) of some adsorbents.

Figure 2. Potential intensity and H2 density in the slit pore with different sizes.
Reproduced with permission from Ref. 36.

a considerable low critical temperature can only allow single


molecular layer adsorption above the critical temperature. When
the surfaces of adsorbents are entirely covered by gas molecules,
the gas molecules cannot further adsorb on the adsorbed gas to
form a second adsorption layer, and thus the excess adsorptive
amount reaches the maximum. The second layer adsorption can
never happen, because the potential fields to trap gas molecules
above the first adsorption layer are from the adsorbed gas molecules instead of the surface of adsorbents, and the kinetic energy
of the gas molecule is always larger than its potential energy above
its critical temperature. From this point of view, it is easy to predict
the H2 adsorption capacity of a material based on its physical property. For example, commercially activated carbon AX-21 with high
surface area of 2800 m2/g and pore volume of 1.6 mL/g (micropore
of 1 mL/g, mesopore of 0.6 mL/g) exhibited H2 gravimetric capacity
as high as 5.2% at 77 K near 35 bar. Meshes (0 0 1) of graphite
exposed in the slit pores of AX-21 supply a lot of adsorptive sites
for H2 molecules. The potential field produced by carbon atoms
in micropores could overlap to enhance the H2 adsorption capacity.
As shown in Figure 2, when the pore size is smaller than 8 , the
divided potential curve comes to focus in the center, and the
strengthening of the intensity leads to an increase of H2 density.36
The isosteric heat of H2 adsorption on AX-21 is within 6.36
7.2 kJ/mol. AX-21 performs outstandingly on gas storage among
all carbon-based materials due to not only its high surface area
but also its microspores. Another carbon material MSC-30 with a
similar surface area (2680 m2/g) and higher mesoporous ratio than
AX-21 exhibited H2 uptake of 4.8% and isosteric heat of about 4 kJ/mol. In mesopores of MSC-30, the potential intensity is weak due
to lack of overlap interaction from the neighboring carbon walls.
According to single layer adsorption mechanism, the adsorptive
phase volume can be calculated from the surface area and H2
kinetic diameter (2.89 ). Calculating the H2 density in adsorptive
phase (qH2), it is obvious that the value (64 g/L) of AX-21 is larger
than that of MSC-30 (61 g/L) and MSC-3000 (57 g/L),37 which
agrees with their H2 isosteric heat order. A series of activated carbon materials (NAC-1.5-n) with different pore sizes and surface
areas showed the similar order of qH2. The H2 density in adsorptive
phase directly reflects potential intensity that determines the gas
uptake. The theoretical H2 total volumetric uptake can be calculated by excess gravimetric uptake, pore volume, skeleton density
(obtained when the volume is measured excluding pores and void

Please cite this article in press as: Liu, J.; et al. Tetrahedron Lett. (2016), http://dx.doi.org/10.1016/j.tetlet.2016.09.085

J. Liu et al. / Tetrahedron Letters xxx (2016) xxxxxx

Figure 3. Structure and organic component of MOF-5, MOF-74, and HKUST-1. Reproduced from Ref. 38.

space), and crystal density (obtained when the volume is measured


including pores). For example, when the skeleton density is 2.2 g/
cm3 and the pore volume is 1.6 mL/g, the crystal density of AX21 is 0.49 g/cm3. Converting excess H2 gravimetric uptake of 5.2%
to total volumetric uptake, the value is 32.5 g/L at 77 K near 35
bar. However, the real packing density (obtained by filling a container with the sample) cannot reach the crystal density without
an effect on the porous structure. When the packing density of
AX-21 reaches 0.7 g/cm3, the H2 uptake is reduced to 3.8% that
resulted from the loss of micropore and surface area. It is impossible to exceed the maximum theoretical volumetric uptake by
increasing the packing density over the crystal density without
changing the pore structure of the adsorbents.
MOFs having high surface areas structured by metal coordination with organic ligands are emerging materials for gas storage
and separation. Some of them have surface areas exceeding
6000 m2/g. MOF-5 (Fig. 3) derived from linking basic zinc acetate
unit, Zn4O(CO2)6, with linear ditopic carboxylates has38 a surface
area of 3534 m2/g, and exhibits H2 uptake of 5% at 77 K near 35
bar, which is comparable to AX-21.
The theoretical H2 total volumetric uptake of MOF-5 was calculated to 39 g/L. Omitting skeleton atom occupation, the H2 density
in the pores can be calculated to evaluate the porous utilization
efficiency. The H2 density in the pore of MOF-5 (41.87 g/L) is less
than its qH2 (53 g/L), indicating that part of pore space is useless
for H2 storage. Comparing with carbon materials such as AX-21,
MOF-5 shows higher volumetric H2 uptake even though having
lower qH2 due to its high exposed atom ratio. FA-ZTC-1, a zeolite
templated carbon material having high surface area of 3300 m2/g
and pore volume of 1.66 mL/g, presents the H2 volumetric uptake
of 43.89 g/L and the qH2 of 76 g/L, both of which are higher than
MOF-5. The reason is that the potential intensity constructed by
graphite in micropores is higher than that constructed by atoms
in MOF-5. In MOF-5, Zn atom is saturated with oxygen and loses
its interaction with H2. The main interactions between H2 and
MOF-5 come from the aromatic ring and OZn4O. While the aromatic
rings of the organic ligand are saturated with hydrogen, resulting
in lower density of p electrons than that produced by the graphite
(0 0 1) surface. In addition, the pore size of MOF-5 is too large for
potential overlap to enhance H2 density in the pores.
In another example, Mg-MOF-74 (Fig. 3) constructed from infinite 31 (or 32) helical rods of Mg3[(O)3(CO2)3] unit shows the H2
uptake of just 2.26 wt %, corresponding to a H2 volumetric uptake
of 37 g/L. But, its qH2 reaches about 82 g/L, exceeding MOF-5 and

even liquefied H2 at 20 K near 1 atm. It was found that the isosteric


heat of H2 adsorption on Mg-MOF-74 (10 kJ/mol)38 is higher than
that on MOF-5 (4.8 kJ/mol)8 at low H2 coverage. This outcome is
because the unsaturated coordinative metal of Mg-MOF-74 has a
strong interaction with H2, while the Zn atom in MOF-5 is saturated with oxygen and loses the affinity with H2. The strong potential field in Mg-MOF-74 may be attributed to one-dimensional
neighboring metaloxygen cluster that would benefit the H2 storage. However, the large occupancy of the skeleton atom and the
presence of useless space in Mg-MOF-74 result in relative lower
H2 volumetric uptake than MOF-5. The ratio of exposed atoms in
Mg-MOF-74 is less than that in MOF-5, and the pore size is larger
than MOF-5, finally leading to lower H2 uptake of Mg-MOF-74 both
in gravimetric and volumetric values than MOF-5. To regulate the
ratio of exposed atoms in Mg-MOF-74, an expanded Mg-MOF-74
was synthesized using 3,30 -dihydroxy[1,10 -biphenyl]-4,40 -dicarboxylic acid as the organic ligand, which showed a higher surface
area of 2740 m2/g and doubled H2 uptake of 4.5 wt % near 35 bar
with the same initial isosteric heat of 10 kJ/mol. The H2 total volumetric uptake and qH2 were calculated to be 32 g/L and 56 g/L
respectively, indicating that the average potential intensity in the
pores of expanded Mg-MOF-74 becomes weak as compared with
normal Mg-MOF-74 on account of introducing more aromatic
rings. With regard to the isosteric heat of these MOF materials, it
was found that, upon increasing the H2 coverage, the isosteric
heats of both MOF materials reduced dramatically from a high
value corresponding to the interaction between metal sites and
H2 to a range of 14 kJ/mol due to the interaction between organic
ligands and H2. Furthermore, the heat at high H2 coverage for these
MOFs is lower than that on porous carbon materials, because the
sp2 carbons of organic ligands in the MOFs are saturated with H
atoms comparing to the porous carbon.
Physical property and H2 uptake of various adsorbents are listed
in Table 1.3952 High BrunauerEmmettTeller (BET) surface areas
of 4530 m2/g and 6420 m2/g for MOF-200 and MOF-210 (Fig. 4)
were determined respectively, which were constructed from the
same cluster of Zn4O(CO2)6 as MOF-5 and BBC (4,40 400 -(benzene1,3,5-triyl-tris(benzene-4,1-diyl))tribenzoate) for MOF-200 and
mixed BTE/BPDC (4,40 ,400 -(benzene-1,3,5,triyl-tris(ethyne-2,1diyl))tribenzoate/biphenyl-4,4-dicarboxylate) for MOF-210. They
presented 14% and 15% H2 uptake respectively, both of which were
higher than that of MOF-177 (Fig. 4) constructed from Zn4O(CO2)6
and BTB (4,40 400 -benzene-1,3,5-triyl-tribenzoate). However, the
volumetric H2 uptake of MOF-200 (36 g/L) and MOF-210 (44 g/L)

Please cite this article in press as: Liu, J.; et al. Tetrahedron Lett. (2016), http://dx.doi.org/10.1016/j.tetlet.2016.09.085

J. Liu et al. / Tetrahedron Letters xxx (2016) xxxxxx

Table 1
Physical property and gas uptake of various adsorbents near 35 bar at 77 K for H2 and 298 K for CH4
Name

Pure gas
AX-21
Maxsorb-3000
LMA-738
Carbon (NAC-1.5-550)
Carbon (NAC-1.5-600)
Carbon (FA-ZTC-1)
MOF-5
MOF-74 (Mg)
MOF-74 (Ni)
Expanded MOF-74 (Mg)
DUT-6
DUT-23 (Co)
DUT-49
DUT-9
HKUST-1
HKUST-1
HKUST-1
IRMOF-6
MIL-101-na (Cr)
mil-101(Cr)
MOF-210
MOF-177
MOF-200
MOF-519
MOF-520
NOTT-101
NU-100
PAF-1
PCN-11
PCN-14
PCN-61
PCN-66
PCN-68
UIO-66
UIO-67
UTSA-76
Zn2(bdc)2(dobco)
COF-1
COF-10
COF-102
COF-103
COF-301@PdCl2
COF-5
COF-6
COF-8

SBET
m2/g

2800
3203
3290
526
1317
3300
3300
1332
1027
2740
4414
4850
5400
5400
1900
1550
1502
2630
3281
2693
5800
4500
6200
2400
3200
2850
6143
5600
1931
1753
3000
4000
5000
970
1877
2282
1450
735
844
3620
3530
1083
1670
825
958

Vp
cm3/g

1.60
1.63
1.43
0.26
0.64
1.66
1.04
0.61
0.44
1.04
2.13
2.03
2.91
2.18
0.65
0.71
0.76
0.92
1.85
1.30
3.60
1.89
3.59
0.94
1.28
1.08
2.82
2.94
0.91
0.87
1.36
1.63
2.13
0.36
0.95
1.09
0.68
0.30
1.44
1.55
1.54
0.42
1.07
0.32
0.69

qcystal

cm3/g

0.48
0.48
0.53
1.40
0.91
0.47
0.59
0.91
1.20
0.59
0.37
0.40
0.30
0.36
0.88
0.88
0.88
0.65
0.42
0.44
0.25
0.43
0.22
0.94
0.58
0.69
0.29
0.32
0.75
0.83
0.56
0.45
0.38
1.32
0.72
0.70
0.82
1.39
0.48
0.43
0.43
1.30
0.58
1.10
1.32

CH4

H2
Uptake
wt %

Uptake
g/L

qH2

Uptake
wt %

V/V (STP)

g/L

5.20
5.30

1.47
2.96
7.30
5.10
2.26

4.50
6.62
7.21
8.00
5.53
3.60

4.70
4.20
5.60
7.50
6.80
8.10

9.95
6.80

6.24
6.65
7.32
2.40
4.60

1.48
3.92
7.24
7.05
4.20
3.30
2.26
3.50

9.8
32.5
33.1

24.1
32.8
42.1
37.2
24.8

32.5
32.2
33.7
32.8
26.0
37.3

37.8
25.3
30.5
27.6
36.9
25.6

36.9
30.5

42.5
36.7
36.0
36.1
39.5

24.7
25.6
37.7
36.8
60.0
25.2
28.3
55.1

70.8
64.0
57.0

96.4
77.5
76.3
53.3
82.0

56.6
51.7
51.3
51.1
57.9
65.3

57.9
44.1
66.9
44.6
52.1
45.1

55.9
41.9

71.7
57.3
50.5
85.4
84.5

68.0
76.8
69.0
68.9
133.7
68.1
103.9
89.4

17.5

19.1

16.2
11.8
11.3

17.0
24.0
15.0

8.5
14.9
15.6

6.0
10.7
11.1

11.4
16.1
20.2

16.7
17.9
17.6
17.6

6.0

21.6
12.2
1.6
3.9
17.5
16.5

2.2
3.8

33.0
143.7

167.1

154.7
168.9
206.2

123.2
131.4
101.8

126.0
206.1
162.1

55.8
68.2
93.7

180.5
155.8
219.9

198.2
232.5
163.9
134.4

126.2

237.3
159.0
45.3
49.8
128.1
121.9

45.9
100.3

was lower than that of MOF-177. Except there is too much useless
space in MOF-200 and MOF-210, the other reason is that there are
too many organic ligands contributing to low potential intensity
for the H2 storage.
The H2 storage in MOFs consists of strong binding with metal
sites and weak binding with organic ligands. Based on single layer
adsorption mechanism, we can understand why the MOFs did not
present overwhelmingly exceptive H2 uptake. Since H2 can only be
adsorbed in one layer, the H2 density in adsorptive phase can obviously reflect the capacity of adsorbent materials. The microspore is
beneficial to form a strong intensity potential field. However, if the
space is divided into many small pores, it would reduce the gravimetric uptake since many atoms fulfill the space. Among these
materials with high qH2, only porous carbon and PCN-61 with
ultrahigh surface areas show an uptake over 6%. The frameworks
with large pores formed by more aromatic ligands have lower
qH2. For PAF-1 (porous aromatic framework), its qH2 is as low as
40 g/L. Thus, the tradeoff between the pore size and surface area
is significant for designing porous materials toward enhanced H2
uptake (Table 1).

Ref.

qCH4

g/cm3
0.42
0.16

0.15

0.13
0.23
0.29

0.09
0.12
0.07

0.14
0.26
0.16

0.06
0.05
0.06

0.13
0.13
0.19

0.23
0.27
0.15
0.12

0.16

0.25
0.22
0.06
0.12
0.13
0.12

0.07
0.10

11,39
39,40
39
12
12
17
13,41
13,42
38
14,16
26
25
31
23
30
43
44
13,45
29
44
21
21
21
46
46
27,47
20
18
48
49
33,34
33
33
19,50
19,32
51
52
22
22
22
22
28
24
22
14

PCN-61 consisting of dimetal paddlewheel clusters and BTEI


(5,50 ,500 -benzene-1,3,5-triyltris(1-ethynyl-2-isophthalate)) with a
special carbyne group shows the highest volumetric H2 uptake
(42 g/L). Unlike HKUST-1 (Fig. 3), H2 is not only adsorbed on open
copper sites of PCN-61, but also has a strong interaction with high
density p electrons around the carbyne group. High qH2 value
(71 g/L) supports this conclusion. When using enlarged ligands to
obtain PCN-66 and PCN-68, they perform worse than PCN-61 on
the H2 storage due to their large pore size, even though both
PCN-66 and PCN-68 have the similar carbyne group (Fig. 5).
It is interesting that UIO-66 and UIO-67 present a similar qH2 as
high as 85 g/L (Fig. 6 and Table 1), meaning that the potential
intensity in UIO-67 constructed by an enlarged ligand than the
ligand used in UIO-66 is not weakened. The most probable reason
is that the H2 adsorptive sites on UIO-66 and UIO-67 do not focus
in the center of octahedron cage, but focus on the tetrahedron cage.
The dehydrated ZrO cluster in UIO-66 and UIO-67 has a strong
interaction with H2. In UIO-66, the tetrahedron cage is too small
to allow more H2 adsorbed, thus reducing the H2 uptake. When
the tetrahedron cage is enlarged in UIO-67, more H2 is allowed

Please cite this article in press as: Liu, J.; et al. Tetrahedron Lett. (2016), http://dx.doi.org/10.1016/j.tetlet.2016.09.085

J. Liu et al. / Tetrahedron Letters xxx (2016) xxxxxx

MOF-177

MOF-200

MOF-210

Figure 4. Structure and organic component of MOF-177, MOF-200, and MOF-210. Reproduced with permission from Ref. 21.

Figure 5. Composition as well as gravimetric uptake (a) and volumetric uptake (b) of PCN-61, PCN-66, and PCN-68 (NOTT-116). Reproduced with permission from Ref. 33.

in the cage to enhance the uptake. The same isosteric heat and qH2
on both materials support these conclusions.
Thus, it is not always a good idea to enlarge surface area and
empty pore volume by simply introducing large organic ligands

in order to enhance the H2 storage. Mesoporous MOFs such as


MIL-101 and DUT-n all present low volumetric H2 uptake and
low qH2 because of low potential intensity. The potential intensity
should be considered as a primary factor. How to achieve high sur-

Please cite this article in press as: Liu, J.; et al. Tetrahedron Lett. (2016), http://dx.doi.org/10.1016/j.tetlet.2016.09.085

J. Liu et al. / Tetrahedron Letters xxx (2016) xxxxxx

Figure 6. Organic ligands, structures and H2 isotherms of UIO-66 and UIO-67 at


77 K. Reproduced with permission from Ref. 19.

face area and useful space with high potential field for the H2 storage using atoms or groups as less as possible is subsequent work.
Some other porous materials, such as COFs structured by
organic ligands without any metal cluster, also possess ultrahigh
surface area (3472 m2/g for COF-102 and 4210 m2/g for COF-103)
on account of their low skeleton density (0.17 g/cm3 for COF-108,
Figure 7). COFs present atomically precise integration of their
building blocks into a porous structure throughout their extended
periodic networks in two or three dimensions under pp stacking
interactions. The main potential field in the pores of COFs is formed
by the atoms on the pore wall. While these atoms usually bond
with hydrogen atoms, the potential intensity is not very high. Isosteric heat of H2 adsorption on COFs is around 4 kJ/mol, indicating a
weak interaction between H2 and COFs. Most of COFs present high
gravimetric H2 uptake (7.24% for COF-102) attributed to their low
density, but having low volumetric uptake (37 g/L for MOF-102).
Two reasons lead to the low volumetric uptake. The first one is
the loss of large delocalized pp electrons. The second one is that
the pore size is too large to form an overlap potential field.
The qH2 of COF-6 was calculated to be 103 g/L, exceeding most
of materials including MOFs and carbons. In COF-6, the C2O2B ring
stacks in a line on which the neighboring O atoms from two layers
can capture H2 efficiently. The potential field from O atoms in six
member ring-based pores can overlap to enhance the potential
intensity, leading to high H2 isosteric heat than other COFs. The
isosteric heat of H2 on COF-6 reaches 7 kJ/mol, which is considered
a high value in COF materials. Comparing with COF-6, COF-1
adopts graphite-type stacking, and the O atoms on the neighboring
layer are not in a line. Such lack of cooperative effect on H2 results
in lower isosteric heat (6.3 kJ/mol) and lower H2 uptake (17 g/L).
Even having high intensity of potential field as COF-6, H2 uptake
of COF-1 is just 26 g/L because of its high crystal density (1.39 g/
cm3). When decreasing the crystal density, the increase of the pore

Figure 7. Upper panel: condensation reactions used to produce COFs. Middle panel: resulted fragments of COFs. Lower panel: atomic connectivity and structures of
crystalline COFs. The topology and the group classification number are indicated for each COF. Inset: the C2O2B and the B3O3 rings formed by condensation reactions.
Reproduced with permission from Ref. 22.

Please cite this article in press as: Liu, J.; et al. Tetrahedron Lett. (2016), http://dx.doi.org/10.1016/j.tetlet.2016.09.085

J. Liu et al. / Tetrahedron Letters xxx (2016) xxxxxx

size leads to cooperative effect in the pores. For example, the pore
size of COF-103 is 1.2 nm with the crystal density of 0.43 g/cm3,
and its H2 uptake is 36.8 g/L. By further increasing the pore size
to 3.2 nm as COF-10, the H2 uptake is just 25.5 L/g. Thus, a tradeoff
between the crystal density and the surface area should be considered for obtaining high H2 uptake. While improving the interaction
between H2 and adsorbents can enhance H2 uptake, none of real
materials have been reported except for some simulated results
(isosteric heat of 20 kJ/mol, H2 uptake of 60 g/L at 298 K for
Pd@COF-103).28,53
Methane storage
Some compounds containing high quantity of hydrogen atoms
such as formic acid, methanol, and CH4 are regarded as substitutes
of H2 during the transition period. It is no doubt that CH4 has the
highest hydrogen-containing ratio. CH4 has a heat value 2.85 times
higher than fossil fuel. It should be noted that CH4 storage is not
the same as natural gas storage. In natural gas, the CH4 component
is 90% with 10% C2+ compounds. The C2+ component can be condensed in the pores of adsorbents under the storage condition and
cannot be dispersed completely above 1 atm, leading to residuals
in the pores. Thus, these C2+ compounds have a significant negative effect on cycling storage capacity. The charge/discharge experimental results indicate that the storage capacity of CH4 on carbon
materials is reduced about 35% resulted from the C2+ residuals.
The negative effect by residuals should be removed by adding
pre-separation process to convert natural gas storage to CH4 storage. Similar to the H2 storage, the materials with high porosity
and surface area are promising in CH4 storage. CH4 molecule has
no dipole and quadrupole because of its high symmetry. The intermolecular force between CH4 and adsorbents is mainly produced

Figure 8. MOF-519 (c) and MOF-520 (d) built from octametallic inorganic
secondary building unit (a) and organic linker (b). In MOF-519, part of the
framework void space is occupied by dangling organic ligands. Reproduced from
Ref. 46.

by instant polarization. The isosteric heat of CH4 on porous carbon


is about 1620 kJ.39,54 The volumetric CH4 uptake on carbon based
materials can reach 140160 V/V, approaching the onboard CH4
storage target of 180 V/V.
MOFs with open metal sites allow the electron transfer from
CH4 sp3 orbital to the empty orbital of metal sites. At low CH4 coverage, the isosteric heat is around 1520 kJ/mol, which is similar to
that of porous carbon. In MOF materials, organic ligands have
weaker interaction with CH4 than metal sites, leading to decreased
isosteric heat upon increasing the CH4 coverage. In a similar potential field, the CH4 uptake is determined by the surface area and
crystal density of the adsorbents. High surface area would result
in high gravimetric CH4 uptake, and high crystal density is related
to high volumetric CH4 uptake. Thus, the tradeoff between these
two factors should be considered.
MOF-519 and MOF-520 (Fig. 8) show different crystal density
(0.94 and 0.58 g/cm3 for MOF-519 and MOF-520, respectively).
Their BET surface areas were estimated to be 2400 m2 g1 and
3290 m2 g1, respectively. The CH4 density in adsorptive phase
(qCH4) is 0.125 g/cm3 for MOF-519 and 0.132 g/cm3 for MOF-520,
corresponding to the uptake of 180 V/V (standard temperature
and pressure (STP)) and 155 V/V (STP), respectively. PCN-11 with
a micropore size of 7 shows high uptake of 198 V/V (STP). For
PCN-14, unbelievable uptake of 232 V/V (STP) was obtained with
high isosteric heat of 30 kJ/mol.49 The adsorptive sites in PCN-14
include copper ions corresponding to an enthalpy of 25 kJ/mol,
and small cages corresponding to an enthalpy of 32 kJ/mol. On
the other hand, the highest isosteric heat in other reports did not
reach this level.55 The same method used for the CH4 adsorption
on graphene surface gave an isosteric heat of 17 kJ/mol. It should
be noted that the CH4 isosteric heat in porous carbon is around
20 kJ/mol, due to the potential overlap in slit pores. Thus, the isosteric heat of 30 kJ/mol for PCN-14 should be over estimated.
The performance of COFs on CH4 storage is not good as their H2
storage. The uptake just reaches 128 V/V (STP) for COF-102 and
121 V/V (STP) for COF-103. Lower performance than MOFs is
attributed to different molecular property. CH4 has only octupole
moment, which has weak Coulombic interaction with polarized
sites. In COFs, the main adsorptive sites are the atoms exposed
on the pore walls, such as oxygen. Carbon and boron atoms are saturated, which have weak interactions with adsorbates. Different to
H2 with quadrupole moment, CH4 does not have interactions with
oxygen atoms from neighboring layers in COFs, resulting in a dramatic decrease of the adsorption performance. Some PCN materials
(PCN-11 and PCN-14) with polycyclic aromatic rings show different CH4 storage capacity as compared with PCN-61 and PCN-66
having a carbyne group due to the same reasons discussed.
Above discussed CH4 uptake is calculated according to the crystal density. In practice, the packing density cannot reach the crystal
density without any structure change. When the packing density is
0.58 g/cm3, the CH4 uptake on HKUST-1 is just 94 V/V (STP), with a
reduction of gravimetric uptake implying a damage of the crystal
integrity. As for AX-21, the packing or shaking density is generally
0.3 g/cm3, and its experimental CH4 uptake amount is 130 V/V
(STP) at 298 K near 35 bar. For MOF-5, its structure is changed
upon compressing to tablets. When its packing density is 0.57 g/
cm3, the surface area drops from 3300 m2/g to 2700 m2/g, indicating the structure collapse. Almost all the MOF materials are not
good for compressing. Thus, there is a large gap between the packing density and crystal density, resulting in a distance toward
onboard CH4 storage target.
The existed materials have not realized the target of onboard
gas storage. Moreover, the negative effect of C2+ residuals provides
also a challenge for cycling storage. Storing gas in gas hydrate is an
alternative technology for natural gas storage, which is free from
the effect of C2+ residuals. Gas hydrate is a kind of clathrate com-

Please cite this article in press as: Liu, J.; et al. Tetrahedron Lett. (2016), http://dx.doi.org/10.1016/j.tetlet.2016.09.085

J. Liu et al. / Tetrahedron Letters xxx (2016) xxxxxx

pound structured by water molecules linking through hydrogen


bonds with the theoretical gas uptake of 180 V/V (STP).56 The center of the cage can be occupied by some small molecules such as
CH4, C2H6, and CO2. In order to keep the gas hydrate state, the
chemical potential corresponding to pressure and temperature in
scale must reach the critical value. Thus, the C2+ component can
be entirely released upon the pressure decrease. Furthermore, feasible nucleation of C2+ component may accelerate the kinetics of
methane hydrate formation. However, both kinetics and gas
hydrate conversion are not satisfactory if the hydrate form is in
bulk water media. One of the solutions is enhancing the gasliquid
contact face by adding surfactants, dispersing water into droplets,
or stirring. But, it is not easy add these methods in a portable cylinder for commercialization purpose. Natural gas storage in wet porous materials can overcome the drawbacks of kinetics and gas
hydrate conversion, and the process can be easily carried out.57
The shortage of this adsorption method is the packing density
faced by all the adsorption storage techniques. We recently proposed a novel mechanism of gas hydrate formation in the nanoporous materials, which may weaken the effect of low packing density
on the storage capacity.57 In this case, the gas hydrate may not be
formed in the nanopores of nanoporous materials such as carbonbased materials. Under the hydrate formation conditions, water in
the nanopores is assembled to gas hydrate crystal to be out of the
pores. Meanwhile, the adsorption capacity of the pores is recovered
after the conversion of water to gas hydrate. By combining the
capacity of the pores and gas hydrate, it would be easy to achieve
the storage target without increasing the packing density. But,
keeping the low temperature about 2 C to make gas hydrate formation is still difficult for commercial onboard storage.
Summary
Alternative clean fuels such as hydrogen and methane have
been sought after, and the demands for effective storage technologies have been increasing. Porous materials are well positioned to
play an important role at the forefront of this research. Intrinsic
properties of porous materials including surface area, pore size,
pore volume, binding energy, and gas storage capacity have important effects on their H2 and CH4 storage capacity, and these factors
interact with each other. In the present work, we propose a new
concept of adsorptive density that is determined by all these factors. Several aspects need to be improved in future studies of the
field. (1) The potential intensity should be considered as one of
the important factors for evaluating the gas storage capability.
The strong potential intensity is not only provided by functional
groups, but also depends on the geometric structure of the materials. (2) The mechanical property of the porous materials should be
enhanced by considering the packing density as a determining factor for the final total storage capacity. (3) For natural gas storage,
the effect of C2+ component impurity on the storage capacity of
materials should be further considered. Although porous materials
show promising performance on the gas storage, a lot of studies
should be conducted for their practical applications in this area.
Acknowledgments
This work is supported by the National Research Foundation
(NRF), Prime Ministers Office, Singapore under its Campus for
Research Excellence and Technological Enterprise (CREATE) ProgrammeSingapore Peking University Research Centre for a Sustainable Low-Carbon Future, the NTU-A*Star Centre of Excellence
for Silicon Technologies under grant no. 11235100003, and Singapore Academic Research Fund (ACFR) Tier 1 (RG112/15). Authors
would also like to thank Yajuan Wei for useful discussions about
the content of this Digest.

References and notes


1. Zhou, L. Renew. Sustain. Energy Rev. 2005, 9, 395408.
2. http://energy.gov/eere/fuelcells/doe-technical-targets-onboard-hydrogenstorage-light-duty-vehicles, 2016 (accessed 29 July 2016).
3. Sakintuna, B.; Lamari-Darkrim, F.; Hirscher, M. Int. J. Hydrogen Energy 2007, 32,
11211140.
4. Choi, Y. J.; Westman, M.; Karkamkar, A.; Chun, J.; Rnnebro, E. C. E. Energy Fuels
2015, 29, 66956703.
5. Durbin, D. J.; Malardier-Jugroot, C. Int. J. Hydrogen Energy 2013, 38, 14595
14617.
6. Langmi, H. W.; Ren, J.; North, B.; Mathe, M.; Bessarabov, D. Electrochim. Acta
2014, 128, 368392.
7. Murray, L. J.; Dinca, M.; Long, J. R. Chem. Soc. Rev. 2009, 38, 12941314.
8. Hu, Y. H.; Zhang, L. Adv. Mater. 2010, 22, E117E130.
9. Zhang, G.; Musgrave, C. B. J. Phys. Chem. A 2007, 111, 15541561.
10. LeSar, R.; Herschbach, D. R. J. Phys. Chem. 1983, 87, 52025206.
11. Bimbo, N.; Xu, W.; Sharpe, J. E.; Ting, V. P.; Mays, T. J. Mater. Des. 2016, 89,
10861094.
12. Sethia, G.; Sayari, A. Carbon 2016, 99, 289294.
13. Wong-Foy, A. G.; Matzger, A. J.; Yaghi, O. M. J. Am. Chem. Soc. 2006, 128, 3494
3495.
14. Deng, H.; Grunder, S.; Cordova, K. E.; Valente, C.; Furukawa, H.; Hmadeh, M.;
Gndara, F.; Whalley, A. C.; Liu, Z.; Asahina, S.; Kazumori, H.; OKeeffe, M.;
Terasaki, O.; Stoddart, J. F.; Yaghi, O. M. Science 2012, 336, 10181023.
15. Zhou, W.; Wu, H.; Yildirim, T. J. Am. Chem. Soc. 2008, 130, 1526815269.
16. Gygi, D.; Bloch, E. D.; Mason, J. A.; Hudson, M. R.; Gonzalez, M. I.; Siegelman, R.
L.; Darwish, T. A.; Queen, W. L.; Brown, C. M.; Long, J. R. Chem. Mater. 2016, 28,
11281138.
17. Masika, E.; Mokaya, R. Prog. Nat. Sci. Mater. Int. 2013, 23, 308316.
18. Ben, T.; Ren, H.; Ma, S.; Cao, D.; Lan, J.; Jing, X.; Wang, W.; Xu, J.; Deng, F.;
Simmons, J. M.; Qiu, S.; Zhu, G. Angew. Chem., Int. Ed. 2009, 48, 94579460.
19. Chavan, S.; Vitillo, J. G.; Gianolio, D.; Zavorotynska, O.; Civalleri, B.; Jakobsen,
S.; Nilsen, M. H.; Valenzano, L.; Lamberti, C.; Lillerud, K. P.; Bordiga, S. Phys.
Chem. Chem. Phys. 2012, 14, 16141626.
20. Farha, O. K.; Yazaydn, A. .; Eryazici, I.; Malliakas, C. D.; Hauser, B. G.;
Kanatzidis, M. G.; Nguyen, S. T.; Snurr, R. Q.; Hupp, J. T. Nat. Chem. 2010, 2, 944
948.
21. 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, 424428.
22. Furukawa, H.; Yaghi, O. M. J. Am. Chem. Soc. 2009, 131, 88758883.
23. Gedrich, K.; Senkovska, I.; Klein, N.; Stoeck, U.; Henschel, A.; Lohe, M. R.;
Baburin, I. A.; Mueller, U.; Kaskel, S. Angew. Chem., Int. Ed. 2010, 49, 84898492.
24. Han, S. S.; Furukawa, H.; Yaghi, O. M.; Goddard, W. A. J. Am. Chem. Soc. 2008,
130, 1158011581.
25. Klein, N.; Senkovska, I.; Baburin, I. A.; Grnker, R.; Stoeck, U.; Schlichtenmayer,
M.; Streppel, B.; Mueller, U.; Leoni, S.; Hirscher, M.; Kaskel, S. Chem.-Eur. J.
2011, 17, 1300713016.
26. Klein, N.; Senkovska, I.; Gedrich, K.; Stoeck, U.; Henschel, A.; Mueller, U.;
Kaskel, S. Angew. Chem., Int. Ed. 2009, 48, 99549957.
27. 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.;
Schrder, M. J. Am. Chem. Soc. 2009, 131, 21592171.
28. Mendoza-Cortes, J. L.; Goddard, W. A.; Furukawa, H.; Yaghi, O. M. J. Phys. Chem.
Lett. 2012, 3, 26712675.
29. Rallapalli, P. B. S.; Raj, M. C.; Senthilkumar, S.; Somani, R. S.; Bajaj, H. C. Environ.
Prog. Sustain. Energy 2016, 35, 461468.
30. Rowsell, J. L. C.; Yaghi, O. M. J. Am. Chem. Soc. 2006, 128, 13041315.
31. Stoeck, U.; Krause, S.; Bon, V.; Senkovska, I.; Kaskel, S. Chem. Commun. 2012,
1084110843.
32. Yoon, M.; Moon, D. Microporous Mesoporous Mater. 2015, 215, 116122.
33. Yuan, D.; Zhao, D.; Sun, D.; Zhou, H.-C. Angew. Chem., Int. Ed. 2010, 49, 5357
5361.
34. Zhao, D.; Yuan, D.; Sun, D.; Zhou, H.-C. J. Am. Chem. Soc. 2009, 131, 91869188.
35. Zhou, L.; Zhou, Y.; Li, M.; Chen, P.; Wang, Y. Langmuir 2000, 16, 59555959.
36. Ihm, Y.; Cooper, V. R.; Peng, L.; Morris, J. R. J. Phys. Condens. Matter 2012, 24,
424205.
37. Zhao, W.; Fierro, V.; Zlotea, C.; Aylon, E.; Izquierdo, M. T.; Latroche, M.; Celzard,
A. Int. J. Hydrogen Energy 2011, 36, 54315434.
38. Mason, J. A.; Veenstra, M.; Long, J. R. Chem. Sci. 2014, 5, 3251.
39. Wang, X.; French, J.; Kandadai, S.; Chua, H. T. J. Chem. Eng. Data 2010, 55, 2700
2706.
40. Juan-Juan, J.; Marco-Lozar, J. P.; Surez-Garca, F.; Cazorla-Amors, D.; LinaresSolano, A. Carbon 2010, 48, 29062909.
41. Kondo, M.; Yoshitomi, T.; Matsuzaka, H.; Kitagawa, S.; Seki, K. Angew. Chem.,
Int. Ed. 1997, 36, 17251727.
42. Wu, H.; Zhou, W.; Yildirim, T. J. Am. Chem. Soc. 2009, 131, 49955000.
43. Moellmer, J.; Moeller, A.; Dreisbach, F.; Glaeser, R.; Staudt, R. Microporous
Mesoporous Mater. 2011, 138, 140148.
44. Senkovska, I.; Kaskel, S. Microporous Mesoporous Mater. 2008, 112, 108115.
45. Eddaoudi, M.; Kim, J.; Rosi, N.; Vodak, D.; Wachter, J.; OKeeffe, M.; Yaghi, O. M.
Science 2002, 295, 469472.
46. Gndara, F.; Furukawa, H.; Lee, S.; Yaghi, O. M. J. Am. Chem. Soc. 2014, 136,
52715274.
47. He, Y.; Zhou, W.; Yildirim, T.; Chen, B. Energy Environ. Sci. 2013, 6, 27352744.

Please cite this article in press as: Liu, J.; et al. Tetrahedron Lett. (2016), http://dx.doi.org/10.1016/j.tetlet.2016.09.085

J. Liu et al. / Tetrahedron Letters xxx (2016) xxxxxx


48. Wang, X.-S.; Ma, S.; Rauch, K.; Simmons, J. M.; Yuan, D.; Wang, X.; Yildirim, T.;
Cole, W. C.; Lpez, J. J.; Meijere, A. D.; Zhou, H.-C. Chem. Mater. 2008, 20, 3145
3152.
49. Wu, H.; Simmons, J. M.; Liu, Y.; Brown, C. M.; Wang, X.-S.; Ma, S.; Peterson, V.
K.; Southon, P. D.; Kepert, C. J.; Zhou, H.-C.; Yildirim, T.; Zhou, W. Chem.-Eur. J.
2010, 16, 52055214.
50. Wiersum, A. D.; Chang, J.-S.; Serre, C.; Llewellyn, P. L. Langmuir 2013, 29, 3301
3309.
51. Li, B.; Wen, H.-M.; Wang, H.; Wu, H.; Tyagi, M.; Yildirim, T.; Zhou, W.; Chen, B.
J. Am. Chem. Soc. 2014, 136, 62076210.

52. 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, 886891.
53. Mendoza-Corts, J. L.; Han, S. S.; Goddard, W. A. J. Phys. Chem. A 2012, 116,
16211631.
54. Himeno, S.; Komatsu, T.; Fujita, S. J. Chem. Eng. Data 2005, 50, 369376.
55. Peng, Y.; Krungleviciute, V.; Eryazici, I.; Hupp, J. T.; Farha, O. K.; Yildirim, T. J.
Am. Chem. Soc. 2013, 135, 1188711894.
56. Wang, W.; Bray, C. L.; Adams, D. J.; Cooper, A. I. J. Am. Chem. Soc. 2008, 130,
1160811609.
57. Zhou, L.; Liu, J.; Su, W.; Sun, Y.; Zhou, Y. Energy Fuels 2010, 24, 37893795.

Please cite this article in press as: Liu, J.; et al. Tetrahedron Lett. (2016), http://dx.doi.org/10.1016/j.tetlet.2016.09.085

You might also like