Hydrogen-Bonded Helical Array, Sodium-Ion-Mediated Head-to-Tail

Hydrogen-Bonded Helical Array, Sodium-Ion-Mediated Head-to-Tail Chain, and ..... The asymmetric unit in 1 contains one bowl-like TCAS tetracobalt(II) ...
0 downloads 0 Views 945KB Size
CRYSTAL GROWTH & DESIGN

Hydrogen-Bonded Helical Array, Sodium-Ion-Mediated Head-to-Tail Chain, and Regular Ionic Bilayer: Structural Diversities of p-Sulfonatothiacalix[4]arene Tetranuclear Cluster Units

2009 VOL. 9, NO. 3 1584–1589

Mingyan Wu, Wei Wei, Qiang Gao, Daqiang Yuan, Yougui Huang, Feilong Jiang,* and Maochun Hong* State Key Laboratory of Structural Chemistry, Fujian Institute of Research on the Structure of Matter, Chinese Academy of Sciences, Fujian, Fuzhou, 350002 China ReceiVed NoVember 3, 2008; ReVised Manuscript ReceiVed December 5, 2008

ABSTRACT: Exploration into the supramolecular chemistry of p-sulfonatothiacalix[4]arene (TCAS) tetranuclear clusters with methylviologen dihexafluorophosphate (MV(PF6)2) and 4,4′-diaminodiphenyl ether (4,4′-Dadpe) has resulted in three interesting structural motifs: {[MV][Co4(TCAS)(µ4-SO4)(H2O)8] · 8H2O}n (1), {[Na(H2O)2][Co4(TCAS)(µ5-SO4)(H2O)7(4,4′-HDadpe)] · 6H2O}n (2), and {[MV][Cu4(TCAS)(µ4-SO4)(H2O)4] · 9H2O}n (3). In complex 1, the assembly of TCAS tetracobalt(II) clusters driven by multiple hydrogen bonding forms supramolecular helices along three crystallographic axes, while TCAS tetracobalt(II) clusters in complex 2 have assembled into a rare head-to-tail chain mediated by sodium ions in the presence of 4,4′-Dadpe. Complex 3 features a regular ionic bilayer structure, in which TCAS tetracopper(II) clusters and MV2+ cations form negative and positive layers, respectively. Introduction Of the major classes of water-soluble supramolecular host compounds, particular attention has been paid to the p-sulfonatocalix[4]arenes for their stable conelike conformation and ability to form a variety of inclusion and/or coordination complexes in both the solution- and solid-state.1 Usually, in these supramolecular complexes the p-sulfonatocalix[4]arenes pack themselves into up-and-down layers, that is, claylike structures.2 Many factors, such as guests with different shapes and sizes, supramolecular interactions, and metal ions, are able to heavily affect the process of assembly. For instance, in the presence of suitable guest molecules, a lot of surprising structural motifs, such as capsules,3 “Ferris wheels”,4 and Russian dolls, are available. The notable example is the spectacular nanometerscale tubules and spheroids reported by Atwood and his coworkers.5 In fact, weak supramolecular interactions, such as π · · · π interactions, C-H · · · π interactions, and especially the hydrogen bonding, can also have an effect on the ultimate species. For example, Atwood reported a stable hydrogen bonding polymer, in which the diprotonated 4,4-bipyridine molecules induce the expected conelike conformation of the p-sulfonatocalix[4]arene into the 1,3-altenate conformation. Additionally, in the process of self-assembly the metal ions also play an important role. They not only can be coordinated by the calixarenes to stabilize and/or extend the structures6 but also form the coordinated complexes with other species to act as the guest complexes.7 Furthermore, the aquated metal ions may hydrogen bond with calixarenes to form interesting structure motifs.8 However, up to now, research is mainly focused on the interactions of the metal ions with the upper rims (sulfonate groups) and the hydrophobic cavities,1-8 while few examples are involved in the interactions between the metal ions and the lower rims (phenolic groups).9 This may be due to the strong acidic reaction environments and the weak coordination ability of phenolic groups to the metal ions. * To whom correspondence should be addressed. E-mail: [email protected]; tel: +86-591-83792460; fax: +86-591-83792460.

Recently, p-sulfonatothiacalix[4]arene (TCAS),10 a new water-soluble calixarene, has attracted much attention. Though being similar to the p-sulfonatocalix[n]arene in shape and structure, p-sulfonatothiacalix[4]arene differs greatly from its counterpart and should be regarded as a new molecular framework rather than a simple substitute of p-sulfonatocalix[n]arene.11 One unique characteristic of p-sulfonatothiacalix[4]arene is that at its lower rim the four sulfur atoms along with four phenolic groups are capable of chelating various divalent transition metal ions to form tetranuclear anionic clusters through synergistic coordination.12 Our research interests largely focus on how these divalent tetranuclear anionic clusters assemble with divalent or the potential divalent cationic complexes for charge balance. Just recently, we obtained two intriguing two-dimensional (2D) includedcoordinationpolymers,{[Co(2,2′-bpz)(H2O)4]2+⊂[Co4(TCAS)(µ4-SO4)(H2O)4]2- · 10.75H2O}n (4) and {[Cu(2,2′bpz)(H2O)3]2+⊂[Cu4(TCAS)(µ4-SO4)(H2O)4]2- · 16H2O}n (5) (2,2′-bpz ) 2,2′-bipyrazine) through self-assembly, in which metal-2,2′-bpz complexes are included in the hydrophobic cavities of TCAS tetranuclear cluster units.13 To further our research, we have investigated the self-assembly between the TCAS tetranuclear cluster units and methylviologen dihexafluorophosphate (MV(PF6)2) or 4,4′-diaminodiphenyl ether (4,4′-Dadpe). Herein, we wish to report the syntheses and novel structures of the hydrogen-bonded helical array {[MV][Co4(TCAS)(µ4-SO4)(H2O)8] · 8H2O}n (1), head-to-tail chain {[Na(H2O)2][Co4(TCAS)(µ5-SO4)(H2O)7(4,4′-HDadpe)] · 6H2O}n (2), and regular ionic bilayer {[MV][Cu4(TCAS)(µ4SO4)(H2O)4] · 9H2O}n (3), and discuss the factors that affect the process of assembly of the TCAS anionic clusters with the above-mentioned cationic complexes. Experimental Section Materials and General Methods. The Na4H4TCAS were prepared according to literature methods,14 while other starting chemicals and solvent were commercially available with analytical purity and were used without further purification. The IR spectra as KBr disk were recorded on a Magna 750 FT-IR spectrophotometer. C, H, and N

10.1021/cg801226p CCC: $40.75  2009 American Chemical Society Published on Web 01/06/2009

p-Sulfonatothiacalix[4]arene Tetranuclear Clusters

Crystal Growth & Design, Vol. 9, No. 3, 2009 1585

Table 1. Crystallographic Details for Complexes 1-3

CCDC deposit no. formula fw crystal system space group a, Å b, Å c, Å R, deg β, deg γ, deg V, Å3 Z Dc, g/cm-3 µ, mm-1 F(000) crystal size, mm3 θ range, deg no. of collected/unique reflns COF final R indices (I > 2σ(I)) R indices (all data)

1

2

3

CCDC-613526 C36H54Co4N2O36S9 1615.07 orthorhombic P212121 14.069(9) 15.909(11) 26.93(3) 90 90 90 6027(8) 4 1.780 1.492 3296 0.60 × 0.60 × 0.50 3.26, 27.48 45924/13735 1.089 0.0301 0.0311

CCDC-613527 C36H51Co4N2NaO36S9 1635.04 monoclinic P21/n 18.513(2) 16.8148(15) 20.190(2) 90 112.333(5) 90 5813.7(11) 4 1.868 1.555 3328 0.35 × 0.30 × 0.25 2.38, 27.48 43409/13230 1.047 0.0449 0.0505

CCDC-613528 C36H48Cu4N2O33S9 1579.46 triclinic P1j 13.7598(8) 14.4813(10) 14.6317(10) 104.327(4) 90.613(4) 95.779(2) 2808.6(3) 2 1.868 1.927 1604 0.25 × 0.22 × 0.10 3.17, 27.48 21827/12657 1.051 0.0461 0.0570

elemental analyses were determined on an Elementary Vario ELIII elemental analyzer. Syntheses. {[MV][Co4(TCAS)(µ4-SO4)(H2O)8] · 8H2O}n, 1. An aqueous solution containing Na4H4TCAS (45 mg, 0.05 mmol), methylviologen dihexafluorophosphate (MV(PF6)2, 22 mg, 0.05 mmol), and cobalt(II) sulfate heptahydrate (56 mg, 0.2 mmol) was alkalified using 0.05 M sodium hydroxide until pH ) 3.0. Over several days, large prismatic orange crystals formed which were suitable for X-ray diffraction. Yield 30 mg, 36.47% based on Na4H4TCAS. Anal. Calcd. For 1, C36H54Co4N2O36S9 (1615.07): C, 26.77%; H, 3.37%; N, 1.73%. Found: C, 26.81%; H, 3.35%; N, 1.74%. IR: 3444.31 s, 3061.71 m, 1644.12 s, 1567.20 m, 1454.55 s, 1189.46 s, 1141.20 s, 1039.64 s, 877.72 m, 824.25 m, 757.60 s, 667.18 s. {[Na(H2O)2][Co4(TCAS)(µ5-SO4)(H2O)7(4,4′-HDadpe)] · 6H2O}n, 2. Na4H4TCAS (45 mg, 0.05 mmol), 4,4′-diaminodiphenyl ether (4,4′Dadpe, 10 mg, 0.05 mmol), and cobalt(II) sulfate heptahydrate (56 mg, 0.2 mmol) was dissolved in 4 mL of water, and the resulting solution was acidfied using 0.05 M HNO3 until pH ) 3.0. About one week later, large prismatic orange crystals formed that were suitable for X-ray diffraction. Yield 28 mg, 34.25% based on Na4H4TCAS. Anal. Calcd. For 2, C36H51Co4N2NaO36S9 (1635.04): C, 26.45%; H, 3.14%; N, 1.71%, Found: C, 26.50%; H, 3.11%; N, 1.72%. IR: 3268.21 s, 1634.92 m, 1574.99 m, 1499.35 s, 1450.15 s, 1203.06 s, 1036.58 s, 877.70 m, 832.94 m, 756.04 s, 611.09 s. {[MV][Cu4(TCAS)(µ4-SO4)(H2O)4] · 9H2O}n, 3. In the very beginning, we tried to synthesize 3 using the preparation method analogous to 1, but what we got was a yellow precipitation formed immediately as soon as MV(PF6)2 in water was mixed to the aqueous solution of Na4H4TCAS and CuSO4 · 6H2O (pH ) 3.0). Thus, we changed our experimental method. A solution of MV(PF6)2 (22 mg, 0.05 mmol) in 5 mL of CH3OH/H2O (4:1) solution was carefully layered on a solution of Na4H4TCAS (45 mg, 0.05 mmol) and CuSO4 · 6H2O (53 mg, 0.20 mmol) in 5 mL of water (pH ) 3.8). Between these two layers, there is a H2O/CH3OH (3 mL, 2:1) buffer. Over a period of ca. three weeks, orange prism crystals formed that were suitable for X-ray diffraction analyses. Yield 26 mg, 32.92% based on Na4H4TCAS. Anal. Calcd. For 3, C36H48Cu4N2O33S9 (1579.46): C, 27.38%; H, 3.06%; N, 1.77%, Found: C, 27.45%; H, 3.01%; N, 1.79%. IR: 3433.68 s, 1641.88 s, 1571.85 m, 1455.19 s, 1190.11 s, 1036.65 s, 879.57 m, 825.33 m, 753.03 m, 621.98 s. X-ray Data Collection and Structure Determination. Data collection for 1-3 was performed on a Rigaku-CCD diffractometer equipped with a graphite monochromated Mo KR radiation (λ ) 0.71073 Å) by using the ω-scan mode at 293 K. All absorption corrections were applied using the CrystalClear program.15 The structures were solved by direct methods, the metal atoms were located from the E-maps, and other non-hydrogen atoms were derived from the successive difference Fourier peaks. The organic hydrogen atoms were positioned geometrically, and allowed to ride on their parent C or N atoms. The hydrogen atoms on the water molecules were located

in difference density maps and were refined as riding using the instruction AFIX 3, and no attempt was made to locate the hydrogen atoms of disorder water molecules. The structure was refined on F2 by full-matrix least-squares using the SHELXTL-97 program package.16 A summary of the crystallographic data of complexes 1-3 was presented in Table 1. CCDC-613526, 613527, 613528 for 1, 2, and 3, respectively.

Results and Discussion Hydrogen Bonded Helical Array. {[MV][Co4(TCAS)(µ4SO4)(H2O)8] · 8H2O}n, 1. Treatment of Na4H4TCAS, MV(PF6)2, and cobalt(II) sulfate in an aqueous solution afforded crystals of 1. The single-crystal X-ray analysis reveals that 1 crystallizes in the chiral space group P212121. The asymmetric unit in 1 contains one bowl-like TCAS tetracobalt(II) cluster, one MV2+ dication, and eight lattice water molecules. The fully deprotoned TCAS retains its cone conformation and captures four cobalt(II) ions at the lower rim through the four bridge sulfur atoms as well as the four phenoxy oxygen atoms (Figure 1). Interestingly, one sulfate dianion in 1 completes the base of the bowl-like tetracobalt(II) cluster in a rare tridentate coordination mode.12 In addition, the water molecules coordinate to each of the four cobalt(II) ions in the equatorial planes to meet the demands for the distorted octahedral coordination environments. At the upper rim, one MV2+ dication is not capsulated by two TCAS molecules as reported before,3h but is intercalated into the hydrophobic pocket of TCAS with its one side, through CH · · · π interactions (closest C · · · aromatic centroid distance, 3.423 Å). This supramolecular assembly is in a slant manner and the angle between the long axis of the MV2+ dication and the plane of the sulfur bridges is 22.52°. However, the other side of the MV2+ dication is out of the cavity since it is too long for the size of the cavity. On crystal packing, the thiacalixarenes of 1 extend their structures through complex hydrogen bonding, which do not arrange themselves into classical bilayers.2 Significant observation on this crystal structure is that the helical assemblies are formed along each of the crystallographic directions. Along the crystallographic a axis, the tetranuclear cluster subunit twists its structure through two similar intermolecular hydrogen bonds (Figure S1, Supporting Information). The lattice water molecule O29 acts as a single donor to the noncoordinated oxygen atom O20 of the sulfate anion, and as an acceptor to the coordinated

1586 Crystal Growth & Design, Vol. 9, No. 3, 2009

Wu et al.

Figure 2. Coordination environment of cobalt(II) ions in complex 2. The hydrogen atoms and lattice water molecules are omitted for clarity. Symmetric codes: A 1/2 + x, 1/2 - y, 1/2 + z. Figure 1. The asymmetric unit of complex 1. For clarity, the lattice water molecules and hydrogen atoms are omitted.

water molecule O21i (O20 · · · H29B-O29 · · · H21Bi-O21i). The separations of O20 · · · O29 and O29 · · · O21i are 2.982 and 2.673 Å, and the bond angles of O20 · · · H29B-O29 and O29 · · · H21Bi-O21i are 147.25 and 173.43°, respectively. Similar to the above O29, the lattice water molecule O31 bridges O24 and O20i (O24-H24B · · · O31-H31B · · · O20i), with the O24 · · · O31 (2.626 Å) and O31 · · · O20i (3.011 Å), and the bond angles for O24-H24B · · · O31 (177.09°) and O31-H31B · · · O20i (156.06°). The supramolecular assemblies are also formed along the crystallographic b axis, which are stabilized by three kinds of intermolecular hydrogen bonds (Figure S2,Supporting Information). The coordinated water molecules O26 and O25 both act as the single donors to interact with O2i and O4i, respectively. Furthermore, the lattice water O29 also acts as a double donor to link O20 and O4i (O20 · · · H29BO29-H29A · · · O4i). The two 21 symmetric superstructures along the orthogonal crystallographic directions are accompanied by a 21 symmetric structure along the third direction (Figure S3,Supporting Information). Along the crystallographic c axis, a coordinated water molecule O22 donates a hydrogen bond to an oxygen atom O5i of sulfonate group, with the O22 · · · O5i separation of 2.708 Å and the O22-H22A · · · O5i bond angle of 168.74°. The above helical packings in 1 are much different from those reported by Raston et al.,17 in which the TCAS molecules arrange themselves into bilayer-type helical chains along one crystallographic direction through C-H · · · π and π-stacking interactions. In 1, the TCAS molecules pack themselves fully by hydrogen-bond interactions and selfassemble into helices along each of the crystallographic directions. Compared with the hydrogen-bonded helices reported in the literature,18 the helical association found in 1 is unprecedented for any other metal-organic complexes because (i) the helical assembly is driven by multiple hydrogen bonding; (ii) the subunit includes a tetracobalt(II) cluster; (iii) the hydrogen bonded helices are found along each of the crystallographic directions; (iv) the participants in hydrogen bond interactions are various, including the lattice water molecules, the aqua ligands, the sulfonate groups, and the noncoordinated oxygen atoms of the sulfate anions. At the molecular level, there must

be an autoresolution process, where hydrogen bonding interactions may play an important role during the crystallization. Comparing 1 with the two included polymers (4 and 5) we reported before,13 the effect of the guest molecules on the assembly of the tetranuclear clusters in 1 is obvious. In 4 and 5, the metal-2,2′-bpz cations are more spherical compared with the MV2+ dication in 1, so that they are well suited to the cyathiform cavity of the TCAS. Meanwhile, the strong hydrogen bonding between the aquo ligands of metal-2,2′-bpz cations and the TCAS further stabilizes the process of including. More importantly, the lengths of metal-2,2′-bpz cations (6.82 and 6.78 Å) are shorter than that of MV2+ dication (9.90 Å) in 1, so that they have less effect on the assembly of tetranuclear clusters. Certainly, in 1 the complex hydrogen bonding also facilitates the breaking-down of the layer-like structures. As a result, the tetranuclear clusters here do not pack into classical up-and-down layers like those in the above-mentioned included polymers, but isolate themselves, as shown in the structure of 1. Head-to-TailChain.{[Na(H2O)2][Co4(TCAS)(µ5-SO4)(H2O)7(4,4′HDadpe)] · 6H2O}n, 2. Slow evaporation of aqueous mixtures of cobalt(II) sulfate, Na4H4TCAS, and 4,4′-Dadpe in a ratio of 4:1:1 afforded a crystalline complex 2. The single-crystal X-ray analysis reveals that 2 crystallizes in the space group P21/n. In the tobacco-pipe-like asymmetric unit, the TCAS retains its cone conformation and captures four cobalt(II) ions at the lower rim through the four bridge sulfur atoms as well as the four phenol oxygen atoms (Figure 2). Interestingly, one sulfate dianion completes the base of the bowl-like tetracobalt(II) cluster. Except Co1, all other three Co(II) ions are in a distorted octahedral SO5 environment and the coordinated spheres are completed by aqua ligands in the equatorial plane. For Co1, the 4,4′HDadpe acts as a terminal ligand rather than a bridging ligand to complete its SO4N coordination sphere and the uncoordinated amino group is protonated for the charge balance. The sodium ion, which is in a distorted octahedral environment of O6, is bound at the axis by two sulfonate groups with an O-Na-O angle 177.13°, while a peculiar feature is that the sodium ion is bound in the equatorial plane by an aqueous ligand (O20A) and an oxygen atom (O26A) of the sulfate anion belonging to the neighboring asymmetric unit. The resulting structure contains a polymeric chain along the line 101, where the calixarenes are parallel with respect to this axis (Figure 3). Certainly, such

p-Sulfonatothiacalix[4]arene Tetranuclear Clusters

Figure 3. Stick representation of the hydrogen-bonded polymeric headto-tail chain along the line [101] for complex 2. The hydrogen bonds are shown in dashed red lines.

uncommon head-to-tail columns have been documented in the unsubstituted calixarenes, in which the larger alkali metal ions such as rubidium and cesium ions interacted with calixarenes through two extreme coordination modes: the exo phenolic oxygen-metal coordination bonding and the endo π-cation bonding.19 Meanwhile, similar head-to-tail self-assemblies can also be obtained from uncomplexed calixarenes through other feeble noncovalent interactions, such as C-H · · · π interactions,20 π-π interactions,21 and hydrogen bond interactions.22 Up to now, there are no examples of such stable head-to-tail metalcalixarene coordination polymers reported in the literature. It should be noted that the sulfate anion adopts a rare η1:µ2:µ2 tridentate coordination mode to bridge four cobalt(II) ions and a sodium ion. Another feature of this complex is that there are rich hydrogen bond interactions. First, the column is stabilized by complex intramolecular and intermolecular hydrogen bonds, which are originated from the lattice water molecules, the aqua ligands, the oxygen atoms of sulfonate groups, and the nitrogen atoms of the 4,4′-Dadpes. Second, the parallel columns in the same direction are linked by two kinds of intermolecular hydrogen bonds into a 2D network in the ac plane (Figure S4, Supporting Information, left). And last, adjacent layers in the different directions are also hydrogen bonded to form a threedimensional (3D) network containing alternate layers (Figure S4,Supporting Information, right). In our previous efforts to control the assembly of the teranuclear clusters, we have introduced bridged ligand 4,4bipyridine into this system and obtained a layer-like coordination polymer 6, in which two monoprotonated 4,4-bipyridine molecules coordinate to two cobalt(II) ions of a tetranuclear cluster in monodentate mode respectively.12a Comparing 2 with 6, though in 2 the bridged ligand 4,4′-Dadpe is also monoprotonated and coordinates to one cobalt(II) ion, the 4,4′-HDadpe forms complex hydrogen bonds with other molecule such as aquo ligands and free water molecules. However, the more important factor is the coordination of sodium(I) ion, which directs the formation of the column-like motif. Regular Ionic Bilayer. [MV][Cu4(TCAS)(µ4-SO4)(H2O)4] · 9H2O}n, 3. Diffusing a methanol solution of MV(PF6)2 into an aqueous solution containing Na4H4TCAS and copper(II) sulfate afforded crystalline complex 3. In the asymmetric unit of 3,

Crystal Growth & Design, Vol. 9, No. 3, 2009 1587

Figure 4. The coordination environment of copper(II) ions in complex 3. The hydrogen atoms and lattice water molecules are omitted for clarity. Symmetric codes: A 2 - x, 1 - y, 1 - z; B 2 - x, 1 - y, -z; C 1 - x, 1 - y, -z; D 1 - x, 1 - y, 1 - z.

there are one divalent tetranuclear cluster anion, one MV2+ dication, and nine lattice water molecules. Like those in complexes 1 and 2, the fully deprotoned TCAS maintains a bowl-like conformation and the phenoxy-bridged cluster of four copper(II) ions are formed at the bottom of the bowl. The sulfate dianion uses its three oxygen atoms to ligate four copper(II) centers and completes the base of the tetranuclear cluster unit. All copper(II) ions are in distorted octahedral environments, and each of them completes the coordination sphere by one aquo ligand and the oxygen atoms from sulfonate groups of adjacent four TCAS (Figure 4). As the counterion, the MV2+ dication twists its two aromatic rings with a dihedral angle 30.60°, which is slightly larger than that in 1 (20.88°). Interestingly, the MV2+ dication does not reside in the cavity of the TCAS as in complex 1 and shows no obvious interactions with the thiacalixarene. On packing, the thiacalixarenes adopt a head-to-tail layer arrangement so that the copper(II) ions at the lower rims can easily be coordinated by the sulfonate groups at the upper rims of adjacent thiacalixarenes. The Cu(II)-O (sulfonato) distances range from 2.011 to 2.444 Å, which arises probably from the Jahn-Teller effect. Thus, each anionic tetranuclear cluster subunit acts as a four-connector and extends the structure into a conventional up and down claylike layer in the ac plane. Careful examination of the aromatic distances between the adjacent thiacalixarenes shows the additional π-π interactions, with the separations of the parallel aromatic planes 3.39, 3.54, 3.35, and 3.31 Å, respectively. In addition, the coordinated water molecules hydrogen bond with the sulfonate groups to stabilize the structure (Figure S5,Supporting Information). The character of complex 3 is that the MV2+ dications also arrange themselves into a positive layer between the anionic layers. As a result, the negative and positive layers stack alternately along the crystallographic b axis (Figure 5). For 1 and 3, the main reason for the formation of their different structural motifs may result from the coordination abilities of cobalt(II) and copper(II) ions. For the Jahn-Teller effect, coordination bonding of copper(II) ions to the sulfonate groups is strong and dominantly induces to form the polymeric up-and-down layer. The MV2+ dication, which cannot break

1588 Crystal Growth & Design, Vol. 9, No. 3, 2009

Wu et al.

conformations from C4V to lower symmetries to contain various guests with different sizes, shapes, and charges. Conclusion Combining the present results and those reported earlier, we can conclude that the assemblies of p-sulfonatothiacalix[4]arene tetranuclear clusters with divalent or potential divalent cationic complexes are heavily affected by not only the metal ions but also by the shape of counterions as well as by noncovalent interand intramolecular forces, such as hydrogen bonding, π-π stacking, and C-H · · · π interactions (Scheme 1). We think the above discussion will help us in the design and construction in the near future of other kinds of complexes based on tetranuclear cluster units and higher symmetric complexes (C3 and C3V), such as nanospheres and porous complexes.

Figure 5. The negative and positive layers stack alternately along the crystallographic b axis in 3.

Scheme 1. The Various Structural Motifs of TCAS Tetranuclear Clusters with Divalent and Potential Divalent Cationic Complexes in the Presence of Different Metal Ions

Acknowledgment. We are thankful for financial support from 973 Program (2006CB932900), the Young Scientist Funds of Fujian Province (2008F305010147), the National Natural Science Foundation of China (20571074), and the Young Scientist Funds of Chinese Academy of Sciences. Supporting Information Available: Crystallographic information files (CIF) for 1-3 are available free of charge via the Internet at http:// pubs.acs.org.

References

the arrangement of the above layer, also forms an anionic layer so as to effectively utilize the space. However, for the cobalt(II) ions the coordination to the sulfonate groups is weak and the intercalation of MV2+ dication into the cavity of p-sulfonatothaicalix[4]arene can easily break the potential layer arrangement. Thus, 1 and 3 display different structural motifs. To compare the conformations of the p-sulfonatothiacalix[4]arenes in the tetranuclear units, we define a plane through four bridge sulfur atoms. Table S1 (Supporting Information) shows the different degrees of the angles (signed by A, B, C, D) between the sulfur atom plane and the four benzene rings of the p-sulfonatothiacalix[4]arenes. For the tetranuclear units, when the guests are water molecules (in 3 and 6) or the [Na(H2O)2]+ cation (in 2), the conelike conformations do not deviate much from ideal C4V symmetries, which may be due to the small size of guests. However, when the guests have aromatic rings, the p-sulfonatothiacalix[4]arenes obviously splay apart of their opposite benzene rings to accommodate the planar guests and decrease their symmetries, which can be seen in compounds 1, 4, and 5. On the whole, even when the conelike conformation is fixed, the p-sulfonatothiacalix[4]arene can also splay apart of its opposite aromatic rings and adjust its conelike

(1) (a) Atwood, J. L.; Barbour, L. J.; Hardie, M. J.; Raston, C. L. Coord. Chem. ReV. 2001, 222, 3. (b) Hardie, M. J.; Raston, C. L. Dalton Trans. 2000, 2483. (c) Guo, D. Sh.; Wang, L. H.; Liu, Y. J. Org. Chem. 2007, 72, 7775. (d) Liu, Y.; Li, Q.; Guo, D. Sh.; Chen, K. CrystEngComm 2008, 10, 675. (e) Zheng, G. L.; Zhang, H. J.; Song, S. Y.; Li, Y. Y.; Guo, H. D. Eur. J. Inorg. Chem. 2008, 1756. (2) (a) Atwood, J. L.; Barbour, L. J.; Dawson, E. S.; Junk, P. C.; Kienzle, J. Supramol. Chem. 1996, 7, 271. (b) Ness, T.; Nichols, P. J; Raston, C. L. Eur. J. Inorg. Chem. 2001, 1993. (c) Atwood, J. L.; Ness, T.; Nichols, P. J.; Raston, C. L. Cryst. Growth Des. 2002, 2, 171. (d) Nichols, P. J; Raston, C. L. J. Chem. Soc., Dalton Trans. 2003, 2923. (e) Dalgarno, S. J.; Hardie, M. J.; Atwood, J. L.; Raston, C. L. Inorg. Chem. 2004, 43, 6351. (3) (a) Liao, W. P.; Bi, Y. F.; Gao, S.; Li, D. Q.; Zhang, H. J.; Dronskowski, R. Eur. J. Inorg. Chem. 2008, 2959. (b) Atwood, J. L.; Barbour, L. J.; Dalgarno, S. J.; Hardie, M. J.; Raston, C. L.; Webb, H. R. J. Am. Chem. Soc. 2005, 126, 13170. (c) Airey, S.; Drljaca, A.; Hardie, S. J.; Raston, C. L. J. Chem. Soc., Chem. Commun. 1999, 1137. (d) Dalgarno, S. J.; Raston, C. L. J. Chem. Soc., Dalton. Trans. 2003, 287. (e) Makaha, M.; Raston, C. L.; Sobolev, A. N.; White, A. H. J. Chem. Soc., Chem. Commun. 2005, 1962. (f) Liu, Y.; Guo, D. Sh.; Zhang, H. Y.; Ding, F.; Chen, K.; Song, H. B. Chem. Eur. J. 2007, 13, 466. (g) Liu, Y.; Li, Q.; Guo, D. Sh.; Chen, K. Cryst. Growth Des. 2007, 7, 1672. (h) Guo, D. Sh.; Su, X.; Liu, Y. Cryst. Growth. Des. 2008, 8, 3514. (4) (a) Drljaca, A.; Hardie, M. J.; Johnson, J. A.; Raston, C. L.; Webb, H. R. J. Chem. Soc., Chem. Commun. 1999, 1135. (b) Dalgarno, S. J.; Hardie, M. J.; Warren, J. E.; Raston, C. L. J. Chem. Soc., Dalton. Trans. 2004, 2413. (5) (a) Orr, G. W.; Barbour, L. J.; Atwood, L. J. Science 1999, 285, 1049. (b) Atwood, J. L.; Barbour, L. J.; Dalgarno, S. J.; Hardie, M. J.; Raston, C. L.; Webb, H. R. J. Am. Chem. Soc. 2004, 126, 13170. (6) (a) Dalgarno, S. J.; Hardie, C. L.; Raston, C. L. Cryst. Growth Des. 2004, 2, 227. (b) Dalgarno, S. J.; Raston, C. L. Chem. Commun. 2002, 2216. (c) Dalgarno, S. J.; Hardie, C. L.; Atwood, J. L.; Warren, J. E.; Raston, C. L. New J. Chem. 2005, 649. (7) (a) Smith, C. B.; Barbour, L. J.; Makha, M.; Raston, C. L.; Sobolev, A. N. Chem. Commun. 2006, 950. (b) Ness, T.; Nichols, P. J.; Raston, C. L. Eur. J. Inorg. Chem. 2001, 1999. (c) Drljaca, A.; Hardie, M. J.; Raston, C. L. J. Chem. Soc., Dalton Trans. 1999, 3639. (d) Drljaca, A.; Hardie, M. J.; Raston, C. L.; Spiccia, L. Chem. Eur. J. 1999, 5, 2298. (e) Nichols, P. J.; Raston, C. L.; Steed, J. W. Chem. Commun. 2001, 1062. (8) (a) Hardie, M. J.; Johnson, J. A.; Raston, C. L.; Webb, H. R. J. Chem. Soc., Chem. Commun. 2000, 849. (b) Atwood, J. L.; Dalgarno, S. J.;

p-Sulfonatothiacalix[4]arene Tetranuclear Clusters

(9)

(10)

(11) (12)

(13)

Hardie, M. J.; Raston, C. L. New J. Chem. 2004, 326. (c) Atwood, J. L.; Barbour, L. J.; Dalgarno, S. J.; Raston, C. L.; Webb, H. R. J. Chem. Soc., Dalton Trans. 2002, 4351. (a) Costanzo, L. D.; Geremia, S.; Randaccio, L.; Purrello, R.; Lauceri, R.; Sciotto, D.; Gulino, F.; Pavone, V. Angew. Chem., Int. Ed. 2001, 40, 4245. (b) Atwood, J. L.; Clark, D. L.; Juneja, R. K.; Orr, G. W.; Robinson, K. D.; Vincent, R. L. J. Am. Chem. Soc. 1992, 114, 7558. (c) Webb, H. R.; Hardie, M. J.; Raston, C. L. Chem. Eur. J. 2001, 7, 3616. (d) Asfari, Z.; Harrowfield, J.; Thuery, P.; Vicens, J. Supramol. Chem 2003, 15, 69. (e) Drljaca, A.; hardie, M. J.; Nesss, T. J.; Raston, C. L. Eur. J. Inorg. Chem. 2000, 2221. (f) Atwood, J. L.; Orr, G. W.; Robinson, K. D. Supramol. Chem. 1994, 3, 89. (a) Lhotak, P. Eur. J. Org. Chem. 2004, 1675. (b) Iki, N.; Suzuki, T.; Koyama, K.; Kabuto, C.; Miyano, S. Org. Lett. 2002, 4, 509. (c) Yuan, D. Q.; Wu, M. Y.; Wu, B. L.; Xu, Y. Q.; Jiang, F. L.; Hong, M. C. Cryst. Growth. Des. 2006, 6, 514. (d) Liu, Y.; Wang, H; Zhang, H. Y.; Wang, L. H. Cryst. Growth. Des. 2005, 5, 231. (e) Liu, Y.; Guo, D. Sh.; Zhang, H. Z.; Kang, S.; Song, H. B. Cryst. Growth. Des. 2006, 6, 1399. (a) Iki, N.; Miyano, S. J. J. Inclusion Phenom. Macrocyclic Chem. 2001, 41, 99. (b) Liu, Y.; Guo, D. Sh.; Zhang, H. Y.; Kang, Sh.; Song, H. B.; Chen, K. Cryst. Growth. Des 2006, 6, 1399. (a) Yuan, D. Q.; Xu, Y. Q.; Hong, M. C.; Bi, W. H.; Zhou, Y. F.; Li, X. Eur. J. Inorg. Chem. 2005, 1182. (b) Guo, Q. L.; Zhu, W. X.; Gao, S.; Ma, S. L.; Dong, S. J.; Xu, M. Q. Inorg. Chem. Commun. 2004, 7, 467. (c) Yuan, D. Q.; Wu, M. Y.; Jiang, F. L.; Hong, M. C. J. Mol. Struct. 2008, 877, 132. Wu, M. Y.; Yuan, D. Q.; Han, L.; Wu, B. L.; Xu, Y. Q.; Hong, M. C. Eur. J. Inorg. Chem. 2006, 526.

Crystal Growth & Design, Vol. 9, No. 3, 2009 1589 (14) Iki, N.; Fujimoto, T.; Miyano, S. Chem. Lett. 1998, 625. (15) CrystalClear Version 1.3.6, RIGAKU/MSC(2004), Rigaku/MSC, 9009 New Trails Drive, The Woodlands, TX77381-5209, USA. (16) (a) Sheldrick, G. M. SHELXS 97, Program for Crystal Structure Solution; University of Go¨ttingen: Germany, 1997. (b) Sheldrick, G. M. SHELXL 97, Program for Crystal Structure Refiniement; University of Go¨ttingen: Germany, 1997. (17) Atwood, J. L.; Dalgarno, S. J.; Hardie, M. J.; Raston, C. L Chem. Commun. 2005, 337. (18) (a) Telfer, S. G.; Sato, T.; Kuroda, R. Angew. Chem., Int. Ed. Engl. 2004, 43, 581. (b) Bailey, P. J.; Gould, R. O.; Harmer, C. N.; Pace, S.; Steiner, A.; Wright, D. S. Chem. Commun. 1997, 1161. (c) Telfer, S. G.; Kuroda, R. Chem. Eur. J. 2005, 11, 57. (d) Sharma, C. V. K.; Rogers, R. D. Chem. Commun. 1999, 83. (e) Beswick, M. A.; Mosquera, M. E. G.; Palmer, J. S.; Raithby, P. R.; Wright, D. S. New J. Chem. 1999, 23, 1033. (19) (a) Hanna, T. A.; Liu, L.; Angeles-Boza, A. W.; Kou, X.; Gutsche, C. D.; Ejsmont, K.; Waston, W. H.; Zakharov, l. N.; Incarvito, C. D.; Rheingold, A. L. J. Am. Chem. Soc. 2003, 125, 6228. (b) Thue´ry, P.; Asfari, Z.; Vicens, J.; Lamare, V.; Dozol, J. F. Polyhedron 2002, 21, 2497. (20) Cecillon, S.; lazar, A.; Danylyuk, O.; Suwinska, K.; Rather, B.; Zaworotko, M. J.; Coleman, A. W. Chem. Commun. 2005, 2442. (21) Makha, M.; Raston, C. L. Chem. Commun. 2001, 2470. (22) Ananchenko, G. S.; Pojarova, M.; Udachin, K. A.; Leek, D. M.; Coleman, A. W.; Ripmeester, J. A. Chem. Commun. 2006, 386.

CG801226P