Your browser doesn't support javascript.
loading
Show: 20 | 50 | 100
Results 1 - 20 de 20
Filter
Add more filters










Publication year range
1.
Microbes Environ ; 38(2)2023.
Article in English | MEDLINE | ID: mdl-37164688

ABSTRACT

Cold-adapted or psychrotrophic fermentative anaerobic bacteria were isolated from rice field soil in a temperate area in Japan using anaerobic enrichment cultures incubated at 5°C. Most isolates were obligately anaerobic, spore-forming rods and affiliated with different lineages of the genus Clostridium based on 16S rRNA gene sequences. The growth temperature ranges and physiological properties of three representative clostridial isolates (C5S7, C5S11T, and C5S18) were examined. Strain C5S7 grew at 0°C, but not at 20°C, and was identified as Clostridium estertheticum, a psychrophile isolated from spoiled, vacuum-packed, chilled meat (blown pack spoilage, BPS). Strain C5S7 produced butyrate, n-butanol, and abundant gases (H2 and CO2) as major fermentation products from the carbohydrates utilized. Strain C5S11T, which was recently described as Clostridium gelidum sp. nov., possessed psychrotrophic properties and grew at temperatures between 0 and 25°C. Strain C5S11T was saccharolytic, decomposed polysaccharides, such as inulin, pectin, and xylan, and produced acetate, butyrate, and gases. Strain C5S18 also grew at 0°C and the optimum growth temperature was 15°C. Strain C5S18 did not ferment carbohydrates and grew in a manner that was dependent on proteinaceous substrates. This strain was identified as the psychrotolerant species, Clostridium tagluense, originally isolated from a permafrost sample. Collectively, the present results indicate that psychrotrophic anaerobic bacteria with different physiological properties actively degrade organic matter in rice field soil, even in midwinter, in a cooperative manner using different substrates. Furthermore, different psychrotrophic species of the genus Clostridium with the ability to cause BPS inhabit cultivated soil in Japan.


Subject(s)
Bacteria, Anaerobic , Oryza , Phylogeny , RNA, Ribosomal, 16S/genetics , Japan , Soil , Clostridium/genetics , Butyrates/metabolism , Carbohydrates , Gases/metabolism , DNA, Bacterial/chemistry
2.
Chemistry ; 17(38): 10708-15, 2011 Sep 12.
Article in English | MEDLINE | ID: mdl-22003514

ABSTRACT

Interchange between the nickel +2 and +3 oxidation states precisely controls the reversible rearrangement of the tris(2-pyridylthio)methanide (tptm) ligand in the organometallic nickel(II) complex [{Ni(µ-Br)-(tptm)}(2)] (2). Oxidation of 2 first gives the corresponding Ni(III) complex [{Ni(µ-Br)(tptm)}(2)][PF(6)](2) (4). However, in solution the tptm ligand in 4 slowly undergoes a rearrangement, in which the N and S atoms of one of the pyridylthiolate arms exchange Ni and C bonding partners, thereby resulting in an "N,S-confused" isomer of tptm in the product, [NiBr(bpttpm)]PF(6) (5; bpttpm= bis(2-pyridylthio)(2-thiopyridinium)-methyl). Reduction of 5 reverses this ligand rearrangement and 2 is reformed quantitatively. The individual steps involved in these unusual ligand rearrangements were investigated by a number of methods, including voltammetric analysis, and a mechanism for this process is proposed. X-ray crystal structure determinations of the key compounds 2, 4 and 5 have been obtained.

3.
Chem Commun (Camb) ; (11): 1314-6, 2008 Mar 21.
Article in English | MEDLINE | ID: mdl-18389118

ABSTRACT

Tris(2-pyridylthio)methane (tptmH) reacts with ZnCl(2) producing the Zn-C containing complex of [ZnCl(tptm)], whose cyclic voltammogram shows an irreversible oxidation peak at 0.2 V vs. E(0')(Fc(+/0)). DFT calculations suggested that 1e(-) oxidation should occur at the tptm ligand resulting in the cleavage of the Zn-C bond, leading to decomposition of the complex.

4.
J Phys Chem A ; 112(12): 2533-42, 2008 Mar 27.
Article in English | MEDLINE | ID: mdl-18288823

ABSTRACT

The photoinduced electron transfer of a series of meta- and para-linked triphenylamine-naphthalimide dyads, N-{3- and 4-[bis(4-R-substituted phenyl)amino]phenyl}-1,8-naphthalimide, 1m,p (R = H), 2m,p (R = Me), 3m,p (R = OMe), and 4m,p (R = NMe2) was investigated in toluene and DMF. The singlet charge-transfer (CT) states were observed in all cases. The decay rates were found to be faster in DMF (tau = 6.5 ps to 100 ps) than those in toluene (tau = 190 ps to 7 ns). The long-lived triplet CT states were observed in toluene for 3 (ca. 10% contribution, tau = 670 ns for 3m, 240 ns for 3p). No long-lived species were detected in DMF. The decay rates were somewhat faster in the para-isomers than in the meta-isomers in most cases. The photolysis of 5 (p-phenylene extended analogue of 3, R = OMe) gave a singlet CT state and a locally excited triplet state on the naphthalimide chromophore.

5.
Dalton Trans ; (30): 3330-4, 2007 Aug 14.
Article in English | MEDLINE | ID: mdl-17893780

ABSTRACT

Quinoline-based, tetradentate nitrogen ligands, N,N'-bis(2-quinolylmethyl)-N,N'-dialkyl-1,2-ethanediamine (alkyl = methyl, bqdmen; ethyl, bqdeen; isopropyl, bqdpen), have been investigated as the supporting ligands for the formation of bis(micro-oxo) dinuclear manganese complexes. Bis(micro-oxo)Mn(2)(iii,iii) complexes and were obtained for bqdmen and bqdeen, respectively, as evidenced by X-ray crystallography, whereas bqdpen did not afford any manganese complexes due to its steric bulk. Complexes and exhibit highly positive Mn(2)(iii,iii)/Mn(2)(iii,iv) and Mn(2)(iii,iv)/Mn(2)(iv,iv) redox couples relative to the corresponding pyridine-ligated (micro-O)(2)Mn(2)(iii,iii) complexes.


Subject(s)
Manganese/chemistry , Nitrogen/chemistry , Organometallic Compounds/chemistry , Quinolines/chemistry , Crystallography, X-Ray , Electrochemistry , Ligands , Molecular Structure , Organometallic Compounds/chemical synthesis , Oxidation-Reduction
8.
Dalton Trans ; (11): 1374-6, 2006 Mar 21.
Article in English | MEDLINE | ID: mdl-16518505

ABSTRACT

Thiacalix[3]pyridine (Py3S3) consists of pyridines and bridging sulfur atoms producing a stable octahedral mononuclear Rh(II) complex [Rh(II)(Py3S3)2]2+ showing mutual Jahn-Teller effect, a metal based reversible redox couple of Rh(III/II) at 0.02 V vs. SCE and a g(perpendicular) > g(||) relationship in EPR measurements.

9.
Inorg Chem ; 45(7): 3048-56, 2006 Apr 03.
Article in English | MEDLINE | ID: mdl-16562961

ABSTRACT

Two new dinuclear Ru(III) complexes containing naphthalene moieties, K[Ru2(dhpta)(mu-O2CCH2-1-naph)2] (1) and K[Ru2(dhpta)(mu-O2CCH2-2-naph)2] (2) (H5dhpta = 1,3-diamino-2-hydroxypropane-N,N,N',N'-tetraacetic acid, naph-1-CH2CO2H = 1-naphthylacetic acid, naph-2-CH2CO2H = 2-naphthylacetic acid), were synthesized. Complex 2 crystallized as an orthorhombic system having a space group of Pbca with unit cell parameters a = 10.6200(5) A, b = 20.270(1) A, c = 35.530(2) A, and Z = 8. EXAFS analysis of 1 and 2 in the solid states and in solution clarified that the dinuclear structures of 1 and 2 were kept in DMSO solutions. Variable-temperature magnetic susceptibility data indicated that the two Ru(III) centers are strongly antiferromagnetically coupled as shown by the large coupling constants, J = -581 cm(-1) (1) and -378 cm(-1) (2). In the cyclic voltammograms of 1 and 2, one oxidation peak and two reduction peaks which were assigned to the redox reaction of the ruthenium moieties were observed in DMF. The large conproportionation constants estimated from the reduction potentials of Ru(III)Ru(III) and Ru(III)Ru(II) indicated the great stability of the mixed-valent state. The mixed-valent species [Ru(III)Ru(II)(dhpta)(mu-O2CCH2-R)2](2-) (R = 1-naph (6) and R = 2-naph (7)) were prepared by controlled potential electrolysis of 1 and 2 in DMF. The electronic absorption spectra of 6 and 7 were similar to that of [Ru(III)Ru(II) (dhpta)(mu-O2CCH3)2](2-) which is a typical Class II type mixed-valent complex. The fluorescence decay of 1 and 2 indicated that there are two quenching processes which come from the excimer and monomer states. The short excimer lifetimes of 1 and 2 were ascribed to the energy transfer from the naphthyl moieties to the Ru centers. The different excimer ratio between 1 and 2 suggested that the excimer formation is affected by the conformation of the naphthyl moieties in the diruthenium(III) complexes.


Subject(s)
Naphthaleneacetic Acids/chemistry , Organometallic Compounds/chemistry , Ruthenium/chemistry , Electrochemistry , Ligands , Magnetic Resonance Spectroscopy , Magnetics , Models, Molecular , Molecular Conformation , Organometallic Compounds/chemical synthesis , Photochemistry , Sensitivity and Specificity , Stereoisomerism , Temperature
10.
Bioorg Med Chem Lett ; 16(12): 3319-22, 2006 Jun 15.
Article in English | MEDLINE | ID: mdl-16564168

ABSTRACT

Several open-chained analogues of UK-2A, a novel antifungal antibiotic isolated from Streptomyces sp. 517-02, were prepared for structure-activity studies. The in vitro antifungal activities of these compounds against Rhodotorula mucilaginosa IFO 0001 and the inhibition of uncoupler-stimulated respiration in bovine heart submitochondrial particles (SMP) were evaluated. Oxidative potentials were measured by cyclic voltammetry. An analogue prepared from dihexyl L-glutamate showed comparable inhibitory activity as UK-2A.


Subject(s)
Antifungal Agents/chemical synthesis , Antifungal Agents/pharmacology , Heart/drug effects , Lactones/chemistry , Streptomyces/chemistry , Animals , Antifungal Agents/chemistry , Cattle , Lactones/chemical synthesis , Lactones/pharmacology , Molecular Structure , Pyridines/chemical synthesis , Pyridines/chemistry , Pyridines/pharmacology , Streptomyces/classification , Structure-Activity Relationship
11.
J Biosci Bioeng ; 100(2): 207-11, 2005 Aug.
Article in English | MEDLINE | ID: mdl-16198266

ABSTRACT

Cross-linked poly-gamma-glutamic acid (C-L gamma-PGA) at 5 microg/ml flocculated bentonite suspension pretreated with polyaluminum chloride (PAC) at 2 microg/ml Al3+-PAC to a transparency of approximately 30% after 30 min and more than 90% after 4 h, while Al3+ concentration in the upper phase of the suspension decreased with incubation time. When pretreated with FeCl3 at 16 microg/ml Fe3+-FeCl3, similar results were obtained. In the case of Escherichia coli suspension, the combination of C-L gamma-PGA and FeCl3 demonstrated a more marked flocculating activity with a satisfactory transparency occurring after 30 min of treatment, accompanied by a decrease in residual Fe3+ concentration. In the above two suspensions pretreated with FeCl3, small visible floats appeared in the early stage of incubation. These floats were found to be due to the direct interaction between FeCl3 and C-L gamma-PGA, indicating the formation of a water-insoluble complex. After allowing the suspension to stand for a long time, elemental analysis and inductively coupled plasma spectroscopy of the precipitates produced suggested that not only the complex was formed due to the interaction between Fe3+ in FeCl3 and COO- in the C-L gamma-PGA molecule, but also Fe2O3 and Fe(OH)3 might be entrapped in this complex. This could be applied to scavenge metal ions including Fe3+ from polluted water.


Subject(s)
Bentonite/chemistry , Escherichia coli/chemistry , Ferric Compounds/chemistry , Polyglutamic Acid/chemistry , Chlorides , Flocculation
12.
Dalton Trans ; (19): 3179-86, 2005 Oct 07.
Article in English | MEDLINE | ID: mdl-16172643

ABSTRACT

A complete series of copper(ii) halide complexes [CuX(tptm)](X = F (), Cl (), Br (), I (); tptm = tris(2-pyridylthio)methyl) with a novel Cu(II)-C(sp(3)) bond has been prepared by the reactions of [Cu(tptm)(CH(3)CN)]PF(6)(.PF(6)) with corresponding halide sources of KF or n-Bu(4)NX (X = Cl, Br, I), and the trigonal bipyramidal structures have been confirmed by X-ray crystallography and/or EPR spectroscopy. The iodide complex easily liberates the iodide anion in acetonitrile forming the acetonitrile complex as a result. The EPR spectra of the complexes showed several superhyperfine structures that strongly indicated the presence of spin density on the halide ligands through the Cu-X bond. The results of DFT calculations essentially matched with the X-ray crystallographic and the EPR spectroscopic results. Cyclic voltammetry revealed a quasi-reversible reduction wave for Cu(II)/Cu(I) indicating a trigonal pyramidal coordination for Cu(I) states. A coincidence of the redox potential for all [CuX(tptm)](0/+) processes indicates that the main oxidation site in each complex is the tptm ligand.

13.
Inorg Chem ; 43(23): 7301-7, 2004 Nov 15.
Article in English | MEDLINE | ID: mdl-15530079

ABSTRACT

New proton and electron donors, M(II)(HL)(2) (M = Ni, Pd, Pt; L = 5,6-diethylpyradzinedithiolate), as well as a proton and electron acceptor, Pt(IV)(L)(2), were prepared and characterized. The pH-dependent cyclic voltammetry of the M(II)(HL)(2) complexes revealed a favorable Gibbs free energy (K(com) > 1) for the proton and electron transfer reactions from M(II)(HL)(2) to M(IV)(L)(2); i.e., the equilibrium for the following reaction lies to the right: M(II)(HL)(2) + M(IV)(L)(2) <==>2M(III)(HL)(L).

14.
Dalton Trans ; (20): 3283-7, 2004 Oct 21.
Article in English | MEDLINE | ID: mdl-15483712

ABSTRACT

The novel (mu-alkoxo)bis(mu-carboxylato)diruthenium complex K[Ru(2)(dhpta)(mu-O(2)C-p-ZnTPP)(2)] 3 was prepared by simple ligand substitution reaction. Strong antiferromagnetic interaction between two Ru(III) ions of 3 was observed with a coupling constant of -425 approximately -404 cm(-1). The cyclic voltammogram of 3 can be explained in terms of superposition of those of ZnTPP-p-CO(2)H and K[Ru(2)(dhpta)(mu-O(2)CPh)(2)] 2, indicating no significant electrochemical interaction. The large conproportionation constant estimated from the reduction potentials for Ru(III)Ru(III) and Ru(II)Ru(III) indicates great stability of the mixed-valence state. The mixed-valence species [Ru(II)Ru(III)(dhpta)(mu-O(2)C-p-ZnTPP)(2)](2-) 4 was prepared by controlled potential electrolysis. The electronic absorption spectrum of 4 was quite similar to that of [Ru(II)Ru(III)(dhpta)(mu-O(2)CCH(3))(2)](2-) which is a typical Class II complex. The fluorescence from the S(2) state of the ZnTPP unit of 3 was significantly (78%) quenched. The electron transfer from the ZnTPP unit to Ru(III) ions in 3 is a plausible mechanism, even though energy transfer could not be ruled out completely. The free energy change for electron transfer, Delta G(CS), was estimated to be ca.-1.1 eV, which is similar to typical values for the reorganization energy lambda in polar solvents. Hence, the electron transfer scheme is situated almost at the top of the Marcus parabola, enabling ultrafast electron transfer.

15.
Org Lett ; 6(20): 3493-6, 2004 Sep 30.
Article in English | MEDLINE | ID: mdl-15387531

ABSTRACT

[structure: see text] A trimeric phenothiazine and its radical cation were prepared, and their structures were elucidated. In contrast to a largely twisted structure in the neutral species, the radical cation had a unique structure deformation that allowed charge-transfer-type conjugation from the outer phenothiazine rings to the central phenothiazine radical cation.

16.
J Inorg Biochem ; 98(1): 105-12, 2004 Jan.
Article in English | MEDLINE | ID: mdl-14659639

ABSTRACT

Two chargeless VO(IV) complexes with 3-hydroxypyridine-2-carboxylic acid (H2hpic), [VO(Hhpic-O,O)(Hhpic-O,N)(H2O)].3H2O (1) and the cyclic tetramer [(VO)4(mu-(hpic-O,O',N))4(H2O)4].8H3O (2), have been synthesized and characterized by elemental analysis, mass, infrared, electronic absorption, electron spin resonance (ESR) spectroscopies, and X-ray crystallography. Their coordination structures are similar to each other (and 1 is readily transformed into 2), but are quite different from that of bis(pyridine-2-carboxylato)oxovanadium(IV). The magnetic susceptibility of 2 indicates the presence of a weak ferromagnetic intramolecular interaction between the V atoms at low temperature, in addition to a weak antiferromagnetic intermolecular interaction. The ESR signal of 2 was broad, while 1 showed an eight-line hyperfine splitting pattern due to coupling of the unpaired electron with the 51V nucleus (I=7/2). The ESR spectrum and cyclic voltammogram of 2 clearly show that the cyclic tetramer remains intact in solution. The insulinomimetic activity of 1 and 2 was evaluated by means of in vitro measurements of the inhibition of free fatty acid release from epinephrine-treated isolated rat adipocytes. While 1 exerted higher insulinomimetic activity than VOSO4, the activity of 2 was significantly lower than that of VOSO4. Hence 2 appears to retain its cyclic structure during the in vitro test. These results indicate that the rational ligand design for VO complexes might be a promising approach to obtain superior insulinomimetic activity.


Subject(s)
Hypoglycemic Agents/chemistry , Hypoglycemic Agents/pharmacology , Picolinic Acids/chemistry , Picolinic Acids/pharmacology , Vanadates/chemistry , Vanadates/pharmacology , Adipocytes/drug effects , Adipocytes/metabolism , Animals , Crystallography, X-Ray , Electrochemistry/methods , Electron Spin Resonance Spectroscopy , Epinephrine/pharmacology , Fatty Acids, Nonesterified/metabolism , Hypoglycemic Agents/chemical synthesis , Inhibitory Concentration 50 , Insulin/pharmacology , Male , Molecular Structure , Organometallic Compounds/chemistry , Organometallic Compounds/pharmacology , Oxidation-Reduction , Picolinic Acids/chemical synthesis , Rats , Rats, Wistar , Vanadates/chemical synthesis
17.
J Biol Inorg Chem ; 8(8): 893-906, 2003 Nov.
Article in English | MEDLINE | ID: mdl-14551811

ABSTRACT

Structures, chemical properties, and in vitro insulinomimetic activities of new vanadyl [oxovanadium(IV), VO(2+)] complexes with five tripodal ligands containing an imidazole functionality were examined. The ligands, N-(carboxymethyl)- N-(4-imidazolylmethyl)amino acids, contain glycine, ( S)- and ( R)-alanine, and ( S)- and ( R)-leucine residues. The molecular structures of the latter four alanine- and leucine-containing complexes were determined by X-ray analysis. The coordination geometry around each vanadium center was octahedral, where an imino nitrogen occupied the apical site and two carboxylate oxygens, an imidazole nitrogen, and a water molecule coordinated in the equatorial plane. The spectroscopic properties of the complexes were characterized by means of IR, electronic absorption, and CD spectra. Acid dissociation constants (p K(a)) and protonation sites of the ligands were determined by a combination of potentiometric titrations and (1)H NMR spectra. The potentiometric study demonstrated that stability constants (log beta) were not so different among the present complexes (14.0-14.9) and a species of molecular complex with a 1:1 metal:ligand ratio existed predominantly at physiological pH 7.4. EPR parameters indicated that the species at pH 7.4 had an octahedral structure similar to the complex in the solid state. On the other hand, an EPR study in phosphate buffer (pH 7.4) suggested that inorganic phosphate coordinated to the vanadium center instead of the imidazole group in the presence of excess phosphate ion. Cyclic voltammograms in the phosphate buffer showed chemically reversible oxidation waves, whereas irreversible oxidation waves were observed in non-coordinating HEPES buffer. Moreover, the oxidation potential of each complex in phosphate buffer was more positive than that in HEPES buffer. Partition coefficients of the present complexes in a n-octanol/saline system were very low, probably due to hydrophilicity of the imidazole group. The in vitro insulinomimetic activities were estimated on the basis of the ability of the complexes to inhibit epinephrine-stimulated free fatty acid release from isolated rat adipocytes. The achiral glycine-derivative complex exhibited the highest insulinomimetic activity, which was higher than that of VOSO(4) as a positive control. Putting our previous observations together, it was found that the vanadyl complexes with tetradentate amino acid derivatives having no alkyl side chain tend to have high in vitro insulinomimetic activity.


Subject(s)
Amino Acids/chemistry , Imidazoles/chemistry , Insulin/chemistry , Vanadates/chemistry , Alanine/chemistry , Alanine/metabolism , Amino Acids/metabolism , Animals , Glycine/chemistry , Glycine/metabolism , Imidazoles/metabolism , Insulin/metabolism , Leucine/chemistry , Leucine/metabolism , Ligands , Male , Molecular Structure , Rats , Rats, Wistar , Solubility/drug effects , Vanadates/metabolism
18.
Inorg Chem ; 42(10): 3187-93, 2003 May 19.
Article in English | MEDLINE | ID: mdl-12739958

ABSTRACT

Rhodium(III) porphyrin complexes, [Rh(4-PyT(3)P)Cl](4) (1) and [Rh(2-PytB(3)P)Cl](2) (2) (4-PyT(3)P = 5-(4-pyridyl)-10,15,20-tritolylporphyrinato dianion, 2-PytB(3)P = 5-(2-pyridyl)-10,15,20-tri(4-tert-butyl)phenylporphyrinato dianion), were self-assembled and characterized by (1)H nuclear magnetic resonance spectroscopy, infrared spectroscopy, and electron spray ionization-mass spectroscopy methods. The spectroscopic results certified that the rhodium porphyrin complexes 1 and 2 have a cyclic tetrameric structure and a cofacial dimeric structure, respectively. The X-ray structure analysis of 1 confirmed the cyclic structure of the complex. The Soret bands of both oligomers were significantly broadened by excitonic interactions between the porphyrin units, compared to those observed for a corresponding analogue of Rh(TTP)(Py)Cl (TTP = 5,10,15,20-tetratolylporphyrinato dianion, Py = pyridine). Stepwise oxidation of the porphyrin rings in the oligomers was observed by cyclic voltammetry. The oligomers 1 and 2 are very stable in solution, and they slowly undergo reactions with pyridine to give corresponding monomer complexes only at high temperatures (approximately 80 degrees C).

19.
Inorg Chem ; 37(19): 4986-4995, 1998 Sep 21.
Article in English | MEDLINE | ID: mdl-11670666

ABSTRACT

A series of cofacially arranged ruthenium(II) porphyrin dimers 1-5 having a variety of axial ligands such as CO, pyridine, and 4-cyanopyridine, were synthesized. Porphyrin tetramers, 6 and 7, which have pyridylporphyrin ligands at the axial positions of the parent cofacial ruthenium(II) dimers, were also prepared. These porphyrin dimers and tetramers were characterized by (1)H NMR spectroscopy, ESI (electrospray ionization)-mass spectroscopy, and elemental analysis. The ruthenium porphyrin dimers and tetramers exhibited characteristic electrochemical and spectroscopic properties caused by interactions between the porphyrin subunits. Stepwise oxidations of the porphyrin rings or the ruthenium ions in the cofacial dimer skeltons were observed in the cyclic voltammograms. The potential differences (DeltaE degrees ' mV) of the oxidation steps were larger than 260 mV for all the porphyrin oligomers. The Soret bands of the cofacial dimers were significantly broadened by excitonic interactions between the two porphyrin subunits. Furthermore, the mixed-valence states of 3-7 showed specific intervalence charge-transfer (IT) bands between the Ru(II) and Ru(III) cores in the near-IR region at around 1500 nm.

20.
Inorg Chem ; 36(8): 1625-1635, 1997 Apr 09.
Article in English | MEDLINE | ID: mdl-11669753

ABSTRACT

A series of ruthenium(II) porphyrin dimers and trimers (carbonyl dimers, 1-4; carbonyl trimers, 5-7, bis(pyridyl) trimers, 8-10), having axial or bridging porphyrin ligands, were synthesized and characterized by (1)H NMR and IR spectroscopy and mass spectrometry. An X-ray structural determination of Ru(II)(OEP)(CO)(H(2)PyP(3)P) (1) (OEP = octaethylporphyrinato dianion, H(2)PyP(3)P = 5-pyridyl-10,15,20-triphenylporphyrinato dianion) was carried out. The axial porphyrin ligand is coordinated to the ruthenium porphyrin subunit obliquely. The Ru-N(Py) bond length is 2.237(4) Å, and the angle between the ruthenium porphyrin macrocycle and the pyridyl ring is 63.23(35) degrees. Crystallographic data for 1 are as follows: chemical formula C(80)H(73)N(9)ORu.CH(2)Cl(2), triclinic, P&onemacr;, a = 14.954(5) Å, b = 25.792(5) Å, c = 10.124(3) Å, alpha = 90.21(2) degrees, beta = 108.43(2) degrees, gamma = 73.39(2) degrees, Z = 2, R(F) = 0.0674. (1)H NMR signals of 2,6- and 3,5-pyridyl protons of the axial ligand porphyrins of the oligomers 1-10 showed significant upfield shifts, indicating that the axial porphyrin subunits are coordinated to the ruthenium porphyrin subunits through the pyridyl group in solution. UV-vis spectra revealed the presence of excitonic interaction between two axial ligand porphyrin subunits in the trimers 8-10. The MLCT bands from the central ruthenium(II) ions to the octaethylporphyrin rings were observed around 450 nm in 8 and 9. Cyclic voltammograms of the carbonyl dimers and trimers showed no redox waves of the ruthenium(II) ions, because the ruthenium(II) oxidation state of these complexes was significantly stabilized by the coordination of the axial CO ligands. On the other hand, bis(pyridyl) trimers exhibit the Ru(III/II) waves in the region of -0.12 to +0.15 V vs Ag/Ag(+) reference electrode.

SELECTION OF CITATIONS
SEARCH DETAIL
...