research letters\(\def\hfill{\hskip 5em}\def\hfil{\hskip 3em}\def\eqno#1{\hfil {#1}}\)

IUCrJ
Volume 4| Part 2| March 2017| Pages 131-135
ISSN: 2052-2525

Metal–organic frameworks for H2 and CH4 storage: insights on the pore geometry–sorption energetics relationship

aDivision of Physical Sciences and Engineering, Advanced Membranes and Porous Materials Center, Functional Materials Design, Discovery and Development Research Group (FMD3), King Abdullah University of Science and Technology (KAUST), Thuwal 4700, Saudi Arabia
*Correspondence e-mail: mohamed.eddaoudi@kaust.edu.sa

Edited by L. R. MacGillivray, University of Iowa, USA (Received 20 September 2016; accepted 29 November 2016; online 10 February 2017)

This study aims to assess the possibility of improving H2 and CH4 binding affinity to the aromatic walls of a designed new Metal–Organic Framework (MOF) through simultaneous dispersive interactions. It is suggested here that desirable H2 and CH4 storage media at low pressures require narrow uniform pores associated with large surface area, a trade-off that is challenging to achieve.

1. Introduction

Metal–Organic Frameworks (MOFs), an emerging class of functional solid-state materials, continue to receive wide scientific interest due to their potential applications in hydrogen storage, gas separation, carbon dioxide capture, enhanced catalysis and drug delivery (Eddaoudi et al., 2001[Eddaoudi, M., Moler, D. B., Li, H., Chen, B., Reineke, T. M., O'Keeffe, M. & Yaghi, O. M. (2001). Acc. Chem. Res. 34, 319-330.]; Moulton & Zaworotko, 2001[Moulton, B. & Zaworotko, M. J. (2001). Chem. Rev. 101, 1629-1658.]; Kitagawa et al., 2004[Kitagawa, S., Kitaura, R. & Noro, S. (2004). Angew. Chem. Int. Ed. 43, 2334-2375.]; Férey et al., 2005[Férey, G., Mellot-Draznieks, C., Serre, C. & Millange, F. (2005). Acc. Chem. Res. 38, 217-225.]; Horcajada et al., 2006[Horcajada, P., Serre, C., Vallet-Regí, M., Sebban, M., Taulelle, F. & Férey, G. (2006). Angew. Chem. 118, 6120-6124.]; Cho et al., 2007[Cho, S. H., Gadzikwa, T., Afshari, M., Nguyen, S. T. & Hupp, J. T. (2007). Eur. J. Inorg. Chem. 2007, 4863-4867.]). MOFs possess unique structural attributes including dual composition, crystallinity and a modular pore system (Rowsell & Yaghi, 2006[Rowsell, J. L. & Yaghi, O. M. (2006). J. Am. Chem. Soc. 128, 1304-1315.]; Belof et al., 2007[Belof, J. L., Stern, A. C., Eddaoudi, M. & Space, B. (2007). J. Am. Chem. Soc. 129, 15202-15210.]; Cho et al., 2007[Cho, S. H., Gadzikwa, T., Afshari, M., Nguyen, S. T. & Hupp, J. T. (2007). Eur. J. Inorg. Chem. 2007, 4863-4867.]; Dincă et al., 2007[Dincă, M., Han, W. S., Liu, Y., Dailly, A., Brown, C. M. & Long, J. R. (2007). Angew. Chem. Int. Ed. 46, 1419-1422.]; Hayashi et al., 2007[Hayashi, H., Côté, A. P., Furukawa, H., O'Keeffe, M. & Yaghi, O. M. (2007). Nat. Mater. 6, 501-506.]; Liu et al., 2007[Liu, Y., Eubank, J. F., Cairns, A. J., Eckert, J., Kravtsov, V. C., Luebke, R. & Eddaoudi, M. (2007). Angew. Chem. Int. Ed. 46, 3278-3283.]; Alkordi et al., 2008[Alkordi, M. H., Liu, Y., Larsen, R. W., Eubank, J. F. & Eddaoudi, M. (2008). J. Am. Chem. Soc. 130, 12639-12641.]; Banerjee et al., 2008[Banerjee, R., Phan, A., Wang, B., Knobler, C., Furukawa, H., O'Keeffe, M. & Yaghi, O. M. (2008). Science, 319, 939-943.]; Llewellyn et al., 2008[Llewellyn, P. L., Bourrelly, S., Serre, C., Vimont, A., Daturi, M., Hamon, L., De Weireld, G., Chang, J.-S., Hong, D.-Y., Kyu Hwang, Y., Hwa Jhung, S. & Férey, G. (2008). Langmuir, 24, 7245-7250.]; Nugent et al., 2013[Nugent, P., Belmabkhout, Y., Burd, S. D., Cairns, A. J., Luebke, R., Forrest, K., Pham, T., Ma, S., Space, B., Wojtas, L., Eddaoudi, M. & Zaworotko, M. J. (2013). Nature, 495, 80-84.]; Belmabkhout et al., 2014[Belmabkhout, Y., Mouttaki, H., Eubank, J. F., Guillerm, V. & Eddaoudi, M. (2014). RSC Adv. 4, 63855-63859.]; Shekhah et al., 2014[Shekhah, O., Belmabkhout, Y., Chen, Z., Guillerm, V., Cairns, A., Adil, K. & Eddaoudi, M. (2014). Nat. Commun 5, 4428.]). Markedly, these attributes are ideal for the assessment and the establishment of the requisite structure–function relationship toward the construction of made-to-order MOFs for a targeted application. In particular, porous MOFs are regarded as prospective adsorbents that can offer practical solutions to the enduring challenges pertaining to the safe storage and efficient use of H2 in mobile applications. Conceivably, MOFs are widely investigated for hydrogen storage due to the ability to control their pore system (functionality and volume) and subsequently impact the H2–MOF interactions and the total H2 uptake (Rowsell & Yaghi, 2005[Rowsell, J. L. & Yaghi, O. M. (2005). Angew. Chem. Int. Ed. 44, 4670-4679.]; Collins & Zhou, 2007[Collins, D. J. & Zhou, H.-C. (2007). J. Mater. Chem. 17, 3154-3160.]; Lin et al., 2007[Lin, X., Jia, J., Hubberstey, P., Schröder, M. & Champness, N. R. (2007). CrystEngComm, 9, 438-448.]; Chen et al., 2008[Chen, B., Zhao, X., Putkham, A., Hong, K., Lobkovsky, E. B., Hurtado, E. J., Fletcher, A. J. & Thomas, K. M. (2008). J. Am. Chem. Soc. 130, 6411-6423.]; Dincă & Long, 2008[Dincă, M. & Long, J. R. (2008). Angew. Chem. Int. Ed. 47, 6766-6779.]; Nouar et al., 2008[Nouar, F., Eubank, J. F., Bousquet, T., Wojtas, L., Zaworotko, M. J. & Eddaoudi, M. (2008). J. Am. Chem. Soc. 130, 1833-1835.]; Kishan et al., 2010[Radha Kishan, M., Tian, J. K., Thallapally, P. K., Fernandez, C. A., Dalgarno, S. J., Warren, J. E., McGrail, B. P. & Atwood, J. L. (2010). Chem. Commun. 46, 538-540.]; Zhou et al., 2012[Zhou, H.-C., Long, J. R. & Yaghi, O. M. (2012). Chem. Rev. 112, 673-674.]). Our group, among others, continue to explore the modularity of MOFs in order to gain better insights on the structure–property relationship and subsequently construct a made-to-order MOF with the ideal gas–MOF interactions and suitable gas uptake for given gas separation/storage applications (Nugent et al., 2013[Nugent, P., Belmabkhout, Y., Burd, S. D., Cairns, A. J., Luebke, R., Forrest, K., Pham, T., Ma, S., Space, B., Wojtas, L., Eddaoudi, M. & Zaworotko, M. J. (2013). Nature, 495, 80-84.]; Belmabkhout et al., 2014[Belmabkhout, Y., Mouttaki, H., Eubank, J. F., Guillerm, V. & Eddaoudi, M. (2014). RSC Adv. 4, 63855-63859.]; Shekhah et al., 2014[Shekhah, O., Belmabkhout, Y., Chen, Z., Guillerm, V., Cairns, A., Adil, K. & Eddaoudi, M. (2014). Nat. Commun 5, 4428.]). Our study on the soc-MOF platform with the underlying square-octahedral (soc) topology (Belof et al., 2007[Belof, J. L., Stern, A. C., Eddaoudi, M. & Space, B. (2007). J. Am. Chem. Soc. 129, 15202-15210.]; Liu et al., 2007[Liu, Y., Eubank, J. F., Cairns, A. J., Eckert, J., Kravtsov, V. C., Luebke, R. & Eddaoudi, M. (2007). Angew. Chem. Int. Ed. 46, 3278-3283.]; Alezi et al., 2015[Alezi, D., Belmabkhout, Y., Suyetin, M., Bhatt, P. M., Weseliński, L. J., Solovyeva, V., Adil, K., Spanopoulos, I., Trikalitis, P. N., Emwas, A. H. & Eddaoudi, M. (2015). J. Am. Chem. Soc. 137, 13308-13318.]; Cairns et al., 2016[Cairns, A. J., Eckert, J., Wojtas, L., Thommes, M., Wallacher, D., Georgiev, P. A., Forster, P. M., Belmabkhout, Y., Ollivier, J. & Eddaoudi, M. (2016). Chem. Mater. 28, 7353-7361.]) indicated that a made-to-order MOF suitable for hydrogen storage at relatively moderate pressures has to be highly porous (high surface area) and concomitantly encompass narrow pores (< 1 nm) and a high localized charged density (polarizable field charges). Notably, a large number of studies on CH4 and H2 storage by MOFs delineated the requirement of high surface area and high heat of adsorption for gas storage (Zhou et al., 2012[Zhou, H.-C., Long, J. R. & Yaghi, O. M. (2012). Chem. Rev. 112, 673-674.]). Hydrogen interactions with metal complexes, clusters or ions, within the inorganic part of a given MOF, are dominated by electrostatic forces between the quadrupole moment of the hydrogen molecule and the inorganic complex. Specifically, such H2-MOF interactions' strengths play a major role in determining the H2 uptake characteristics and hence are the subject of considerable theoretical and experimental investigations (Rowsell & Yaghi, 2005[Rowsell, J. L. & Yaghi, O. M. (2005). Angew. Chem. Int. Ed. 44, 4670-4679.]; Collins & Zhou, 2007[Collins, D. J. & Zhou, H.-C. (2007). J. Mater. Chem. 17, 3154-3160.]; Hirscher & Panella, 2007[Hirscher, M. & Panella, B. (2007). Scr. Mater. 56, 809-812.]; Lin et al., 2007[Lin, X., Jia, J., Hubberstey, P., Schröder, M. & Champness, N. R. (2007). CrystEngComm, 9, 438-448.]; Dincă & Long, 2008[Dincă, M. & Long, J. R. (2008). Angew. Chem. Int. Ed. 47, 6766-6779.]; Nouar et al., 2008[Nouar, F., Eubank, J. F., Bousquet, T., Wojtas, L., Zaworotko, M. J. & Eddaoudi, M. (2008). J. Am. Chem. Soc. 130, 1833-1835.]; Murray et al., 2009[Murray, L. J., Dincă, M. & Long, J. R. (2009). Chem. Soc. Rev. 38, 1294-1314.]; Zhou et al., 2012[Zhou, H.-C., Long, J. R. & Yaghi, O. M. (2012). Chem. Rev. 112, 673-674.]). In particular, the weaker, dispersive interactions between H2 molecules and the organic linkers in MOFs, best represented by benzene ring moieties, have been theoretically investigated (Hübner et al., 2004[Hübner, O., Glöss, A., Fichtner, M. & Klopper, W. (2004). J. Phys. Chem. A, 108, 3019-3023.]; Bhatia & Myers, 2006[Bhatia, S. K. & Myers, A. L. (2006). Langmuir, 22, 1688-1700.]; Lochan & Head-Gordon, 2006[Lochan, R. C. & Head-Gordon, M. (2006). Phys. Chem. Chem. Phys. 8, 1357-1370.]; Düren et al., 2009[Düren, T., Bae, Y.-S. & Snurr, R. Q. (2009). Chem. Soc. Rev. 38, 1237-1247.]; Han et al., 2009[Han, S. S., Mendoza-Cortés, J. L. & Goddard, W. A. III (2009). Chem. Soc. Rev. 38, 1460-1476.]) and experimentally documented (Rosi et al., 2003[Rosi, N. L., Eckert, J., Eddaoudi, M. Vodak, D. T., Kim, J., O'Keeffe, M. & Yaghi, O. M. (2003). Science, 300, 1127-1129.]). Recent studies demonstrate that such interactions could, in principle, be enhanced through chemical modifications of the organic linkers, providing a prospective strategy for a material designer to fine-tune the organic building blocks and subsequently enhance H2 sorption characteristics of the MOF (Lochan & Head-Gordon, 2006[Lochan, R. C. & Head-Gordon, M. (2006). Phys. Chem. Chem. Phys. 8, 1357-1370.]). Nevertheless, to the best of our knowledge, no experimental synthetic studies have been published to address the potential to improve the H2 and CH4 binding affinity to the walls of MOFs through simultaneous dispersive interactions, acting additively, between the gas molecules and multiple aromatic rings placed at optimal interaction distance(s) within a specific geometry. Therefore, we opted to explore this approach separately, regardless of the degree of porosity, which could potentially pave the way for the rational design of MOF adsorbents as suitable and effective gas storage media. Computational studies revealed moderate binding affinities for the H2 molecule towards benzene and various aromatic rings (Hübner et al., 2004[Hübner, O., Glöss, A., Fichtner, M. & Klopper, W. (2004). J. Phys. Chem. A, 108, 3019-3023.]; Bhatia & Myers, 2006[Bhatia, S. K. & Myers, A. L. (2006). Langmuir, 22, 1688-1700.]; Lochan & Head-Gordon, 2006[Lochan, R. C. & Head-Gordon, M. (2006). Phys. Chem. Chem. Phys. 8, 1357-1370.]; Düren et al., 2009[Düren, T., Bae, Y.-S. & Snurr, R. Q. (2009). Chem. Soc. Rev. 38, 1237-1247.]; Han et al., 2009[Han, S. S., Mendoza-Cortés, J. L. & Goddard, W. A. III (2009). Chem. Soc. Rev. 38, 1460-1476.]). Such interactions are mostly dispersive and within the range of 3.4–4.0 kJ mol−1 for a H2 molecule interacting with the benzene ring of terephthalic acid (Lochan & Head-Gordon, 2006[Lochan, R. C. & Head-Gordon, M. (2006). Phys. Chem. Chem. Phys. 8, 1357-1370.]). The aforementioned binding enthalpy is remotely below the estimated and debated target for efficient H2 storage materials, range of 15–20 kJ mol−1 (room temperature at pressures up to 30 bar; Han et al., 2009[Han, S. S., Mendoza-Cortés, J. L. & Goddard, W. A. III (2009). Chem. Soc. Rev. 38, 1460-1476.]) or 21–32 kJ mol−1 (−20°C and pressure range 1–100 bar; Lochan & Head-Gordon, 2006[Lochan, R. C. & Head-Gordon, M. (2006). Phys. Chem. Chem. Phys. 8, 1357-1370.]). Herein, we set to investigate if such interactions could be additive and hence can lead to enhanced interactions between a H2 molecule and multiple aromatic rings in a tailored MOF adsorbent. As a test model, we envision a molecular square constructed of four benzene rings interacting simultaneously with a single H2 molecule, residing in the center of the square, as a potential model for a material with enhanced H2 binding affinity.

2. Experimental

The solvothermal reaction of Pb(NO3)2 and 4,4′-sulfonyldibenzoic acid in N,N-dimethylformamide (DMF) yields colorless crystals of 1 (Fig. 1[link]). The as-synthesized compound was characterized and formulated by single-crystal X-ray diffraction studies as [Pb2(C14H8O6S)2]·DMF (1). The phase purity of 1 was confirmed by similarities between its calculated and experimental powder X-ray diffraction patterns (PXRD, supporting information).

[Figure 1]
Figure 1
Crystal structure of 1 (top), the Pb—CO2 rod-shaped infinite SBU (middle), and the coordination mode of the carboxylic linker and the Pb(II) ion (below). Pb (green), C (gray), S (yellow), O (red), H (white).

A sample of 1 was activated for sorption studies by solvent exchange in acetonitrile, where complete removal of the DMF guest molecules was confirmed by IR spectroscopy, see the supporting information. The activated sample was found to be stable up to 400°C as confirmed by TGA studies, see the supporting information.

3. Results and discussion

The crystal structure of 1 revealed square-like channels running through the a-axis. The distinctive shape of the ditopic ligand molecule (dihedral Ph—SO2—Ph angle of 103.93°) complemented by the coordination sphere around Pb(II) (CO2—Pb—CO2 dihedral angle of 78.73°) facilitated the construction of the MOF containing square-like channels (Table 1[link]). In the crystal structure of 1, infinite CO2—Pb(II) secondary building units (SBUs) (Rosi et al., 2005[Rosi, N. L., Kim, J., Eddaoudi, M., Chen, B., O'Keeffe, M. & Yaghi, O. M. (2005). J. Am. Chem. Soc. 127, 1504-1518.]) are observed and resulting from coordination of the carboxylate linkers in the bridging bis-bidentate mode to Pb(II) ions. Each carboxylate group is coordinated to three Pb(II) ions, enabling the formation of the CO2—Pb(II) infinite coordination chains, along the a-axis. Each Pb(II) ion is coordinated to six oxygen atoms from bridging carboxylate groups (O—Pb bond distances of 2.434–2.815 Å). Additionally, each Pb(II) ion is coordinated to an oxygen atom from the nearest sulfone group (O—Pb bond distance of 2.865 Å).

Table 1
Selected geometric parameters (Å, °)

Pb1—O3i 2.434 (6) Pb1—O4i 2.441 (6)
O3—Pb1—O4i 72.0 (2) O2—S3—C6ii 108.1 (3)
O3i—Pb1—O4i 117.1 (2) O1—S3—C6ii 107.6 (4)
Symmetry codes: (i) [x, -y+{1\over 2}, z]; (ii) [x, -y-{1\over 2}, z].

The resultant connectivity of the Pb ions by the organic linkers facilitated the construction of parallelogram, square-like, shaped channels running along the a-axis, held together in the bc-plane through the sulfone–Pb coordination. Such orthogonal bridging interactions resulted in square-like, guest-accessible, channels in 1. The surface area of 1 as probed by N2 and Ar at 77 K and 87 K (supporting information) were estimated to be 165 and 169 m2 g−1, respectively. The resulting square-like channels in 1 encompass a periodic array of aromatic rings with a relatively short interplanar distance between opposing rings (8.448 Å, centroid-to-centroid). Essentially, the periodically aligned aromatic rings delimiting the pore system dictate the pore aperture size and its maximum opening to be around ∼ 4 Å (excluding the nearest van der Waals surfaces). These special structural features encountered in 1 (narrow one-dimensional channels aligned with a periodic array of aromatic rings) inspired us to explore and further investigate the potential effect(s) of the pore system (size, geometry and functionality) on the H2 interactions with the aromatic walls. Indeed, the observed H2 adsorption properties of the present MOF are remarkable and unique. Of special note are the observed H2 adsorption isotherms with a sharp steepness (type I isotherm shape particularly at 77 K), which, to the best of our knowledge, are scarce for physical adsorbents. Equally interesting is the observed steady H2 isosteric heat of adsorption (Qst) at 9 kJ mol−1 (Fig. 2[link]). In fact, the H2 adsorption isotherms for 1 demonstrate rapid saturation at early dosing stages and nearly linear behaviour for heats of adsorption throughout the entire H2 adsorption loading, two highly desirable features for H2 storage applications. The sharp step in the H2 adsorption isotherm can be translated to sorption sites saturation by H2 molecules at moderate pressures. This behaviour could be attributed to the equal distribution of H2 sorption sites with uniform binding affinities within the framework, most probably on the surfaces of aromatic rings present in 1. Interestingly, such uniform interactions were also observed for methane adsorption in 1 (Fig. 3[link]), with a relatively high and steady Qst of 25 kJ mol−1 over the entire loading range, thus confirming the interesting structural aspects of 1 for enhanced gas–solid adsorbent interactions.

[Figure 2]
Figure 2
(a) Variable-temperature H2 adsorption isotherms and (b) Qst of H2 adsorption in 1.
[Figure 3]
Figure 3
(a) Variable-temperature CH4 adsorption isotherms and (b) Qst of CH4 adsorption in 1.

Furthermore, high-pressure gas adsorption studies conducted on 1 revealed interesting adsorption behaviour of selected gases (Fig. 4[link]). The recorded uptakes for O2 and N2 in 1 were comparable. This is in contrast to commonly observed preferential N2 uptake, compared with O2, in numerous examples of zeolites (Talu et al., 1996[Talu, O., Li, J., Kumar, R., Mathias, P. M., Moyer, J. D. & Schork, J. M. (1996). Gas Sep. Purif. 10, 149-159.]; Hutson et al., 1999[Hutson, N. D., Rege, S. U. & Yang, R. T. (1999). AIChE J. 45, 724-734.]; Agha et al., 2005[Agha, R. K., De Weireld, G. & Frère, M. (2005). Adsorption 11, 179-182.]). In the case of most MOFs and zeolites, preferential N2 sorption is attributed to stronger interactions between N2 molecules and the material surface due to a larger quadrupole moment of N2. The interaction/adsorption sites in 1 are dominated by the periodic array of aromatic rings aligned in the one-dimensional channels, thus explaining the weak quadrupole interactions between the gas molecules and the framework and the subsequent comparable N2 and O2 uptakes. Similarly, the nature of the gas/framework interactions is also reflected in the observed comparable uptake of CH4 and CO2 (two distinct molecules with and without quadrupole moment) in 1 (Fig. 4[link]), emphasizing the dominance of dispersive interactions between adsorbed gas molecules and the aromatic walls of the MOF porous material. It is noteworthy that porous materials having similar uptakes for CO2 and CH4 in a wide range of pressure is uncommon behaviour in adsorption on porous materials (Nugent et al., 2013[Nugent, P., Belmabkhout, Y., Burd, S. D., Cairns, A. J., Luebke, R., Forrest, K., Pham, T., Ma, S., Space, B., Wojtas, L., Eddaoudi, M. & Zaworotko, M. J. (2013). Nature, 495, 80-84.]). Additionally, the adsorption of C2H6 and C3+ on 1 (supporting information) showed to some extent the same behaviour, uncommon for most MOF materials. The aforementioned results demonstrate that molecules with various degrees of high polarizabilities probe the surface of 1 in the same way. Specifically, the interaction potential of 1 with different molecules, having different chemical–physical properties like CO2, CH4, C2H6 and C3+, is governed mainly by dispersive (non-electrostatic) interactions.

[Figure 4]
Figure 4
High-pressure sorption isotherms for different gases at 298 K in 1.

4. Conclusions

In conclusion, we present an unprecedented experimental illustration of uptake–energetics relationships for H2 and CH4 adsorption in a novel MOF with narrow one-dimensional channels aligned with a periodic array of aromatic rings. The newly synthesized material represents a model material for pinpointing the interplay between porosity/storage capacity and the strength of host–guest interactions. This approach supports the impact of narrow channels with a periodic array of aromatic rings, controlling the access to prospective larger pores within a targeted porous material, for the effective gas adsorption. The present results pave the way to additional experimental and theoretical investigations in order to further assess the extent of additive dispersive interactions in enhancing H2 and CH4 binding affinity in gas storage materials, in general, and MOFs, in particular. Noticeably, the distinct sharp step at relatively low pressures in the H2 adsorption isotherm at 77 K and the relatively high Qst for CH4 storage was achieved despite the detrimental reduction in the MOF overall porosity. Conceivably, from the present study, the optimal combination of narrow pores/windows (< 1 nm diameter) with a suitable and uniform charge density in the pores (coordinatively unsaturated metal sites and polar functional groups) operating synergistically could play a significant role in promoting the storage of H2 and CH4 at moderate pressures and ambient temperatures.

Supporting information


Computing details top

Program(s) used to refine structure: XL (Sheldrick, 2008); molecular graphics: Olex2 (Dolomanov et al., 2009); software used to prepare material for publication: Olex2 (Dolomanov et al., 2009).

(kme002_0m) top
Crystal data top
0.5(C28H16O12Pb2S2)Z = 4
Mr = 511.45F(000) = 952
Orthorhombic, PnmaDx = 2.291 Mg m3
a = 5.8740 (4) ÅCu Kα radiation, λ = 1.54178 Å
b = 13.0820 (7) ŵ = 23.70 mm1
c = 19.2970 (12) ÅT = 100 K
V = 1482.85 (16) Å30.1 × 0.05 × 0.05 mm
Data collection top
3462 measured reflectionsθmax = 64.9°, θmin = 4.6°
1232 independent reflectionsh = 63
1184 reflections with I > 2σ(I)k = 1415
Rint = 0.036l = 1921
Refinement top
Refinement on F20 restraints
Least-squares matrix: fullHydrogen site location: inferred from neighbouring sites
R[F2 > 2σ(F2)] = 0.039H-atom parameters constrained
wR(F2) = 0.103 w = 1/[σ2(Fo2) + (0.0528P)2 + 43.9503P]
where P = (Fo2 + 2Fc2)/3
S = 0.89(Δ/σ)max = 0.001
1232 reflectionsΔρmax = 2.62 e Å3
106 parametersΔρmin = 1.36 e Å3
Special details top

Geometry. All esds (except the esd in the dihedral angle between two l.s. planes) are estimated using the full covariance matrix. The cell esds are taken into account individually in the estimation of esds in distances, angles and torsion angles; correlations between esds in cell parameters are only used when they are defined by crystal symmetry. An approximate (isotropic) treatment of cell esds is used for estimating esds involving l.s. planes.

Fractional atomic coordinates and isotropic or equivalent isotropic displacement parameters (Å2) top
xyzUiso*/Ueq
Pb11.26391 (7)0.2500000.68779 (2)0.0151 (2)
S30.3739 (5)0.2500000.50639 (15)0.0167 (6)
C30.6732 (15)0.0162 (6)0.6252 (4)0.0172 (17)
C20.7896 (15)0.0231 (7)0.5685 (5)0.0217 (19)
H20.9317440.0052810.5550270.026*
C40.4628 (16)0.0233 (6)0.6449 (5)0.023 (2)
H40.3841550.0048430.6834540.028*
C10.6964 (16)0.1049 (6)0.5312 (5)0.0200 (18)
H10.7756550.1333960.4928360.024*
C60.4864 (14)0.1437 (6)0.5512 (4)0.0177 (18)
C70.7806 (14)0.1025 (7)0.6662 (5)0.0182 (19)
C50.3673 (15)0.1051 (6)0.6075 (5)0.0213 (18)
H50.2245110.1333200.6205980.026*
O10.4718 (15)0.2500000.4370 (4)0.0238 (19)
O20.1290 (16)0.2500000.5157 (5)0.025 (2)
O30.9737 (10)0.1311 (4)0.6476 (3)0.0228 (14)
O41.1630 (11)0.3617 (4)0.7843 (3)0.0234 (13)
Atomic displacement parameters (Å2) top
U11U22U33U12U13U23
Pb10.0169 (3)0.0118 (3)0.0167 (3)0.0000.00047 (17)0.000
S30.0251 (15)0.0097 (12)0.0152 (14)0.0000.0029 (12)0.000
C30.022 (4)0.010 (4)0.019 (5)0.002 (3)0.004 (4)0.003 (3)
C20.023 (4)0.015 (4)0.027 (5)0.000 (4)0.002 (4)0.000 (4)
C40.032 (5)0.014 (4)0.024 (5)0.002 (4)0.001 (4)0.004 (4)
C10.027 (4)0.012 (4)0.021 (5)0.002 (3)0.005 (4)0.001 (4)
C60.025 (4)0.006 (4)0.022 (5)0.001 (3)0.004 (4)0.003 (3)
C70.023 (5)0.011 (4)0.020 (5)0.007 (3)0.010 (4)0.001 (4)
C50.024 (5)0.014 (4)0.026 (5)0.005 (4)0.000 (4)0.000 (4)
O10.037 (5)0.017 (4)0.017 (4)0.0000.002 (4)0.000
O20.028 (5)0.015 (4)0.032 (5)0.0000.001 (4)0.000
O30.027 (3)0.017 (3)0.025 (3)0.003 (3)0.003 (3)0.001 (3)
O40.032 (3)0.019 (3)0.019 (3)0.003 (3)0.003 (3)0.007 (3)
Geometric parameters (Å, º) top
Pb1—O3i2.435 (6)C3—C41.392 (13)
Pb1—O32.435 (6)C3—C71.516 (12)
Pb1—O4i2.441 (6)C2—C11.401 (13)
Pb1—O42.441 (6)C4—C51.408 (12)
S3—C61.765 (8)C1—C61.389 (12)
S3—C6ii1.765 (8)C6—C51.389 (12)
S3—O11.458 (9)C7—O31.247 (11)
S3—O21.450 (10)C7—O4iii1.268 (12)
C3—C21.390 (13)
O3i—Pb1—O379.4 (3)C4—C3—C7120.2 (8)
O3—Pb1—O4117.1 (2)C3—C2—C1119.7 (8)
O3i—Pb1—O472.0 (2)C3—C4—C5119.8 (8)
O3i—Pb1—O4i117.1 (2)C6—C1—C2118.9 (8)
O3—Pb1—O4i72.0 (2)C1—C6—S3119.0 (6)
O4i—Pb1—O473.6 (3)C1—C6—C5122.2 (8)
C6ii—S3—C6103.9 (5)C5—C6—S3118.8 (6)
O1—S3—C6107.6 (4)O3—C7—C3116.8 (8)
O1—S3—C6ii107.6 (4)O3—C7—O4iii127.0 (8)
O2—S3—C6ii108.1 (4)O4iii—C7—C3116.2 (8)
O2—S3—C6108.1 (4)C6—C5—C4118.5 (8)
O2—S3—O1120.4 (5)C7—O3—Pb1137.5 (6)
C2—C3—C4121.0 (8)C7iv—O4—Pb1131.6 (6)
C2—C3—C7118.8 (8)
Symmetry codes: (i) x, y+1/2, z; (ii) x, y1/2, z; (iii) x1/2, y+1/2, z+3/2; (iv) x+1/2, y+1/2, z+3/2.
 

Acknowledgements

We acknowledge funding from King Abdullah University of Science and Technology (KAUST).

References

First citationAgha, R. K., De Weireld, G. & Frère, M. (2005). Adsorption 11, 179–182.  Web of Science CrossRef Google Scholar
First citationAlezi, D., Belmabkhout, Y., Suyetin, M., Bhatt, P. M., Weseliński, L. J., Solovyeva, V., Adil, K., Spanopoulos, I., Trikalitis, P. N., Emwas, A. H. & Eddaoudi, M. (2015). J. Am. Chem. Soc. 137, 13308–13318.  Web of Science CSD CrossRef CAS PubMed Google Scholar
First citationAlkordi, M. H., Liu, Y., Larsen, R. W., Eubank, J. F. & Eddaoudi, M. (2008). J. Am. Chem. Soc. 130, 12639–12641.  Web of Science CrossRef PubMed CAS Google Scholar
First citationBanerjee, R., Phan, A., Wang, B., Knobler, C., Furukawa, H., O'Keeffe, M. & Yaghi, O. M. (2008). Science, 319, 939–943.  Web of Science CSD CrossRef PubMed CAS Google Scholar
First citationBelmabkhout, Y., Mouttaki, H., Eubank, J. F., Guillerm, V. & Eddaoudi, M. (2014). RSC Adv. 4, 63855–63859.  Web of Science CrossRef CAS Google Scholar
First citationBelof, J. L., Stern, A. C., Eddaoudi, M. & Space, B. (2007). J. Am. Chem. Soc. 129, 15202–15210.  Web of Science CrossRef PubMed CAS Google Scholar
First citationBhatia, S. K. & Myers, A. L. (2006). Langmuir, 22, 1688–1700.  Web of Science CrossRef PubMed CAS Google Scholar
First citationCairns, A. J., Eckert, J., Wojtas, L., Thommes, M., Wallacher, D., Georgiev, P. A., Forster, P. M., Belmabkhout, Y., Ollivier, J. & Eddaoudi, M. (2016). Chem. Mater. 28, 7353–7361.  Web of Science CSD CrossRef CAS Google Scholar
First citationChen, B., Zhao, X., Putkham, A., Hong, K., Lobkovsky, E. B., Hurtado, E. J., Fletcher, A. J. & Thomas, K. M. (2008). J. Am. Chem. Soc. 130, 6411–6423.  Web of Science CSD CrossRef PubMed CAS Google Scholar
First citationCho, S. H., Gadzikwa, T., Afshari, M., Nguyen, S. T. & Hupp, J. T. (2007). Eur. J. Inorg. Chem. 2007, 4863–4867.  Web of Science CrossRef Google Scholar
First citationCollins, D. J. & Zhou, H.-C. (2007). J. Mater. Chem. 17, 3154–3160.  Web of Science CrossRef CAS Google Scholar
First citationDincă, M., Han, W. S., Liu, Y., Dailly, A., Brown, C. M. & Long, J. R. (2007). Angew. Chem. Int. Ed. 46, 1419–1422.  Google Scholar
First citationDincă, M. & Long, J. R. (2008). Angew. Chem. Int. Ed. 47, 6766–6779.  Google Scholar
First citationDüren, T., Bae, Y.-S. & Snurr, R. Q. (2009). Chem. Soc. Rev. 38, 1237–1247.  Web of Science PubMed Google Scholar
First citationEddaoudi, M., Moler, D. B., Li, H., Chen, B., Reineke, T. M., O'Keeffe, M. & Yaghi, O. M. (2001). Acc. Chem. Res. 34, 319–330.  Web of Science CrossRef PubMed CAS Google Scholar
First citationFérey, G., Mellot-Draznieks, C., Serre, C. & Millange, F. (2005). Acc. Chem. Res. 38, 217–225.  Web of Science PubMed Google Scholar
First citationHan, S. S., Mendoza-Cortés, J. L. & Goddard, W. A. III (2009). Chem. Soc. Rev. 38, 1460–1476.  Web of Science CrossRef PubMed CAS Google Scholar
First citationHayashi, H., Côté, A. P., Furukawa, H., O'Keeffe, M. & Yaghi, O. M. (2007). Nat. Mater. 6, 501–506.  Web of Science CSD CrossRef PubMed CAS Google Scholar
First citationHirscher, M. & Panella, B. (2007). Scr. Mater. 56, 809–812.  Web of Science CrossRef CAS Google Scholar
First citationHorcajada, P., Serre, C., Vallet-Regí, M., Sebban, M., Taulelle, F. & Férey, G. (2006). Angew. Chem. 118, 6120–6124.  CrossRef Google Scholar
First citationHübner, O., Glöss, A., Fichtner, M. & Klopper, W. (2004). J. Phys. Chem. A, 108, 3019–3023.  Google Scholar
First citationHutson, N. D., Rege, S. U. & Yang, R. T. (1999). AIChE J. 45, 724–734.  Web of Science CrossRef CAS Google Scholar
First citationKitagawa, S., Kitaura, R. & Noro, S. (2004). Angew. Chem. Int. Ed. 43, 2334–2375.  Web of Science CrossRef CAS Google Scholar
First citationLin, X., Jia, J., Hubberstey, P., Schröder, M. & Champness, N. R. (2007). CrystEngComm, 9, 438–448.  Web of Science CrossRef CAS Google Scholar
First citationLiu, Y., Eubank, J. F., Cairns, A. J., Eckert, J., Kravtsov, V. C., Luebke, R. & Eddaoudi, M. (2007). Angew. Chem. Int. Ed. 46, 3278–3283.  Web of Science CSD CrossRef CAS Google Scholar
First citationLlewellyn, P. L., Bourrelly, S., Serre, C., Vimont, A., Daturi, M., Hamon, L., De Weireld, G., Chang, J.-S., Hong, D.-Y., Kyu Hwang, Y., Hwa Jhung, S. & Férey, G. (2008). Langmuir, 24, 7245–7250.  Web of Science CrossRef PubMed CAS Google Scholar
First citationLochan, R. C. & Head-Gordon, M. (2006). Phys. Chem. Chem. Phys. 8, 1357–1370.  Web of Science CrossRef PubMed CAS Google Scholar
First citationMoulton, B. & Zaworotko, M. J. (2001). Chem. Rev. 101, 1629–1658.  Web of Science CrossRef PubMed CAS Google Scholar
First citationMurray, L. J., Dincă, M. & Long, J. R. (2009). Chem. Soc. Rev. 38, 1294–1314.  Web of Science CrossRef PubMed CAS Google Scholar
First citationNouar, F., Eubank, J. F., Bousquet, T., Wojtas, L., Zaworotko, M. J. & Eddaoudi, M. (2008). J. Am. Chem. Soc. 130, 1833–1835.  Web of Science CSD CrossRef PubMed CAS Google Scholar
First citationNugent, P., Belmabkhout, Y., Burd, S. D., Cairns, A. J., Luebke, R., Forrest, K., Pham, T., Ma, S., Space, B., Wojtas, L., Eddaoudi, M. & Zaworotko, M. J. (2013). Nature, 495, 80–84.  Web of Science CSD CrossRef CAS PubMed Google Scholar
First citationRadha Kishan, M., Tian, J. K., Thallapally, P. K., Fernandez, C. A., Dalgarno, S. J., Warren, J. E., McGrail, B. P. & Atwood, J. L. (2010). Chem. Commun. 46, 538–540.  Web of Science CSD CrossRef Google Scholar
First citationRosi, N. L., Eckert, J., Eddaoudi, M. Vodak, D. T., Kim, J., O'Keeffe, M. & Yaghi, O. M. (2003). Science, 300, 1127–1129.  Web of Science CSD CrossRef PubMed CAS Google Scholar
First citationRosi, N. L., Kim, J., Eddaoudi, M., Chen, B., O'Keeffe, M. & Yaghi, O. M. (2005). J. Am. Chem. Soc. 127, 1504–1518.  Web of Science CSD CrossRef PubMed CAS Google Scholar
First citationRowsell, J. L. & Yaghi, O. M. (2005). Angew. Chem. Int. Ed. 44, 4670–4679.  Web of Science CrossRef CAS Google Scholar
First citationRowsell, J. L. & Yaghi, O. M. (2006). J. Am. Chem. Soc. 128, 1304–1315.  Web of Science CSD CrossRef PubMed CAS Google Scholar
First citationShekhah, O., Belmabkhout, Y., Chen, Z., Guillerm, V., Cairns, A., Adil, K. & Eddaoudi, M. (2014). Nat. Commun 5, 4428.  Web of Science CSD CrossRef PubMed Google Scholar
First citationTalu, O., Li, J., Kumar, R., Mathias, P. M., Moyer, J. D. & Schork, J. M. (1996). Gas Sep. Purif. 10, 149–159.  CrossRef CAS Web of Science Google Scholar
First citationZhou, H.-C., Long, J. R. & Yaghi, O. M. (2012). Chem. Rev. 112, 673–674.  Web of Science CrossRef CAS PubMed Google Scholar

This is an open-access article distributed under the terms of the Creative Commons Attribution (CC-BY) Licence, which permits unrestricted use, distribution, and reproduction in any medium, provided the original authors and source are cited.

IUCrJ
Volume 4| Part 2| March 2017| Pages 131-135
ISSN: 2052-2525