You are on page 1of 7

Microporous and Mesoporous Materials 119 (2009) 331–337

Contents lists available at ScienceDirect

Microporous and Mesoporous Materials


journal homepage: www.elsevier.com/locate/micromeso

Microwave synthesis of hybrid inorganic–organic materials including porous


Cu3(BTC)2 from Cu(II)-trimesate mixture
You-Kyong Seo a,b, Geeta Hundal a, In Tae Jang a, Young Kyu Hwang a,*, Chul-Ho Jun b, Jong-San Chang a,*
a
Research Center for Nanocatalysts, Korea Research Institute of Chemical Technology (KRICT), P.O. Box 107, Yusung, Daejeon 305-600, Republic of Korea
b
Department of Chemistry, Center for Bioactive Molecular Hybrid, Yonsei University, Seodaemoonku, Seoul 120-749, South Korea

a r t i c l e i n f o a b s t r a c t

Article history: Using a single, improved synthetic process, three different compounds comprising two known phases
Received 21 August 2008 such as [Cu3(BTC)2(H2O)3] (1), [Cu2(OH)(BTC)(H2O)]  2nH2O (2) and a new phase [Cu(BTC–H2)2-
Received in revised form 15 October 2008 (H2O)2]  3H2O (3) have been prepared by microwave-irradiation (MW) of identical reaction mixture of
Accepted 31 October 2008
Cu(II) salt and trimesic acid, benzene-1,3,5-tricarboxylic acid (BTC–H3) at different temperatures. The
Available online 9 November 2008
microwave synthesis of (1) has been compared to its conventional hydrothermal synthesis. It has been
found that by using MW synthesis (1) can be obtained in a much shorter synthesis time with improved
Keywords:
yield and physical properties. The effect of other synthetic parameters such as solvent, concentration of
MOF
CuBTC
reactant mixtures and reaction time on the product (1) has been also studied. At high temperature, the
Microwave synthesis compound (2) was obtained in the same solution. The new phase (3) has been characterized by elemental
Temperature effect analysis, thermogravimetric analysis, IR spectroscopy, scanning electron microscopy and powder XRD
Solvent effect analysis.
Ó 2008 Elsevier Inc. All rights reserved.

1. Introduction of their formation, which is dependent on synthetic conditions


such as reaction temperature, pH, solvent, kinetic/thermodynamic
The metal-organic frameworks (MOFs) have received consider- factors and reaction pathways [27]. We have, therefore, investi-
able attention in the recent years [1–5] for being excellent exam- gated the effects of temperature, solvent and starting precursor
ples of rational designing of porous solids, with specific surface materials and optimized synthesis conditions to get Cu3(BTC)2
areas and micropore volumes that surpass many of the zeolites, (H2O)3 phase in high yields and purity by CE method. Two such
metal phosphates and activated carbons [6–8]. The pursuit is being ventures [28,29] have produced two entirely new coordination
made fervently because of their potential applications in gas stor- polymers having the molecular formula [Cu2(OH)(BTC)(H2O)] 
age [9–11], catalysis [12,13], separations [14], drug delivery [15] 2H2O and [Cu(BTC–H)(H2O)3] on changing the solvent system from
and molecular recognition [16]. Cu3(BTC)2(H2O)3  xH2O (named water/ethanol to water in a reaction time of 24 h and 12 h, respec-
as HKUST-1) is one of the first robust metal-organic framework tively, at 120 °C.
(MOF) materials with a microporous structure that is reminiscent The microwave-assisted hydrothermal synthesis (henceforth
of zeolite frameworks [17]. The HKUST-1 is one of the highly cited MW) is known to be advantageous over CE synthesis because of ra-
MOFs because it has a large surface area, high pore volume, high pid heating, faster kinetics, phase purity, higher yield and better
chemical stability, high Lewis acidity and lability of coordinated reliability and reproducibility [30–34]. In addition to this, it pro-
water molecules. These properties enable it to be a potential can- vides an efficient way to control particle size distribution, phase
didate for adsorption, gas storage applications and catalysis [18– selectivity, and macroscopic morphology in the synthesis of nano-
25]. porous materials as well as inorganic solids [35–38]. Despite these
Ever since the synthesis of Cu3(BTC)2(H2O)3 was reported [17] facts, to the best of our knowledge, there are only a few reports on
by Chui et al. the material has been synthesized a number of times the use of microwave for synthesis of hybrid materials [39–43].
[18–25] by conventional hydrothermal methods (henceforth CE) or Here, we present the use of microwaves for the synthesis of
electrochemical method [26] so as to optimize the synthetic condi- [Cu3(BTC)2(H2O)3] (1) in a much shorter time and lower tempera-
tions and improve the purity of the product (Table 1). However, ture than reported for the original synthesis via CE [17]. Various
MOFs show considerable structural diversity and chemical trends factors which may affect the purity, yield, morphology and crystal-
linity of (1) thus formed, have been investigated and the reaction
* Corresponding authors.
conditions have been optimized. In addition to the above, we also
E-mail addresses: ykhwang@krict.re.kr (Y.K. Hwang), jschang@krict.re.kr report here the synthesis of an already known MOF material
(J.-S. Chang). [Cu2(OH)(BTC)(H2O)]  2H2O (2), and another new material [Cu

1387-1811/$ - see front matter Ó 2008 Elsevier Inc. All rights reserved.
doi:10.1016/j.micromeso.2008.10.035
332 Y.-K. Seo et al. / Microporous and Mesoporous Materials 119 (2009) 331–337

Table 1
A summary of various syntheses of Cu3(BTC)2 and their characteristics.

MOF SBET (m2/g) Reaction temperature (°C) Reaction time (h) Solvent used Method Reference
(1) 692 180 12 H2O, EtOH CE [17]
(1) 1500 110 18 H2O, EtOH CE [24]
(1) 964/1333 150 18 H2O, EtOH CE [24]
*
(1) 120 12 H2O, EtOH CE [18]
(1) 1200–1400 110 15 H2O, EtOH CE [19]
(1) 1239 150 12 DMF APS [23]
*
(2) 120 24 H2O CE [28]
*
(1) 110 18 H2O, EtOH CE [20]
*
(1) 130 12 H2O, EtOH CE [24]
*
Cu(BTC)(H2O)3 120 12 H2O CE [29]
(1) 1392 140 0.5 H2O, EtOH MW This work
(1) 1656a 140 1 EtOH MW This work
(1) 1300–1500 20–23 4 MtOH EC [26]
a
Surface area was obtained in high concentrated solution (3 times comparing to unit concentration).
b
EC denotes electrochemical method and APS denotes ambient pressure synthesis method, respectively.
*
Values are not available.

(BTC–H2)2(H2O)2]  3H2O (3), synthesized from identical reaction monochromator. The crystal size and morphology were examined
mixtures by changing the reaction temperature only in a shorter using a scanning electron microscope (SEM, JEOL JSM-840 A). Ther-
time giving higher yields and phase purities. mogravimetric analyses (TGA) were performed with TA instrument
model, TGA Q500 V6.7 with 10 °C/min in N2 atmosphere, on sam-
2. Experimental ples which were placed in a chamber having saturated NH4Cl vapor
for 24 h. The BET analyses were performed with N2 adsorption–
2.1. Synthesis of Cu–MOF compounds based on Cu(II) and BTC–H3 desorption isotherms at liquid nitrogen temperature (77 K) after
dehydration under vacuum at 423 K for 12 h using Micromeritics
Various reaction conditions used for the syntheses and their Tristar 3000. The specific surface areas were evaluated using the
outcomes are listed in Table 2. For having best yields and purity Brunauer-Emmett-Teller (BET) method in the p/p0 range of 0.05–
of the phases, the optimized synthetic conditions for small scale 0.2. IR spectra were recorded as KBr pellets on a Nicolet Magna-
preparation of compounds (1) and (2) under microwave-irradia- 560 IR spectrophotometer. Elemental analyses were done on a CE
tion are as follows. Typically, an exact amount of H3BTC (2 mmol) instruments EA-1110 elemental analyzer and on Jobin-Yvon Ulti-
and copper (II) nitrate trihydrate, Cu(NO3)2  3H2O (3.65 mmol) ma-C (ICP-AES) for Cu.
were dissolved in 24 mL of a 1:1 (w/w) mixture of water:ethanol
and stirred magnetically for 10 min (Standard concentration). The 3. Results and discussion
resulting mixture was then loaded into a Teflon autoclave, sealed
and placed in a microwave oven (MARS-5, CEM). The autoclave 3.1. Temperature-dependent synthesis of Cu–MOFs
was heated for 60 min at 140 °C (1) and for 10 min at 170 °C for
(2), respectively, in the microwave power of 300 W. After the com- Since the first synthetic report of a porous 3D coordination
pletion of the reaction, the mixtures were allowed to cool down to polymer Cu3(BTC)2 at 180 °C by Chui et al. [17], (1) has been pre-
room temperature, filtered and washed with water/ethanol mix- pared by many groups using CE at temperatures varying from
ture for several times and were dried at 100 °C overnight. For the 110 to 180 °C (Table 1). The optimized reaction conditions were gi-
sake of comparison, (1) was also prepared by CE following the ven by Schlichte et al. who showed that Cu2O free samples could be
known method [17] using the above mixtures at 120 and 140 °C, obtained after heating the synthesis mixture for 12 h at 120 °C. The
respectively, for 12 h (Table 2). higher temperature synthesis of the above materials is generally
beneficial for crystallographic investigations, but it is mostly
2.2. Characterization accompanied by impurities of Cu2O. Since 140 °C proved to be
the best temperature in our case for MW synthesis of (1), we pre-
Powder X-ray diffraction patterns of all samples were obtained pared the CE samples (for comparison) both at 120 and 140 °C. As
by a Rigaku diffractometer (D/MAX IIIB, 2 kW) using Ni-filtered Cu the XRD patterns of both samples were exactly same (Fig. 1a–d),
Ka-radiation (40 kV, 30 mA, k = 1.5406 Å) and a graphite crystal most characterization techniques were applied to 140 °C samples

Table 2
Temperature effect in synthesis of Cu–MOFs.

No. Method Temperature (°C) Time (min) Water/ethanol (w/w, g) Yields (%) Crystal structure
1 MW 120 60 12:12 No product (1)
2 MW 130 60 12:12 19 (1)
3 MW 140 60 12:12 88 (1)
4 MW 150 10 12:12 84 (3)
5 MW 160 10 12:12 79 (2)b
6 MW 170 10 12:12 92 (2)
7 CE 140 760 12:12 85 (1)
8 CE 120 760 12:12 95 (1)
* Concentrations used in all cases are standard unit one (H BTC 2 mmol, Cu(NO ) 3H O 3.65 mmol in 24 ml of solvents).
3 3 3 2
b
Indicates some impurity in the product formed.
Y.-K. Seo et al. / Microporous and Mesoporous Materials 119 (2009) 331–337 333

MW even after heating for 1 h. Microwave synthesis at both 130


(h)
and 140 °C gives (1) but better yields are obtained at the latter
(Table 2). Heating at 140 °C starts giving (1) after 5 min but maxi-
(g) mum yield is obtained after 60 min treatment. A 10 min heating at
150 °C results in the new phase (3), which starts changing into yet
another phase [Cu2(OH)(BTC)(H2O)]  2H2O (2) on heating at
Intensity / a.u.

(f ) 160 °C, completely becomes (2) at 170 °C and remains in that


phase on increasing the reaction time further (Fig. 1e–h). The
(e)
SEM photographs of the products formed at these four tempera-
tures ( Fig. 2) clearly show the differences in the morphologies.
(d) Phase (3) consists of aggregates (10 lm) of small particles with
indefinite shapes. However, its XRD pattern is incomparable to
(c) both (1) and (2) as well as to the other three known phases of Cu(II)
(b) with BTC obtained at ambient temperature [44,45]. Both (1) and
(a)
(2) are crystalline with the former having octahedron-shaped sin-
gle crystals of >1 lm in size and the latter having plate like crystals
5 10 15 20 25 30 of an average size 5 lm. For (1) both CE and MW form turquoise
2 θ / degree blue crystals with different particle sizes (not shown). Under CE
conditions, SEM image of Cu–BTC shows octahedron-shaped crys-
Fig. 1. XRD patterns of (a) simulated structure of Cu3(BTC)2(H2O)3 (1) and samples
prepared at (b) CE 120 °C, (c) CE 140 °C, (d) MW 140 °C, (e) MW 150 °C, (f) MW tals with large particle size (P20 lm) whereas MW method yields
160 °C, (g) MW 170 °C and (h) simulated pattern of [Cu2(OH)(BTC)(H2O)]  2H2O (2) a more irregular shaped product with the smaller particle size
for 1 h MW and for 12 h CE, respectively. (610 lm).
The thermal stability was studied using TGA (Fig. 3), which
only, unless specified otherwise. No attempts were made to syn- again shows that (1) prepared by both CE and MW are similar.
thesize [Cu2(OH)(BTC)(H2O)]  2H2O by the reported CE [28] and The first step in the curves is associated with the loss of physio-
(2) has been characterized by comparison with its simulated XRD sorbed water and its exact height depends on the initial degree
pattern and TGA analysis. Powder XRD patterns of typical samples of hydration of the material. Both the as-synthesized samples show
of (1) and (2) are given in Fig. 1a–d and are compared with those a continuous weight loss (30%) up to a temperature of 110 °C cor-
recorded for (1) prepared by CE and also with those calculated responding to a theoretical loss of (29.7%) for 15 water molecules
from their known crystal structures [18,28]. The patterns show per Cu3 unit. The results are in accordance with those found by
good overall agreement with the simulated ones; however, some Schlichte et al. who have demonstrated a higher capacity for sol-
deviations in relative intensities are expected because of variations vent molecules in the interior of the pores [18]. The second weight
in the degree of hydration [18]. For example, the intensity of (1 1 1) loss (40.0 %) starting at 280 °C disintegrates the metal-organic
reflection in (1) increases manifolds in the samples prepared at structure (39.5% loss) with a complete transformation into Cu2O
140 °C. and CuO. The TGA profile for (2) shows an initial weight loss of
The above results indicated that the temperature plays a crucial 14.0%, in two steps corresponding to three water molecules
role in MW synthesis. There was no product obtained at 120 °C in (13.3%). The compound disintegrates at 282 °C with a weight loss

Fig. 2. Scanning electron micrographs of Cu3(BTC)2(H2O)3 (1) synthesized at (a) 140 °C, (b) 150 °C, (c)160 °C and (d) 170 °C for 10 min under MW, respectively.
334 Y.-K. Seo et al. / Microporous and Mesoporous Materials 119 (2009) 331–337

100 and concentration of synthesis mixture have been compared vis-


à-vis CE as well as MW methods to obtain (1) in high yield and
phase purity.

80
3.2.1. Role of solvent
The nature of the phase formed is very much dependent on the
Weight loss / %

(d) (water:ethanol) solvent ratio used. In the reference synthesis of (1)


60 reported by Chui et al. a 1:1 (w/w) mixture of water/ethanol was
used and later this method of preparation was optimized by
Schlichte et al. [4] with the same solvent system at lower tempera-
(c) ture of 120 °C. Chen et al. [28] produced [Cu2(OH)(BTC)(H2O)] 
40 (b)
2H2O (2) on changing the solvent system from water/ethanol to
water alone in 24 h, whereas Gascon et al. [10] have reported the
(a) formation of a completely nonporous material, Cu(BTC–H)(H2O)3
20 [44] in 12 h, with all other conditions being same as used by
Schlichte et al. The use of a different solvent i.e. DMF has been re-
ported by Krawiec et al. [23] in their ambient pressure synthesis
100 200 300 400 500 600
o of (1) involving refluxing and repeated heating at various tempera-
Temperature / C
tures to get (1). To the best of our knowledge, there has not been get
Fig. 3. TGA curves of compound (1) synthesized at (a) 140 °C by CE, (b) 140 °C by any report of a systematic variation of water/ethanol ratio in the
MW, compound (2) at (c) 170 °C and compound (3) at (d) 150 °C by MW, synthesis of (1). Since it is now known that a 1:1 water/ethanol
respectively. mixture gives (1) and water only yields (2) in a CE synthesis, we
therefore decided to study the effect of varying the water/ethanol
ratios in the MW synthesis and observe the resulting changes in
of 37.8%, which is comparable with a theoretical loss of 36.7% the physicochemical properties in the synthesis of (1).
accounting for decomposition into Cu2O and CuO. The chemical A complete deprotonation of H3–BTC is required for its mult-
formula [Cu(BTC–H2)2(H2O)2]  3H2O of (3) was determined from identicity to produce three dimensional, rigid and thermally stable
elemental and TGA analyses. The experimental and calculated data systems. H3BTC is less soluble in water than in water/ethanol sys-
for (3) (for C18H14O14Cu in parenthesis); C, 42.08 (41.8), H, 2.75 tem, leading to different products depending on the amount of dis-
(2.71), Cu, 12.9 (12.3) agree well with the molecular formula solved acid and degree of its deprotonation. Therefore the ratio of
Cu(BTC–H2)2(H2O)2. The TGA curve of (3), however indicates three water/ethanol used in the reaction turned out to be critical. The
molecules of water of crystallization. It shows an initial loss of 3.4% synthesis yields of (1) in the pure form when the water:ethanol ra-
due to one water molecule (calc. 3.1%) at 84 °C, followed by a loss tios were varied from 1:5 to 1:1 (Fig. 4). On increasing the amount
of four more water molecules up to 300 °C, in steps of two each of water to 2:1 gives impurity in the phase which is evident in the
with a loss of 7.3% both times (calc. 6.5% and 6.9%). The structure form of an extra peak at 2h = 10.4 and a shoulder at 13.8°. The reac-
disintegrates at 309 °C to leave Cu2O and CuO with 52.7% loss (calc. tion in pure ethanol also yields the product (1) but impurity peak is
53.1%). The IR spectrum of (1) prepared by both methods are sim- seen at 2h = 12.9°. Increasing the ratio of water/ethanol 20:4 (5:1)
ilar providing another evidence of them being the same material gives the phase Cu2(OH)(BTC)(H2O) (2). Further increase in
(not shown). The broad absorption band near 3500 cm 1 which is water:ethanol ratio to 24:0 changes it into the new phase (3). From
due to water molecules (both coordinated and lattice) gradually Figs. 1 and 4, it is quite evident that in MW synthesis, the phase (2)
decreases in intensity due to dehydration upon heating the sample. may be prepared either by increasing the reaction temperature to
To confirm the permanent porosity, we obtained the N2 adsorp- 170 °C or by just changing the solvent mixture to 20:4 water:eth-
tion–desorption isotherms of the samples obtained from CE and anol instead of a 1:1 mixture in MW at 140 °C synthesis. Similarly
MW at 196 °C after evacuation at 150 °C for 12 h. The type I nitro- (3) could be obtained by changing the temperature to 150 °C or
gen physiosorption isotherms of (1) prepared by MW at 140 °C for
30 min give BET surface area of 1392 m2/g (total pore volume of
0.56 cm3/g). As listed in Table 1, these values are much higher than
the BET surface area of 974 m2/g for (1) prepared by CE by us and
those reported in literature [17,18,20] using similar concentra-
tions. The values, however, look similar to those reported in the lit- (e)
erature [19,22–24]. However, in the synthesis of (1) they have used
different concentrations than used by us and thus are incompara-
Intensity / a.u.

ble. Among results of similar concentrations as used in our case,


the increase in surface area and small particle size (from SEM) of (d)
our samples signify that the MW synthesis accelerates the nucle-
ation and crystal growth steps and produces smaller particles with (c)
larger surface areas.
(b)
3.2. Factors affecting the formation of phases

From the above discussion it has been established that MW can (a)
be used to produce (1) and (2) in a very short time. Since our main
5 10 15 20 25 30
aim is to find an economical method for the preparation of (1), we
2 θ / degree
therefore examined systematically the factors which could affect
the yield, morphology, particle size, surface area and porosity of Fig. 4. XRD patterns of the products formed with change in water/ethanol ratio of
(1) being formed by MW. The factors like solvent, reaction time (a) 24:0, (b) 20:4, (c) 16:8, (d) 8:16 and (e) 0:24.
Y.-K. Seo et al. / Microporous and Mesoporous Materials 119 (2009) 331–337 335

running the reaction in a water only system at 140 °C. However,


Gascon et al. [29] and Chen et al. [28] have reported the synthesis
of one dimensional phase Cu(BTC–H)(H2O)3 and (2), respectively, (f )
in similar CE hydrothermal syntheses at 120 °C, in the absence of
ethanol (Table 1). The phase (3) is nevertheless a new phase (cf.
XRD patterns) can also be verified from a comparison of the IR

Intensity / a.u.
(e)
spectra and TGA curves for these samples. The IR spectra of MW
at 140 °C (water:ethanol 24:0) sample is identical to that obtained
for MW at 150 °C (water:ethanol 1:1), i.e. (3) (Fig. 5) whereas those
of MW at 140 °C (1) and MW at 170 °C (2), both (water:ethanol (d)
1:1) are very much different to each other. A comparison of (1)
and (3) reveals that IR spectrum of the latter shows additional
peaks than those present in (1). The most significant things of these
peaks are the stretching VC@O, VC–O and bending O–H vibrational fre- (c)
quencies seen at 1710, 1193 and 1232 cm 1, respectively, indicat-
5 10 15 20 25 30
ing the presence of a carboxylic acid group [46,47]. The VO–H band
is much broader than (1) and shows a low energy component at 2 θ / degree
2550 cm 1, which is characteristic of a strongly H-bonded O–H of
the carboxylic acid [46]. From these bands, it may be concluded Reaction time / h
that BTC group in phase (3) is only partially deprotonated. Simul- 0 20 40 60 100 110 120
taneously, there are more peaks seen in the 1300–1600 cm 1 re- 100
gion than in (1). As these peaks stand for bridging bidentate
(a)
coordination of carboxylate group, therefore a greater number of
80
these peaks indicate a rather more unsymmetrical kind of coordi-
nation by the different carboxylic acid/carboxylate groups to dif-
(b)
ferent metal ions, in comparison with a symmetrical dimeric 60
Yield / %

copper (II) carboxylate type of structural units found in (1). The


TGA curves for MW at 160 °C, MW at 170 °C and MW at 140 °C
(water:ethanol 20:4) samples are identical resulting all to be the 40
phase (2). Similarly, TGA curve of MW at 140 °C (water:ethanol
24:0) sample is identical to that obtained for MW at 150 °C
(water:ethanol 1:1) sample demonstrating these to be the same 20
phase (3).

3.2.2. Effect of reaction time 0


With the optimum temperature being 140 °C for the MW syn- 0 20 40 60 100 110 120
thesis of (1) in the pure form, we further studied the effect of reac- Reaction time / min
tion time on the products being formed. Fig. 6 shows that MW the Fig. 6. Yields of MOF according to the reaction time under (a) MW and (b) CE. XRD
yield of (1) increases linearly with increase in crystallisation time patterns of Cu3(BTC)2(H2O)3 prepared by MW at 140 °C for (c) 60 min, (d) 30 min,
from 5 to 30 min, with a larger increase (50%) between 10 and (e) 10 min and (f) 5 min.
30 min. From 30 to 60 min the increase is there but to a relatively
smaller extent (20%). It is worth noting that the yields are incred-
ibly low when the crystallization time is of the order of 5 min. time, although the total decrease is only 30%. The SEM image of
However, in CE the yield decreases with an increase in reaction MW products at various reaction times shows that the dimensions
of the crystals formed increase (from 10 to 20 lm) with increasing
crystallization times (Fig. 7a and b), whereas reverse is true in the
case of high temperature (HT) synthesis. In CE (Fig. 7c and d), the
crystal size is decreasing from 25 lm (for crystallization time
12 h) to a constant value of 10 lm (for 24 h and 48 h). Both,
(b) the decrease in yield and particle size may be explained by consid-
ering that in CE synthesis the nucleation and crystal growth are
Transmittance / a.u.

slow processes and thus it takes much longer time. Due to long
crystallization times (12–48 h), the crystals formed remain in con-
tact with the solvent for a much longer period. This may result in
creating equilibrium of the kind, Solute(crystallized) M Solute(dissolved),
(a)
i.e. between the amount of solute crystallized and that remaining
in the solution. Under the effect of heating in CE conditions, espe-
cially at low pH as in the existing case (<2), the reaction may shift
largely towards right depending on time duration, before it reaches
equilibrium. In MW, however the heating is fast and uniform, cre-
ating nuclei throughout the solution which quickly grow to crys-
3000 2500 2000 1500 1000 500 tals. Initially, with increasing time the number of nuclei formed
-1 increases to a greater extent and hence there is a big jump in yield
Wavenumber / cm
from 5 to 30 min and as both the nucleation and crystallization
Fig. 5. FT–IR spectrum of (a) compound (3) and (b) compound (1), the arrows point steps are accelerated [35–43] by MW effect and therefore all the
towards the significant additional peaks in (3). nuclei once formed grow more with longer crystallization time.
336 Y.-K. Seo et al. / Microporous and Mesoporous Materials 119 (2009) 331–337

Fig. 7. SEM images of Cu3(BTC)2(H2O)3 at 140 °C for (a) 5 min, (b) 30 min by MW, (c) 12 h, and (d) 24 h by CE, respectively. Concentration of solution is (e) 3 times and (f) 5
times by MW at 140 °C, for 1 h.

From 30 to 60 min, an increase in nucleation is there but is rela-


tively less, resulting in a lesser but definite increase in the yield
accompanied by larger crystals as well. Since the maximum reac- 600
tion time given is 1 h only, MW method does not suffer from back (b)
of dissolution of crystals in the solution (see Fig. 8).
500
-1

(c)
3.2.3. Effect of concentration change in reactant
Volume adsorbed / ml g

Although the MW method usually does not yield crystals with 400 (a)
size adequate enough for single crystal studies, it produces uni-
form and faster nucleation [13] leading to faster crystal growth
300
and therefore the crystal sizes may be varied by changing the reac-
tion concentrations. Thus smaller particles may be produced by
reducing the concentration of the starting solution and vice versa. 200
Gascon et al. [29] have reported changes in yield and size of the
crystals with concentration of the reactants in a typical CE synthe- 100
sis of (1). They observed that the yield of the product formed was
independent of the concentration of the reactants (close to 98%),
0
however the crystal size was highly dependent on it. More concen-
0.0 0.2 0.4 0.6 0.8 1.0
trated reaction mixtures produced larger crystals (Fig. 7c and d).
P/P 0
The yield becomes quantitative with increase in concentration of
the reactants by 3 times (98%), subsequently becoming invariant Fig. 8. N2 isotherm of Cu3(BTC)2(H2O)3 prepared by MW at 140 °C for 1 h with
on a further increase up to 5 times (99%) of the initial concentra- different solute concentration. Concentration of solution is a (a) standard, (b) 3
tion (88%). The SEM images of the products (Fig. 7e and f) clearly times and (c) 5 times.
Y.-K. Seo et al. / Microporous and Mesoporous Materials 119 (2009) 331–337 337

show an increase in crystal size from 20 lm to 50 lm on increas- [4] M.C. Hong, Y.J. Zhao, W.P. Su, R. Cao, M. Fujita, Z.Z. Zhou, S.C. Albert, J. Am.
Chem. Soc. 122 (2000) 4819.
ing the concentrations from 3 to 5 times, and this when compared
[5] C.J. Kepert, T.J. Prior, M.J. Rosseinsky, J. Am. Chem. Soc. 122 (2000) 5158.
to Fig. 7a and b gives a linear relationship between concentration [6] S.L. James, Chem. Soc. Rev. 32 (2003) 276.
and crystal size. The present results are consistent with those ob- [7] M. Eddaoudi, D.B. Moler, H.L. Li, B.L. Chen, T.M. Reineke, M. O’Keeffe, O.M.
tained by high-throughput screening in synthesis of HKUST-1 Yaghi, Acc. Chem. Res. 34 (2001) 319.
[8] S. Kaskel, in: F. Schüth, K.S.W. Sing, J. Weitkamp (Eds.), Handbook of Porous
[48]. The BET surface areas measured by N2 adsorption isotherms Solids, vol. 2, Wiley-VCH, Weinheim, 2002, pp. 1190–1249.
however showed a non-linear increase with concentration. The [9] X. Zhao, B. Xiao, A. Fletcher, K.M. Thomas, D. Bradshaw, M.J. Rosseinsky,
BET surface areas (SBET) and pore volume of (1) were found to be Science 306 (2004) 1012.
[10] M. Latroche, S. Surblé, C. Serre, C. Mellot-Draznieks, P.L. Llewellyn, J.-H. Lee, J.-
1304, 1656, 1577 m2/g and 0.56, 0.81, 0.74 ml/g for concentrations S. Chang, S.H. Jhung, G. Férey, Angew. Chem. Int. Ed. 45 (2006) 8227.
of the reactants 1, 3 and 5 times of the original concentration, [11] M. Dincă, J.R. Long, J. Am. Chem. Soc. 127 (2005) 9376.
respectively. The increase in size from 610 to 20 lm on changing [12] J.S. Seo, D. Whang, H. Lee, S.I. Jun, J. Oh, Y.J. Jeon, K. Kim, Nature 404 (2000) 982.
[13] Y.K. Hwang, D.-Y. Hong, J.-S. Chang, S.H. Jhung, Y.-K. Seo, J. Kim, A. Vimont, M.
from 1 to 3 times concentration is consistent with an increase in Daturi, C. Serre, G. Férey, Angew. Chem. Int. Ed. 47 (2008) 4144.
the SBET from 1304 to 1656 m2/g. The subsequent increase in sur- [14] R. Kitaura, K. Seki, G. Akiyama, S. Kitagawa, Angew. Chem. Int. Ed. 42 (2003)
face area may be attributed to good crystallinity in 3 times en- 428.
[15] P. Horcajada, C. Serre, M. Vallet-Regí, M. Sebban, F. Taulelle, G. Férey, Angew.
hanced concentration. The surface area in the sample with 5 Chem. Int. Ed. 45 (2006) 5974.
times concentration decreased because of small sized spherical [16] S. Kitagawa, R. Kitaura, S.-I. Noro, Angew. Chem. Int. Ed. 43 (2004) 2334.
impurities of Cu2O (5–10 lm), which are conspicuous in Fig. 7f [17] S.S.-Y. Chui, S.M.-F. Lo, J.P.H. Charmant, A.G. Orpen, I.D. Williams, Science 283
(1999) 1148.
(see the arrow points).
[18] K. Schlichte, T. Kratzke, S. Kaskel, Micropor. Mesopor. Mater. 73 (2004) 81.
In spite of the formation of metal-organic frameworks, which [19] L. Alaerts, E. Séguin, H. Poelman, F. Thibault-Starzyk, P.A. Jacobs, E. De Vos,
could not be established, acceleration in nucleation and crystalliza- Chem. Eur. J. 12 (2006) 7353.
tion steps are very much noticeable in MW synthesis because of [20] H. Dathe, E. Peringer, V. Roberts, A. Jentys, J.A. Lercher, C.R. Chimie 8 (2005)
753.
the fast dissolution and deprotonation of the BTC–H3 and enhanced [21] A.R. Millward, O.M. Yaghi, J. Am. Chem. Soc. 127 (2005) 17998.
condensation of metal-oxygen networks under microwave-irradia- [22] Q.M. Wang, D. Shen, M. Bülow, M.L. Lau, S. Deng, F.R. Flitch, N.O. Lemcoff, J.
tion condition [38,39] by hot spots and superheating effects. Semanscin, Micropor. Mesopor. Mater. 55 (2002) 217.
[23] P. Krawiec, M. Kramer, M. Sabo, R. Kunschke, H. Fröde, S. Kaskel, Adv. Eng.
Mater. 8 (2006) 293.
[24] A. Vishnyakov, P.I. Ravikovitch, A.V. Neimark, M. Bulow, Q.M. Wang, Nano Lett.
4. Conclusions 3 (2003) 713.
[25] M. Hartmann, S. Kunz, D. Himsl, O. Tangermann, Langmuir 28 (2008) 8634.
We have successfully synthesized three different phases of [26] U. Mueller, M. Schuber, F. Teich, H. Puetter, K. Schierle-Arndt, J. Pastré, J. Mater.
Chem. 16 (2006) 626.
Cu(II)-(BTC–H3) system, starting from identical reaction mixtures
[27] A.K. Cheetham, C.N.R. Rao, R.K. Feller, Chem. Commun. (2006) 4780.
of Cu(NO3)2 and BTC–H3 by irradiating them in MW for different [28] J. Chen, T. Yu, Z. Chen, H. Xiao, G. Zhou, L. Weng, B. Tu, D. Zhao, Chem. Lett. 32
reaction temperatures from 140 to 170 °C, in a 1:1 water:ethanol (2003) 590.
[29] J. Gascon, S. Aguado, F. Kapteijn, Micropor. Mesopor. Mater. 113 (2008) 132.
mixture. The products (1–3) could also be obtained by running
[30] Y.K. Hwang, J.-S. Chang, S.-E. Park, D.S. Kim, Y.-U. Kwon, S.H. Jhung, J.-S.
the reaction at a fixed temperature of 140 °C in MW and varying Hwang, M.S. Park, Angew. Chem. Int. Ed. 44 (2005) 556.
the water:ethanol ratio only. The procedure gives a highly useful [31] S. Komarneni, R.K. Rajha, H. Katuski, Mater. Chem. Phys. 61 (1999) 50.
MOF material Cu3(BTC)2 (1) in its dehydrated form, in a consider- [32] G. Tompsett, W.C. Conner, K.S. Yngvesson, Chem. Phys. Chem. 7 (2006) 296.
[33] W.S. Ahn, K.K. Kang, K.Y. Kim, Catal. Lett. 72 (2001) 229.
ably lesser time in comparison to CE. Various factors which influ- [34] S.-E. Park, J.-S. Chang, Y.K. Hwang, D.S. Kim, S.H. Jhung, J.-S. Hwang, Catal.
ence the yield, quality and morphology of the product (1) have Survey Asia 8 (2004) 91.
also been investigated and optimized. Efforts are currently being [35] Y.K. Hwang, J.-S. Chang, Y.-U. Kwon, S.-E. Park, Micropor. Meospor. Mater. 68
(2004) 21.
made to adapt the procedure of this batch method for a continuous [36] S.H. Jhung, J.-S. Chang, Y.K. Hwang, S.-E. Park, J. Mater. Chem. 14 (2004) 280.
synthesis of (1) Cu3(BTC)2 with a kilogram scale in an hour. [37] Y.K. Hwang, T.-H. Jin, J.M. Kim, Y.-U. Kwon, S.-E. Park, J.-S. Chang, J. Nanosci.
Nanotechnol. 6 (2006) 1786.
[38] S.H. Jhung, T. Jin, Y.K. Hwang, J.-S. Chang, Chem. Eur. J. 13 (2007) 4410.
Acknowledgments [39] S.H. Jhung, J.-H. Lee, J.W. Yoon, C. Serre, G. Férey, J.-S. Chang, Adv. Mater. 19
(2007) 121.
[40] S.H. Jhung, J.-H. Lee, P.M. Forster, G. Férey, A.K. Cheetham, J.-S. Chang, Chem.
This work was supported by MKE through the Research Center Eur. J. 12 (2006) 7899.
for Nanocatalysts and Institutional Research Program. Dr. Geeta [41] Z. Ni, R.I. Masel, J. Am. Chem. Soc. 128 (2006) 12394.
Hundal gratefully acknowledges KOFST for awarding the Brain Pool [42] S.H. Jhung, J.-H. Lee, H.-S. Chang, Bull. Kor. Chem. Soc. 26 (2005) 880.
[43] E.R. Parnham, R.E. Morris, Acc. Chem. Res. 40 (2007) 1005.
fellowship at KRICT, Korea.
[44] R. Pech, J. Pickardt, Acta Cryst. C44 (1988) 992.
[45] Q.-W. Zhang, G.-X. Wang, Z. Krist.-New Cryst. Struc. 221 (2006) 101.
References [46] D.H. Williams, I. Fleming, Spectroscopic Methods in Organic Chemistry, fourth
ed., McGraw-Hill Publishers, UK, 1988.
[1] O.M. Yaghi, M. O’Keeffe, N.W. Ockwig, H.K. Chae, M. Eddaoudi, J. Kim, Nature [47] K. Nakamoto, Infrared and Raman Spectra of Inorganic and Coordination
423 (2003) 705–714. Compounds, fourth ed., John-Wiley & Sons Publishers, New York, 1986.
[2] S. Kitagawa, R. Matsuda, Coord. Chem. Rev. 251 (2007) 2490. [48] E. Biemmi, S. Christain, N. Stock, T. Bein, Micropor. Mesopor. Mater. 117 (2009)
[3] G. Férey, Chem. Soc. Rev. 37 (2008) 191. 111.

You might also like