You are on page 1of 11

RSC Advances

PAPER
View Article Online
View Journal

Published on 18 April 2013. Downloaded by FAC DE QUIMICA on 05/06/2013 20:57:08.

Cite this: DOI: 10.1039/c3ra40726h

An alkaline one-pot metathesis reaction to give a [Cu3(BTC)2] MOF at r.t., with free Cu coordination sites and enhanced hydrogen uptake properties3{
pez Nu pezn Sandra Loera-Serna,a Lourdes Lo ez,a Jorge Flores,a Roxana Lo b b n* Simeon and Hiram I. Beltra
A one-pot metathesis reaction between trimesic acid sodium salt and Cu(NO3)2 to yield [Cu3(BTC)2] (HKUST-1) in moderate to good yields, even at room temperature (r.t.) is presented, herein, for the first time. To determine if the basic nature of the metathesis reaction afforded any new physicochemical properties to HKUST-1, three variables were monitored: a) the energy source either from i) ultrasound (US), ii) stirring at room temperature (SRT) or iii) solvothermal (ST) treatment, b) the reaction solvent and c) the reaction time. The metathesis reaction provided DMF?H2O, EtOH?H2O or MeOH?H2O solvates. The compounds were thoroughly characterized by FTIR, XRD, nitrogen adsorption BET surface areas and TGA/ DSC data, which indicated the formation of [Cu3(BTC)2] MOFs in eight out of ten synthetic entries. Lowpressure hydrogen uptake experiments revealed values as high as 2.44 wt% for the [Cu3(BTC)2] EtOH?H2O solvates. The [Cu3(BTC)2] MeOH?H2O and [Cu3(BTC)2] EtOH?H2O solvates possess high surface areas, the latter being the metathesis prepared MOF with the least H2OACu coordination according to TGA, DSC and a thermal desorption gas chromatographymass spectrometry (GC-MS) coupled technique. This resulted in a material with less chemisorbed H2O and better H2 adsorption capability.

Received 11th February 2013, Accepted 15th April 2013 DOI: 10.1039/c3ra40726h www.rsc.org/advances

Introduction
Metalorganic frameworks (MOFs) and their study have become important research and development fields not only in chemistry, but also in general science and technology.13 Therefore, most of the important related work is aimed at finding compounds designed to possess very large pores4,5 and high surface areas in order to load these materials with atoms, molecules or even biomolecules.3,57 Due to these loading possibilities, wide applications of MOFs have emerged in different fields, such as in catalysis,712 drug delivery,13 guest adsorption (molecular recognition),14 optical applications,15,16 sensor technologies and 8 gas storage,7,1620 among others.1,3,7,9,14,21

sicas, DCBI, Universidad Auto noma Metropolitana, Departamento de Ciencias Ba Unidad Azcapotzalco, Av. San Pablo No. 180, Reynosa Tamaulipas, Azcapotzalco, 02200, D.F., Mexico b noma Metropolitana, Departamento de Ciencias Naturales, DCNI, Universidad Auto lvaro Unidad Cuajimalpa, Pedro Antonio de los Santos 84, San Miguel Chapultepec, A n, 11850, D.F., Mexico. E-mail: hbeltran@correo.cua.uam.mx; Obrego hbeltran75@gmail.com 3 Electronic supplementary information (ESI) available: XRD (Fig. S1), FTIR (Fig. S2) and TGA (Fig. S3) analyses. See DOI: 10.1039/c3ra40726h n on the occasion of his { This paper is dedicated to Professor Norberto Farfa 58th birthday. These authors contributed equally to the results presented in this manuscript.

Researchers have rationalised the hierarchically obtained networks in order to augment the pore size in MOFs, by: i) the use of extensions of in the linkers, which often retain the symmetry of the system;5,7,22,23 or ii) replacement of the metallic knots, which varies the bond angles,7,23 and thus changes the symmetry and the reticular density of the system.24 The aim is to always try and circumvent interpenetration.7,25 Synthetic approaches and the rationale behind them are very important for enhancing or modulating the physicochemical properties of the prepared materials.11,17,20,26,27 Related references cover various methodologies involving high temperature or Soxhlet separation in order to activate and remove20,28 physically or chemically confined solvent molecules. Methods also include the use of acidic reagents29,30 or combinations to get neutral media3032 and even surfactants to control the crystallite size,17 as well as variation of energy source.2729,3335 In particular, the aim of this contribution is to use a low temperature,11,31 alkaline assisted one-pot method to obtain [Cu3(BTC)2] (some synonyms of which are: [Cu3(C9H3O6)2], [Cu3(TMA)2], HKUST-1 or MOF-199, where BTC = 1,3,5-benzenetricarboxylate and also TMA = trimesic acid carboxylate) due to its pioneering preparation,29 investigate the possibility of improving its physicochemical properties,11,2729,3336 e.g. through careful balancing of the occupancy of X A Cu sites,11,26 and to its potential applica-

This journal is The Royal Society of Chemistry 2013

RSC Adv.

View Article Online

Paper tions.1,3,721 Some of the described variations in the preparation method of [Cu3(BTC)2] produced important differences in the surface area, which mainly involved variations of the energy source or nature of the solvent(s) and a small collection of them is presented in Table 1. As seen, the first three procedures shown in this table led to some of the more narrow specific surface area values of the material. By comparison, the wider surface area values obtained have been found for the latter six procedures, reaching BET/Langmuir values as high as 17131820/2260 m2 g21, most of them using post synthesis activation procedures.33,3739 Moreover, some of the highest values for the hydrogen uptake for [Cu3(BTC)2] MOFs have been reported as between 2.182.67 wt% at low pressure,40,41 the higher adsorption always correlating with the availability of free Cu coordination sites.38,42 The latter is clear evidence that the synthetic procedure is important for optimizing desired physicochemical properties, such as specific surface area, pore size and volume, but more importantly the careful balance of the occupancy of XACu sites,11,26 thus leading to an enhancement of its potential applications, until final technological goals such as H2 uptake can be achieved. The presented low temperature, alkaline assisted one-pot strategy employs simple, but well-chosen reagents to obtain a metalorganic framework through a metathesis reaction.4346 The efficient displacing of the reaction course to products in metathesis methodologies has been acquired by two important means: i) phase change (precipitation, neutralization, gasification or combinations of the three) of the products; and ii) generation of products with ionic energy contributions, such as in the formation of inorganic salts.45 If both i) and ii) metathesis criteria are applied, low temperature, fast, easy and thus efficient chemical transformations are obtained, that could be applicable to a myriad of different chemical procedures.45 In this contribution, the fine analysis of experimental evidence gave us decisive information indicating that this new alkaline synthetic procedure modifies the physicochemical properties of the resulting [Cu3(BTC)2]

RSC Advances material. Therefore, now it is feasible to prepare this MOF with modulated amounts of physisorbed (molecules placed into the channels) or chemisorbed (molecules occupying CurX coordination sites) water molecules and with high surface areas straight from the reaction vessel without any post-synthetic steps as will be shown in the forthcoming sections.

Published on 18 April 2013. Downloaded by FAC DE QUIMICA on 05/06/2013 20:57:08.

Experimental section
General information Available reagents and solvents were used as received. Trimesic acid (TMA or BTC), dimethylformamide (DMF) and methanol were obtained from Sigma-Aldrich, absolute ethanol from JT-Baker, and copperII nitrate from Fluka. Synthetic procedure for the formation of [Cu3(BTC)2] The one-pot preparation of [Cu3(BTC)2] MOF was carried out as follows, using six equivalents of copperII nitrate and four equivalents of trimesic acid. 0.5 g (2.38 mmol) of trimesic acid and 0.6 g (7.14 mmol) of sodium bicarbonate were poured into a reaction flask containing 150 mL of distilled water and the mixture was stirred until complete dissolution of the reagents. This step was used in order to obtain triply activated carboxylate from trimesic acid which more easily reacts with copper(II) nitrate and hence the reaction media in this contribution was basic in nature unlike in other procedures where the reaction medium was acidic or neutral, please see references used in Table 1 for further details. Then, a solution containing 0.86 g, (3.57 mmol) of Cu(NO3)2?3H2O and 40 mL of solvent, was added a dropwise. The [Cu3(BTC)2] product started to precipitate from the solution as a blue solid, which was independent of the energy source applied to the reaction flask. The [Cu3(BTC)2] or the resulting compound was isolated by centrifugation and the blue precipitate was washed twice with 5 mL of the particular solvent mixture employed for the synthesis. Yields of [Cu3(BTC)2] or resulting compound varied from 52.5 to 72.6%. Dry yields were obtained by weight loss

Table 1 Variations of the method for preparation/activation of [Cu3(BTC)2] MOFs and the surface areas obtained.ac

Method of preparation Solvothermal29 Electrochemical27 Mechanochemical28 Microwave34 Solvothermal35 Electrochemical33d Mechanochemical37d Solvothermal11,26 Solvothermal38,39

Solvent(s) H2O?EtOH MeOH Solventless H2O?EtOH H2O?EtOH MeOH Solventless H2O?EtOH H2O?EtOH?DMF

TSynth/TActivation [uC] 180/n.d. 40/220 r.t./200 140170/150 150/110 r.t./120 r.t./250 r.t./n.a.e r.t./r.t.f 95/95g 85/170h

Surface area [m2 g21] 692/917.6e 897 692 1656 1500 1649 1820 758 1713 1310 2260b

a Results are for BET surface area values. b Langmuir surface area values. c TSynth/Activation, synthesis temperature and activation temperature; r.t. = room temperature, mainly for mecanochemical preparation. n.d. = not determined. d Two experimental conditions were described with different surface area values for each. e No activation procedure was carried out. f The sample was washed with a EtOH/H2O solvent mixture to displace acetic acid molecules from the pores. g The sample was activated at 95 uC but at a reduced pressure of 1 6 1024 mbar. h The sample was washed with CHCl3 solvent prior to thermal activation.

RSC Adv.

This journal is The Royal Society of Chemistry 2013

View Article Online

RSC Advances
Table 2 Preparation conditions and selected results obtained for each synthesis of [Cu3(BTC)2] and related materials

Paper

Entry 1 2 3 4 5 6 7 8 9 10

Energy source (solvent) US (DMF/H2O) US (H2O) US (EtOH/H2O) US (MeOH/H2O) SRT (MeOH/H2O) SRT (MeOH/H2O) SRT (EtOH/H2O) SRT (EtOH/H2O) ST (MeOH/H2O) SRT + ST (MeOH/H2O)

Reaction time 20 min 20 min 20 min 20 min 5 min 12 h 5 min 12 h 12 h 12 + 12 h

Yielda (%) 72.6 57.9 68.6 56.2 56.8 64.1 52.5 69.1 52.5 57.7

SBET (m2 g21)


b

XRD 1.27
b

I[331]

/I[420]

H2 uptake (wt%)
b b

20 891 485 570 1771 520 1732


b

1.00 0.85 1.78 1.08 1.12 1.30


b

1.41 0.68 1.00 1.50 0.82 2.44


b

Published on 18 April 2013. Downloaded by FAC DE QUIMICA on 05/06/2013 20:57:08.

799

1.0

1.15

a Determined from gravimetric analysis of activated and dried samples at 275 uC to avoid solvent occlusion.25 b (not determined). US denotes ultrasound, SRT denotes stirred at room temperature and ST denotes solvothermal conditions. The XRD I[331]/I[420] ratio is a measure of the hydration of HKUST-1.11

from wet as synthesized samples and the differential gravimetric determination was developed after activation of samples at ca. 240 uC, to avoid decomposition as clearly stated in the original work of Chui et al.29 Results were very near to those obtained from thermogravimetric analysis. The designed synthetic alternatives depending on the experimental conditions are depicted as entries in Table 2 and their variations are explained as follows. Energy source variations (SRT) Stirred at room temperature; the synthetic mixture was stirred at room temperature for the needed time, in accordance with Table 2. (US) Ultrasound; the synthetic mixture was placed under ultrasonic irradiation during the needed time (see Table 2) at a frequency of 750 KHz (40% power). (ST) Solvothermal; the synthetic mixture was poured into a Teflon container, then placed into an autoclave. The reaction was performed under autogenous pressure for 12h at 100 uC. At the end of the reaction time, the autoclave was cooled down to room temperature. Fourier transform infrared (FT-IR) spectra The FT-IR spectra were recorded between the range of 4000 and 650 cm21, with a Bruker Tensor-27 FT-IR spectrophotometer, using KBr mixtures of each of the isolated samples and an ATR technique. Low pressure N2 and H2 adsorption measurements All adsorption measurements were conducted using a BELSORP-max (BELL Japan Inc.) system at 2196 uC. Samples were degassed under dynamic conditions (extra-dry air flow) over 24 h at 100 uC prior to N2 or H2 adsorption measurements. BET specific surface areas were calculated from the N2 adsorption isotherms. Powder X-ray diffraction (XRD) A powder diffractometer (Philips XPERT PRO) coupled to a copper anode X-ray tube was used to determine that [Cu3(BTC)2] was obtained and in due course its crystalline characteristics. The Ka1 radiation (45 KV, 40 mA, l = 1.5406 )

was selected with a diffracted beam monochromator, a scan speed of 0.01 s per step and a time step of 0.9 s. Thermogravimetric analysis (TGA) The thermogravimetric degradation of each sample was followed with a TA TGA Q500 (TA Instruments, USA) instrument using a heating rate of 5 uC min21 and a thermal range defined from 25 to 500 uC performed under a N2 atmosphere. Differential scanning calorimetry (DSC) The thermograms of each sample were acquired with a DSC 4000 Perkin Elmer calorimeter, using ca. 34 mg per sample per run, employing a scanning rate of 10 uC min21 with a thermal range defined from 0 to 250 uC. All DSC manipulations, determination of temperatures and enthalpies from thermograms were developed within the Pyris Software ver. 10.1 from Perkin Elmer, Inc. Analyses were run in duplicate and enthalpies are expressed in Joules per gram of the employed sample. Thermal programmed desorption or decomposition, gas chromatography-mass spectrometry (TPD-GC-MS) TPD-GC-MS experiments were carried out to follow the desorption/decomposition of selected HKUST-1 materials. Fresh solids (50 mg) were placed in a quartz reactor and heated from room temperature to 250 uC (10 uC min21). Helium (50 mL min21) was used as a carrier and the composition of the outlet gas was monitored using a quadrupole mass spectrometer (HPR-20 QIC Gas Analysis System, Hiden). Only selected masses were analyzed, in this case 18 (H2O). Entry 1. US energy source, 20 min reaction time, formation of [Cu3(BTC)2] DMF?H2O solvate in 72.6 (dry basis) % yield. IR (n, cm21, ATR): 38002684 (OH), 3090, 2980 (CH), 1703, 1650, 1589, 1565, 1450, 1422, 1379 (COOCu2), 1112, 1041 (C O), 941, 764, 731 (CCO2). XRD (u, 2h (count number)[hkl]): 7.2 (365)[200], 9.9 (470)[220], 12.1 (1117)[222], 13.9 (335)[400], 15.2 (122)[331] 15.4 (96)[420], 16.8 (130)[422], 17.9 (191)[511], 19.6 (304)[440], 20.7 (183)[600], 21.9 (148)[620], 23.9 (100)[444], 24.6 (148)[711], 26.5 (235)[731], 29.8 (183)[751], 35.7 (161)[773], 39.6 (165)[1131], 42.0 (148)[1151], 47.5 (130)[995]. TGA (% wt loss,

This journal is The Royal Society of Chemistry 2013

RSC Adv.

View Article Online

Paper Tinterval [uC]): 10, 2575; 7, 75300; 30, 300350. DSC (Tm [uC], DH [J g21]): 86.5, 553.7; 204.9, 13.2. Entry 2. US energy source, 20 min reaction time, formation of [CuBTC?3H2O] coordination polymer47 in 57.9 (dry basis) % yield. IR (n, cm21, ATR): 38002600 (OH), 3552, 3447 (OH), 1636, 1622, 1579, 1507, 1441, 1388, 1365 (COOH, COOCu), 1093, 1031 (CO), 936, 807, 755, 726 (CCO2). XRD (u, 2h (count number)[hkl]): 8.1 (948)[010], 9.4 (701)[011], 12.3 (515)[012], 16.1 (221)[020], 16.5 (251)[021], 18.5 (281)[004], 24.5 (284)[101 ], 26.6 (209)[111 ], 26.3 (174)[031], 27.1 (315)[103 ], 28.6 (231)[112], 30.4 (268)[113 ], 31.9 (212)[121], 35.8 (160)[105 ], 37.0 (164)[105], 42.5 (235)[044], 48.6 (159)[019]. BET surface area (m2 g21): 20. TGA (% wt loss, Tinterval [uC]): 8% 2560, 2% 60100, 8% 100150, 0% 150275, 33% 275350. DSC (Tm [uC], DH [J g21]): 88.4, 26.4; 157.3, 0.2; 205.6, 10.9. Entry 3. US energy source, 20 min reaction time, formation of [Cu3(BTC)2] EtOH?H2O solvate in 68.6 (dry) % yield. IR (n, cm21, ATR): 38202695 (OH), 3580 (OH), 3109, 2976 (CH), 1703, 1645, 1617, 1564, 1449, 1373 (COOCu2), 1277, 1114, 1042 (CO), 937, 760, 727 (CCO2). XRD (u, 2h (count number)[hkl]): 7.0 (147)[200], 9.8 (300)[220], 11.9 (765)[222], 13.7 (312)[400], 15.1 (141)[331], 15.3 (141)[420], 16.7 (165)[422], 17.7 (259)[511], 19.3 (347)[440], 20.5 (176)[600], 22.6 (147)[620], 23.6 (129)[444], 24.4 (182)[711], 26.2 (241)[731], 29.7 (212)[751], 35.5 (171)[773], 39.5 (188)[1131], 41.7 (141)[1151], 47.5 (141)[995]. BET surface area (m2 g21): 891. TGA (% wt loss, Tinterval [uC]): 12, 2575; 4, 75275; 32, 275330. Hydrogen uptake (wt%): 1.41. DSC (Tm [uC], DH [J g21]): 88.3, 623.9; 180.4, 11.7. Entry 4. US energy source, 20 min reaction time, formation of [Cu3(BTC)2] MeOH?H2O solvate in 56.2 (dry basis) % yield. IR (n, cm21, ATR): 38042556 (OH), 3569 (OH), 3115, 2953 (CH), 1706, 1644, 1615, 1568, 1443, 1377 (COOCu2), 1266, 1185, 1118, 1094 (CO), 937, 755, 731 (CCO2). XRD (u, 2h (count number)[hkl]): 7.0 (290)[200], 9.8 (500)[220], 11.9 (1280)[222], 13.7 (450)[400], 14.9 (169)[331], 15.3 (200)[420], 16.7 (230)[422], 17.8 (370)[511], 19.3 (440)[440], 20.5 (210)[600], 21.6 (170)[620], 23.7 (170)[444], 2424.4 (290)[711], 26.2 (310)[731], 29.6 (250)[751], 35.5 (180)[773], 39.5 (190)[1131], 41.8 (150)[1151], 47.5 (150)[995]. BET surface area (m2 g21): 485. TGA (% wt loss, Tinterval [uC]): 13, 2565; 7, 65150; 8, 150280; 32, 280330. Hydrogen uptake (wt%): 0.68. DSC (Tm [uC], DH [J g21]): 83.3, 568.3; 184.2, 8.9. Entry 5. SRT energy source, 5 min. reaction time, formation of [Cu3(BTC)2] MeOH?H2O solvate in 56.8 (dry basis) % yield. IR (n, cm21, ATR): 38002618 (OH), 3590 (OH), 3109, 2980 (CH), 1703, 1646, 1617, 1560, 1450, 1374 (COOCu2), 1274, 1112, 1045 (CO), 935, 759, 731 (CCO2). XRD (u, 2h (count number)[hkl]): 6.2 (161)[111], 7.1 (285)[200], 9.9 (489)[220], 12.0 (927)[222], 13.8 (336)[400], 14.9 (190)[331], 15.2 (107)[420], 16.8 (175)[422], 17.9 (292)[511], 19.5 (307)[440], 20.6 (161)[600], 22.4 (117)[620], 23.8 (139)[444], 24.6 (161)[711], 26.3 (219)[731], 29.7 (204)[751], 35.7 (175)[773], 39.6 (190)[1131], 41.9 (161)[1151], 47.5 (131)[995]. BET surface area (m2 g21): 570. TGA (% wt loss, Tinterval [uC]): 11, 2580; 5, 80280; 30, 280335. Hydrogen uptake (wt%): 1.00. DSC (Tm [uC], DH [J g21]): 87.3, 500.6; 185.7, 5.8. Entry 6. SRT energy source, 12 h reaction time, formation of [Cu3(BTC)2] MeOH?H2O solvate in 64.1 (dry basis) % yield. IR (n, cm21, ATR): 38182702 (OH), 3575 (OH), 3112, 2983 (C

RSC Advances H), 1705, 1649, 1616, 1589, 1570, 1545, 1451, 1376 (COOCu2), 1276, 1192, 1112, 1042 (CO), 940, 762, 730 (CCO2). XRD (u, 2h (count number)[hkl]): 7.1 (92)[200], 9.9 (402)[220], 12.1 (652)[222], 13.9 (392)[400], 15.1 (137)[331], 15.5 (127)[420], 16.9 (204)[422], 17.9 (219)[511], 19.5 (351)[440], 20.7 (148)[600], 21.8 (117)[620], 23.8 (127)[444], 24.6 (158)[711], 26.4 (219)[731], 29.9 (290)[751], 35.7 (163)[773], 39.6 (153)[1131], 41.9 (122)[1151], 47.5 (117)[995]. BET surface area (m2 g21): 1771. TGA (% wt loss, Tinterval [uC]): 5, 2555; 5, 5595; 10, 95140; 3, 140275; 30, 275350. Hydrogen uptake (wt%): 1.50. DSC (Tm [uC], DH [J g21]): 74.6, 62.5; 138.1, 55.8; 175.4, 60.8. Thermal desorption GC-MS (% m/z=18 uma, Tinterval [uC]): physisorbed, 18.7, 5095; chemisorbed, 81.3, 95 190. Entry 7. SRT energy source, 5 min reaction time, formation of [Cu3(BTC)2] EtOH?H2O solvate and [CuBTC?3H2O] coordination polymer in 52.5 (dry basis) % yield (3 : 1 crystalline mixture determined in XRD). IR (n, cm21, ATR): 38052650 (O H), 3569, 3496 (OH), 3073, 2950 (CH), 1707, 1635, 1622, 1616, 1576, 1561, 1547, 1464, 1442, 1389, 1372, 1262, 1246, 1187, 1116, 1029 (CO), 968, 888, 752, 730 (CCO2). XRD (u, 2h (count number)[hkl]): 7.0 (200)[200], 8.1 (100)[010], 9.4 (140)[011], 9.8 (260)[220], 11.9 (844)[222], 12.3 (113)[012], 13.7 (293)[400], 15.0 (140)[331], 15.3 (125)[420], 16.7 (133)[422], 17.7 (233)[511], 18.4 (107)[004], 19.3 (360)[440], 20.5 (193)[600], 21.6 (133)[620], 23.6 (127)[444], 24.4 (220)[711], 24.5 (140)[101 ], 26.2 (353)[731], 27.1 (153)[103 ], 29.7 (240)[751], 35.5 (187)[773], 39.5 (193)[1131], 41.8 (153)[1151], 47.4 (193)[995]. BET surface area (m2 g21): 520. TGA (% wt loss, Tinterval [uC]): 18, 2560; 3, 6080; 4, 80110; 4, 110 160; 2, 160220; 3, 220280; 33, 280345. Hydrogen uptake (wt%): 0.82. DSC (Tm [uC], DH [J g21]): 87.8, 379.9; 119.6, 17.9; 180.6, 62.7. Entry 8. SRT energy source, 12 h reaction time, formation of [Cu3(BTC)2] EtOH?H2O solvate in 69.1 (dry basis) % yield. IR (n, cm21, ATR): 38002700 (OH), 3575 (OH), 3094, 2978 (CH), 1704, 1648, 1591, 1548, 1448, 1419, 1375 (COOCu2), 1113, 1044 (CO), 939, 763, 731 (CCO2). XRD (u, 2h (count number)[hkl]): 7.0 (327)[200], 9.8 (429)[220], 11.9 (1186)[222], 13.7 (358)[400], 14.9 (133)[331], 15.2 (102)[420], 16.7 (174)[422], 17.8 (256)[511], 19.3 (389)[440], 20.5 (225)[600], 21.5 (133)[620], 23.6 (112)[444], 24.4 (194)[711], 26.26 (286)[731], 29.6 (225)[751], 35.5 (153)[773], 39.5 (174)[1131], 41.8 (133)[1151], 47.5 (123)[995]. BET surface area (m2 g21): 1732. TGA (% wt loss, Tinterval [uC]): 10, 2570; 9, 70315; 30, 315350. Hydrogen uptake (wt%): 2.44. DSC (Tm [uC], DH [J g21]): 87.8, 513.6; 174.5, 1.6. Thermal desorption GC-MS (% m/ z = 18 uma, Tinterval [uC]): physisorbed, 55.1, 5095; chemisorbed, 44.9, 95190. Entry 9. ST energy source, 12 h reaction time, formation of [Cu(OH)(BTC)(H2O)]?2H2O48 in 52.3 (dry basis) % yield. IR (n, cm21, ATR): 38002550 (OH), 3551, 3500, 3489 (OH), 3097 (CH), 1698, 1638, 1617, 1580, 1508, 1484, 1432, 1399, 1381, 1360, 1114, 1093 (CO), 939, 830, 800, 751, 721 (CCO2). XRD (u, 2h (count number)[hkl]): 9.7 (282)[010], 10.2 (615)[002], 13.9 (461)[012], 17.4 (203)[013], 18.8 (118)[021], 22.4 (227)[014], 25.6 (110)[101 ], 26.2 (123)[101], 27.8 (235)[024], 29.0 (389)[103 ], 30.0 (158)[112], 31.9 (116)[121], 34.4 (134)[034], 38.9 (125)[124], 39.7 (128)[132], 43.4 (151)[126 ], 44.2 (152)[134]. TGA (% wt loss, Tinterval [uC]): 5, 25100; 7, 100140; 2, 140275; 40, 275345. DSC (Tm [uC], DH [J g21]): 78.6, 36.6; 165.0, 11.6; 205.3, 33.3.

Published on 18 April 2013. Downloaded by FAC DE QUIMICA on 05/06/2013 20:57:08.

RSC Adv.

This journal is The Royal Society of Chemistry 2013

View Article Online

RSC Advances Entry 10. SRT + ST energy source, 12SRT h + 12ST h reaction time, formation of [Cu3(BTC)2] MeOH?H2O solvate in 57.7 (dry basis) % yield. IR (n, cm21, ATR): 38132680 (OH), 3574 (O H), 3113 (CH), 1703, 1646, 1617, 1563, 1548, 1449, 1374 (COOCu2), 1272, 1255, 1193, 1113 (CO), 939, 760, 731 (C CO2). XRD (u, 2h (count number)[hkl]): 7.1 (183)[200], 9.9 (271)[220], 12.0 (893)[222], 13.8 (329)[400], 15.0 (132)[331], 15.4 (132)[420], 16.8 (154)[422], 17.8 (212)[511], 19.4 (285)[440], 20.6 (146)[600], 21.7 (124)[620], 23.8 (124)[444], 24.5 (176)[711], 26.3 (249)[731], 29.8 (198)[751], 35.7 (161)[773], 39.5 (190)[1131], 41.8 (132)[1151], 47.4 (139)[995]. BET surface area (m2 g21): 799. TGA (% wt loss, Tinterval [uC]): 12, 2575; 8, 75275; 30, 275340. Hydrogen uptake (wt%): 1.15. DSC (Tm [uC], DH [J g21]): 94.8, 588.5.

Paper the aqueous solvent mixture (Scheme 1). It is not a normal ion + ion A ion + ion metathesis reaction, nevertheless it fulfills the general metathesis statements of i) a double metal displacement, ii) a precipitation and concomitant solubilization of the products; and in this particular case iii) a change in the nature of the chemical bonding, proceeding from ion (Na3BTC) + ion (Cu(NO3)2) A ion (NaNO3) + coordinative covalent nature [Cu3(BTC)2] MOF in the products. In order to explore if this new methodology provided [Cu3(BTC)2] MOFs with modified physicochemical properties, the following was varied: a) the energy source applied to the reaction vessel; b) the solvent and thus the nature of the reaction medium; and c) the reaction time.27 The application of this alkaline metathesis strategy allowed [Cu3(BTC)2] MOFs to be obtained even using low energy sources, such as stirring at room temperature (SRT, entries 58 and 10, Table 2), ultrasound (US, entries 1, 3 and 4, Table 2) and solvothermal (ST, entry 10, Table 2). The alkaline metathesis reaction also provided [Cu3(BTC)2] MOFs in short reaction times, e.g. at instantaneous addition of reactants (61.8%, not shown in Table 2 but determined for comparison purposes), 5 min (67.6%, entry 5, Table 2), 30 min (71.5%, not shown in Table 2, but determined for comparison purposes), 6 h (76.1%, not shown in Table 2, but determined for comparison purposes), with 12 h providing 83.2% (entry 6, Table 2) yield. The best yield was obtained for the 20 min ultrasonic (US) irradiation with a DMF?H2O solvent mixture, providing 87.5% yield (entry 1, Table 2). All of the latter results were measured according to our general purification procedure using centrifugation followed by a couple of mixture solvent washings and oven drying (60 uC) overnight. Hence, occlusion of high boiling temperature solvents such as DMF is expected. Therefore, when DMF was employed as a co-solvent the surface areas of the materials were very low, but when low temperature boiling solvents were employed, such as EtOH and MeOH, variable degrees of surface areas were obtained (see Table 2). To determine if these synthetic variations produced structural or physicochemical differences in the material, further characterization techniques were employed and the results are discussed as follows. XRD analysis Fig. S1 (in the ESI3) shows the collected X-ray diffraction patterns of the obtained compounds prepared by each of the synthetic entries that are gathered in Table 2. All the diffraction peaks as well as their shape and intensities are almost coincident with those described elsewhere for [Cu3(BTC)2] MOFs.29 In particular, for entries 1, 36, 8 and 10 (Fig. S1, ESI3) there are two evident peaks at the [200] and [400] crystallographic planes, which are characteristic for COO2 type fragments present in this structure, as well as other peaks that are characteristic for the chemical structure of this MOF. Nevertheless in entries 2 and 9, the diffraction patterns were different from [Cu3(BTC)2] and indicated the formation of other compounds. In the case of entry 2, the synthetic procedure that solely used H2O as reaction medium, it is not possible to yield [Cu3(BTC)2] MOFs, but a [CuBTC?3H2O] coordination polymer instead. This is even if an alkaline

Published on 18 April 2013. Downloaded by FAC DE QUIMICA on 05/06/2013 20:57:08.

Results and discussion


Synthetic methodology Herein, a low temperature, one-pot metathesis reaction has been employed to yield [Cu3(BTC)2] MOFs (see Scheme 1) using a 6 : 4 equivalent ratio of tri-hydrated copper(II) nitrate (Cu(NO3)2?3H2O) and sodium trimesate (C6H3(COONa)3, obtained in situ (see experimental section). The alkaline activation step for the trimesic acid used to conduct this reaction, for a metathesis type methodology, is one of the major differences between this procedure and other variations present in the general literature. The alkali was used to obtain three sodium carboxylates from the trimesic acid reagent, which more easily reacts with copper nitrate and hence the reaction media in this contribution was basic in nature. As is expected from a metathesis reaction, this simple experimental procedure generated the desired MOF as stated in the entries of Table 2, yielding the [Cu3(BTC)2] MOF in 67.6 to 87.5%. It is worthwhile mentioning that this is the first time this type of alkaline metathesis methodology has been employed to obtain a metalorganic framework. Significantly, the resultant synthetic procedure fulfills the expectation that [Cu3(BTC)2] MOFs be easy to obtain, in two important ways: i) [Cu3(BTC)2] MOFs precipitated (phase change) in the reaction media, and ii) the formation of NaNO3 salt occurred, but it did not intervene in the purification process because it dissolved in

Scheme 1 Low temperature, one-pot metathesis reaction for the formation of [Cu3(BTC)2] MOFs starting from an alkali activated trimesate.

This journal is The Royal Society of Chemistry 2013

RSC Adv.

View Article Online

Paper metathesis reaction is employed and a comparison of the corresponding diffraction pattern with that of this compound has been stated by Gascon et al.47 For the case of entry 9, which is the ST (MeOH?H2O) 12 h reaction, we have compared the diffraction pattern of this species with related compounds and have found that the [Cu(OH)(BTC)(H2O)]?2H2O material, described by Chen et al.,48 provided the same diffraction peaks and XRD intensities. Thus the formation of this compound instead of the HKUST-1 was proven. The only diverse change between this solvothermal synthetic methodology and the rest is the employment of alkalis, through the activation of trimesic acid as sodium trimesate. Thus an increase of pH in the reaction medium occurs, probably banning the normal course of the reaction to yield the desired MOF. For entry 7, which is the SRT (EtOH?H2O) 5 min reaction, there is evidence for the formation of an approximate 3 : 1 mixture of entry 1 and entry 2 crystalline phases (see Fig. S1, ESI3). This result again shows that the [Cu3(BTC)2] compound is obtained in a very fast process that promptly passes through a [CuBTC?3H2O] coordination polymer to yield the desired MOF. The latter was supported because the following of this reaction to reach the conditions described for entry 8 provided the same diffraction pattern for both of the resulting compounds. Again, the presence of an augmented pH in the synthesis causes the formation of very reactive intermediates that swiftly facilitate the production of [Cu3(BTC)2]. Another synthetic advantage is that during all of the XRD characterization of the samples presented herein, evidence of any impurities could be observed, e.g. the formation of Cu2O or other copper oxides (such as CuO) and this also avoided the retention of any NaNO3, trimesic acid, sodium trimesate or Cu(NO3)2 starting materials. Moreover, Schlichte et al.,11 experimentally determined that the ratio between the intensities in the reflections present at [331] and [420] crystallographic planes can be employed as a tracker of the hydration degree in HKUST-1 crystalline structures. Using this method, the ratios for the HKUST-1 materials obtained were measured and placed in Table 2. Interestingly, the measurements have shown that entries 1, 5 and 8 were the most dehydrated materials straight from synthesis, with XRD I[331]/I[420] ratios of 1.27, 1.78 and 1.30 respectively. Meanwhile the rest show a similar value for this ratio, implying that very hydrated samples were obtained and measured, in comparison to where there resulted ratios of less than the unity for the case of entry 4. This result indicates that the synthetic conditions are very important for providing (de)hydrated samples of [Cu3(BTC)2] material straight from reaction vessel with this metathesis procedure. FT-IR analysis In the present case, the FT-IR spectra of the obtained materials have been analyzed, (see Fig. S2, ESI3) and have shown broad bands between 38002700 cm21, which are characteristic of OH groups in water with a large number of hydrogen bonds, thus the amount of water absorbed into the pores was high. In entries 1, 36, 8 and 10 (see Table 2) the formation of [Cu3(BTC)2] MOFs is evident from the presence of C(OCu)OCu metallic esters with bands at 1704, 1648, 1591, 1548, 1448, 1419 and 1375 cm21 indicating an isobidentate mode of

RSC Advances coordination towards the two Cu sites in a symmetric fashion. The FTIR bands present in entry 2 clearly show that formation of [Cu3(BTC)2] is not occurring, due to the presence of bands corresponding to COOCu and COOH, and also an absence of the band at 1703 cm21 (which is always present in [Cu3(BTC)2]). These are both characteristic of the [CuBTC?3H2O] coordination polymer, where only two of the three carboxylates are present as Cu esters and the remaining one does not participate in the coordination and thus remains free as its COOH. As previously observed in the XRD, the FTIR spectra of entry 7 again indicated the presence of a mixture of [Cu3(BTC)2] and the [CuBTC?3H2O] coordination polymer. This was more evident in the carbonyl region of the spectra, which has shown bands for C(OCu)OCu but also for COOCu and COOH. Moreover, in agreement with the XRD data, entry 9 was identified as having a [Cu(OH)(BTC)(H2O)]?2H2O crystalline phase, which is shown here by the characteristic band of an alkaline OH group at ca. 3600 cm21, which is more pronounced for this entry. Additionally, in every entry where [Cu3(BTC)2] is obtained, there are CO bands that belong to small amounts of methanol or ethanol at approximately 1115 1110 cm21 and 10461040 cm21 that should be occluded within the porous net, again indicating that a 3D coordination polymer is formed and the occlusion of solvent molecules easily happened. BET surface areas In those entries where the [Cu3(BTC)2] MOF material is obtained, the synthetic procedure described herein did indeed provide physicochemical enhancements. This is in comparison to the material obtained during the pioneering work of Chui, where the BET specific surface area for HKUST-1 was found to be 692 m2 g21. Meanwhile, the results provided in this study gave better BET specific surface areas of 1771 and 1732 m2 g21, for entries 6 and 8 in Table 2, respectively. They also have practically the same area value and both were obtained using a low temperature procedure. These results and the rest of the nitrogen (77 K) adsorption isotherms are present in Fig. 1. All of these materials were degassed for this characterization at 100 uC. TGADSC analysis The thermogravimetric analysis (TGA) of the material obtained in each entry of Table 2 was determined, showing the characteristic weight loss depending on the solvent mixture at ca. 100 uC and the decomposition of the compound at around 320 uC (Fig. S3, ESI3).49 All of the compounds analyzed have shown weight loss ca. 100 uC, with the compound in entry 9 being the one that released the smallest amount of solvent molecules below this temperature. TGA of the material obtained from entry 6 shows an important weight loss between 100 and 150 uC, which is related to the TGA of the material from entry 2, and to a lesser extent the same occurs for the mixture of crystalline phases present in entry 7. These findings are evidence that the release of chemisorbed water molecules is feasible for these compounds at these temperatures and thus have more metallic sites occupied at low temperatures. TGA results for those entries where HKUST-1 was obtained,

Published on 18 April 2013. Downloaded by FAC DE QUIMICA on 05/06/2013 20:57:08.

RSC Adv.

This journal is The Royal Society of Chemistry 2013

View Article Online

RSC Advances

Paper
Table 3 Weight loss (WL) and cumulative weight loss (CWL) mainly due to loss of solvent molecules as observed in TGA for each synthetic entry where [Cu3(BTC)2] was obtained

Entry 1 3 4

Energy source (solvent) US (DMF/H2O) US (EtOH/H2O) US (MeOH/H2O) SRT (MeOH/H2O) SRT (MeOH/H2O) SRT (EtOH/H2O)

WL (T range) % (uC) 10 (2575) 7 (75300) 12 (2575) 4 (75275) 13 (2565) 7 (65150) 8 (150280) 11 (2580) 5 (80280) 5 (2555) 5 (5595) 10 (95140) 3 (140275) 18 (2560) 3 (6080) 4 (80110) 4 (110160) 2 (160220) 3 (220280) 10 (2570) 9 (70315) 12 (2575) 8 (75275)

CWL (%) 17 16 28 16 23

Published on 18 April 2013. Downloaded by FAC DE QUIMICA on 05/06/2013 20:57:08.

5 6

33

8 10

SRT (EtOH/H2O) SRT + ST (MeOH/H2O)

19 20

Fig. 1 N2 adsorption isotherms at 77 K for [Cu3(BTC)2] MOFs for the entries shown in Table 2.

compiled in Table 3, gave very low weight losses (less than 28%) even when reaching temperatures of 275/315 uC taking into account pure samples (not counting entry 7). As observed in the XRD and FTIR, entry 7 again indicated the presence of a mixture of [Cu3(BTC)2] and the [CuBTC?3H2O] coordination polymer. Again in agreement with the XRD and FTIR data, entry 9 was identified as having a [Cu(OH)(BTC)(H2O)]?2H2O crystalline phase with the TGA clearly showing a small weight loss of ca. 4% below 100 uC and a more pronounced loss of ca. 7% above 100 uC, which was very different in comparison to the TGA results obtained from the HKUST-1 samples. The differential scanning calorimetry (DSC) of the material obtained from each entry in Table 2 was determined. Their thermograms are shown in Fig. 2 and their most important physicochemical data are gathered in Table 4. In general, the thermograms present two main peaks with the low temperature peak corresponding to physisorbed solvent molecules and the high temperature peak relating to chemisorbed solvent molecules.50 In some cases a third peak was observed, but it mainly corresponds to non-MOF materials, but nevertheless the differences are discussed as follows. As before, analysis of the thermograms obtained for entries 2 and 9 concluded that reticular materials were absent, this is due to their low DH values for developing transitions and the appearance of a third

peak. For the rest, the characteristic thermogram of HKUST-1, as determined by Wenge et al.26, was observed. Nonetheless smooth variations in the transition area and positioning of the Tm values were determined. For the analysis of peak 1, from data in Table 4 it can be determined that the material in entry 6 occludes few solvent molecules. A low transition temperature (Tm) value of 74.6 uC was found along with a very low DH value of 62.53 J g21 reinforcing that a low quantity of physisorbed molecules were present within the net. By comparison the other HKUST-1 samples respective Tm values for peak 1 were an average of 87 uC, but the DH values average around 550 J g21, denoting that there were high amounts of easily removable physisorbed molecules. For the analysis of peak 2, the material in entry 6 again shows a low Tm value of 138.1 uC, but in this case it possesses the highest DH value in the series, of 55.8 J g21, showing that the solvent molecules were hard to be remove due to the coordination toward free Cu sites. Comparatively, the other HKUST-1 materials possess higher Tm values with the lowest being 174.5 uC and corresponding to the lowest DH value of 1.6 J g21 (entry 8, Table 4), reinforcing that the synthetic conditions used to obtain this MOF were efficient for retaining a high number of free Cu sites in the net. H2 uptake The H2 uptake was found to be highly dependent on the synthetic procedure used, and most importantly it has been shown herein that it is also dependent on the number of water molecules that occupy the CurX sites of the MOF. An important fact that emphasizes the latter is that, using the same solvent system for both methods, the sample prepared according to entry 8 (2.44 wt%), synthesized by the SRT

This journal is The Royal Society of Chemistry 2013

RSC Adv.

View Article Online

Paper

RSC Advances

Published on 18 April 2013. Downloaded by FAC DE QUIMICA on 05/06/2013 20:57:08.

Fig. 3 H2 adsorption isotherms at 77 K for the [Cu3(BTC)2] MOFs prepared in entries 3, 6 and 8.

method, could adsorb more H2 (ca. 1 wt% more), than the sample prepared according to entry 3 (1.41 wt%), synthesized by the US method (see Fig. 3). In addition, the remaining solvent present near these CurX moieties and affects physicochemical character of the cavity. A noteworthy point is that the [Cu3(BTC)2] material in entry 8 is the one that shows the highest H2 uptake of 2.44 wt%, denoting an important enhancement when compared to the [Cu3(BTC)2] MOFs in entry 6, that only absorbed 1.50 wt% (see Fig. 3). Thus a material with more free CurX sites, depending on the synthetic conditions, and a physisorbed solvent depleted cavity should be better for H2 adsorption. Water desorption GC-MS analysis In order to more strongly state the finer differences of the studied materials, the thermal water desorption from the MOFs obtained under the synthetic conditions used in entries 6 and 8 was carried out by using gas chromatography coupled with a mass spectrometry technique that followed the release of molecules or fragments of 18 g mol21 mass (in this case these are mostly water molecules). As has been depicted in Fig. 4, the MOF compound obtained by using the conditions in entry 6 contained less physisorbed water molecules (water molecules that are desorbed below 100 uC) and more chemisorbed water molecules (water molecules that are desorbed above 100 uC). In comparison, the compound obtained using the conditions in entry 8 contains more physisorbed water molecules and less chemisorbed water molecules. Therefore, the easy water removal from the latter material is achieved at low temperatures and it also contains a large quantity of free Cu coordination sites in the material, thus facilitating better hydrogen uptake. This is for two reasons, there are more free channels as well as more free Cu coordination sites. General remarks The XRD pattern analysis provided clear evidence that [Cu3(BTC)2] MOFs are formed almost instantaneously even

Fig. 2 DSC thermograms for samples 110 (related to the entries in Table 2).

Table 4 Tm and DH obtained from DSC analysis for each synthetic entry as described in Table 2

Entry 1 2 3 4 5 6 7 8 9 10

Peak 1 Tm(uC) DH (J g 86.5 88.4 88.3 83.3 87.3 74.6 87.8 87.8 78.6 94.8 553.7 26.4 623.9 568.3 500.6 62.5 379.9 513.6 36.6 588.5
21

Peak 2 ) Tm (uC) DH (J g 204.9 157.3 180.4 184.2 185.7 138.1 119.6 174.5 165.0 13.2 0.2 11.7 8.9 5.8 55.8 17.9 1.6 11.6
21

Peak 3 ) Tm (uC) DH (J g21) 205.6 175.4 180.6 205.3 10.9 60.8 62.7 33.3

RSC Adv.

This journal is The Royal Society of Chemistry 2013

View Article Online

RSC Advances

Paper combination of stirring at room temperature followed by a solvothermal treatment (entry 10, SRT + ST). Nevertheless, the physicochemical properties obtained for the resulting materials gave poorer results in comparison to that obtained for entry 6 or analogously for entry 8. These results could indicate that, depending on the acidity or basicity of the reacting mixture, the synthetic procedure is favored by ST or SRT, respectively. In our case, the basicity of our employed system improved the generation of HKUST-1 using a SRT method at low temperatures. In order to rationalize the experimental findings, it was important to note that results rely on the XACuCurX coordination sites in the cavities of the [Cu3(BTC)2] material. As shown in other contributions, depending on experimental conditions, these sites can be occupied by various solvent molecules, such as DMF, H2O, NH3, pyridine, etc. Hence, the correct employment of an energy source, such as sonication, led to an important and fast chemical functionalization of the material, as is the case for the [Cu3(BTC)2] MOFs, where the coordination sites are rapidly occupied, mainly by water molecules. As such, according to our experimental data, we have obtained 1 H2O?DMF solvate, 3 H2O?EtOH solvates, and 5 H2O?MeOH solvates (see entries in Table 2). Since, the H2O?DMF solvate provided a very low surface area and was inactive structure (Table 2) it was not used for the hydrogen uptake experiments. As stated elsewhere, if the objective is to achieve H2 storage capacities, the ruling factor is the affinity of the cavities in the MOF, due to free Cu coordination sites, toward the absorption of H2 molecules, as well as the availability of free channels into the material. This clue displaces the normal consensus that a network possessing high porosity will therefore have high storage capacities. This is the present case for the [Cu3(BTC)2] H2O?EtOH solvate material obtained using precise experimental conditions (entry SRT(EtOH/H2O) in Table 2) with an H2 uptake of 2.44 wt% and a BET specific surface area of 1732 m2 g21. This surface area is obtained through a low temperature solvent removal procedure and is equivalent to that obtained through the activated-mechanochemical approach shown in the work of Klimakow et al.37 Indeed, this can be considered the material with the highest BET specific surface area obtained below 100 uC (the entry with conditions of SRT in MeOH/H2O from Table 2 was quite similar having a surface area of 1771 m2 g21, but with an H2 uptake of just 1.50 wt%). It was the highest H2 scavenger in this series due to its better molecular recognition and affinity towards H2 guests. Another important experimental finding is that via the employment of this methodology, a high temperature activation step (above 100 uC) of the resulting material is no longer needed, which has been a major setback in many other contributions to this area. Finally, we consider that the applicability of this procedure for the preparation of other MOFs is an important field of research.

Published on 18 April 2013. Downloaded by FAC DE QUIMICA on 05/06/2013 20:57:08.

Fig. 4 The thermal water desorption GC-MS spectra for compounds prepared in entries 6 and 8.

under mildly energetic conditions. However, the time required to obtain good surface areas was found to be 12 h using SRT (entries 6 and 8 of Table 2) as will be discussed as follows. The best BET specific area was 1771 m2 g21 for the SRT (12 h) method using a MeOH/H2O solvent mixture (entry 6, Table 2), reaching values comparable to that of the commercially available sample material of 1836 m2 g21, but without the need for high temperature (more than 100 uC) solvent removal. Moreover, also using the SRT (12 h) method, but with EtOH/ H2O as the solvent mixture, gave a MOF compound with an H2 uptake capacity with 2.44 wt% (entry 8, Table 2). Here it is worth mentioning that the material formed under the conditions in entry 6 gave the best average results for BET specific area and yield. Two further synthetic efforts were developed, in order to improve these findings. In entries 9 and 10 of Table 2, we tried to improve the synthetic procedure of entry 6 by using solvothermal conditions (entry 9, ST) or by a

Conclusions
Main highlights from this new synthetic alkaline metathesis strategy and the experimental findings are that even using low

This journal is The Royal Society of Chemistry 2013

RSC Adv.

View Article Online

Paper energetic procedures, the reaction to obtain [Cu3(BTC)2] (HKUST-1) MOFs easily takes place using the correct selection of starting materials and thus shows the advantage of this strategy for application in MOF preparation and optimization. Due to the latter, the current synthetic procedure represents one of the easiest, cheapest and most effective ways, up until now, for the preparation of metalorganic frameworks. Our findings present clear evidence that the [Cu3(BTC)2] MOF is formed almost instantaneously even using very mildly energetic conditions, but the time required to reach desired physicochemical properties was found to be 12 h of stirring at room temperature. Anyway, from now on this new strategy should be considered a low energy preparation procedure for HKUST-1 MOF with the highest degree of free Cu sites and low temperature solvent removal channels in order to improve hydrogen storage capacities as well as many other applications within this Cu sites vacancies. We have observed that in CurOH2, the coordination positions can be successfully protected from the chemisorption of water depending on the co-solvent used for the synthesis. For the H2O?MeOH solvates, the low pressure H2 uptake varied from 0.8 to 1.50 wt% depending on the synthetic conditions, while in comparison the H2 uptake of the H2O?EtOH solvates varied from 0.82 to 2.44 wt%. This finding presented clear evidence that low temperature solvent desorbed MOF materials with freely available Cu coordination sites would lead to important compounds with high throughput hydrogen storage capacities, even at low pressure values, just using the correct selection of both synthetic methodology and EtOH water solvent mixture for [Cu3(BTC)2] MOFs which minimizes the chemisorption of water in the metallic sites. It is feasible to prepare this MOF with modulated amounts of physisorbed (molecules placed into the channels) or chemisorbed (molecules occupying CurX coordination sites) water molecules and with high surface area straight from the reaction vessel without any post-synthetic steps. Finally, we consider that the application of this procedure (or variations on this procedure) for the preparation of other MOFs is an important field of research.

RSC Advances 3 J. R. Long and O. M. Yaghi, Chem. Soc. Rev., 2009, 38, 12131214. 4 H. Furukawa, N. Ko, Y. B. Go, N. Aratani, S. B. Choi, E. Choi, A. O. Yazaydin, R. Q. Snurr, M. OKeeffe, J. Kim and O. M. Yaghi, Science, 2010, 329, 424428. 5 H. Deng, S. Grunder, K. E. Cordova, C. Valente, ndara, A. C. Whalley, H. Furukawa, M. Hmadeh, F. Ga Z. Liu, S. Asahina, H. Kazumori, M. OKeeffe, O. Terasaki, J. F. Stoddart and O. M. Yaghi, Science, 2012, 336, 10181023. 6 M. Eddaoudi, J. Kim, N. Rosi, D. Vodak, J. Wachter, M. OKeeffe and O. M. Yaghi, Science, 2002, 295, 469472. 7 O. K. Farha and J. T. Hupp, Acc. Chem. Res., 2010, 43, 11661175. 8 W. Yang, J. Feng, S. Song and H. Zhang, ChemPhysChem, 2012, 13, 27342738. 9 Z. Wang, G. Chen and K. L. Ding, Chem. Rev., 2009, 109, 322359. 10 D. Farrusseng, S. Aguado and C. Pinel, Angew. Chem., Int. Ed., 2009, 48, 75027513. 11 K. Schlichte, T. Kratzke and S. Kaskel, Microporous Mesoporous Mater., 2004, 73, 8188. 12 Y. Fu, D. Sun, M. Qin, R. Huang and Z. Li, RSC Adv., 2012, 2, 33093314. 13 P. Horcajada, C. Serre, G. Ferey, P. Couvreur and R. Gref, Med. Sci., 2010, 26, 761767. 14 B. L. Chen, S. C. Xiang and G. D. Qian, Acc. Chem. Res., 2010, 43, 11151124. 15 C.-Y. Sun, C. Qin, C.-G. Wang, Z.-M. Su, S. Wang, X.L. Wang, G.-S. Yang, K.-Z. Shao, Y.-Q. Lan and E.-B. Wang, Adv. Mater., 2011, 23, 5629. 16 F. Gandara, N. Snejko, A. d. Andres, J. R. Fernandez, J. C. Gomez-Sal, E. Gutierrez-Puebla and A. Monge, RSC Adv., 2012, 2, 949955. 17 Y.-N. Guo, Y. Li, B. Zhi, D. Zhang, Y. Liu and Q. Huo, RSC Adv., 2012, 2, 54245429. 18 N. L. Rosi, J. Eckert, M. Eddaoudi, D. T. Vodak, J. Kim, M. OKeeffe and O. M. Yaghi, Science , 2003 , 300 , 11271129. 19 J. L. C. Rowsell, E. C. Spencer, J. Eckert, J. A. K. Howard and O. M. Yaghi, Science, 2005, 309, 13501354. 20 S. S. Kaye, A. Dailly, O. M. Yaghi and J. R. Long, J. Am. Chem. Soc., 2007, 129, 1417614177. 21 D. Bradshaw, J. B. Claridge, E. J. Cussen, T. J. Prior and M. J. Rosseinsky, Acc. Chem. Res., 2005, 38, 273282. 22 N. Klein, I. Senkovska, K. Gedrich, U. Stoeck, A. Henschel, U. Mueller and S. Kaskel, Angew. Chem., Int. Ed., 2009, 48, 99549957. 23 D. J. Tranchemontagne, J. L. Mendoza-Cortes, M. OKeeffe and O. M. Yaghi, Chem. Soc. Rev., 2009, 38, 12571283. 24 I. F. Hernandez-Ahuactzi, H. Hopfl, V. Barba, P. RomanBravo, L. S. Zamudio-Rivera and H. I. Beltran, Eur. J. Inorg. Chem., 2008, 27462755. 25 B. L. Chen, M. Eddaoudi, S. T. Hyde, M. OKeeffe and O. M. Yaghi, Science, 2001, 291, 10211023. 26 Q. Wenge, W. Yu, L. Chuanqiang, Z. Zongcheng, Z. Xuehong, Z. Guizhen, W. Rui and H. Hong, Chin. J. Catal., 2012, 33, 986992. 27 M. Schlesinger, S. Schulze, M. Hietschold and M. Mehring, Microporous Mesoporous Mater., 2010, 132, 121127. 28 H. Yang, S. Orefuwa and A. Goudy, Microporous Mesoporous Mater., 2011, 143, 3745.

Published on 18 April 2013. Downloaded by FAC DE QUIMICA on 05/06/2013 20:57:08.

Acknowledgements
We acknowledge CONACyT, UAM, SEP-PIFI, SEP-PROMEP and TWAS for financial support. LLN acknowledges a university scholarship from CONACyT. We acknowledge the kind revision and amendments to this contribution done by all the reviewers as well as some of our colleagues.

References
1 M. Eddaoudi, D. B. Moler, H. L. Li, B. L. Chen, T. M. Reineke, M. OKeeffe and O. M. Yaghi, Acc. Chem. Res., 2001, 34, 319330. 2 O. M. Yaghi, H. L. Li, C. Davis, D. Richardson and T. L. Groy, Acc. Chem. Res., 1998, 31, 474484.

RSC Adv.

This journal is The Royal Society of Chemistry 2013

View Article Online

RSC Advances 29 S. S. Y. Chui, S. M. F. Lo, J. P. H. Charmant, A. G. Orpen and I. D. Williams, Science, 1999, 283, 11481150. 30 J. Li, S. Cheng, Q. Zhao, P. Long and J. Dong, Int. J. Hydrogen Energy, 2009, 34, 13771382. 31 D. J. Tranchemontagne, J. R. Hunt and O. M. Yaghi, Tetrahedron, 2008, 64, 85538557. 32 H. Li, M. Eddaoudi, M. OKeeffe and O. M. Yaghi, Nature, 1999, 402, 276279. 33 U. Mueller, M. Schubert, F. Teich, H. Puetter, K. SchierleArndt and J. Pastre, J. Mater. Chem., 2006, 16, 626636. 34 Y. K. Seo, G. Hundal, I. T. Jang, Y. K. Hwang, C. H. Jun and J. S. Chang, Microporous Mesoporous Mater., 2009, 119, 331337. 35 A. Vishnyakov, P. I. Ravikovitch, A. V. Neimark, M. Bulow and Q. M. Wang, Nano Lett., 2003, 3, 713718. 36 Z. Q. Li, L. G. Qiu, T. Xu, Y. Wu, W. Wang, Z. Y. Wu and X. Jiang, Mater. Lett., 2009, 63, 7880. 37 M. Klimakow, P. Klobes, A. F. Thunemann, K. Rademann and F. Emmerling, Chem. Mater., 2010, 22, 52165221. 38 A. G. Wong-Foy, A. J. Matzger and O. M. Yaghi, J. Am. Chem. Soc., 2006, 128, 34943495. 39 J. L. C. Rowsell and O. M. Yaghi, J. Am. Chem. Soc., 2006, 128, 13041315. 40 P. Krawiec, M. Kramer, M. Sabo, R. Kunschke, H. Frode and S. Kaskel, Adv. Eng. Mater., 2006, 8, 293296. 41 J. L. C. Rowsell and O. M. Yaghi, J. Am. Chem. Soc., 2006, 128, 13041315.

Paper 42 K.-S. Lin, A. K. Adhikari, C.-N. Ku, C.-L. Chiang and H. Kuo, Int. J. Hydrogen Energy, 2012, 37, 1386513871. 43 H. I. Beltran, R. Esquivel, M. Lozada-Cassou, M. A. Dominguez-Aguilar, A. Sosa-Sanchez, J. L. SosaSanchez, H. Hopfl, V. Barba, R. Luna-Garcia, N. Farfan and L. S. Zamudio-Rivera, Chem.Eur. J., 2005, 11, 27052715. 44 H. I. Beltran, R. Esquivel, A. Sosa-Sanchez, J. L. SosaSanchez, H. Hopfl, V. Barba, N. Farfan, M. G. Garcia, O. Olivares-Xometl and L. S. Zamudio-Rivera, Inorg. Chem., 2004, 43, 35553557. 45 R. H. Crabtree and D. M. P. Mingos, Comprehensive Organometallic Chemistry III, Volumes 113, Elsevier, 2003. 46 J. L. Sosa-Sanchez, A. Sosa-Sanchez, N. Farfan, L. S. Zamudio-Rivera, G. Lopez-Mendoza, J. P. Flores and H. I. Beltran, Chem.Eur. J., 2005, 11, 42634273. 47 J. Gascon, S. Aguado and F. Kapteijn, Microporous Mesoporous Mater., 2008, 113, 132138. 48 J. X. Chen, T. Yu, Z. X. Chen, H. P. Xiao, G. Q. Zhou, L. H. Weng, B. Tu and D. Y. Zhao, Chem. Lett., 2003, 32, 590591. 49 B. D. Chandler, D. T. Cramb and G. K. H. Shimizu, J. Am. Chem. Soc., 2006, 128, 1040310412. 50 S. Devautour-Vinot, G. Maurin, F. Henn, C. Serre and G. Ferey, Phys. Chem. Chem. Phys., 2010, 12, 1247812485.

Published on 18 April 2013. Downloaded by FAC DE QUIMICA on 05/06/2013 20:57:08.

This journal is The Royal Society of Chemistry 2013

RSC Adv.

You might also like