Supramolecular Networks in Crystals of Metal (II) Complexes with

Water-Soluble Salicylaldehyde-2-sulfobenzoylhydrazone Anion. Ligand. La-Mei Wu, Han-Bing Teng, Xi-Chun Feng, Xian-Bing Ke, Qi-Feng Zhu, Jiang-Tao Su,...
0 downloads 0 Views 435KB Size
Supramolecular Networks in Crystals of Metal(II) Complexes with Water-Soluble Salicylaldehyde-2-sulfobenzoylhydrazone Anion Ligand La-Mei Wu, Han-Bing Teng, Xi-Chun Feng, Xian-Bing Ke, Qi-Feng Zhu, Jiang-Tao Su, Wen-Jin Xu, and Xian-Ming Hu*

CRYSTAL GROWTH & DESIGN 2007 VOL. 7, NO. 7 1337-1342

State Key Laboratory of Virology, College of Pharmacy, Wuhan UniVersity, Wuhan, 430072, China ReceiVed February 27, 2007; ReVised Manuscript ReceiVed April 12, 2007

ABSTRACT: A new water-soluble salicylaldehyde-2-sulfobenzoylhydrazone Schiff base anion ligand (HL-) has been designed and synthesized. The reaction of HL- with Zn(OAc)2‚2H2O, CoCl2‚6H2O, MnCl2‚4H2O, NiSO4‚7H2O, and Cu(OAc)2‚H2O or CuBr2 led to the formation of five novel binuclear complexes, namely, [Zn2(L2-)2(H2O)4](H2O)4 (1), [Co2(L2-)2(H2O)4](H2O)4 (2), [Mn2(L2-)2(H2O)4](H2O)4 (3), [Ni2(L2-)2(H2O)4](H2O)2 (4), and [Cu2(L2-)2(H2O)2](H2O)12 (5), which were characterized by IR spectroscopy, elemental analyses, and X-ray crystallography. Structural investigations of complexes 1-5 revealed that the introduction of a hydrophilic group -SO3- to the benzoyl ring of salicylaldehyde-benzoylhydrazone facilitates intermolecular infinite hydrogenbonding interactions, which lead to the formation of supramolecular networks and a dodecameric water cluster in the crystalline solid 5. Moreover, the electronegative -SO3- group in the ligand acts as a counteranion to balance the charge of frameworks, causing the ligand to coordinate metal ions through the oxygen of a protonated amide functionality rather than that of the deprotonated form and keeping additional anions out of these crystallization frameworks. Introduction Hydrogen bonds are a particularly important force for organizing crystal packing1 because they are strong and directional.2 Many different self-complementary hydrogenbonding groups such as -COOH,3 -CONH2,4 and -B(OH)25 can be used to control association in the solid state and to produce networks with predictable structural features and properties. Hydrogen-bonding interactions among these groups and water molecules also contribute to the formation of water clusters.6 Recent work has revealed the usefulness of the -SO3group in crystal engineering.7 Tetrahedral -SO3- groups can form strong hydrogen-bonding interactions with various hydrogenbonding donor molecules in different directions to enhance crystal stability and improve water solubility. Furthermore, the electronegative sulfonate group -SO3- not only can adopt many bridging coordination modes to construct new coordination frameworks but can also act as counteranion7b that balances the charge on frameworks when neutral ligands are employed. A report on water clusters8 based on the hydrogen-bonding interactions between -SO3- group and water molecules has appeared. Salicylaldehyde-acylhydrazone Schiff base and the metal complexes thereof show a wide spectrum of biological activities.9 Some of metal complexes can selectively cleave DNA.10 Salicylaldehyde-benzoylhydrazone and its metal complexes appeared to be unusually potent inhibitors of DNA synthesis and cell growth in a variety of human and rodent cell lines,9a-9e and cytotoxicity against a human adenocarcinoma cell line. A quantitative structure-activity relationship analysis revealed ligand cytotoxicity is strongly correlated with electronic and transport factors and can be modeled by treating each “half” of the molecule as an isolated unit. Activity increases when substituents in the benzoyl ring were electron withdrawing, whereas, for the salicylaldehyde ring, electron donation was required.9a * To whom correspondence should be addressed. E-mail: wlm52875@ 163.com. Fax: +86-2768754629. Tel: +86-2768753532.

To improve the force for organizing crystal packing and the water solubility as well as the bioactivity of the salicylaldehydebenzoylhydrazone and its metal complexes, we introduced a -SO3- group to the benzoyl ring of salicylaldehyde-benzoylhydrazone with the intention of synthesis of a water-soluble ligand. On the basis of the capacity of the -SO3- group to form strong hydrogen-bonding interactions with different hydrogenbonding donors to enhance crystal stability, it is expected the water-soluble ligands could coordinate with different metal ions in water and form strong hydrogen-bonding interactions with water molecules, so as to lead to the formation of crystals with sufficient size and quality for single-crystal X-ray diffraction. It is also expected that the electron-withdrawing group -SO3in the ligand could increase the cytotoxicity and the bioavailability of the ligand and its metal complexes. A new watersoluble salicylaldehyde-2-sulfobenzoylhydrazone ligand triethyl amine salt (HL‚HNEt3) has been synthesized and allowed to react with Zn(OAc)2‚2H2O, CoCl2‚6H2O, MnCl2‚4H2O, NiSO4‚ 7H2O, and Cu(OAc)2‚H2O or CuBr2 in water to provide five new complexes [Zn2(L2-)2(H2O)4](H2O)4 (1), [Co2(L2-)2(H2O)4] (H2O)4 (2),[Mn2(L2-)2(H2O)4](H2O)4 (3),[Ni2(L2-)2(H2O)4](H2O)2 (4), and [Cu2(L2-)2(H2O)2](H2O)12 (5) (see Scheme 1). Experimental Section Materials and Measurement. The starting materials were purchased from commercial sources and used without further purification. 1H NMR spectra were obtained at a Varian Mercury 300 MHz spectrometer. Chemical shifts are reported in δ relative to TMS. IR spectra were measured as KBr pellets on a Perkin-Elmer (Spectrum One) spectrometer in the range of 400-4000 cm-1. Elemental analyses (C, H, N) were carried out on a VarioEL III (German) instrument. Preparation of HL‚HNEt3. 2-Sulfobenzoic cyclic anhydride (SA) was prepared as reported in the literature.12 In methylene chloride (30 mL) 9.23 g, 0.05 mol of the anhydride was dissolved followed by dropwise addition of 50 mL of methanol at room temperature. The mixture was then refluxed for 30 min. After removal of the solvent under vacuum, the residue was dissolved in the hydrazine (30 mL, 85 wt % solution in water), and the mixture was refluxed for 7 h. After removal of the solvent and hydrazine under vacuum, the product was crystallized from ethanol. The crystalline material was dissolved in 100

10.1021/cg070196f CCC: $37.00 © 2007 American Chemical Society Published on Web 06/02/2007

1338 Crystal Growth & Design, Vol. 7, No. 7, 2007 Scheme 1.

Wu et al.

Synthesis of the New Ligand Triethyl Amine Salt HL‚HNEt3 and Its M(II) Coordination Polymers

mL of hot ethanol followed by addition of 20 g (0.16 mol) of salicylaldehyde and 1.01 g of triethyl amine, and the mixture was refluxed for 30 min. After filtration of the byproduct and removal of the solvent under vacuum, the residue was purified by flash column chromatography to yield HL‚HNEt3 in 60% yield (total). 1HNMR (DMSO, ppm): 12.69 (s, 1 H, -NH), 11.33 (s, 1 H, -OH), 8.35 (s, 1 H, -NCH), 7.57-7.89 (q, 4 H, -C6H4), 6.93-7.52 (q, 4 H, -C6H4), 3.07 (q, 6 H, -NCH2), 1.16 (t, 9 H, -CH3). IR (KBr, cm-1): 3446 (br), 3209 (m), 3149 (m), 3062 (m), 2988 (m), 1688 (vs), 1620 (m), 1611 (m), 1342 (s), 1216 (s), 1186 (s), 1019 (s), 740 (s), 616 (s). Anal. Calcd for C20H27N3O5S: C, 56.99; H, 6.46; N, 9.97. Found: C, 56.78; H, 6.62; N, 10.12. Preparation of 1. A water solution (5 mL) of HL‚HNEt3 (25.3 mg, 0.06 mmol) was slowly diffused into a water solution (5 mL) of Zn(OAc)2‚2H2O (21.9 mg, 0.1 mmol). Colorless crystals of 1 were collected after approximately 4 days. Yield: 75%. IR (cm-1, KBr pellet): 3492 (s, br), 3383 (m), 3227 (m), 3022 (m), 1654 (m), 1624 (m), 1600 (vs), 1581 (s), 1385 (m), 1202 (s), 1162 (m), 1022 (s), 768 (m), 617 (m). 1H NMR (DMSO, ppm): 13.83 (s, 1 H, -NH) 8.31 (s, 1 H, -NCH), 7.61-7.92 (q, 4 H, -C6H4), 6.46-7.20 (q, 4 H, -C6H4). Anal. Calcd for C28H36Zn2N4O18S2: C, 36.89; H, 3.39; N, 6.15. Found: C, 36.78; H, 3.18; N, 6.23. Preparation of 2. A water solution (5 mL) of HL‚HNEt3 (25.3 mg, 0.06 mmol) was slowly diffused into a water solution (5 mL) of CoCl2‚ 6H2O (23.8 mg, 0.1 mmol). Red crystals of 2 were collected after approximately two weeks. Yield: 71%. IR (cm-1, KBr pellet): 3447 (vs, br), 3231 (m), 1650 (m), 1621 (m), 1600 (s), 1585 (m), 1385 (m), 1198 (s),1160 (w), 1021 (m), 769 (w), 616 (w). Anal. Calcd for C28H36Co2N4O18S2: C, 37.42; H, 4.04; N, 6.23. Found: C, 37.25; H, 4.21; N, 6.12. Preparation of 3. A water solution (5 mL) of HL‚HNEt3 (25.3 mg, 0.06 mmol) was slowly diffused into a water solution (5 mL) of MnCl2‚ 4H2O (19.8 mg, 0.1 mmol), and then about 5.6 mg of KOH in 2 mL of water was added. Yellow crystals of 3 were collected after approximately 3 days. Yield: 51%. IR (cm-1, KBr pellet): 3489 (vs, br), 3383 (m), 1654 (m), 1618 (m), 1599 (s), 1585 (m), 1386 (m), 1199 (s),1161 (w), 1021 (m), 768 (w), 619 (w). Anal. Calcd for C28H36Mn2N4O18S2: C, 37.76; H, 4.07; N, 6.29. Found: C, 37.64; H, 3.93; N, 6.16. Preparation of 4. A water solution (5 mL) of HL‚HNEt3 (25.3 mg, 0.06 mmol) was slowly diffused into a water solution (5 mL) of NiSO4‚ 7H2O (28.8 mg, 0.1 mmol), and then about 5.6 mg of KOH in 2 mL of water was added. Green crystals of 4 were collected after ap-

proximately 2 days. Yield: 42%. IR (cm-1, KBr pellet): 3587 (m), 3454 (vs, br), 1634 (m), 1601 (s), 1585 (m), 1542 (m), 1444 (m), 1397 (m), 1243 (s), 1158 (w), 1018 (m), 772 (w), 611 (w). Anal. Calcd for C28H32Ni2N4O16S2: C, 39.01; H, 3.74; N, 6.50. Found: C, 39.26; H, 3.84; N, 6.43. Preparation of 5. A water solution (5 mL) of HL‚HNEt3 (25.3 mg, 0.06 mmol) was slowly diffused into a water solution (5 mL) of Cu(OAc)2‚H2O (19.9 mg, 0.1 mmol). Green crystals of 5 were collected after approximately four weeks. Yield: 65%. IR (cm-1, KBr pellet): 3440 (vs, br), 1654 (m), 1619 (s), 1599 (m), 1384 (w), 1180 (s), 1142 (m), 1023 (m), 753 (m), 617 (m). We substituted CuBr2 for Cu(Ac)2‚ H2O, and same crystal could be obtained after approximately 3 days. Single-Crystal Structure Determination. Single-crystal X-ray diffraction measurements for complexes 1-5 were carried out on a Bruker Smart APEX CCD-based diffractometer equipped with a graphite crystal monochromator for data collection at 292(2) K. The determinations of unit cell parameters and data collections were performed with Mo-KR radiation (λ ) 0.71073 Å), and unit cell dimensions were obtained with least-squares refinements. The program Bruke SAINT7 was used for reduction date. All structures were solved by direct methods using SHELXS-97 (Sheldrick, 1990) and refined with SHELXL-97 (Sheldrick, 1997);13 non-hydrogen atoms were located in successive difference Fourier syntheses. The final refinement was performed by full matrix least-squares methods with anisotropic thermal parameters for non-hydrogen atoms on F2. The hydrogen atoms were treated by a mixture of independent and constrained refinement. Crystallographic data and experimental details for structural analyses are summarized in Table 1.

Results and Discussion Structural Analysis of 1-3. Compound 1 is very stable at room temperature outside of the mother liquor. As shown in Figure 1, the complex 1 is binuclear molecule and the structure is composed of Zn(II) ions and tridentate Schiff base anion ligands L2- in a 1:1 ratio. The coordination geometry around each metal center is distorted octahedral with two Schiff base ligands in the basal plane acting as tridentate ligands through the imine-N, amide-O and the bridging phenolate-O atoms, which make two five-membered and two six-membered chelate rings. The two water molecules O6w, O7w are in the axial sites.

Table 1. Crystallographic Data for 1-5

empirical formula fw cryst syst a (Å) b (Å) c (Å) R (deg) β (deg) γ (deg) V (Å3) space group Z value F (calcd) (g/cm3) µ (Mo KR) (mm-1) T (K) R1; wR2 (I > 2 σ(I)) R1; wR2 (all) goodness-of-fit

1

2

3

4

5

C28H36Zn2N4O18S2 911.52 triclinic 9.2466(16) 10.2483(18) 10.6261(19) 108.031(3). 94.958(3) 106.695(2) 900.0(3) P1h 1 1.682 1.532 292(2) 0.0226; 0.0611

C28H36Co2N4O18S2 898.59 triclinic 9.2399(7) 10.2621(7) 10.5794(8) 108.0320(10) 95.2900(10) 105.9930(10) 899.51(11) P1h 1 1.659 1.122 292(2) 0.0240; 0.0636

C28H36Mn2N4O18S2 890.61 triclinic 9.1452(11) 10.3790(12) 10.7212(12) 107.715(2) 95.322(2) 103.939(2) 925.54(19) P1h 1 1.598 0.876 292(2) 0.0448; 0.1002

C28H32Ni2N4O16S2 862.12 triclinic 7.4869(7) 9.9503(9) 12.2777(11) 70.2410(10) 76.7820(10) 79.4670(10) 832.42(13) P1h 1 1.720 1.339 292(2) 0.0239; 0.0590

C28H48Cu2N4O24S2 1015.90 triclinic 7.6787(9) 11.1756(13) 12.6173(15) 77.358(2) 76.905(2) 84.316(3) 1027.6(2) P1h 1 1.642 1.229 292(2) 0.0352; 0.1041

0.0243; 0.0620 1.045

0.0255; 0.0646 1.032

0.0730; 0.1371 0.959

0.0271; 0.0610 1.067

0.0386; 0.1112 1.091

Supramolecular Networks in Crystals of Metal(II) Complexes

Crystal Growth & Design, Vol. 7, No. 7, 2007 1339

Figure 1. ORTEP diagram of 1.

Figure 3. The same packing pattern and intermolecular hydrogenbonded networks of 1-3.

Figure 2. Self-assembly of the complex 1 via hydrogen bonds into a 3-D network (O, red; C, gray; N, blue; Zn, green; S, yellow). Hydrogen atoms are omitted for clarity.

It is noteworthy that the oxygen atom of -SO3- group in L2is not coordinated to Zn(II) ions. The -SO3- group has strong intramolecular hydrogen-bonding interactions with coordinated O6w [O(3)‚‚‚O(6) distance 2.731(2)Å, O(3)‚‚‚H(6A)-O(6) bond angle 161(2)°] (black broken line in Figure 2). Additionally, as expected, each molecule of 1 interacts with two neighbors to form hydrogen-bonded one-dimensional (1-D) chains (sky-blue broken line in Figure 2) by intermolecular hydrogen-bonding interactions between coordinated water and the -SO3- group. The O‚‚‚O distance within the chains is 2.854(2) and 2.8368(19) Å, respectively (O(4) #3‚‚‚H(6B)-O(6) bond angle 168(2)°, O(4) #4‚‚‚H(7B)-O(7) bond angle 170(2)°). Furthermore, the chains of dinuclear complex 1 do not associate directly by forming lateral hydrogen bonds. Instead, they are linked to the two-dimensional (2-D) hydrogen-bonded structure by intervening molecules of lattice water O8w (red broken line in Figure 2), which acts as a single hydrogen-bonding acceptor from O7w(O‚‚‚O distance 2.764(2)Å, O(8) #5‚‚‚H(7A)-O(7) bond angle 173(2)°) and a double hydrogen-bonding donor to O2 and O5 atoms of nearby chains (O‚‚‚O distance 2.928 (2) and 2.878 (2)Å, O(2) #1‚‚‚H(8A)-O(8) bond angle 134.60(2)° and O(5) #1‚‚‚H(8B)-O(8) bond angle 164.77(2)°). O8w is presumably included because it is a better acceptor and donor of lateral hydrogen bonds than the chains themselves for it can form three hydrogen bonds in different directions that benefits the interchain self-assembly. Unlike -COOH,3 -CONH2,4 and -B(OH)2 groups5 that can only form 1-D hydrogen-bonded chains or 2-D hydrogen-bonded sheets, -SO3- is tetrahedral, so 2-D hydrogen-bonded motifs of 1 can associate further to three-dimensional (3-D) networks (purple broken line in Figure 2) by intervening molecules of

Figure 4. Self-assembly of the complex 4 via hydrogen bonds into 2-D network. (O, red; C, gray; N, blue; Ni, green; S, yellow). The oxygen atoms of lattice water O8w are ball style and hydrogen atoms are omitted for clarity.

lattice water O9w. These lateral interactions reinforce structures by adding hydrogen bonds and extending them in new directions. The single-crystal X-ray structure revealed that complexes 2 and 3 have similar molecular structures and the same packing pattern in addition to intermolecular hydrogen-bonded networks compared to complex 1 (Figure 3). In the three complexes, the organic components define a single packing arrangement that persists over a range of structures and the metal ions serve as a component that can be exchanged, which is helpful for design of crystalline modular materials.15 This type of modularity is crucial in the design of crystalline materials because it provides a handle with which to alter physical properties without changing the structure of the material. Structural Analysis of 4. As shown in Figure 4, the molecular structure of complex 4 is similar to that of complex 1, although its packing pattern and intermolecular hydrogenbonded networks are different. Compared with the 3-D hydrogenbond networks structure in 1, compound 4 is linked to a 2-D network structure by intermolecular hydrogen-bonding interactions. Each molecule of dinuclear complex 4 interacts with two neighbors to form hydrogen-bonded 1-D chains (mauve broken

1340 Crystal Growth & Design, Vol. 7, No. 7, 2007

Wu et al.

Figure 6. Self-assembly of the complex 5 via hydrogen bonds into a 3-D network. Hydrogen atoms are omitted for clarity.

Figure 5. ORTEP diagram of 5.

line in Figure 4) by intermolecular hydrogen-bonding interactions between the coordinated water and the -SO3- group. The O‚‚‚O distance within the chains is 2.738(2) Å (O(6)-H(11) ‚‚‚O(4)#3 bond angle 169(3)°). The chains of dinuclear complex 4 associate directly by forming lateral hydrogen bonds. They are linked to a 2-D sheet hydrogen-bonded structure by the intermolecular hydrogen-bonding interactions between the nitrogen of an amide and the -SO3- group (red broken line in Figure 4), and the N‚‚‚O distance between the chains is 2.809(2) Å (N(2)-H(6)‚‚‚O(3)#2 bond angle 171(2)°). The lattice water O8w acts as a single hydrogen-bonding acceptor from O7w and a single hydrogen-bonding donor to O3 atom of the same molecule (sky-blue broken line in Figure 4), and it is different from the lattice water O8w and O9w in compound 1. Structural Analysis of 5. Crystals of 5 is unstable outside of the mother liquor and loses guest solvent molecules slowly. As shown in Figure 5, each Cu(II) center in compound 5 is a five-coordinated distorted square-pyramidal coordination sphere, which differs from the six-coordinated octahedral geometry seen in complexes 1-4. One dinegative tridentate ligand binds the metal ion via the imine-N, the amide-O, and the bridging phenolate-O atoms, and the fourth site is satisfied by the oxygen atom of a water molecule to form a square-plane, and the bridging phenolate-O atom of the other tridentate ligand occupies the axial coordination site. The coordination mode is different from that of the other complexes.11e,11f,11h The extensive hydrogen-bonding interactions that link 5 into multidimensional network structure are all indirect. Two adjacent -SO3- groups are linked to the hydrogen-bonded structure by intervening molecules of lattice water O8w and O7w. Four water molecules (two O8w and two O7w) link two adjacent -SO3groups into a cagelike hydrogen-bonded net approximately along the c-axis (sky-blue broken line in Figure 6), the lattice water O8w and O7w are the hydrogen-bonding bridges ([O‚‚‚O distance 2.739(3)Å, O(3) #4‚‚‚H(8A)-O(8) bond angle 164(4)°], [O‚‚‚O distance 2.850(3)Å, O(2)#7‚‚‚H(8B)-O(8) bond angle 173(5)°], [O‚‚‚O distance 2.872(3) Å, O(2) #1‚‚‚H(7A)O(7) bond angle 173(4)°], [O‚‚‚O distance 2.970(3) Å, O(1)#5‚‚‚H(7B)-O(7) bond angle 179(4)°]). Furthermore, the cagelike hydrogen-bonded nets are linked approximately along the a-axis by the intermolecular hydrogen bonds (O‚‚‚N distance 2.796(3) Å, O(7) #7‚‚‚H(2)-N(2) bond angle 156.1°), O7w is the hydrogen-bonding bridge acting as three-connecting spacers, just like water molecules O9w in compound 1. In addition, approximately along the b-axis, the cagelike nets are linked

Figure 7. A dodecameric water cluster and the surrounding environment: black broken lines are the hydrogen bonds among dodecameric water clusters, and sky-blue broken lines are the hydrogen bonds linking dodecameric water cluster and six units of compound 5.

together through dodecameric water clusters (black broken line in Figure 6) intermolecular hydrogen-bonding interactions. Of particular interest here is the structure of a dodecameric water cluster in the crystalline solid 5. The cluster consists of a combination of cyclic tetramer and cyclic pentamer. To our knowledge, this dodecameric water cluster was not found in the 12 hydrogen-bonded water molecules reported previously.16 The dodecameric water cluster inside the six units of compound 5 is connected on two sides by two copper(II) ions [O6w‚‚‚Cu, 1.9254(17) Å] and hydrogen bonded to six oxygen atoms of sulfonate groups [O8w‚‚‚O2#7, 2.850(3) Å, O8w‚‚‚O(3)#4, 2.739(3) Å, O9w‚‚‚O(1)#5, 2.859(3) Å] (sky-blue broken line in Figure 7). In the dodecameric water cluster (mauve ball linked by black broken line in Figure 7), six water molecules are connected to another six centrosymmetrically related water molecules forming a spiro cyclic dodecameric water cluster. Atoms O11w and O12w are hydrogen bonded to O11w′ and O12w′ forming a cyclic planar tetramer unit, atoms O11w and O11w′ link a cyclic nonplanar pentamer unit along the opposite direction, respectively. A closer look at the water cluster (see Figure 8) reveals that coordinated water O6w acts as a double hydrogen-bonding donor, and other water molecules act as both hydrogen-bonding donors and hydrogen-bonding acceptors. O8w and O11w participate in four hydrogen bonds in a tetrahedral arrangement, and the O‚‚‚O distances range from 2.645(5) to 3.096(5) Å with an average distance of 2.806 Å, comparable to 2.77-2.84 Å in the ice II phase17 and 2.776 Å in the 2D supramolecular icelike layer containing (H2O)12 rings in the organic compound,16c and 2.703(6) to 3.100(6) Å with an average distance of 2.782 Å in an infinite chain containing an (H2O)12 cluster encapsulated in a supramolecular open framework.16e However, they are shorter than those observed in liquid water (2.854 Å).18 The O‚‚‚O

Supramolecular Networks in Crystals of Metal(II) Complexes

Crystal Growth & Design, Vol. 7, No. 7, 2007 1341

Conclusions

Figure 8. A polymeric water layer formed by the oxygen atoms of sulfonate groups (purple ball) and lattice water O7w along the a direction.

Figure 9. Stacking diagram of complex 5, showing 3-D hydrogenbonded network structure and water cluster in channels viewed in the crystallographic bc plane (sky-blue broken lines are the hydrogen bonds). Hydrogen atoms are omitted for clarity.

distances within the pentamer with an average distance of 2.701 Å are shorter than those observed in the cyclic water pentamer (2.706(3)-2.918(3) Å),19 and the O‚‚‚O distances within cyclic planar tetramer with an average distance of 3.047 Å are longer than those observed in the cyclic water tetramer 2.731(1)-2.80(1) Å20a and 2.647(13)-3.01(3) Å.20b The dodecameric water cluster is linked to a layer along the a-axis by the intermolecular hydrogen-bonding interactions between oxygen atoms of the -SO3- groups and lattice water O7w (Figure 8) in the 3-D network and fit very well in the cavity of the 3D net (Figure 9). The intermolecular long water cluster makes complex 5 unstable outside of the mother liquor, and it loses water molecules slowly. Thermogravimetric (TG) Analyses. TG analysis of 1 reveals that its first weight loss appears at ca. 70 °C due to the loss of lattice water and coordinating water molecules (calcd. 15.8%; found 15.3%). The second weight loss occurs between 370 °C and 453 °C and is characteristic of the decomposition of the organic moiety. Complex 4 shows the first weight loss at between 40 and 87 °C with the loss of lattice water and coordinating water molecules (calcd. 12.5%; found 12.7%), and the second weight loss occurs between 250 and 360 °C and is characteristic of the decomposition of the organic moiety. Complex 5 shows the first weight loss with the loss of lattice water and coordinating water molecules (calcd. 17.7%; found 16.5%) from room temperature to 205 °C, and the second weight loss occurs between 280 and 301 °C and is characteristic of the decomposition of the organic moiety (see Supporting Information).

(1) The -SO3- group can increase the water solubility of the ligand and the metal complexes. We are currently performing DNA binding and the cytotoxicity studies on the ligand and their metal complexes to reveal how the electronic effect and the solubility of substituents affect their bioactivity. (2) The water-soluble ligand can coordinate to different metal ions to form crystals with sufficient size and quality for singlecrystal X-ray diffraction easily, because the tetrahedral -SO3group can form characteristic motifs incorporating six hydrogen bonds that extend in different directions, and on crystallization in the presence of suitable guest ions lattice water can divert -SO3- from its normal tendency to form hydrogen bonds in different directions and thereby leads to the formation of highdimensional supramolecular networks. (3) One essential requirement for the metal complexes with biological activities is the formation of electrically neutral species at physiological pH so as to enable easy perfusion into tissues with minimal renal loss. The -SO3- group is negatively charged and acts as counteranion to balance these coordination frameworks charge; various anionic counterions (OAc-, Cl-, SO42-, and Br-) involved were excluded from these frameworks. These electrically neutral crystals with fixed components and similar packing arrangement are helpful for the design of electrically neutral metal complexes. (4) A dodecameric water cluster was found in the crystalline solid 5. This water cluster demonstrates that the sulfobenzoylhydrazone compound is a promoter to induce the aggregation of water molecules. Supporting Information Available: Table of selected bond lengths and bond angles for complexes 1-5, table of hydrogen bonds for complexes 1, 4, 5, thermogravimetric analyses plots of compounds 1, 4, and 5, and X-ray crystallographic information files (CIF) for compounds 1-5. This material is available free of charge via the Internet at http://pubs.acs.org.

References (1) (a) Nishio, M. CrystEngComm 2004, 6, 130. (b) Jeffrey, G. A. Cryst. ReV. 2003, 9, 135. (c) Hardie, M. J. In Supramolecular Assembly Via Hydrogen Bonds II; Mingos, D. M. P., Ed.; Springer-Verlag: Berlin, 2004; Vol. 111, pp 139-174. (d) Desiraju, G. R. Acc. Chem. Res. 2002, 35, 565. (e) Gavezzotti, A. J. Mol. Struct. 2002, 615, 5. (f) Prins, L. J.; Reinhoudt, D. N.; Timmerman, P. Angew. Chem., Int. Ed. 2001, 40, 2382. (2) (a) Steiner, T. Angew. Chem., Int. Ed. 2002, 41, 48. (b) Moulton, B.; Zaworotko, M. J. Chem. ReV. 2001, 101, 1629. (c) Prins, L. J.; Reinhoudt, D. N.; Timmerman, P. Angew. Chem., Int. Ed. 2001, 40, 2382. (3) (a) Aakeroy, C. B.; Hussain, I.; John, D. Cryst. Growth Des. 2006, 6, 474. (b) Sledz, M.; Janczak, J.; Kubiak, R. J. Mol. Struct. (THEOCHEM) 2001, 77, 595. (4) Cobbledick, R. E.; Small, R. W. H. Acta Crystallogr. 1972, B28, 2893. (5) (a) Aakeroy, C. B.; Desper, J.; Levin, B. CrystEngComm 2005, 7, 102. (b) Rodriguez-Cuamatzi, P.; Arillo-Flores, O. I.; BernalUruchurtu, M. I.; Hopfl, H. Cryst. Growth Des. 2005, 5, 167. (c) Rodriguez-Cuamatzi, P.; Vargas-Diaz, G.; Maris, T.; Wuest, J. D.; Hopfl, H. Acta Crystallogr. 2004, E60, o1316. (d) RodriguezCuamatzi, P.; Vargas-Diaz, G.; Hopfl, H. Angew. Chem., Int. Ed. 2004, 43, 3041. (e) Fournier, J.-H.; Maris, T.; Wuest, J. D.; Guo, W.; Galoppini, E. J. Am. Chem. Soc. 2003, 125, 1002. (6) (a) Zhao, B.; Cheng, P.; Chen, X.; Cheng, C.; Shi, W.; Liao, D.; Yan, S.; Jiang, Z. J. Am. Chem. Soc. 2004, 126, 3012. (b) Yang, J.; Ma, J.-F.; Liu, Y.-Y.; Ma, J.-C.; Jia, H.-Q.; Hu, N.-H. Eur. J. Inorg. Chem. 2006, 1208. (c) Li, Y.; Jiang, L.; Feng, X.-L.; Lu, T.-B. Cryst. Growth Des. 2006, 6, 1074. (d) Sreenivasulu, B.; Vittal, J. J. Angew. Chem., Int. Ed. 2004, 43, 5769. (e) Luan, X. J.; Chu, Y. C.; Wang, Y. Y.; Li, D. S.; Liu, P.; Shi, Q. Z. Cryst. Growth Des. 2006, 6, 812. (f) Ng, M. T.; Deivaraj, T. C.; Klooster, W. T.; McIntyre, G. J.;

1342 Crystal Growth & Design, Vol. 7, No. 7, 2007

(7)

(8) (9)

(10)

(11)

Vittal, J. J. Chem. Eur. J. 2004, 10, 5853. (g) Zhu, W.-H.; Wang, Z.-M.; Gao, S. Dalton Trans. 2006, 765. (h) Das, S.; Bharadwaj, P. K. Cryst. Growth Des. 2006, 6, 187. (i) Mukherjee, A.; Saha, M. K.; Nethaji, M.; Chakravarty, A. R. Chem. Commun. 2004, 716. (j) Carballo, R.; Covelo, B.; Lodeirob, C.; Vazquez-Lopeza, E. M. CrystEngComm 2005, 7, 294. (k) Rodriguez-Cuamatzi, P.; VargasDaz, G.; Hopfl, H. Angew. Chem., Int. Ed. 2004, 43, 3041. (l) Blanton, W. B.; Gordon-Wylie, S. W.; Clark, G. R.; Jordan, K. D.; Wood, J. T.; Geiser, U.; Collins, T. J. J. Am. Chem. Soc. 1999, 121, 3551. (m) Barbour, L. J.; Orr, G. W.; Atwood, J. L. Chem. Commun. 2000, 859. (a) Burke, N. J.; Burrows, A. D.; Mahon, M. F.; Warren, J. E. Cryst. Growth Des. 2006, 6, 546. (b) Li, F.-F.; Ma. J.-F.; Song, S.-Y.; Yang, J. Cryst. Growth Des. 2006, 6, 209. (c) May, L. J.; Shimizu, G. K. H. Chem. Mater. 2005, 17, 217. (d) Prakash, M. J.; Radhakrishnan, T. P. Cryst. Growth Des. 2005, 5, 1831. (e) Cote, A. P.; Ferguson, M. J.; Khan, K. A.; Enright, G. D.; Kulynych, A. D.; Dalrymple, S. A.; Shimizu, G. K. H. Inorg. Chem. 2002, 41, 287. (f) Makinen, S. K.; Melcer, N. J.; Parvez, M.; Shimizu, G. K. H. Chem. Eur. J. 2001, 7, 5176. (g) Yuan, D. Q.; Wu, M. Y.; Wu, B. L.; Xu, Y. Q.; Jiang, F. L.; Hong, M. C. Cryst. Growth Des. 2006, 6, 514. (h) Dalgarno, S. J.; Atwood, J. L.; Raston, C. L. Cryst. Growth Des. 2006, 6, 174. Liua, Q.-Y.; Xu, L. CrystEngComm 2005, 7, 87. (a) Ainscough, E. W.; Brodie, A. M.; Denny, W. A.; Finlay, G. J.; Gothe, S. A.; Ranford, J. D. J. Inorg. Biochem. 1999, 77, 125. (b) Pickart, L.; Goodwin, W. H.; Burgua, W.; Murphy, T. B.; Johnson, D. K. Biochem. Pharmacol. 1983, 32, 3868. (c) Aruffo, A. A.; Murphy, T. B.; Johnson, D. K.; Rose, N. J.; Schomaker, V. Acta Crystallogr. Sect. C. 1984, 40, 1164. (d) Johnson, D. K.; Murphy, T. B.; Rose, N. J.; Goodwin, W. H.; Pickart, L. Inorg. Chim. Acta. 1982, 67, 159. (e) Aruffo, A. A.; Murphy, T. B.; Johnson, D. K.; Rose, N. J.; Schomaker, V. Inorg. Chim. Acta. 1982, 67, L52. (f) Melnyk, P.; Leroux, V.; Sergheraerta, C.; Grellier, P. Bioorg. Med. Chem. Letters. 2006, 16, 31. (g) Mohan, M.; Kumar, A.; Kumar, M.; Jha, N. K. Inorg. Chim. Acta 1987, 136, 65. (h) Koh, L. L.; Kon, O. L.; Loh, K. W.; Long, Y. C.; Ranford, J. D.; Tan, A. L. C.; Tjan, Y. Y. J. Inorg. Biochem. 1998, 72, 155. (a) Sylvain, R.; Bernier, J. L.; Waring, M. J. J. Org. Chem. 1996, 61, 2326. (b) Santanu, B.; Subbrangsu, S. M. J. Chem. Soc. Chem. Commun. 1995, 2489. (c) Gravert, D. J.; Griffin, J. H. J. Org. Chem. 1993, 58, 820. (a) Rath, S. P.; Rajak, K. K.; Chakravorty, A. Inorg. Chem. 1999, 38, 4376. (b) Plass, W.; Pohlmann, A.; Yozgatli, H.-P. J. Inorg. Biochem. 2000, 80, 181. (c) Rajak, K. K.; Baruah, B.; Rath, S. P.; Chakravorty, A. Inorg. Chem. 2000, 39, 1598. (d) Das, S.; Pal, S. J.

Wu et al.

(12)

(13) (14) (15)

(16)

(17) (18) (19) (20)

Mol. Struct. 2005, 753, 68. (e) Sangeetha, N. R.; Baradi, K.; Gupta, R.; Pal, C. K.; Manivannan, V.; Pal, S. Polyhedron 1999, 18, 1425. (f) Ainscough, E. W.; Brodie, A. M.; Dobbs, A. J.; Ranford, J. D.; Waters, J. M. Inorg. Chim. Acta 1998, 267, 27. (g) Sangeetha, N. R.; Pal, S. Polyhedron 2000, 19, 1593. (h) Iskande, M. F.; Khalil, T. E.; Werner, R.; Haase, W.; Svoboda, I.; Fuess, H. Polyhedron 2000, 19, 949. (i) Iskander, M. F.; El-Sayed, L.; Salem, N. M. H.; Haase, W.; Linder, H. J.; Foro, S. Polyhedron 2004, 23, 23. (j) Yin, H. D.; Hong, M.; Li, G.; Wang, D. Q. J. Organomet. Chem. 2005, 690, 3714. (k) Jagst, A.; Sanchez, A.; Vazquez-Lopez, E. M.; Abram, U. Inorg. Chem. 2005, 44, 5738. (l) Rath, S. P.; Rajak, K. K.; Mondal, S.; Chakravorty, A. J. Chem. Soc. Dalton Trans. 1998, 2097. (m) Caruso, U.; Centore, R.; Panunzi, B.; Roviello, A.; Tuzi, A. Eur. J. Inorg. Chem. 2005, 2747. (n) Ainscough, E. W.; Brodie, A. M.; Ranford, J. D.; Waters, J. M. Inorg. Chim. Acta. 1995, 236, 83. (o) Das, S.; Pal, S. J. Organomet. Chem. 2006, 691, 2575. Jacks, T. E.; Belmont, D. T.; Briggs, C. A.; Horne, N. M.; Kanter, G. D.; Karrick, G. L.; Krikke, J. J.; McCabe, R. J.; Mustakis, J. G.; Nanninga, T. N.; Risedorph, G. S.; Seamans, R. E.; Skeean, R.; Winkle, D. D.; Zennie, T. M. Org. Proc. Res. DeV. 2004, 8, 201. Sheldrick, G. M. SHELXTL V5.1; Madison, WI, 1998. Banthia, S.; Samanta, A. Cryst. Growth Des. 2006, 6, 360. (a) Yigit, M. V.; Biyikli, K.; Moulton, B.; MacDonald, J. C. Cryst. Growth Des. 2006, 6, 63. (b) Luo, T.-J. M.; MacDonald, J. C.; Palmore, G. T. R. Chem. Mater. 2004, 16, 4916. (c) MacDonald, J. C.; Dorrestein, P. C.; Pilley, M. M.; Foote, M. M.; Lundburg, J. L. J. Am. Chem. Soc. 2000, 122, 11692. (a) Nishikiori, S.; Iwamoto, T. J. Chem. Soc. Chem. Commun. 1993, 1555. (b) Janiak, C.; Scharman, T. G. J. Am. Chem. Soc. 2002, 124, 14010. (c) Ma, B. Q.; Sun, H. L.; Gao, S. Angew. Chem., Int. Ed. 2004, 43, 1374. (d) Ghosh, S. K.; Bharadwaj, P. K. Angew. Chem., Int. Ed. 2004, 43, 3577. Jeffrey, G. A. An Introduction to Hydrogen Bonding; Oxford University Press: Oxford, U.K., 1997; pp 160-180. Speedy, R. J.; Madura, J. D.; Jorgensen, W. L. J. Phys. Chem. 1987, 91, 909. Ma, B.-Q.; Sun, H.-L.; Gao, S. Chem. Commun. 2004, 2220. (a) Ng, M. T.; Deivaraj, T. C.; Klooster, W. T.; McIntyre, G. J.; Vittal, J. J. Chem. Eur. J. 2004, 10, 5853. (b) Supriya, S.; Manikumari, S.; Raghavaiah, P.; Das, S. K. New J. Chem. 2003, 27, 218.

CG070196F