Вы находитесь на странице: 1из 13

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

Metal-organic frameworks

Stuart L. James
School of Chemistry, David Keir Building, Queens University Belfast, Stranmillis Road, Belfast,
Northern Ireland UK BT9 5AG. E-mail: s.james@qub.ac.uk

Received 13th February 2003


First published as an Advance Article on the web 27th June 2003

Metal-organic frameworks are a recently-identified class of which has lead to their applications in detergents and water-
porous polymeric material, consisting of metal ions linked softeners. Advances in inorganic porous frameworks based on
together by organic bridging ligands, and are a new other elements have been reviewed recently.2a Such materials
Published on 27 June 2003 on http://pubs.rsc.org | doi:10.1039/B200393G

development on the interface between molecular coordina- include metal phosphates, sulfides and chlorides. An early
tion chemistry and materials science. A range of novel report also exists of the use of cyanide as a bridging group, to
structures has been prepared which feature amongst the form the mixed-metal open framework
Downloaded by University of Virginia on 08 September 2012

largest pores known for crystalline compounds, very high K2Zn3[Fe(CN)6]2.xH2O, in which tetrahedral Zn2+ ions link
sorption capacities and complex sorption behaviour not together octahedral hexacyanoferrates.2b The analogy between
seen in aluminosilicate zeolites. The development of syn- this compound and zeolites was pointed out by the authors, who
described the structure as having large cavities occupied by K+
thetic approaches to these materials and investigations of
counterions and water molecules. Since the early 1990s,
their properties are reviewed.
research into materials with polymeric, sometimes porous,
structures based on metal ions and organic bridging ligands has
increased greatly.3 Early papers by Robson,4 Moore,5 Yaghi,6
1 Introduction and Zaworotko7 pointed out the rich possibilities for new
material structures and properties offered by these coordination
Over the past fifty years, research into porous materials has polymers. In principle, through the wide choice of metal, and
resulted in a number of applications, which have had a direct the infinite choice and design of ligands, a broad range of
impact on domestic life and large scale industrial processes. structural, magnetic, electrical, optical, and catalytic properties
Zeolites, which are porous aluminosilicates, are the prime might be rationally incorporated into such materials. The
examples.1 The porosity of these materials allows guest number of papers which feature the term coordination
molecules to diffuse into the bulk structure, and the shape and polymer has increased dramatically over the last few years (see
size of the pores leads to shape- and size-selectivity over the Figure 1). Use of the term metal-organic framework is also
guests which may be incorporated. Once incorporated, chemical
transformations of the guests may take place. Large scale
examples of the latter are catalytic cracking and xylene
isomerisation. Ion-exchange is another property of zeolites

Stuart James was born in Pakistan in 1968, obtained his PhD in


1992 (with Professor Paul Pringle, Bristol, on diphosphine
complexes for catalysis), then was awarded a Royal Society
Post-Doctoral Fellowship with Jean-Pierre Sauvage (Uni-
versit Louis Pasteur, Strasbourg) to work on cyclometallated Fig. 1 Graph showing numbers of papers which feature the terms
diruthenium compounds. Further post-doc experience was coordination polymer (triangles) or metal-organic framework (squares)
gained with Professors Gerard van Koten (Utrecht) and Paul in their titles, abstracts or keywords (source: ISI Web of Science).
Raithby (Cambridge) during which he became interested in
metal-organic polymers. In 1996 he was appointed as a Fixed- increasing, although perhaps delayed by a few years. The first
Term Lecturer at Imperial term is used widely to describe any extended structure based on
College, London, under the metals and organic bridging ligands, whereas the second is
guidance of Professor Mike normally, though not exclusively, applied to such structures
Mingos, where he worked on which exhibit porosity. The establishment of porosity in these
hydrogen bonding between polymeric metal-organic structures has been a challenging but
metal-containing molecules central goal during the development of this field since, by
and instigated research into analogy with zeolites, it opens up possibilities for new chemical
phosphine-based coordina- separations, ion exchange, sensing and possibly even catalytic
tion cages. Since 1999, he has behaviour. This review attempts to give an accessible summary
held a Lectureship at Queens of the field at a stage where trends can be identified in synthetic
University Belfast, where his approaches, characterisation of material properties is becoming
main interests are in coor- more detailed, and some properties have been observed which
dination cages and polymers, point toward possible longer term applications. It has not been
and playing the guitar. possible to include every contribution to the topic. Instead I
have tried to include illustrative examples of key aspects.

276 Chem. Soc. Rev., 2003, 32, 276288 DOI: 10.1039/b200393g


This journal is The Royal Society of Chemistry 2003
View Online

2 Synthesis
2.1 The first problem predicting network geometry

In order to make a coordination polymer it is, in principle, only


necessary to react a potentially bridging ligand with a metal ion
which has more than one vacant or labile site (Figure 2). Either
infinite extended polymeric or discrete closed oligomeric
structures can arise, depending on the nature of the system used.
In particular, if divergent metal sites are coordinated (e.g.
mutually trans positions, or all four sites of a tetrahedral ion)
then an extended polymer results. The most detailed structural
information will be gained from single crystal X-ray crystallog-
raphy. Indeed, without knowing the crystal structure, it is going
to be very hard to interpret any subsequently observed
properties. Therefore, synthesis in this area is strongly directed
toward obtaining X-ray quality single crystals. Amorphous Fig. 3 1,2-di(4-pyridyl)ethane and the sheet structure of the anti-anti-
solid phases and gels have not been greatly investigated, despite gauche isomer of [Co2(NO3)4(L)3]n (L = 1,2-di(4-pyridyl)ethane). Re-
the potential for novel chemistry here too. Avoidance of an produced with permission from ref. 8a.
Published on 27 June 2003 on http://pubs.rsc.org | doi:10.1039/B200393G

amorphous product is favoured if the metalligand combination


used is highly labile, i.e. bond formation is rapidly reversible, so gauche, or gauche-gauche-anti, each of which propagated into
Downloaded by University of Virginia on 08 September 2012

that initially formed (kinetic) products, which may be mixtures, contrasting network topologies, one of which is shown in Figure
have the opportunity to rearrange and give a single thermody- 3. It should also be noted that other structure-determining
namically favoured product. This is a sort of self-assembly factors seem to be very closely balanced, so that the concentra-
approach. Amongst the more labile metal ions are Cu+, Cu2+, tion, type of anion, temperature etc. are all variables which can
Ag+, Cd2+, Zn2+, Co2+ and Ni2+ and these feature prominently affect the overall structure obtained. An additional problem is
in the field of coordination polymers. A potential problem with that a single metal coordination geometry can also propagate
labile metal ions is, however, that they often do not impose as into more than one type of network.8b
strong a preference for a given geometry as do other ions. This
can therefore lead to a lack of predictability over the structure of
the network obtained. Another potential problem is flexibility in 2.2 Approaches to the first problem
the bridging ligand. Clearly, if a ligand has a number of possible
conformations, again, framework geometry will be hard to Use of rigid ligands. Orientational freedom of the ligand
predict. In a study of ligand flexibility, using 1,2-bis(4- lone pairs may be reduced by using rigid backbones, such as in
pyridyl)ethane (Figure 3), which can adopt gauche or anti 4,4A-bipyridine (1) (Figure 4). Although this ligand can in
conformations, it was shown that a variety of network principle rotate about the central CC bond, importantly this
geometries can indeed occur. Reaction of this ligand with rotation does not affect the mutual orientation of the two lone
Co(NO3)2 under a range of conditions lead to the isolation of pairs. Most of the early work on coordination polymers (and
three different isomeric polymers, all of formula [Co2(N- much of that still done) employs rigid ligands such as this.
O3)4(L)3]n, which also had similar asymmetric units in their Examples of rigid 3- and 4-connecting N-donor ligands which
crystal structures,8a and the same trigonal bipyramidal N3O2 have also been used are given in Figure 4. A rigid potentially
coordination sets at the metals. The three products differed in 4-connecting ligand is hexamethylenetetramine (3) as em-
having ligand conformations that were either all anti, anti-anti- ployed by Ciani et al.9 An early example of a rigid four-

Fig. 2 The building block, or modular, principle behind forming coordination polymers.

Chem. Soc. Rev., 2003, 32, 276288 277


View Online
Published on 27 June 2003 on http://pubs.rsc.org | doi:10.1039/B200393G
Downloaded by University of Virginia on 08 September 2012

Fig. 4 Examples of rigid ligands used in the synthesis of coordination polymers.

connecting ligand is tetra(4-cyanophenyl)methane (4).4a Exam-


ples of various rigid carboxylates which have been used are also
shown, and Lin has made use of the hybrid pyridine-carboxylate
ligand isonicotinate, 4-NC5H4CO22.10

Exploitation of recurring coordination motifs. Despite the


fact that coordination at the metal is often quite flexible, it is
sometimes possible to identify recurring structural motifs either
from existing polymeric structures, or from the much greater
number of discrete model complexes. A structural motif which
is common for a variety of metals in discrete complexes and
which has been found to recur in coordination polymers11 is the
dimetallic tetracarboxylate M2(O2CR)4 unit (Figure 5, above).
Several bridging di- and tri-carboxylates have been shown to
react with Cu2+ ions to give networks based on this unit.11 An
example is that based on Cu2+ and 1,3,5-benzenetricarboxylate
(6) (the trianion of trimesic acid) (Figure 5, below).11d When a
given structural unit which assembles from a number of metal
ions and ligands can be identified, it is sometimes referred to as
a secondary building unit, or SBU11ac,12 (the ligands, metal
ions and anions themselves are the primary building units). This
is analogous to aluminosilicate zeolite chemistry, where nine
SBUs based on tetrahedral AlO4 and SiO4 primary building
units have been classified.1a The dicopper tetracarboxylate unit
can be described as a square SBU11 with regard to the
disposition of the R groups.
Tetrahedral N4 coordination at individual Cu(I) and Ag(I)
centres with non-bulky N-donor bridging ligands has occurred
in many cases, in particular in the large number of diamondoid
ML2 networks based on these metals with rod-like bidentate
pyridine- and nitrile-derived ligands (Figure 6, above).13ad
However, despite being crystallised under similar conditions to Fig. 5 The Cu2(carboxylate)4 square SBU and structure of Cu362.
Reproduced with permission from ref. 11d.
diamondoid networks, the copper network based on 1,4-bis(4-
pyridyl)butadiyne (10) proved to be the exception to the rule,
giving instead a ladder-type structure with Cu(I) centres
coordinated to three of the pyridine-type ligands and one
acetonitrile ligand.
This product, of overall stoichiometry {Cu2(NCCH3)(10)3}
was even obtained if 10 was present in large excess.14a Also, the
only crystallographically characterised example of trigonal-

278 Chem. Soc. Rev., 2003, 32, 276288


View Online

planar Cu(pyridine)3-type coordination was observed in a weeks in their supernatant, they slowly transformed via
polymeric complex Cu(1)1.5(NO3)(H2O)1.5.14b With low coor- dissolution and recrystallisation into a second product, found to
dination numbers it is also possible for metalmetal bonding to be the polymer shown in Figure 7. It is interesting that the
occur between d10 ions. An example is in the polymer precursor complex and this polymer are formally related to each
Ag(1)(NO3).14c In this case, linear chains formed by NAgN other by ring-opening polymerisation (ROP). Ring-opening
coordination were cross-linked by AgAg bonds of 2.970(2) polymerisation is well-known in main-group chemistry, but is a
(Figure 6, below). novel development for coordination rings and cages, the key
difference being that dative covalent bonds rather than ordinary
covalent bonds are involved. Simultaneously, a similar struc-
tural relationship was suggested for related gold complexes,
although the solution-state precursor was not fully character-
ised.15b The ability to form discrete complexes as well as
polymers is due to the rotational freedom of the lone pairs of
these ligands. Given the highly labile nature of the metal ions
used, it may seem unlikely that very long chain structures would
be sufficiently stable in solution to result in crystallisation of the
observed linear polymers. Perhaps the existence of a ROP
relationship actually indicates that ROP occurs at the surface of
the growing polymeric crystal. In this way, ring-opened cages
Published on 27 June 2003 on http://pubs.rsc.org | doi:10.1039/B200393G

would become trapped as relatively insoluble polymeric


material. It will be interesting to see if the concept of ROP leads
to increased structural predictability and control in coordination
Downloaded by University of Virginia on 08 September 2012

polymers, and whether it sheds more light on the actual


coordination polymer crystallisation processes. Although few
studies have involved examination of the solution state
precursors to coordination polymers, in at least one case15c the
precursors, although discrete coordination cages, which must
have undergone ring-opening in the process of producing the
final polymer, were not structurally related to the final
polymeric product by simple ROP. Therefore if ROP was
indeed occurring, there would need to be an additional cage
intermediate to those observed by NMR spectroscopy.

Effect of the anion. Anions may play key roles in


determining the polymer structure. Apart from obvious differ-
ences in volume and shape which must be accommodated in the
crystal lattice, the different coordinating abilities of anions, and
their ability to bridge between metal centres (i.e. to act as
bridging ligands themselves) may be important. For example,
the helical staircase structure of [Ag(12)(NO3)] (12 = bis(4-

Fig. 6 A representation of a diamondoid M(L)2 stoichiometry structure


(above) reproduced with permission from ref. 23, and lower-coordinate
M(1) stoichiometry chains with AgAg contacts in Ag(1)NO3 as reported pyridyl)tetrazine) was ascribed to the ability of NO32 to bridge
in ref. 14c (below). Figure reproduced with permission from ref. 23. between Ag+ centres, whereas fluoro anions BF42 and PF62
coordinate much more weakly and so metalanion bonding
interactions in [Ag(12)(fluoroanion)] were not thought to be
structure determining.16 The latter two structures have solvent
ROP relationships. Recently, where bridging bidentate coordinated to the metals in place of the anions and the linear
phosphines were used, coordination polymers were obtained chains pack in a simple parallel manner (Figure 8).
which had a clear structural relationship with their discrete
solution-based precursors. In particular, triply-bridged com-
plexes [Ag2(11)3][anion]2 (Figure 7) were characterised in 2.3 The second problem interpenetration
solution by NMR spectroscopy and molecular weight determi-
nation.15a They were also obtained as crystals and characterised Having established the possibility of forming a two- or three-
by X-ray crystallography. If these crystals were left for several dimensional coordination network, with some prediction as to

Fig. 7 Formation of discrete [Ag2(11)3(OTf)2] cages and their formal ROP conversion to [Ag2(11)3(OTf)2]H (P = PPh2).

Chem. Soc. Rev., 2003, 32, 276288 279


View Online
Published on 27 June 2003 on http://pubs.rsc.org | doi:10.1039/B200393G
Downloaded by University of Virginia on 08 September 2012

Fig. 9 Twofold interpenetration of adamantoid cavities of diamondoid


networks as in M(CN)2 (M = Zn, Cd) (above) and fivefold interpenetration
in Cu(13)2BF4 (13 = 1,2-trans-bipyridylethene), (below). Reproduced with
permission from ref. 13c.

Fig. 8 Contrasting network topologies resulting from the use of different 2.4 Approaches to the second problem
anions: parallel chains in [Ag(12)(PF6)] (above) and helical staircase in
[Ag(12)(NO3)] (below). Reproduced with permission from ref. 16.
Robson et al. demonstrated a way of avoiding interpenetration
in a key early paper in the field of coordination polymers.4b
its connectivity, use of long bridging ligands could, it seems, Neutral frameworks Cd(CN)2 and Zn(CN)2 were found to exist
lead to structures with very large open cavities, far larger than as doubly interpenetrated diamondoid nets. However, a charged
those in aluminosilicate zeolites. However, as pointed out by framework would also have to accommodate counterions. The
Robson,4b and later demonstrated in a series of diamondoid framework salt [N(CH3)4][CuZn(CN)4] was therefore prepared,
structures by a number of groups,13ad what typically happens and, as hoped, the bulky ammonium counterions present within
instead is that a second polymer network interpenetrates the the adamantoid cavities, prevented interpenetration. Impor-
first. The situation is illustrated in Figure 9. Such mutual tantly, for charge balance, only half of the adamantane cavities
interpenetration13e,f will typically fill up the potential cavities, in the mixed-metal framework were required to be filled by
reducing or even removing porosity. In fact, increasing the cations, and the remaining cavities were apparently empty
length of the bridging ligand leads to larger and larger natural (there was no evidence of guests from X-ray crystallography).
cavity sizes, but the result is that increasing numbers of It can also be reasoned that if strong enough interactions
networks then mutually interpenetrate. For example, although between the framework and guest molecules exist, then
the cyanide-bridged network in Cd(CN)2 exhibits twofold framework cavities may preferentially fill with guests than with
interpenetration (Figure 9, above),4a the use of 1 as a bridge additional polymer strands. This was suggested to have
gave fourfold interpenetrated structures Cu(1)2PF613a and occurred in Ni(1)2.5(H2O)2(ClO4)21.5(1)2H2O,19 in which X-
Ag(1)2O3SCF3.13b Increasing the ligand length further with a ray crystallography revealed large square Ni4(1)4 pores of
trans-alkenediyl spacer in 1,2-trans-bipyridylethene (13) gave a dimensions ca. 11 by 11 . M4(1)4 squares had previously been
fivefold interpenetrated network in Cu(13)2BF4,13c (Figure 9, shown by Robson et al. to be large enough for interpenetration
below) and ninefold interpenetration occurred in Ag(14)2PF6 to occur in the polymer Zn(1)2(H2O)2SiF6.20 However, in the Ni
(14 = 4,4A-biphenyldicarbonitrile).17 polymer these pores were in fact filled with hydrogen-bonded
guest aggregates comprising perchlorate anions, water mole-
cules and additional, uncoordinated 1, and there was no network
interpenetration. Hydrogen bonding was found to occur both
within these aggregates and between the aggregates and the
aquo ligands of the nickel centres, suggesting significant
interaction with the framework. Related to this point is a series
An alternative way for a structure to self-fill was pointed out of 2-dimensional square grid polymers exemplified by [Ni-
by Robson and co-workers.18 The polymer [{Cu2(O2CCH- (1)2(NO3)2]2pyrene.21 Zaworotko and coworkers showed that
3)4}3(2)2]2MeOH (2 = tri(4-pyridyl)triazine) consists of aromatic molecules such as naphthalene, pyrene, anisole and
chickenwire sheets with very large 3.43.7 nm hexagonal others, could fill up the cavities in the metalorganic polymer.
windows (Figure 10). Despite these features seeming easily The aromatic guests were in contact with each other through
large enough for interpenetration to occur, the structure was edge-to-face aromatic interactions, so that they could actually
instead found to self-fill by distorting its layers from planarity. be described as forming a second, non-covalent, network which
By adopting puckered conformations, the layers allowed their interpenetrated the coordination polymer (Figure 11). There
vertices to protrude into the windows of adjacent layers. were also short (face-to-face) contacts between the pyrene

280 Chem. Soc. Rev., 2003, 32, 276288


View Online
Published on 27 June 2003 on http://pubs.rsc.org | doi:10.1039/B200393G
Downloaded by University of Virginia on 08 September 2012

Fig. 11 Mutually orthogonal interpenetrating coordination and molecular


aromatic networks in [Ni(1)2(NO3)2]2pyrene. Reproduced with permission
from ref. 21.

example of an approximately cubic terbium dicarboxylate


network, using 4,4A-azodibenzoate (15) as the dicarboxylate

linker, where the octahedral Tb2(carboxylate)6 SBU shown in


Figure 12 was adopted.22
This SBU is a structural unit which had previously been
characterised in discrete form in terbium complexes of
monocarboxylates. Use of the long linker 12 gave rise to very
Fig. 10 Large hexagonal rings in the chickenwire layers of [Cu2(O2CCH- large cubic cavities (ca. 20 on each side) but the structure was
3)4(2)2]2MeOH (2 = tripyridyltriazine) and schematic showing how self- found to be only doubly interpenetrating, and retained a large
filling can occur by puckering of layers rather than by interpenetration. open volume which was filled by numerous DMSO guests. The
Reproduced with permission from ref. 18.
explanation for the existence of this free volume was that,
despite the very large cavity size, the large bulk of the SBU
guests and the bipyridine ligands of the coordination frame-
vertex only allowed a maximum of two networks to mutually
work. There was no interpenetration of coordination networks
interpenetrate, leaving a relatively large volume unoccupied.
in these compounds.
This assertion was supported by a geometric model of two
It is also possible to design frameworks in which inter-
interpenetrating cubic frameworks with large spherical vertices,
penetration is forbidden by the inherent dimensions of the
which enabled the maximum degree of interpenetration, n, to be
framework itself. This relies on the fact that in order to
related to the diameter of the SBU, d, the length of the linker, l,
interpenetrate, the size of the vertex or diameter of the polymer
and the van der Waals radius of the atom joining the SBU and
strand must be less than the size of the cavity. The use of bulkier
linker, d/2 (Figure 12).
vertices could therefore give rise to cavities, as long as there is
also control over the natural cavity size of the structure (this
approach assumes that, even when forbidden from inter-
penetrating, the metal and ligands will still adopt the expected
overall topology). Metal clusters and cages, as opposed to single This relationship was extrapolated to display the conse-
metal ions, are therefore of interest as building units. Yaghi et quences of d and l for the percentage free volume in maximally
al. introduced this idea to coordination networks with the interpenetrating cubic structures (Figure 12, bottom).

Chem. Soc. Rev., 2003, 32, 276288 281


View Online

the same phenomenon,15c where the network was very bulky


indeed. The X-ray crystal structure of [Ag4(17)3(OTf)4] re-
vealed hexagonal layers with very large 72-membered rings of
diameter 1618.4 nm. Although accurately defining the dimen-
sions of the irregularly-shaped layers is somewhat difficult, it
seems that the large thickness and depth of the layers would
prevent identical frameworks from orthogonal interpenetration.
This was also supported by a simplified geometric model.
Published on 27 June 2003 on http://pubs.rsc.org | doi:10.1039/B200393G

Effect of concentration. In perhaps the most dramatic


Downloaded by University of Virginia on 08 September 2012

examples of metal-organic frameworks seen so far, Yaghi et al.


succeeded in producing a series of materials with square pore
windows of diameter up to 19.1 , and pore diameters of up to
28.6 .12b The calculated densities of these materials were as
low as 0.21g cm23 the lowest of any crystalline material.
These compounds consisted of rigid dicarboxylate linkers and
octahedral Zn4O(carboxylate)6 SBUs connected into cubic
frameworks (Figure 13). Whilst doubly interpenetrating frame-
works were obtained in some cases, their non-interpenetrating
counterparts could be obtained by conducting the synthesis
under more dilute conditions.

3 Properties
3.1 Structural stability and the sorption and exchange of
liquids and vapours

Fig. 12 Bulky pseudo-octahedral Tb2(carboxylate)6 SBU (top), geometrical Assuming that interpenetration has been controlled and that a
model used to relate vertex bulk and spacer length to the degree of material with an appreciable cavity size has resulted, the
interpenetration, n, in interpenetrated cubes (middle), and the extrapolation question of stability arises. In addition to any required
of this relationship to show how percentage free volume relates to SBU counterions, solvent molecules normally become included
diameter, d, and linker length, l, reproduced with permission from ref.
22.
inside any cavity which is large enough to accept them and they
may even template the structure around them. A frequent
problem is that networks are not stable to loss of the solvent
Use of bulky ligands. Use of bulky connecting ligands guests. Here are described some of the frameworks which have
should also decrease the tendency to interpenetrate. In this shown evidence of stability to guest loss, or at least the ability
approach, window sizes to cavities should be reduced, and the to reversibly exchange one guest for another without structural
polymer strand which would need to thread through that collapse. Also included in the discussion are examples where
window is made more bulky. Schrder et al.13d explored this by the framework loses crystallinity on guest loss, but regains it on
using 2,7-diazapyrene (16) to prepare the diamondoid network re-exposure to guests.
[Cu(16)2][PF6]. Only threefold interpenetration occurred, com- The high kinetic stability of zeolites is consistent with the
pared to fourfold interpenetration in analogous networks based high energy barrier to the breaking of strong SiO and AlO
on 1.13a,b bonds. Coordination bond strengths are lower and so MOFs are
expected to be less thermally stable. This does generally seem to
be the case, and typically decomposition occurs at lower
temperatures than for zeolites. Framework stability could be
maximised by ensuring good hard-hard or soft-soft matching
between the ligand and metal to maximise bond strengths,
which might help explain why some of the highest decomposi-
tion temperatures for coordination polymers have been ob-
Although 16 is slightly shorter than 1, which would also served for lanthanide carboxylates.24 However, in studies where
decrease the propensity for interpenetration, it was concluded TGA has been the sole technique used to determine the
that the major factor was its greater lateral steric bulk. It has also decomposition temperature, it is possible that phase changes
been pointed out that nitriles, being less bulky than pyridines, and loss of long range order occur before weight is lost.
give greater numbers of interpenetrating strands for similar As a general strategy, Robson suggested exploiting the
ligand lengths, probably also for steric reasons.13d,23 Phosphine chelate effect to increase the stability of porous coordination
ligands bearing bulky phenyl groups may have also given rise to polymers.25 Ligand 18 was proposed as a phenanthroline-type

282 Chem. Soc. Rev., 2003, 32, 276288


View Online

structure with nitromethane molecules and perchlorate anions


occupying the pores. Stability studies were encouraging for this
chelate approach. A single crystal was left open to the
atmosphere, which led to apparent exchange of the nitro-
methane guests for atmospheric water molecules. Significantly,
X-ray diffraction revealed that the product was, however, still
crystalline, even though a change from a tetragonal to a cubic
unit cell had occurred, as well as an overall decrease in cell
volume. Re-exposure to nitromethane caused regeneration of
the original cell volume, although the cell remained cubic, and
a second exposure to the atmosphere lead again to the smaller
cubic cell. This last change was once again associated with
exchange of nitromethane for water. The chelate framework
was robust enough to persist throughout each of these guest
exchanges.
Moore reported an early example of partial guest loss from a
stable porous coordination network.5a The polymer
[Ag(19)(O3SCF3)]2C6H6, contained large cavities with sixteen
Published on 27 June 2003 on http://pubs.rsc.org | doi:10.1039/B200393G
Downloaded by University of Virginia on 08 September 2012

benzene molecules per unit cell, four of which were disordered.


Thermogravimetric analysis revealed mass losses correspond-
ing to four (110 C) and then twelve (145 C) benzene
molecules. X-ray powder diffraction showed that the frame-
work remained intact after the first solvent loss, but not after the
second. Changes in diffraction intensity indicated that it was the
disordered benzenes that had been lost in the first step.
There have also been several metal-organic frameworks
reported to undergo complete, or almost complete, desolvation
with retention of the framework structure, even if character-
isation of the desolvated form by single crystal X-ray crys-
tallography has not been included (for exceptions to the latter
see below). Some of the desolvated materials have also been
found to resorb small molecules, sometimes with considerable
selectivity. Reaction between 1,3,5-benzenetricarboxylic acid
(H36) and Co, Ni or Zn salts gave isostructural hydrated
products, M3(6)212H2O, with zig-zag polymeric chains and H-
bonding between the chains.26 Partial dehydration gave materi-
als which were formulated as monohydrates M3(6)2.H2O after
elemental analysis. X-ray powder diffraction showed these
Fig. 13 Series of cubic frameworks based on dicarboxylate linkers and
materials to have poor crystallinity. Nevertheless, re-exposure
octahedral Zn4O(carboxylate)6 SBUs, reproduced with permission from ref. to water regenerated the original structures, perfectly in the case
12b. of the Zn compound, but imperfectly for the Co and Ni
compounds. The Zn monohydrate also reacted with ammonia to
chelate which could generate metal-organic polymers via its generate a material which elemental analysis suggested was
three divergent chelating sites. Zn3(6)22H2O10NH3. There was no reaction with larger
potential ligands such as pyridine, which was ascribed to size
constraints imposed by a proposed channel structure based on
the original dodecahydrate. Under different synthetic condi-
tions, in the presence of pyridine, H3BTC reacted with
Co(NO3)2 to give a different polymer, Co(H6)(NC5H5)2(2/
3)NC5H5.6 Pyridine was found coordinated at the Co centres
forming Co-H6-pyridine layers, but also as an uncoordinated
guest, occupying channel spaces in the framework. By heating
A silver coordination polymer [Ag(18)(ClO4)]2CH3NO2 to 200 C the uncoordinated pyridines were removed and the
was obtained and structurally characterised. It had a porous resulting material gave an X-ray powder diffraction pattern

Chem. Soc. Rev., 2003, 32, 276288 283


View Online

whose most intense lines were identical in intensity and position


to those of the starting material, supporting some retention of
the original structure. The material was able to resorb pyridine
to regenerate the original material. Amongst other potential
guests the material showed selectivity toward aromatic mole-
cules, as deduced from competitive sorption experiments using
solvent mixtures such as benzene/nitromethane or benzonitrile/
acetonitrile. This selectivity was ascribed to the ability of
aromatic guests to enter into favourable p-stacking interactions
with the framework.
The silver-triphosphine framework [Ag4(17)3(OTf)4]15c
which has a layered chickenwire structure and unusually wide
channels (1.61.8 nm) also showed evidence of retaining its
structural integrity after guests had been removed. Thermal
decomposition was followed by XRPD, which gave a decom-
position temperature between 170 and 200 C. Guest ethanol
molecules in the as-synthesised material could be exchanged for
other liquids (e.g. diethyl ether) by immersion of the crystals in
an excess of that liquid. After immersion in dichloromethane,
Published on 27 June 2003 on http://pubs.rsc.org | doi:10.1039/B200393G

followed by heating to 170 C for 24 hours at 0.01 mmHg, a


sample of the material was dissolved in CD3CN and 1H NMR
spectroscopy used to ascertain the presence of any guests. There
Downloaded by University of Virginia on 08 September 2012

was no ethanol or dichloromethane detected, although traces of


water were present. XRPD showed no evidence of amorphous
phases, or crystalline phases other than the original.
Examples of stable frameworks which have been structurally
characterised by single crystal X-ray diffraction in both solvated
and desolvated forms are still quite rare. Rosseinsky et al.
described the single crystal structures of [Ni2(1)3(NO3)4] in Fig. 14 Reversible sorption of ethanol in [Ni2(1)3(NO3)4], which was
both its ethanol-loaded and empty forms.27a This framework characterised by X-ray crystallography in both solvated and desolvated
consisted of tongue-and-groove bilayers, with CHO forms, reproduced with permission from ref. 27a.
hydrogen bonds, involving nitrato and bipyridine ligands
connecting the layers. The material had channels of ca. 6 by 3 The fact that samples which, after saturation with methanol and
which, in the as-synthesised state, were occupied by ethanol subsequent removal from the methanol vapour, did not undergo
molecules hydrogen bonded to nitrato ligands. The ethanol this structural change suggested that it was surface-adsorbed
molecules could be completely removed by heating to 100 C, methanol which was essential to the isomerisation. However,
and then reintroduced by exposure of the desolvated material to observations of individual crystals using an optical microscope
ethanol vapour. Single-crystal X-ray crystallography revealed showed that they retained their shape during the process,
that only slight structural readjustment occurred in the frame- suggesting that it was not due to dissolution and recrystallisa-
work during this process (Figure 14). tion.
The overall framework connectivity was retained on desolva- Yaghi et al. reported single crystal structures of Zn4(O)(5)3 (5
tion, with a decrease in the unit cell volume of only 2.4%. A = 1,4-benzenedicarboxylate) in both solvated and desolvated
detailed thermodynamic and kinetic analysis of sorption in the states, which revealed no change in the unit cell parameters,
desolvated material was also made.27b It revealed complex indicating more inflexible zeolite-like sorption and desorption
behaviour with steps in the sorption isotherms. For methanol behaviour.27d
sorption, steps occurred at high loading which were associated Fujita et al. reported a stable, non-interpenetrated square grid
with slowing down in the adsorption kinetics at these points. framework which was obtained by reacting the extended
This was interpreted as being due to the speedy initial bipyridyl-type ligand 4,4A-di(4-pyridyl)biphenyl (20) with
occupation of nitrate sites (hydrogen bond acceptors) by
methanol molecules at low loadings, followed by a structural
adjustment of the framework to allow further methanol sorption
to occur after the initial nitrate sites had been saturated.
If the framework synthesis was conducted in methanol
solvent, a different structure [Ni(1)3(NO3)4]2CH3OH was
obtained, with similar pseudo-square pyramidal coordination at Ni(NO3)2 in a mixture of methanol and o-xylene solvents, and
Ni, but connected to form closed-packed metal-bipyridine had the layered structure depicted in Figure 15.27e
ladders.27c In sorption studies of this material, considerable
framework flexibility was discovered. Although the desolvated
compound had a pore window of cross-sectional area 12.3 2,
toluene, of minimum cross-sectional area 26.6 2 could still be
sorbed into the structure. This is in marked contrast to sorption
in zeolites, where the pore window size is taken as a more rigid
indicator of which guests may be admitted.1 Whilst consider-
able flexing of the framework must have occurred to allow
access of the guest through the windows, Monte Carlo
calculations showed that once the guest had been included Fig. 15 Sliding of layers induced by guest sorption in Ni(20)2(NO3)2 (guests
relatively little structural expansion was required to accom- are omitted), reproduced with permission from ref. 27e.
modate it within the pores. The ethanol-templated tongue-and-
groove structure could be converted to the methanol templated Rectangular channels running orthogonal to the layers were
ladder structure by standing in a static methanol vapour phase. occupied by an ordered array of o-xylene and methanol

284 Chem. Soc. Rev., 2003, 32, 276288


View Online

molecules, which were lost on heating to 70150 C. The


framework retained its integrity up to 300 C. A single crystal
structure determination on the empty material revealed that
guest loss led to only very slight readjustments of the
framework.
However, after immersion in mesitylene, a distinct change
occurred and a new structural form was obtained, which was
related to the original by a dramatic sliding of the layers over
each other. The channels in the latter material were much wider
than in the original, and mesitylene was included within them.
The structural change was monitored by using X-ray diffraction
on a single crystal of the original material sealed in a capillary
tube which also contained some mesitylene. This experiment
supported the assertion that a genuine crystal-to-crystal trans-
formation had taken place, rather than dissolution and re-
crystallisation. A second example of a crystal-to-crystal struc-
tural change was also reported in a Zn(II) based framework with
the bridging ligand 2.27f In this case, a doubly interpenetrated
network was obtained which had continuous channels occupied
Published on 27 June 2003 on http://pubs.rsc.org | doi:10.1039/B200393G

by either nitrobenzene or cyanobenzene molecules, depending


on which had been used as solvent in the synthesis. On exposure
to the atmosphere, the crystals turned from colourless to yellow,
Downloaded by University of Virginia on 08 September 2012

and single crystal X-ray diffraction revealed that the guests had
been lost. Also, significant changes in unit cell parameters had
occurred with a reduction in volume of 2023%, and associated
reduction in parameters such as inter-network metalmetal
distances. Re-immersion of this desolvated condensed mate-
rial in nitrobenzene regenerated the original solvent-filled
expanded framework (Figure 16).

Fig. 17 Conversion of [Zn3(5)3(H2O)3]4DMF to [Zn(5)(H2O)]DMF by


partial loss of the water and DMF content, which requires considerable
rearrangement of the metal ions and bridging ligands into distinctly
different connectivity, reproduced with permission from ref. 28a.
Fig. 16 Representation of the structural contraction and expansion induced
by guest sorption in the interpenetrated structure of (ZnI2)3(2)2, reproduced
with permission from ref. 27f.

Whilst single crystal X-ray structures have been obtained for


both solvated and desolvated forms of certain frameworks,
supporting the assertion that permanent porosity can indeed be
achieved in metal-organic frameworks, the observation of were occupied by water molecules which could be removed. In
reversible guest sorption does not in itself indicate that a this case the empty material showed an only slightly altered and
material has permanent porosity. Work by Wright et al.28a on a still sharp XRPD pattern compared to that of the hydrated
series of zinc polymers based on 1,4-benzenedicarboxylate (5) material, supporting retention of the structure with only a slight
highlighted the fact that metal-organic polymers are potentially change in inter-layer distances. The original diffraction pattern
highly reactive and adaptable materials. On partial loss of the H- was regenerated on exposure to water or methanol. Inter-
bonding water and DMF content from [Zn3(5)3(H2O)3]4DMF estingly, the methanol sorption isotherm showed hysteresis
to give [Zn(5)(H2O)]DMF, very dramatic changes in structure (Figure 18), consistent with sorption and desorption being
occurred, and in fact the two polymeric structures were associated with structural readjustment. Methane, despite being
topologically quite distinct from each other (Figure 17). The smaller, was not sorbed by this material.
structural change must have involved considerable metal
ligand bond breaking and reforming, i.e., loss of the hydrogen
bonding molecules caused actual reconstruction of the metal
ions and ligands into a new and distinct network topology.
Therefore, it should be borne in mind that, even though
molecules may be reversibly sorbed in a metal-organic polymer
material, the included molecules might not in all cases be most
accurately regarded simply as guests within the pores of fixed
frameworks, but instead as integral components of dense but
structurally adaptable materials. Loss of such guests can even
lead to materials with greatly reduced crystallinity which may,
however, undergo perfect recrystallisation to the inclusion
compound on re-exposure to the guest in vapour form.
Fig. 18 The hysteretic methanol sorption isotherm of Cu(II) framework
Kitagawa reported a Cu-based framework with the 1,2-dipyr- based on 21, reproduced with permission from ref. 28b.
idylglycol (21) bridging ligand.28b The alcohol groups were not
used in coordination to Cu, but instead provided hydrophilic Related to the issues of stability and guest inclusion is the
surfaces to channels of 4 3 6 cross-section. The channels ability to functionalise a framework after it has been formed.

Chem. Soc. Rev., 2003, 32, 276288 285


View Online

Lee and co-workers29 reported a remarkable series of structures


based on silver ions and bridging trinitriles, which also bore
diverse functional groups and spanned a range of sizes. They
reported good evidence that one of these channelled structures
bearing alcohol functionalities could be converted to the
corresponding ester functionalised framework by exposing it to
trifluoroacetic anhydride vapour, and that this occurred without
rearrangement of the framework.

3.2 Other properties

Chirality. Chiral metal-organic frameworks are interesting


because they could be applied to enantioselective separations,
especially so because of the difficulties in preparing enantiopure
traditional zeolite frameworks. The fact that metal-organic
frameworks are prepared in a modular or building block
fashion means that chirality can be introduced simply by
choosing chiral building blocks, most conveniently as the
Published on 27 June 2003 on http://pubs.rsc.org | doi:10.1039/B200393G

organic ligands. Rosseinsky et al. reported Ni frameworks


based on bridging 6, which also contained diols coordinated at
the nickel centres.30a The 1,2-ethanediol-containing network
Downloaded by University of Virginia on 08 September 2012

had monodentate diols, and exhibited fourfold interpenetration.


Despite this interpenetration, pseudo-tetrahedral pores of diam-
eter 12 accounted for 28% of the crystal volume, and were
interconnected to form continuous channels. The helical nature
of the networks meant that the crystals were inherently chiral,
bulk samples consisting of 50:50 ratios of enantiomeric crystals.
Use of racemic 1,2-propanediol as the alcohol led to a closely- Fig. 19 The trigonal prismatic SBU and chiral channel structure of Zn3(m3-
related helical structure, but with chelating diols at the nickel O)(22)6502H3O12H2O, reproduced with permission from ref. 30b.
centres and only twofold network interpenetration. Individual
crystals of this material were again found to grow as single most zeolites.27d Strikingly, Kitagawa reported that
enantiomers, and, interestingly, each crystal contained only one Cu(1)2(SiF6) was able to sorb very large quantities of
enantiomer of the diol. This suggested that the diol co-ligand methane.31a In fact, relative to framework weight, more
had exerted a templating effect on the handedness of the methane was sorbed by this material than by zeolite 5A, which
coordination networks. This clearly pointed to the possibility of was taken as the optimum conventional zeolite for methane
harnessing the chirality of organic building blocks to produce sorption (see Figure 20).
enantiopure porous metal-organic frameworks. Kim and co-
workers used an enantiopure ligand 22 derived from D-tartaric

acid to prepare a porous framework on reaction with


Zn(NO3)2.30b The distorted trigonal prismatic SBU in Figure 19
was adopted, and the structure exhibited chiral channels.
Pyridinium groups pointed into the channels, and these could
undergo exchange of protons for other cations. The chirality of
the material was manifested in its ability to discriminate to some
degree between enantiomers of [Ru(2,2A-bipy)3]2+. Immersion
of the framework in a methanolic solution of racemic [Ru(2,2A-
bipy)3]2+ led to a change in the colour of the crystals from
colourless to yellow, and a 66% enantiomeric excess of the D
form of the complex was found to have been included. The
material was further used to demonstrate enantioselective
catalysis (see below).

Gas Sorption. Sorption of liquids and vapours has already


been discussed above in the section on stability. Sorption of
gases has also been used to investigate the effective porosity and Fig. 20 [Cu(1)2(SiF6)] and comparison of its methane sorption isotherm
apparent surface area of frameworks, and to probe the with that of zeolite 5A, reproduced with permission from ref. 31a.
possibility of future applications in gas storage. For example,
Yaghi reported N2 sorption by Zn4(O)(5)3 at 78 K and found a A great impetus for studying this kind of sorption is the need
reversible type 1 isotherm similar to isotherms observed for for low-pressure storage media for gases such as methane and

286 Chem. Soc. Rev., 2003, 32, 276288


View Online

hydrogen, due to the potential applications of the latter as Catalysis. Catalysis is often cited as a desirable characteristic
clean fuels. Subsequently, Yaghi et al. reported still greater of metal-organic frameworks, and a justification for general
weight-for-weight methane sorption of 240 cm3 g21 in a Zn- research into these materials. However, there are still relatively
carboxylate framework.12b When compared to a conventional few reports of actual catalytic activity, and detailed investiga-
gas cylinder, at a pressure of 205 atm, the results suggested that tions are particularly lacking. In 1994 Fujita reported that the
this material could store 70% as much methane in the same channelled Cd framework [Cd(1)2(NO3)2] would catalyse the
volume, but at a lower (safer) pressure of 36 atm. cyanosilylation of aldehydes.34 In particular, treatment of
Very strikingly, sorbed oxygen molecules were recently benzaldehyde and cyanotrimethylsilane with a powdered sus-
investigated within a metal-organic framework at 90 K.31b The pension of [Cd(1)2(NO3)2] in dichloromethane at 40 C over 24
O2 molcules were located by X-ray crystallography and found hours gave 2-(trimethylsiloxy)phenylacetonitrile in 77% yield.
to exist as (O2)2 van der Waals dimers at an intermolecular Importantly, these workers had considered that the reaction
distance of 3.28(4) , close to that in solid O2, despite the study might be occurring due to partial dissolution of the framework
being conducted at well above the pure O2 freezing point of 54.4 into catalytically active soluble discrete complexes. Accord-
K. This illustrates the fact that metal-organic frameworks are ingly, they tested the activity of a supernatant dichloromethane
providing unusual new microenvironments. phase and found it gave no activity, supporting the contention
that catalysis was occurring at the solid. Also, no activity was
observed for the separate framework component precursors
Ion exchange and solubility. When crystals of the silver-
Cd(NO3)2 and 1. Some selectivity was observed in the catalytic
bipyridyl polymer [Ag(1)(NO3)]14c were immersed in aqueous
behaviour, in that 2-tolualdehyde was a more active substrate
Published on 27 June 2003 on http://pubs.rsc.org | doi:10.1039/B200393G

solutions of anions such as PF62, MoO42, BF42 or SO422 they


than 3-tolualdehyde, and although a- and b-naphthaldehyde
were found to undergo anion exchange, i.e. the aqueous anion,
were good substrates, there was almost no conversion of
which was in excess, was incorporated into the crystals and the
9-anthraldehyde. This selectivity was taken to be due to the
Downloaded by University of Virginia on 08 September 2012

crystals nitrate anions were released into solution. For PF62,


limited cavity size of the material. The selective inclusion of
the exchange was almost complete after 6 hours, and the crystals
ortho-dihalobenzenes in preference to meta or para isomers was
had become opaque, with XRPD giving a new diffraction
also reported, supporting the materials selective inclusion
pattern. Re-exposure to a nitrate-containing solution gave back
characteristics. The cadmium centres are distorted octahedral
transparent crystals with the original XRPD pattern. This
with N4O2 coordination sets, and, if coordination to Cd is
process was later studied32 by microscopy, including TEM and
involved in the catalysis, it implies that there is dissociation of
AFM, and with spectroscopic examination of the supernatant.
the terminal nitrate anions to give active Lewis acid sites. The
The exchange process was concluded to occur via the solution
enantiopure Zn-based framework of Kim and coworkers30b (see
phase, i.e. there was dissolution of the original polymer and
above) has chiral channels with pyridinium functional groups
subsequent crystallisation of the new anion-exchanged poly-
protruding into the channel space. Accordingly, these re-
mer, rather than ion diffusion and exchange within an inert solid
searchers tested for catalytic activity in transesterification
framework (as occurs with aluminosilicate zeolites). Since
reactions. They observed catalytic activity and even a modest
coordination polymers are normally obtained as crystals from
ca. 8% enantiomeric excess in the product when racemic
solution, and good quality X-ray crystal structures imply that in
1-phenyl-2-propanol was used as the solvent. There was also
the solid the chains are very long indeed, coordination polymers
evidence that the pyridinium groups of the channels could be
have often been assumed to be effectively insoluble. This is
alkylated with, for example methyl iodide, and this protected
often supported by visual observation of crystal stability on
nonacidic framework showed little or no activity.
immersion in given solvents. This latest work shows how
important it is to examine the supernatant solution. Clearly the
stability of coordination polymers to leaching will be a critical
point with regard to potential applications in ion-exchange or 4 Conclusions
water/solvent purification.
There is growing evidence to support the optimistic forecasts of
Luminescence and nonlinear optical (NLO) properties. the earliest researchers in this field. MOFs are a diverse and
The combination of size- or shape-selective sorption with rapidly growing class of functional material. The use of organic
luminescence is of interest since it could in principle lead to bridging ligands has allowed the rational design and functional-
selective sensor materials. Puddephatt designed luminescence isation of structures, allowing their properties to be rationally
into gold-phosphine coordination polymers by targeting the tuned. Demonstrations have been made of chiral frameworks
Au(PR3)3+ structural unit, and preparing the 2-D chickenwire and unusual sorption behaviour distinct from that seen in
network polymer [Au2(Ph2P(CH2)4PPh2)3][Au(CN)2]2, which aluminosilicate zeolites. Despite the newness of the field,
had the desired trigonal AuP3 centres.15b Strong blue-purple dramatic demonstrations, especially of gas sorption capacities,
solid state luminescence was observed for this polymer, similar have been reported which suggest that applications may occur.
to that of the monomeric model complex [Au(PPh3)3]+. Researchers are now patenting their work.35 In addition, MOFs
However, the polymer was not porous and no inclusion are providing new confined environments in which to study
behaviour was reported. The presence of lanthanide centres in novel inclusion effects. Lastly, the unusually large cavity sizes
[Tb(5)2] also lead to luminescence.24a This material was able to that can be accessed with these crystalline compounds, now
sorb water and ammonia, each of which gave rise to products encroaching into the mesoporous regime of 2 nm and above,
with different luminescence decay constants. suggests that MOFs may contribute to the bottom-up approach
Lin et al. exploited the fact that diamondoid networks lack to nanoscale synthesis. For example, they might act as structural
centres of symmetry to prepare two materials which exhibited templates for the formation of nanowires and other shapes of
second order nonlinear optical (NLO) behaviour.33 Odd diverse materials within their pores.
numbers of interpenetrating diamondoid networks are also non-
centrosymmetric, and the latter is a requirement for second
order NLO activity. Frameworks exhibiting 3- and 5-fold References
interpenetration based on Zn(II) or Cd(II) and mixed pyridine-
carboxylate ligands, were prepared, and found in one case to 1 (a) Molecular Sieves, R. Szostak, Van Nostrand Reinhold, New York,
give a second order NLO response of three times that of 1989; (b) Natural Zeolites, G. Gottardi and E. Galli, Springer-Verlag,
potassium dideuterophosphate (KDP). Berlin, 1985.

Chem. Soc. Rev., 2003, 32, 276288 287


View Online

2 (a) A. K. Cheetham, G. Ferey and T. Loiseau, Angew. Chem. Int. Ed., 15 (a) E. Lozano, M. Nieuwenhuyzen and S. L. James, Chem. Eur. J., 2001,
1999, 38, 3269; (b) P. Gravereau, M. Garnier and A. Hardy, Acta 7, 2644; (b) M. C. Brandys and R. J. Puddephatt, J. Am. Chem. Soc.,
Crystallogr. Sect. B, 1979, 35, 2843. 2001, 123, 4839; (c) X. Xu, M. Nieuwenhuyzen and S. L. James, Angew.
3 For highlight articles and earlier reviews see: (a) B. Moulton and M. J. Chem. Int. Ed., 2002, 41, 764.
Zaworotko, Chem. Rev., 2001, 101, 1629; (b) M. J. Zaworotko, Angew. 16 M. A. Withersby, A. J. Blake, N. R. Champness, P. Hubberstey, W. S.
Chem. Int. Ed., 2000, 39, 3052; (c) B. Moulton and M. J. Zaworotko, Li and M. Schrder, Angew. Chem., Int. Ed. Engl., 1997, 36, 2327.
Curr. Opin. Solid State Mater. Sci., 2002, 6, 117; (d) M. Eddaoudi, D. 17 K. A. Hirsch, D. Venkataraman, S. R. Wilson, J. S. Moore and S. Lee,
B. Moler, H. Li, B. Chen, T. M. Reineke, M. OKeeffe and O. M. Yaghi, J. Chem. Soc., Chem. Commun., 1995, 2199.
Acc. Chem. Res., 2001, 34, 319; (e) M. J. Zaworotko, Nature, 1999, 402, 18 S. R. Batten, B. F. Hoskins, B. Moubaraki, K. S. Murray and R. Robson,
242; (f) A. N. Khlobystov, A. J. Blake, N. R. Champness, D. A. Chem. Commun., 2000, 1095.
Lemenovskii, A. G. Majouga, N. V. Zyk and M. Schrder, Coord. 19 O. M. Yaghi, H. L. Li and T. L. Groy, Inorg. Chem., 1997, 36, 4292.
Chem. Rev., 2001, 222, 155; (g) C. Janiak, Angew. Chem., Int. Ed. Engl., 20 R. W. Gable, B. F. Hoskins and R. Robson, J. Chem. Soc., Chem.
1997, 36, 1431; (h) O. M. Yaghi, H. Li, C. Davis, D. Richardson and T. Commun., 1990, 1677.
L. Groy, Acc. Chem. Res., 1998, 31, 474. 21 K. Biradha, K. V. Domasevitch, B. Moulton, C. Seward and M. J.
4 (a) B. F. Hoskins and R. Robson, J. Am. Chem. Soc., 1989, 111, 5962; Zaworotko, Chem. Commun., 1999, 1327.
(b) B. F. Hoskins and R. Robson, J. Am. Chem. Soc., 1990, 112, 1546; 22 T. M. Reineke, M. Eddaoudi, D. Moler, M. OKeeffe and O. M. Yaghi,
(c) B. F. Abrahams, B. F. Hoskins, D. M. Michail and R. Robson, J. Am. Chem. Soc., 2000, 122, 4843.
Nature, 1994, 369, 727. 23 A. J. Blake, N. R. Champness, P. Hubberstey, W.-S. Li, M. A.
5 (a) D. Venkataraman, G. B. Gardner, S. Lee and J. S. Moore, J. Am. Withersby and M. Schrder, Coord. Chem. Rev., 1999, 183, 117.
Chem. Soc., 1995, 117, 11600; (b) G. B. Gardner, D. Venkataraman, J. 24 (a) T. M. Reineke, M. Eddaoudi, M. Fehr, D. Kelley and O. M. Yaghi,
S. Moore and S. Lee, Nature, 1995, 374, 792. J. Am. Chem. Soc., 1999, 121, 1651; (b) T. M. Reineke, M. Eddaoudi,
6 O. M. Yaghi, G. M. Li and H. L. Li, Nature, 1995, 378, 703. M. OKeeffe and O. M. Yaghi, Angew. Chem. Int. Ed., 1999, 38, 2590;
Published on 27 June 2003 on http://pubs.rsc.org | doi:10.1039/B200393G

7 S. Subramanian and M. J. Zaworotko, Angew. Chem., Int. Ed. Engl., (c) L. Pan, E. B. Woodlock, X. T. Wang and C. Zheng, Inorg. Chem.,
1995, 34, 2127. 2000, 39, 4174; (d) V. Kiritsis, A. Michaelides, S. Skoulika, S. Golhen
8 (a) T. L. Hennigar, D. C. MacQuarrie, P. Losier, R. D. Rogers and M. and L. Ouahab, Inorg. Chem., 1998, 37, 3407.
Downloaded by University of Virginia on 08 September 2012

J. Zaworotko, Angew. Chem., Int. Ed. Engl., 1997, 36, 972; (b) As 25 B. F. Abrahams, P. A. Jackson and R. Robson, Angew. Chem. Int. Ed.,
pointed out by a referee, both 2D (6,3) sheets and 3D (10,3) networks 1998, 37, 2656.
can arise from trigonal metal coordination. This referee also drew 26 O. M. Yaghi, H. Li and T. L. Groy, J. Am. Chem. Soc., 1996, 118,
attention to the mysterious predominance of diamondoid networks for 9096.
tetrahedral metal coordination, when other unobserved networks 27 (a) C. J. Kepert and M. J. Rosseinsky, Chem. Commun., 1999, 375; (b)
(Lonsdaleite and quartz) seem equally feasible. A. J. Fletcher, E. J. Cussen, T. J. Prior, M. J. Rosseinsky, C. J. Kepert
9 L. Carlucci, G. Ciani, D. M. Proserpio and S. Rizzato, J. Solid State and K. M. Thomas, J. Am. Chem. Soc., 2001, 123, 10001; (c) E. J.
Chem., 2000, 152, 211 and references therein. Cussen, J. B. Claridge, M. J. Rosseinsky and C. J. Kepert, J. Am. Chem.
10 M. E. Chapman, P. Ayyappan, B. M. Foxman, G. T. Yee and W. B. Lin, Soc., 2002, 124, 9574; (d) H. Li, M. Eddaoudi, M. OKeeffe and O. M.
Cryst. Growth Des., 2001, 1, 159 and references therein. Yaghi, Nature, 1999, 402, 276; (e) K. Biradha, Y. Hongo and M. Fujita,
11 (a) S. A. Bourne, J. J. Lu, A. Mondal, B. Moulton and M. J. Zaworotko, Angew. Chem. Int. Ed., 2002, 41, 3395; (f) K. Biradha and M. Fujita,
Angew. Chem. Int. Ed., 2001, 40, 2111; (b) B. L. Chen, M. Eddaoudi, T. Angew. Chem. Int. Ed., 2002, 41, 3392.
M. Reineke, J. W. Kampf, M. OKeeffe and O. M. Yaghi, J. Am. Chem. 28 (a) M. Edgar, R. Mitchell, A. M. Z. Slawin, P. Lightfoot and P. A.
Soc., 2000, 122, 11559; (c) B. L. Chen, M. Eddaoudi, S. T. Hyde, M. Wright, Chem. Eur. J., 2001, 7, 5168; (b) R. Kitaura, K. Fujimoto, S.-i.
OKeeffe and O. M. Yaghi, Science, 2001, 291, 1021; (d) S. S.-Y. Chui, Noro, M. Kondo and S. Kitagawa, Angew. Chem. Int. Ed., 2002, 41,
S. M.-F. Lo, J. P. H. Charmant, A. G. Orpen and I. D. Williams, Science, 133.
1999, 283, 1148. 29 Y.-H. Kiang, G. B. Gardner, S. Lee, Z. Xu and E. B. Lobkovsky, J. Am.
12 (a) J. Kim, B. L. Chen, T. M. Reineke, H. L. Li, M. Eddaoudi, D. B. Chem. Soc., 1999, 121, 8204.
Moler, M. OKeeffe and O. M. Yaghi, J. Am. Chem. Soc., 2001, 123, 30 (a) C. J. Kepert, T. J. Prior and M. J. Rosseinsky, J. Am. Chem. Soc.,
8239; (b) M. Eddaoudi, J. Kim, N. Rosi, D. Vodak, J. Wachter, M. 2000, 122, 5158; (b) J. S. Seo, D. Whang, H. Lee, S. I. Jun, J. Oh, Y. J.
OKeeffe and O. M. Yaghi, Science, 2002, 295, 469. Jeon and K. Kim, Nature, 2000, 404, 982.
13 For examples see: (a) L. R. MacGillivray, S. Subramanian and M. J. 31 (a) S.-i. Noro, S. Kitagawa, M. Kondo and K. Seki, Angew. Chem. Int.
Zaworotko, J. Chem. Soc., Chem. Commun., 1994, 1325; (b) L. Ed., 2000, 39, 2081; (b) R. Kitaura, S. Kitagawa, Y. Kubota, T. C.
Carlucci, G. Ciani, D. M. Proserpio and A. Sironi, J. Chem. Soc., Chem. Kobayashi, K. Kindo, Y. Mita, A. Matsuo, M. Kobayashi, H. C. Chang,
Commun., 1994, 2755; (c) A. J. Blake, N. R. Champness, S. S. M. T. C. Ozawa, M. Suzuki, M. Sakata and M. Takata, Science, 2002, 298,
Chung, W. S. Li and M. Schrder, Chem. Commun., 1997, 1005; (d) A. 2358.
J. Blake, N. R. Champness, A. N. Khlobystov, D. A. Lemonovskii, W. 32 A. N. Khlobystov, N. R. Champness, C. J. Roberts, S. J. B. Tendler, C.
S. Li and M. Schrder, Chem. Commun., 1997, 1339.For reviews of Thompson and M. Schrder, CrystEngComm, 2002, 4, 426.
interpenetration see (e) S. R. Batten and R. Robson, Angew. Chem. Int. 33 O. R. Evans, R.-G. Xiong, Z. Y. Wang, G. K. Wong and W. Lin, Angew.
Ed., 1998, 37, 1460; (f) S. R. Batten, CrystEngComm, 2001, 67. Chem. Int. Ed., 1999, 38, 536.
14 (a) A. J. Blake, N. R. Champness, A. Khlobystov, D. A. Lemenovskii, 34 M. Fujita, Y. J. Kwon, S. Washizu and K. Ogura, J. Am. Chem. Soc.,
W. S. Li and M. Schrder, Chem. Commun., 1997, 2027; (b) O. M. 1994, 116, 1151.
Yaghi and H. L. Li, J. Am. Chem. Soc., 1995, 117, 10401; (c) O. M. 35 See for example: U.S. Patent no. 5 648 508, July 15, 1997, to Nalco;
Yaghi and H. L. Li, J. Am. Chem. Soc., 1996, 118, 295. U.S. Patent no. 6 372 932, April 16 2002, to University of Liverpool.

288 Chem. Soc. Rev., 2003, 32, 276288

Вам также может понравиться