Вы находитесь на странице: 1из 3

ChemComm Dynamic Article Links

Cite this: Chem. Commun., 2012, 48, 10642–10644

www.rsc.org/chemcomm COMMUNICATION
Instant MOFs: continuous synthesis of metal–organic frameworks by
rapid solvent mixingw
Miquel Gimeno-Fabra,a Alexis S. Munn,b Lee A. Stevens,a Trevor C. Drage,a
David M. Grant,a Reza J. Kashtiban,c Jeremy Sloan,c Edward Lester*a and
Richard I. Walton*b
Received 22nd June 2012, Accepted 7th September 2012
DOI: 10.1039/c2cc34493a

A continuous flow reactor allows the preparation of porous metal– such as microwaves8 and ultrasound,9 which often give
organic framework materials with crystallisation induced by rapid enhanced crystallisation rates, in some cases allowing some
mixing of streams of preheated water and solutions of reagents in control over the form of the product, for example, giving
organic solvent: this gives high volume production (132 g h!1) with nanocrystals of MOFs. Other methods that may provide
crystallite size of the products from nanoscale to micron. versatile approaches for preparing the materials include electro-
chemical deposition of films,10,11 and direct mechanochemical
Metal–organic frameworks (MOFs) are extended structures of reaction in the solid-state,12 which eliminates solvent, thereby
metal ions or groups of metals (clusters, chains or sheets) simplifying processing of products. These new methods, how-
connected by polydentate organic ligands.1 In many cases they ever, are not always applicable to all combinations of metal
present architectures with considerable porosity once any precursors and ligand. In this communication we report the use
occluded solvent is removed. MOFs are thus attracting intense of a high-throughput, continuous-flow method that allows rapid
interest as highly functionalisable porous hosts to rival zeolites, synthesis by inducing crystallisation from a stream of dissolved
with potential applications in fields such as gas adsorption, reagents by rapid mixing with a pre-heated water flow. The
storage and sensing;2 catalysis;3 and drug delivery.4 One desir- technology has already been proved suitable for the controlled
able characteristic of MOFs is their ability to adsorb gases such production of various inorganic nano-materials, such as oxide
as CO2, H2 and CH4.5 This makes them interesting industrially, particles.13 We herein demonstrate its first use in MOF produc-
where they may engender novel approaches to energy applica- tion by study of two porous frameworks that are of importance
tions and to environmental remediation technologies.6 for their properties in gas storage and separation.
MOFs are typically synthesised using batch solvothermal Fig. 1 shows a schematic of the reactor design. This setup uses
reactions whereby solutions of chemical precursors are mixed a ‘‘counter current mixing reactor’’,14 consisting of a pipe-in-pipe
and heated in a vessel, with pressure generated if sealed, for concentric arrangement, where preheated water is introduced
periods of hours or even days.7 This synthesis approach is through the inner pipe (downflow), meeting the precursor stream
versatile due to the large combination of possible organic (upflow, injected at room temperature) at the mixing point.
ligands and metal salt precursors, the choice of solvent and
the range of temperatures possible; yet it is not always possible
to control crystal growth since a given material may only
crystallise under certain, precise combinations of reaction
conditions. There have been a number of recent reports of novel
approaches to MOF synthesis that overcome some of the limita-
tions of the simple batch solvothermal approach. This includes
the use of energy sources other than conventional heating,

a
Department of Chemical and Environmental Engineering,
University of Nottingham, University Park, Nottingham, NG9 2RD,
UK. E-mail: edward.lester@nottingham.ac.uk;
Tel: +44 (0)115 951 4974
b
Department of Chemistry, University of Warwick, Coventry,
CV4 7AL, UK. E-mail: r.i.walton@warwick.ac.uk;
Tel: +44 (0)2476523241
c
Department of Physics, University of Warwick, Coventry,
CV4 7AL, UK
w Electronic supplementary information (ESI) available: Further
XRD, SEM and TEM analysis, and results of BET and TGA. See Fig. 1 Schematic representation of the continuous synthesis system
DOI: 10.1039/c2cc34493a used for the instant production of MOFs.

10642 Chem. Commun., 2012, 48, 10642–10644 This journal is c The Royal Society of Chemistry 2012
This arrangement facilitates instant, powerful and uniform
mixing of the reactant streams and short residence times. The
final suspension of particles is evacuated through the top of
the reactor; this is then rapidly cooled as it passes through a
water jacket cooler and is collected after a manual back
pressure regulator set to maintain the system at 250 bar.
The first material we consider is the copper(II) carboxylate
HKUST-1. This material has composition Cu3(BTC)2(H2O)"xH2O
(BTC = 1,3,5-benzenetricarboxylate) and has been one of the
most extensively studied MOFs since being reported by Chui
et al. in 1999.15 It presents high chemical stability, high surface
Fig. 3 SEM images of HKUST-1 prepared at 300 1C from reagent
area and pore volume together with coordination properties
solutions of (a) 0.03 M Cu2+ and (b) with 0.0075 M Cu2+.
towards various adsorbate gases.16 The material was also the first
MOF to be manufactured on an industrial scale, an electro- Table 1 Textural properties of HKUST-1 and Basolite C300 (Sigma-
chemical process with reaction times of 3 to 12 hours,11 and has Aldrich). Vp is total pore volume at 0.95 p/po, and Vm micropore
been commercialised under the name of Basolites C300. Powder volume by t-plot analysis (Harkins and Jura)
X-ray diffraction of a typical as-prepared HKUST-1 product
Sample BETSA (m2 g!1) Vp (cm3 g!1) Vm (cm3 g!1)
from the continuous reactorz is shown in Fig. 2, confirming the
high crystallinity and phase-purity of the material, even after a HKUST-1 1950 0.80 0.77
Basolite C300 1694 0.72 0.70
residence time at the point of reaction of less than 1 s. Batch
synthesis of HKUST-1 is typically performed up to 150 1C,7 but
using a downflow of water at 300 1C allows rapid crystallisation. and a commercially available sample of the same composition
The powder X-ray pattern for material prepared at a reaction (Basolite C300) have Type I isotherms, indicative of micro-
temperature of 400 1C, shows collapse to Cu2O, indicating an porous materials (see ESIw). Both samples have equivalent
upper limit to reaction temperature (see ESIw). SEM images of textural properties, Table 1, although our material has BET
the materials prepared at 300 1C show some variation in crystal and Langmuir surface areas that exceed the commercial
size and shape when the dilution and flow rate of the reagents is sample and are at the high end of the range of values typically
changed, Fig. 3, with generally larger crystals seen at lower reported for this material.16 Adsorption studies of other gases
dilution. It is noteworthy that the crystal morphology is not carried by thermogravimetric analysis also indicates that
typical of the octahedral crystals usually reported for HKUST-1.7 HKUST-1 has similar adsorption properties and performance
Reported in situ studies of HKUST-1 crystallisation show to the commercially available material, with typical CO2
that nucleation occurs instantaneously but continues late into uptake (298 K, 1 bar) of 25% by mass, which compares well
the crystal growth period;17 indeed a range of crystal sizes are with reported values for samples prepared in batch (ESIw).
often formed in samples from batch synthesis, consistent with The second MOF we studied was the Ni(II) carboxylate
the continuous nucleation for crystal growth. We can thus CPO-27.18 This is constructed from edge-shared metal octahedra
propose that an advantage of our synthesis lies in the short linked by 2,5-dihydroxycarboxylate ligands to give an open
period of crystallisation: this prevents reaction before the structure with B11 Å diameter, one-dimensional channels and
heated downflow is encountered and then crystal growth accessible metal sites. The structure is also known for other
occurs only at the nucleation sites formed during the residence metals, including Fe, Co and Zn, and these various forms have
at the reaction point. Textural characterisation by N2 sorption been used in gas storage, e.g., for dihydrogen,18 H2S,19 selec-
isotherms using a ASAP 2420 instrument show both HKUST tive binding of oxygen over nitrogen,20 and various toxic
gases.21 We effected crystallisation by using a similar approach
as for HKUST: the metal salt (Ni(II) acetate) and organic
linker dissolved in DMF (upflow) and crystallisation induced
by mixing with pre-heated water (downflow).2 Powder XRD,
Fig. 4, of the CPO-27 samples confirms the sample identity.
All peaks in the powder XRD can be indexed on the published
crystal structure18 but there is significant peak broadening due to
small particle size. Scherrer analysis gives 12 nm particle diameter
for the material prepared at the highest concentration of reagents,
and 15 nm for those prepared at the lower dilution. Fig. 5 shows
results of electron microscopy of materials prepared at 300 1C at
two different dilutions of reagents. Nanocrystalline powders are
formed at the highest concentration at 300 1C, but with increasing
dilution at 200 1C, larger, needle-like crystallites are seen, con-
sistent with the morphology seen in batch reactions using various
heating methods (conventional, microwave and ultrasound).22
Fig. 2 Profile fit (Fm3% m) to powder X-ray pattern of HKUST-1 Even the larger crystals are less than 20 nm in cross section,
sample prepared at 300 1C with 0.03 M Cu2+. consistent with the broadened diffraction profile.

This journal is c The Royal Society of Chemistry 2012 Chem. Commun., 2012, 48, 10642–10644 10643
organic linker and metal precursors, with the potential for
fine tuning of crystal morphology for scalable production.

Notes and references


z Continuous synthesis of HKUST-1: Downflow: 20 mL min!1 deionized
water (at 100–400 1C). Upflow: 10 mL min!1 benzene-1,3,5-tricarboxylic
acid (4.20 g, 0.03 M, Aldrich) and copper(II) nitrate hemipentahydrate
(4.82 g, 0.03 M, Aldrich). These were stirred for 15 minutes in 660 mL
of solvent consisting of equal parts of N,N 0 -dimethylformamide
(Fluka) and ethanol (Fluka) before being pumped into the system.
Continuous synthesis of CPO-27(Ni): Downflow: 20 mL min!1 deionized
water (at 100–400 1C). Upflow: 10 mL min!1 nickel acetate (3.19 g,
0.15 M, Aldrich) and 2,5 dihydroxyterephthalic acid (1.26 g, 0.075 M,
Aldrich) were dissolved in at least 85 ml N,N 0 -dimethylformamide
(Fluka). In both cases the solid was allowed to settle and the cloudy
concentrate separated by decantation and dried at 70 1C.
1 See, for example: C. J. Kepert, in Porous Materials, ed. D. O’Hare,
D. W. Bruce and R. I. Walton, John Wiley and Sons, Chichester,
Fig. 4 Profile fits (R3% ) to powder X-ray diffraction pattern of CPO- 2010; Functional Metal-Organic Frameworks: Gas Storage, Separa-
tion and Catalysis, ed. M. Schröder, Springer, Heidelberg, 2010;
27(Ni) prepared at two different temperatures and dilutions. Note the
Metal-Organic Frameworks Design and Application, ed. L. R.
effect of particle morphology on preferred orientation. MacGillivray, John Wiley and Sons, Hoboken, 2010;
D. Farrusseng, Metal-Organic Frameworks, Wiley-VCH,
Weinheim, 2011; H. Zhou, J. R. Long and O. M. Yaghi, Chem.
Rev., 2012, 112, 673(MOF Special Issue).
2 J.-R. Li, R. J. Kuppler and H.-C. Zhou, Chem. Soc. Rev., 2009,
38, 1477.
3 D. Farrusseng, S. Aguado and C. Pinel, Angew. Chem., Int. Ed.,
2009, 48, 7502.
4 P. Horcajada, R. Gref, T. Baati, P. K. Allan, G. Maurin,
P. Couvreur, G. Férey, R. E. Morris and C. Serre, Chem. Rev.,
2012, 112, 1232–1268.
5 M. Shah, M. C. McCarthy, S. Sachdeva, A. K. Lee and
H. K. Jeong, Ind. Eng. Chem. Res., 2012, 51, 2179.
6 A. U. Czaja, N. Trukhan and U. Müller, Chem. Soc. Rev., 2009,
38, 1284.
7 N. Stock and S. Biswas, Chem. Rev., 2012, 112, 933–969.
8 S. H. Jhung, J.-H. Lee and J.-S. Chang, Bull. Korean Chem. Soc.,
2005, 26, 880; J. Klinowski, F. A. A. Paz, P. Silva and J. Rocha,
Dalton Trans., 2011, 40, 321.
9 W. J. Son, J. Kim and W. S. Ahn, Chem. Commun., 2008, 6336.
10 A. Martinez-Joaristi, J. Juan-Alcañiz, P. Serra-Crespo, F. Kapteijn
and J. Gascon, Cryst. Growth Des., 2012, 12, 3489–3498.
11 U. Müller, H. Puetter, M. Hesse and H. Wessel, International
Patent, WO 2005/049892, 2005.
12 T. Friščić, J. Mater. Chem., 2010, 20, 7599.
13 E. Lester, P. Blood, J. Denyer, D. Giddings, B. Azzopardi and
M. Poliakoff, J. Supercrit. Fluids, 2006, 37, 209; E. Lester,
G. Aksomaityte, J. Li, S. Gomez, J. Gonzalez-Gonzalez and
M. Poliakoff, Prog. Cryst. Growth Charact. Mater., 2012, 58, 3.
14 E. Lester and B. Azzopardi, International Patent, WO/2005/077505,
2005.
15 S. S. Y. Chui, S. M. F. Lo, J. P. H. Charmant, A. G. Orpen and
Fig. 5 Electron microscopy of CPO-27(Ni) (a) HR-TEM of sample I. D. Williams, Science, 1999, 283, 1148.
prepared at 300 1C at high concentration. Aligned pores viewed along 16 J. C. Liu, J. T. Culp, S. Natesakhawat, B. C. Bockrath, B. Zande,
S. G. Sankar, G. Garberoglio and J. K. Johnson, J. Phys. Chem. C,
[001] are visible at I and detail (I 0 ). FFT of I is inset top left, HR-TEM 2007, 111, 9305.
simulation and model are inset middle right and bottom. (b) SEM 17 D. Zacher, J. N. Liu, K. Huber and R. A. Fischer, Chem.
image of 200 1C sample prepared at lower concentration. Commun., 2009, 1031; F. Millange, M. Medina, N. Guillou,
G. Férey, K. M. Golden and R. I. Walton, Angew. Chem., Int. Ed.,
2010, 49, 763.
The textural properties of the CPO-27(Ni), as studied by
18 P. D. C. Dietzel, P. A. Georgiev, J. Eckert, R. Blom, T. Strässle
BET nitrogen sorption analysis, show a typical Type I iso- and T. Unruh, Chem. Commun., 2010, 4962.
therm with a slight mesopore character at high relative pres- 19 P. K. Allan, P. S. Wheatley, D. Aldous, M. I. Mohideen, C. Tang,
sure (ESIw). A surface area of 1030 m2 g!1 and total and J. A. Hriljac, I. L. Megson, K. W. Chapman, G. De Weireld,
S. Vaesen and R. E. Morris, Dalton Trans., 2012, 41, 4060.
micropore volume 0.67 and 0.40 cm3 g!1 respectively, are 20 E. D. Bloch, L. J. Murray, W. L. Queen, S. Chavan,
calculated. A CO2 uptake of 19.57 wt% was measured. These S. N. Maximoff, J. P. Bigi, R. Krishna, V. K. Peterson,
values match reported data for samples from conventional F. Grandjean, G. J. Long, B. Smit, S. Bordiga, C. M. Brown
synthesis. The method of synthesis we report should be and J. R. Long, J. Am. Chem. Soc., 2011, 133, 14814.
21 T. G. Glover, G. W. Peterson, B. J. Schindler, D. Britt and
applicable to a wide range of MOFs, given the scope for O. Yaghi, Chem. Eng. Sci., 2011, 66, 163.
varying solvent or solvent mixtures along with choice of 22 E. Haque and S. H. Jhung, Chem. Eng. J., 2011, 173, 866.

10644 Chem. Commun., 2012, 48, 10642–10644 This journal is c The Royal Society of Chemistry 2012

Вам также может понравиться