You are on page 1of 23

Subscriber access provided by FLORIDA STATE UNIV

Article
Experimental Study of CO, CH, and Water Vapor Adsorption
2 4

on a Dimethyl-Functionalized UiO-66 Framework


Himanshu Jasuja, and Krista S. Walton
J. Phys. Chem. C, Just Accepted Manuscript • DOI: 10.1021/jp311857e • Publication Date (Web): 07 Mar 2013
Downloaded from http://pubs.acs.org on March 16, 2013

Just Accepted

“Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted
online prior to technical editing, formatting for publication and author proofing. The American Chemical
Society provides “Just Accepted” as a free service to the research community to expedite the
dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts
appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been
fully peer reviewed, but should not be considered the official version of record. They are accessible to all
readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered
to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published
in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just
Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor
changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers
and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors
or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

The Journal of Physical Chemistry C is published by the American Chemical Society.


1155 Sixteenth Street N.W., Washington, DC 20036
Published by American Chemical Society. Copyright © American Chemical Society.
However, no copyright claim is made to original U.S. Government works, or works
produced by employees of any Commonwealth realm Crown government in the course
of their duties.
Page 1 of 22 The Journal of Physical Chemistry

1
2
3
4
5
6
7
8
Experimental Study of CO2, CH4, and Water Vapor
9
10
11
12 Adsorption on a Dimethyl-Functionalized UiO-66
13
14
15
16 Framework
17
18
19
20
21 Himanshu Jasuja and Krista S. Walton*
22
23
24
25
School of Chemical and Biomolecular Engineering, Georgia Institute of Technology, 311 Ferst
26
27 Drive NW, Atlanta, Georgia 30332
28
29
30 KEYWORDS: UiO-66; water stability; hydrophobicity; CO2 capture; selectivity
31
32
33
ABSTRACT: A water resistant, highly robust dimethyl-functionalized UiO-66 analogue (UiO-
34
35
36 66-DM) has been synthesized and characterized using N2 adsorption, powder X-ray diffraction,
37
38 1
H nuclear magnetic resonance, Fourier transform infrared spectroscopy, and thermo-gravimetric
39
40
41
analysis followed by mass spectroscopy. High-pressure (up to 20 bar) single-component
42
43 adsorption isotherm measurements of CO2 and CH4 have been performed on UiO-66-DM at
44
45 different temperatures (293-308 K). Adsorption isotherm data were modeled using the Toth
46
47
48 equation, and isosteric heats of adsorption were calculated using the Clausius-Clapeyron
49
50 equation. The Ideal Adsorbed Solution Theory (IAST) was used to calculate mixture
51
52 selectivities. Water adsorption experiments show that water adsorption loadings in UiO-66-DM
53
54
55 are reduced by almost 50% compared to the parent material. In the low pressure region,
56
57 functionalization of the benzenedicarboxylic ligand by 2,5-dimethyl leads to higher interactions
58
59
60
ACS Paragon Plus Environment
1
The Journal of Physical Chemistry Page 2 of 22

1
2
3
with both CO2 and CH4 compared to the parent UiO-66. CO2/ CH4 selectivity calculated from
4
5
6 IAST is higher for UiO-66 in the low pressure region (<5 bar) but for pressures > 5 bar, UiO-66-
7
8 DM is more CO2 selective than UiO-66.
9
10
11
12
13
14 1. INTRODUCTION
15
16
17 The synthesis and characterization of porous materials for use as adsorbents and catalysts have
18
19
20 been an active area of research for decades due to numerous uses in applications ranging from
21
22 petrochemicals and catalysis to adsorption-based selective separations.1 Traditional porous
23
24
materials such as zeolites and similar oxide-based materials have been a focal point since 1960.2
25
26
27 However, new developments in synthetic zeolites have been limited due to the relatively small
28
29 size of the pores, difficulty in the tuning of these pores, and difficulty in chemical modification
30
31
32
of the surface. Metal-organic frameworks (MOFs), which are made up of metal clusters linked to
33
34 each other by organic ligands,3 have provided a significant breakthrough in this regard.4 These
35
36 hybrid materials have extremely high surface areas and pore volumes, uniform pore sizes, and
37
38
39 chemical functionalities.5 Moreover, their pore sizes and chemical functionalities can be tuned by
40
41 modifying the metal group or organic linker.6,7 Hence, MOFs have attracted considerable interest
42
43 for their possible use in a variety of applications.8-10
44
45
46
47 The practical use of MOFs in many industrial applications such as natural gas sweetening
48
49 hinges on the framework stability under harsh conditions such as high pressures up to 60 bar11
50
51
52
and stability upon exposure to water vapor and acid gases.12 Novel MOFs are appearing at a very
53
54 rapid pace,9 and a few water stable examples have emerged recently. Cavka et al. was first to
55
56 synthesize a zirconium (IV)-based MOF (UiO-66) with exceptional stability.13 This stability has
57
58
59
60
ACS Paragon Plus Environment
2
Page 3 of 22 The Journal of Physical Chemistry

1
2
3
been attributed to the highly oxophilic nature of zirconium (IV), leading to the formation of a
4
5
6 very stable inorganic brick [Zr6O4(OH)4]. These Zr6-octahedra are bound to twelve 1,4-
7
8 benzenedicarboxylic acid (BDC) ligands (each Zr atom is 8-coordinated) leading to a 3D
9
10
11
arrangement of micro pores (6 Å). In 2010, Cohen et al.3 first reported the synthesis of amino-,
12
13 bromo-, nitro-, and naphthalene-functionalized UiO-66 MOFs. Later in the same year, Kandiah
14
15 et al.14 reported that amino-, bromo-, and nitro-functionalized UiO-66 MOFs also retain high
16
17
18 thermal and chemical stabilities similar to the parent material. Cmarik et al.15 observed high
19
20 stability in other UiO-66 variants including methoxy and naphthyl functional groups. Hence, the
21
22 UiO-66 framework is quite robust and is exceedingly open to isoreticular functionalization
23
24
25 without losing its high hydrothermal and chemical stabilities.
26
27
28 Schoenecker et al. showed that UiO-66 is quite hydrophilic, i.e., it adsorbs a large amount of
29
30
31
water vapor at 90% RH (22.42 mmol/g), which is higher than the water adsorption capacities of
32
33 traditional adsorbents such as zeolites 5A and 13X and BPL carbon.16 Hence, the UiO-66
34
35 framework will be most effective for use in separations under humid conditions only when its
36
37
38 affinity for water is reduced; otherwise, the pores will fill with water vapor, thus reducing the
39
40 adsorption capacity of the target gases, e.g., CO2 capture from flue gas/ natural gas streams. The
41
42 pore properties and host-guest interactions (hydrophilicity in case of water) can be systematically
43
44
45 tuned in MOFs through pre-synthesis ligand functionalization or post-synthetic modification
46
47 (PSM). Thus, their adsorption capacity and selectivity toward a targeted gas can be enhanced.
48
49
50
The idea of incorporating hydrophobic functional groups into a MOF structure to reduce water
51
52
53 adsorption has been suggested in many literature reports.15, 17-22 However, only a few of these
54
55 reports have examined CO2 and CH4 adsorption as well. Ideally, functionalized MOFs should
56
57
58
enhance the adsorption of the target gas without also enhancing the adsorption of water vapor
59
60
ACS Paragon Plus Environment
3
The Journal of Physical Chemistry Page 4 of 22

1
2
3
and other mixture components. Cmarik et al.15 evaluated CO2, CH4, N2, and water vapor
4
5
6 adsorption in several variants of UiO-66. The nonpolar naphthyl-functionalized material was
7
8 found to suppress water adsorption compared to the parent MOF, but the CO2 loadings were also
9
10
11
lowered. Cai et al.22 studied the adsorption capacity of CO2 and CH4 on the alkyl-functionalized
12
13 MOFs CuMBTC and CuEBTC, with respect to CuBTC or HKUST-1, and found that while water
14
15 adsorption is significantly lower, the CO2 loadings at low pressure were unaffected.
16
17
18
19 Here we report a detailed experimental study of CO2, CH4, and water vapor adsorption on a
20
21 new dimethyl-functionalized version of UiO-66 (UiO-66-DM, Fig. 1). This MOF was
22
23 developed simultaneously by Huang et al.,23 and our procedure is distinct from their report. In
24
25
26 their work, the water effect on the structure was only qualitatively examined, and no CH4
27
28 adsorption data were reported. For comparison with UiO-66-DM, we have also reproduced
29
30
31
adsorption isotherms for UiO-66 (Figs. 3, 5). Water vapor adsorption isotherms (Fig. 3) at 298 K
32
33 were measured up to 90 % RH, and the gas adsorption isotherms at different temperatures (293-
34
35 308 K, Figs. 5, S14-16) were measured up to 20 bar. The results show that upon
36
37
38 functionalization with hydrophobic dimethyl moieties, water vapor adsorption has been reduced
39
40 by almost 50% compared to the parent material. In fact, the water adsorption loading is now
41
42 even less than the water vapor loadings for commercial adsorbents such as zeolites 5A and
43
44
45 13X.16 Moreover, as expected, this functionalization do not render any adverse effect on the high
46
47 hydrothermal, chemical, and mechanical stability of UiO-66 framework (Figs. S7-S11). At high
48
49
pressures (> 5 bar) UiO-66-DM has better CO2/CH4 selectivity compared to UiO-66, which is
50
51
52 promising for natural gas sweetening processes where operating pressures can reach up to 60 bar.
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
4
Page 5 of 22 The Journal of Physical Chemistry

1
2
3
4
5
6
7 DMBDC
8
9
10
11
12 (DMBDC)
13
14
15
16
17
18 Figure 1. Synthesis of dimethyl-functionalized UiO-66 (UiO-66-DM).
19
20
21
Intensity (counts)

10000
22
23
24 6400
25
26
27 3600
28
29 1600
30
31
32 400
33
34
35 0
36 10 15 20 25 30 35 40 45 50
2Theta (°)
37
38
39 Figure 2. Comparison between PXRD patterns of as-synthesized UiO-66-DM, UiO-66, and the
40
41 theoretical pattern of UiO-66 simulated from single crystal data.
42
43
44 2. EXPERIMENTAL SECTION
45
46
47
48 2.1 MOF Synthesis. All of the employed chemicals were commercially available and used as
49
50 received without further purification from the following sources: Sigma-Aldrich, N,N′-
51
52
dimethylformamide (DMF); TCI America, 2,5-dimethyl 1,4-benzene-dicarboxylic acid
53
54
55 (DMBDC); Acros, zirconium (IV) chloride (ZrCl4), 1,4-benzene-dicarboxylic acid (BDC).
56
57
58
59
60
ACS Paragon Plus Environment
5
The Journal of Physical Chemistry Page 6 of 22

1
2
3
UiO-66. UiO-66 was synthesized and activated as reported by Schoenecker et al.16
4
5
6
7 UiO-66-DM. Synthesis of UiO-66-DM was performed using a slightly modified procedure of
8
9 Lillerud et al.13 reported for the parent UiO-66 MOF. UiO-66-DM was synthesized by
10
11 dissolving 0.681 mmol of zirconium (IV) chloride (ZrCl4) and 0.681 mmol of 2,5-dimethyl-1,4-
12
13
14 benzenedicarboxylic acid (DMBDC) in 26.5 mL of DMF at room temperature. The resulting
15
16 mixture was divided equally into two 20 mL scintillation vials. These vials were kept in a sand
17
18
19
bath inside an oven at 393 K for 24 h. The final product was cooled to room temperature and
20
21 then washed with DMF three times to remove the unreacted reactants before activation at 473 K.
22
23 Similar to the parent material, UiO-66-DM crystallizes as intergrown crystals that are too small
24
25
26 for structure determination by single crystal diffraction.
27
28
29 2.2 Characterization. Successful synthesis of UiO-66-DM was confirmed using 1H nuclear
30
31 magnetic resonance (NMR), Fourier transform infrared (FT-IR) spectroscopy, thermo-
32
33
34 gravimetric analysis (TGA) followed by mass spectroscopy (MS), powder X-ray diffraction
35
36 (PXRD), and N2 adsorption. Refer to the supporting information (SI) for more details.
37
38
39
2.3 Water Vapor Adsorption Isotherm Measurements. An Intelligent Gravimetric
40
41
42 Analyzer (IGA-3 series, Hiden Analytical Ltd.) was used to obtain water vapor adsorption
43
44 isotherms. A sample size of approximately 45 mg was used for collecting the water vapor
45
46
47 adsorption isotherm. Dry air was used as the carrier gas, with a percentage of it being bubbled
48
49 through a vessel filled with deionized water. Two mass flow controllers were used to vary the
50
51 ratio of saturated air and dry air so that the relative humidity (RH) could be controlled. Due to
52
53
54 water condensation in the equipment at higher humidities, experiments were conducted up to
55
56 90% RH. The total gas flow rate was set at 200 cc/min, and each adsorption/desorption step was
57
58
59
60
ACS Paragon Plus Environment
6
Page 7 of 22 The Journal of Physical Chemistry

1
2
3
allowed enough time (maximum time of 24 h) to approach equilibrium. UiO-66-DM was
4
5
6 activated in situ under vacuum at 473 K for 12 h to remove residual solvent molecules before
7
8 starting the adsorption measurements. UiO-66-DM was regenerated by heating at 473 K under
9
10
11
vacuum for 12 h after the water adsorption isotherm measurement and prior to PXRD and BET
12
13 analysis.
14
15
16 2.4 High Pressure Gas Adsorption Isotherm Measurements. An Intelligent Gravimetric
17
18
19 Analyzer (IGA-1 series, Hiden Analytical Ltd.) was used to collect pure gas (CO2 and CH4)
20
21 adsorption isotherms at various temperatures (293-308 K, Figs. S14-16) and pressures up to 20
22
23 bar. UiO-66-DM was activated in situ at 473 K under vacuum until no further weight loss was
24
25
26 observed. After activation, the system was maintained under vacuum, and the temperature was
27
28 adjusted to the desired value. A sample size of approximately 30 mg was used for the
29
30
31
measurements, and a maximum equilibration time of 30 minutes was used for each point in the
32
33 isotherm.
34
35
36 3. RESULTS AND DISCUSSION
37
38
39
40 3.1 Structure Characterization and Physical Properties. The PXRD pattern obtained for
41
42 the as-synthesized UiO-66-DM is compared with the pattern simulated from the single-crystal
43
44 structure of UiO-66 in Fig. 2. These patterns are consistent with each other, confirming that UiO-
45
46
47 66-DM has the same topology as the parent UiO-66. Evidence of the presence of the dimethyl
48
49 moieties on the BDC linker was obtained by characterizing the UiO-66-DM with 1H NMR and
50
51
52
FTIR spectroscopy. 1H NMR spectrum of activated (under vacuum at 473 K, 12 h) UiO-66-DM
53
54 displayed resonances associated with the DMBDC ligand (Fig. S1). The peak at 1650 cm-1 in the
55
56 FTIR pattern corresponds to C=O in DMF/DEF.24 Figure S2 shows peaks at 1660 cm-1 and 3430
57
58
59
60
ACS Paragon Plus Environment
7
The Journal of Physical Chemistry Page 8 of 22

1
2
3
cm-1 in the FTIR spectrum of as-synthesized UiO-66-DM which corresponds to C=O in DMF
4
5
6 and O-H in both the inorganic brick [Zr6O4(OH)4] and physisorbed water, respectively. Hence,
7
8 an absence of the peak at 1650 cm-1 in the FTIR spectrum of activated UiO-66-DM (Fig. S3)
9
10
11
confirms the removal of DMF molecules after activation. Moreover, the peak at 3430 cm-1 has
12
13 also disappeared for activated UiO-66-DM (Fig. S3), which confirms the removal of physisorbed
14
15 water and dehydroxylation of the Zr6O4(OH)4 cornerstone (each Zr atom is 8-coordinated) into
16
17
18 Zr6O6 (each Zr atom is 7-coordinated) during activation. However, Figure S3 is now showing a
19
20 peak at ~3680 cm-1 which also represents OH-groups on the Zr6O4(OH)4 corner stone. However,
21
22 intensity of this peak is small. This is consistent with the observation made by Valenzano et al.
23
24
25 for the parent UiO-66.25 Thus, heating the as-synthesized UiO-66-DM at 473 K under vacuum is
26
27 appropriate for activation. Furthermore, a peak is observed at 2930 cm-1 that corresponds to the
28
29
stretching frequency of the C-H bond in the dimethyl moieties attached to the benzene ring.
30
31
32 Hence, FTIR also confirms the presence of DMBDC ligand in this MOF.
33
34
35 The thermal stability of UiO-66-DM was examined using TGA. The decomposition during the
36
37
38 TGA experiment was followed by mass spectrometry (MS). Figure S4 shows that UiO-66-DM
39
40 decomposes at approximately 723 K, which is less than the reported decomposition temperature
41
42 of UiO-66 (813 K)13 but still higher than the typical value of 623 K observed for many other
43
44
45 MOFs.13 Dingemans et al.18 synthesized methyl functionalized MOF-5, and also observed a
46
47 similar decrease in the thermal stability. UiO-66-NH2 and UiO-66-NO2 have also been shown to
48
49
decompose at a lower temperature compared to UiO-66.14 However, the origin of the differences
50
51
52 in these thermal stabilities is still unclear. Similar to UiO-66,25 the TGA curve for as-synthesized
53
54 UiO-66-DM in Fig. S4 shows three weight loss regions before framework decomposition that
55
56
57
58
59
60
ACS Paragon Plus Environment
8
Page 9 of 22 The Journal of Physical Chemistry

1
2
3
corresponds to dehydration (physisorbed water, T < 373 K), solvent removal (DMF, 373 K < T <
4
5
6 453 K), and dehydroxylation of the Zr6O4(OH)4 cornerstone into Zr6O6 (453 K < T < 553 K).
7
8
9 In Fig. S5, around the decomposition temperature of UiO-66-DM, we see MS signals mainly
10
11 due to CO2 (at m/z value of 12, 28, 44), and p-xylene (at m/z value of 91, 106). This suggests
12
13
14 bond cleavage between the linker and the inorganic brick as well as cleavage of the bond
15
16 between p-xylene ring and the terminal carboxyl groups. The absence of an MS signal for DMF
17
18
19
(at m/z value of 38, 42, 73) confirms that heating the as-synthesized UiO-66-DM at 473 K under
20
21 vacuum is appropriate for activation. We also observe a weight loss in the temperature range of
22
23 473-573 K at m/z value of 18 due to release of water for the activated sample. This shows that
24
25
26 the activated material rehydroxylates or physisorbs water upon exposure to moisture present in
27
28 the air during the transfer of the sample to the TGA system. Wiersum et al.26 also showed that
29
30 activated/dehyroxylated UiO-66 rehydroxylates upon exposure to water vapor.
31
32
33
34
Table 1. Adsorption Loadings at 90% Relative Humidity and BET Surface Areas Before and
35
36 After Water Exposure.
37
38
39 Pore Pore Loading,
40 Material Volume† Diameter 90% RH Surface Area* (m2/g)
41 (cm3/g) (Å) (cm3 H2O/g) Before After % loss
42 ‡
UiO-66 0.52 ~6 0.43 1160 1136 2
43
44 UiO-66-DM 0.40 <6 0.24 811 797 2
45
46 *BET Analysis
47 †
Obtained from the Dubinin-Astakov model of N2 adsorption at 77K
48 ‡
49
Reported from literature16
50
51
52 As expected, N2 adsorption (Fig. S12) at 77 K on the activated UiO-66-DM sample shows that
53
54 functionalization by the dimethyl moieties leads to reduced porosity (Table 1) compared to the
55
56 parent MOF. The BET surface area of UiO-66-DM is calculated to be 811 m2/g, with a DA pore
57
58
59
60
ACS Paragon Plus Environment
9
The Journal of Physical Chemistry Page 10 of 22

1
2
3
volume of 0.40 cm3/g. Entry to the internal pores of the parent UiO-66 MOF is limited by
4
5
6 triangular windows of 6 Å.13 However, in UiO-66-DM, functionalization by the dimethyl
7
8 moieties should further reduce the window opening.
9
10
11 3.2 High Structural Stability. The sensitivity of MOFs under humid conditions is well
12
13
14 known, but water adsorption studies on MOFs are still lacking compared to other adsorbates.21,27-
15
16 31
Recently, we measured the water vapor adsorption isotherms of several well-known MOFs
17
18
19
including UiO-66 and it’s several variants and found that this family of MOFs is uniformly
20
21 stable in the presence of water.15,16 In the current study, water vapor adsorption measurements
22
23 for UiO-66-DM and subsequent stability tests were performed to compare with UiO-66.
24
25
26
27 In Fig. 3, UiO-66-DM shows significantly lower water adsorption loadings compared to UiO-
28
29 66 and, is also even lower than the water vapor loadings for traditional porous materials such as
30
31 zeolites 5A and 13X and BPL carbon.16 Among the UiO-66 variants studied by Cmarik et al.,15
32
33
34 the naphthyl functionalized MOF displayed the lowest water adsorption loadings but still
35
36 adsorbed approximately 13 mmol/g at 40% RH compared to only 6 mmol/g adsorbed in UiO-66-
37
38
DM under the same conditions. The desorption isotherms of these Zr-MOFs exhibit hysteresis
39
40
41 with a portion of the water being retained. At 0% RH, UiO-66 and UiO-66-DM retain 2.42
42
43 mmol/g, and 1.75 mmol/g, respectively. For comparison, UiO-66-NO2 was shown to retain
44
45
46
almost 6 mmol/g of water after desorption,15 and UiO-66-NH2 retains approximately 3.3
47
48 mmol/g.16 Wiersum et al.26 showed that the amount of water retained by UiO-66 after the first
49
50 adsorption cycle corresponds to partial rehydroxylation of the sample, and full rehydroxylation
51
52
53 was possible only after the third cycle. This, the same explanation of partial rehydroxylation can
54
55 be extended to the functionalized UiO-66 MOFs as well. However, we do expect a different
56
57 extent of rehydroxylation to occur after the first cycle for UiO-66-NO2, −NH2, and −DM
58
59
60
ACS Paragon Plus Environment
10
Page 11 of 22 The Journal of Physical Chemistry

1
2
3
compared to UiO-66 because the amount retained at 0 % RH appears to be related to the polarity
4
5
6 or hydrophilicity of the functional group. Thus, the amount of water retained in the MOFs at 0%
7
8 RH follows the trend UiO-66-NO2 > −NH2 > −DM ~ −Naphthyl ~ parent structure.
9
10
11 25
12
Loading (mmol H2O/ g MOF)
13
20
14
15
16 15
17
18 10
19
UiO-66-DM
20 5 UiO-66*
21
22
23 0
0 20 40 60 80 100
24
25 Relative Humidity (%)
26
27
28 Figure 3. Comparison of water vapor adsorption isotherms at 298 K for the desolvated
29
30
compounds UiO-66-DM and UiO-66 (closed symbols– adsorption; open symbols– desorption).
31
32
33 *Obtained from our previous work16
34
35
36 PXRD patterns for UiO-66-DM before and after water exposure are shown in Fig. 4. Similar
37
38
39 to the previously studied variants,15,16 UiO-66-DM also does not lose its crystallinity and is quite
40
41 robust as there is negligible loss in its BET surface area (shown in Table 1) even after exposure
42
43 to high levels of humidity. Moreover, functionalization with dimethyl moieties does not render
44
45
46 any adverse effect on the high chemical and mechanical stability of the framework (Figs. S8-
47
48 S11), in spite of the slightly lower thermal stability. The volumetric water loadings (cm3/g) at
49
50 90% RH are shown in Table 1. It is expected that water should completely fill the pores at
51
52
53 saturation, but for UiO-66, the water volume is slightly lower than the pore volume obtained
54
55 from nitrogen adsorption. This may be due to fact that only during water adsorption does the Zr
56
57
58
inorganic unit undergo rehydroxylation from [Zr6O6] to [Zr6O4(OH)4]. Thus, it is difficult to
59
60
ACS Paragon Plus Environment
11
The Journal of Physical Chemistry Page 12 of 22

1
2
3
directly compare pore volumes obtained from water adsorption with those obtained by N2
4
5
6 adsorption. The water volume obtained for UiO-66-DM is 40% lower than the pore volume
7
8 obtained by N2 adsorption. Similar to UiO-66, this decrease in volume relative to the N2 pore
9
10
11
volume will have contributions from the rehydroxylation process. However, adsorption
12
13 saturation has not been reached at 90% RH, so water does not fully fill the pores of UiO-66-DM.
14
15 Thus, the volume should be much lower than the N2 volume.
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41 Figure 4. PXRD patterns for as-synthesized (middle), water vapor-exposed (top), and
42
43
44 regenerated (bottom) UiO-66-DM
45
46
47 Zeolite 13X has often been employed commercially for CO2 separation from gas streams due
48
49
50
to its high CO2 adsorption capacity under dry conditions, but also must be regenerated at high
51
52 temperature (623 K) to retain this capacity and has a high affinity for moisture.32,33 Counter to
53
54 this, UiO-66-DM can be regenerated at relatively lower temperatures (473 K). Moreover, at 28
55
56
57 % RH zeolite 13X32 adsorbs 15.11 mmol H2O/g which is ~ 3.5 times more than the water
58
59
60
ACS Paragon Plus Environment
12
Page 13 of 22 The Journal of Physical Chemistry

1
2
3
adsorption in UiO-66-DM at 30 % RH. Since the pore volume and surface area of zeolite 13X
4
5
6 (0.25 cm3/g, 725 m2/g)34,35 are lower than UiO-66-DM (Table 1), we expect UiO-66-DM to have
7
8 a higher capacity for gases such as CO2 at higher pressures and outperform zeolite 13X,
9
10
11
especially under humid environments. Water adsorption studies on other MOFs, e.g. the DMOF
12
13 or Zn-BDC-DABCO family,36,37 also show much lower water loadings than zeolite 13X, but
14
15 many of these MOFs lose their crystallinity after exposure to humidity, with the exception of the
16
17
18 tetramethyl and anthracene derivatives. In general, the key to developing high performance
19
20 adsorbents is finding a balance between adsorption capacity and selectivity for the target; these
21
22 two characteristics are often inverse related.
23
24
25
26 3.3 CO2 and CH4 Adsorption. Inspired by the high stability and significantly low water
27
28 loadings of UiO-66-DM, pure-component CO2 and CH4 adsorption isotherms were measured.
29
30 The CO2 and CH4 adsorption isotherms for UiO-66 and UiO-66-DM shown in Fig. 5 exhibit type
31
32
33 I isotherms with no hysteresis. CO2 is more strongly adsorbed than CH4 because it has a
34
35 relatively high quadrupole moment while CH4 is nonpolar. At high pressure (Fig. 5a), the
36
37 adsorption loadings are higher for the parent material since it has a higher pore volume.
38
39
40 However, Fig. 5b shows that CO2 and CH4 loadings at low pressures are slightly enhanced in
41
42 UiO-66-DM due to the presence of increased van der Waals interactions from the dimethyl
43
44
45
moieties; the enhancement is relatively greater for CH4 than CO2. This can be more easily seen
46
47 by comparing the values of Henry’s constants (KH) for CH4 and CO2 adsorption on UiO-66-DM
48
49 and UiO-66 in Table 2. The Toth parameters (Eq S1)38 were used to calculate Henry’s constants
50
51
52 (Figs. 6, S14-15). Results show that functionalization with dimethyl moieties leads to an increase
53
54 in the value of KH (Table 2) by a factor of 2 for CO2 and a factor of 3 for CH4 compared to the
55
56 parent MOF. This is disadvantageous from CO2/CH4 selectivity point of view in the low-pressure
57
58
59
60
ACS Paragon Plus Environment
13
The Journal of Physical Chemistry Page 14 of 22

1
2
3
region. These results are different than the behavior observed for UiO-66-Naphthyl where the
4
5
6 Henry’s constant is essentially unchanged compared to the parent material.15 Steric factors are
7
8 clearly at play in these examples since the BDC in UiO-66-DM is functionalized on both sides of
9
10
11
the benzene ring and on one side for UiO-66-Naphthyl.
12
13
14
15
(a)
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32 (b)
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49 Figure 5. Experimental adsorption-desorption isotherms for CO2 and CH4 in UiO-66, and UiO-
50
51
52 66-DM at 303 K (closed symbols-adsorption; open symbols-desorption): (a) 0−20,000 mbar and
53
54 (b) 0−2000 mbar. Lines connecting the adsorption points are to faciliate viewing. *Reported
55
56
from literature.26
57
58
59
60
ACS Paragon Plus Environment
14
Page 15 of 22 The Journal of Physical Chemistry

1
2
3
4
5
6
7
8 Table 2. Toth Equation Parameters, Henry’s Constants (KH), and Low Coverage CO2/ CH4
9
10 Selectivities at 303 K for UiO-66 and UiO-66-DM.
11
12
13
14 KHCO2 = Ns*b KHCH4 = Ns*b Ideal
15 MOF CO2 CH4
Ns b t Ns b t (mmol/g*mbar) (mmol/g*mbar) Selectivity
16 KH(CO2)/KH(CH4)
17
18 UiO-66 10.11 2.06E-04 0.76 42.95 2.28E-05 0.35 2.08E-03 9.78E-04 2.12
19
20 UiO-66-DM 07.71 4.92E-04 0.61 10.53 2.58E-04 0.36 3.79E-03 2.72E-03 1.39
21
22
23 The ideal adsorbed solution theory (IAST)39 has been applied to calculate CO2/CH4 selectivity
24
25
26
for an equimolar mixture at 303 K in UiO-66 and UiO-66-DM (Fig 6). Up to 5000 mbar, the CO2
27
28 selectivity over CH4 is higher in UiO-66, but above 5000 mbar, UiO-66-DM has the higher
29
30 selectivity. At low pressure, CH4 molecules are adsorbed in the tetrahedral cages of UiO-66-DM
31
32
33 at a larger concentration compared to CO2 due to enhanced van der Waals interactions with the
34
35 dimethyl moieties. However, in UiO-66 under the same conditions, a larger concentration of CO2
36
37 molecules is adsorbed in the tetrahedral cages.40 Now, when pressure increases beyond 5000
38
39
40 mbar, the CO2/CH4 selectivity is higher for UiO-66-DM because of the competitive adsorption
41
42 behavior in which adsorbed CO2 molecules push CH4 molecules to the octahedral cages. Good
43
44
CO2/ CH4 selectivity at high pressure is advantageous for processes such as natural gas
45
46
47 sweetening where the operating pressure can range up to 60 bar.11 Cu-BTC has been shown to
48
49 have a higher CO2/CH4 selectivity (5-9 for 50:50 CO2/ CH4 mixture)41 compared to UiO-66-DM,
50
51
52
but water adsorbs strongly in Cu-BTC, and it is eventually unstable under humid conditions.
53
54 CO2/CH4 selectivity of UiO-66-DM at high pressures is comparable to other well-known MOFs
55
56 such as MIL-53 (Al, Cr).42,43
57
58
59
60
ACS Paragon Plus Environment
15
The Journal of Physical Chemistry Page 16 of 22

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25 Figure 6. CO2/CH4 selectivity for equimolar CO2/CH4 gas mixture in UiO-66, and UiO-66-DM
26
27 at 303 K, calculated using IAST. Lines connecting the points are to faciliate viewing. *Pure
28
29
component excess loading data for CO2 and CH4 adsorption in UiO-66 is taken from literature.26
30
31
32
33
34 The isosteric heat of adsorption was calculated by applying the Clausius-Clapeyron equation
35
36
37 (Eq S2)38 to adsorption isotherms measured at three temperatures (Figs. S14,15). The isosteric
38
39 heats of adsorption (Fig. 7) calculated for CH4 and CO2 on UiO-66-DM at a loading of 0.05
40
41 mmol/g are 30.9 kJ/mol, and 46.8 kJ/mol, respectively. These values are quite high compared to
42
43
44 UiO-66 and its other functionalized versions.15,40,44 The high stability of UiO-66-DM, low
45
46 interaction with water, and good CO2/CH4 selectivity at high pressures compared to UiO-66
47
48 make it a promising adsorbent for adsorption separations at these conditions.
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
16
Page 17 of 22 The Journal of Physical Chemistry

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
Figure 7. Isosteric heat of adsorption for CO2 and CH4 in UiO-66-DM as a function of adsorbate
21
22
23 loading. Qst is calculated using the Clausius-Clapeyron equation.
24
25
26 4. CONCLUSION
27
28
29
30 In summary, the synthesis and characterization of a new water resistant, highly robust dimethyl-
31
32 functionalized UiO-66 analogue (UiO-66-DM) has been carried out. Our study shows that
33
34 functionalization by dimethyl moieties significantly reduces the water loading by almost a factor
35
36
37 of 2 compared to UiO-66 without altering the high stability of the framework. Moreover, the
38
39 water adsorption loadings and regeneration temperature of UiO-66-DM are significantly lower
40
41 than the benchmark material zeolite 13X. UiO-66-DM also performs better than the parent MOF
42
43
44 for CO2/ CH4 separation at high pressures. While amine groups have been a topic of intense
45
46 focus for CO2 separations, this work shows that nonpolar functional groups may also play an
47
48
49
important role in enhancing CO2 adsorption while lowering interactions with water. UiO-66-DM
50
51 has low affinity for water, is highly robust, and has a good CO2/CH4 selectivity at high pressures.
52
53 In general, understanding the trade-offs between water adsorption behavior and high selectivity
54
55
56 and high capacity for the target will be key to developing new adsorbents for CO2 capture.
57
58
59
60
ACS Paragon Plus Environment
17
The Journal of Physical Chemistry Page 18 of 22

1
2
3
ASSOCIATED CONTENT
4
5
6 Supporting Information.
7
8 1
H NMR spectrum, FT-IR spectra, TG-MS curves, PXRD patterns, and N2 adsorption data. This
9
10
11
material is available free of charge via the Internet at http://pubs.acs.org.
12
13
14
15 AUTHOR INFORMATION
16
17 Corresponding Author
18
19
20 *E-mail: krista.walton@chbe.gatech.edu. Fax: +1-404-894-2866. Tel: +1-404-894-5254.
21
22
23 Notes
24
25
The authors declare no competing financial interest.
26
27
28
29
30
31 ACKNOWLEDGMENT
32
33 This material is based upon work supported by Army Research Office PECASE Award
34
35
36 W911NF-10-1-0079 and Contract W911NF-10-1-0076.
37
38
39
40 REFERENCES
41
42
43
(1) Davis, M. E. Ordered Porous Materials for Emerging Applications. Nature 2002, 417,
44 813-821.
45
46 (2) Ferey, G. Hybrid Porous Solids: Past, Present, Future. Chem. Soc. Rev. 2008, 37, 191-
47 214.
48
49 (3) Garibay, S. J.; Cohen, S. M. Isoreticular Synthesis and Modification of Frameworks with
50 the UiO-66 Topology. Chem. Commun. 2010, 46, 7700-7702.
51
52 (4) Ferey, G.; Mellot-Draznieks, C.; Serre, C.; Millange, F. Crystallized Frameworks with
53 Giant Pores: Are There Limits to the Possible? Acc. Chem. Res. 2005, 38, 217-225.
54
55 (5) Li, J.-R.; Kuppler, R. J.; Zhou, H.-C. Selective Gas Adsorption and Separation in Metal-
56
57 Organic Frameworks. Chem. Soc. Rev. 2009, 38, 1477-1504.
58
59
60
ACS Paragon Plus Environment
18
Page 19 of 22 The Journal of Physical Chemistry

1
2
3
(6) Eddaoudi, M.; Kim, J.; Rosi, N.; Vodak, D.; Wachter, J.; O'Keeffe, M.; Yaghi, O. M.
4
5 Systematic Design of Pore Size and Functionality in Isoreticular MOFs and Their Application in
6 Methane Storage. Science 2002, 295, 469-472.
7
8 (7) Karra, J. R.; Walton, K. S. Effect of Open Metal Sites on Adsorption of Polar and
9 Nonpolar Molecules in Metal-Organic Framework Cu-BTC. Langmuir 2008, 24, 8620-8626.
10
11 (8) Keskin, S.; Van Heest, T. M.; Sholl, D. S. Can Metal-Organic Framework Materials Play
12 a Useful Role in Large-Scale Carbon Dioxide Separations? ChemSusChem 2010, 3, 879-891.
13
14 (9) Keskin, S.; Kizilel, S. Biomedical Applications of Metal Organic Frameworks. Ind. Eng.
15
16
Chem. Res. 2011, 50, 1799-1812.
17
18
(10) Kitagawa, S.; Kitaura, R.; Noro, S. Functional Porous Coordination Polymers. Angew.
19 Chem.-Int. Edit. 2004, 43, 2334-2375.
20
21 (11) Herm, Z. R.; Krishna, R.; Long, J. R. CO2/CH4, CH4/H2 and CO2/CH4/H2 Separations at
22 High Pressures Using Mg2(dobdc). Microporous and Mesoporous Mater. 2012, 151, 481.
23
24 (12) Han, S.; Huang, Y.; Watanabe, T.; Dai, Y.; Walton, K. S.; Nair, S.; Sholl, D. S.;
25 Meredith, J. C. High-Throughput Screening of Metal-Organic Frameworks for CO2 Separation.
26 ACS Comb. Sci. 2012, 14, 263-267.
27
28 (13) Cavka, J. H.; Jakobsen, S.; Olsbye, U.; Guillou, N.; Lamberti, C.; Bordiga, S.; Lillerud,
29
30
K. P. A New Zirconium Inorganic Building Brick Forming Metal Organic Frameworks with
31 Exceptional Stability. J. Am. Chem. Soc. 2008, 130, 13850-13851.
32
33 (14) Kandiah, M.; Nilsen, M. H.; Usseglio, S.; Jakobsen, S.; Olsbye, U.; Tilset, M.; Larabi, C.;
34 Quadrelli, E. A.; Bonino, F.; Lillerud, K. P. Synthesis and Stability of Tagged UiO-66 Zr-MOFs.
35 Chem. Mater. 2010, 22, 6632-6640.
36
37 (15) Cmarik, G. E.; Kim, M.; Cohen, S. M.; Walton, K.S. Tuning the Adsorption Properties of
38 UiO-66 via Ligand Functionalization. Langmuir 2012, 28, 15606-15613.
39
40 (16) Schoenecker, P. M.; Carson, C. G.; Jasuja, H.; Flemming, C. J. J.; Walton, K. S. Effect of
41
42 Water Adsorption on Retention of Structure and Surface Area of Metal–Organic Frameworks.
43 Ind. Eng. Chem. Res. 2012, 51, 6513-6519.
44
45 (17) Wu, T.; Shen, L.; Luebbers, M.; Hu, C.; Chen, Q.; Ni, Z.; Masel, R. I. Enhancing the
46 Stability of Metal-Organic Frameworks in Humid Air by Incorporating Water Repellent
47 Functional Groups. Chem. Commun. 2010, 46, 6120-6122.
48
49 (18) Yang, J.; Grzech, A.; Mulder, F. M.; Dingemans, T. J. Methyl Modified MOF-5: A Water
50 Stable Hydrogen Storage Material. Chem. Commun. 2011, 47, 5244-5246.
51
52
(19) Nguyen, J. G.; Cohen, S. M. Moisture-Resistant and Superhydrophobic Metal-Organic
53
54 Frameworks Obtained via Postsynthetic Modification. J. Am. Chem. Soc. 2010, 132, 4560-4561.
55
56 (20) Paranthaman, S.; Coudert, F.-X.; Fuchs, A. H. Water Adsorption in Hydrophobic MOF
57 Channels. Phys. Chem.Chem. Phys. 2010, 12, 8123-8129.
58
59
60
ACS Paragon Plus Environment
19
The Journal of Physical Chemistry Page 20 of 22

1
2
3
(21) Ma, D.; Li, Y.; Li, Z. Tuning the Moisture Stability of Metal-Organic Frameworks by
4
5 Incorporating Hydrophobic Functional Groups at Different Positions of Ligands. Chem.
6 Commun. 2011, 47, 7377-7379.
7
8 (22) Cai, Y.; Zhang, Y. D.; Huang, Y. G.; Marder, S. R.; Walton, K. S. Impact of Alkyl-
9 Functionalized BTC on Properties of Copper-Based Metal-Organic Frameworks. Cryst. Growth
10 Des. 2012, 12, 3709-3713.
11
12 (23) Huang, Y.; Qin, W.; Li, Z.; Li, Y. Enhanced Stability and CO2 Affinity of a UiO-66 Type
13
Metal-Organic Framework Decorated with Dimethyl Groups. Dalton Trans. 2012, 41, 9283-
14
15 9285.
16
17 (24) Mu, B.; Huang, Y.; Walton, K. S. A Metal-Organic Framework with Coordinatively
18 Unsaturated Metal Centers and Microporous Structure. Crystengcomm 2010, 12, 2347-2349.
19
20 (25) Valenzano, L.; Civalleri, B.; Chavan, S.; Bordiga, S.; Nilsen, M. H.; Jakobsen, S.;
21 Lillerud, K. P.; Lamberti, C. Disclosing the Complex Structure of UiO-66 Metal Organic
22 Framework: A Synergic Combination of Experiment and Theory. Chem. Mater. 2011, 23, 1700-
23
1718.
24
25
(26) Wiersum, A. D.; Soubeyrand-Lenoir, E.; Yang, Q.; Moulin, B.; Guillerm, V.; Yahia, M.
26
27 B.; Bourrelly, S.; Vimont, A.; Miller, S.; Vagner, C.; Daturi, M.; Clet, G.; Serre, C.; Maurin, G.;
28 Llewellyn, P. L. An Evaluation of UiO-66 for Gas-Based Applications. Chem. Asian J. 2011, 6,
29 3270-3280.
30
31 (27) Greathouse, J. A.; Allendorf, M. D. The Interaction of Water with MOF-5 Simulated by
32 Molecular Dynamics. J. Am. Chem. Soc. 2006, 128, 10678-10679.
33
34 (28) Kaye, S. S.; Dailly, A.; Yaghi, O. M.; Long, J. R. Impact of Preparation and Handling on
35
the Hydrogen Storage Properties of Zn4O(1,4-benzenedicarboxylate)3 (MOF-5) J. Am. Chem.
36
37 Soc. 2007, 129, 14176-14177.
38
39 (29) Kondo, A.; Daimaru, T.; Noguchi, H.; Ohba, T.; Kaneko, K.; Kanob, H. Adsorption of
40 Water on Three-Dimensional Pillared-Layer Metal-Organic Frameworks. J. Colloid Interface
41 Sci. 2007, 314, 422-426.
42
43 (30) Li, Y.; Yang, R. T. Gas Adsorption and Storage in Metal-Organic Framework MOF-177.
44 Langmuir 2007, 23, 12937-12944.
45
46 (31) Low, J. J.; Benin, A. I.; Jakubczak, P.; Abrahamian, J. F.; Faheem, S. A.; Willis, R. R.
47
48
Virtual High Throughput Screening Confirmed Experimentally: Porous Coordination Polymer
49 Hydration. J. Am. Chem. Soc. 2009, 131, 15834-15842.
50
51 (32) Franchi, R. S.; Harlick, P. J. E.; Sayari, A. Applications of Pore-Expanded Mesoporous
52 Silica. 2. Development of a High-Capacity, Water-Tolerant Adsorbent for CO2. Ind. Eng. Chem.
53 Res. 2005, 44, 8007-8013.
54
55 (33) Cavenati, S.; Grande, C. A.; Rodrigues, A. E. Adsorption Equilibrium of Methane,
56 Carbon Dioxide, and Nitrogen on Zeolite 13X at High Pressures. J. Chem. Eng. Data 2004, 49,
57
58
1095-1101.
59
60
ACS Paragon Plus Environment
20
Page 21 of 22 The Journal of Physical Chemistry

1
2
3
(34) Cortes, F. B.; Chejne, F.; Carrasco-Marin, F.; Moreno-Castilla, C.; Perez-Cadenas, A. F.
4
5 Water Adsorption on Zeolite 13X: Comparison of the Two Methods Based on Mass
6 Spectrometry and Thermogravimetry. Adsorption 2010, 16, 141-146.
7
8 (35) Lee, J. S.; Kim, J. H.; Kim, J. T.; Suh, J. K.; Lee, J. M.; Lee, C. H. Adsorption Equilibria
9 of CO2 on Zeolite 13X and Zeolite X/Activated Carbon Composite. J. Chem. Eng. Data 2002,
10 47, 1237-1242.
11
12 (36) Jasuja, H.; Huang, Y.; Walton, K. S. Adjusting the Stability of Metal–Organic
13
Frameworks under Humid Conditions by Ligand Functionalization. Langmuir 2012, 28, 16874-
14
15 16880.
16
17 (37) Jasuja, H.; Burtch, N. C.; Huang, Y.; Cai, Y.; Walton, K. S., Kinetic Water Stability of an
18 Isostructural Family of Zinc-Based Pillared Metal–Organic Frameworks. Langmuir 2012, 29,
19 633-642.
20
21 (38) Duong, D. D. In Adsorption Analysis: Equilibria and Kinetics. Imperial College Press,
22 London 1998.
23
24 (39) Myers, A. L.; Prausnitz, J. M. Thermodynamics of Mixed-Gas Adsorption. AICHE J.
25
26
1965, 11, 121-127.
27
28
(40) Yang, Q. Y.; Wiersum, A. D.; Jobic, H.; Guillerm, V.; Serre, C.; Llewellyn, P. L.;
29 Maurin, G. Understanding the Thermodynamic and Kinetic Behavior of the CO2/CH4 Gas
30 Mixture within the Porous Zirconium Terephthalate UiO-66(Zr): A Joint Experimental and
31 Modeling Approach. J. Phys. Chem. C 2011, 115, 13768-13774.
32
33 (41) Hamon, L.; Jolimaitre, E.; Pirngruber, G. D. CO2 and CH4 Separation by Adsorption
34 Using Cu-BTC Metal-Organic Framework. Ind. Eng. Chem. Res. 2010, 49, 7497-7503.
35
36 (42) Hamon, L.; Llewellyn, P. L.; Devic, T.; Ghoufi, A.; Clet, G.; Guillerm, V.; Pirngruber, G.
37
38
D.; Maurin, G.; Serre, C.; Driver, G.; Beek, W. V.; Jolimai tre, E.; Vimont, A.; Daturi, M.;
39 Férey, G. R. Co-adsorption and Separation of CO2−CH4 Mixtures in the Highly Flexible MIL-
40 53(Cr) MOF. J. Am. Chem. Soc. 2009, 131, 17490-17499.
41
42 (43) Finsy, V.; Ma, L.; Alaerts, L.; De Vos, D. E.; Baron, G. V.; Denayer, J. F. M. Separation
43 of CO2/CH4 Mixtures with the MIL-53(Al) Metal–Organic Framework. Microporous and
44 Mesoporous Mater. 2009, 120, 221-227.
45
46 (44) Yang, Q.; Wiersum, A. D.; Llewellyn, P. L.; Guillerm, V.; Serred, C.; Maurin, G.
47
48
Functionalizing Porous Zirconium Terephthalate UiO-66(Zr) for Natural Gas Upgrading: A
49 Computational Exploration. Chem. Commun. 2011, 47, 9603-9605.
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
21
The Journal of Physical Chemistry Page 22 of 22

1
2
3
TOC GRAPHIC
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
22

You might also like