H2O-Involved Hydrogen Bonds in Pseudo-Double-Decker

Nov 14, 2008 - H2O-Involved Hydrogen Bonds in Pseudo-Double-Decker Supramolecular Structure of 1,8,15,22-Tetrasubstituted Phthalocyaninato Zinc ...
0 downloads 0 Views 704KB Size
CRYSTAL GROWTH & DESIGN

H2O-Involved Hydrogen Bonds in Pseudo-Double-Decker Supramolecular Structure of 1,8,15,22-Tetrasubstituted Phthalocyaninato Zinc Complex Renjie Li, Yuexing Zhang, Yang Zhou, Shuai Dong, Xianyao Zhang, Yongzhong Bian, and Jianzhuang Jiang*

2008 VOL. 8, NO. 12 4454–4459

Department of Chemistry, Shandong UniVersity, Jinan 250100, China ReceiVed April 3, 2008; ReVised Manuscript ReceiVed September 14, 2008

ABSTRACT: A novel dimeric supramolecular structure, {Zn[Pc(R-OC5H11)4] · H2O}2, formed from two molecules of 1,8,15,22tetrakis(3-pentyloxy)phthalocyaninato zinc complex was revealed by X-ray single crystal analysis. With the help of the oxygen atom from the alkoxy substituent attached at the nonperipheral position of the phthalocyanine ring, each H2O forms two hydrogen bonds, connecting two phthalocyanine molecules to form a pseudo-double-decker supramolecular structure in the crystals with a ring-to-ring separation of 3.728 Å. This, to the best of our knowledge, represents the first example of supramolecular structure formed from phthalocyanine complexes with transition metals via H2O-involved hydrogen bonding interaction. To enhance understanding of the existence of hydrogen bonds in the solid-state crystal structure of this compound, theoretical calculations on the stabilization energy in a system composed of two Zn[Pc(R-OC5H11)4] moieties as well as in the supramolecular structure {Zn[Pc(ROC5H11)4] · H2O}2 have been performed using the density functional theoretical method. Comparison of the calculated stabilization energy between these two systems together with the natural bond orbital analysis over the later supramolecular structure reveals the dominant H2O-involved hydrogen bonding interaction over the π-π interaction in {Zn[Pc(R-OC5H11)4] · H2O}2. Introduction Since their early synthesis at the beginning of last century, phthalocyanines have been important industrial dyes and pigments.1 In recent years, they have been employed as charge carriers in photocopiers and laser printers and materials for optical storage.2,3 Many potential applications are expected for these molecular materials in the area of oxidation catalysts,4 solar cell functional materials,5 gas sensors,6 nonlinear optical limiting devices,7 photodynamic therapy agents,8 antimycotic material,9 and corrosion inhibitors.10 It is well-known that phthalocyanines can form complexes with more than 70 elements, including virtually all the metals in the Periodic Table. Their optical, electrochemical, and other physicochemical properties have been found to depend not only on the metal center and substituents at the phthalocyanine ring but also on the molecular ordering in the solid-state form. As a consequence, intensive and extensive studies have been conducted on the single crystal structures of various kinds of phthalocyanine compounds, including metal-free phthalocyanines and phthalocyaninato complexes with main group metals, transition metals, rare earth metals, and actinide metals, in particular since the 1980s.11 However, H2O- and organic solventinvolved hydrogen bonding interaction, which is common and plays important role in large biological molecules and supramolecular materials,12 has rarely been reported for the phthalocyaninato metal complexes, most probably because of the intrinsic intense π-π interaction between planar phthalocyanine rings, except the phthalocyaninato complexes with main group IIA metal of Mg and Be.13 In contrast, the relatively weakened π-π interaction between porphyrin molecules associated with their smaller conjugated system as well as the substituents at the meso- and/or β-positions of porphyrin rings, which is not strong enough to exclude the formation of H2O- or organic solvent-involved hydrogen bonding between porphyrin molecules, has resulted in relatively numerous reports for the * E-mail: [email protected].

porphyrin supramolecular structures formed, depending on the hydrogen bonding interaction.14 We describe herein a novel dimeric supramolecular structure, {Zn[Pc(R-OC5H11)4] · H2O}2, formed from two molecules of 1,8,15,22-tetrakis(3-pentyloxy)phthalocyaninato zinc complex. X-ray single crystal analysis reveals that two water molecules exist between two Zn[Pc(R-OC5H11)4] molecules. With the help of the oxygen atom from the alkoxy substituent attached at the nonperipheral position of phthalocyanine ring, each H2O forms two hydrogen bonds, connecting two phthalocyanine molecules to form a pseudo-double-decker supramolecular structure in the crystals with a ring-to-ring separation of 3.728 Å. This, to the best of our knowledge, seems to represent the first example of supramolecular structure formed from phthalocyanine complexes with transition metals via H2O-involved hydrogen bonding interaction. To further enhance understanding of the formation of H2O-involved hydrogen bonds, theoretical calculations on the stabilization energy in a system composed of two Zn[Pc(ROC5H11)4] moieties as well as in the supramolecular structure {Zn[Pc(R-OC5H11)4] · H2O}2, and the natural bond orbital analysis over the later supramolecular structure, have been performed using the density functional theoretical method. Experimental Section General. n-Pentanol was distilled from sodium. Dichloromethane for voltammetric studies was freshly distilled from CaH2 under nitrogen. Column chromatography was carried out on silica gel columns (Merck, Kieselgel 60, 70-230 mesh). All other reagents and solvents were used as received. The compounds of 3-(3-pentyloxy)phthalonitrile and H2Pc(R-OC5H11)4 were prepared following reported procedures.15,16 1 H NMR spectra were recorded on a Bruker DPX 300 spectrometer (300 MHz) in CDCl3. Spectra were referenced internally using the residual solvent resonances (δ 7.26) relative to SiMe4. Electronic absorption spectra were recorded on a Hitachi U-4100 spectrophotometer. IR spectra were recorded as KBr pellets using a BIORAD FTS165 spectrometer with 2 cm-1 resolution. MALDI-TOF mass spectra were taken on a Bruker Daltonics BIFLEX III MALDI-TOF mass spectrometer with R-cyano-4-hydroxycinnamic acid as matrix. Elemen-

10.1021/cg800342b CCC: $40.75  2008 American Chemical Society Published on Web 11/14/2008

1,8,15,22-Tetrasubstituted Phthalocyaninato Zn Complex Table 1. Crystallographic Data for {Zn[Pc(r-OC5H11)4] · H2O}2 molecular formula M crystal system space group a/Å b/Å c/Å β/deg 3 U/Å Z Dc/mg-1m-3 µ/mm data collection range/deg reflections measured independent reflections parameters R1 [I > 2σ(I)] wR2 [I > 2σ(I)] goodness of fit

C104H116N16O10Zn2 1880.95 monoclinic P2(1)/c 16.1943(2) 20.7600(3) 14.4784(2) 94.1770(10) 4854.62(11) 4 1.287 0.561 1.60-25.00 43705 8548 (Rint ) 0.1046) 603 0.0500 0.1317 1.026

tal analyses were performed by the Institute of Chemistry, Chinese Academy of Sciences. Electrochemical measurements were carried out with a BAS CV50W voltammetric analyzer. The cell comprised inlets for a glassy carbon disk working electrode of 3.0 mm in diameter and a silverwire counter electrode. The reference electrode was Ag/Ag+, which was connected to the solution by a Luggin capillary whose tip was placed close to the working electrode. It was corrected for junction potentials by being referenced internally to the ferrocenium/ferrocene (Fc+/Fc) couple [E1/2 (Fc+/Fc) ) 0.5 V vs SCE]. Typically, a 0.1 mol dm-3 solution of [Bu4N][ClO4] in CH2Cl2 containing 0.5 mmol dm-3 of sample was purged with nitrogen for 10 min, and then the voltammograms were recorded at ambient temperature. The scan rate was 20 and 10 mV s-1 for CV and DPV, respectively. Computational Details. The primal input structure of {Zn[Pc(ROC5H11)4] · H2O}2 dimer was imported from the crystal structure. The {Zn[Pc(R-OC5H11)4] · H2O}2 dimer structure is first optimized using the DMol3 module of Cerial2 software17 with the GGA and PBD functions, and then the single point energy of every molecule is determined to calculate the stabilization energy between the two rings in the dimeric molecule. The stabilization energy of the {Zn[Pc(R-OC5H11)4]}2 dimer without hydrogen-bonding involved-water molecules is also calculated using the same method. Using the hybrid density functional B3LYP (Becke-Lee-Young-Parr composite of exchange-correlation functional) method18 and 6-31G(d) basis set,19 natural bond orbital analysis was carried out using a full natural bond orbital analysis (NBO) population method involved in the Gaussian 03 suite program20 on the basis of the primal molecular structure imported from the crystal in the IBM P690 system in Shandong Province High Performance Computing Centre. Zn[Pc(r-OC5H11)4]. A mixture of Zn(OAc)2 · 2H2O (44 mg, ca. 0.20 mmol) and 1,8,15,22-tetrakis(3-pentyloxy)phthalocyanine (86 mg, 0.10 mmol) in n-pentanol (4 mL) was heated to reflux under nitrogen for ca. 4 h. The solvent was then removed in vacuo, and the residue was subjected to chromatography on a silica gel column using CHCl3 as eluent. The crude product was purified by repeated chromatogrphy followed by recrystallization from CHCl3/MeOH, giving a dark green powder, 78 mg, 85%. 1H NMR (CDCl3) δ 8.73-8.74 (d, 4 H, Ha), 7.94-7.99 (t, 4 H, Hb), 7.49-7.51 (d, 4 H, Hc), 4.78 (br, 4 H, OCH), 2.18-2.27 (m, 8 H, CH2), 2.06-2.10 (m, 8 H, CH2), 1.25-1.29 (t, 24H, CH3); UV-vis (CHCl3) [λmax/nm (log ε)] 318 (4.65), 354 (4.45), 634 (4.58), 705 (5.36); MS (MALDI-TOF) an isotopic cluster peaking at m/z 922.0 (calcd for M+ 922.5). Anal. Calcd for C52H56N8O4Zn · H2O: C, 66.41; H, 6.22; N, 11.91. Found: C, 66.38; H, 6.18; N, 11.90%. X-ray Crystallography. Crystal data and details of data collection and structure refinement are given in Table 1. Data were collected on a Bruker SMART CCD diffractometer with an Mo KR sealed tube (λ ) 0.71073 Å) at 298 K, using a ω scan mode with an increment of 0.5°. Preliminary unit cell parameters were obtained from 36 frames. Final unit cell parameters were derived by global refinements of reflections obtained from integration of all the frame data. The collected frames were integrated using the preliminary cell-orientation matrix. The SMART software was used for collecting frames of data, indexing reflections, and determination of lattice constants; SAINT-PLUS for

Crystal Growth & Design, Vol. 8, No. 12, 2008 4455 integration of intensity of reflections and scaling;21 SADABS for absorption correction;22 and SHELXL for space group and structure determination, refinements, graphics, and structure reporting.23 The hydrogen atoms of the water were found from the E-map, and the other H atoms in this compound were obtained geometrically. These H atoms were included in the subsequent least-squares refinement as fixed contributors. The final refinement with anisotropic temperature factors for non-H atoms led to an R1 value of 0.05 for this compound. CCDC-662453 contains the supplementary crystallographic data for this paper. These data can be obtained free of charge from the Cambridge Crystallographic Data Centre via www.ccdc.cam.ac.uk/ data_request/cif.

Results and Discussion Synthesis, Spectroscopic, And Electrochemical Characterization of Monomeric Complex Zn[Pc(r-OC5H11)4]. Pure isomer of metal-free 1,8,15,22-tetrakis(3-alkoxy)phthalocyanine H2Pc(R-OC5H11)4 was prepared according to published procedures.15,16 Metalation of this pure isomer of H2Pc(R-OC5H11)4 with Zn(OAc)2 · 2H2O was performed in refluxing n-pentanol, yielding the target monomeric compound Zn[Pc(R-OC5H11)4] with very high yield, 85%.24 To confirm the nature of this compound, the newly prepared phthalocyaninato zinc sample was characterized by elemental analysis and various spectroscopic and electrochemical methods. The MALDI-TOF mass spectra of this compound clearly showed an intense signal for the molecular ion (M)+. The diamagnetic zinc(II) compound gave well-resolved 1H NMR spectra in CDCl3, Figure S1 (Supporting Information). Figure S2 (Supporting Information) gives the UV-vis spectrum of Zn[Pc(R-OC5H11)4]. The spectrum shows a typical Soret band at 318 nm. The Q-band appears around 705 nm as a very strong absorption with a weak vibronic band around 634 nm. The weak absorption band in the region of 350-450 nm is common for alkoxy-substituted phthalocyanines which may be attributed to an n-π* transition.25 The IR spectrum for the single crystal sample of Zn[Pc(ROC5H11)4] dispersed in KBr powder is shown in Figure S3 (Supporting Information). An intense IR band observed at 1337 cm-1 is the marker IR band for phthalocyanine dianion.26 In addition, a group of absorption bands was also seen in the range of 2876-2965 cm-1 in the IR spectrum, which are attributed to the C-H stretching vibrations of the 3-pentyloxy groups. It is worth noting that the bands with medium intensity at 3440 and 3051 cm-1 are characteristic for the O-H stretching of the coordinated water molecule.13,27 In addition, the weak bands at 3066 and 3033 cm-1 are attributed to the O-H · · · N(aza) and O-H · · · O[R-OC5H11)4], respectively.13 These results confirm the H2O-involved hydrogen bonds present in the solid-state single crystal of the phthalocyaninato zinc complex as clearly revealed by the single crystal X-ray diffraction analysis as detailed below. The redox behavior of Zn[Pc(R-OC5H11)4] was studied by cyclic voltammetry (CV) and differential pulse voltammetry (DPV) in CH2Cl2. Typical cyclic voltammetry is displayed in Figure S4 (Supporting Information). Under our experimental conditions, zinc(II) Zn[Pc(R-OC5H11)4] has been revealed to undergo two one-electron oxidations and three one-electron reductions according to the CV and more clearly the DPV results, (Table S1, Supporting Information). All these redox processes can be attributed to the successive removal of electrons from and addition of electrons to the ligand-based orbitals, respectively, as the divalent zinc cannot be oxidized or reduced under the present conditions. The potential difference between Oxd1 and Red1 for Zn[Pc(R-OC5H11)4] is 1.34 V, reflecting the HOMO-LUMO gap.

4456 Crystal Growth & Design, Vol. 8, No. 12, 2008

Li et al.

Scheme 1. Formation of Dimeric Supramolecular Structure {Zn[Pc(r-OC5H11)4] · H2O}2

Structural Studies. The molecular structure of the zinc phthalocyanine Zn[Pc(R-OC5H11)4] · H2O was clearly established by X-ray diffraction analysis. Single crystals suitable for crystallographic analysis were obtained from the toluene solution of Zn[Pc(R-OC5H11)4], Scheme 1. The compound crystallizes in the monoclinic system with a P2(1)/c space group with two pairs of enantiomeric molecules in a unit cell like Pb[Pc(ROC5H11)4].28 In each monomeric Zn[Pc(R-OC5H11)4] unit, the zinc ion is coordinated with four isoindole nitrogen atoms of the phthalocyanine ligand, Pc(R-OC5H11)4. In addition, it also coordinates with an oxygen atom of a water molecule. The coordination polyhedron of the zinc is thus essentially a slightly distorted square pyramid. However, due to the hydrogen bond formed between a zinc atom and an oxygen atom of a water molecule, the divalent zinc ion does not situate in the central hole of Pc(ROC5H11)4 but sits 0.346 Å above the N(isoindole)4 plane. As a result, the substituted Pc(R-OC5H11)4 ring adopts a conformation that is domed toward the zinc cation (φ ) 9.50°). As shown in Figures 1 and S5 (Supporting Information), the two phthalocyanine molecules are bound to each other via four H2O-involved hydrogen bonds. The zinc atom of one molecule binds to a water molecule which also forms two hydrogen bonds with an aza-nitrogen atom and a neighboring oxygen atom from an alkoxy group of another macrocyclic molecule, leading to the formation of a pseudo-double-decker supramolecular structure {Zn[Pc(R-OC5H11)4] · H2O}2. The distance, 2.745 Å, between the water oxygen and aza-nitrogen atom of another Zn[Pc(R-OC5H11)4] molecule is a bit of shorter than that observed for the general hydrogen-bonded O · · · N distance, 2.80 Å,29 indicating the formation of effective hydrogen bonds

Figure 1. Pseudo-double-decker supramolecular structure in crystal of {Zn[Pc(R-OC5H11)4] · H2O}2. Hydrogen atoms have been omitted for clarity.

between these two atoms. This is additionally supported by the large O-H · · · N hydrogen bond angle, 172.42°, while for the O-H · · · O hydrogen bonds in the pseudo-double-decker supramolecular structure, the distance between the water oxygen atom and oxygen atom from the 3-pentyloxy of another Zn[Pc(ROC5H11)4] molecule is 2.727 Å and the O-H · · · O hydrogen bond angle amounts to 124.06°, both of which also indicate the effective hydrogen bonding interaction existing between the two phthalocyanine rings. Nevertheless, the ring-to-ring separation of these two virtually parallel N(isoindole)4 mean planes of phthalocyanine rings in the pseudo-double-decker supramolecular structure is revealed to be 3.728 Å, which is larger than that found in sandwich bis(phthaolocyaninato) rare earth complexes, of 2.71-2.83 Å,30 but a bit of smaller than that between the two Pc(R-OC4H9)8 rings in the slipped quadrupledecker supramolecular structure {SmIII(Pc)[Pc(RIII OC4H9)8]}Na2{Sm (Pc)[Pc(R-OC4H9)8]}, 4.23 Å,31 revealing the relatively intense π-π interaction between the two Pc(ROC5H11)4 rings in the pseudo-double-decker supramolecular structure. As mentioned in the Introduction, except for the phthalocyaninato complexes with a main group IIA metal of Mg and Be,13 H2O- and organic solvent-involved hydrogen bonding interaction has not yet been reported for phthalocyaninato complexes with any other metals. Due to the low eletronegativity and H+ standard electrode potential, both Mg and Be can easily coordinate to H2O using their outer p atomic orbitals.32 As a consequence, even after the formation of phthalocyaninato metal complexes, the H2O molecule coordinated with the central Mg/ Be is able to form hydrogen bonds with the aza-nitrogen atoms from other phthalocyanine ligands easily. However, metals other than Mg and Be in particular the transition metals with outer d or f orbitals are harder to form intense coordination bonds with H2O. As a result, the lack of H2O coordinated with the central transition metal of phthalocyaninato complexes prevents the formation of hydrogen bonds. The π-π interaction between phthalocyanine ligands in these metal complexes therefore dominates the crystallization procedure from solvent and induces the formation of supramolecular structures without any hydrogen bonding interactions.33 However, in the present case, the side 3-pentyloxy groups attached at the nonperipheral positions of the phthalocyanine ring are able to form strong H2O-involved hydrogen bonds, which are believed to play an important role in promoting the formation of coordination bond between the H2O molecule and central zinc ion. As a total result, both the coordination interaction between H2O and the central Zn ion and the two hydrogen bonds between H2O and an aza-nitrogen atom and a neighboring oxygen atom from an alkoxy group of another macrocyclic molecule are responsible for the formation of pseudo-double-decker supramolecular structure {Zn[Pc(ROC5H11)4] · H2O}2. It is also worth noting that the hydrogen bond

1,8,15,22-Tetrasubstituted Phthalocyaninato Zn Complex

Crystal Growth & Design, Vol. 8, No. 12, 2008 4457

Figure 2. Packing arrangement of pseudo-double-decker supramolecular structure in the crystal of {Zn[Pc(R-OC5H11)4] · H2O}2.

formed between the water oxygen atom and the oxygen atom from the side 3-pentyloxy group of the second phthalocyanine molecule prevents the formation of an additional hydrogen bond of the water molecule with the aza-nitrogen atom of the third phthalocyanine ligand or with the oxygen atom of the bulky 3-pentyloxy group of the third phthalocyanine ligand, resulting in the formation of a pseudo-double-decker supramolecular structure {Zn[Pc(R-OC5H11)4] · H2O}2 instead of the infinite onedimensional chains or two-dimensional network in unsubstituted phthalocyaninato complexes of the main group IIA metals Mg and Be, Figures 2 and S6, Supporting Information.13 To the best of our knowledge, the present result represents the first example of phthalocyaninato complexes with a transition metal, for which H2O-involved hydrogen bonds formed with the help of side 3-pentyloxy groups dominate the intermolecular interaction and lead to the formation of a pseudo-double-decker supramolecular structure. To help understand the interaction in the {Zn[Pc(ROC5H11)4] · H2O}2 supramolecular structure, theoretical calculations using the density functional theory method were carried out to calculate the stabilization energy. The calculation reveals that the stabilization energy arisen from both π-π interaction and hydrogen bonding interaction is as high as 64.0 kcal/mol in the supramolecular structure {Zn[Pc(R-OC5H11)4] · H2O}2. It is noteworthy that in the optimized structure, the distance between the two macrocycles in {Zn[Pc(R-OC5H11)4] · H2O}2 is 3.857 Å, which corresponds well with the experimental result of 3.728 Å. For the purpose of comparison, a system composed of two Zn[Pc(R-OC5H11)4] molecules taken from the crystal structure of this work by removing the water molecules is also optimized using the same method as for {Zn[Pc(ROC5H11)4] · H2O}2. The calculated stabilization energy between two Zn[Pc(R-OC5H11)4] moieties with a ring-to-ring distance of 3.857 Å is only 6.4 kcal/mol, significantly smaller than that in {Zn[Pc(R-OC5H11)4] · H2O}2. As a result, H2O-involved hydrogen bond interaction dominates the intermolecular interaction for Zn[Pc(R-OC5H11)4] and is responsible for the formation of the pseudo-double-decker supramolecular structure of {Zn[Pc(R-OC5H11)4] · H2O}2. To further understand the bond nature in the {Zn[Pc(ROC5H11)4] · H2O}2, natural bond orbital (NBO) analysis over

{Zn[Pc(R-OC5H11)4] · H2O}2 was carried out on the basis of crystal structure using the Gaussian 03 program.20 The relevant atoms are labeled in Figure S7 (Supporting Information). It is noteworthy that the NBO analysis is able to provide information for both Lewis and the non-Lewis (Rydberg) structures, however, which corresponds to the hypothetical Lewis structure. As a result, NBO methodology defines three kinds of electrons located in bonds (BD)(electron pairs centered on two atoms), lone pairs (LP)(electron pairs centered on one atom), and core pairs (CR)(electron pairs centered on the core of one atom), respectively. In line with the previous report on the NBO analysis of Ni[Pc(tBu)4] [H2Pc(tBu)4 ) 3(4),12(13),21(22),30(31)tetra(tert-butyl)phthalocyanine],34 the calculated NBO results over the pseudo-double-decker supramolecular structure {Zn[Pc(ROC5H11)4] · H2O}2 reveal that the interaction between zinc atom and isoindole nitrogen atoms (Niso) is through electronic delocalization between bond orbitals from LP(1) Niso to LP(6) Zn instead of through BD. The LP(1) Niso presents about 29% s character and 71% p character, revealing its nature of approximate sp2 hybrid orbitals, while the LP(6) Zn presents a pure s orbital character. These results listed in ( Table S2, Supporting Information) indicate that the Zn-Niso bonds in {Zn[Pc(R-OC5H11)4] · H2O}2 are actually coordinate bonds from the sp2 hybrid orbitals of isoindole nitrogen atoms to the almost unoccupied s orbital of the zinc atom. The electronic delocalization degree can be estimated by second-order perturbation theory from the energy E(2) associated with the donor-acceptor interaction between two NBOs. The higher the E(2) value, the stronger the electronic delocalization degree of the bond, and as a consequence the stronger the donor-acceptor interaction. It is noteworthy that in each Zn[Pc(R-OC5H11)4] system of the pseudo-double-decker supramolecular structure, the E(2) value of one coordinate bond between Zn and one isoindole nitrogen which locates in the isoindole segment with its side-chain oxygen atom forming a hydrogen bond with one water molecule, 43.89 kcal/mol, is a little larger than the remaining three Zn-Niso bonds, 42.42, 42.88, and 42.90 kcal/mol, respectively, revealing the more intense coordination interaction of this bond over the others. This is consistent with the shorter bond length of this Zn-Niso bond in comparison with the three other ones as revealed by the single crystal X-ray diffraction results. In

4458 Crystal Growth & Design, Vol. 8, No. 12, 2008

addition, the Zn-O(water) coordinate bond in {Zn[Pc(ROC5H11)4] · H2O}2 is predicted to be through electronic delocalization between bond orbitals from LP(2) O(water) to LP(6) Zn with LP(2) O(water) composed of about 28% s character and 72% p character, which gives a donor-acceptor interaction energy E(2) of 27.33 kcal/mol. The quite large E(2) value of Zn-O(water) bond indicates the very strong interaction between Zn and the O(water) atom, corresponding well with the quite short Zn-O(water) bond length, 2.137Å. As mentioned above, the two hydrogen bonds formed between H2O with an azanitrogen atom and a neighboring oxygen atom from an alkoxy group of another macrocyclic molecule play an important role in the formation of pseudo-double-decker supramolecular structure {Zn[Pc(R-OC5H11)4] · H2O}2. This is additionally supported by the NBO analysis results. The electronic delocalization between bond orbitals from LP(1) Naza9 to BD*(1)(antibonding orbital) O127-H247(water) induces the donor-acceptor interaction energy of E(2) as high as 21.13 kcal/mol, which is even comparable with the E(2) of Zn-O(water) coordinate bond, indicating the very strong hydrogen bond formed between H2O and Naza9 atom of another macrocyclic molecule. This result is corresponding well with the small distance between the water oxygen and aza-nitrogen atom of another Zn[Pc(R-OC5H11)4] molecule and the large O-H · · · N hydrogen bond angle as mentioned above. However, the donor-acceptor interaction energy E(2) resulting from electronic delocalization between bond orbitals from LP(1,2) O(alkoxy) to BD*(1) O127H247,248(water) is much smaller than from LP(1) Naza9 to BD*(1)(antibonding orbital) O127-H247(water), 4.48 vs 21.13 kcal/mol, indicating the much weaker hydrogen bonding interaction between H2O and a neighboring oxygen atom from an alkoxy group of another macrocyclic molecule than that between H2O and an aza-nitrogen atom of another phthalocyanine ring. This is also rationalized by the larger distance between O(alkoxy) and O127(water) as well as the unsuitable relative orientation of hydrogen atoms of H2O and O(alkoxy). Conclusion In summary, X-ray diffraction analysis over a single crystal of Zn[Pc(R-OC5H11)4] reveals the existence of H2O-involved hydrogen bonds formed with the help of side 3-pentyloxy groups, which dominate the intermolecular interaction and induce the formation of a pseudo-double-decker supramolecular structure in the crystal of this complex. This appears to represent the first example of phthalocyaninato transition metal complexes, for which H2O-involved hydrogen bonds instead of the π-π interaction dominate the formation of supramolecular structure in the solid state. This result is rationalized by density functional theoretical calculations on the stabilization energy and NBO analysis of supramolecular structure {Zn[Pc(R-OC5H11)4] · H2O}2.

Li et al.

(2) (3) (4) (5) (6) (7)

(8) (9) (10) (11)

(12)

(13) (14)

(15) (16) (17) (18) (19)

(20)

Acknowledgment. Financial support from the National Science Foundation of China, Ministry of Education of China, and Shandong University is gratefully acknowledged. Supporting Information Available: 1H NMR, UV-vis, IR, CV, ORTEP diagrams, NBO analysis, and other information. This material is available free of charge via the Internet at http://pubs.acs.org.

References (1) (a) Lever, A. B. P.; Leznoff, C. C. Phthalocyanine: Properties and Applications; VCH: New York, 1989-1996; Vols. 1-4. (b) McKeown, N. B. Phthalocyanines Materials: Synthesis, Structure and Function; Cambridge University Press: New York, 1998. (c) Kadish,

K. M.; Smith, K. M.; Guilard, R. The Porphyrin Handbook; Academic Press: San Diego, 2000-2003; Vols. 1-20. (a) Gregory, P. High-Technology Applications of Organic Colorants; Plenum Press: New York, 1991. (b) Gregory, P. J. Porphyrins Phthalocyanines 2000, 4, 432. (a) Ao, R.; Kilmmert, L.; Haarer, D. AdV. Mater. 1995, 7, 495. (b) Birkett, D. Chem. Ind. 2000, 178. Moser, F. H.; Thomas, A. L. The Phthalocyanines; Vols. 1 and 2, Manufacture and Applications; CRC Press: Boca Raton, FL, 1983. (a) Wo¨hrle, D.; Meissner, D. AdV. Mater. 1991, 3, 129. (b) Eichhorn, H. J. Porphyrins Phthalocyanines 2000, 4, 88. (a) Wright, J. D. Prog. Surf. Sci. 1989, 31, 1. (b) Snow, A. W.; Barger, W. R. In Phthalocyanines: Properties and Applications; Leznoff, C. C.; Lever, A. B. P., Eds.; VCH: New York, 1989; pp 341. (a) Nalwa, H. S.; Shirk, J. S. In Phthalocyanines: Properties and Applications; Leznoff, C. C.; Lever, A. B. P., Eds.; VCH: New York, 1996; p 79. (b) Shirk, J. S.; Pong, R. G. S.; Flom, S. R.; Heckmann, H.; Hanack, M. J. Phys. Chem. 2000, 104, 1438. (c) de la Torre, G.; Va´zquez, P.; Agullo´lo´Pez, F.; Torres, T. J. Mater. Chem 1998, 8, 1671. (a) Luk’yanets, E. A. J. Porphyrins Phthalocyanines 1999, 3, 424. (b) H. Hasrar, H.; van Lier, J. E. Chem. ReV. 1999, 99, 2379. Cosomelli, B.; Roncuccin, G.; Dei, D.; Fantetti, L.; Ferroin, F.; Ricci, M.; Spinelli, D. Tetrahedron 2003, 59, 10025. Beltra´n, H. I.; Esquive, R.; Sosa-Sa´nchez, A.; Sosa-Sa´nchez, J. L.; Ho¨pfl, H.; Barba, V.; Farfan, N.; Garcia, M. G.; Olivares-Xometl, O.; Zamudio-Rivera, L. S. Inorg. Chem. 2004, 43, 3555. (a) Nohr, R. S.; Wynne, K. J. Chem. Commun. 1981, 1210. (b) Wynne, K. J. Inorg. Chem. 1985, 24, 1339. (c) Dirk, C. W.; Inabe, T.; Schoch, K. F.; Marks, T. J. J. Am. Chem. Soc. 1983, 105, 1539. (d) Deneux, M. M.; Benlian, D.; Pierrot, M.; Fournel, A.; Sorbier, J. P. Inorg. Chem. 1985, 24, 1878. (e) Cian, A. D.; Moussavi, M. M.; Fischer, J.; Weiss, R. Inorg. Chem. 1985, 24, 3162. (f) Orhashi, Y.; Kobayashi, N.; Ohno, H.; Tsuchida, E.; Matsuda, H.; Nakanishi, H.; Kato, M. Synth. Met. 1987, 19, 751. (a) Jeffrey, G. A. An Introduction to Hydrogen Bonding; Oxford University Press: New York, 1997. (b) Desiraju, G. R.; Steiner, T. The Weak Hydrogen Bond; OxfordUniversity Press: Oxford, 1999. (c) Dumas, J. M.; Gomel, M.; Guerin, M. Molecular Interactions Involving Organic Halides. In The Chemistry of Functional Groups, Supplement D; Patai, S.; Rappoport, Z., Eds.; John Wiley & Sons Ltd.: New York, 1983; Chapter 21. (d) Legon, A. C. Angew. Chem., Int. Ed. 1999, 38, 2686. (e) Suzuki, S.; Green, P. E.; Bumgarner, R. E. Science 1992, 257, 942. (a) Mizuguchi, J.; Mochizuki, M. Z. Kristallogr. NCS 2002, 217, 244. (b) Mizuguchi, J. Z. Kristallogr. NCS 2002, 217, 251. (c) Janczak, J.; Idemori, Y. M. Polydedron 2003, 22, 1167. (a) Barkigia, K. M.; Palacio, M.; Sun, Y.; Nogues, M.; Renner, M. W.; Varret, F.; Battioni, P.; Mansuy, D.; Fajer, J. Inorg. Chem. 2002, 41, 5647. (b) Boitrel, B.; Halime, Z.; Michaudet, L.; Lachkar, M.; Toupet, L. Chem. Commun. 2003, 2670. (a) Bian, Y.; Wang, R.; Jiang, J.; Lee, C.-H.; Wang, J.; Ng, D. K. P. Chem. Commun. 2003, 1194. (b) Li, R.; Zhang, X.; Zhu, P.; Li, X.; Ng, D. K. P.; Kobayashi, N.; Jiang, J. Inorg. Chem. 2006, 45, 2327. Liu, W.; Lee, C.-H.; Li, H.-W.; Lam, C.-K.; Wang, J.; Mak, T. C. W.; Ng, D. K. P. Chem. Commun. 2002, 628. One Molecular Simulation Software, Inc., see http://www.accelrys.com. (a) Lee, C.; Yang, W.; Parr, R. G. Phys. ReV. B 1998, 37, 785. (b) Becke, A. D. J. Chem. Phys. 1993, 98, 5648. (a) Ditchfield, R.; Hehre, W. J.; Pople, J. A. J. Chem. Phys. 1971, 54, 724. (b) Hehre, W. J.; Ditchfield, R.; Pople, J. A. J. Chem. Phys. 1972, 56, 2257. (c) Hariharan, P. C.; Pople, J. A. Mol. Phys. 1974, 27, 209. (d) Gordon, M. S. Chem. Phys. Lett. 1980, 76, 163. (e) Hariharan, P. C.; Pople, J. A. Theor. Chim. Acta 1973, 28, 213. Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Montgomery, J. A., Jr.; Vreven, T.; Kudin, K. N.; Burant, J. C.; Millam, J. M.; Iyengar, S. S.; Tomasi, J.; Barone, V.; Mennucci, B.; Cossi, M.; Scalmani, G.; Rega, N.; Petersson, G. A.; Nakatsuji, H.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Klene, M.; Li, X.; Knox, J. E.; Hratchian, H. P.; Cross, J. B.; Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Ayala, P. Y.; Morokuma, K.; Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Zakrzewski, V. G.; Dapprich, S.; Daniels, A. D.; Strain, M. C.; Farkas, O.; Malick, D. K.; Rabuck, A. D.; Raghavachari, K.; Foresman, J. B.; Ortiz, J. V.; Cui, Q.; Baboul, A. G.; Clifford, S.; Cioslowski, J.; Stefanov, B. B.; Liu, G.; Liashenko, A.; Piskorz, P.; Komaromi, I.; Martin, R. L.; Fox, D. J.; Keith, T.; Al-Laham, M. A.; Peng, C. Y.; Nanayakkara, A.; Challacombe, M.; Gill, P. M. W.; Johnson, B.; Chen,

1,8,15,22-Tetrasubstituted Phthalocyaninato Zn Complex

(21) (22) (23) (24) (25)

(26) (27) (28)

W.; Wong, M. W.; Gonzalez, C.; Pople, J. A. Gaussian 03, Revision B.05; Gaussian, Inc.: Pittsburgh, PA, 2003. SMART and SAINT for Windows NT Software Reference Manuals, Version 5.0, Bruker Analytical X-Ray Systems, Madison, WI, 1997. Sheldrick, G. M. SADABS - A Software for Empirical Absorption Correction, University of Go¨ttingen, 1997. SHELXL Reference Manual, Version 5.1, Bruker Analytical X-Ray Systems, Madison, WI, 1997. Kasuga, K.; Asano, K.; Lin, L.; Sugimori, T.; Handa, M.; Abe, K.; Kikkawa, T.; Fujiwara, T. Bull. Chem. Soc. Jpn. 1997, 70, 1859. (a) Jiang, J.; Arnold, D. P.; Yu, H. Polyhedron 1999, 18, 2129. (b) Lu, F.; Bao, M.; Ma, C.; Zhang, X.; Arnold, D. P.; Jiang, J. Spectrochim. Acta A 2003, 59, 3273. (c) Bao, M.; Pan, N.; Ma, C.; Arnold, D. P.; Jiang, J. Vibr. Spectrosc. 2003, 32, 175. (d) Zhang, X.; Zhang, Y.; Jiang, J. Vibr. Spectrosc. 2003, 33, 153. Jiang, J.; Bao, M.; Rintoul, L.; Arnold, D. P. Coord. Chem. ReV. 2006, 250, 424, and references therein. Sidorov, A. N.; Kotlyar, I. P. Opt. Spectrosc 1961, 11, 175 (Engl. Transl.). Bian, Y.; Li, L.; Dou, J.; Cheng, D. Y. Y.; Li, R.; Ma, C.; Ng, D. K. P.; Kobayashi, N.; Jiang, J. Inorg. Chem. 2004, 43, 7539.

Crystal Growth & Design, Vol. 8, No. 12, 2008 4459 (29) Pimentel, G. C.; McClellan. A. L. In The Hydrogen Bond; Freeman, W. H., Ed.; San Francisco, 1960, p 289. (30) Jiang, J.; Kasuga, K.; Arnold, D. P. In Supramolecular PhotosensitiVe and ElectroactiVe Materials; Nalwa, H. S., Ed.; Academic Press: New York, 2001; chapter 2, pp 113-210. (31) Zhang, H.; Wang, R.; Zhu, P.; Han, J.; Lu, F.; Lee, C.-H.; Ng, D. K. P.; Cui, X.; Ma, C.; Jiang, J. Inorg. Chem. 2004, 43, 4740. (32) (a) Shipman, L. L.; Cotton, T. M.; Norris, J. R.; Katz, J. J. J. Am. Chem. Soc. 1976, 98, 8222. (b) Okada, K.; Unhara, K.; Ozaki, Y. Photochem. Photobiol. 1993, 57, 958. (c) Kubiak, R.; Waskowska, A.; Pietraszko, A.; Bukowska, E. Inorg. Chim. Acta 2005, 358, 453. (33) (a) Takano, S.; Naito, T.; Inable, T. Chem. Lett. 1998, 1249. (b) Ballirano, P.; Caminiti, R.; Ercolani, C.; Maras, A.; Orru, M. A. J. Am. Chem. Soc. 1998, 120, 12798. (c) Matsuda, M.; Naito, T.; Inabe, T.; Hanasaki, N.; Tajima, H. J. Mater. Chem. 2001, 11, 2493. (d) Bench, B. A.; Beveridge, A.; Sharman, W. M.; Diebold, G. J.; Van Lier, J. E.; Gorun, S. M. Angew. Chem., Int. Ed. 2002, 41, 747. (34) Berrı´os, C.; Ca´rdenas-Jiro´n, G. I.; Marco, J. F.; Gutie´rrez, C.; UretaZaartu, M. S. J. Phys. Chem. A 2007, 111, 2706.

CG800342B