Polyhedron 25 (2006) 1753–1762 www.elsevier.com/locate/poly

Synthesis, structures and fluorescence of nickel, zinc and cadmium complexes with the N,N,O-tridentate Schiff base N-2-pyridylmethylidene-2-hydroxy-phenylamine Arpi Majumder a, Georgina M. Rosair b,*, Arabinda Mallick c, Nitin Chattopadhyay c, Samiran Mitra a,* a

Department of Chemistry, Inorganic Chemistry Section, Jadavpur University, S.C. Mallik Road, Kolkata, West Bengal 700 032, India b Department of Chemistry, Heriot-Watt University, Edinburgh EH14 4AS, UK c Department of Chemistry, Physical Chemistry Section, Jadavpur University, S.C. Mallik Road, Kolkata, West Bengal 700 032, India Received 19 August 2005; accepted 14 November 2005 Available online 18 January 2006

Abstract Mono-, tri- and dinuclear neutral complexes [Ni(HL)(L)] Æ (ClO4) Æ 0.16(H2O) (1), [ZnLZn(OOCCH3)4ZnL] (2) and [Cd2L2(OCH3CO)2(H2O)2] (3) have been obtained from the reaction between the potentially tridentate N,N,O-donor Schiff base ligand HL, where HL = N-2-pyridylmethylidene-2-hydroxy-phenylamine with nickel, zinc or cadmium salts, respectively. The ligand has been prepared by 1:1 condensation of pyridine-2-carboxaldehyde and 2-aminophenol. The ligand and metal complexes were characterised by elemental analysis, spectroscopic studies such as IR, UV–Vis, 1H NMR, fluorescence, electrochemical and magnetic susceptibility measurement. The structures of the three complexes have been determined by single-crystal X-ray diffraction. The nickel ions in 1 show a distorted mer-octahedral geometry. In 2, the terminal zinc ions have coordination geometry midway between square pyramidal and trigonal bipyramidal, whilst the central zinc ion has slightly distorted octahedral geometry. The trinuclear unit is held together by bridging deprotonated phenolic oxygen atoms from the Schiff base and acetate groups. In 3, two monocapped-octahedron cadmium ions are held together by l2-diphenoxo bridges. Among the three synthesised complexes, 1 is nonfluorescent while the other two can serve as potential photoactive materials as indicated from their characteristic fluorescence properties.  2005 Elsevier Ltd. All rights reserved. Keywords: Mononuclear Ni(II); Trinuclear Zn(II); Dinuclear heptadentate Cd(II); l2-Diphenoxo bridged complexes; Tridentate Schiff base ligand; Crystal structures; Fluorescence

1. Introduction Schiff base ligands have been extensively studied in coordination chemistry mainly due to their facile syntheses, easily tunable steric, electronic properties and good solubility in common solvents [1]. Transition metal complexes with oxygen and nitrogen donor Schiff bases are of particular interest [2] because of their ability to possess unusual configurations, be structurally labile and their sensitivity to molecular environments [3]. Schiff base ligands have pro*

Corresponding authors. Tel.: +91 33 2668 2017; fax: +91 33 2414 6414. E-mail address: [email protected] (S. Mitra).

0277-5387/$ - see front matter  2005 Elsevier Ltd. All rights reserved. doi:10.1016/j.poly.2005.11.029

ven to be very effective in constructing supramolecular architectures such as coordination polymers, double helixes, and triple helicates [4]. Schiff bases can accommodate different metal centres involving various coordination modes allowing successful synthesis of homo- and heterometallic complexes with varied stereochemistry [5]. This feature is employed for modelling active sites in biological systems [6]. Additionally, they have wide applications in fields such as antibacterial, antiviral, antifungal agents [7], homogeneous or heterogeneous catalysis [8] and magnetism [9]. Schiff bases are potential anticancer drugs [10] and when administered as their metal complexes, the anticancer activity of these complexes is enhanced in

1754

A. Majumder et al. / Polyhedron 25 (2006) 1753–1762

comparison to the free ligand [11]. It has been shown that Schiff base complexes derived from 4-hydroxysalicylaldehyde and amines have strong anticancer activity, e.g., against Ehrlich ascites carcinoma (EAC) [12]. The p-system in a Schiff base often imposes geometrical constraints as well as affecting electronic structure [13]. Schiff bases derived from a large number of carbonyl compounds and amines have been used [3], however, the studies on their optical properties, such as fluorescence, are rare. The variety of possible Schiff-base metal complexes with a wide choice of ligands, and coordination environments, has prompted us to undertake research in this area. As a part of our continuing work on Schiff base complexes [5,14,15], nickel(II), zinc(II) and cadmium(II) complexes with notable photo-activity were prepared and structurally characterised. The chemistry of nickel Schiff base complexes has a strong role in bioinorganic chemistry and redox enzyme systems [6a,6b]. Morrow and Kolasa reported the cleavage of plasmid DNA by square planar nickel-salen [bis-(salicylidene)ethylenediamine] in the presence of either magnesium mono peroxypthalic acid (MPPA) or iodosulbenzene [16]. Zinc is an important transition metal in biological systems. Zinc-containing carboxylate-bridged complexes [2] have varied structural motifs in hydrolytic metalloenzymes, such as phosphatases and aminopeptidases. The catalytic role of Zn comprises Lewis acid activation of the substrate, generation of a reactive nucleophile (Zn–OH) and stabilisation of the leaving group [17]. There is substantial interest in the coordination chemistry of cadmium complexes because of the toxic environmental impact of cadmium. The mobilisation and immobilisation of cadmium in the environment, in organisms, and in some technical processes (such as in ligand exchange chromatography) have been shown to depend significantly on the complexation of the metal centre by chelating nitrogen donor ligands [18]. In this paper, we describe the preparation and structures of nickel(II), zinc(II) and cadmium(II) complexes with the related Schiff base ligand N-2-pyridylmethylidene-2hydroxy-phenylamine displaying strong fluorescent emission at room temperature. 2. Experimental 2.1. Physical techniques The infrared spectra of the complexes were recorded on a Perkin–Elmer RX 1 FT-IR spectrophotometer with a KBr disc. Elemental analyses were carried out using a Perkin–Elmer 2400 II elemental analyser. Electrochemical study for complex 1 was performed on a CH 600A cyclic voltammeter instrument in methanol, with tetrabutylammonium perchlorate as the supporting electrolyte. 1H NMR spectral measurements were carried out on a Bruker FT300 MHz spectrometer with TMS as an internal reference. Magnetic susceptibility was measured on a powdered sample in a vibrating sample magnetometer for 1 using mercury(tetrathiocyanato)cobaltate as the standard. The

electronic spectra of the complexes were recorded on a Perkin–Elmer Lambda 40 (UV–Vis) spectrophotometer in methanol. Steady state fluorescence measurements were performed using a Spex fluorolog II spectrofluorimeter. For spectroscopic measurements, the concentration of the solutions were ca. 2 · 105 M. Fluorescence quantum yields were determined against phenosafranine in MeOH (uf = 0.20) [19]. 2.2. Materials All the chemicals and solvents used for the syntheses were of reagent grade. Ni(ClO4)2 Æ 6H2O, Zn(CH3COO)2 Æ 2H2O, Cd(CH3COO)2 Æ 2H2O, pyridine-2-carboxaldehyde (Fluka) and 2-aminophenol (Merck) were used as received. Caution! Although no problems were encountered in this work, perchlorate salts are potentially explosive. They should be prepared in small quantities and handled with care. 2.3. Synthesis 2.3.1. Synthesis of the Schiff base ligand The ligand HL [N-2-pyridylmethylidine-2-hydroxyphenylamine] was synthesised by stirring pyridine 2-carboxaldehyde (0.380 ml, 4 mmol) and 2-aminophenol (0.436 g, 4 mmol) in 10 ml methanol, resulting in a deep red solution containing the liquid ligand (HL) and this was used for the preparation of the complex without further purification. 1H NMR (Scheme 1): 10.09 (s, 1H) phenolic –OH, disappears upon shaken with D2O; 8.84 (s, 1H) H–C@N; 8.73 (d, 1H, J = 4.44 Hz) C14–H; 8.20 (d, 1H, J = 7.95 Hz) C11–H; 7.97 (t, 1H, J = 8.4 Hz) C13–H; 7.92 (t, 1H, J = 14.7 Hz) C12–H; 7.54 (t, 1H, J = 5.7 Hz) C5–H; 7.03 (d, 1H, J = 8.28 Hz) C3–H; 6.93 (t, 1H, J = 7.62 Hz) C4–H. 2.3.2. Synthesis of [Ni(HL)(L)] Æ (ClO4) Æ 0.16(H2O) (1) A methanolic solution of Ni(ClO4)2 Æ 6H2O (0.365 g, 1 mmol) was added to a stirred solution (190 ml, 2 mmol) of the above ligand. Green hexagonal crystals of 1 appeared after 2 days. Yield: 0.42 g (75.6%). Anal. Calc. for C24H19.33ClN4NiO6.16: C, 51.74; H, 3.47; N, 10.06; Cl, 6.37; Ni, 10.54. Found: C, 51.29; H, 3.42; N, 10.48; Cl, 6.2; Ni, 10.43%.

H

H C13

H

C14 O1

N15

N

C12 C11

C10 C9

C

H

HO

H C2

N8

C3

C7

N

C4 C6

H H Scheme 1.

C5

H

H

A. Majumder et al. / Polyhedron 25 (2006) 1753–1762

2.3.3. Synthesis of [ZnLZn(OOCCH3)4ZnL] (2) A methanolic solution of Zn(CH3COO)2 Æ 2H2O (0.219 g, 1 mmol) was added to a stirred solution (0.095 ml, 1 mmol) of the above ligand. Red hexagonal crystals of complex 2 appeared after 2 days. Yield: 0.61 g (74.2%). Anal. Calc. for C32H30N4O10Zn3: C, 46.44; H, 3.62; N, 6.77; Zn, 23.72. Found: C, 46.31; H, 3.49; N, 6.68; Zn, 23.45%. 1H NMR (Fig. 4B): 8.81 (s, 1H) H– C@N; 8.73 (d, 1H, J = 4.11 Hz) C14B–H; 8.64 (s, 1H) H–C@N; 8.42 (d, 1H, J = 4.34 Hz) C11B–H; 8.13 (t, 1H, J = 7.48 Hz) C13B–H; 7.77 (d, 1H, J = 7.62 Hz) C3B–H; 7.72 (d, 1H, J = 4.95 Hz) C6B–H; 7.45 (t, 1H, J = 6.81 Hz) C4B–H; 6.75 (t, 1H, J = 7.56 Hz) C5B–H; 3.49 (s, 1H) CH3 (acetate). 2.3.4. Synthesis of [Cd2L2(OCH3CO)2(H2O)2] (3) A methanolic solution of Cd(CH3COO)2 Æ 2H2O (0.266 g, 1 mmol) was added to a stirred solution (0.095 ml, 1 mmol) of the above ligand. Red cubic crystals of 3 obtained after 2 days. Yield: 0.59 g (76%). Anal. Calc. for C28H28Cd2N4O8: C, 43.22; H, 3.60; N, 7.20; Cd, 28.92. Found: C, 43.10; H, 3.49; N, 6.01; Cd, 28.51%. 1H NMR: 8.94 (s, 1H) H–C@N; 8.88 (d, 1H, J = 4.56 Hz) C12–H; 8.81 (d, 1H, J = 4.59 Hz) C9–H; 8.05 (t, 1H, J = 4.90 Hz) C10–H; 7.93 (t, 1H, J = 8.64 Hz) C11–H; 7.58 (d, 1H, J = 3.96 Hz) C2–H; 7.50 (d, 1H, J = 7.38 Hz) C5–H. 2.4. X-ray crystallography A crystal of 1 was mounted with glue on glass fibres and ˚ ) at data were collected with Mo Ka radiation (0.7107 A

1755

293 K on a Rigaku AFC7 diffractometer equipped with a Mercury CCD detector. The space group is non-centrosymmetric, but the Flack x parameter could not be determined reliably. Checks for the presence of missed crystallographic symmetry were made using the checkcif procedure employed by the IUCr (http://journals. iucr.org/services/cif/checkcif.html). No correlation coefficients greater than 0.5 were found between potentially symmetry related parameters but some atoms are highly anisotropic. No solution was found in the higher symmetry space group (C2/m). A single red coloured crystal of 2 was mounted with vacuum grease on glass fibres on a Bruker Nonius X8 Apex2 diffractometer. Data were collected with ˚ ) at 100 K with an Oxford Mo Ka radiation (0.7107 A Cryosystems Cryostream. No significant crystal decay was found for both 1 and 2. The structures of both 1 and 2 were solved by direct methods and crystallographic computing was performed using the SHELXTL suite of programs [20]. All non-hydrogen atoms were refined with anisotropic displacement parameters and most H atoms were constrained to idealised geometries and refined with riding isotropic displacement parameters. Phenolic and water H atoms in 1 were refined freely but with O–H distances ˚ , respectively. Further restrained to 0.98(2) and 0.90(2) A details are given in Table 1. The occupancy of the water molecule for 1 was estimated at 50% based on obtaining an equivalent isotropic displacement parameter of 0.132 which is not unreasonable for a solvent molecule in a structure determined at room temperature. The X-ray single crystal data for 3 was collected at 293 K on an Enraf-Nonius CAD-4 MACH 3

Table 1 Crystallographic data for the complexes 1, 2 and 3 Compound

1

2

3

Chemical formula Formula weight Crystal system T (K) Space group ˚) a (A ˚) b (A ˚) c (A a () b () c () ˚ 3) V (A Z Reflections collected Independent reflections Observed data Unique data (Rint) Dcalc (Mg/m3) Absorption coefficient (mm1) F(0 0 0) Crystal size (mm) h Range for data collection () R indices (all data) Final R indices [I > 2r(I)] ˚ 3) Largest difference in peak and hole (e A

C24H19NiN4O2 Æ ClO4 Æ 0.16(H2O) 556.59 monoclinic 293(2) C2 31.292(6) 12.047(2) 19.679(4) 90.00 99.70(3) 90.00 7312(3) 12 11 902 7896 4799 0.0592 1.517 0.954 3428 0.14 · 0.14 · 0.10 1.54–24.11 R1 = 0.1397, wR2 = 0.2421 R1 = 0.0895, wR2 = 0.2054 0.795 and 0.498

C32H30Zn3N4O10 826.75 triclinic 100(2) P 1 7.923(3) 10.828(3) 19.419(7) 99.006(18) 95.227(19) 99.799(17) 1609.3(10) 2 50 088 11 474 9619 0.0257 1.706 2.283 840 0.36 · 0.22 · 0.18 1.07–32.63 R1 = 0.0367, wR2 = 0.0790 R1 = 0.0281, wR2 = 0.0754 0.722 and 0.411

C28H28Cd2N4O8 777.34 triclinic 293(2) P 1 7.438(3) 9.440(2) 11.895(3) 106.38(5) 97.78(3) 113.01(6) 708.5(4) 1 2988 2844 2736 0.0173 1.813 1.558 384 0.33 · 0.28 · 0.05 2.52–26.16 R1 = 0.0280, wR2 = 0.0709 R1 = 0.0267, wR2 = 0.0700 0.461 and 1.571

1756

A. Majumder et al. / Polyhedron 25 (2006) 1753–1762

diffracto-meter using graphite-monochromated Mo Ka ˚ ): orientation matrices and unit radiation (k = 0.71073 A cell parameters from the setting angles of 25 centred medium-angle reflections; collection of the diffraction intensities by x scans (data corrected for absorption using refined methods) [21]; Tmin = 0.5936, Tmax = 0.8731. The structure was solved by direct methods and subsequently refined by full-matrix least-squares procedures on F2 with allowance for anisotropic thermal motion of all non-hydrogen atoms employing the WinGX package [22] with the relevant programs (SIR-97 [23], SHELXL-97 [24], ORTEP-3 [25]) implemented therein. Further details are given in Table 1. 3. Results and discussion 3.1. Description of the crystal structures 3.1.1. [Ni(HL)(L)] Æ (ClO4) Æ 0.16(H2O) (1) In the molecular structure of 1, there are three crystallographically independent molecules in the asymmetric unit (atoms labeled A, B and C) along with three perchlorate anions and half of one water molecule, which is shown in Fig. 1. Selected bond distances and angles are listed in Table 2. Despite several attempts, we could not obtain better data on 1 so the structure of 1 will not be discussed in great detail. The nickel complexes associate as hydrogen bonded dimers involving phenolic hydrogen atoms. One pair, molecules B and C, is shown in Fig. 2. Molecule A forms a hydrogen bonded dimer with another molecule A related by a twofold axis. Packing diagram with intermolecular hydrogen bonding interactions is depicted in Fig. 3 (Table 3). Evidence for the presence of phenolic hydrogen atoms came from infrared spectroscopy and

Table 2 ˚ ) and angles () for 1 Selected bond distances (A Ni1–N8A Ni1–N1A

1.953(14) 2.091(14)

Ni1–N23A Ni1–O15A

2.013(15) 2.097(11)

Ni1–N30A Ni1–O16A

2.083(13) 2.147(11)

Ni2–N8B Ni2–N1B

2.030(12) 2.114(13)

Ni2–N23B Ni2–O15B

2.006(13) 2.053(10)

Ni2–N30B Ni2–O16B

2.081(12) 2.136(10)

Ni3–N8C Ni3–N1C

2.004(14) 2.099(14)

Ni3–N23C Ni3–O15C

2.006(13) 2.075(10)

Ni3–N30C Ni3–O16C

2.112(12) 2.117(10)

N8A–Ni1–N23A N23A–Ni1–N30A N23A–Ni1–N1A N8A–Ni1–O15A N30A–Ni1–O15A N8A–Ni1–O16A N30A–Ni1–O16A O15A–Ni1–O16A

176.8(6) 77.9(7) 102.2(6) 79.0(5) 95.2(4) 103.9(5) 156.6(6) 89.9(4)

N8A–Ni1–N30A N8A–Ni1–N1A N30A–Ni1–N1A N23A–Ni1–O15A N1A–Ni1–O15A N23A–Ni1–O16A N1A–Ni1–O16A

99.5(6) 79.5(5) 89.8(5) 99.3(5) 158.5(5) 78.8(5) 93.7(5)

N8B–Ni2–N23B N23B–Ni2–N30B N23B–Ni2–N1B N8B–Ni2–O15B N30B–Ni2–O15B N8B–Ni2–O16B N30B–Ni2–O16B O15B–Ni2–O16B

177.7(6) 79.7(6) 103.7(6) 81.7(5) 93.7(4) 104.4(5) 157.5(5) 90.8(4)

N8B–Ni2–N30B N8B–Ni2–N1B N30B–Ni2–N1B N23B–Ni2–O15B N1B–Ni2–O15B N23B–Ni2–O16B N1B–Ni2–O16B

98.0(5) 76.4(6) 94.6(5) 98.4(5) 157.4(5) 77.9(5) 89.5(4)

N8C–Ni3–N23C N8C–Ni3–N1C N8C–Ni3–O15C N1C–Ni3–O15C N30C–Ni3–O16C N8C–Ni3–N30C N23C–Ni3–N1C N23C–Ni3–O15C

173.8(6) 80.1(7) 79.1(6) 158.5(6) 157.8(6) 93.6(6) 99.1(6) 102.2(5)

N8C–Ni3–O16C N1C–Ni3–O16C N23C–Ni3–N30C N30C–Ni3–N1C N30C–Ni3–O15C N23C–Ni3–O16C O15C–Ni3–O16C

108.6(5) 91.1(5) 80.3(6) 92.0(4) 94.5(4) 77.5(5) 90.5(4)

Fig. 2. Perspective view of three crystallographically independent and one symmetry related units of the nickel complex along with the three perchlorate ions and water molecule in the unit cell.

Fig. 1. Two of the three crystallographically independent molecules of 1 with atom numbering scheme showing the hydrogen-bonded dimeric arrangement. Lattice water and perchlorate molecules are omitted for clarity.

these hydrogen atoms were located in the difference Fourier map. Two tridentate Schiff base ligands (of which one of the phenolic groups is deprotonated) are bonded to the nickel(II) ion in a tridentate fashion resulting in a mer octahedral geometry. The largest deviations from octahedral geometry arise from the geometric constraints of this chelating ligand, for example, the trans angle involving N30 and O16 which is in the range of 156.6(6)–157.8(6) for the three independent molecules. A similar situation is seen for N1 and O15 (see Table 2).

A. Majumder et al. / Polyhedron 25 (2006) 1753–1762

1757

Fig. 3. Packing diagram of complex 1 with intermolecular hydrogen bonding interactions.

Table 3 ˚ and ) Hydrogen bonds for 1 (A D–H  A

d(D–H)

d(H  A)

d(D  A)

\(DHA)

O15B–H15B  O16C O15C–H15C  O16B O15A–H15A  O16A#1 O1W–H2W  O3S#2 O1W–H1W  O9S

0.97(2) 0.98(2) 0.99(2) 0.96 0.81

1.52(7) 1.53(10) 1.47(7) 2.51 2.30

2.421(14) 2.466(14) 2.424(15) 3.263(18) 2.872(19)

152(13) 156(23) 161(18) 136.00 129.00

Fig. 4. Molecular structure of complex 2 with two crystallographically independent molecules A and B showing atom numbering scheme. Zn2C and Zn2D are generated by the symmetry operations 1  x, y, z and 1x, 2  y, 1  z, respectively.

Symmetry transformations used to generate equivalent atoms: #1: x  1, y, z; #2: x  1/2, y + 1/2, z.

3.1.2. [ZnLZn(OOCCH3)4ZnL] (2) The crystal structure of 2 consists of two crystallographically independent molecules A and B which lie on the centres of inversion. Fig. 4 gives an ORTEP view of the complex 2 with the atomic labeling scheme. Selected bond lengths and angles are presented in Table 4. The trinuclear complex is built up of two mononuclear ZnL moieties which are linked through bridging acetate and phenolate groups to the central Zn atom. This structure is analogous to a previously reported manganese-complex [26]. The co-ordination geometry around the terminal Zn centres may be regarded as midway between trigonal bipyramidal and square pyramidal as described by the s parameter [27] 0.45 and 0.43 for Zn2A and Zn2B, respectively. The equatorial plane of the two terminal Zn atoms are formed by the bridging acetate (O2A, O5A) and imine nitrogen atom (N8A) while the pyridine nitrogen (O15) and phenolic oxygen atom (O1) of the Schiff base lie in the axial position. ˚ ), –pyridine The Zn2A–imine nitrogen (2.0852(13) A ˚ ˚ ˚) (2.2081(13) A), –acetate (1.9647(12) A and 1.9701(12) A ˚ and –phenolic oxygen (2.0724(11) A) bond distances are in the range observed for similar systems [28]. The observed Zn1A–O1A–Zn2A angle is 108.41(5). The deviations from ideal trigonal bipyramidal and square pyramidal geometries

Table 4 ˚ ) and angles () for 2 Selected bond distances (A Zn1A–O1A Zn1A–O3A Zn2A–N8A Zn2A–O1A Zn2A–O2A Zn2A–O5A Zn1A–O4A Zn2A–N15A O1A–Zn1A–O3A O1A–Zn1A–O3A#1 O1A–Zn1A–O4A#1 O3A–Zn1A–O4A O1A–Zn1A–O4A O1A–Zn1A–O1A#1 O2A–Zn2A–O1A O5A–Zn2A–O1A O2A–Zn2A–O5A O2A–Zn2A–N8A O2A–Zn2A–N15A O5A–Zn2A–N8A O5A–Zn2A–N15A N8A–Zn2A–N15A O1A–Zn2A–N15A O3A–Zn1A–O4A#1 O1A–Zn2A–N8A

2.0584(11) 2.0874(12) 2.0852(13) 2.0724(11) 1.9647(12) 1.9701(12) 2.1645(13) 2.2081(13) 89.67(5) 90.33(5) 91.92(5) 92.82(5) 88.08(5) 180.00(6) 103.89(5) 98.72(5) 111.92(5) 125.75(5) 88.24(5) 121.46(5) 98.84(5) 75.41(5) 153.04(4) 87.18(5) 78.00(5)

Zn1B–O1B Zn1B–O3B Zn1B–O4B Zn2B–O1B Zn2B–O2B Zn2B–O5B Zn2B–N8B Zn2B–N15B O1B–Zn1B–O3B O1B–Zn1B–O3B#2 O1B–Zn1B–O4B#2 O3B–Zn1B–O4B O1–Zn1B–O4B O3B#2–Zn1B–O3B O2B–Zn2B–O1B O5B–Zn2B–O1B O5B–Zn2B–O2B O2B–Zn2B–N8B O2B–Zn2B–N15B O5B–Zn2B–N8B O5B–Zn2B–N15B N8B–Zn2B–N15B O1B–Zn2B–N15B O3B–Zn1B–O4B#2 O1B–Zn2B–N8B

2.0557(11) 2.0838(13) 2.1335(12) 2.0613(11) 1.9797(13) 1.9432(12) 2.0996(14) 2.2004(12) 89.52(5) 90.48(5) 89.33(5) 93.53(5) 90.67(5) 180.0 101.62(5) 102.55(5) 112.34(6) 126.95(5) 92.51(5) 119.56(5) 93.00(5) 75.18(5) 152.95(5) 86.47(5) 77.90(5)

Symmetry transformations used to generate equivalent atoms: #1: x + 1, y, z; #2: x + 1, y + 2, z + 1.

1758

A. Majumder et al. / Polyhedron 25 (2006) 1753–1762

are indicated by the bond angles at Zn2A and Zn2B shown in Table 4, e.g., the largest angle at each metal centre is 153.04(4) and 152.95 for A and B, respectively. However, the coordination geometry of the central Zn(II) atoms deviates very slightly from an ideal octahedron. They are bound by four equatorial oxygen atoms belong to the bridging acetates and the other two from the phenolic oxygens, which lie trans to each other. The distance between two neighbouring zinc atoms in ˚ and double that (6.736 A ˚ ) between molecule A is 3.368 A the terminal Zn centres because the Zn  Zn  Zn angle is 180. In molecule B these distances are slightly shorter, ˚ and 6.679 A ˚ , respectively. Few crystal structures 3.339 A of trinuclear Zn(II) complexes have been described in the literature. These have either a linear or a bent structure [28]. The planes between the bridging acetate groups are 66.96(18) and 57.03(15) for molecules A and B, respectively. The difference could stem from the significant distortion of the zinc coordination geometry from ideal tetrahedral for Zn(II). This deviation from ideal geometry may encourage the molecule to try different relative arrangement of the acetate bridges because it has to try and compromise between what is ideal and what is energetically and sterically possible. Two acetate anions are coordinated to two zinc ions as l-acetato-O,O 0 syn–syn mode whereas in the related Mn system [26], the syn–anti mode is observed too. In the crystal packing the neutral molecules stack along the crystallographic a axis. Adjacent stacks are connected via intermolecular C–H  O hydrogen bonds involving the Schiff base ligands (imine hydrogen and other aromatic hydrogens) or acetate hydrogens and the acetate oxygens with C  O distances in the range of ˚ (Fig. 5, Table 5). Acetate oxygen 3.047(2)–3.487(2) A atoms in both A and B participate in C–H  O hydrogen bonding, although all the contacts are long. In the arrange˚ is long, the ment, C13B–H13B  O5A the H  O of 2.55 A ˚) contact observed for O5B is even longer (H  O 2.62 A and intermolecular rather than intramolceular (see Fig. 5).

Table 5 ˚ and ) C–H  O contacts for 2 (A D–H  A #1

C3A–H3A  O3A C3B–H3B  O3B#2 C9A–H9A  O2A#3 C9B–H9B  O2B#4 C13B–H13B  O5A C19B–H19E  O3B#5

d(D–H)

d(H  A)

d(D  A)

\(DHA)

0.95 0.95 0.95 0.95 0.95 0.98

2.58 2.60 2.53 2.49 2.55 2.47

3.233(2) 3.173(2) 3.229(2) 3.0465(19) 3.487(2) 3.447(2)

125.9 119.3 130.6 117.8 168.0 172.0

Symmetry transformations used to generate equivalent atoms: #1: x + 1, y, z; #2: x + 1, y + 2, z + 1; #3: x + 1, y + 1, z; #4: x + 1, y + 1, z + 1; #5: 1 + x, y, z.

3.1.3. [Cd2L2(OCH3CO)2(H2O)2] (3) A perspective view of the complex 3 with an atom numbering scheme is shown in Fig. 6. Selected bond distances and angles are summarised in Table 6. The complex sits on a crystallographically imposed centre of inversion, forming the bridged binuclear structure with both cadmium centres being heptacoordinate. A coordination number of seven is relatively uncommon in Cd(II) complexes [29]. The local coordination around the Cd(II) ion can be

Fig. 6. The coordination environment of the Cd(II) ions complex 3 with atom numbering scheme showing intramolecular hydrogen bonding interactions.

Table 6 ˚ ) and angles () for complex 3 Selected bond distances (A Cd1–O1A Cd1–O2 Cd1–O3

Fig. 5. Crystal packing arrangement for complex 2 showing intermolecular hydrogen bonding interactions.

O1A–Cd1–O1 O1–Cd1–N2 O1–Cd1–O2 O1A–Cd1–N1 N2–Cd1–N1 O1A–Cd1–O4 N2–Cd1–O4 N1–Cd1–O4 O1–Cd1–O3 O2–Cd1–O3 O4–Cd1–O3

2.245(2) 2.348(3) 2.599(3)

Cd1–O1 Cd1–N1

73.32(9) 69.80(9) 134.73(8) 142.96(8) 69.83(9) 78.83(10) 90.07(10) 85.41(8) 85.38(8) 52.40(8) 163.88(7)

2.324(2) 2.414(3)

Cd1–N2 Cd1–O4

O1A–Cd1–N2 O1A–Cd1–O2 N2–Cd1–O2 O1–Cd1–N1 O2–Cd1–N1 O1–Cd1–O4 O2–Cd1–O4 O1A–Cd1–O3 N2–Cd1–O3 N1–Cd1–O3

2.333(3) 2.448(2) 142.49(7) 87.42(10) 123.96(10) 137.85(8) 79.60(10) 83.03(8) 133.67(9) 87.13(10) 96.41(10) 110.67(8)

A. Majumder et al. / Polyhedron 25 (2006) 1753–1762

best described as a distorted monocapped octahedron with a CdO5N2 chromophore. Its equatorial plane is described by two nitrogen atoms (pyridine nitrogen, N1 and imine nitrogen, N2) from the Schiff base and two bridging phenolic oxygen atoms O1 and O1A, with r.m.s. deviation of ˚ . The axial positions are occupied by one oxygen 0.042 A atom (O3) of acetate group and a water molecule (O4) with a trans angle of 163.88(7) (O3–Cd1–O4) which is consistent with related systems [30,31]. The capping atom O2 ˚ ) deviates by 1.997 A ˚ from the plane (Cd1–O2, 2.348(3) A defined by the atoms N1, N2, O1, O1A. The Cd atom sits ˚ off this plane in the direction of the acetate group. 0.274 A The acetate group is inclined at 91.7 to the Cd2O2 unit which is constrained by symmetry to be planar. The Cd–O distances in this unit are asymmetric (Cd1– ˚ and Cd1–O1A 2.245(2) A ˚ ). The Cd1–N O1, 2.324(2) A ˚ ˚ (2.333(3) A and 2.414(3) A) distances are in accord with those values previously reported for seven- and eightcoordinated cadmium(II) complexes [29], although they are slightly longer than those distances normally found in four-, five- and six-coordinated cadmium(II) complexes. Such lengthening may be explained on the basis of the increased coordination number of cadmium [31]. The axial Cd1–O distances (Cd1–O4 and Cd1–O3 are 2.448(2) and ˚ , respectively) are significantly longer than the 2.599(3) A equatorial bond distances. Cd–O2 and Cd–O3 bond lengths lie within the range of Cd–Ocarboxylate bond distances ˚ ) reported for Cd(II)–Ocarboxylate coor(2.209(2)–2.879(2) A dination polymers [32]. The C7–N2 (imino group, ˚ ) is indicative of a C–N double bond. The sum 1.273(3) A of the angles at the phenoxide oxygen is almost exactly 360 (solid angle at O1 is 359.83) indicating no pyramidal oxygen distortion whilst the Cd1–O–Cd1A angle of 106.68(9) is very similar to the value observed for diphenoxo-bridged cadmium(II) complexes [33]. The long ˚ indicates that there is no Cd  Cd separation of 3.666 A interaction between the two metal atoms. There are significant intramolecular hydrogen bonds shown in Fig. 6 involving the axial acetate oxygen atom (O3) and the water ligand (Table 7). The metal complexes are linked as a ladder through intermolecular hydrogen bonding (Fig. 7) interactions. These give rise to columns parallel to the a axis. These columns are linked via C– H  O hydrogen bonds involving the other acetate O atom (O2) and a pyridyl hydrogen (H11). The shortest intermo-

Table 7 ˚ and ) Hydrogen bonds for 3 (A D–H  A

d(D–H)

d(H  A)

d(D  A)

\(DHA)

O4–H4A  O3#2 O4–H4B  O3#1 C7–H7  O4 C11–H11  O2

0.87 0.85 0.93 0.93

1.94 2.16 2.59 2.54

2.783(3) 2.982(3) 3.441(4) 3.436(4)

163.6 163.1 152.00 163.00

Symmetry transformations used to generate equivalent atoms: #1: x + 1, y + 1, z + 1; #2: x + 1, y, z; #3: x + 1, y + 1, z; #4: x + 1, y + 1, z + 1; #5: x + 1, y, z.

1759

Fig. 7. An illustration of the infinite one-dimensional chain through the O–H  O (acetate–water) hydrogen bonding interactions found in 3.

˚ between the columns lecular Cd  Cd distance is 7.377 A ˚ within the columns. and 7.438 A 3.2. Infrared spectra The infrared spectra of the three complexes 1, 2 and 3 are very much consistent with the structural data presented in this paper. Peaks at 3650–3300 cm1 for 1 and 3500– 3400 cm1 for 3 are attributable to O–H stretching vibrations of solvated or coordinated water molecules and indicate the presence of hydrogen bonding. The sharp band due to the phenolic OH groups appears at about 3220 cm1 [34] in complex 1. Several sharp weak peaks observed for the complexes in the range 3180–2855 cm1 likely to be due to aromatic stretches [35]. The absence of m(N–H) at 3435 cm1 in the IR spectra of the metal complexes, suggests that the ligand loses one proton on complexation, thus acting as a uninegative ligand [30a]. All the complexes display peaks at 1645–1617 cm1 which is assigned to the C@N stretch of the coordinated Schiff base ligands [36]. Strong well resolved-sharp bands in the regions 1605–1597, 1485– 1460, 1445–1420, 1055–1040 and 1015–1005 cm1 are assigned to the coordinated pyridine ring [35]. Bands appearing at 465, 355, 459, 365, 456 and 378 cm1 correspond to m(Ni–N), m(Ni–O), m(Zn–N), m(Zn–O), m(Cd–N), and m(Cd–O), respectively. Peaks at 766 cm1 (C–H stretching) and 586 cm1 (pyridyl out-of-ring deformation) are also observed for the complexes. The m(C@O), m(C–O) modes are present as two very strong bands at about 1460, 1250 cm1, respectively, for complex 2. The asymmetric and symmetric stretching vibrations of the acetate groups appear at 1597 and 1448 cm1, respectively, for complex 2. The difference between masym(COO) and msym(COO) (Dm = 149 cm1), which is smaller than 164 cm1 observed

A. Majumder et al. / Polyhedron 25 (2006) 1753–1762

in ionic acetate, reflects the bidentate bridging co-ordination mode [30a]. The characteristic strong bands of carboxylate groups appeared at 1560 (for asymmetric stretching) and 1416 cm1 (for symmetric stretching) for complex 3 [35] having the separation value (Dm = 144 cm1) between masym(COO) and msym(COO) bands suggests the presence of chelating acetate group linked with the metal centre for complex 3 [25]. Broad intense bands at ca. 1100 cm1 due to ClO4  shows no splitting, indicating the absence of coordinated ClO4  in 1 [33,37]. 3.3. Electronic spectra The electronic spectrum of 1 was recorded in methanol and displayed three strong absorption bands at about 245, 382 and 640 nm, which are assigned to the spin allowed transitions 3A2g ! 3T2g, 3A2g ! 3T1g, 3A2g ! 3T2g(P), respectively [38]. The last transition appeared as shoulder [38]. The cadmium and zinc complexes show only the charge-transfer transitions which can be assigned to charge transfer from the ligand to the metal and vice versa. The transitions of very strong intensity at 367–394 nm and 266– 269 nm have been attributed to the charge transfer from the pyridine and imino nitrogens to the metal centres [36]. 3.4. Electrochemical study The cyclic voltametric study of complex 1 was performed using methanol as solvent and tetrabutylammonium perchlorate as supporting eletrolyte at a scan rate of 50 mV s1. One irreversible reductive response at 0.68 V versus SCE was tentatively assigned to the reduction of coordinated ligand. One irreversible oxidation response was found at 0.82 V versus SCE on the positive side of SCE which may be attributable to Ni(II)–Ni(III) oxidation. 3.5. Fluorescence spectral study Few reports have appeared so far on the prospective use of fluorescence characteristics on transition metal complexation of Schiff base ligands. In this work, steady-state fluorescence studies have been employed as independent evidence of complexation between the ligand and the metal ions. It is important to mention here that among the three complexes, 1 is nonfluorescent and the other two have characteristic fluorescence emissions. The fluorescence spectra of the ligand and the complexes 2 and 3 in MeOH are shown in Fig. 8. Both the complexes show broad emission bands indicating charge transfer nature of the transitions. The fluorescence quantum yields of both the complexes are found to be considerably higher than that of the ligand itself. Significant differences in positions of emission maximum and fluorescence quantum yields of 2 and 3 from those of the ligand establish the complexation process. The two photophysical parameters of the ligand, complexes 2 and 3 are given in Table 8.

Fluorescence intensity (arb.unit)

1760

550

600 650 Wavelength (nm)

700

750

Fig. 8. Fluorescence spectra of ligand (——), 2 (h) and 3 (s) in MeOH (kexc = 500 nm). Probe concentrations are different around 105 M.

Table 8 Photophysical parameters of ligand, 2 and 3 in MeOH System

Emission maximum (nm)

Fluorescence quantum yield

Ligand 2 3

560 635 630

2 · 104 1 · 103 1 · 103

Quenching of fluorescence of a ligand by transition metal ions during complexation is a rather common phenomenon which is explained by processes such as magnetic perturbation, redox-activity, electronic energy transfer, etc. [39,40]. Enhancement of fluorescence through complexation is, however, of much interest as it opens up the opportunity for photochemical applications of these complexes [41,42]. Factors like a simple binding of ligand to the d10 metal ions [42], an increased rigidity in structure of the complexes [43], a restriction in the photoinduced electron transfer (PET) [41,44], etc. are assigned to the increase in the photoluminescence. In the present case the first two factors seem to be responsible for the enhanced fluorescence. 3.6. Magnetic study The effective magnetic moment at 20 C for complex 1 is 2.77 BM, which are nearer to the spin-only value of nickel(II) ion relative to mercury(tetrathiocyanato)cobaltate as the standard. 4. Conclusion The Schiff base ligand HL reacts with nickel(II), zinc(II) and cadmium(II) to form three types of complexes. On reaction with nickel(II) ions, HL forms six-coordinate octahedral complex with 1:2 metal:ligand stoichiometry. In the case of 2, the two terminal zinc ions are coordinated with

A. Majumder et al. / Polyhedron 25 (2006) 1753–1762

the tridentate Schiff base ligand along with the bridging bidentate acetate groups to form an intermediate trigonal bipyramidal-square pyramidal structure while the central zinc ion resides on a centre of symmetry and is surrounded by four oxygen atoms from four bridging bidentate acetate and two phenolic oxygen atoms. Cadmium(II) forms a dimeric 1:1 seven-coordinate complex where the deprotonated phenolic oxygen atoms from the Schiff base ligand bridge the two metal centres. The difference in preferred coordination number between nickel (mainly 4 or 6), zinc (4–6) and cadmium (2–8) inspires significant differences in their respective structures. Additionally, the metal coordination geometry is also varied, such that 1 possesses regular octahedral geometry whilst 3 contains the less common monocapped octahedron. However, for 2, regular octahedral exists along side intermediate trigonal bipyramidalsquare pyramidal coordination geometry. Furthermore, 2 and 3 possess an intense fluorescence property at room temperature, which is not observed for 1. It is suggested that complexes 2 and 3 exhibit potential applications as photoactive materials. 5. Supplementary material Crystallographic data for the structural analysis have been deposited with the Cambridge Crystallographic Data Centre, CCDC Nos. 234749, 265683 and 234750 for 1, 2 and 3, respectively. Copies of this information may be obtained free of charge from The Director, 12 Union Road, Cambridge CB2 1EZ, UK (fax: +44 1223 336 033; e-mail: [email protected] or www: http://www.ccdc. cam.ac.uk). Acknowledgements Dr. Bappaditya Bag, Department of Chemistry, Jadavpur University, Kolkata 700 032, India, is gratefully acknowledged for his valuable suggestion in the field of the synthesis of the Schiff base ligand. We thank Dr. Alex Slavin, St. Andrews University for data collection on 1. Our thanks is extended to Prof. L. Dahlenburg, Institut fu¨r Anorganische Chemie, Universita¨t Erlangen-Nu¨nberg, Egerland Strasse-1, 91058, Erlangen, Germany for the structural study of 3. References [1] Q. Shi, L. Xu, J. Ji, Y. Li, R. Wang, Z. Zhou, R. Cao, M. Hong, A.S.C. Chan, Inorg. Chem. Commun. 7 (2004) 1254. [2] (a) Z.-L. You, H.-L. Zhu, W.-S. Liu, Z. Anorg. Allg. Chem. 630 (2004) 1617; (b) Z.-L. You, H.-L. Zhu, Z. Anorg. Allg. Chem. 630 (2004) 2754. [3] A. Golcu, M. Tumer, H. Demirelli, R.A. Wheatley, Inorg. Chim. Acta 358 (2005) 1785. [4] (a) R. Ziessel, Coord. Chem. Rev. 216–217 (2001) 195; (b) M. Albrecht, Chem. Rev. 101 (2001) 3457. [5] C.R. Choudhury, S.K. Dey, N. Mondal, S. Mitra, S.O.G. Mahalli, K.M.A. Malik, J. Chem. Crystallogr. 31 (2002) 57.

1761

[6] (a) A.F. Kolodziej, Prog. Inorg. Chem. 41 (1994) 493; (b) D.X. West, H. Gebremedhin, R.J. Butcher, J.P. Jasinski, A.E. Liberta, Polyhedron 12 (1993) 2489; (c) R.K. Parashar, R.C. Sharma, A. Kumar, G. Mohan, Inorg. Chim. Acta 151 (1988) 201; (d) J.R. Lancaster (Ed.), The Bioinorganic Chemistry of Nickel, VCH, New York, 1988. [7] S. Chandra, X. Sangeetika, Spectrochim. Acta A 60 (2004) 147. [8] (a) E. Fujita, B.S. Brunschwig, T. Ogata, S. Yanagida, Coord. Chem. Rev. 132 (1994) 195; (b) E. Kimura, S. Wada, M. Shiyonoya, Y. Okazaki, Inorg. Chem. 33 (1994) 770; (c) B. De Clercq, F. Verpoort, Macromolecules 35 (2002) 8943; (d) T. Opstal, F. Verpoort, Angew. Chem. Int. Ed. 42 (2003) 2876; (e) B. De Clercq, F. Lefebvre, F. Verpoort, Appl. Catal. A 247 (2003) 345. [9] S.L. Lambert, C.L. Spiro, R.R. Gagne, D.N. Hendriekson, Inorg. Chem. 21 (1982) 68. [10] (a) D. Kessel, A.F.A. Sayyab, E.M.H. Jaffar, A.H.H.A. Lanil, Iraqi J. Sci. 22 (1981) 312; (b) E.M. Hodnett, W.J. Dunn, J. Med. Chem. 13 (1970) 768. [11] (a) E.M. Hodnett, W.J. Dunn, J. Med. Chem. 15 (1972) 339; (b) J. Chakraborty, R.N. Patel, J. Indian Chem. Soc. 73 (1996) 191. [12] W. Zishen, L. Zhiping, Y. Zhenhuan, Transition Met. Chem. 18 (1993) 291. [13] E. Szlyk, A. Surdykowski, M. Barwiolek, E. Larsen, Polyhedron 21 (2002) 2711. [14] (a) C.R. Choudhury, S.K. Dey, S. Mitra, N. Mondal, J. Ribas, K.M.A. Malik, Bull. Chem. Soc. Jpn. 77 (2004) 959; (b) R. Karmakar, C.R. Choudhury, G. Bravic, J.-P. Sutter, S. Mitra, Polyhedron 23 (2004) 949; (c) S.K. Dey, N. Mondal, M.S. El Fallah, R. Vicente, A. Escuer, X. Solans, M.F. Bardia, T. Matsushita, V. Gramlich, S. Mitra, Inorg. Chem. 43 (2004) 2427; (d) P. Talukder, A. Datta, S. Mitra, G. Rosair, M.S. El Fallah, J. Ribas, Dalton Trans. (2004) 4161. [15] (a) C.R. Choudhury, S.K. Dey, R. Karmakar, C. De Wu, C.-Z. Lu, M.S. El Fallah, S. Mitra, New J. Chem. (2003) 1360; (b) M.K. Saha, D.K. Dey, B. Samanta, A.J. Edwards, W. Clegg, S. Mitra, Dalton Trans. (2003) 488; (c) N. Mondal, M.K. Saha, B. Bag, S. Mitra, G. Rosair, M.S. El Fallah, Polyhedron 20 (2001) 579; (d) N. Mondal, S. Mitra, V. Gramlich, S.O.G. Mahalli, K.M.A. Malik, Polyhedron 20 (2001) 135. [16] J.R. Morrow, K.A. Kolasa, Inorg. Chim. Acta 195 (1992) 245. [17] A. Erxleben, J. Hermann, J. Chem. Soc., Dalton Trans. (2000) 569. [18] (a) N.H.C. Cooke, R.L. Viavattene, R. Eksteen, W.S. Wong, G. Davies, B.L. Karger, J. Chromatogr. 149 (1978) 391; (b) J.N. Le Page, W. Lindner, G. Davies, D.E. Seitz, B.L. Karger, Anal. Chem. 51 (1979) 433; (c) W. Lindner, Naturwissenschaften 67 (1980) 354. [19] M.F. Broglia, M.L. Gomez, S.G. Bertolotti, H.A. Montejano, C.M. Previtali, J. Photochem. Photobiol. A 173 (2005) 115. [20] G.M. Sheldrick, SHELXTL version 5.1: Program for the Solution and Refinement of Crystal Structures, Bruker AXS P4, Inc., Madison, WI, USA, 1999. [21] N. Walker, D. Stuart, Acta Crystallogr., Sect. A 39 (1983) 158. [22] L.J. Farrugia, J. Appl. Crystallogr. 32 (1999) 837. [23] A. Altomare, M.C. Burla, M. Camalli, G.L. Cascarano, C. Giacovazzo, A. Guagliardi, A.G.G. Moliterni, G. Polidori, R. Spagna, J. Appl. Crystallogr. 32 (1999) 115. [24] G.M. Sheldrick, SHELXL-97 (Release 97-2), Universita¨t Go¨ttingen, Go¨ttingen, Germany, 1998. [25] L.J. Farrugia, J. Appl. Crystallogr. 30 (1997) 565. [26] H. Asada, K. Hayashi, S. Negoro, M. Fujiwara, T. Matsushita, Inorg. Chem. Commun. 6 (2003) 193.

1762

A. Majumder et al. / Polyhedron 25 (2006) 1753–1762

[27] A.W. Addison, T.N. Rao, J. Reedijk, J. Van Rijn, G.C. Verschoor, J. Chem. Soc., Dalton Trans. (1984) 1349. [28] (a) M. Gembicky´, P. Baran, R. Bocˇa, H. Fuess, I. Svoboda, M. Valko, Inorg. Chim. Acta 305 (2000) 75; (b) S.R. Korupoju, N. Mangayarkarasi, S. Ameerunisha, E.J. Valente, P.S. Zacharias, J. Chem. Soc., Dalton Trans. (2000) 2845. [29] S. Sen, M.K. Saha, P. Kundu, S. Mitra, C. Kruger, J. Bruckmann, Inorg. Chim. Acta 288 (1999) 118. [30] (a) M.T.H. Tarafder, A. Kasbollah, K.A. Crouse, A.M. Ali, B.M. Yamin, H.-K. Fun, Polyhedron 20 (2001) 2363; (b) S. Gao, J.-W. Liu, L.-H. Huo, H. Zhao, J.-G. Zhao, Acta Crystallogr., Sect. E 60 (2004) m1875; (c) E. Yang, Z.-J. Li, J. Zhang, Y.-B. Chen, Y.-G. Yao, Acta Crystallogr., Sect. C 60 (2004) m457. [31] (a) A.F. Cameron, R.H. Nuttall, D.W. Taylor, J. Chem. Soc., Chem. Commun. (1971) 129; (b) A.F. Cameron, R.H. Nuttall, D.W. Taylor, J. Chem. Soc., Dalton Trans. (1972) 1608; (c) A.F. Cameron, R.H. Nuttall, D.W. Taylor, J. Chem. Soc., Dalton Trans. (1972) 1603. [32] W. Clegg, J.T. Cressey, A. McCamley, B.P. Straughan, Acta Crystallogr., Sect. C 51 (1995) 234.

[33] X.-Y. Xu, Q.-H. Luo, M.-C. Shen, X.-Y. Huang, Q.-J. Wu, Polyhedron 16 (1997) 915. [34] J. Sanmartin, M.R. Marmejo, A.M.G. Deibe, I.M. Rivas, A.R. Fernandez, J. Chem. Soc., Dalton Trans. (2000) 4174. [35] K. Nakamoto, Infrared and Raman Spectra of Inorganic and Coordination Compounds, Wiley, New York, 1970. [36] S.S. Tandon, S. Chander, L.K. Thompson, Inorg. Chim. Acta 300– 302 (2000) 683. [37] H. Keypour, S. Salehzadeh, R.G. Pritchard, R.V. Parish, Polyhedron 19 (2000) 1633. [38] A.B.P. Lever, Inorganic Electronic Spectroscopy, 2nd ed., Elsevier, Amsterdam, 1984. [39] A.W. Varnes, R.B. Dodson, E.L. Wehry, J. Am. Chem. Soc. 94 (1972) 94. [40] J.A. Kemlo, T.M. Sheperd, Chem. Phys. Lett. 47 (1977) 158. [41] N. Chattopadhyay, A. Mallick, S. Sengupta, J. Photochem. Photobiol. A 177 (2005) 55. [42] P. Purkayastha, G.K. Patra, D. Datta, N. Chattopadhyay, Indian J. Chem. A 39 (2000) 375. [43] L.-Z. Cai, W.-T. Chen, M.-S. Wang, G.-C. Guo, J.-S. Huang, Inorg. Chem. Commun. 7 (2004) 611. [44] G. Hennrich, H. Sonnenschein, U.R. Genger, J. Am. Chem. Soc. 121 (1999) 5073.

Synthesis, structures and fluorescence of nickel, zinc ...

... West Bengal 700 032, India. Received 19 August 2005; accepted 14 November 2005. Available online 18 January 2006 ... +91 33 2668 2017; fax: +91 33 2414 6414. E-mail address: ... comparison to the free ligand [11]. It has been shown ...

467KB Sizes 0 Downloads 222 Views

Recommend Documents

Synthesis of new zinc(II) phthalocyanine conjugates with ... - Arkivoc
Dec 14, 2016 - E-mail: [email protected]. Dedicated to Prof. Jacek Młochowski on the occasion of his 80th anniversary. Received 06-27-2016.

Effect of levels and frequency of zinc sulphate ...
Effect of levels and frequency of zinc sulphate application on yied and DTPA - Zn availability in maize-sunflower cropping system. Madras Agric. J. 92 (4-6) : 252-257 April-June 2005. R. SHANMUGASUNDARAM AND P.SAVITHRI. Department of Agronomy, Tamil

Best BOOKDownload Nickel and Dimed
turning received wisdom on its head with intelligence, clarity and verve.With some 12 ... laws of supply and demand have been reversed. Rental prices skyrocket ...

Benzil fluorescence and phosphorescence emissions
Oct 26, 2001 - temperatures in liquid solutions have been as-. 26 October ... preparation of the mother solution, spectropho- tometric .... B 101 (1997) 5353.

Hindustan Zinc -
... FY08E + FY07E cash of ~Rs 106 / shr). Financial summary. Y/E Mar. Sales. (Rs mn). PAT. (Rs mn). Consensus. EPS* (Rs ). EPS. (Rs ). Change. YoY (%). P/E.

Program of Fluorescence Microscopy Workshop 2015
E-‐mail : [email protected]. Tel./Fax : +216 70 527 882. Program of Fluorescence Microscopy Workshop 2015. Monday March 2nd, 2015. 8:30 – 16:00 Registration (Entrance Hall, research building). Conference (Venue: Theatre D). President of sessio

Multiphoton-Excited Fluorescence of Silicon ... - Semantic Scholar
May 15, 2017 - to be an enabling technology for minimally invasive in vivo molecular ... within the second near-IR transmission window of tissue .... for use in the IR. The pulse width of the excitation laser beam is determined as a function of wavel

WQ_COSPBO04a-Gamble-Gulch-Cadmium-and-Zinc ...
WQ_COSPBO04a-Gamble-Gulch-Cadmium-and-Zinc-TMDL-w-Cover-Letter.pdf. WQ_COSPBO04a-Gamble-Gulch-Cadmium-and-Zinc-TMDL-w-Cover-Letter.

short communication high concentrations of zinc in the fangs and ...
insects from several orders using an electron microprobe. They determined the zinc concentration, averaged over the whole mandible, with wet chemical ...

Zinc,Zirconium.pdf
ZN009600 Thickness: 6.0mm, Purity: 99.99+%, Size: 50 mm to 75 mm. Ermine Business Park, Huntingdon PE29 6WR England. Telephone 0800 731 4653 : Fax ...

Concentration dependence of fluorescence and lasing efficiency in Cr ...
In gain-switched operation, the highest output energy was obtained with the .... the pump beam was focused at the center of the Cr2+:ZnSe samples to a spot ...

Multiphoton-Excited Fluorescence of Silicon ... - Semantic Scholar
May 15, 2017 - for use in the IR. The pulse width of the excitation laser beam is determined as a function of wavelength using the autocorrelator, substituting a silicon photodiode for the. (Ga,As)P detector ... A spectrum at a location 2 μm distant

Nickel-And-Dimed-Acting-Edition.pdf
There was a problem previewing this document. Retrying... Download. Connect more apps... Try one of the apps below to open or edit this item. Nickel-And-Dimed-Acting-Edition.pdf. Nickel-And-Dimed-Acting-Edition.pdf. Open. Extract. Open with. Sign In.

SYNTHESIS AND CHARACTERIZATION OF ...
1 Faculty of Chemical Technology, Hanoi University of Technology. 2 Institute of .... their different degrees of ionization depending on pH values. Actually, the ...

Synthesis, spectral characteristics and electrochemistry of ... - Arkivoc
studied representatives of electron-injection/hole-blocking materials from this class is .... Here, the diagnostic peak comes from C2 and C5 carbon atoms of the.

Gold catalyzed synthesis of tetrahydropyrimidines and ... - Arkivoc
Dec 21, 2017 - or the replacement of hazardous organic solvents with environmentally benign solvents has received ..... Replacement of p-MeOC6H4 8c or t-Bu 8i by other hydrophobic groups such as o,p-. Me2 8d ..... Jones, W.; Krebs, A.; Mack, J.; Main

Synthesis and Decomposition of Processes in Organizations.
Edwin L. Cox School of Business, Southern Methodist University, Dallas, Texas ... Owen Graduate School of Management, Vanderbilt University, Nashville, ...

Synthesis and spectroscopic characterization of double ... - Arkivoc
Dec 4, 2016 - Such derivatives are used as reagents in organic synthesis and, due to their interest from the biological point of view, in the preparation of ...

Synthesis and antiviral activity of 4 - Arkivoc
Jun 25, 2017 - Chemical shifts are reported in ppm (δ) with reference to internal standard TMS. The signals ... with a Thermo Scientific, model Flash 1112EA apparatus and Eagar xperience software. ... (400 MHz, DMSO-d6, δ ppm): 0.92 (s, 6H), 1.38-

Synthesis and physicochemical properties of merocyanine ... - Arkivoc
Mar 30, 2017 - dyes find wide use in many areas of human activity: optoelectronics, photovoltaics, biology, and medicine. 2,15,16. Thermophotoresistors ...

Synthesis and spectroscopic characterization of double ... - Arkivoc
Dec 4, 2016 - with the elaboration at positions 2, 3 or 6, depending on the application ..... CHaHbO), 4.32 (dd, J 5.9, 11.7 Hz, 1H, CHaHbO), 4.80 (d, J2.0 Hz, ...