Spontaneous resolution of a 3D chiral polyoxometalate-based polythreaded framework consisting of an achiral ligand

Ya-Qian Lan , Shun-Li Li , Zhong-Min Su *, Kui-Zhan Shao , Jian-Fang Ma *, Xin-Long Wang and En-Bo Wang *
Institute of Functional Material Chemistry, Key Laboratory of Polyoxometalate Science of the Ministry of Education, Faculty of Chemistry, Northeast Normal University, Changchun, 130024, China. E-mail: zmsu@nenu.edu.cn; jianfangma@yahoo.com.cn; wangenbo@public.cc.jl.cn; Tel: +86 431 85099108

Received (in Cambridge, UK) 18th September 2007 , Accepted 6th November 2007

First published on 14th November 2007


Abstract

Two enantiomerically 3D chiral POM-based architectures have been constructed based on the achiral ligand bbi, [V10O26]4− polyoxoanion and mixed valence Cu(I/II) without a chiral auxiliary, and they represent the first examples of enantiomerically 3D POM-based compounds using achiral ligands.


The design and construction of chiral polyoxometalates (POMs) is currently of great interest because of their intriguing potential applications, for instance in medicine and asymmetric heterogeneous catalysis.1 To date, only a limited number of chiral POM-based compounds have been observed, and most of them are discrete or low-dimensional structures.2 The rational synthesis of chiral high-dimensional frameworks constructed from POM units is still a great challenge.

In general, chiral compounds can be obtained by two routes. The first method is based on the use of chiral species (chiral organic linkers or chiral metal complexes) as structure directing agents.3 This method has been developed for the formation of chiral POM-based compounds. Recently, thanks to the work of Pope, Hill, Yamase, and Kortz, much prominent work has been performed using this approach,4 and our group has obtained two chiral POM-based 3D architectures from this approach.5 One of them is constructed from [BW12O40]5−clusters and copper–proline acid complexes.5a The chirality of the amino acid is transferred to the whole framework through the bonding of the metal centers to the POM cluster. The other was achieved using a typical Waugh heteropolymolybdate [MnMo9O32]6− as a chiral polyoxoanion building block,5b thereby making the chirality of the polyoxoanion transfer to the whole framework.

The second approach to creating chiral compounds is based on the use of achiral ligands under spontaneous resolution without any chiral auxiliary,6 which yields a conglomerate. A conglomerate is a mechanical and racemic mixture of chiral crystals, in which each crystal is enantiopure and crystallizes in a chiral space group.7 Using this method to synthesize chiral POM-based 3D architectures is more difficult given the following considerations: (i) How is a chiral molecular unit generated from achiral components without a chiral auxiliary? (ii) It is much more difficult to achieve interlinking of the chiral molecular units into a homochiral structure of higher dimensionality and to induce spontaneous resolution. (iii) How can the high symmetry of the POM be destroyed and the structure be crystallised in a non-centrosymmetric space group? Therefore, the search for suitable achiral organic molecules, metal cations, and POM building units to be applied in the construction of chiral 3D frameworks is one of the most challenging issues in synthetic chemistry and materials science.

Based on the previous literature,8 we chose the flexible ligand 1,1′-(1,4-butanediyl)bis(imidazole) (bbi) as a achiral ligand to synthesize chiral POM-based 3D frameworks. Fortunately, compounds [Cu(bbi)2V10O26][Cu(bbi)]2·H2O (L-1 and D-1) can be isolated by hydrothermal methods and separated manually. Single-crystal X-ray diffraction analyses§ reveal that the structures of compounds L-1 and D-1 are enantiomers. Their unit-cell dimensions, volumes, related bond distances, and angles are only slightly different. To the best of our knowledge, they represent the first example of enantiomerically pure chiral 3D POM-based polythreaded frameworks which are constructed from achiral ligands under spontaneous resolution without any chiral auxiliary. Each of them contains one [V10O26]4− polyoxoanion, three kinds of Cu cations and six kinds of bbi ligands (bbi1, bbi2, bbi3, bbi4, bbi5 and bbi6) in the crystallographically independent unit (Fig. S1, S2). The bbi3–6 ligands are imposed on the crystallographically twofold symmetry. The [V10O26]4− polyoxoanion is an oblate spherical mixed valence structure, which has eight vertex-sharing V(V)O4 tetrahedra forming a cyclic octavanadate and two V(IV)O5 tetragonal pyramids positioned above and below as crowns of cyclic octavanadate with V–O distances similar to those in the reported complexes (Fig. S3).9 Due to the strong Jahn–Teller effect of the d9 electronic configuration,10 the Cu1(II) center adopts a distorted octahedral geometry which is completed by four nitrogen atoms from two bbi1 and two bbi2 ligands, and two terminal oxygen atoms from different [V10O26]4− polyoxoanions (Fig. S6a), while Cu2(I) is linearly coordinated by two nitrogen atoms from the bbi3 and bbi4 ligands. Cu3(I) exhibits a similar coordination environment to Cu2(I) except that the two nitrogen atoms are from the bbi5 and bbi6 ligands. The bbi1 and bbi2 units act as bis-monodentate bridging ligands, and coordinate to Cu1 ions to generate a 2D waved (4,4) sheet which is pillared by [V10O26]4− polyoxoanions to extend to a distorted 3D α-Po topology framework (Fig. 1, Fig. S4). It is interesting that there exists a flower-like subunit in which two bbi1 and two bbi2 ligands act as petals and the [V10O26]4− polyoxoanion acts as the stamen (Fig. S5). This kind of connection mode results in the formation of two kinds of channels (A and B) along the c axis. Cu2 and Cu3 are coordinated by the bbi3 and bbi4, and bbi5 and bbi6 ligands respectively to form two kinds of distorted helical chains. Two Cu2–bbi chains pass through channel A, and two Cu3–bbi chains are inserted in channel B along the crystallographic c axis (Fig. 2). As a result, the unusual chiral 3D polythreaded framework has been formed, by two helical chains threading through a distorted α-Po skeleton (Fig. S7a). Polythreaded structures which have been defined by Ciani and co-workers are characterized by the presence of closed loops, as well as of elements that can thread through the loops, and can be considered as extended periodic analogues of molecular rotaxanes and pseudorotaxanes.11 A few coordination compounds exhibiting polythreaded network architectures have been reported.12 However, a 3D chiral polyoxometalate-based polythreaded framework has not been observed. On further study of the structure, there are Cu–O interactions between CuI ions and two terminal oxygen atoms from the [V10O26]4− polyoxoanions (Fig. S6b). So the overall structures of L-1 and D-1 are 3D (3,4,6)-connected frameworks based on CuI ions as three-connected nodes, [V10O26]4− polyoxoanions as four-connected nodes and CuII ions as six-connected nodes (Fig. S6). The Schläfli symbols for L-1 and D-1 are (6.82)2(65.8) (44.68.83) (Fig. S7b).


(a) Ball–stick representation of the (4,4) sheet formed by bbi ligands and Cu(ii). (b) Polyhedral and ball–stick representation of each [V10O26]4−polyanion linking two Cu1 cations.
Fig. 1 (a) Ball–stick representation of the (4,4) sheet formed by bbi ligands and Cu(II). (b) Polyhedral and ball–stick representation of each [V10O26]4−polyanion linking two Cu1 cations.

Space-filling representation of the two kinds of channels in the α-Po framework (left), and view of the two kinds of Cu(i)–bbi helical chains in the channels along the different directions (right) in d-1.
Fig. 2 Space-filling representation of the two kinds of channels in the α-Po framework (left), and view of the two kinds of Cu(I)–bbi helical chains in the channels along the different directions (right) in D-1.

Even though spontaneous resolution was discovered as early as in 1848 by Louis Pasteur,13 it is still a rare phenomenon and cannot be predicted a priori because the laws of physics determining the processes are not yet fully understood.14 What is the reason that L-1 and D-1 can be constructed from achiral bbi ligands and high symmetry [V10O26]4− polyoxoanions by spontaneous resolution upon crystallization without any enantiopure chiral auxiliary? Further study into the nature of these intricate architectures reveals that the bbi ligands in these compounds adopt two conformations. The bbi1, bbi2, bbi4, and bbi5 adopt a TGT conformation, and the bbi3 and bbi6 adopt GTG conformation (T = trans, G = gauche, Fig. S2). Each CuI is linked linearly by the bbi ligands with different GTG and TGT conformations to generate conformationally chiral chains with a C2 axis (left-handed helices in L-1, right-handed helices in D-1) (Fig. 3a). In addition, [V10O26]4− polyoxoanions are linked by two CuII ions via covalent linkages and two CuI ions through Cu–O interactions (Fig. S6b). The asymmetrical coordination mode without mirror or inversion symmetry between the [V10O26]4− polyoxoanions and the Cu cations not only destroys the symmetry of the [V10O26]4− polyoxoanions but also generates the [CuIICuI2V10O26] chain with a 2-fold screw axis passing through the CuII ions and the center of the [V10O26]4− polyoxoanions (Fig. 8a). Four distorted helical chains with conformational chirality are linked by the [CuIICuI2V10O26] chain via Cu–O interactions, which are related by the 2-fold screw axis and hence of the same chirality. As a result, the POM-based chiral units [CuIICuI2(bbi)2V10O26] are formed (left-handed in L-1, right-handed in D-1) (Fig. 3b, Fig. S8). The adjacent POM-based chiral units are extended into the whole framework by the bbi1 and bbi2 ligands with the same conformation (TGT) bridging CuII cations (Fig. S9). All chiral units are related by translation or 2-fold rotation, so the whole structure is homochiral and spontaneous resolution occurs. The generation of the crystals of L-1 and D-1 represents two new examples of induction and transfer of chirality in the absence of chiral auxiliaries, a progression from simple achiral species to homochiral chains, to POM-based homochiral units [CuIICuI2(bbi)2V10O26] and finally to whole chiral crystals. Obviously, the twist of the bbi ligands induces the initial generation of the chirality of the distorted helical chains, and the asymmetrical coordination mode destroys the high symmetry of the POMs. Chiral preservation in the whole structure is achieved via non-covalent Cu–O interactions between the CuI ions and the [V10O26]4− polyoxoanions. The non-covalent interactions, such as Cu–O interactions, may play a crucial role in the process of spontaneous resolution.


(a) View of bbi–Cu(i) distorted helical chains (left-handed helix in l-1, right-handed helix in d-1). (b) Ball–stick representation of the chiral subunit [CuIICuI2(bbi)2V10O26]∞ (left in l-1, right in d-1; dashed lines represent Cu(i)–O interactions).
Fig. 3 (a) View of bbi–Cu(I) distorted helical chains (left-handed helix in L-1, right-handed helix in D-1). (b) Ball–stick representation of the chiral subunit [CuIICuI2(bbi)2V10O26] (left in L-1, right in D-1; dashed lines represent Cu(I)–O interactions).

The CD spectra in DMSO have been investigated and display strong Cotton effects at 234 and 258 nm for L-1, and 233 and 259 nm for D-1, respectively (Fig. S10–S11). However, electrospray ionization mass spectra of L-1 and D-1 in DMSO show that they decompose to some small fragments (Fig. S12). The interesting results from the CD spectra indicate these small fragments retain their chirality, which may suggest there exist stable interactions among small fragments in solution.

In summary, we have prepared and characterized two enantiomerically 3D chiral POM-based architectures based on the achiral ligand bbi, [V10O26]4− polyoxoanion and mixed valence Cu(I/II) without a chiral auxiliary. To our knowledge, compounds L-1 and D-1 represent the first example of enantiomerically 3D POM-based compounds using achiral ligands, and they are also unique 3D chiral POM-based polythreaded frameworks which are constructed from two helical chains threading through the distorted α-Po skeleton. The successful isolation of this species not only provides an intriguing example of enantiomerically pure architectures but also may open up possibilities for synthesizing new chiral POM-based materials with particular functions by using achiral ligands.

We are thankful for financial support from the Program for Changjiang Scholars and Innovative Research Team in University and the Science Foundation for Young Teachers of Northeast Normal University (No. 20070309).

Notes and references

  1. (a) A. Müller, F. Peters, M. T. Pope and D. Gatteschi, Chem. Rev., 1998, 239 CrossRef; (b) D. A. J. Judd, H. Nettles, N. Nevins, J. P. Snyder, D. C. Liotta, J. Tang, J. Ermolieff, R. F. Schinazi and C. L. Hill, J. Am. Chem. Soc., 2001, 123, 886 CrossRef CAS.
  2. (a) T. A. Sullens, R. A. Jensen, T. Y. Shvareva and T. E. Albrecht-Schmitt, J. Am. Chem. Soc., 2004, 126, 2676 CrossRef CAS; (b) S. Bareyt, S. Piligkos, B. Hasenknopf, P. Gouzerh, E. Lacôte, S. Thorimbert and M. Malacria, Angew. Chem., Int. Ed., 2003, 42, 3404 CrossRef CAS; (c) M. Sadakane, M. H. Dickman and M. T. Pope, Inorg. Chem., 2001, 40, 2715 CrossRef CAS; (d) V. Soghomonian, Q. Chen, R. C. Haushalter, J. Zubieta and C. J. O'Connor, Science, 1993, 259, 1596; (e) M. H. Alizadeh, K. T. Holman, M. Mirzaei and H. Razavi, Polyhedron, 2006, 25, 1567 CrossRef CAS; (f) S. T. Zheng, J. Zhang, J. Q. Xu and G. Y. Yang, J. Solid State Chem., 2005, 178, 3740 CrossRef CAS; (g) S. T. Zheng, J. Zhang and G. Y. Yang, J. Mol. Struct., 2005, 752, 25 CrossRef CAS.
  3. (a) A. von Zelewsky and U. Knof, Angew. Chem., Int. Ed., 1999, 38, 302 CrossRef; (b) C. J. Kepert, T. J. Prior and M. J. Rosseinsky, J. Am. Chem. Soc., 2000, 122, 5158 CrossRef CAS; (c) T. J. Prior and M. J. Rosseinsky, Inorg. Chem., 2003, 42, 1564 CrossRef CAS; (d) T. Ezuhara, K. Endo and Y. Aoyama, J. Am. Chem. Soc., 1999, 121, 3279 CrossRef CAS; (e) Y. Cui, H. L. Ngo and W. B. Lin, Chem. Commun., 2003, 1388 RSC; (f) R.-G. Xiong, X.-Z. You, B. F. Abrahams, Z.-L. Xue and C.-M. Che, Angew. Chem., Int. Ed., 2001, 40, 4422 CrossRef CAS.
  4. (a) F. B. Xin and M. T. Pope, J. Am. Chem. Soc., 1996, 118, 7731 CrossRef CAS; (b) M. Inoue and T. Yamase, Bull. Chem. Soc. Jpn., 1995, 68, 3055 CAS; (c) M. Inoue and T. Yamase, Bull. Chem. Soc. Jpn., 1996, 69, 2863 CAS; (d) D. L. Long, P. Kögerler, L. J. Farrugia and L. Cronin, Chem.–Asian J., 2006, 1, 352 CrossRef CAS; (e) D. L. Long, E. Burkholder and L. Cronin, Chem. Soc. Rev., 2007, 36, 105 RSC; (f) C. Streb, D. L. Long and L. Cronin, Chem. Commun., 2007, 471 RSC; (g) U. Kortz, M. G. Savelieff, F. Y. A. Ghali, L. M. Khalil, S. A. Maalouf and D. I. Sinno, Angew. Chem., Int. Ed., 2002, 41, 4070 CrossRef CAS; (h) X. K. Fang, T. M. Anderson and C. L. Hill, Angew. Chem., Int. Ed., 2005, 44, 3540 CrossRef CAS; (i) X. K. Fang, T. M. Anderson, Y. Hou and C. L. Hill, Chem. Commun., 2005, 5044 RSC; (j) L. M. Zheng, T. Whitfield, X. Wang and A. J. Jacobson, Angew. Chem., Int. Ed., 2000, 39, 4528 CrossRef CAS.
  5. (a) H. Y. An, E. B. Wang, D. R. Xiao, Y. G. Li, Z. M. Su and L. Xu, Angew. Chem., Int. Ed., 2006, 45, 904 CrossRef CAS; (b) H. Q. Tan, Y. G. Li, Z. M. Zhang, C. Qin, X. L. Wang, E. B. Wang and Z. M. Su, J. Am. Chem. Soc., 2007, 129, 10066 CrossRef CAS.
  6. (a) T. Matsuura and H. Koshima, J. Photochem. Photobiol., C, 2005, 6, 7 CrossRef CAS; (b) D. K. Kondepudi, R. Kaufman and N. Singh, Science, 1990, 250, 975 CrossRef CAS; (c) X. D. Chen, M. Du and T. C. W. Mak, Chem. Commun., 2005, 4417 RSC; (d) M. Sasa, K. Tanaka, X.-H. Bu, M. Shiro and M. Shionoya, J. Am. Chem. Soc., 2001, 123, 10750 CrossRef CAS; (e) U. Siemeling, I. Scheppelmann, B. Neumann, A. Stammler, H. G. Stammler and J. Frelek, Chem. Commun., 2003, 2236 RSC; (f) F. M. Tabellion, S. R. Seidel, A. M. Arif and P. J. Stang, Angew. Chem., Int. Ed., 2001, 40, 1529 CrossRef CAS; (g) Y. T. Wang, M. L. Tong, H. H. Fan, H. Z. Wang and X. M. Chen, Dalton Trans., 2005, 424 RSC; (h) Q. Z. Sun, Y. Bai, G. J. He, C. Y. Duan, Z. H. Lin and Q. J. Meng, Chem. Commun., 2006, 2777 RSC; (i) S. Jayanty and T. P. Radhakrishnan, Chem.–Eur. J., 2004, 10, 2661 CrossRef CAS.
  7. (a) L. Pérez-García and D. B. Amabilino, Chem. Soc. Rev., 2002, 31, 342 RSC; (b) L. Pérez-García and D. B. Amabilino, Chem. Soc. Rev., 2007, 36, 941 RSC; (c) J. Jacques, A. Collet and S. H. Wilen, Enantiomers, Racemates and Resolutions, Krieger Publishing Company, Malabar, Florida, 1994 Search PubMed.
  8. (a) P. C. M. Duncan, D. M. L. Goodgame, S. Menzer and D. J. Williams, Chem. Commun., 1996, 2127 RSC; (b) L. Ballester, I. Baxter, P. C. M. Duncan, D. M. L. Goodgame, D. A. Grachvogel and D. J. Williams, Polyhedron, 1998, 17, 3613 CrossRef CAS; (c) J. F. Ma, J. F. Liu, Y. Xing, H. Q. Jia and Y. H. Lin, J. Chem. Soc., Dalton Trans., 2000, 2403 RSC; (d) B. X. Dong, J. Peng, C. J. Gómez-García, S. Benmansour, H. Q. Jia and N. H. Hu, Inorg. Chem., 2007, 46, 5933 CrossRef CAS.
  9. (a) A. Bino, S. Cohen and C. Heitner-Wirguin, Inorg. Chem., 1982, 21, 429 CrossRef CAS; (b) Y. Hayashi, N. Miyakoshi, T. Shinguchi and A. Uehara, Chem. Lett., 2001, 170 CrossRef CAS; (c) Y. G. Li, Y. Lu, G. Y. Luan, E. B. Wang, Y. B. Duan, C. W. Hu, Ni. H. Hu and H. Q. Jia, Polyhedron, 2002, 21, 2601 CrossRef CAS.
  10. M. Fujita, Chem. Soc. Rev., 1998, 27, 417 RSC.
  11. (a) L. Carlucci, G. Ciani and D. M. Proserpio, Coord. Chem. Rev., 2003, 246, 247 CrossRef CAS; (b) S. R. Batten and R. Robson, Angew. Chem., Int. Ed., 1998, 37, 1460 CrossRef.
  12. (a) L. Carlucci, G. Ciani and D. M. Proserpio, Chem. Commun., 1999, 449 RSC; (b) M. Du, X. J. Jiang and X. J. Zhao, Chem. Commun., 2005, 5521 RSC; (c) X. L. Wang, C. Qin, E. B. Wang, Y. G. Li, Z. M. Su, L. Xu and L. Carlucci, Angew. Chem., Int. Ed., 2005, 44, 5824 CrossRef CAS.
  13. L. Pasteur, Ann. Chim. Phys., 1848, 24, 442 Search PubMed.
  14. (a) E.-Q. Gao, Y.-F. Yue, S.-Q. Bai, Z. He and C.-H. Yan, J. Am. Chem. Soc., 2004, 126, 1419 CrossRef CAS; (b) C. He, Y. G. Zhao, D. Guo, Z. H. Lin and C. Y. Duan, Eur. J. Inorg. Chem., 2007, 3451 CrossRef CAS.

Footnotes

Electronic supplementary information (ESI) available: Bond valence sums, additional figures, TGA, CD, ESI-MS and UV-Vis spectra. See DOI: 10.1039/b714413j
Synthesis of compound 1: A mixture of NH4VO3 (0.117 g, 1 mmol), bbi (0.095 g, 0.5 mmol), Cu(NO3)2·3H2O (0.121 g, 0.5 mmol) and H2O (10 mL) was adjusted to approximately pH 5 with dilute Et3N solution and hydrochloric acid and stirred for 1 h, and then transferred to and sealed in a 25 mL Teflon-lined stainless steel container, which was heated at 150 °C for 72 h and then cooled to room temperature at a rate of 10 °C h−1. Black crystals of 1 were collected in 42.6% yield based on Cu(NO3)2·3H2O. Elemental analysis (%) calcd for compound 1: C, 25.35; H, 3.08; N, 11.83. Found: C, 25.41; H, 3.01; N, 11.89%. IR (cm−1): 3121 (w), 1635 (m), 1522 (m), 1450 (w), 1278 (w), 1233 (w), 1104 (m), 890 (s), 828 (s), 655 (s).
§ Crystal data for C40H58N16Cu3V10O27 (L-1): orthorhombic, space groupI222, Mr = 1895.04, a = 23.0670(10) Å, b = 23.1760(10) Å, c = 24.9480(9) Å, V = 13337.2(9) Å3, Z = 8, μ = 2.353 mm−1, Dc = 1.888 Mg m−3, F(000) = 7544, T = 293(2) K, 37072 reflections collected, 13213 unique with Rint = 0.0596, R1 = 0.0539, wR2 = 0.1121 (I > 2σ(I)), GOF = 1.045. Crystal data for C40H58N16Cu3V10O27 (D-1): orthorhombic, space groupI222, Mr = 1895.04, a = 23.0670(14) Å, b = 23.1760(14) Å, c = 24.9480(14) Å, V = 13337.2(9) Å3, Z = 8, μ = 2.353 mm−1, Dc = 1.888 Mg m−3, F(000) = 7544, T = 293(2) K, 33818 reflections collected, 11790 unique with Rint = 0.0841, R1 = 0.0549, wR2 = 0.1107 (I > 2σ(I)), GOF = 1.006. Data collection was performed on a Bruker Smart Apex CCD diffractometer (Mo Kα, λ = 0.71073 Å). Empirical absorption corrections were applied. The Flack parameters of 0.017(17) and 0.02(2) for L-1 and D-1 indicate that the absolute configurations are correct. CCDC 658057 (L-1) and CCDC 658056 (D-1). For crystallographic data in CIF or other electronic format see DOI: 10.1039/b714413j

This journal is © The Royal Society of Chemistry 2008