Sie sind auf Seite 1von 14

Chemie Ingenieur Technik 2007, 79, No.

DOI: 10.1002/cite.200700062

Metal-Organic Frameworks (MOFs)

767

Paddle-Wheel Zinc Carboxylate Clusters as Building Units for MetalOrganic Frameworks


Sergei I. Vagin, Anna K. Ott, and Bernhard Rieger*
Dedicated to Professor Dr.-Ing. J. Weitkamp on the occasion of his 65th birthday
Growth of interest to a field of materials chemistry such as the metal-organic frameworks (MOFs) demands the comprehensible surveying of new results, and the reviews systematizing and highlighting one or another trend in the MOF research appear periodically. Structural peculiarities of coordination polymers constructed by zinc paddle-wheel clusters and the investigation of their microporosity are reviewed in detail to emphasize those features of the materials, which are interesting both for the pure and applied chemists. Keywords: Clusters, Metal-Organic Frameworks (MOFs), Microporosity Received: April 10, 2007; Accepted: April 18, 2007

Introduction

Within the last decades, the interest towards coordination polymers as microporous materials, often called metal-organic frameworks (MOFs), has immensely grown, as can be seen from a number of publications appearing annually and devoted to this field [1]. Such a high concernment is mostly caused by a colossal potential of these materials for numerous applications in modern science and technology. The overwhelming majority of these applications are, however, based on the ability of MOFs to behave as hosts for certain molecules. MOFs and the related organic-inorganic hybrid materials have already been tested as microporous materials for the storage of gases [2], as catalysts, sometimes showing the enantioselectivity [3], as sensors for special classes of molecules [4], as active materials for non-linear optics [5], as organic magnets [6], as materials for selective sorption from the gaseous and liquid mixtures [7], etc. The versatility of attractive features manifested by this class of materials is mostly due to the unlimited possibilities for modification and fine-tuning their structures and properties. In general, MOFs and other coordination polymers can be considered as materials assembled with high order from the so called secondary building units

(SBUs) containing metal ions or their clusters [8], which are linked by polytopic or polydentate organic fragments (ligands or linkers) to build multidimensional nets or frameworks. The process of MOFs self-assembling is driven by the formation of metal ion to ligand coordination bonds as well as weaker hydrogen bonds and Van-der-Waals interactions between non-metallic components. It is obvious, that variation of only the SBUs nature should lead to a high number of different coordination polymers with diverse properties, while the design of organic linkers allows the chemists to multiply their number and functionality nearly to infinite. In support of the said above, already several hundreds structures of homoleptic and heteroleptic coordination polymers based on unsubstituted terephthalic acid were reported in CSD up to now, and the number of all MOFs described to date is incommensurably higher. It is clear, that nowadays a comprehensive review covering all the reported MOFs would be nearly impossible to accomplish, and there is more sense in doing a detailed survey devoted to a narrow field of this type of chemistry, which would be useful both for the specialists and for those who begins the work in this topic or is just interested in. The aim of the present survey was inspired by recent publications [9], and is to summarize the achieve-

Within the last decades, the interest towards coordination polymers as microporous materials, often called metalorganic frameworks (MOFs), has immensely grown.

The process of MOFs self-assembling is driven by the formation of metal ion to ligand coordination bonds as well as weaker hydrogen bonds and Van-der-Waals interactions between non-metallic components.

2007 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

www.cit-journal.de

768

bersichtsbeitrge

Chemie Ingenieur Technik 2007, 79, No. 6

Abbreviation DMF

Name N,N-dimethylformamide Dimethylsulfoxide

Structural formula

N O
O S

ments in coordination chemistry of polymeric zinc carboxylates, in particularly constructed by binuclear Zn clusters with paddle-wheel geometry.

DMSO

Paddle-Wheel Clusters

HMPA

Hexamethylphosphor-triamide

O N P N N
O

1,2-Dimethoxyethane DABCO Triethylenediamine {1,4-diazabicyclo[2,2,2]octane} Pyridine

O
N

Py

4-MeOPy

4-Methoxypyridine

Lut

3,4-Lutidine

N
iQ Isoquinoline

N
Q Quinoline

N
BIPY 4,4-Bipyridine

BPE

Trans-1,2-bis(4-pyridyl)ethylene 1,2-Bis-(4-pyridyl)ethane

N N

N N

DPNI

N,N-bis(4-pyridyl)1,4,5,8-naphthalenetetra-carboxydiimide

O N N O

O N O N

Usually, binuclear metal clusters in which two metal ions are bridged by four carboxylic groups in syn-syn mode are called in the literature as paddle-wheel. A typical example of paddle-wheel complex is copper(II) acetate hydrate [10], in which four acetate-anions are each coordinated as dimonodentate ligands to two copper(II) ions forming the D4h-symmetrical structure having the appearance of a paddle-wheel with four blades, in axial positions of which the molecules of water are coordinated (see Fig. 1). The similar constructions employing different bridging carboxylates are known for nearly all metal ions in periodic system, starting from magnesium(II) in its N,N-diphenylcarbamato hexamethylphosphortriamide complex [Mg2(Ph2NCO2)4(HMPA)2] [11] and finishing with bismuth(II) in its trifluoroacetates (see Tab. 1 for chemical structures of ligands referred to in this review) [12]. Alteration of the size and geometry of coordination sphere from one metal in the periodic system to another results in some structural variations observed for their paddle-wheel clusters, such as absence of the axial ligands or presence of the extra ones, formation of the multiple metal-metal bonds, etc. This is often realized for the metals of 5-th and 6-th periods, among which, perhaps, the complexes of molybdenum, rhodium and ruthenium are investigated to the utmost. For these metals, the paddle-wheel structures with lowered symmetry are also known. The symmetry lowering to C2v or D2h can occur, for example, when two different bridging ligands are used for the pad-

1,4-Bis(1,2,4-triazol1-yl)-butane OAc Acetate

N N

N N N

Trifluoroacetate

O O

O O

F F F
Figure 1. Structural model of copper(II) acetate monohydrate paddle-wheel complex.

Table 1. Structure and name abbreviation of ligands referred to in the text.

www.cit-journal.de

2007 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Chemie Ingenieur Technik 2007, 79, No. 6

Metal-Organic Frameworks (MOFs)

769

dle-wheel construction [13], whereas C4v-symetrical cluster will arise if the axial ligands alone or together with the oxidation states of metal ions are non-equivalent [14]. For the metals with high coordination numbers such as lanthanides even higher decrease of the symmetry can take place due to the coordination of additional axial ligands and distortion of the paddle-wheel structure (see for example [15]). Paddle-wheel structural motifs can be found both in the discrete (molecular) coordination compounds and in the infinite coordination polymers such as 1D chains, 2D nets, and 3D frameworks. 2D and 3D MOFs based on the carboxylate paddle-wheel SBUs are known, for example, for cobalt ([Co2(2,6-NDC)2(BPE)]n nC6H6nH2O, where 2,6-NDC denotes 2,6naphthalenedicarboxylate and BPE denotes trans-1,2-bis(4-pyridyl)ethylene) [16], copper ([Cu3(BTC)2(H2O)3]n, where BTC denotes benzene-1,3,5-tricarboxylate) [17], cadmium ([Cd2(3,3-BPDC)2(DPE)]n, where 3,3-BPDC denotes biphenyl-3,3-dicarboxylate) [18], lanthanides [19] and other metals besides zinc. The latter, among few other ions, which also exhibit the coordination number 6, is considered to have a flexible coordination sphere. Zinc can easily adopt the coordination geometry as well as form clusters to fit the minimum energy according to the requirements of ligands and solvent surrounding in addition to other conditions. This becomes apparent from the number of different zinc carboxylate clusters observed in MOFs, for example. Already terephthalic acid along or in combination with other ligands gives a variety of MOFs based on single zinc ions [20], different binuclear Zn2 clusters including paddle-wheel structures [21], diverse trinuclear- [22], tetranuclear- [23], pentanuclear- [24], and clusters with higher Zn ion number, finishing with the infinite l(OCO) and l-O bent bridged Zn-rods [25]. As was already said above, we will try here to summarize the peculiarities of MOFs utilizing Zncarboxylate paddle-wheel clusters as SBUs. MOFs constructed by binuclear Zn2 clusters in which less then three carboxylate groups are bridging two metal ions in syn-syn mode {e.g., (l-OCO)2(l-O)Zn2} will not be considered further as being only distantly related to paddle-wheel structures.

Abbreviation

Name Crotonate

Structural formula

Tiglate

O O

Benzoate

O O

2-Chlorobenzoate

O O

Cl O O

Fumarate

O O O O O O H2N

O O O O O -

BDC

1,4-Benezenedicarboxylate (terephthalate) 2-Aminoterephthalate

ABDC

TBDC Tetramethylterephthalate

O O

O O
F F O O

F4BDC

Tetrafluoroterephthalate

O O

CB-BDC

Dihydrocyclobuta[1,2-b]terephthalate

O O

O O
O O O

1,3-BDC

1,3-Benezenedicarboxylate (isophthalate)

BTC

Benzene-1,3,5-tricarboxylate (trimesitate)

O
2,6-NDC 2,6-Naphthalenedicarboxylate

O O O O

Table 1. Continued.

2007 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

www.cit-journal.de

770

bersichtsbeitrge

Chemie Ingenieur Technik 2007, 79, No. 6

Abbreviation 1,4-NDC

Name 1,4-Naphthalenedicarboxylate

Structural formula

Coordination Polymers with Zinc Paddle-Wheel SBUs

BPDC Biphenyl-4,4dicarboxylate

O O

O O

O O

3.1 Three-Blade Paddle-Wheels: Structure


Two types of binuclear zinc clusters will be referred further as paddle-wheels (PW), namely those symmetrically bridged by three (threeblade PW-3) and by four (four-blade PW-4) carboxylates in syn-syn mode. Clusters of the first type (see Fig. 2) are relatively often observed in coordination polymers based on monocarboxylic acids, both aliphatic (crotonic, tiglic, etc. [26]) and aromatic (benzoic, 2-chlorobenzoic [27]), and more rare as molecular units [28]. Extension of such clusters into 1D polymeric chain occurs via axial coordination either with a hydroxide (l-OH, [27b]) or with a carboxylate group in syn-anti bridging mode, which makes such polymeric chains structurally related to infinite zinc rod clusters. All zinc ions have tetrahedral coordination sphere in these polymers. Anti-anti carboxylate bridging mode can also take place, as for example in case of catena-(tris-(l2-trifluoroacetato)-l2-trifluoroacetato-bis-(dimethoxyethane)-dizinc(II)) [29], where it is probably promoted by the presence of additional axial ligands (dimethoxyethane) at the PW cluster. All zinc ions have pseudo-octahedral coordination sphere in this polymer, with Zn-Zn distance in the cluster being the highest among all reported PW-3 Zn compounds. Dense packing of such 1D polymers in the bulk of the crystals results in no noticeable porosity, and no properties related to host-guest interactions are reported for these materials. Similar Zn2 cluster structures were observed in several 3D coordination polymers, which were formed from polycarboxylic acids such as trimesic (H2BTC) [30], adamantanetetracarboxylic (H2ATC) [31], tetramethylterephthalic (H2TBDC) [32a], and 2-aminoterephthalic (H2ABDC) [32b]. In the case of a 3D coordination polymer [Zn2(TBDC)2(H2O)1.5(DMF)0.5 (DMF)(H2O)]n, (DMF = N,N-dimethylformamide) known also as MOF-47 [32a], the threeblade PW-3 Zn2 SBU is formed by four acid molecules and one additional neutral axial ligand, e.g., water. Three carboxylate groups of three different acid molecules are bridging two zinc ions in a dimonodentate fashion, and the fourth TBDC is axially coordinated to one of the zinc ions in a monodentate anti mode. Extension of the structure based on this SBU to form the infinite polymer reveals two types of connectivity of TBDC: one unit exhibits a bisdimonodentate coordination to zinc centers, whereas the other TBDC unit coordinates in

O O

3,3-BPDC

Biphenyl-3,3dicarboxylate

O O
DPA Biphenyl-2,2dicarboxylate (diphenate)

O O
CDC 1,4-Cubanedicarboxylate

O O -

O O

ATC

Adamantanetetracarboxylate

O O O O O

O O O

MTB

4,4,4,4-Tetraphenyl-methanetetracarboxylate

O O

O O

O
Ph2NCO2 N,N-diphenylcarbamate

N O
Table 1. Continued.

www.cit-journal.de

2007 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Chemie Ingenieur Technik 2007, 79, No. 6

Metal-Organic Frameworks (MOFs)

771

dimonodentate-monodentate fashion. This leads to the double layer structural motif of the polymer with the tetrahedral Zn2-based SBU (see Fig. 3) [32a]. If axial ligands are not considered, the symmetry of such cluster would ideally be D3h with Zn-O-C-O-Zn bound atoms lying nearly in one plane. This is not the rule, however, and the symmetry of such clusters can be often lowered due to the noticeable out-of-plane torsion and in-plane tilting of bridging carboxylic groups. Additionally, depending on the axial ligands, different connectivity of PW-3 Zn2 SBUs can be observed. Thus, in [Zn2(BTC)(NO3)(H2O)(C2H5OH)5]n [30a], called also MOF-4, the axial positions in Zn2 cluster are occupied by non-bridging nitrate and ethanol ligands, resulting in the triangular connectivity of this SBU via the BTC units (see Fig. 4). Combined with the triangular connectivity of the BTC linkers, this leads to a 3D porous network, guest binding properties of which will be discussed in the subsequent chapter. Coordination polymer [Zn4(ABDC)3(NO3)2 (H2O)2]n (PNMOF-3) [32b] shows zinc PW-3 cluster structure similar to MOF-4 with axially coordinated nitrate and aqua ligands. The zinc ions in the cluster are syn-syn bridged by three ABDC units (paddle-wheel blades), and the clusters are linked through linear ABDC struts into infinite hexagonal grid. The staking of grids generates large channels (14.9 ) orthonormal to the layer, which are filled with solvent. The structure of the PW-3 cluster in MOF-35 ([Zn2(ATC)(C2H5OH)2(H2O)2]n) [31] is similar to that of MOF-47, with the only difference that ATC unit in axial position coordinates via a carboxylic group in monodentate syn mode, and the opposite axial position is occupied by ethanol instead of water, resulting in a distorted tetrahedral shape of such SBU. Assembly of such SBUs via tetrahedral four-connecting ATC linker gave a 3D-network with

Figure 2. Structural model of a typical zinc three-blade paddle-wheel cluster extended into 1D polymeric chain via the axial bridging carboxylate in syn-anti coordination mode (left), and the possible dimonodentate (non-chelating) coordination modes of acetate (right). Pink is zinc, green is any metal, red is oxygen, cyan is carbon.

Figure 3. Structure of zinc PW-3 SBU in MOF-47, its representation as a tetrahedron and extension into a 3D structure [32a]. Reproduced by permission of The Royal Society of Chemistry.

Figure 4. PW-3 (a) and its representation with triangular building units (b) in MOF-4. Reprinted in part with permission from [8]. Copyright 2001 American Chemical Society.

Figure 5. PW-3 in MOF-35 (left), its representation with polygons (middle) and the assembly into a framework with hexagonal channels (right). Reprinted in part with permission from [31]. Copyright 2001 American Chemical Society.

2007 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

www.cit-journal.de

772

bersichtsbeitrge

Chemie Ingenieur Technik 2007, 79, No. 6

internal cavities of 5 filled with ethanol and water guest molecules (see Fig. 5). Another interesting network USF-4 ([Zn6(BTC)4(iQ)4(MeOH)2(MeOH)8(C6H5Cl)]n, where iQ denotes isoquinoline) was recently reported [30b], formed by a combination of triangular, square and pseudo-tetrahedral building blocks (ternary 3,4-connected net) in a ratio 4:1:2, respectively. The square and tetrahedral units were represented by PW-4 and PW-3 zinc clusters correspondingly, and the triangular unit was BTC. The structure of the PW-3 SBU in this coordination polymer differs from that in MOF-35 only by the nature of neutral ligands at one of the axial positions.

3.2 Guest Binding Properties of the Three-Blade PW-Based MOFs


The tendency of nature to avoid empty volume in the course of self-organization in condensed matter, e.g., in the crystal growth processes, is quite understandable as it corresponds to the lowering of total energy in the system. For this reason all potentially porous MOFs are formed as networks filled with guest molecules, and this is also true for the described above MOF-4, MOF-35, MOF-47, PNMOF-3, and USF-4. Among these five, MOF-4 is the only studied in detail for the guest removal and binding. Although MOF-35 and PNMOF-3 were found to have internal cavities with guest molecules, no further investigation on the guest removal and related processes in these materials was undertaken. For MOF-47 only the thermogravimetric analysis (TGA) was carried out, showing a decrease of weight (21 %) in the range 50 to 90 C and corresponding to the loss of one and half DMF and one water molecule per structural unit [32a]. Coordination polymer USF-4 is reported to possesses 32 % free volume filled with guest molecules such as methanol and chlorobenzene. TGA and temperature dependent PXRD (powder X-ray diffraction) spectra were measured for this material, with the latter indicating a substantial loss of crystallinity already at 130 C. As-synthesized MOF-4 ([Zn2(BTC)(NO3) (H2O)(C2H5OH)5]n) has channels filled with unbound ethanol and water molecules (one of each per formula unit), removal of which would lead to open pores with approximately 9 diameter (see Fig. 6). Further removal of bound ethanol molecules would extend the channels to 14 in diameter and lead to coordinative unsaturated zinc moieties, which would be highly desirable for some applications, e.g., catalysis and selective

Figure 6. View along the diagonal axis in the cubic cell of MOF-4 [8], showing the open channels through several elementary cells. Hydrogen atoms and unbound solvent molecules are omitted for clarity, disorder of the nitrate ligand is maintained. Created from the data supplied by the Cambridge Crystallographic Data Centre.

The tendency of nature to avoid empty volume in the course of selforganization in condensed matter, e.g., in the crystal growth processes, is quite understandable as it corresponds to the lowering of total energy in the system.

binding of small molecules. However, the pore constriction is likely to take place upon removal of guest molecules from this material, as no N2, Ar, or CO2 sorption isotherms could be measured for it [33]. On the other hand, an uptake of ethanol expressed by Type V isotherm was found for this material, related to coordinative sorption mechanism followed by pores filling (see Fig. 7). Such behavior can be referred, according to classification of Kitagawa [34], to flexible microporous coordination polymer of third generation with healing or breathing type pores, although the structure of the guest-free MOF-4 material was not investigated. Partially desolvated MOF-4 with a composition Zn2(BTC)(NO3)(H2O)0.5(C2H5OH) seems to maintain the network structure of as-synthesized material, since its PXRD pattern remains the same with only slight broadening of the lines [30a].

Figure 7. Ethanol sorption isotherm of MOF-4 after complete removal of guest and bound solvent molecules. Reprinted in part with permission from [33]. Copyright 2000 American Chemical Society.

www.cit-journal.de

2007 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Chemie Ingenieur Technik 2007, 79, No. 6

Metal-Organic Frameworks (MOFs)

773

As-synthesized MOF-4 exhibits specific guest exchange upon immersing into multicomponent toluene solutions. Only low molecular weight alcohols (homologues) and DMF can participate in such exchange processes in contrast to many other tested compounds, as was confirmed by solid state 13C NMR of MOF samples and by GC analysis of the solutions before and after interaction with MOF-4. Host-guest interactions in MOFs based on four-blade PW-4 zinc clusters are not less interesting, but first let us consider some structural aspects of such coordination polymers.

3.3 Structural Features of MOFs with Four Blade Paddle-Wheel SBUs


The paddle-wheel PW-4 cluster of zinc in molecular coordination compounds is observed relatively often and was reported for the first time already more than 20 years ago for tetrakis-(l-crotonato)-bis-(quinoline)-dizinc [35], with the structure similar to that of copper(II) acetate monohydrate. The average distance between two zinc ions in PW-4 Zn2 cluster is noticeably shorter than in the three-blade PW-3 clusters (approx. 2.9 to 3.0 versus 3.2 to 4.0 ) and each zinc ion has square-pyramidal coordination sphere corresponding to a coordination number five. It is not surprising that this structural unit became a relatively common building block in coordination polymers, as it can act as two-connecting (bipolar SBU), four-connecting (square and pseudo-square SBU) or six-connecting (octahedral and pseudo-octahedral SBU) structur-

al fragment (see Fig. 8). Two-connectivity is achieved if monocarboxylic acids are bridging two zinc ions in the cluster, while bidentate neutral molecules, e.g., 1,2-bis-(4-pyridyl)ethane or 1,2-bis-(4-pyridyl)-ethylene, are acting as linkers at axial positions, thus leading to 1D polymeric chains [36]. Four-connecting Zn2 PW-4 are usually constructed by polycarboxylic acids and non-bridging (typically monodentate) axial ligands. Combination of such building units with linear ditopic connectivity of corresponding dicarboxylic acids results in square grid 2D-planar frameworks with 44 regular tiling topology [37], represented, for example, by MOF-2 ([Zn2(BDC)2(H2O)2(DMF)2]n, where H2BDC denotes terephthalic acid [38]) and its close analogues [Zn2(BDC)2(DMSO)2(DMSO)5]n [39] (where DMSO denotes dimethylsulfoxide) and [Zn2(BDC)2(DMF)2]n [40]. The nature of axial ligands and of inclusion molecules in these three parent MOFs appeared to be essential in defining their bulk structure, e.g., by controlling via hydrogen bonding and Van-derWaals interaction the degree of rectangular distortion of 2D layers and the deviation of PW-4 symmetry from ideal (D4h), as well as determining the offsets of neighboring layers and the distances between them [40]. Other examples of 2D 44 tiling nets based on zinc paddle-wheel PW-4 clusters are: [Zn2(ABDC)2(DMF)2(C6H5Cl)0.5]n (MOF-46, where H2ABDC denotes 2-amino-terephthalic acid) [32a], [Zn2(2,6-NDC)2(Lut)2]n (where 2,6H2NDC denotes 2,6-naphthalenedicarboxylic acid, Lut denotes 3,4-lutidine) [41] and its analogues [Zn2(2,6-NDC)2(DMF)2(C6H5Cl)]n

The paddle-wheel PW-4 cluster of zinc in molecular coordination compounds is observed relatively often and was reported for the first time already more than 20 years ago.

Figure 8. Connectivity of zinc paddle-wheel clusters and the simplest framework topologies based on them.

2007 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

www.cit-journal.de

774

bersichtsbeitrge

Chemie Ingenieur Technik 2007, 79, No. 6

Figure 9. Representation of nSBUs in [Zn2(1,3-BDC)2(L)2(solv)x]n with cone (a), partial cone (b), 1,2-alternate (c) and 1,3-alternate (d) shapes [45b]. Axial ligands are omitted for clarity. Created from the data supplied by the Cambridge Crystallographic Data Centre.

Coordination polymers constructed from zinc paddlewheels and nonlinear dicarboxylic acids can give several net topologies different from planar 44-grid motif.

(MOF-105) [42], [Zn2(2,6-NDC)2(DMF)2 (solv)x]n (where solv denotes DMF, C6H6, toluene or p-xylene) [43], as well as [Zn2(CDC)2(DMF)2(DMF)2(C6H5Cl)]n (MOF104, where H2CDC denotes 1,4-cubanedicarboxylic acid) [42], [Zn2(CB-BDC)2(H2O)2 (H2O)3(DMF)1.8]n (MOF-103, where H2CBBDC denotes dihydrocyclobuta[1,2-b]terephthalic acid) [42] and [Zn2(BPDC)2(DMSO)2(DMSO)4]n (where H2BPDC denotes 4,4-biphenyldicarboxylic acid) [44]. The latter shows especially big grid cavities (approx. 1515 ), which create upon staggered stacking of 2D sheets the rectangular channels (approx. 7.515 ) filled with DMSO molecules. In case of MOF-103 and -105 a noticeable corrugation of 2D-layers was observed, caused by the lowered symmetry of organic linkers [42].

Figure 10. 2D-trigonal network constructed from zinc isophthalate paddle-wheel clusters shown as rectangles. Reproduced with permission from [45b].

Figure 11. Structure of a 2D-layer in [Zn2(DPA)2(DABCO)]n with disordered DABCO ligands [47]. Created from the data supplied by the Cambridge Crystallographic Data Centre.

Coordination polymers constructed from zinc paddle-wheels and non-linear dicarboxylic acids, such as isophthalic (1,3-H2BDC) with 120 linking angle, can give several net topologies different from planar 44-grid motif. Thus, strongly ruffled or undulating 2D 44-connected nets can be formed in course of assembling zinc paddle-wheel PW-4 clusters by isophthalate, as represented, for example, by a series of parent metal-organic frameworks [Zn2(1,3BDC)2(L)2(solv)x]n, where L can be pyridine or some of its substituted derivatives as well as isoquinoline, and solv denotes benzene, nitrobenzene or o-dichlorobenzene [45]. However, besides the difference in axial ligands and guest molecules, some structural variation of 2D-layers was observed in these coordination polymers. The latter can be easily understood when considering the tetragonal supra-SBUs (nSBUs) constructed by four paddle-wheel clusters, the shapes of which resemble to calix[4]arenes and for which four atropoisomers with cone-, partial cone-, 1,2-alternate- and 1,3alternate-form can be drawn (see Fig. 9). Three types of 2D-layers with different kind of undulation were observed in the mentioned above MOFs, composed either from pure 1,2-alternate shaped nSBUs, pure partial cone shaped nSBUs, or from combination of cone and 1,3alternate shaped nSBUs. Upon staking the layers, this becomes at least a partial reason for different potentially solvent-accessible areas calculated for these polymers [45b]. Combination of isophthalate and Zn2 PW-4 can also lead to an undulated trigonal 2D network (Kagome lattice, see Fig. 10) with 3262 topology, as observed for [Zn2(1,3-BDC)2(4-MeOPy)2(guest)x]n (where 4-MeOPy denotes 4methoxypyridine and the nature of guest is not specified). This network can be represented as constructed by triangular nSBUs consisting of three paddle-wheel clusters connected by 1,3BDC. Staking of such 2D-layers generates hexagonal channels with effective diameter 9.3 , and potential free volume of this MOF is estimated to be 46.3 % upon removal of all guest molecules [45b]. Three-dimensional coordination polymer with complex 65.8 net topology constructed from 1,3-BDC and Zn2 PW-4, namely [Zn2(1,3-BDC)2(Q)2(C6H5NO2)]n (where Q denotes quinoline), was also reported [45b, 46]. Although the compound does not possess notable pores, the potential solvent-accessible area was calculated to be 17.3 % upon removal of guest nitrobenzene molecules. Diphenic acid (H2DPA) belongs to non-linear ditopic linkers as well. So far, only one coordination polymer was reported to be based on Zn2 paddle-wheel clusters formed by DPA and triethylenediamine (DABCO) as additional li-

www.cit-journal.de

2007 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Chemie Ingenieur Technik 2007, 79, No. 6

Metal-Organic Frameworks (MOFs)

775

gand, namely [Zn2(DPA)2(DABCO)]n.[47] The structure of the compound is described as composed of two-dimensional layers. These are formed by zinc PW-4 clusters recumbent in one plane and linearly bridged by DABCO via axial positions into 1D chains, which are interlinked by DPA above and below the plane. Formally, paddle-wheel zinc clusters in this MOF can be considered as capable for six-connectivity, but their assembling leads to a 2D net with 44 topology due to specific geometry and flexibility of the diphenate linker (see Fig. 11). Zinc paddle-wheel PW-4 MOFs based on the carboxylic acids with the connectivity higher than two are also already known. Four-connecting tetrahedral 4,4,4,4-tetraphenylmethanetetracarboxylic acid (H2MTB) gave a 3D-framework (see Fig. 12) MOF-36 ([Zn2(MTB)(H2O)2 (DMF)6(H2O)5]n) upon assembling with square-connecting paddle-wheel zinc clusters terminated axially with aqua ligands [31]. Three-connecting trimesic acid as a triangular SBU led to complex 3D coordination framework USF-3 (see Fig. 13) composed from triangular, square and tetrahedral SBUs in ratio 4:2:1, respectively, where squares are represented by Zn2 PW-4 clusters with axially coordinated isoquinoline. The tetrahedra are bis-(l-carboxylate)l-oxo bridged dizinc clusters with additional isoquinoline or methanol ligands and chelating carboxylate group at each zinc ion [30b]. MOFs, in which zinc PW-4 clusters implement the function of six-connecting SBUs and lead to three-dimensional frameworks, have being described only relatively recently, with the first example being [Zn2(BDC)2(DABCO)(DMF)4(H2O)0.5]n.[48]. This framework is formed by distorted square grid 2D-layers of Zn2 clusters bridged by unusually bent terephthalate dianions, and the clusters are pillared by DABCO via the axial positions to built a compressed primitive cubic 3D structure with interconnecting channels. Removal of guest DMF molecules results in a relaxation of the strained 2D layers into perfect square grid, whereas inclusion of benzene leads to the shrinkage of the pores and formation of 2D rhombic grid (see Fig. 14). Such behavior of this coordination polymer, observed by X-ray crystal analysis and supported by PXRD measurements, fits with Kitagawa classification for third generation MOFs with guest-exchange deformation type pores or induced-fit type pores. Additionally, a series of coordination polymers filled with guest molecules and congeneric with the described above [Zn2(BDC)2(DABCO)(DMF)4(H2O)0.5]n was reported, in which the nature of bridging dicarboxylate as well as of pillaring ligand was varied. Using DABCO

Figure 12. Simplified presentation of MOF-36 structure, in which PW-4 clusters are shown as squares. Right view along the c-axis. Reprinted in part with permission from [31]. Copyright 2000 American Chemical Society.

Figure 13. Structural fragment of USF-3 and its representation with polyhedra and polygons, axial ligands are omitted for clarity. Reproduced with permission from [30b].

Figure 14. Structures of a 3D-framework in the as-synthesized MOF [Zn2(BDC)2 (DABCO)(DMF)4(H2O)0.5]n (a), its guest free (b) and benzene-filled (c) forms with disordered DABCO pillars [48]. Guest molecules and hydrogen atoms are omitted for clarity. Created from the data supplied by the Cambridge Crystallographic Data Centre.

as pillars allowed to prepare Zn2 PW-4 3D MOFs based on tetramethyl- (H2TBDC) or tetrafluoroterephthalic acid (H2F4BDC), 1,4naphthalenedicarboxylic acid (1,4-H2NDC), as well as 1:1 combination of terephthalic and tetramethylterephthalic acids [9a]. Applying 4,4bipyridine (BIPY) as a pillaring ligand led to new zinc paddle-wheel frameworks assembled via terephthalic, tetramethylterephthalic, fumaric and 2,6-naphthalenedicarboxylic acid [9, 21b]. MOFs with larger pillaring molecules, e.g., N,N-bis(4-pyridyl)-1,4,5,8-naphthalenetetracarboxydiimide (DPNI), and terephthalic or 4,4-biphenyldicarboxylic acid as Zn2 PW-4 forming struts are also described [9b]. One of the important features of the above MOFs is their tendency to form interpenetrated struc-

2007 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

www.cit-journal.de

776

bersichtsbeitrge

Chemie Ingenieur Technik 2007, 79, No. 6

3.4 Sorption Pproperties of Zinc Paddle-Wheel PW-4 Coordination Frameworks


Many of the up to date reported coordination polymers based on zinc four blade paddlewheel SBUs were described exclusively respecting their single crystal structures and net topologies with no further investigation on their guest removal-binding behavior. In a better case, their potential free volumes were calculated and their TGA and temperature variable PXRD characteristics were measured and discussed. It is worth to mention that the highest reported temperatures, at which some zinc paddle-wheel MOFs still maintain microcrystallinity, are in the range of 300 to 400 C, above which the decarboxylation processes usually take place. Among paddle-wheel MOFs studied in detail with the object to guest removal-binding response is MOF-2 (see above), as-synthesized composition of which is expressed by the formula [Zn2(BDC)2(H2O)2(DMF)2]n. Removal of water and DMF subsequently from this material results in the crystalline phase formulated as [Zn2(BDC)2]n, which is stable in the temperature range 190 to 315 C and is different from the parent material according to PXRD. It was proposed that removal of axially coordinated H2O and included DMF molecules, which provide a hydrogen bonding framework and connect the layers in MOF-2, results in a closer contact between Zn2 clusters of neighboring layers with formation of coordination bonds between the carboxylate group in one layer and Zn ion in another adjacent layer (see Fig. 16) [8, 38]. Such linking should lead to a quasi 3D-network with open channels, and the authors reveal a Type I isotherms for N2 and CO2 sorption at 77 K and 195 K, respectively, with the Langmuir apparent surface calculated to be 270 m2/g and 310 m2/g, respectively. Besides, the material can also absorb dichloromethane, chloroform, benzene and cyclohexane, with the latter showing a noticeable absorption-desorption hysteresis [33]. Resolvation of [Zn2(BDC)2]n by 1:1 mixture H2O:DMF resulted in appearing of only several PXRD lines belonging to the parent solid. Irreversible reaction of [Zn2(BDC)2]n with water to yield a nonporous phase was also reported [33]. Surprisingly, practically the same material [Zn2(BDC)2]n prepared by another group showed no Type I sorption of N2 at 77 K, but its subsequent resolvation by moist air and DMF resulted in full restoration of the original structure equivalent to MOF-2 [50]. [Zn2(BDC)2(DABCO)(DMF)4(H2O)0.5]n [48] is another example of MOF with well-studied

Figure 15. Double (left) and triple (right) interpenetrated structures of [Zn2(BDC)2(BIPY)]n and [Zn2(2,6-NDC)2(BIPY)]n, respectively. Reproduced with permission from [9a].

Many of the up to date reported coordination polymers based on zinc four blade paddle-wheel SBUs were described exclusively respecting their single crystal structures and net topologies with no further investigation on their guest removal-binding behavior.

tures upon increasing the length of struts (see Fig. 15). For example, BIPY pillared MOFs are all interpenetrated except the one formed by H2TBDC. Because of the bulkiness of TBDC (high rotational volume, disordering in MOF crystal), the latter MOF has too small windows in the zinc carboxylate 2D square grid to allow interpenetration. Interestingly, [Zn2(2,6NDC)2(BIPY)]n was reported to have both three-fold [9a] and two-fold interpenetrated structures [9b]. MOFs formed with DPNI pillars are double-interpenetrated [9b]. Three-fold interpenetrated structure without guest molecules was also reported, namely a MOF formed by zinc terephthalate paddle-wheel PW-4 and 1,4-bis(1,2,4-triazol-1-yl)butane as pillaring ligand [49]. Structural investigations for all these metalorganic frameworks by either single crystal Xray analysis or by PXRD reveal that most of them retain a 3D network upon removal of guest solvent molecules, forming open interconnected channels. The sizes of the channels cross-sections are dependent on the geometry of struts and the degree of interpenetration, and thus can be tuned in a reasonable scope. For example, the channels can be varied from nearly absent in triply-interpenetrated [Zn2(2,6-NDC)2(BIPY)]n up to 10 to 11 in internal diameter and 7.5 in openings for [Zn2(BDC)2(DABCO)]n [9a]. Comparison of the properties concerning the porosity and guest-host interactions in described above MOFs will be given in the following chapter.

Figure 16. Possible structural changes upon solvent removal from MOF-2. Reprinted in part with permission from [38]. Copyright 1998 American Chemical Society.

www.cit-journal.de

2007 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Chemie Ingenieur Technik 2007, 79, No. 6

Metal-Organic Frameworks (MOFs)

777

sorption properties. The guest-free framework shows permanent porosity, as confirmed by N2 and H2 sorption measurements at 77 K. The sorption of N2 follows Langmuir (Type I) isotherm with a BET surface area of 1450 m2/g and Langmuir surface area of 2090 m2/g [9a]. The hydrogen sorption at 1 bar pressure reaches 225 mL/g or 2.0 % by weight, which correspond to 5.7 H2 molecules per formula unit, and the sorption curve still indicated unsaturation under these conditions (see Fig. 17). Guest-free MOFs isoreticular to [Zn2 (BDC)2(DABCO)]n with compositions [Zn2 (BDC)(TBDC)(DABCO)]n, [Zn2(F4BDC)2(DABCO)]n, [Zn2(1,4-NDC)2(DABCO)]n, [Zn2 (TBDC)2(DABCO)]n and [Zn2(TBDC)2(BIPY)]n show comparable Langmuir/(BET) surface areas ranging from 1740/(1120) m2/g to 1400/ (920) m2/g and comparable hydrogen uptake under standard temperature and pressure (1.7 to 2.1 % by weight) [9a]. However, at low relative pressures of H2 the difference in hydrogen absorption by these isoreticular MOFs becomes more distinct, which was related to the discrepancy in the shape and size of frameworks channels rather than in the chemical nature of organic linkers [9a]. Doubly interpenetrated MOF [Zn2 (BDC)2(BIPY)(guests)x]n, called also MOF508a when as-synthesized or MOF-508b when being guest-free [21b], shows mutual displacement of interpenetrating frameworks upon removal of guest molecules. Together with distortion of paddle-wheel clusters, this leads to a reduction of the calculated potential free volume from MOF-508a to MOF-508b by 16.7 % (dense form of MOF). However, these changes appeared to be reversible, since MOF-508b can be resolvated to give exactly the same PXRD pattern as MOF-508a, and the drying-resolvating procedure can be repeated many times without loss of crystallinity. The sorption isotherms for N2 and H2 at 77 K or CO2 at 195 K measured on MOF-508b samples all show a noticeable hysteresis, which was related to re-

Figure 17. Gas sorption isotherms of guest free [Zn2(BDC)2(DABCO)]n. under standard temperature and pressure. Reproduced with permission from [48].

versible open-dense framework transformations (see Fig. 18). Apparent Langmuir surface area of the guest-free compound was found to be approximately 950 m2/g (BET surface area of 660 m2/g was measured by another group for the identical compound under the same conditions [9b]), and the hydrogen uptake at standard temperature and pressure was 90 mL/g or 0.8 % by weight [21b]. Additionally, the compound was tested as a sorbent in GC experiments on separation of natural gas, presenting good preliminary results. It also exhibited efficient GC separation of some branched and linear alkanes, showing the higher retention times for the latter because of the better fit of linear molecules to the apertures of the channels (approx. 44 in MOF-508a). Another framework isomorphic to MOF-508 also exhibits permanent porosity upon removal of guest molecules. As found from N2 sorption measurements, doubly interpenetrated guestfree [Zn2(2,6-NDC)2(DPNI)]n [9b] has BET surface area of 420 m2/g. Besides the studies of guest sorption properties, practically no other specific behavior of zinc paddle-wheel based MOFs was reported.

Figure 18. Left representation of open (a) and dense (b) phases of MOF-508 (guest molecules are omitted). Right N2 (triangles), CO2 (circles) and H2 (squares) adsorption (filled symbols) and desorption (open symbols) isotherms measured on MOF-508b. Reproduced with permission from [21b].

2007 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

www.cit-journal.de

778

bersichtsbeitrge

Chemie Ingenieur Technik 2007, 79, No. 6

Besides the studies of guest sorption properties, practically no other specific behavior of zinc paddle-wheel based MOFs was reported.

One can mention only the examination of redox activity of [Zn2(2,6-NDC)2(DPNI)]n caused by the nature of pillaring ligand [9b].

3.5 Paddle-Wheels Based MOFs: Some Aspects of Synthesis


The manner in which the assembling of zinc carboxylate coordination polymer occurs depends strongly on the applied reaction conditions, such as the solvent and co-solvent nature, reaction time and temperature, concentrations, ratios and nature of reactants and basic additives if used, etc. The reaction conditions can influence both the structure of SBUs and the type of network adopted in the process of polymer assembling. In general, the exact structure of coordination polymer can not be predicted unambiguously. In some particular cases application of the same reaction conditions to a series of organic linkers results in maintaining the SBU and framework topologies in formed MOFs including those based on zinc paddle-wheel [2a, 9a], whereas in other cases minor variations of a co-ligand structure results in the same type (PW-4) of zinc SBU but leads to drastically different net topologies [45b]. On the other hand, different co-solvents alone can already lead to a change in the SBU structure and consequently in the framework topology [30b, 36b]. The influence of a co-solvent is often connected with its ability to modulate and stabilize the framework structure by means of the fitted inclusion into the pores as guest upon MOF assembling (a template effect) [30b, 43], participating in the network of hydrogen bonds and Van-derWaals interactions. By the example of terephthalic acid and its substituted derivatives, the bulkiness of the dicarboxylate was shown to influence the preferential assembling of zinc paddle-wheels, leading either to PW-4 or to PW-3 type of clusters [32a]. Basicity of the reaction medium, as well as the nature (coordination ability and strength) of applied base was shown to affect the structures of zinc carboxylate SBUs and of resulting coordination frameworks [49, 30a]. The lower pH led to MOF assembled via single zinc ions, whereas higher pH value or presence of stronger base with lower coordination ability, e.g., triethylamine instead of pyridine, resulted in formation of PW-4 zinc clusters. One should mention that depending on the additives, reaction time or ratio of the reagents, also the mixtures of coordination polymers showing different structures, including those with zinc paddle-wheel SBUs, can be simultaneously formed during the synthesis [32b, 36b,

In general, the exact structure of coordination polymer can not be predicted unambiguously.

47, 50]. Possible reasons are the comparable thermodynamics of formation or existence of a kinetic control for the reaction products. It is, therefore, essential to always verify the uniformity of the bulk product by comparing its PXRD pattern with the one calculated from the data on single crystals picked up for the X-ray analysis. Additionally, one should keep in mind the relative flexibility of zinc-based coordination polymers, which becomes apparent from a number of reported solid-state interconversions of MOFs at ambient temperature, involving even the SBU rearrangement [40, 50]. The interconversion can take place both in the reaction solution and upon exposure to the atmosphere, for example, because of the absorption of the moisture or loss of the inclusion solvents. Interconversion of zinc MOF in solvent at increased temperature to form zinc paddle-wheel based coordination polymer is also reported, which is promoted by the addition of a complementary ligand DABCO [47]. Systematization of the reported procedures for the preparation of Zn PW MOFs indicates that several very general synthetic methodologies for coordination polymers can lead to the frameworks of interest. One of the methods involves refluxing the corresponding carboxylic acid with freshly prepared zinc hydroxide in a polar solvent such as water or alcohol, or mixture of solvents, followed by slow cooling of the reaction mixture and, when necessary, reduction of the solvent volume to form crystals of coordination polymer. The method was typically applied to prepare 1D polymeric chains constructed by zinc monocarboxylates threeblade paddle-wheel clusters [26, 27a]. Slightly modified procedure in which zinc nitrate is used as Zn source, and therefore addition of base such as pyridine is required, was applied to prepare 2D polymer based on zinc isophthalate PW-4 [45a]. Unfortunately, no yields were reported for the above preparations, and the effectiveness of the method from synthetic point of view can not be estimated. A variation of the above method is the hydrothermal synthesis, where water heated above 100 C in autoclave is applied as a reaction medium [27b, 41, 49]. Solvothermal synthesis is more often used method, in which substituted formamides are utilized as solvent. The reactants are usually zinc nitrate and carboxylic acid, often in combination with other ligands. The reaction is carried out at the elevated temperatures resulting in slow thermal decomposition of formamides to give amines, which serve as base to slowly generate the corresponding MOF-building carboxylates. The yields of zinc PW-4 based MOFs in solvothermal synthesis are typically high [9a, 21b].

www.cit-journal.de

2007 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Chemie Ingenieur Technik 2007, 79, No. 6

Metal-Organic Frameworks (MOFs)

779

The most common method for Zn PW-4 MOF preparation is based on the diffusion controlled basification of the reactants solution, typically carried out at ambient temperature. One approach involves the absorption of volatile amines by the reaction solution from the vapor phase, and another is based on the layering technique. The yields in this method vary from very low to moderate [30b, 31, 45b].

[4]

[5]

Conclusions
[6]

Whereas the binuclear paddle-wheel clusters of different metals as building blocks for coordination polymers have been already widely employed, the zinc analogs have received higher attention only relatively recently. The grown interest towards MOFs constructed from zinc paddle-wheel clusters is to a great extent connected with the recently applied pillaring approach to prepare new 3D architectures with high microporosity, good sorption properties, and thermal stability. Such coordination polymers can be prepared in very high yield, they can be respectively functionalized and their structure can be flexibly tuned to fit the desired application by means of altering the nature of carboxylate and pillaring ligands. This can be considered as an additional advantage of such MOFs over many other coordination frameworks. However, the contribution of earlier works on zinc paddle-wheel coordination compounds is not less valuable, as they lead to understanding the regularities in clusters and networks formation.

[7]

[8]

[9]

[10]

[11]

[12] [13]

Prof. B. Rieger (rieger@tum.de), Dr. S. Vagin, A. K. Ott, Waker-Lehrstuhl fr Makromolekulare Chemie, Technische Universitt Mnchen, Lichtenbergstrae 4, D-85748 Garching, Germany

[14]

[15]

[16]

References
[1] a) S. L. James, Chem. Soc. Rev. 2003, 32, 276 288. DOI: 10.1039/b200393g; b) N. W. Ockwig, O. Delgado-Friedrichs, M. OKeeffe, O. M. Yaghi, Acc. Chem. Res. 2005, 38, 176182. DOI: 10.1021/ar020022l [2] a) M. Eddaoudi, J. Kim, N. Rosi, D. Vodak, J. Wachter, M. OKeeffe, O. M. Yaghi, Science 2002, 295, 469472. DOI: 10.1126/science.1067208; b) J. L. C. Rowsell, O. M. Yaghi, J. Am. Chem. Soc. 2006, 128, 13041315. DOI: 10.1021/ja056639q [3] a) C.-D. Wu, A. Hu, L. Zhang, W. Lin, J. Am. Chem. Soc. 2005, 127, 89408941. [17]

[18]

[19]

DOI: 10.1021/ja052431t; b) D. N. Dybtsev, A. L. Nuzhdin, H. Chun, K. P. Bryliakov, E. P. Talsi, V. P. Fedin, K. Kim, Angew. Chem. 2006, 118, 930934. DOI: 10.1002/ange.200503023 T. M. Reineke, M. Eddaoudi, M. Fehr, D. Kelley, O. M. Yaghi, J. Am. Chem. Soc. 1999, 121, 16511657. DOI: 10.1021/ja983577d. a) O. R. Evans, R.-G. Xiong, Zh. Wang, G. K. Wong, W. Lin, Angew. Chem. 1999, 111, 557559. DOI: 10.1002/(SICI)15213757(19990215)111:4<557::AID-ANGE557>3.0. CO;2-Z; b) Y.-T. Wang, H.-H. Fan, H.-Z. Wang, X.-M. Chen, Inorg. Chem. 2005, 44, 41484150. DOI: 10.1021/ic0504137 E. Coronado, J. R. Galan-Mascaros, C. J. Gomez-Garcia, V. Laikhin, Nature 2000, 408, 447. DOI: 10.1038/35044035 D. N. Dybtsev, H. Chun, S. H. Yoon, D. Kim, K. Kim, J. Am. Chem. Soc. 2004, 126, 3233. DOI: 10.1021/ja038678c M. Eddaoudi, D. B. Moler, H. Li, B. Chen, T. M. Reineke, M. OKeeffe, O. M. Yaghi, Acc. Chem. Res. 2001, 34, 319. DOI: 10.1021/ar000034b a) H. Chun, D. N. Dybtsev, H. Kim, K. Kim, Chem. Eur. J. 2005, 11, 35213529. DOI: 10.1002/chem.200401201; b) B.-Q. Ma, K. L. Mulfort, J. T. Hupp, Inorg. Chem. 2005, 44, 49124914. DOI: 10.1021/ic050452i a) J. N. Van Niekerk, F. R. L. Schoening, Acta Cryst. 1953, 6, 227232. DOI: 10.1107/S0365110X53000715; b) G. M. Brown, R. Chidambaram, Acta Cryst. B 1973, B 29, 23932403. DOI: 10.1107/S0567740873006758 K.-C. Yang, C.-C. Chang, C.-S. Yeh, G.-H. Lee, S.-M. Peng, Organometallics 2001, 20, 126137. DOI: 10.1021/om000735c E. V. Dikarev, B. Li, Inorg. Chem. 2004, 43, 34613466. DOI: 10.1021/ic049937h F. A. Cotton, C. Y. Liu, C. A. Murillo, Inorg. Chem. 2004, 43, 22672276. DOI: 10.1021/ic035433s S. Takamizawa, E. Nakata, T. Saito, T. Okamura, N. Ueyama, Inorg. Chem. Commun. 2003, 6, 12391242. DOI: 10.1016/S13877003(03)00240-5 G.-F. Liu, Z.-P. Qiao, H.-Z. Wang, X.-M. Chen, G. Yang, New J. Chem. 2002, 26, 791795. DOI: 10.1039/b109869a E.-Y. Choi, K. Park, C.-M. Yang, H. Kim, J.-H. Son, S. W. Lee, Y. H. Lee, D. Min, Y.-U. Kwon, Chem. Eur. J. 2004, 10, 55355540. DOI: 10.1002/chem.200400178 S. S.-Y. Chui, S. M.-F. Lo, J. P. H. Charmant, A. G. Orpen, I. D. Williams, Science 1999, 283, 11481150. DOI: 10.1126/science.283.5405.1148 R. Wang, M. Hong, D. Yuan, Y. Sun, L. Xu, J. Luo, R. Cao, A. S. C. Chan, Eur. J. Inorg. Chem. 2004, 3743. DOI: 10.1002/ejic.200300375 a) X. Guo, G. Zhu, Q. Fang, M. Xue, G. Tian, J. Sun, X. Li, S. Qiu, Inorg. Chem. 2005, 44, 38503855. DOI: 10.1021/ic0500457; b) T. M. Reineke, M. Eddaoudi, D. Moler, M. OKeeffe, O. M. Yaghi, J. Am. Chem. Soc. 2000, 122, 48434844. DOI: 10.1021/ja000363z

The most common method for Zn PW-4 MOF preparation is based on the diffusion controlled basification of the reactants solution, typically carried out at ambient temperature.

2007 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

www.cit-journal.de

780

bersichtsbeitrge

Chemie Ingenieur Technik 2007, 79, No. 6

[20] a) G. Guilera, J. W. Steed, Chem. Commun. 1999, 15631564. DOI: 10.1039/a903163d; b) T. Ohmura, W. Mori, M. Hasegawa, T. Takei, T. Ikeda, E. Hasegawa, Bull. Chem. Soc. Japan 2003, 76, 13871395. DOI: 10.1246/bcsj.76.1387 [21] a) J. Tao, M.-L. Tong, X.-M. Chen, J. Chem. Soc., Dalton Trans. 2000, 36693674. DOI: 10.1039/b005438k; b) B. Chen, C. Liang, J. Yang, D. S. Contreras, Y. L. Clancy, E. B. Lobkovsky, O. M. Yaghi, S. Dai, Angew. Chem. 2006, 118, 14181421. DOI: 10.1002/ange.200502844 [22] a) H. Li, C. E. Davis, T. L. Groy, D. G. Kelley, O. M. Yaghi, J. Am. Chem. Soc. 1998, 120, 21862187. DOI: 10.1021/ja974172g; b) C. A. Williams, A. J. Blake, P. Hubberstey, M. Schroeder, Chem. Commun. 2005, 54355437. DOI: 10.1039/b509929c [23] H. Li, M. Eddaoudi, M. OKeeffe, O. M. Yaghi, Nature 1999, 402, 276279. DOI: 10.1038/46248 [24] X.-L. Wang, C. Qin, E.-B. Wang, Z.-M. Su, L. Xua, S. R. Batten, Chem. Commun. 2005, 47894791. DOI: 10.1039/b506398a [25] N. L. Rosi, J. Kim, M. Eddaoudi, B. Chen, M. OKeeffe, O. M. Yaghi, J. Am. Chem. Soc. 2005, 127, 15041518. DOI: 10.1021/ja045123o [26] a) W. Clegg, I. R. Little, B. P. Straughan, Acta Cryst., C 1986, C42, 919920. DOI: 10.1107/S0108270186094052; b) W. Clegg, J. T. Cressey, D. R. Harbron, B. P. Straughan, J. Chem. Cryst. 1994, 24, 211217. DOI: 10.1007/BF01672412 [27] a) W. Clegg, D. R. Harbron, P. A. Hunt, I. R. Little, B. P. Straughan, Acta Cryst., C 1990, C46, 750753. DOI: 10.1107/ S0108270189010152; b) S.-Y. Yang, L.-S. Long, R.-B. Huang, L.-S. Zheng, S. W. Ng, Acta Cryst., E 2005, E61, m1340-m1342. DOI:10.1107/S1600536805018404 [28] a) J. Klunker, M. Biedermann, W. Schaefer, H. Hartung, Z. Anorg. Allg. Chem. 1998, 624, 15031508. DOI: 10.1002/(SICI)15213749(199809)624:9<1503::AID-ZAAC1503>3.0. CO;2-D; b) D. J. Darensbourg, J. R. Wildeson, J. C. Yarbrough, Inorg. Chem. 2002, 41, 973 980. DOI: 10.1021/ic0107983 [29] D. B. DellAmico, D. Boschi, F. Calderazzo, L. Labella, F. Marchetti, Inorg. Chim. Acta 2002, 330, 149154. DOI: 10.1016/S00201693(01)00739-3 [30] a) O. M. Yaghi, C. E. Davis, G. Li, H. Li, J. Am. Chem. Soc. 1997, 119, 28612868. DOI: 10.1021/ja9639473; b) Z. Wang, V. C. Kravtsov, M. J. Zaworotko, Angew. Chem. 2005, 117, 29372940. DOI: 10.1002/ange.200500156 [31] J. Kim, B. Chen, T. M. Reineke, H. Li, M. Eddaoudi, D. B. Moler, M. OKeeffe, O. M. Yaghi, J. Am. Chem. Soc. 2001, 123, 82398247. DOI: 10.1021/ja010825o [32] a) M. E. Braun, C. D. Steffek, J. Kim, P. G. Rasmussen, O. M. Yaghi, Chem. Commun. 2001, 25322533. DOI: 10.1039/b108031h; b) A. L. Grzesiak, F. J. Uribe, N. W. Ockwig, O. M. Yaghi, A. M. Matzger, Angew. Chem. Int. Ed. 2006, 45, 25532556. DOI: 10.1002/anie.200504312

[33] M. Eddaoudi, H. Li, O. M. Yaghi, J. Am. Chem. Soc. 2000, 122, 13911397. DOI: 10.1021/ja9933386 [34] K. Uemura, R. Matsuda, S. Kitagawa, J. Solid State Chem. 2005, 178, 24202429. DOI: 10.1016/j.jssc.2005.05.036 [35] W. Clegg, I. R. Little, B. P. Straughan, J. Chem. Soc., Dalton Trans. 1986, 12831288. DOI: 10.1039/DT9860001283 [36] a) G. Li, H. Hou, Z. Li, X. Meng, Y. Fan, New J. Chem. 2004, 28, 15951599. DOI: 10.1039/b406983h; b) N. L. Toh, M. Nagarathinam, J. J. Vittal, Angew. Chem. 2005, 117, 22772281. DOI: 10.1002/ange.200462673 [37] A. F. Wells, Structural Inorganic Chemistry, Clarendon Press, Oxford, 5th ed, 1984. [38] H. Li, M. Eddaoudi, T. L. Groy, O. M. Yaghi, J. Am. Chem. Soc. 1998, 120, 85718572. DOI: 10.1021/ja981669x [39] S.-Y. Yang, L.-S. Long, R.-B. Huang, L.-S. Zheng, S. W. Ng, Acta Cryst, E 2005, E61, m1671m1673. DOI: 10.1107/S1600536805023986 [40] H. F. Clausen, R. D. Poulsen, A. D. Bond, M.-A. S. Chevallier, B. B. Iversen, J. Solid State Chem. 2005, 178, 33423351. DOI: 10.1016/j.jssc.2005.08.013 [41] K. O. Kongshaug, H. Fjellvag, J. Solid State Chem. 2002, 166, 213218. DOI: 10.1006/ jssc.2002.9584 [42] M. Eddaoudi, J. Kim, D. Vodak, A. Sudik, J. Wachter, M. OKeeffe, O. M. Yaghi, Proc. Nat. Acad. Sci. USA 2002, 99, 49004904. DOI: 10.1073/pnas.082051899 [43] R. N. Devi, M. Edgar, J. Gonzalez, A. M. Z. Slawin, D. P. Tunstall, P. Grewal, P. A. Cox, P. A. Wright, J. Phys. Chem. B 2004, 108, 535543. DOI: 10.1021/jp0306949 [44] J. Tao, X. Yin, R. Huang, L. Zheng, S. W. Ng, Inorg. Chem. Commun. 2002, 5, 975977. DOI: 10.1016/S1387-7003(02)00623-8 [45] a) S. A. Bourne, J. Lu, A. Mondal, B. Moulton, M. J. Zaworotko, Angew. Chem. 2001, 113, 21692171. DOI: 10.1002/15213757(20010601)113:11<2169::AID-ANGE2169> 3.0.CO;2-6; b) H. Abourahma, G. J. McManus, B. Moulton, R. D. Bailey-Walsh, M. J. Zaworotko, Macromol. Symp. 2003, 196, 213227. DOI: 10.1002/masy.200390162 [46] B. Moulton, H. Abourahma, M. W. Bradner, J. Lu, G. J. McManus, M. J. Zaworotko, Chem. Commun. 2003, 13421343. DOI: 10.1039/ b301221b [47] P. D. C. Dietzel, R. Blom, H. Fjellvag, Dalton Trans. 2006, 586593. DOI: 10.1039/b509746k [48] D. N. Dybtsev, H. Chun, K. Kim, Angew. Chem. 2004, 116, 51435146. DOI: 10.1002/ange.200460712 [49] X.-L. Wang, C. Qin, E.-B. Wang, Z.-M. Su, Chem. Eur. J. 2006, 12, 26802691. DOI: 10.1002/chem.200501242 [50] M. Edgar, R. Mitchell, A. M. Z. Slawin, P. Lightfoot, P. A. Wright, Chem. Eur. J. 2001, 7, 51685175. DOI: 10.1002/1521-3765(20011203)7:23 <5168::AID-CHEM5168>3.0.CO;2-S

www.cit-journal.de

2007 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Das könnte Ihnen auch gefallen