Sunteți pe pagina 1din 4

LETTERS

PUBLISHED ONLINE: 3 MAY 2009 | DOI: 10.1038/NMAT2445

Controlling interpenetration in metalorganic


frameworks by liquid-phase epitaxy
Osama Shekhah1 , Hui Wang1 , Markos Paradinas2 , Carmen Ocal2 , Bjrn Schpbach3 ,
Andreas Terfort3 , Denise Zacher4 , Roland A. Fischer4 and Christof Wll1 *
Metalorganic frameworks (MOFs) are highly porous materials
generally consisting of two building elements: inorganic
coupling units and organic linkers14 . These frameworks offer
an enormous porosity, which can be used to store large
amounts of gases and, as demonstrated in more recent
applications5,6 , makes these compounds suitable for drug
release. The huge sizes of the pores inside MOFs, however, also
give rise to a fundamental complication, namely the formation
of sublattices occupying the same space. This interpenetration
greatly reduces the pore size and thus the available space
within the MOF structure7 . We demonstrate here that the
formation of the second, interpenetrated framework can be
suppressed by using liquid-phase epitaxy on an organic
template. This success demonstrates the potential of the
step-by-step method to synthesize new classes of MOFs not
accessible by conventional solvothermal methods.
Metalorganic frameworks (MOFs) were developed to provide
the largest amount of coordination space possible using the least
amount of material. In general, inorganic coupling units are
combined with organic linkers, yielding molecular cages with
vertices defined by the ligands and corners defined by the inorganic
coupling units14 . Whereas the first application proposed for
these highly porous materials was the storage of gas molecules,
in particular hydrogen4 , the fascinating properties of this new
class of materials have inspired a huge variety of other potential
applications, including separation8,9 , catalysis10 , drug release5,6
and the embedding of metal nanoparticles for applications in
catalysis and sensor technology1113 . Several applications require
a controlled deposition of MOFs on supports. This is a difficult
task considering the solvothermal synthesis conditions and has only
recently been accomplished1416 .
The two key parameters for comparing MOF materials are
the volume of a single pore in the network and the length
of the backbones defining the framework. Recent intense effort
has led to the development of MOFs with a surface area of
5,900 m2 g1 (ref. 17), making these inorganicorganic hybrids
the most highly porous material available. For many researchers,
the most compelling challenge with regard to MOF synthesis is
probing the limits of this molecular architecture. One goal, for
example, is to build a material with the lowest density possible,
yet that is still rigid enough to survive the removal of the solvent
molecules in which the synthesis is carried out. In principle, the
simplest way to enlarge the cavities or channels of a MOF with a
given topology is to increase the length of the rigid organic linkers
without changing the coupling units, also called the secondary
building units4 (SBUs). The coupling chemistry of the linkers to the

Interpenetrated network

Bulk
synthesis

Epitaxy

SURMOF

Figure 1 | Representation of MOF synthesis concepts. Using conventional


synthesis, often two equivalent networks (coloured red and green) are
formed at the same time. Using liquid-phase epitaxy, the equivalence of
these two networks is lifted by the presence of the substrate (dotted line in
the schematic diagram on the right) and the formation of interpenetrated
networks is suppressed, yielding SURMOFs containing only one network.

SBUs should also remain unchanged. This approach is, however,


hampered by the fact that the extremely high porosity of MOFs also
gives rise to a fundamental complication: if the voids become large
enough, the pores in the MOFs can accommodate the SBUs and the
organic linkers in addition to the guest molecules. These units may
then form a second (or higher) network coexisting with the first
(see Fig. 1)7,8,18 . The presence of such interpenetrating networks
greatly reduces the pore size and thus the available space within
the MOF and leads in many cases to the collapse of the network
on solvent removal3,7 .
Not surprisingly, structures favouring interpenetration are
commonly favoured for simple cubic networks as well as for
sterically non-demanding SBUs and linkers19 . A well-known
example is the isoreticular series of MOFs called IRMOF-n, which
are based on IRMOF-1 consisting of the carboxylate-bridged
unit [Zn4 O]6+ ligated with 1,4-benzenedicarboxylate (bdc). Using
4,4-biphenyldicarboxylate (bpdc) instead of bdc leads to the
formation of interpenetrated [Zn4 O(bpdc)3 ] (IRMOF-9; ref. 20).
Even the phase-pure synthesis of the parent compound IRMOF-1
is complicated by interpenetration21 , and even the recently
reported IRMOF-0, [Zn4 O(adc)3 ], (adc: acetylenedicarboxylic
acid) apparently exists only as an interpenetrated structure22 .

1 Ruhr-Universitt

Bochum, Lehrstuhl fr Physikalische Chemie 1, 44780 Bochum, Germany, 2 Institut de Cincia de Materials de Barcelona, Campus de la
UAB, 08193 Bellaterra, Spain, 3 Philipps-Universitt Marburg, Fachbereich Chemie, 35032 Marburg, Germany, 4 Ruhr-Universitt Bochum, Organometallics
and Materials Chemistry, 44780 Bochum, Germany. *e-mail: woell@pc.rub.de.
NATURE MATERIALS | VOL 8 | JUNE 2009 | www.nature.com/naturematerials

2009 Macmillan Publishers Limited. All rights reserved.

481

LETTERS

NATURE MATERIALS DOI: 10.1038/NMAT2445


001

38.0

Out-of-plane XRD

SURMOF-1 on PBMT SAM

37.8

MOF-508a
MOF-508b
SURMOF-1 on PBMT SAM
B

37.2

B
36.8 A
36.6

24,000

002

37.0

36.4

I (a.u.)

RIU 103

37.6
37.4

10,000

20,000
Time (s)

26,000

28,000

30,000

30,000

40,000

10

Figure 2 | SPR signal as a function of time recorded in situ during


sequential injections. Zinc acetate (A), ethanol (B) and H2 bdc + 4,40 -bipy
(C) in the SPR cell containing a pyridine-terminated SAM substrate.

Nevertheless, for a given combination of SBUs and linkers, it


is sometimes possible, but difficult, to suppress interpenetration
by varying the solvothermal reaction conditions in a trial and
error fashion. For example, one can choose a different solvent,
a removable template or work at very high dilution. Researchers
should bear in mind that the MOF designer is always at the mercy
of intermolecular forces that limit predictability4 .
From a topological point of view, interpenetration is hard to
avoid because the different networks bear translational symmetry
and are therefore equivalent within the bulk. Here, we propose a
novel approach to prevent this interpenetration by using liquidphase epitaxial growth on a two-dimensional (2D) template organic
surface. This approach was shown to yield very homogeneous films
of [Cu3 (btc)2 ] (HKUST-1; btc: 1,3,5-benzenetricarboxylate)14 . The
presence of an appropriately functionalized substrate breaks the
symmetry of the bulk MOF, rendering the different interpenetrating
networks inequivalent, as illustrated in Fig. 1. Only the network
having a plane of coupling units in common with the organic
surface of the substrate is favoured, whereas the growth of other
networks is strongly suppressed.
We demonstrate this novel principle for the case of the
microporous MOF-508 [Zn(bdc)-(4, 40 -bipy)0.5 ] (4,40 -bipy:
4,40 -bipyridine). This MOF is conventionally synthesized by a
solvothermal reaction of H2 bdc, 4,40 -bipy and Zn(NO3 )2 6H2 O in
dimethylformamide/ethanol at 90 C for 24 h (ref. 23). Depending
on the presence or removal of the solvent, one of the two known
MOF polymorphs (MOF-508a and MOF-508b) is obtained, both of
which consist of two interpenetrating, pillared, paddle-wheel-type
networks23 . Interpenetration results from 2D zinc carboxylate
layers, which have a primitive square framework, held apart by
4,40 -bipy poles. In this way, 1D channels are formed with a
width of 11.0 11.0 2 , permitting a double interpenetration that
results in four much smaller 1D channels each of 4.0 4.0 2
in cross-section23 .
The unconventional approach proposed here to obtain the so
far unknown, non-interpenetrated version of MOF-508 is not
based on a homogeneous mixture of the reactants in solution but
instead the ethanolic solutions of the components. Zinc acetate
and the mixture of the organic ligands (H2 bdc + 4,40 -bipy) were
kept separated in two beakers and an appropriately functionalized
organic surface was immersed alternately in the two different
solutions with intermittent rinsing14,24 . The organic surface was
prepared by fabricating a self-assembled monolayer (SAM) from
4,4-pyridyl-benzenemethanethiol (PBMT) on an Au substrate.
The surface of this SAM mimics the upper rim of a layer of
4,40 -bipyridine poles in the MOF-508a structure and thus should
permit the epitaxial growth of this material.
482

20

30

2 ()

Figure 3 | Out-of-plane XRD pattern. [Zn(bdc)-(bipy)0.5 ] SURMOF-1


sample (40 cycles) grown on a pyridine-terminated SAM from PBMT (in
blue). The XRD patterns for the two possible bulk phases are shown
for comparison.

As demonstrated by the surface plasmon resonance spectroscopy


(SPR) data shown in Fig. 2, the step-by-step synthesis leads
to a layer-by-layer growth of an organicinorganic polymer
on the substrate (a surface-mounted metalorganic framework,
SURMOF). Figure 2 clearly shows that after an initiating period,
each step of this liquid-phase epitaxy leads to roughly the same
amount of deposited material. The growth of this SURMOF on
the PBMT SAM was also characterized using infrared reflection
absorption spectroscopy: the spectrum shows an increase in the
intensity of the characteristic bands, mainly the anti-symmetric and
symmetric stretching bands of the carboxylate groups of the SBUs of
the SURMOF with increasing the number of immersion cycles. The
infrared reflection absorption spectroscopy spectrum also shows
the absence of any bands of the solvent (ethanol) or unreacted
species of H2 bdc and 4,40 bipy or Zn(OH)2 in the SURMOF (see
Supplementary Information S1).
To demonstrate that the deposited organicinorganic polymer
corresponds to a highly ordered MOF material, X-ray diffraction
(XRD) was carried out. In Fig. 3, top, we show the corresponding
diffraction pattern recorded in the out-of-plane diffraction modus
for MOF material of the likely composition [Zn(bdc)(4,40 -bipy)0.5 ]
with 40 deposition cycles deposited on the pyridine-terminated
organic surface. The two sharp diffraction peaks indicate the
presence of highly oriented material with a large lattice constant.
We assign the two diffraction peaks based on the corresponding
XRD bulk data for the two different polymorphs of MOF-508,
which are shown in the bottom of Fig. 3. Clearly, the positions
of the two Bragg peaks in the XRD data of the SURMOF fit
only with MOF-508a.
However, when comparing the experimental XRD pattern for
the SURMOF (Fig. 4, top) to a simulation of the diffraction pattern
expected for an oriented MOF layer of the type MOF-508a on
the surface (Fig. 4, middle), we see a major deviation: the relative
intensities of the [100] and [200] peaks seen in the SURMOF XRD
data are greatly different from those seen for MOF-508a (ref. 15).
Again note, that infrared spectroscopy reveals that the constituents
of this lattice are the same units as present in MOF-508a; the
presence of solvent and other adsorbed species can be excluded.
As the thickness of the SURMOF is too small to enable a direct
structure determination using XRD, we have further characterized
the SURMOF by determining its porosity through measuring the
Kr BrunauerEmmettTeller (BET) surface area. To this end, we
have first determined the thickness of a 55-cycle SURMOF (one
cycle consists of immersion in the Zn acetate and then in the
H2 bdc + 4,40 -bipy mixture), using atomic force microscopy. By

NATURE MATERIALS | VOL 8 | JUNE 2009 | www.nature.com/naturematerials


2009 Macmillan Publishers Limited. All rights reserved.

LETTERS

NATURE MATERIALS DOI: 10.1038/NMAT2445

SURMOF experimental

I (a.u.)

Calculated
no interpenetration

H
C
N

001

002

Zn

Calculated MOF-508a
5

10

15

20

2 ()

Figure 4 | Out-of-plane XRD pattern. Bottom: Calculated diffraction


pattern expected for an oriented MOF layer of type MOF-508a on the
surface including interpenetration. Middle: Calculated diffraction pattern
expected for an oriented MOF layer of type MOF-508a (no
interpenetration) on the surface. Top: SURMOF-1 sample (40 cycles) grown
on a pyridine-terminated SAM.

deliberately removing material, the thickness of the SURMOF


was found to be around 40 nm (data shown in Supplementary
Information); that is, one deposition cycle results in a thickness
gain of half a unit cell. The atomic force microscopy data show
a rather flat, homogeneous surface similar to data reported for a
previous liquid epitaxy growth study for another (non-penetrating)
MOF system, HKUST-I (ref. 24). Using this information for the
layer thickness, we obtain a value of 627 15 m2 cm3 for the inner
surface of the SURMOF from the Kr BET measurement, which is
slightly more than half the Kr BET value for the interpenetrated
MOF-508a, 930 15 m2 cm3 (the corresponding N2 BET value
amounts to 821 m2 cm3 ; ref. 25).
The findings reported above demonstrate that our novel
synthesis route yields a crystalline framework with the same
composition and unit-cell parameters as MOF-508a but with only
about half the inner surface/volume area. In addition the SURMOF
is solvent-free. The simplest explanation for these findings is
that the SURMOF is the non-interpenetrated and solvent-free
analogue of MOF-508a. A proposed model structure for such a
highly porous structure is shown in Fig. 5. A calculation of the
corresponding XRD diffraction peak intensities reveals an excellent
agreement (see Fig. 4). The fact that the BET surface per volume
for the interpenetrated and desolvated MOF-508a and MOF-508b
is less than double that of the non-interpenetrated and solventfree framework of our SURMOF is expected, because the two
interpenetrating networks in MOF-508a and MOF-508b will be so
close in space, that the inner surface available for Kr adsorption will
be less than double the value of a single framework. When converted
to surface area per weight, we obtain values of 1,010 m2 g1 , which
is substantially more than the value of 660 m2 g1 (ref. 25) for the
interpenetrated MOF-508a and MOF-508b.
We explain the suppression of the second, interpenetrating
lattice in the SURMOF by the pyridine-terminated organic surface
acting as a nucleation template. The other sub-lattice does not
match with this template (see Fig. 1), and it therefore cannot
nucleate at the surface. Our results suggest that nucleation of
the second lattice at later points in time of the liquid-phase
epitaxy synthesis is not sufficient to enable the formation of
the interpenetrated sublattice, which would yield MOF-508a. In
addition, we point out that we used Zn acetate as the source
for the Zn2+ rather than Zn(NO3 )2 6H2 O. Zn acetate consists of
paddle-wheel units [Zn2 (CH3 COO)4 (H2 O)2 ]. Thus, the SURMOF
growth takes advantage of the so-called controlled SBU approach
that is typically used in the synthesis of MILs (ref. 17).

Figure 5 | Proposed model structure. A solvent-free, non-penetrated


pseudomorph of MOF-508a [Zn(bdc)-(4,40 -bipy)0.5 ], SURMOF-1 grown
along the [001] direction.

In future work it will be investigated whether liquid-phase


epitaxy is a general method to suppress interpenetration in MOFs
with large pores. By further analysing and optimizing the growth
conditions, it may become possible to use the surfaces of small
particles, including MOF particles, for the deposition of SURMOFs.
Received 3 November 2008; accepted 8 April 2009;
published online 3 May 2009

References
1. Ferey, G. Microporous solids: From organically templated inorganic
skeletons to hybrid frameworks... Ecumenism in chemistry. Chem. Mater. 13,
30843098 (2001).
2. Hoskins, B.-F. & Robson, R. Design and construction of a new
class of scaffolding-like materials comprising infinite polymeric
frameworks of 3-D-linked molecular rodsa reappraisal of the
Zn(Cn)2 and Cd(Cn)2 structures and the synthesis and structure
of the diamond-related frameworks [N(Ch3 )4 ][Cuiznii(Cn)4 ] and
Cui[4,40 ,400 ,4000 -tetracyanotetraphenylmethane]Bf4.Xc6h5no2. J. Am.
Chem. Soc. 112, 15461554 (1990).
3. Kitagawa, S., Kitaura, R. & Noro, S. Functional porous coordination polymers.
Angew. Chem. Int. Ed. 43, 23342375 (2004).
4. Rowsell, J. & Yaghi, O. Metal organic frameworks, a new class of porous
materials. Microporous Mesoporous Mater. 73, 314 (2004).
5. Horcajada, P. et al. Flexible porous metalorganic frameworks for a controlled
drug delivery. J. Am. Chem. Soc. 130, 67746780 (2008).
6. Horcajada, P. et al. Metalorganic frameworks as efficient materials for drug
delivery. Angew. Chem. Int. Ed. 45, 59745978 (2006).
7. Yaghi, O.-M. A tale of two entanglements. Nature Mater. 6, 9293 (2007).
8. Snurr, R.-Q., Hupp, J.-T. & Nguyen, S.-T. Prospects for nanoporous
metalorganic materials in advanced separations processes. AICHE J. 50,
10901095 (2004).
9. Maji, T.-K., Uemura, K., Chang, H.-C., Matsuda, R. & Kitagawa, S. Expanding
and shrinking porous modulation based on pillared-layer coordination
polymers showing selective guest adsorption. Angew. Chem. Int. Ed. 43,
32693272 (2004).
10. Seo, J.-S. et al. A homochiral metalorganic porous material for enantioselective
separation and catalysis. Nature 404, 982986 (2000).
11. Hermes, S., Schroder, F., Amirjalayer, S., Schmid, R. & Fischer, R.-A. Loading of
porous metalorganic open frameworks with organometallic CVD precursors:
Inclusion compounds of the type [LnM](a)@MOF-5. J. Mater. Chem. 16,
24642472 (2006).
12. Hermes, S. et al. Metal@MOF: Loading of highly porous coordination polymers
host lattices by metal organic chemical vapor deposition. Angew. Chem. Int. Ed.
44, 62376241 (2005).
13. Allendorf, M.-D. et al. Stress-induced chemical detection using flexible
metalorganic frameworks. J. Am. Chem. Soc. 130, 14404 (2008).

NATURE MATERIALS | VOL 8 | JUNE 2009 | www.nature.com/naturematerials

2009 Macmillan Publishers Limited. All rights reserved.

483

LETTERS

NATURE MATERIALS DOI: 10.1038/NMAT2445

14. Shekhah, O. et al. Step-by-step route for the synthesis of metalorganic


frameworks. J. Am. Chem. Soc. 129, 1511815119 (2007).
15. Biemmi, E., Scherb, C. & Bein, T. Oriented growth of the metalorganic
framework Cu3 (BTC)2 (H2 O)3 xH2 O tunable with functionalized
self-assembled monolayers. J. Am. Chem. Soc. 129, 80548055 (2007).
16. Hermes, S., Schrder, F., Chelmowski, R., Wll, C. & Fischer, R.-A. Selective
nucleation and growth of metalorganic open framework thin films on
patterned COOH/CF3 -terminated self-assembled monolayers on Au(111).
J. Am. Chem. Soc. 127, 1374413745 (2005).
17. Ferey, G. et al. A chromium terephthalate-based solid with unusually large pore
volumes and surface area. Science 309, 20402042 (2005).
18. Batten, S.-R. & Robson, R. Interpenetrating nets: Ordered, periodic
entanglement. Angew. Chem. Int. Ed. 37, 14601494 (1998).
19. Eddaoudi, M. et al. Modular chemistry: Secondary building units as a basis for
the design of highly porous and roboust metalorganic carboxylate frameworks.
Acc. Chem. Res. 34, 319330 (2001).
20. Eddaoudi, M. et al. Systematic design of pore size and functionality in
isoreticular MOFs and their application in methane storage. Science 295,
469472 (2002).
21. Hafizovic, J. et al. The inconsistency in adsorption properties and powder XRD
data of MOF-5 is rationalized by framework interpenetration and the presence
of organic and inorganic species in the nanocavities. J. Am. Chem. Soc. 129,
36123620 (2007).
22. Tranchemontagne, D. J., Hunt, J. R. & Yaghi, O. M. Room temperature
synthesis of metalorganic frameworks: MOF-5, MOF-74, MOF-177,
MOF-199, and IRMOF-0. Tetrahedron 64, 85538557 (2008).

484

23. Chen, B.-L. et al. A microporous metalorganic framework for


gas-chromatographic separation of alkanes. Angew. Chem. Int. Ed. 45,
13901393 (2006).
24. Munuera, C., Shekhah, O., Wang, H., Wll, C. & Ocal, C. The controlled
growth of oriented metal organic frameworks on functionalized surfaces
as followed by scanning force microscopy. Phys. Chem. Chem. Phys. 10,
72577261 (2008).
25. Ma, B.-Q., Mulfort, K.-L. & Hupp, J.-T. Microporous pillared paddle-wheel
frameworks based on mixed-ligand coordination of zinc ions. Inorg. Chem. 44,
49124914 (2005).

Acknowledgements
We acknowledge financial support by the EU through the FP6 STREP
initiative SURMOF.

Author contributions
O.S. and H.W. prepared the SURMOFs investigated in this study. A.T. and B.S.
synthesized the organothiols used to fabricate the SAMs. M.P. and C.O. carried out the
AFM work. Data analysis was carried out by O.S., D.Z., R.F. and C.W. The work was
directed by O.S., R.F. and C.W. All authors contributed equally in writing the manuscript.

Additional information
Supplementary information accompanies this paper on www.nature.com/naturematerials.
Reprints and permissions information is available online at http://npg.nature.com/
reprintsandpermissions. Correspondence and requests for materials should be
addressed to C.W.

NATURE MATERIALS | VOL 8 | JUNE 2009 | www.nature.com/naturematerials


2009 Macmillan Publishers Limited. All rights reserved.

S-ar putea să vă placă și