Cu(II) Coordination Polymer with New

Nov 8, 2012 - In recent years, the design of new coordination polymers has seen an enormous development, becoming a hugely popular research topic that...
0 downloads 0 Views 3MB Size
Communication pubs.acs.org/crystal

Unique Mixed-Valence Cu(I)/Cu(II) Coordination Polymer with New Topology of Bitubular 1D Chains Driven by 1,3,5-Triaza-7phosphaadamantane (PTA) Piotr Smoleński,*,† Julia Kłak,† Dmytro S. Nesterov,‡ and Alexander M. Kirillov*,‡ †

Faculty of Chemistry, University of Wrocław, ul. F. Joliot-Curie 14, 50-383 Wroclaw, Poland Centro de Química Estrutural, Complexo I, Instituto Superior Técnico, Technical University of Lisbon, Av. Rovisco Pais, 1049-001 Lisbon, Portugal



S Supporting Information *

ABSTRACT: The new one-dimensional (1D) mixed-valence copper(I/II) coordination polymer [Cu3(μ-HCOO)2(HCOO)3(μ-PTA)3(PTA)]n·nH2O (1) has been obtained from copper(II) formate, copper(0) powder, and 1,3,5-triaza7-phosphaadamantane (PTA) in MeCN medium. It has been characterized by IR, 1 H, and 31P{H} NMR and EPR spectroscopy, and elemental, differential thermal, powder, and single-crystal X-ray diffraction analysis. The latter features the intricate bitubular 1D chains composed of the novel tetracopper(II) {Cu4(μHCOO)4(HCOO)6}2− blocks that are multiply assembled by copper(I) {μ3Cu(PTA)4}+ units. The topological analysis of 1 discloses a binodal 3,3connected network with the unprecedented topology described by the point symbol of (4·6·8)2(42·6). The magnetic susceptibility of 1 has been studied and described by a model that considers independent contributions of tetracopper(II) {Cu4(μ-HCOO)4(HCOO)6}2− blocks, with global antiferromagnetic exchange between the “outer” (J = −23(2) cm−1) and “inner” (J′ = −112(3) cm−1) pairs of copper atoms. Compound 1 widens the still limited family of PTA-driven coordination polymers, also representing the first homometallic mixed-valence derivative bearing this cagelike aminophosphine building block.

I

found that the N-methylated PTA cores can act as spacers between copper atoms of different valences in a discrete complex.11c As an extension of these studies, the current work aimed at the preparation of a mixed-valence copper(I/II)-PTA coordination polymer, namely by probing the coordination ability of soft P- and hard N-donor atoms of the PTA cages toward the Cu(I) and Cu(II) centers, respectively. To this end, it should be mentioned that a mixed-valence homometallic compound bearing PTA itself has not yet been described before the present study, in spite of a great number of structurally characterized PTA derivatives reported up to date.6 Another objective of the present work consisted in extending the still very limited application of PTA based materials in molecular magnetism.7 Thus, we report herein the facile synthesis and full characterization, X-ray crystal structure, and magnetic behavior of the topologically unique 1D copper(I/II) coordination polymer [Cu3(μ-HCOO)2(HCOO)3(μ-PTA)3(PTA)]n·nH2O (1), which also represents the first homometallic mixed-valence compound with PTA. Thus, the reaction of copper(II) formate and copper(0) powder with PTA, in MeCN medium at ∼25

n recent years, the design of new coordination polymers has seen an enormous development, becoming a hugely popular research topic that crosses the areas of crystal engineering, coordination, and materials chemistry.1,2 From both synthetic and structural aspects, the search for new multidentate organic building blocks toward the construction of functional coordination polymers continues to attract increasing attention, often resulting in novel types of framework materials with intricate topologies and unusual properties.1−4 In contrast to the very common application of N- and/or O-donor organic ligands for the design of coordination polymers,1,2 the use of N,P-aminophosphine building blocks still remains poorly explored.1,5,6 In this regard, the cagelike aminophosphine 1,3,5-triaza-7-phosphaadamantane (PTA) represents a particularly interesting molecule with diamondoid geometry and up to four potential coordination sites.7 However, in spite of the recognized uses in aqueous organometallic chemistry,7 PTA and derived ligands are still rarely employed in crystal engineering,8,9 which is primarily associated with the synthetic difficulties in achieving multiple N,P-coordination modes of PTA cages.7−10 In pursuit of our general research on the PTA chemistry,11 we have recently synthesized new Ag(I)8 and Cu(I)9 coordination polymers, driven by PTA or derived cagelike ligands, which showed interesting structural, topological, photochemical, and biological features.8,9 Besides, we have © 2012 American Chemical Society

Received: August 2, 2012 Revised: October 24, 2012 Published: November 8, 2012 5852

dx.doi.org/10.1021/cg301107k | Cryst. Growth Des. 2012, 12, 5852−5857

Crystal Growth & Design

Communication

°C, leads to the generation of 1 (Scheme 1),12 which has been isolated as a yellow-green air-stable crystalline solid. The

molecular structure of 1 has been established by single crystal X-ray diffraction13 and supported by elemental, differential thermal, and powder X-ray diffraction analysis, IR, NMR, and EPR spectroscopy.12 The IR spectrum of 1 exhibits a set of vibrations in the 1450−550 cm−1 range typical for PTA moieties,11 as well as several highly intense bands due to the vas(COO) [1617−1569 cm−1] and vs(COO) [1340−1303 cm−1] vibrations of η1-, μ2η1-, and μ2-η2-formate ligands.14a Other characteristic vibrations concern the weak v(CH) bands in the 2970−2880 cm−1 range and a medium intensity v(H2O) band with a maximum at 3450 cm−1.14b The differential thermal analysis of 1 under a N2 atmosphere shows five endothermic effects in the 70−370 °C temperature range (Figure S1 of the Supporting Information). The effects in the 70−130, 130−165, and 165−180 °C intervals correspond to the multistep elimination of the crystallization H2O molecule and decomposition of five formate ligands (mass loss of ∼19%). A strong endothermic process in the 180−260 °C range with a maximum at 235 °C is due to the decomposition of PTA ligands (mass loss of ∼65%), whereas the effect in the 270−370 °C without the mass loss is associated with some phase transitions. The total mass loss of ∼84% is

Scheme 1. Synthesis and Structural Formula of 1 (numbers 1, 2, 3, and 4 Correspond to Extensions of Polymeric Motifs)

Figure 1. Structural representations of 1 showing: (a) ellipsoid plot (50% probability) with partial atom labeling scheme, (b) chain fragment composed of repeating Cu4 and Cu6 cyclic motifs with polyhedral representation of the coordination environments around copper atoms, and (c) front and (d) side views of an infinite 1D bitubular metal−organic chain. H atoms and crystallization H2O molecules (a−d) and terminal formate ligands (c, d) are omitted for clarity. Color codes: Cu (green), N (blue), P (orange), O (red), C (cyan). Selected distances (Å): Cu1−N2 2.275(2), Cu1−O1 1.937(2), Cu1−O3 1.990(2), Cu1−O3i 2.032(2), Cu1−O5ii 1.956(2), Cu2−P1 2.2439(7), Cu2−P2 2.2469(8), Cu2−P3 2.2645(8), Cu2−P4 2.2293(8), Cu3−N5 2.085(2), Cu3−N12iv 2.102(2), Cu3−O6 2.051(2), Cu3−O7 2.063(2), Cu3−O9 1.987(2), Cu1···Cu1i 3.1623(6), Cu1···Cu2 6.6518(6), Cu2···Cu3 6.2838(5), Cu1···Cu3ii 5.1854(6) Å. Symmetry codes: (i) −x, −y, −z; (ii) −1−x, −y, −z; (iii) 1 + x, y, z; (iv) −1 + x, y, z. 5853

dx.doi.org/10.1021/cg301107k | Cryst. Growth Des. 2012, 12, 5852−5857

Crystal Growth & Design

Communication

observed in PTA derivatives.7−10 Moreover, polymer 1 is a unique example wherein PTA cages act as spacers between the homometallic centers of different valences. The calculation of the bond valence sum (BVS)17,18 for copper centers in 1 was also performed using the Cu−X bond constants derived previously (see the Supporting Information).18 The BVS analysis resulted in the values of 1.80 (for Cu1), 1.89 (for Cu3), and 1.12 (for Cu2), thus confirming the formal oxidation states of +2 for the Cu1 and Cu3 atoms, and +1 for the Cu2 atom. The obtained BVS values are consistent with the previously reported data for copper coordination compounds18,19 and are in agreement with the oxidation states derived from the charge balance, coordination environments, and magnetic properties of 1. To better understand the intricate structure of bitubular 1D chains in 1, we have performed their topological analysis with the TOPOS software package20 by applying the concept of the simplified underlying net.3,4 Thus, by reducing μ-PTA and μHCOO moieties to their centroids and eliminating all terminal ligands, the network of 1 can be described as an underlying net (Figure 2) constructed from the formally 4-connected Cu1

close to the calculated value (82%), assuming a possible reduction of Cu(II) and Cu(I) to metallic copper during the decomposition of 1.14c In spite of the presence of the paramagnetic Cu(II) centers, the 1H NMR spectrum of 1 in D2O reveals signals of the terminal PTA ligand P-bound to a Cu(I) atom, which are typical for the methylene protons of the NCH2N and PCH2N moieties.11a The 31P{1H} NMR spectrum of 1 in D2O also shows a very broad singlet at δ −85.3, similar to that reported for [Cu(PTAH)4](NO3)5 (PTAH = N-protonated form of PTA).11a It should be mentioned that the observation of resonances in the NMR spectra of Cu(II) compounds bearing PTA and other phosphine ligands is explained by their somewhat remote location from the paramagnetic metal center.11c In contrast, similar ligands connected directly to Cu(II) ions show no signals in the NMR spectra.15 The X-ray crystal structure of 1 (Figure 1) is composed of three symmetry nonequivalent Cu atoms (monovalent Cu2 and divalent Cu1 and Cu3), four PTA and five formate ligands, and one crystallization H2O molecule per formula unit. Within the copper(I) {Cu(PTA)4}+ units, the four-coordinate Cu2 atoms adopt a distorted {CuP4} tetrahedral environment, filled by the P1−P4 atoms of three bridging (P1, P2, P4) and one terminal (P3) PTA ligands, with the bond distances ranging from 2.2293(4) [Cu2−P4] to 2.2645(8) [Cu2−P3] Å. The fivecoordinate Cu3 atoms possess rather distorted {CuN2O3} trigonal bipyramidal geometries taken by the N5 and N12iv atoms from bidentate P,N-coordinated PTA spacers [Cu3−N5 2.085(2), Cu3−N12iv 2.102(2) Å] in apical sites, and three O atoms from bridging (O6) and terminal (O7, O9) formate ligands in equatorial positions [Cu3−O6 2.051(2), Cu3−O7 2.063(2), and Cu3−O9 1.987(2) Å] (hereinafter, all the symmetry codes are those of Figure 1). The five-coordinate Cu1 atoms exhibit distorted {CuNO4} square pyramidal environments, the basal sites of which are occupied by four O atoms from three μ-HCOO (O3, O3i, O5ii) linkers [Cu1− O3 1.990(2), Cu1−O3i 2.032(2), Cu1−O5ii 1.956(2) Å] and one terminal formate ligand (O1) [Cu1−O1 1.937(2) Å], while the axial position is taken by the N2 atom of μ-PTA moiety [Cu1−N2 2.275(2) Å] (Figure 1). The adjacent Cu1 atoms are doubly bridged by the μ2-η1formate ligands forming the {Cu2(μ-HCOO)2(HCOO)2}2− units with the planar Cu1−O3−Cu1i−O3i core [Cu1···Cu1i 3.1623(6) Å]. Two Cu3 centers are additionally bound to this core via the μ2-η2-formate moieties in a syn-anti fashion, thus furnishing the unique tetracopper(II) {Cu4(μ-HCOO)4(HCOO)6}2− blocks wherein the Cu1···Cu3ii and Cu3ii···Cu3iii separations are 5.1854(6) and 12.2233(7) Å, respectively. Interestingly, the observed herein arrangement of four Cu atoms via mixed μ2-η2- and μ2-η1-formate linkers has not yet been detected in any metal-formate compound, as confirmed by the search in the Cambridge Structural Database (CSD).6 In 1, these tetracopper(II) blocks are multiply assembled by the copper(I) {μ3-Cu(PTA)4}+ units into an infinite 1D chain network (Figure 1c). Each of the bitubular chains in 1 (Figure 1d) can alternatively be described as a collection of repeating tetracopper ∼Cu2−Cu3−Cu1−Cu1∼ and hexacopper ∼Cu2− Cu1−Cu3−Cu2−Cu1−Cu3∼ ring motifs (labeled as 4 and 6 in Figure 1b, respectively). The bonding parameters of 1 are comparable to those found in other Cu(I)-PTA9a,11a and Cu(II)-formate16 compounds. A noteworthy feature of 1 also concerns the presence of both bridging and terminal PTA ligands, since mixed coordination modes are still rarely

Figure 2. Topological representations of the simplified underlying 1D network in 1: front (a) and side (b) views of the topologically unique binodal 3,3-connected net with the point symbol of (4·6·8)2(42·6). Color codes: 3-connected Cu2 and Cu3 nodes and 4-connected Cu1 nodes (green); centroids of 2-connected PTA (cyan) and HCOO (red) linkers.

nodes (these are better considered at the 3-connected nodes due to the double linkage of adjacent Cu1 atoms), the 3connected Cu2 and Cu3 nodes, and the 2-connected μ-PTA and μ-HCOO linkers. The topological analysis of the resulting underlying net discloses a binodal 3,3-connected network with a new topology described by the point symbol of (4·6·8)2(42·6), wherein the notation (4·6·8) corresponds to the topologically equivalent Cu2 and Cu3 nodes, while the notation (42·6) refers to the Cu1 nodes. The hitherto undocumented character of the present type of topology has been supported by the search in different databases.20,21 Interestingly, by treating the {Cu2(μ-HCOO)2} moieties as the 4-connected nodes, the network of 1 can alternatively be considered as a binodal 3,4-connected net (Figure S2 of the Supporting Information) with the unknown topology described by the point symbol of (3·6·7)4(32·62·72). Herein, the notations (3·6·7) and (32·62·72) correspond to the topologically equal 3connected Cu2 and Cu3 nodes, and the 4-connected (Cu1)2 nodes, respectively. Although binodal 3,3- and/or 3,4connected networks are frequently encountered in 2D and 3D metal−organic structures,3,4 new topologies are very rarely 5854

dx.doi.org/10.1021/cg301107k | Cryst. Growth Des. 2012, 12, 5852−5857

Crystal Growth & Design

Communication

The magnetic properties of 1 were investigated over the 1.8− 300 K temperature range. Plots of the magnetic susceptibility χm and χmT product vs T (χm is the molar magnetic susceptibility for 4 Cu(II) ions) are given in Figure 3b. The absence of a maximum in the χm curve may indicate that the possible antiferromagnetic coupling is weak. At 300 K, the χmT equals 1.47 cm3 mol−1 K, which roughly corresponds to the value expected for the four uncoupled copper(II) ions [χmT = 3(Nβ2g2/3k)S(S + 1) = 1.49 cm3 mol−1 K with g = 2.1 and S = 1 /2, where N, β, g, k, S, and T have their usual meaning.24 Upon cooling, the χmT decreases continuously, reaching a plateau with a value of 0.846 cm3 mol−1 K in the 50−3 K temperature range. This plateau would denote that the ground state is dominated by an antiferromagnetic coupling. The slight decrease of the χmT below 3 K is most likely due to the weak antiferromagnetic interactions between the tetracopper(II) entities. To confirm the nature of the ground state of 1, we have investigated the variation of the magnetization (M) with respect to the field (H), at 2 K. The results are shown in Figure 3c, where the molar magnetization M (per Cu4 entities) is expressed in μB units. The magnetization curve for the four uncoupled spins S = 1/2 was derived from the equation M = gβSNBs(x), where Bs(x) is the Brillouin function and x = gβSH/ kT.24 As can be seen (Figure 3c), the calculated magnetization level is significantly higher than that observed for 1. This feature agrees with the global antiferromagnetic coupling within the four copper(II) ions. The magnetic susceptibility of 1 has been analyzed by a model that takes into consideration the interactions within the tetracopper(II) {Cu4(μ-HCOO)4(HCOO)6}2− entities (Figure 3a). Hence, this model analyzes the global susceptibility, χm (eq 1), as the sum of two independent contributions, namely one due to the tetracopper(II) blocks with the S = 1/2 spins [χm(tetr)] and the other one due to eventually paramagnetic monomeric impurities [χm(para)], in addition to a possible temperature independent term [χ(Nα)], with a typical value for copper(II) ion of 60 × 10−6 cm3 mol−1.

found in 1D coordination polymers. To this end, the structure of 1 contributes to the identification and classification of metal−organic materials with novel topologies, also showing that rather intricate topologies can be observed even in lowdimensional networks. The EPR spectra of powdered samples of 1 recorded in the X-band are quite different at 293 and 77 K (Figure S3 of the Supporting Information). At 293 K, only a broad isotropic signal with the g value of ∼2.16 has been observed, whereas at 77 K a rhombic spectrum with g1 = 2.06, g2 = 2.18, and g3 = 2.26 has been detected.22 This spectral shape could be compatible with the presence of two magnetically nonequivalent copper(II) centers in the ligand fields of the distorted square pyramidal {Cu(1)NO4} and trigonal bipyramidal {Cu(3)N2O3} geometries (Figure 3a).23 Thus, the resulting crystal g-tensor does not reflect the local crystal field symmetry of the copper(II) compound.

χm = (1 − x)χm(tetr) + xχm(para) + χNα

(1)

The χm(tetr) contribution derives from the Heisenberg Hamiltonian for the Cu4 unit with S = 1/2 (eq 2):25 H = −J(S1̂ S2̂ + S3̂ S4̂ ) − J ′S2̂ S3̂

(2)

where J is the exchange coupling constant between the “outer” pair (Cu1···Cu3) of copper atoms and J′ is that between the “inner” pair (Cu1···Cu1) (Figure 3a). An exact solution for such a type of model has been previously described.25 The magnetic susceptibility per tetracopper(II) unit, χm(tetr), derived from the van Vleck formula assuming an equal g value for the four copper(II) ions, is given by eq 3: Ng 2β 2 4k(T − Θ) 10 exp(2u) + 2 exp( −2u) + 4 exp( −2υ) 5 exp(2u) + 3 exp( −2u) + 6 exp( − 2υ) + exp(− 4υ)

χm(tetr) =

Figure 3. (a) Structural fragment showing the tetracopper(II) {Cu4(μHCOO)4(HCOO)6}2− block in 1 with the polyhedral representation of the coordination environments around Cu atoms and the magnetic coupling scheme (color codes are those of Figure 1). (b) Temperature dependence of the experimental χmT (○) and χm (●) for 1 (χm per 4 Cu(II) ions); the solid line is the calculated curve derived from eq 1. (c) Field dependence of the magnetization (M per Cu4 entities) for 1 (○); the solid line is the Brillouin function curve for the system of four uncoupled spins with S = 1/2 and g = 2.0.

(3)

where υ = J/kT and u = J′/kT. The contribution χm(para) is expected to follow a Curie−Weiss model for the S = 1/2 spins (eq 4): 5855

dx.doi.org/10.1021/cg301107k | Cryst. Growth Des. 2012, 12, 5852−5857

Crystal Growth & Design χm(para) = Ng 2β 2 /4kT

Communication

topology, this compound is the unique mixed-valence derivative wherein PTA cages act as spacers between the Cu(I) and Cu(II) centers. Besides, the magnetic susceptibility studies have disclosed the antiferromagnetic nature of 1, thus opening up the application of Cu-PTA derivatives in molecular magnetism. The nature and magnitude of the magnetic couplings have been discussed on the basis of structural parameters and compared with those of related carboxylato-bridged copper(II) compounds. The features observed herein should be accounted for in further developments, namely focused on the exploration of aminophosphine PTA cages as versatile diamondoid P,Nbuilding blocks toward the design of functional metal−organic materials.

(4)

The paramagnetic monomeric impurity χm(para) was introduced to obtain satisfactory results.24 In addition, a Curie− Weiss term (Θ) was included in eq 3 to account for a slight decrease of the χmT in the lowest temperature domain. A leastsquares fitting of the magnetic data in the whole temperature range leads to J = −23(2) cm−1, J′ = −112(3) cm−1, g = 2.19(1), Θ = −0.31(1) K, and χm(para) = 3.06(1)%, as indicated by the solid curve in Figure 3b. The criterion used for the determination of the best fit was based on the minimization of the sum of squares of the deviation, R = ∑(χexpT − χcalcT)2/ ∑(χexpT)2 [R = 7.36 × 10−6]. The calculations show the global antiferromagnetic exchange between the “outer” and “inner” pairs of copper atoms. It is worthwhile to note that the magnitude of J [−23(2) cm−1] is rather high considering the quite long distance of 5.1854(6) between the respective Cu(II) centers. The magnitude and sign of the exchange coupling J in the carboxylato-bridged copper(II) complexes depends on the type of bridging mode, the coordination geometry of copper(II) centers, the Cu···Cu separations, and the bond angles at bridging atoms.26−28 In contrast to the rather common μ2-η2coordination of carboxylate ligands (Cu−O−C−O−Cu), the μ 2-η 1-modes (Cu−O−Cu) are rarely observed in Cucarboxylate derivatives. Consequently, only a few examples of systems with the μ2-η1-COO ligands have been magnetostructurally characterized.26 In this respect, it may be recalled that the dinuclear bis(hydroxo)- or bis(alkoxo)-bridged copper(II) compounds show a correlation between J and the Cu−O−Cu angle.28 The antiferromagnetic interaction is usually found when the Cu−O−Cu angle is larger than 97.5°, whereas a ferromagnetic interaction is expected at a smaller Cu−O−Cu angle.29a In 1, the Cu1−O3−Cu1 angles are equal to 103.67(8)°, thus suggesting an antiferromagnetic exchange. As mentioned above, the magnetic behavior may be understood in terms of the nature of the orbitals involved in the exchange interactions, together with the structural characteristics of the bridging network (conformation of the bridge). In 1, the Cu3 atoms possess a distorted trigonal bipyramidal geometry with the dz2 ground state, whereas the Cu1 centers are square pyramidal with the dx2−y2 ground state. It is also known that a change in electron density of the magnetic orbital can have a pronounced effect on the sign and magnitude of a magnetic exchange interaction.27,29 When the copper(II) geometry is close to square pyramidal, a strong antiferromagnetic interaction is often predominant, and the reduction of an antiferromagnetic contribution is observed when the geometry becomes closer to trigonal bipyramidal. The antiferromagnetic coupling between the Cu1···Cu1i centers is a result of the interaction of the dx2−y2 orbitals positioned in the square base via the formate oxygen atoms. The overlap between the magnetic orbitals of the Cu1 and Cu3 atoms through the synanti carboxylate bridge is significantly weaker, presumably arising from a more trigonal character of the Cu3 environments and the electron delocalization on the dz2 magnetic orbitals. In summary, the present work has extended the use of PTA as an important but still underexplored cagelike building block in crystal engineering, resulting in the easy synthesis and full characterization of the new 1D coordination polymer [Cu3(μHCOO)2(HCOO)3(μ-PTA)3(PTA)]n·nH2O (1) driven by μPTA and μ-formate linkers. Apart from featuring an intricate bitubular chain network with a hitherto undocumented



ASSOCIATED CONTENT

S Supporting Information *

Materials and methods, refinement details for X-ray analysis, supporting references, differential thermal analysis plot (Figure S1), additional topological representations (Figure S2), EPR spectra (Figure S3), powder X-ray diffraction patterns (Figure S4), and crystallographic file in CIF format for 1. This material is available free of charge via the Internet at http://pubs.acs.org.



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected] (A.M.K.), piotr.smolenski@chem. uni.wroc.pl (P.S.), Phone: +351 218419207. Fax: +351 218464455. Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This work was supported by the KBN program (Grant No. N204 280438), Poland, and by the Foundation for Science and Technology (FCT) (Projects PTDC/QUI-QUI/121526/2010 and PEst-OE/QUI/UI0100/2011), Portugal. We thank Mrs. K. Ptak for her contribution to this research and Mr. M. Siczek for X-ray measurement.



REFERENCES

(1) For recent state-of-the-art books on the topic, see: (a) Design and Construction of Coordination Polymers; Hong, M.-C., Chen, L., Eds.; Wiley: New York, 2009. (b) Batten, S. R.; Turner, D. R.; Neville, S. M. Coordination Polymers: Design, Analysis and Application; Royal Society of Chemistry: London, 2009. (c) Metal-Organic Frameworks: Design and Application; MacGillivray, L. R., Ed.; Wiley-Interscience: New York, 2010. (d) Functional Metal-Organic Frameworks: Gas Storage, Separation and Catalysis; Schroder, M., Ed.; Springer: New York, 2010. (e) Metal-Organic Frameworks: Applications from Catalysis to Gas Storage; Farrusseng, D., Ed.; Wiley-VCH: Weinheim, 2011. (2) For selected recent reviews, see: (a) Stock, N.; Biswas, S. Chem. Rev. 2012, 112, 933. (b) Paz, F. A. A.; Klinowski, J.; Vilela, S. M. F.; Tomé, J. P. C.; Cavaleiro, J. A. C.; Rocha, J. Chem. Soc. Rev. 2012, 41, 1088. (c) Leong, W. L.; Vittal, J. J. Chem. Rev. 2011, 111, 688. (d) Meek, S. T.; Greathouse, J. A.; Allendorf, M. D. Adv. Mater. 2011, 23, 249. (e) Kirillov, A. M. Coord. Chem. Rev. 2011, 255, 1603. (f) Janiak, C.; Vieth, J. K. New J. Chem. 2010, 34, 2366. (g) Fromm, K. M.; Sagué, J. L.; Mirolo, L. Macromol. Symp. 2010, 291−292, 75. (h) Perry, J. J., IV; Perman, J. A.; Zaworotko, M. J. Chem. Soc. Rev. 2009, 38, 1400. (i) Tranchemontagne, D. J.; Mendoza-Cortes, J. L.; O’Keeffe, M.; Yaghi, O. M. Chem. Soc. Rev. 2009, 38, 1257. (j) Qiu, S.; Zhu, G. Coord. Chem. Rev. 2009, 253, 2891. (3) (a) Blatov, V. A.; Proserpio, D. M. In Modern Methods of Crystal Structure Prediction; Oganov, A. R., Ed.; Wiley: 2010; pp 1−28. 5856

dx.doi.org/10.1021/cg301107k | Cryst. Growth Des. 2012, 12, 5852−5857

Crystal Growth & Design

Communication

(b) Blatov, V. A.; O’Keeffe, M.; Proserpio, D. M. CrystEngComm 2010, 12, 44. (c) Alexandrov, E. V.; Blatov, V. A.; Kochetkova, A. V.; Proserpio, D. M. CrystEngComm 2011, 13, 3947. (4) O’Keeffe, M.; Yaghi, O. M. Chem. Rev. 2012, 112, 675. (5) James, S. L. Chem. Soc. Rev. 2009, 38, 1744. (6) See the Cambridge Structural Database (CSD, version 5.33, May 2012): Allen, F. H. Acta Crystallogr. 2002, B58, 380. (7) For reviews on PTA chemistry, see: (a) Bravo, J.; Bolãno, S.; Gonsalvi, L.; Peruzzini, M. Coord. Chem. Rev. 2010, 254, 555. (b) Phillips, A. D.; Gonsalvi, L.; Romerosa, A.; Vizza, F.; Peruzzini, M. Coord. Chem. Rev. 2004, 248, 955. (8) (a) Kirillov, A. M.; Wieczorek, S. W.; Guedes da Silva, M. F. C.; Sokolnicki, J.; Smoleński, P.; Pombeiro, A. J. L. CrystEngComm 2011, 13, 6329. (b) Lis, A.; Guedes da Silva, M. F. C.; Kirillov, A. M.; Smoleński, P.; Pombeiro, A. J. L. Cryst. Growth Des. 2010, 10, 5244. (c) Kirillov, A. M.; Wieczorek, S. W.; Lis, A.; Guedes da Silva, M. F. C.; Florek, M.; Król, J.; Staroniewicz, Z.; Smoleński, P.; Pombeiro, A. J. L. Cryst. Growth Des. 2011, 11, 2711. (9) (a) Jaremko, Ł.; Kirillov, A. M.; Smoleński, P.; Pombeiro, A. J. L. Cryst. Growth Des. 2009, 9, 3006. (b) Kirillov, A. M.; Smoleński, P.; Haukka, M.; Guedes da Silva, M. F. C.; Pombeiro, A. J. L. Organometallics 2009, 28, 1683. (10) (a) Lidrissi, C.; Romerosa, A.; Saoud, M.; Serrano-Ruiz, M.; Gonsalvi, L.; Peruzzini, M. Angew. Chem., Int. Ed. 2005, 44, 2568. (b) Serrano-Ruiz, M.; Romerosa, A.; Sierra-Martin, B.; FernandezBarbero, A. Angew. Chem., Int. Ed. 2008, 47, 8665. (c) Tu, X.; Truong, H.; Nichol, G. S.; Zheng, Z. Inorg. Chim. Acta 2010, 363, 4189. (d) Koutmos, M.; Georgakaki, I. P.; Tsiolis, P.; Coucouvanis, D. Z. Anorg. Allg. Chem. 2008, 634, 255. (e) Mohr, F.; Falvello, L. R.; Laguna, M. Eur. J. Inorg. Chem. 2006, 3152. (11) (a) Kirillov, A. M.; Smoleński, P.; Guedes da Silva, M. F. C.; Pombeiro, A. J. L. Eur. J. Inorg. Chem. 2007, 2686. (b) Kirillov, A. M.; Smoleński, P.; Ma, Z.; Guedes da Silva, M. F. C.; Haukka, M.; Pombeiro, A. J. L. Organometallics 2009, 28, 6425. (c) Kirillov, A. M.; Filipowicz, M.; Guedes da Silva, M. F. C.; Kłak, J.; Smoleński, P.; Pombeiro, A. J. L. Organometallics 2012, 31, doi:10.1021/om3005564. (d) Jaremko, Ł.; Kirillov, A. M.; Smoleński, P.; Lis, T.; Pombeiro, A. J. L. Inorg. Chem. 2008, 47, 2922. (e) Smoleński, P.; Pombeiro, A. J. L. Dalton Trans. 2008, 87. (12) Synthesis and analytical data of [Cu3(μ-HCOO)2(HCOO)3(μPTA)3(PTA)]n·nH2O (1). To a solution of Cu(HCOO)2·H2O (5.0 mmol, 0.858 g) in acetonitrile (100 mL) was added an excess of metallic copper as a fine powder (30 mmol, 1.906 g). The reaction mixture was stirred at room temperature (rt, ∼25 °C) for several hours, resulting in a gradual color change from green to colorless, and then was filtered off. Solid PTA (10 mmol, 1.571 g) was added to the obtained filtrate, leading to an off-white suspension. This was stirred overnight at rt and filtered off. The filtrate was left in a vial for several days to evaporate in air at rt, producing yellow-green X-ray quality single crystals, which were collected and dried in air to furnish 1 in 35.4% yield (1.409 g), based on the copper(II) formate. Anal. Calcd for C29H55Cu3N12O11P4 (Mr = 1062.4 g/mol), %: C, 32.79; H, 5.22; N, 15.82. Found, %: C, 32.81; H, 5.13; N, 15.99. 1H NMR (D2O, rt) δ/ppm: 4.11 (s, 6H, PCH2N), 4.57 (s br, 6H, NCH2N). 31P{1H} NMR (D2O, rt) δ/ppm: −85.3 (s br). IR (KBr, cm−1): 3450 (m br) ν(H2O), 2924 (m) and 2904 (m) ν(CH), 1617 (vs br) and 1569 (s) νas(COO), 1447 (w), 1413 (w), 1340 (s) and 1303 (s) νs(COO), 1238 (m), 1102 (m), 1017 (s), 972 (m), 947 (m), 818 (w), 788 (m), 743 (w), 619 (m), and 584 (m). (13) Crystal data: 1: C29H55Cu3N12O11P4, M = 1062.4, triclinic, space group P1,̅ a = 12.1664(4), b = 13.0392(5), c = 13.1697(5) Å, α = 93.064(3), β = 107.195(3), γ = 96.601(3)°, V = 1974.20(12) Å3, T = 100(2) K, Z = 2, Dcalcd = 1.787 Mg m−3, μ = 1.838 mm−1, F(000) = 1096, Θ = 4.82 − 30.00, GoF = 0.908, 18498 reflections measured, 11246 unique (Rint = 0.0343), R1 (7121 reflections with I > 2σ(I)) = 0.0379, wR2 = 0.0782, largest difference peak, hole: 1.82, −1.12 e Å−3. (14) (a) Gibson, D. H.; Ding, Y.; Miller, R. L.; Sleadd, B. A.; Mashuta, M. S.; Richardson, J. F. Polyhedron 1999, 18, 1189. (b) Nakamoto, K. Infrared and Raman Spectra of Inorganic and

Coordination Compounds, 5th ed.; Wiley: New York, 1997. (c) Galwey, A. K.; Jamieson, D. M.; Brown, M. E. J. Phys. Chem. 1974, 78, 2664. (15) Tu, X.; Nichol, G. S.; Zheng, Z. J. Cluster Sci. 2009, 20, 93. (16) (a) Boonmak, J.; Youngme, S.; Chaichit, N.; van Albada, G. A.; Reedijk, J. Cryst. Growth Des. 2009, 9, 3318. (b) Rebilly, J.-N.; Bacsa, J.; Rosseinsky, M. J. Chem. Asian J. 2009, 4, 892. (c) Kong, D.; Zon, J.; McBee, J.; Clearfield, A. Inorg. Chem. 2006, 45, 977. (d) Casarin, M.; Corvaja, C.; Di Nicola, C.; Falcomer, D.; Franco, L.; Monari, M.; Pandolfo, L.; Pettinari, C.; Piccinelli, F. Inorg. Chem. 2005, 44, 6265. (17) O’Keeffe, M.; Brese, N. E. J. Am. Chem. Soc. 1991, 113, 3226. (18) Shields, G. P.; Raithby, P. R.; Allen, F. H.; Motherwell, W. D. S. Acta Crystallogr. 2000, B56, 455. (19) (a) Das, O; Adarsh, N. N.; Paul, A.; Paine, T. K. Inorg. Chem. 2010, 49, 541. (b) Li, B.; Zhao, J.-W.; Zheng, S.-T.; Yang, G.-Y. Inorg. Chem. 2009, 48, 8294. (20) Blatov, V. A. IUCr CompComm Newsl. 2006, 7, 4. (21) The Reticular Chemistry Structure Resource (RCSR) Database; O’Keeffe, M.; Peskov, M. A.; Ramsden, S. J.; Yaghi, O. M. Acc. Chem. Res. 2008, 30, 1782. (22) (a) Hathaway, B. J. In Comprehensive Coordination Chemistry; Wilkinson, G., Gill, R. D., McCleverty, J. A., Eds.; Pergamon Press: Oxford, 1987; Vol. 5. (b) Hathaway, B. J. J. Chem. Soc., Dalton Trans. 1972, 1196. (23) Pilbrow, J. R. Transition Ion Electron Paramagnetic Resonance; Clarendon Press: Oxford, 1990. (24) (a) Kahn, O. Molecular Magnetism; VCH Publishers: New York, 1993. (25) Rubenacker, G. V.; Drumheller, J. E.; Emerson, K.; Willett, R. D. J. Magn. Magn. Mater. 1986, 54−57, 1483. (26) (a) Costes, J.-P.; Dahan, F.; Laurent, J.-P. Inorg. Chem. 1985, 24, 1018. (b) Chiari, B.; Hatfield, W. E.; Piovesana, O.; Tarantelli, T.; Ter Haar, L. W.; Zanazzi, P. F. Inorg. Chem. 1983, 22, 1468. (c) Chiari, B.; Helms, J. H.; Piovesana, O.; Tarantelli, T.; Zanazzi, P. F. Inorg. Chem. 1986, 25, 870. (d) Escrivà, E.; Server-Carrió, J.; Lezama, L.; Folgado, J.-V.; Pizarro, J. L.; Ballesteros, R.; Abarca, B. J. Chem. Soc., Dalton Trans. 1997, 2033. (27) Youngme, S.; Chailuecha, C.; Albada, G. A.; Pakawatchai, C.; Chaichit, N.; Reedijk, J. Inorg. Chim. Acta 2005, 358, 1068. (28) (a) Youngme, S.; Phatchimkun, J.; Wannarit, N.; Chaichit, N.; Meejoo, S.; Albada, G. A.; Reedijk, J. Polyhedron 2008, 27, 304. (b) Barquín, M.; González Garmendia, M. J.; Larrínaga, L.; Pinilla, E.; Torres, M. R. Inorg. Chim. Acta 2006, 359, 2424. (c) Das, A.; Todorov, I.; Dey, S. K.; Mitra, S. Inorg. Chim. Acta 2006, 359, 2041. (d) Rodríguez, M.; Llobet, A.; Corbella, M.; Müller, P.; Usón, M. A.; Martell, A. E.; Reibenspies, J. J. Chem. Soc., Dalton Trans. 2002, 2900. (e) Chiari, B.; Piovesana, O.; Tarantelli, T.; Zanazzi, P. F. Inorg. Chem. 1993, 32, 4834. (f) Astheimer, H.; Nepveu, F.; Walz, L.; Haase, W. J. Chem. Soc., Dalton Trans. 1985, 315. (g) Walz, L.; Paulus, H.; Haase, W.; Langhof, H.; Nepveu, F. J. Chem. Soc., Dalton Trans. 1983, 657. (h) Laurent, J.-P.; Bonnet, J.-J.; Nepveu, F.; Astheimer, H.; Walz, L.; Haase, W. J. Chem. Soc., Dalton Trans. 1982, 2433. (29) (a) Crawford, V. H.; Richardson, H. W.; Wasson, J. R.; Hodgson, D. J.; Hatfield, W. E. Inorg. Chem. 1976, 15, 2107. (b) Haddad, M. S.; Wilson, S. R.; Hodgson, D. J.; Hendrickson, D. N. J. Am. Chem. Soc. 1981, 103, 384.

5857

dx.doi.org/10.1021/cg301107k | Cryst. Growth Des. 2012, 12, 5852−5857