Sei sulla pagina 1di 10

View Article Online / Journal Homepage / Table of Contents for this issue

PAPER www.rsc.org/materials | Journal of Materials Chemistry

Synthesis and characterization of the interpenetrated MOF-5†


Bi Chen,a Xiujian Wang,b Qianfeng Zhang,c Xiaoyong Xi,a Jingjing Cai,a Huang Qi,a Si Shi,a Jie Wang,a
Dan Yuana and Min Fang*ad
Received 28th October 2009, Accepted 1st February 2010
First published as an Advance Article on the web 12th March 2010
DOI: 10.1039/b922528e
Published on 12 March 2010. Downloaded by Seoul National University on 4/11/2019 4:18:47 AM.

MOF-5 is an important metal–organic framework and has been intensely studied, especially in its
hydrogen storage properties. In this study, we obtained the interpenetrated MOF-5 materials (MOF-5-
int) using N,N0 -dimethylformamide (DMF) or N,N0 -diethylformamide (DEF) as solvents. The
Langmuir surface area of MOF-5-int determined by N2 adsorption is 950–1100 m2 g1, much lower
than the non-penetrated MOF-5 (3000 m2 g1). However, it can store 1.54–1.82 wt% by volumetric
method hydrogen at 77 K and 1 atm, which is higher than the amount stored by the non-penetrated
MOF-5. The MOF-5-int was also characterized by XRD-powder diffraction, thermogravimetric
analysis (TGA), nitrogen adsorption/desorption analysis, scanning electron microscope (SEM) and
X-ray single-crystal structure diffraction. In addition, we found grinding greatly facilitates the
decomposition of the MOF-5-int material by H2O to a nonporous phase ZnBDC$xH2O (within 2–5
min, BDC ¼ 1,4-benzenedicarboxylate), even under low humidity (30%), which calls for careful
handling of the MOF-5 material. The effects of the water content, reaction time, reaction temperature,
molar ratio of Zn(NO3)2 to H2BDC, addition of H2O2, rapid stirring and dilution on the synthesis of
MOF-5-int were studied and the synthetic conditions were optimized. Moreover, Hafizovic et al. (J.
Am. Chem. Soc., 2007, 129, 3612) found the intensity ratio of the powder XRD peak at 9.7 to that at
6.8 (referred to as the R1 value) of MOF-5 can be used to predict its porosity. The lower the intensity
ratio, the more porous it is. In this study, we showed that MOF-5-int can have a very low R1 value but
also a low porosity. The low specific surface area (SSA) is mainly due to its interpenetrated structure
instead of the entrapped zinc species or the mesopores in the material, as previously proposed in the
literature, and associated with the characteristic, very strong peak at 13.8 in its XRD-powder
diffraction pattern. A high R2 value (the ratio of the intensity of the peak at 13.8 to that at 6.8 )
suggests an interpenetrated structure, especially when the R1 value is low. In addition, we found that
although entrapped ZnO or solvent molecules can increase the R1 value, and a low R1 value implies no
zinc species or solvent molecules entrapped in the MOF-5 framework, a high R1 value does not
necessarily suggest the presence of entrapped molecules.

1. Introduction (2900–4000 m2 g1) were usually prepared in the expensive N,N0 -


diethylformamide (DEF) (e.g. IRMOF-12c) and materials with
MOF-5 is the archetype metal–organic framework and has been SSA 600–1300 m2 g1 were usually obtained in dimethylforma-
subjected to numerous studies in the past few years, especially in mide (DMF) (e.g. MOCP-L2a synthesized at room temperature).
hydrogen storage.1 A broad set of conditions and procedures The PXRD (Powder X-ray Diffraction) peak positions of these
have been reported for MOF-5 synthesis.1a, 2 Using Zn(NO3)2$4– materials are essentially the same, except for the peak intensities.
6H2O, MOF-5 materials with high specific surface areas (SSA) Hafizovic et al.3 rationalized this phenomenon based on their
single-crystal structures and proposed that the difference in their
a
Department of Chemistry, Nanjing Normal University, Nanjing 210097, adsorption properties is due to the presence of zinc species
China. E-mail: fangmin@njnu.edu.cn trapped in the cavities and lattice interpenetration in the latter
b
School of Chemistry and Chemical Engineering, Guangxi Normal
University, Guilin 541004, China materials. They concluded that pore filling effects from the zinc
c
Department of Applied Chemistry, Anhui University of Technology, species (and partly the solvent molecules) are also responsible for
Ma’anshan, Anhui 243002, China the pronounced variations in PXRD peak intensities, and the
d
State Key Laboratory of Coordination Chemistry, Nanjing University, intensity ratio of the PXRD peak at 9.7 to the peak at 6.8
Nanjing 210093, China (referred to as the R1 value) can be used diagnostically to predict
† Electronic supplementary information (ESI) available: The PXRD
patterns of the MOF-5 synthesized using DEF (Fig. S1), the the adsorption properties of a MOF-5 type material. That is, the
MOF-5-int before and after the adsorption measurement, the MOF-5 lower the intensity ratio, the more porous the material is. Tsao
materials synthesized according to Huang’s method at room et al.4 also found that the low intensity ratio correlates with
temperature with and without adding H2O2 solution, the MOF-5-int a relatively high surface area. Recently, Yaghi et al.2g reported
prepared in a much diluted reactant mixture (1/4 of the original), the
packing of the MOF-5-int structure, the N2 adsorption isothermal plot a MOF-5 material with a Langmuir surface area of 3909 m2 g1
and the pore size distribution. See DOI: 10.1039/b922528e using Zn(OAc)2$2H2O at room temperature (r.t.). Using

3758 | J. Mater. Chem., 2010, 20, 3758–3767 This journal is ª The Royal Society of Chemistry 2010
View Article Online

Zn(NO3)2$6H2O, Zhao et al.2h synthesized MOF-5 with a Lang-


muir surface area of 2517 m2 g1 in DMF. Calleja et al.5 studied
the differences of IRMOF-1 and MOCP-L, the relatively low gas
adsorption capacity of which was attributed to the blocking of
pores by ZnO impurity.
In this study, we synthesized interpenetrated MOF-5 materials
(denoted as MOF-5-int) using DMF or DEF with 70–90% yield.
The R1 value can be very low, but the material has a relatively
low Langmuir surface area of 950–1100 m2 g1. We found the
Published on 12 March 2010. Downloaded by Seoul National University on 4/11/2019 4:18:47 AM.

intensity of the PXRD peak at 13.8 correlates to the inter-


penetrated structure, especially when the R1 value is low. In
addition, we found grinding greatly facilitates the decomposition
of the MOF-5-int material by H2O to a nonporous phase
ZnBDC$xH2O (within 2–5 min) even under low humidity (30%).
The MOF-5-int materials stored 1.54–1.82% hydrogen at 77 K
and 1 atm, which is higher than the reported 1.32% of the non-
penetrated MOF-5 material (denoted as MOF-5-non) with
a BET of 3000 m2 g1.1b, 6 In addition, we report herein the effect
of the H2O content, reaction temperature, reaction time, molar
ratios and addition of H2O2 on the synthesis of the MOF-5-int
material in DMF, and its synthetic conditions were optimized.

2. Experimental section
Fig. 1 Powder XRD patterns of simulated (a) MOF-5-int and (c) non-
Synthesis of the interpenetrated MOF-5 interpenetrated MOF-5 based on its crystal structure by Mercury 1.4.2
software, (b) the MOF-5-int A synthesized using DMF, (d) the non-
All chemicals were obtained commercially and used without penetrated MOF-5 synthesized by Panella et al.2d digitalized using
further purification. DMF solvent (Sinopharm Chemical Reagent WinDig 2.5, (e) the MOF-5-int B synthesized using DEF.
Co., Ltd, Analytical reagent) was pre-dried over activated mole-
cular sieves (Sinopharm Chemical Reagent Co., Ltd., 4 A,  Na-A)
for 3 d and without further distillation. DEF (99%) was purchased
from Acros, Belgium. Zn(OAc)2$2H2O (Analytical regent) was We attempted to synthesize non-interpenetrated MOF-5
purchased from Guangdong XiLong Chemical Co., Ltd. using DEF according to the literature,8,9 with a slight modi-
In a typical synthesis of the MOF-5-int material with a low R1 fication. CHCl3 was used to exchange the DEF solvent in the
value, 0.4399 g (1.4789 mmol) of Zn(NO3)2$6H2O (Aldrich, pores. Instead, MOF-5-int materials were synthesized. The
reagent grade, dry appearance) in 10.0 mL DMF (pre-dried) was detailed experimental data and powder XRD (PXRD) patterns
mixed with 0.1843 g (1.120 mmol) of H2BDC (Sinopharm in a 2q range of 5–50 are given in the ESI† (Table S1 and
Chemical Reagent Co., Ltd, Analytical reagent) in 10.0 mL of Fig. S1, respectively). One such MOF-5-int (denoted as B)
the same DMF, and small amounts of water (180 mL) were with 71% yield (based on eqn (1)) was synthesized with
introduced. The molar ratio of Zn(NO3)2 : H2BDC :- a molar ratio of Zn(NO3)2$6H2O : H2BDC : DEF of
DMF : H2O ¼ 1.0 : 0.76 : 175 : 12.8.7 The mixture was stirred 1 : 0.40 : 85 at 105  C for 24 h and characterized by PXRD
for 10 min and then transferred to a Teflon-lined autoclave and diffraction, TGA, N2 adsorption/desorption analysis (Lang-
heated for 48 h at 120  C. The autoclave was then cooled to r.t. muir surface area: 1052 m2 g1, BET: 797 m2 g1), elemental
outside the oven. The cubic-like, colorless crystals obtained were analysis (found (calc): C 35.34 (35.42); H, 1.57 (1.49)) and low
washed with the pre-dried DMF three times and dried at 150  C pressure H2 adsorption/desorption analysis (1.79 wt% at 77 K
for 12 h in an oven, resulting in a white solid of 0.2846 g (84.9% and 1 atm).
yield based on Zn(NO3)2 and eqn (1)) which was stored in We also attempted to synthesize MOF-5 with a high surface
a desiccator containing silica gel. For long term storage, it was area in DMF according to the method of Zhao et al.2h DMF
stored in a Schlenck vessel under argon. Unless specified, the solvent was degassed by argon for 3–4 h before use. Zhao
above synthetic conditions were applied.7 The product obtained et al. exchanged the product with chloroform at 70  C for 3 d
has a characteristic PXRD pattern, as shown in Fig. 1b, having and dried it under vacuum at 90  C. We dried the product as
low R1 values and high R2 values, and is denoted as the MOF-5- described for the MOF-5-int A. The product obtained (deno-
int A, which was also characterized by TGA and N2 adsorption/ ted as C, 74% yield based on eqn (1)) has an interpenetrated
desorption analysis (Langmuir surface area: 954 m2 g1, BET: structure (section 3.2 and 3.3). It was characterized by PXRD,
724 m2 g1) and low pressure H2 adsorption analysis (1.54 wt% at TGA, elemental analysis (found (calc): C 37.47 (37.44); H,
77 K and 1 atm). 1.69 (1.57)), N2 adsorption/desorption analysis (Langmuir
surface area: 990 m2 g1, BET: 756 m2 g1) and low pressure
4Zn(NO3)2$6H2O + 3H2BDC / Zn4O(BDC)3 H2 adsorption/desorption analysis (1.82 wt% at 77 K and
+ 23H2O + 8HNO3 (1) 1 atm).

This journal is ª The Royal Society of Chemistry 2010 J. Mater. Chem., 2010, 20, 3758–3767 | 3759
View Article Online

Characterization

XRD patterns were obtained with D/max 2500VL/PC instru-


ment using Cu-Ka1 radiation. A small angle XRD pattern was
obtained with the same diffractometer. Thermogravimetric (TG)
analyses were carried out in a flowing O2 atmosphere
(15 mL min1) with a heating rate of 5  C min1, using Diamond
TG/DTA instrument. The sample weight was 5–6 mg. Scanning
electron microscope (SEM) images of the MOF-int material were
Published on 12 March 2010. Downloaded by Seoul National University on 4/11/2019 4:18:47 AM.

recorded using a JSM-5610LV high resolution scanning electron


microscope.
The porosities and specific surface areas were measured with
a micromeritics ASAP 2020 gas sorption and porosimetry
apparatus using nitrogen gas at 77 K. Before the measurements,
the samples were heated at 250  C under vacuum for 6–20 h. The
hydrogen adsorption measurement (volumetric method) was
measured at 77 K under 0–100 kPa pressure using the same gas
sorption apparatus. Before the hydrogen-uptake measurement,
the sample was heated at 250  C in a vacuum for 10–20 h.
The crystal structure of the interpenetrated MOF-5 was
determined by single crystal XRD diffraction using a Bruker
Smart Apex II diffractometer. The data set was collected at room
temperature. Fig. 2 Powder XRD patterns of (a) the as-synthesized MOF-5-int A
sample dried at 90  C for 2 h, (b) sample (a) continued to be heated at
300  C for 6 h, (c) the as-synthesized sample (in DMF) using
3. Results and discussion Zn(NO3)2$0.45H2O10 dried at 90  C for 2 h, reactant molar ratio:
3.1 Powder XRD Zn(NO3)2$0.45H2O : H2BDC : DMF ¼ 1 : 0.40 : 50, and (d) sample (c)
continued to be heated at 300  C for 6 h under vacuum.
As mentioned in the introduction section, Hafizovic et al.3
proposed that the R1 value can be used diagnostically to predict
the adsorption properties of a MOF-5-type material. The lower Fig. 2(a) and (c). The PXRD of the sample dried at 90  C showed
the value, the more porous the material is. The PXRD peaks a splitting and high intensity peak at around 9.7 . When the same
of the MOF-5-int we obtained (Fig. 1b) have the same 2q values sample was continually heated at 300  C under vacuum for 6 h,
as those of the reported MOF-5 materials (Fig. 1),2d,3 and their the splitting disappeared and the R1 value decreased greatly
sharpness and high intensity indicate high crystallinity of the (Fig. 2(b) and (d)). This suggests that drying at 90  C for 2 h
materials, consistent with the SEM images of the MOF-5-int A, could not remove most of the solvent, while at 150  C for 12 h or
which reveal no amorphous component (Fig. 10, section 3.4). R1 at 300  C for 6 h under vacuum usually can. The reason why the
values are similar to or lower (usually 0.17, lowest value: 0.059) splitting disappears is probably that the structure changed into
than the values (0.21–0.50) of the MOF-5 materials having high a more symmetric structure after the solvent was removed, due to
surface areas.2e,3 However, the Langmuir specific surface area is the rather flexible structure of MOF-5.3
not 3000 m2 g1, but only 950–1100 m2 g1, which was found to be When we used a different zinc source, the Zn(NO3)2$0.45-
mainly interpenetrated MOF-5 (see section 3.2, 3.3). H2O,10 the R1 value (Fig. 2(d)) of the product was not as low as
Besides the low R1 value, the peak at 13.8 is very high, which that of the product obtained by using Zn(NO3)2$6H2O
corresponds to the reflections of (400) planes of the cubic MOF-5 (Fig. 2(b)). We think that this relatively high value originated
structure, while the intensity of this peak of the non-penetrated from the entrapped zinc species, such as Zn(OH)2, in the cavities,
MOF-5 is low.3 Since (400) planes are half the (200) planes, the which started to dehydrolyze to ZnO at 125  C. Some of the
high peak intensity might suggest an interpenetrated structure, Zn(NO3)2$6H2O might have hydrolyzed to Zn(OH)2 during the
supported by other experimental results (see section 3.2). In the heating of Zn(NO3)2$6H2O. This is supported by a similar
interpenetrated structure, the 13.8 peak is due to the reflections synthetic work carried out by Liu et al.11 Their TGA result
of (2, 0, 4), and the 6.8 peak is due to (1, 0, 2) planes. The indicated only 47% weight loss during the decomposition of the
simulated XRD pattern of the non-penetrated MOF-5 based on desolvated MOF-5 framework due to the entrapped ZnO in the
its crystal structure is similar to its PXRD pattern, while the framework, while the theoretical weight loss for material without
simulated XRD pattern of the interpenetrated MOF-5 based on ZnO in the pore is 57.7% (see section 3.2). Thus, the existance of
the crystal structure (solvents were excluded) has different peak zinc species does increase the R1 value, as found by Hafizovic
intensities from that of its PXRD pattern, as shown in Fig. 1. Our et al.3
attempt to synthesize non-interpenetrated MOF-5 in DEF also After gas adsorption analysis (activated at 250 for 6–20 h), the
resulted in MOF-5-int (Fig. 1(e)), showing a strong peak at 13.8 . R1 value increased and the relative intensity of the peak at 13.8
A typical product, after washing with DMF thrice, was usually in the PXRD patterns also slightly changed (Fig. S2 and S3 in the
dried at 150  C in an oven for 12 h. If the solvent is not ESI†). After such an experiment, the structure of the product
completely removed, the R1 value will be high, as shown in should not change. So, such changes could be caused by the slight

3760 | J. Mater. Chem., 2010, 20, 3758–3767 This journal is ª The Royal Society of Chemistry 2010
View Article Online

grinding of the sample during the PXRD experiments (Fig. 12 in and TGA evidence. Since our synthetic method is very similar to
section 3.5). their method (see section 3.6), the product we obtained could also
During our synthesis and characterization of the MOF-5-int contain entrapped Zn(OH)2. Thermogravimetric (TG) analyses
C, we found that a high R1 value does not always suggest of two synthesized MOF-5-int A samples (referred to as A1 and
entrapped ZnO or solvent molecules. In an attempted synthesis A2, obtained under slightly different drying conditions using
of MOF-5 with high surface area according to the method of DMF) were carried out in a flowing O2 atmosphere exactly as
Zhao et al.,2h the MOF-5-int C had a high R1 value (1.2), but Hafizovic et al. did,3 and are shown in Fig. 3 together with their
contained no ZnO or solvent molecules according to its TGA XRD-power diffraction patterns. The PXRD patterns show that
curve (see section 3.2) and elemental analysis. When the intensity the R1 value of sample A1 is higher than that of sample A2.
Published on 12 March 2010. Downloaded by Seoul National University on 4/11/2019 4:18:47 AM.

of the PXRD peak at 9.7 is high, the intensities of the peaks after Consistently, the TG results indicate that sample A1 contained
13.8 also increase. more solvent than sample A2 did. Only about 5% of DMF
solvent was in sample A2, indicating that simply heating the
3.2 TGA of the MOF-5-int recovered sample at 150  C in the oven had removed most
trapped solvent. In regard to sample A1, the 4% weight loss in the
Hafizovic et al.3 found that the high R1 value is caused by the temperature range 20–90  C is assigned to the H2O desorption,
entrapped solvent and Zn(OH)2 (in its non-interpenetrated another weight loss of 13.0% in the range of 100–300  C was
MOF-5 component) based on X-ray single crystal diffraction assigned to DMF desorption. The rather sharp weight loss of
these two samples took place after 375  C, corresponding to the
structural decomposition to ZnO.
Zn(OH)2, if trapped in the pores, would dehydrate at 125  C.
Thus, before 375  C, it has completely converted to ZnO. The
calculated theoretical weight loss for the desolvated, original
MOF-5 upon decomposition to ZnO is 57.7%, whereas the
observed weight losses for samples A1 and A2 are 56.1% and
58.0%, respectively, indicating that there is no zinc species in
sample A2 and only a small amounts of zinc species in sample A1.
The 56.1% corresponds to a dehydrated formula of Zn4O-
(C8H4O4)3$0.27ZnO. The extra ZnO could exist in the pores of
the non-penetrated MOF-5 component or outside of the frame-
work.
In contrast, the MOF-5 material prepared by Hafizovic et al.3
had a much lower weight loss of 46%, due to the trapped zinc
species, corresponding to a dehydrated formula of Zn4O-
(C8H4O4)3$2ZnO. While they only observed 0.5 Zn(OH)2 (later
Fig. 3 TGA curves of the MOF-5-int A samples (A1 (a) and A2 (b)) became ZnO before 200  C) in each cage, corresponding to
synthesized in DMF in flowing O2 (15 mL min1) with a heating rate of a formula of Zn4O(C8H4O4)3$0.5ZnO and no Zn(OH)2 in the
5  C min1. The inset figure is the corresponding XRD-powder diffrac- interpenetrated MOF-5 crystals, consistent with its much smaller
tion patterns of the two samples. pore sizes (the trapped zinc species found in the non-penetrated
MOF-5 was estimated as 7.5 A). 3 Therefore, the extra ZnO might

Fig. 4 TGA curves of the MOF-5-int B synthesized in DEF in flowing Fig. 5 TGA curves of the MOF-5-int C in flowing O2 (15 mL min1)
O2 (15 mL min1) with a heating rate of 5  C min1. The inset figure is the with a heating rate of 5  C min1. The inset figure is the corresponding
corresponding XRD-powder diffraction pattern. XRD-powder diffraction pattern of the sample.

This journal is ª The Royal Society of Chemistry 2010 J. Mater. Chem., 2010, 20, 3758–3767 | 3761
View Article Online

exist outside of the MOF-5 structures, and one cannot deduce desolvated MOF-5 upon decomposition to ZnO is 54.6%. The
how much interpenetrated MOF-5 is present in their case. formula deduced based on TGA data is Zn4O(BDC)3$0.54ZnO.
The TGA curve of the MOF-5-int B synthesized in DEF is The calculated C and H percentage based on the formula is
shown in Fig. 4. The sample started to decompose at 264  C, 35.42% and 1.49%, consistent with the elemental analysis result
much lower than the MOF-5-int synthesized in DMF (375  C). (35.34% and 1.57%).
However, it is stable under 250  C for 6 h, as indicated by its In an attempt to synthesise MOF-5 with high surface area in
PXRD pattern obtained after the N2 gas adsorption/desorption DMF according to the method of Zhao et al.,2h we obtained the
experiment (Fig. S3 in the ESI†). The weight loss for the MOF-5-int C with a high R1 value instead (Fig. 5). The weight
loss for the desolvated MOF-5 upon decomposition to ZnO is
Published on 12 March 2010. Downloaded by Seoul National University on 4/11/2019 4:18:47 AM.

57.5% (Fig. 5), very close to the theoretical value (57.7%), indi-
cating that there is no ZnO trapped in the cages. The calculated C
and H percentage based on the formula is 37.44% and 1.57%,
consistent with the elemental analysis results (37.47% and
1.69%). Thus, although entrapped ZnO or solvent molecules can
increase the R1 value, a high R1 value does not necessarily suggest
the presence of trapped molecules in the pore.

3.3 N2 and H2 adsorption/desorption analysis

The low SSA of a MOF-5 material is either due to included zinc


species, the interpenetrated structure,3 or the mesopores in the
material which were identified by Tsao et al.4 The SSA contri-
buted by the mesopores in the MOF-5 material prepared by Tsao
et al. is 9.8–10.5%. DFT methods revealed that very small
amounts of pores with diameters larger than 20 A  are present in
the MOF-5-int materials (A, B and C). The surface area
contributed by pores with sizes ranging from 2–233 nm in A, B
and C are only 6 m2 g1 (0.8%), 1 m2 g1 (0.1%) and 0 m2 g1
(Table 1 and Fig. 8), respectively, indicating no or very few
mesopores are present. Thus, the presence of mesopores should
not be the cause of the small SSA. Since the MOF-5-int we
obtained contained no or very small amounts of zinc species (see
section 3.2), it is reasonable to conclude that the interpenetrated
structure caused the small SSA.
An interpenetrated structure will reduce the pore sizes. The
pore size distributions of the MOF-5-int materials (A, B and C),
calculated by Horvath–Kawazoe method (referred to as H–K
method), contained pores with size ranged from 4.7–8.0 A.  The
Fig. 6 Pore size distribution of the MOF-5-int materials A (a), B (b) and mean pore diameters are in the range of 5.3–5.7 A  (Fig. 6 and
C (c) calculated by H–K method. Table 1). In contrast, Saha et al.12 reported the pore size

Table 1 Comparison of properties of various MOF-5-int materials obtained

Pore Mean pore ZnO H2 uptake


SLang/m2 g1a SBET/m2 g1b SDFT/m2 g1c vol./cm3 g1d e
diameter/A R1 R2 content/molf (wt%) g h
Pore size/A

A 954 725 759 0.40 5.7 0.13 1.3 (0–0.27) 1.54% H–K: 4.8–8.0 (4.8, 5.5, 6.9)
 652
<8.0 A:
 747
<10 A: DFT: 5.1, 6.8, 8.0, 19
 753
<20 A:
B 1052 797 951 0.40 5.3 0.14 0.68 0.54 1.79% H–K: 4.7–8.0 (4.7, 5.0, 7.0)
 654;
<5.9 A: DFT: 5.6, 7.5, 15, 19
 885
<8.0 A:
 950
<20 A:
C 990 756 866 0.35 5.3 1.2 1.1 0 1.82% H–K: 4.7–8.0 (4.7, 5.3, 7.0)
 364
<3.9 A:
 446;
6.4–7.3 A: DFT: 3.9, 7.3
 866
<8.0 A:
a
Langmuir surface area. b BET surface area. c Surface area obtained using DFT method together with surface areas contributed by certain size pores.
d
Pore volume obtained by H–K method. e Obtained by H–K method. f Moles of ZnO/mole of Zn4O(BDC)3 obtained based on TGA result (see Section
3.2). g wt% H2/absorbed at 77 K and 1 atm. h Pore sizes deduced based on pore size distribution curves (Fig. 6 and 7).

3762 | J. Mater. Chem., 2010, 20, 3758–3767 This journal is ª The Royal Society of Chemistry 2010
Published on 12 March 2010. Downloaded by Seoul National University on 4/11/2019 4:18:47 AM. View Article Online

Fig. 8 Cumulative surface area of the MOF-5-int A (a), B (b) and C (c)
calculated by DFT methods.

Fig. 7 Pore size distribution of the MOF-5-int materials A (a), B (b) and
C (c) calculated by DFT methods.

distribution of the MOF-5-non (Langmuir surface area: 3917 m2


g1) calculated by the same method, showing only one peak at 8.7
 with a mean pore diameter of 8.6 A.
A,  The pore sizes deduced by

DFT methods are less than 10 A (Fig. 7), much smaller than the
14.0–15.9 A  of a MOF-5 material with high surface area
Fig. 9 H2 adsorption/desorption isotherms of the MOF-5-int A (a), B
(2517 m g1) reported by Zhao et al.2h According to the DFT
2

 account for 86, 93 and 100% (b, c) and C (d, e) at 77 K and low pressure.
method, pores smaller than 8.0 A
surface area in A, B, and C, respectively (Table 1).These pores
also account for the majority of the pore volumes (Fig. S6–8 in reached 1.54–1.82 wt% at 77 K and 100 kPa, which is greater
the ESI†). The above evidence again confirms their inter- than the 1.32 wt%, reported under the same conditions, of the
penetrated structures. non-penetrated MOF-5.1b, 6 Considering its much smaller specific
The N2 adsorption/desorption studies also revealed fine surface area, this is significant, indicating that like the inter-
structures of these materials. The pore distribution curves of A penetrated Cu-MOF PCN-6,13 catenation of the MOF-5 material
and B are similar, but are different to that of C. DFT methods is favorable for hydrogen storage. We found MOF-5-int mate-
reveal that there were pores smaller than 3.9 A  in C, which rials with relatively high surface areas have better H2 adsorption
account for 42% surface area, while sample A and B contained
few or no pores smaller than 4.7 A  (Fig. 8). These findings
suggest that the intensity of the PXRD peak at 9.7 , if not caused
by trapped species, may have correlations with pore structures of
the material.
Hydrogen adsorption/desorption analyses of MOF-5-int
materials (A, B and C) were carried out at 77 K and 0–100 kPa,
and the isothermal adsorption/desorption curves are given in
Fig. 9. As Rowsell et al. found under the same conditions with
a non-penetrated MOF-5 material, only a small part of the
complete isotherm is measured due to the high condensation
pressure of hydrogen at 77 K, and the absence of plateaus indi- Fig. 10 SEM images of the MOF-5-int crystals (the molar ratio of the
cates that surface saturation is not achieved. The adsorption synthetic mixture: 1.0 Zn(NO3)2 : 0.74 H2BDC : 12.9 H2O : 765 DMF).

This journal is ª The Royal Society of Chemistry 2010 J. Mater. Chem., 2010, 20, 3758–3767 | 3763
View Article Online

properties, as indicated in Fig. 9 and Table 1. The H2 adsorption/


desorption curves of B and C are almost identical, and the
adsorptions and desorptions are reversible.

3.4 Single crystal X-ray diffraction and SEM study

The MOF-5-int A consists of colorless cubic-like crystals


(Fig. 10a), and some crystals are interpenetrated with each other
as shown in the SEM image (Fig. 10b). Single crystals of the
Published on 12 March 2010. Downloaded by Seoul National University on 4/11/2019 4:18:47 AM.

MOF-5-int A were picked out and X-ray single crystal diffraction


analyses carried out. They were found to have the inter-
penetrated framework as obtained by Hafizovic et al.3 In the
pores were found disordered H2O molecules, causing a high Q
value in the resolved structure. The packing of the MOF-5-int is
given in the ESI† (Fig. S5), showing reduced pore sizes with an
opening from 3–9 A  compared with the non-penetrated MOF-5

structure (8–13 A).

3.5 Conversion of MOF-5 materials under grinding

In order to find out whether the high intensity of the PXRD peak
at 13.8 is due to the preferred orientation of the crystals, we
ground the MOF-5-int. To our surprise, we found that the des-
olvated MOF-5-int converted to another phase, the nonporous 2
within 2–5 min, even under low humidity (30%) (Fig. 11). Upon
increasing the grinding time, more unidentified peaks appeared.
In a parallel experiment, the MOF-5-int from the same product
was ground in a dry box for the same period of time, and the
PXRD pattern shows the product is still MOF-5, but with less
crystallinity and different peak intensities (Fig. 12). The relative
intensity of the peak at 13.8 to 6.7 even increased, not
decreased, indicating that the high peak intensity is not due to the
preferred orientation of the crystals. The intensity of the peak at
9.7 also increased, which could be due to the adsorption of
solvent molecules in the dry box, or the structure change due to
the grinding. Thus, we concluded that the conversion in air is due
to the reaction of the MOF-5 with the moisture in the air. The
MOF-5-int converted to 2 in air with a speed much slower than

Fig. 11 PXRD patterns of MOF-5 after grinding for (a) 0 (showing MOF-5),
(b) 1 min, (c) 2 min (showing phase 2), (d) 5 min and (e) 12 min in air. (f) The
phase N reported by Thirumurugan and Rao14 (g) ZnBDC$xH2O (1 < x < 2)
reported by Mertens et al.15 (h) The MOCP-H reported by Huang et al.2a (i)
MOF-5 after being exposed to air for 24 h reported by Long et al.2f (j) Simu-
Fig. 12 The powder XRD patterns of (a) the synthesized inter-
lated PXRD pattern of MOF-69c based on its crystal structure. (f)–(i) Were
penetrating MOF-5 material; (b) after sample (a) was ground for 5 min in
obtained by digitalizing curves published in the literature using WinDig 2.5.
dry box.

3764 | J. Mater. Chem., 2010, 20, 3758–3767 This journal is ª The Royal Society of Chemistry 2010
View Article Online

under grinding, but similar to that reported for the non-pene- 3.6 Effects of reaction conditions on the synthesis of the MOF-
trated MOF-5.2f The non-penetrated MOF-5, since it has the 5-int A
same framework composition, is very likely to also have the same
When one obtains a MOF-5 material with high R1 value, TGA
property when ground. Phase 2 is also observed by Huang et al.2a
and N2 adsorption experiments must be carried out to determine
They referred it as MOCP-H. Using the same conditions to
whether it has an interpenetrated structure. Thus, we think that
synthesize MOCP-H, MOF-5 was obtained by Panella and
optimizing the synthetic conditions for the MOF-5-int A would
Hirscher1a and us, as indicated by its PXRD pattern (Fig. S9 in
be worthwhile, since A always has a low R1 value, indicating no
the ESI†). So the MOCP-H is actually a MOF-5 material,
zinc species or solvent molecules, and a high peak at 13.8 ,
however, when taking the PXRD spectrum, it was ground in
Published on 12 March 2010. Downloaded by Seoul National University on 4/11/2019 4:18:47 AM.

indicating an interpenetrated structure, and can be conveniently


order to mount the sample and turned into phase 2. In our
synthesized using the inexpensive DMF solvent. The reproduc-
study, a few prepared samples were first thought to be 2 based
ibility of the synthesis of A is good as long as dry DMF (or fresh
on their PXRD pattern, but after retaking the PXRD carefully
analytically pure DMF (H2O # 0.1%)) and dry Zn(NO3)2$6H2O
without much friction when mounting the sample, they were
(from Aldrich) were used. The synthetic procedure of the MOF-
found to be actually MOF-5. The Langmuir surface area of 2
5-int A is similar to that of Hafizovic et al. with some modifi-
obtained after 5 min grinding of a MOF-5-int sample is only
cation. Small amounts of water were introduced to the system,
33 m2 g1, indicating that 2 is nonporous. This discovery
the ratio of Zn(NO3)2 to H2BDC is 1 : 0.75 instead of 1 : 0.5, and
indicates that MOF-5 materials should not be ground in air,
the reaction time is 48 h instead of 21 h. We found that the
even for short time. However, in a recent paper studying the H2
addition of water is necessary to get the interpenetrated MOF-5,
storage via the Spillover method, the MOF-5 material had been
as shown in Fig. 13. Without adding water, an unknown phase
ground with catalysts for 1 h.1e Also, the previously reported
was obtained. Adding too much water also resulted in impurities.
IR spectrum of MOF-5 might not be that of MOF-5, but that
We also notice that the water content affects the intensity of the
of phase 2.
peak at 20.6 . The optimized ratio of Zn(NO3)2 to H2O using dry
Long et al.2f observed that their desolvated non-penetrated
DMF is 1 : 12–13. Mertens et al.15 have identified water and acid
MOF-5 material converted to 2 within 24 h in air (Fig. 11i).
reducing processes for MOF-5 formation in DEF solvent. In the
They propose that 2 is isostructural with MOF-69c. However,
synthesis of MOF-5 in DMF, similar processes may also occur.
the simulated XRD-powder diffraction pattern of MOF-69c
They proposed that low water content and temperature of the
(Fig. 11j) based on its crystal structure is different from that of
reaction solution are the most important factors to arrive at
2. This phase was probably the phase N reported by Thir-
MOF-5, and if the water content in the standard synthesis is less
umurugan and Rao,14 which was identified as ZnBDC$xH2O
than 0.6 mol L1, MOF-5 is formed. In this study, the H2O
(1 < x < 2)(Fig. 11g) by Mertens et al.15 It is interesting that
concentration in some cases reaches 0.9–1.0 mol L1, exceeding
MOF-5 with DEF in the pore will convert into MOF-69c, while
the proposed value.
without solvent in the pore it will convert into ZnBDC$xH2O
The optimized reaction time for synthesizing MOF-5-int A is
(1 < x < 2).
30–48 h (Fig. 14). Shorter (12 h) or longer times (96 h) increased
the R1 value. The amount of product also increased with
increased reaction time from 12 h to 48 h and no apparent
increase after 48 h. These results might suggest that

Fig. 13 PXRD patterns of the interpenetrated MOF-5 synthesized using


dried DMF (molecular sieves for 3 d) and applying different water
contents. The syntheses were carried out by adding (a) 0 mL H2O; (b)
100 mL H2O; (c) 180 mL H2O; (d) 200 mL H2O, (e) 230 mL H2O; (f) 300 mL
H2O. The molar ratios of Zn(NO3)2 : H2BDC : DMF : H2O are
(a)1 : 0.71 : 134 : 6, (b)1 : 0.75 : 144 : 9.1, (c) 1 : 0.77 : 147 : 11.7, (d) Fig. 14 PXRD patterns of products synthesized at 120  C for (a) 12 h,
1 : 0.75 : 148 : 12.3, (e) 1 : 0.73 : 151 : 13.4, (f) 1 : 0.75 : 149 : 15.6. (b) 30 h, (c) 48 h, and (d) 96 h.

This journal is ª The Royal Society of Chemistry 2010 J. Mater. Chem., 2010, 20, 3758–3767 | 3765
Published on 12 March 2010. Downloaded by Seoul National University on 4/11/2019 4:18:47 AM. View Article Online

Fig. 17 PXRD patterns of products synthesized at 130  C for 4 h


Fig. 15 PXRD patterns of the synthesized interpenetrated MOF-5
without adding water (a), adding three drops of H2O2 solution (35%)
at 80 (a), 100 (b), 120 (c), 140  C (d). The molar ratio of
(120 mL) (b), adding 240 mL H2O2 (c), and adding 245 mL H2O (d). (a) and
Zn(NO3)2 : H2BDC : H2O : DMF is kept as 1 : 0.75 : 13.0 : 154.
(b) were synthesized using DMF as obtained. The molar ratio
non-penetrated MOF-5 crystals were obtained first, in which zinc of Zn(NO3)2 : H2BDC : H2O : DMF is (a) 1 : 0.38 : 6 : 134, (b)
1 : 0.38 : 9.3 : 134, (c) 1 : 0.37 : 12.9 : 134, (d) 1 : 0.37 : 12.9 :134. Prod-
species were trapped inside. These trapped zinc species can be
ucts were dried at 180  C for 12 h.
further used (probably in situ) to synthesize penetrated MOF-5
crystals. It has been reported that the zinc species ZnO can be unidentified species. We noticed that although there were
used as a zinc source to synthesize MOF-5.16 On the other hand, mesopores in the MOF-5-int (Table 1), the material could have
a relatively high R1 value could only suggest some structural no zinc species in these pores as indicated by the TGA results
characteristics of the formed MOF-5 material (see section 3.2 discussed above. This evidence, together with the high yield of
and 3.3). Reaction temperature is also an important factor on the the product, suggests that the reaction might occur stoichio-
synthesis of MOF-5-int (Fig. 15). 120  C is the ideal temperature metrically according to the 1 : 0.75 molar ratio of Zn(NO3)2 to
for its synthesis; lower temperatures resulted in unknown species, H2BDC (eqn (1)).
higher temperatures results in high R1 values. Cheng et al.17 found that addition of a small amount of H2O2
Keeping the ratio of Zn(NO3)2 : DMF : H2O in the reaction favored the growth of higher quality samples (larger crystals),
mixtures as 1 : 154 : 12.9, we found the optimized molar ratio of larger pore volume and higher specific surface area MOF-5
Zn(NO3)2 : H2BDC to synthesize MOF-5-int was 1 : 0.75 materials. We applied the synthetic molar ratio of Tsao et al.
(Fig. 16), 1 : 0.5 resulted in a product with a high R1 value. (Zn(NO3)2 : H2BDC : DMF ¼ 1 : 0.37–0.38 : 134), same
Increasing the ratio to a value greater than 1 resulted in temperature and reaction time (130  C and 4 h), but using
a Teflon-lined autoclave in an oven instead of glassware. We
found that by adding 3 drops of H2O2, the R1 value decreased
(Fig. 17a–b). To check whether this is caused by the acid or the
water introduced, equal amounts of water were introduced in
another parallel experiment (Fig. 17c–d). An even lower R1 value
was obtained by adding H2O, and the increase of the acidity of
the system seems to slow down the reaction process, since the
peak at 9.7 is as high as in the case applying a short reaction time
(12 h, Fig. 14a). This might explain why the addition of H2O2 was
found to be favorable to produce large crystals.
We also found that rapid stirring during the reaction cannot
prevent the formation of the interpenetrated structure, as indi-
cated by the high intensity of the peak at 13.8 . Reducing the
concentration to about 1/4 of the original value
(Zn(NO3)2 : H2BDC : H2O : DMF ¼1 : 0.75 : 11.0 : 665) still
lead to the interpenetrated structures, as shown by its PXRD
pattern (Fig. S4 in the ESI†). Recently, Yaghi et al. reported that
a MOF-5 material with a surface area as high as those prepared
Fig. 16 PXRD patterns of the interpenetrated MOF-5 synthesized using in DEF can be synthesized using DMF as the solvent and using
different molar ratios of Zn(NO3)2 : H2BDC. (a)1 : 0.5, (b)1 : 0.75, (c) Zn(OAc)2$2H2O at room temperature (r.t.).2g However, we were
1 : 1, (d) 1 : 1.5 while the ratio of Zn(NO3)2 : DMF : H2O is kept as unable to repeat their results, the phase we obtained was very
1 : 154 : 12.9. similar to what Huang et al. obtained with Zn(NO3)2$6H2O.2a

3766 | J. Mater. Chem., 2010, 20, 3758–3767 This journal is ª The Royal Society of Chemistry 2010
View Article Online

4. Conclusions M. Eddaoudi, D. T. Vodak, J. Kim, M. O’Keeffe and


O. M. Yaghi, Science, 2003, 300, 1127; (d) J. L. C. Rowsell,
We have synthesized the interpenetrated MOF-5 materials J. Eckert and O. M. Yaghi, J. Am. Chem. Soc., 2005, 127, 14904;
(e) Y. Li and R. T. Yang, J. Am. Chem. Soc., 2006, 128, 8136; (f)
using DMF or DEF as the solvent and fully characterized K. Sillar, A. Hofmann and J. Sauer, J. Am. Chem. Soc., 2009, 131,
them. We have found that the high PXRD ratio of the peak at 4143.
9.7 to that at 6.8 (R1 value) is not the only indictor of gas 2 (a) L. Huang, H. Wang, J. Chen, Z. Wang, J. Sun, D. Zhao and
adsorption capacity and surface area. The intensity of the peak Y. Yan, Microporous Mesoporous Mater., 2003, 58, 105; (b) H. Li,
M. Eddaoudi, M. O’Keeffe and O. M. Yaghi, Nature, 1999, 402,
at 13.8 is also an indicator. A high intensity of 13.8 suggests 276; (c) M. Eddaoudi, J. Kim, N. Rosi, D. Vodak, J. Wachter,
an interpenetrated structure, especially when the R1 value is M. O’Keeffe and O. M. Yaghi, Science, 2002, 295, 469; (d)
Published on 12 March 2010. Downloaded by Seoul National University on 4/11/2019 4:18:47 AM.

low, which indicates no trapped molecules in the pores. The B. Panella, M. Hirscher, H. Putter and U. Muller, Adv. Funct.
Mater., 2006, 16, 520; (e) Z. Ni and R. Masel, J. Am. Chem. Soc.,
MOF-5-int materials stored 1.54–1.82 wt% hydrogen at 77 K
2006, 128, 12394; (f) S. S. Kaye, A. Dailly, O. M. Yaghi and
under 100 kPa, the highest among the published MOF-5 J. R. Long, J. Am. Chem. Soc., 2007, 129, 14176; (g)
materials. We also noticed that grinding the MOF-5-int mate- D. J. Tranchemontagne, J. R. Hunt and O. M. Yaghi, Tetrahedron,
rial could turn it into a nonporous phase in a few minutes. 2008, 64, 8553; (h) Z. Zhao, Z. Li and Y. S. Lin, Ind. Eng. Chem.
Res., 2009, 48, 10015.
Some other MOFs might also have similar properties and calls 3 J. Hafizovic, M. Bjorgen, U. Olsbye, P. D. C. Dietzel, S. Bordiga,
for much more careful handling of them. The effect of water C. Prestipino, C. Lamberti and K. P. Lillerud, J. Am. Chem. Soc.,
content, reaction time, reaction temperature, molar ratio of 2007, 129, 3612.
Zn(NO3)2 to H2BDC, and the addition of H2O2 on the 4 C.-S. Tsao, M.-S. Yu, T.-Y. Chung, H.-C. Wu, C.-Y. Wang,
K.-S. Chang and H.-L. Chen, J. Am. Chem. Soc., 2007, 129, 15997.
synthesis of MOF-5-int was studied. The synthetic conditions 5 G. Calleja, J. A. Botas, M. G. Orcajo and M. Sanchez-Sanchez, J.
for the MOF-5-int A were optimized. Dilution does not lead to Porous Mater., 2010, 17, 91.
a non-penetrated structure. The factor controlling the catena- 6 D. Zhao, D. Yuan and H.-C. Zhou, Energy Environ. Sci., 2008, 1, 222.
7 The interpenetrated MOF-5 with low R1 value can also be made
tion behavior of MOF-5 might be the solvent effect or temp- without drying the DMF solvent beforehand. After mixing 0.5210 g
lating effect. Larger solvent molecules or species, such as DEF (1.751 mmol) of Zn(NO3)2$6H2O (Aldrich, regent grade, dry
or zinc species, might fill the pores of the non-penetrated appearance) and 0.2170 g (1.306 mmol) of H2BDC in 20.0 mL of
structure and prevent the further formation of non-penetrated the newly opened analytical pure DMF (H2O # 0.1%), small
amounts of water (180 mL) were purposely introduced. The molar
species. This was supported by the recent study carried out by ratio of Zn(NO3)2 : H2BDC : DMF : H2O ¼ 1.0 : 0.75 : 148 : 11.7.
Choi et al.18 who used 1-methyl-2-pyrrolidone-synthesized non- The rest of the procedures were the same, resulting in white powder,
penetrated MOF-5 under microwave radiation. In addition, we 0.27 g (80% yield based on Zn(NO3)2).
8 (a) S. Hermes, F. Schroder, S. Amirjalayer, R. Schmid and
found that although entrapped ZnO or solvent molecules can
R. A. Fischer, J. Mater. Chem., 2006, 16, 2464; (b) K. Schrock,
increase the R1 value, and a low R1 value implies no zinc F. Schroder, M. Heyden, R. A. Fischer and M. Havenith, Phys.
species or solvent molecules entrapped in the MOF-5 frame- Chem. Chem. Phys., 2008, 10, 4732.
work, a high R1 value does not necessarily suggest the presence 9 J. Gonzalez, R. N. Devi, D. P. Tunstall, P. A. Cox and P. A. Wright,
Microporous Mesoporous Mater., 2005, 84, 97.
of trapped molecules in the cage. 10 To check the effect of zinc source, we heated up 8.0 g Zn(NO3)2$6H2O
at 150  C for 1 h, obtaining a white, sticky solid; according to the lost
Acknowledgements content of H2O, we referred it as Zn(NO3)2$0.45H2O.
11 Y. Liu, Z. Ng, E. A. khan, H.-K. Jeong, C.-B. Ching and z. Lai,
We thank the National Natural Science Foundation of China Microporous Mesoporous Mater., 2009, 118, 296.
12 D. Saha, Z. Wei and S. Deng, Sep. Purif. Technol., 2009, 64, 280.
(Project 20771060) and the Natural Science Foundation of
13 S. Ma, J. Eckert, P. M. Forster, J. W. Yoon, Y. K. Hwang,
Education Department of Jiangsu Province (No. 06KJB150058) J.-S. Chang, C. D. Collier, J. B. Parise and H.-C. Zhou, J. Am.
for financial support of this research. We thank Dr Xueqin An, Chem. Soc., 2008, 130, 15896.
Dr Yaoming Zhou, Mr. Gang Li and Dr Hongbin Du for their 14 A. Thirumurugan and C. N. R. Rao, J. Mater. Chem., 2005, 15, 3852.
15 S. Hausdorf, J. Wagler, R. Mossig and F. O. R. L. Mertens, J. Phys.
help to this project. Chem. A, 2008, 112, 7567.
16 E. Biemmi, S. Christian, N. Stock and T. Bein, Microporous
References Mesoporous Mater., 2009, 117, 111.
17 S. Cheng, S. Liu, Q. Zhao and J. Li, Energy Convers. Manage., 2009,
1 (a) B. Panella and M. Hirscher, Adv. Mater., 2005, 17, 538; (b) 50, 1314.
J. L. C. Rowsell, A. R. Milward, K. S. Park and O. M. Yaghi, J. 18 J.-S. Choi, W.-J. Son, J. Kimb and W.-S. Ahn, Microporous
Am. Chem. Soc., 2004, 126, 5666; (c) N. L. Rosi, J. Eckert, Mesoporous Mater., 2008, 116, 727.

This journal is ª The Royal Society of Chemistry 2010 J. Mater. Chem., 2010, 20, 3758–3767 | 3767

Potrebbero piacerti anche