首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
In this study, we address the effect of the cis-double bond in 1,2-dioleoyl-sn-glycero-3-phosphoethanolamide-N-[methoxy(polyethylene glycol)-2000, DOPE PEG2000 (DP), on the Langmuir monolayer of C18 fatty acids—namely, stearic acid (SA), oleic acid (L1), linoleic acid (L2), and linolenic acid (L3)—with the same head group but different degrees of saturation on their hydrocarbon chains. Negative values of Gibbs free energy of mixing (ΔG mix) were obtained throughout the investigated ranges of the unsaturated C18 fatty-acid (L1, L2 and L3) mixed systems, indicating that very strong attractions occurred between molecules in the monolayers. The bend and kink effects from the cis-double bond(s) in the hydrocarbon chain affected the membrane fluidity and molecular packing in the monolayers, which resulted in a greater interaction between unsaturated C18 fatty acids and DP. The most thermodynamically stable mole composition of unsaturated C18 fatty acids to DP was observed at 50:1; this ratio is suggested to be the best mole ratio and will be subsequently used to prepare DP–C18 fatty-acid nanoliposomes. The presence of cis-double bonds in both hydrocarbon chains of DOPE in DP also created an imperfection in the membrane structure of lipid-drug delivery systems, which is expected to enhance lipid-based systems for antibody conjugation and drug encapsulation.  相似文献   

2.
A new thiophene-functionalized benzimidazolium salt (2) has been prepared by reacting N-methylbenzimidazole with 2-bromomethylthiophene (1), which in turn was obtained by bromination of 2-thiophenemethanol with PBr3. Subsequent reaction of salt 2 with Pd(OAc)2 afforded the cis-configured bis(carbene) Pd(II) complex (cis-3), which in solution exists as an inseparable mixture of cis-anti and cis-syn-rotamers in a 3.5:1 ratio. All new compounds have been fully characterized by spectroscopic and spectrometric methods. A preliminary catalytic study shows that cis-3 is highly active in the Suzuki-Miyaura coupling of aryl bromides with phenylboronic acid in/on water as environmentally benign reaction media.  相似文献   

3.
1-Alkynyl-dimethyl(triorganophosphine)gold(III) complexes of the type cis-Me2(Ph3P)Au-CC-R with R = H, Me, Ph (1-3) have been prepared from the cis-Me2(Ph3P)AuX (X = Cl, I) complexes and lithium alkynyls. The crystal structures of 1 and 2 have been determined together with those of the reference compounds cis-Me2(Ph3P)AuX (X = Cl, I) and cis-Me2(Me3P)AuI. The molecules have a standard square planar geometry and are not associated into oligomers. Due to the different hybridization of the carbon orbitals, the Au-C(CR) bonds are found significantly shorter than the Au-CH3 bonds. Compounds 1-3 are stable colourless, crystalline solids at 20 °C but decompose on heating with selective (cis) reductive elimination of ethane and formation of the gold(I) alkynyls (Ph3P)Au-CC-R thus retaining the stronger gold-alkynyl bonds. Two complexes of this type have also been prepared by conventional routes from (R3P)AuX complexes and the crystal structures of (Me3P)Au-CC-Ph and [(p-Tol)3P]Au-CC-H have been determined. The former with the small Me3P ligand is associated into two different trimers via aurophilic bonding and further aggregated into chains via weak inter-trimer contacts, while the latter is a monomer owing to the steric bulk of the (p-Tol)3P ligand.  相似文献   

4.
Starting from previously reported cis-Ru(MeL)2Cl2, where MeL is 4,4,4′,4′-tetramethyl-2,2′-bisoxazoline, cis-Ru(MeL)2Br2 (1), cis-Ru(MeL)2I2 (2), cis-Ru(MeL)2(NCS)2 · H2O (3), cis-Ru(MeL)2(N3)2 (4) and cis-[Ru(MeL)2(MeCN)2](PF6)2 · (CH3)2CO (5) are synthesised. The X-ray crystal structures of complexes 1, 2, 3 and 5 have been determined. All the five new complexes have been characterized by FTIR, ESIMS and 1H NMR. In cyclic voltammetry in acetonitrile at a glassy carbon electrode, the complexes display a quasireversible Ru(II/III) couple in the range 0.32-1.71 V versus NHE. The Ru(II/III) potentials yield a satisfactorily linear correlation with Chatt’s ligand constants PL for the monodantate ligands. From the intercept and by comparing the known situation in Ru(2,2′-bipyridine)2L2, it is concluded that MeL, a non-aromatic diimine, is significantly more π-acidic than 2,2′-bipyridine.  相似文献   

5.
The reaction of [Ru(CO)2Cl2]n with bis(2-pyridylmethyl)amine (bpma) in refluxing ethanol followed by anion exchange yields two products: cis,fac-[Ru(bpma)(CO)2Cl]PF6 (1a, 71%) and trans,fac-[Ru(bpma)(CO)2Cl]PF6 (1b, 29%). Reaction of 1a with AgBF4 in acetone, followed by acetonitrile and then anion exchange gave cis,fac-[Ru(bpma)(CO)2(CH3CN)](PF6)2 (2a). In the same way, 1b afforded trans,fac-[Ru(bpma)(CO)2(CH3CN)](PF6)2 (2b). Reaction of depolymerized [Ru(CO)2Cl2]n with bpma in ethanol at room temperature afforded cis,cis-[Ru(η2-bpma)(CO)2Cl2] (3). In refluxing ethanol, 3 was converted to cis,fac-[Ru(bpma)(CO)2Cl]Cl (1a-Cl). Heating 3 in chlorobenzene afforded 1b-Cl, exclusively; heating 3 in ethylene glycol gave mainly 1a-Cl. Heating 1a-Cl in ethanol resulted in no isomerization, but heating in chlorobenzene gave a mixture of 3 and 1b-Cl. Anion exchange for PF6 with 1a-Cl and 1b-Cl afforded 1a and 1b, respectively, whereas anion exchange for BPh4 afforded 1a-BPh4. Compounds 1a, 1b, 2a and 3 have been structurally characterized.  相似文献   

6.
Condensation of 1,2,3,4,6-penta-O-acetyl-á-l-idopyranose (1) with phenol yielded phenyl 2,3,4,6-tetra-O-acetyl-α- (2) and α-l-idopyranoside (4). Deacetylation of 2 and 4 afforded phenyl α and β-l-idopyranosides (3 and 5), respectively, the structures of which were verified by periodate oxidation studies. A platinum-catalyzed oxidation of 3 and 5 produced the amorphous phenyl α- and β-L-idopyranosiduronic acids (9 and 11), respectively, which were isolated as the crystalline cyclohexylammonium salts. Phenyl β- and α-d-glucopyranosiduronic acids are apparent minor byproducts of the catalytic oxidations, resulting from an inversion at C-5. p-Nitrophenyl α-d-mannopyranosiduronic acid and p-nitrophenyl α- and β-d-galactopyranosiduronic acids are also described.  相似文献   

7.
The effect of fatty acids (FAs) (C12–C24) on the functioning of winter wheat (Triticum aestivum L.) mitochondria was studied. Such fatty acids as C12:0, C16:0, and C18:0 and unsaturated FAs, such as C18:1 (n-9 cis), C18:1 (n-12 cis), C18:2 (n-9, 12), (18:3, n-3), and C22:1 (n-9 cis) caused efficient uncoupling of oxidative phosphorylation in mitochondria, i.e., an increase in the nonphosphorylating respiration rate and a decrease in the respiratory control value. It was established that C16:0 had the strongest uncoupling effect among all saturated FAs, and C18:3, among unsaturated FAs. The uncoupling effect of saturated FAs is provided by the ADP/ATP-antiporter, while plant uncoupling proteins play an important role in the uncoupling effect of unsaturated FAs. In addition, unsaturated, as well as saturated FAs might serve as oxidative substrates for mitochondria. It was concluded that the role of FAs in energetic metabolism of winter wheat seedlings consisted of uncoupling of oxidative phosphorylation and of serving as substrates for oxidation.  相似文献   

8.

Background

Retinal dehydrogenases (RALDHs) catalyze the dehydrogenation of retinal into retinoic acids (RAs), which are required for embryogenesis and tissue differentiation. This study sought to determine the detailed kinetic properties of 2 mouse RALDHs, namely RALDH3 and 4, for retinal isomer substrates, to better define their specificities in RA isomer synthesis.

Methods

RALDH3 and 4 were expressed in Escherichia coli as His-tagged proteins and affinity-purified. Enzyme kinetics were performed with retinal isomer substrates. The enzymatic products were analyzed by high pressure liquid chromatography.

Results

RALDH3 oxidized all-trans retinal with high catalytic efficiency (Vmax/Km = 77.9) but did not show activity for either 9-cis or 13-cis retinal substrates. On the other hand, RALDH4 was inactive for all-trans retinal substrate, exhibited high activity for 9-cis retinal oxidation (Vmax/Km = 27.4), and oxidized 13-cis retinal with lower catalytic efficiency (Vmax/Km = 8.24). β-ionone, a potent inhibitor of RALDH4 activity, suppressed 9-cis and 13-cis retinal oxidation competitively with inhibition constants of 0.60 and 0.32, respectively, but had no effect on RALDH3 activity. The divalent cation MgCl2 activated 13-cis retinal oxidation by RALDH4 by 3-fold, did not significantly influence 9-cis retinal oxidation, and slightly activated RALDH3 activity.

Conclusions

These data extend the kinetic characterization of RALDH3 and 4, providing their specificities for retinal isomer substrates.

General significance

The kinetic characterization of RALDHs should give useful information in determining amino acid residues that are involved in the specificity for retinal isomers and on the role of these enzymes in the synthesis of RAs in specific tissues.  相似文献   

9.
Heterotrophically grown cell suspension cultures of soya (Glycine max L.) were incubated with two different mixed substrates consisting of positional isomers of either cis-[1-14C]octadecenoic acids (8 to 15) or trans-[1-14C]octadecenoic acids (8 to 16), each with known composition. With both substrates, about one-fourth of the radioactivity supplied was incorporated into the diacylglycerophosphocholines, while another one-fourth of the radioactivity was almost equally distributed between diacylglycerophos-phoethanolamines and triacylglycerols. All the positional isomers of cis-and trans-octadecenoic acids supplied to the cells were readily incorporated into various classes of glycerolipids. None of the octadecenoic acids was isomerized, elongated or desaturated during incubation. From the cis-octadecenoic acids, only the naturally occurring 9-isomer (oleic acid) was preferentially incorporated into position 2 of diacylglycerophosphocholines, diacylglycerophospho-ethanolamines, and triacyglycerols; all the other isomers exhibited a strong affinity for position 1 of the glycerophospholipids and positions 1 and 3 of the triacylglycerols. From the trans-octadecenoic acids, only the 9-isomer (elaidic acid) was preferentially incorporated into position 2 of diacylglycerophospho-cholines and triacylglycerols; all the other isomers preferred position 1 and positions 1 and 3, respectively, of these lipids. In diacylglycerophospho-ethanolamines, however, each of the trans-octadecenoic acids, including the 9-isomer, exhibited a strong affinity for position 1. Apparently, the enzymes involved in the incorporation of exogenous monounsaturated fatty acids into membrane lipids of plant cells can recognize the preferred substrate in a mixture of closely related isomers.  相似文献   

10.
The “amidate-hanging” Pt mononuclear complexes, which can easily bind a second metal ion with the non-coordinated oxygen atoms in the amidate moieties, have been synthesized and characterized by 1H NMR, MS, IR spectroscopy, and single crystal X-ray analysis. Five new complexes with various amidate ligands and co-ligands, cis-[Pt(PVM)2(en)] · 4H2O (1, PVM = pivaloamidate, en = ethylenediamine), cis-[Pt(PVM)2(NH2CH3)2] · H2O (2), cis-[Pt(PVM)2(NH2tBu)2] (3), cis-[Pt(TCM)2(NH3)2] (4, TCM = trichloroacetamidate), and cis-[Pt(BZM)2(NH3)2] (5, BZM = benzamidate), were successfully synthesized by direct base hydrolysis of the corresponding Pt nitrile complexes, cis-[Pt(NCR)2(Am)2]2+ (P1, P2, P3, and P5) (NCR = nitrile, Am = amine). These nitrile complexes were obtained by introducing nitriles into the Pt aqua complexes, cis-[Pt(OH2)2(Am)2](ClO4)2, whereas introduction of trichloronitrile into [Pt(OH2)2(NH3)2](ClO4)2 induced more facilitated water nucleophilic attack to afford [Pt(TCM)(NH(COH)CCl3)(NH3)2](ClO4) (P4). The base treatments of the precursor complexes (P1-5) lead to produce “amidate-hanging” Pt mononuclear complexes (1-5) without geometry isomerization. The 195Pt chemical shifts for 1-5 exhibit subtle differences of the Pt electron densities among them.  相似文献   

11.
We have examined the role of different solvents in the crystallisation process of cis-octahedral, diphenyltin(IV)-bis-cupferronato complex, Ph2Sn(cupf)2 (1), where . The Mössbauer spectra of frozen chloroform solution of 1 revealed the presence of cis and trans isomers. This cis-trans isomerisation was investigated by Mössbauer spectroscopy and the results inspired the synthesis of two new heptacoordinated derivatives: Ph2Sn(cupf)2(H2O) (2) and Ph2Sn(cupf)2(EtOH) · EtOH (3). In both compounds, the O-donor solvent molecules (H2O, EtOH) form novel Sn-O bonds with the Ph2Sn(IV) centre of 1, consequently the phenyl groups attached to tin undergo an intramolecular rearrangement. Compound 2 contains O-H ? O hydrogen bonded infinite chains. In compound 3, O-H ? O hydrogen-bonds and short O ? O contacts assemble the complexes and uncoordinated solvent molecules into dimeric supramolecules. These solvents have structure-determining roles at both molecular and supramolecular levels: at molecular level the coordination of solvent determines intramolecular rearrangement by changing the conformation of the parent unsolvated complex, whilst at supramolecular level they control the association of solvated molecules via hydrogen bonds.  相似文献   

12.
Using a phosphorus based Mannich condensation reaction the new pyridylphosphines {5-Ph2PCH2N(H)}C5H3(2-Cl)N (1-Cl) and {2-Ph2PCH2N(H)}C5H3(5-Br)N (1-Br) have been synthesised in good yields (60% and 88%, respectively) from Ph2PCH2OH and the appropriate aminopyridine. The ligands 1-Cl and 1-Br display variable coordination modes depending on the choice of late transition-metal complex used. Hence P-monodentate coordination has been observed for the mononuclear complexes AuCl(1-Cl) (2), AuCl(1-Br) (3), RuCl2(p-cymene)(1-Cl) (4), RuCl2(p-cymene)(1-Br) (5), RhCl2(Cp)(1-Cl) (6), RhCl2(Cp)(1-Br) (7), IrCl2(Cp)(1-Cl) (8), IrCl2(Cp)(1′-Cl) (8′), IrCl2(Cp)(1-Br) (9), cis-/trans-PdCl2(1-Cl)2 (10), cis-/trans-PdCl2(1-Br)2 (11), cis-PtCl2(1-Cl)2 (12) and cis-PtCl2(1-Br)2 (13). Reaction of Pd(Me)Cl(cod) (cod = cycloocta-1,5-diene) with either 1 equiv. of 1-Br or the known pyridylphosphines 1′-Cl, 1-OH or 1-H gave the P/N-chelate complexes Pd(Me)Cl(1-Br-1-H) (14)-(17). All new compounds have been fully characterised by spectroscopic and analytical methods. Furthermore the structures of 4, 5, 10 and 16 · (CH3)2SO have been elucidated by single crystal X-ray crystallography. A crystal structure of the dinuclear metallocycle trans,trans-[PdCl2{μ-P/N-{Ph2PCH2N(H)}C5H4N}]2 · CHCl3, 18 · CHCl3, has also been determined. Here 1-H bridges, using both P and pyridyl N donors, two dichloropalladium centres affording a 12-membered ring with the PdCl2 units adopting a head-to-tail arrangement.  相似文献   

13.
Reduction of RuQ3 (1a, Q = 8-quinolinolato) with Zn/Hg in the presence of various π-acceptor ligands in ethanol affords RuQ2L2 (L2 = (dimethylsulfoxide)2 (2); (4-picoline)2 (3); N,N′-dimethyl-1,4-diazabuta-1,3-diene, dab (4); cyclooctadiene, COD (5); norborna-2,5-diene, nbd (6)). Compound 6 is isolated as an equimolar mixture of cis,trans (6a) and trans,cis (6b) isomers, which can be separated by column chromatography. DFT calculations have been performed on 6a and 6b. Oxidation of 3 and 6b affords the corresponding ruthenium(III) species 7 and 8, respectively. The structures of 2, 3, 4 and 6 have been determined by X-ray crystallography.  相似文献   

14.
Plankton filament cyanobacteria Prochlorothrix hollandica is characterized by a very high content of C14 and C16 fatty acids (FA) in the lipid membranes. Depending on culturing conditions of the cyanobacteria, total concentrations of myristic and myristoleic acids can reach 35% and those of palmitic and palmitoleic acids can reach 60% of all esterified FA cells. In P. hollandica, a variety of monounsaturated FA is represented by myristoleic and palmitic acids, and by hexadecenoic (C16:1) acid with olefin bond of cis-configuration, located in the Δ4 position. The process of intensive culturing for P. hollandica cells to yield a maximal biomass in order to isolate the pure drug of myristoleic acid derivative has been optimized. The use of a threestage purification gives 30 mg of chromatographically pure myristoleic acid methyl ester from 17 g of P. hollandica raw biomass (dry mass is 3 g), which is 1% of dry cell mass.  相似文献   

15.

Background

Trans fatty acids are produced either by industrial hydrogenation or by biohydrogenation in the rumens of cows and sheep. Industrial trans fatty acids lower high-density lipoprotein (HDL) cholesterol, raise low-density lipoprotein (LDL) cholesterol, and increase the risk of coronary heart disease. The effects of trans fatty acids from ruminants are less clear. We investigated the effect on blood lipids of cis-9, trans-11 conjugated linoleic acid (CLA), a trans fatty acid largely restricted to ruminant fats.

Methodology/Principal Findings

Sixty-one healthy women and men were sequentially fed each of three diets for three weeks, in random order, for a total of nine weeks. Diets were identical except for 7% of energy (approximately 20 g/day), which was provided either by oleic acid, by industrial trans fatty acids, or by a mixture of 80% cis-9, trans-11 and 20% trans-10, cis-12 CLA. After the oleic acid diet, mean (± SD) serum LDL cholesterol was 2.68±0.62 mmol/L compared to 3.00±0.66 mmol/L after industrial trans fatty acids (p<0.001), and 2.92±0.70 mmol/L after CLA (p<0.001). Compared to oleic acid, HDL-cholesterol was 0.05±0.12 mmol/L lower after industrial trans fatty acids (p = 0.001) and 0.06±0.10 mmol/L lower after CLA (p<0.001). The total-to–HDL cholesterol ratio was 11.6% higher after industrial trans fatty acids (p<0.001) and 10.0% higher after CLA (p<0.001) relative to the oleic acid diet.

Conclusions/Significance

High intakes of an 80∶20 mixture of cis-9, trans-11 and trans-10, cis-12 CLA raise the total to HDL cholesterol ratio in healthy volunteers. The effect of CLA may be somewhat less than that of industrial trans fatty acids.

Trial Registration

ClinicalTrials.gov NCT00529828  相似文献   

16.
Palladium(II) and platinum(II) complexes with N-alkylpyridylpyrazole-derived ligands, 2-(1-ethyl-5-phenyl-1H-pyrazol-3-yl)pyridine (L1) and 2-(1-octyl-5-phenyl-1H-pyrazol-3-yl)pyridine (L2), cis-[MCl2(L)] (M = Pd(II), Pt(II)), have been synthesised. Treatment of [PdCl2(L)] (L = L1, L2) with excess of ligand (L1, L2), pyridine (py) or triphenylphosphine (PPh3) in the presence of AgBF4 and NaBPh4 produced the following complexes: [Pd(L)2](BPh4)2, [Pd(L)(py)2](BPh4)2 and [Pd(L)(PPh3)2](BPh4)2. All complexes have been characterised by elemental analyses, conductivity, IR and NMR spectroscopies. The crystal structures of cis-[PdCl2(L2)] (2) and cis-[PtCl2(L1)] (3) were determined by a single crystal X-ray diffraction method. In both complexes, the metal atom is coordinated by one pyrazole nitrogen, one pyridine nitrogen and two chlorine atoms in a distorted square-planar geometry. In complex 3, π-π stacking between pairs of molecules is observed.  相似文献   

17.
The reaction of [RuCl3(2mqn)NO] (H2mqn=2-methyl-8-quinolinol) with 2-chloro-8-quinolinol (H2cqn) afforded cis-1 [RuCl(2cqn)(2mqn)NO] (the oxygen of 2cqn is trans to the NO) (complex 1), cis-1 [RuCl(2cqn)(2mqn)NO] (the oxygen of 2mqn is trans to the NO) (complex 2) and a 1:1 mixture of cis-2 [RuCl(2cqn)(2mqn)NO] (the oxygen of 2mqn is trans to the NO) and cis-2 [RuCl(2cqn)(2mqn)NO] (the oxygen of 2cqn is trans to the NO) (complex 3). The reaction was compared with that of [RuCl3(2mqn)NO] with 8-quinolinol (Hqn) or 5-chloro-8-quinolinol (H5cqn). Photoirradiation reaction of complex 1 at room temperature in deaerated CH2Cl2 in the presence of NO gave trans-[RuCl(2cqn)(2mqn)NO] (the Cl is trans to the NO) and complex 2 with recovery of complex 1. The reaction was contrasted with that of cis-1 [RuCl(qn)(2mqn)NO] or cis-1 [RuCl(5cqn)(2mqn)NO]. The crystal structure of complex 1 was determined by X-ray diffraction. The reactions were examined under consideration of atomic charge of the phenolato oxygen in 8-quinolinol and its derivatives calculated at the restricted Hartree-Fock/6-311G** level.  相似文献   

18.
19.
Reaction of cis-[Ru(acac)22-C8H14)2] (1) (acac = acetylacetonato) with two equivalents of PiPr3 in THF at −25 °C gives trans-[Ru(acac)2(PiPr3)2], trans-3, which rapidly isomerizes to cis-3 at room temperature. The poorly soluble complex [Ru(acac)2(PCy3)2] (4), which is isolated similarly from cis-[Ru(acac)22-C2H4)2] (2) and PCy3, appears to exist in the cis-configuration in solution according to NMR data, although an X-ray diffraction study of a single crystal shows the presence of trans-4. In benzene or toluene 2 reacts with PiPr3 or PCy3 to give exclusively cis-[Ru(acac)22-C2H4)(L)] [L = PiPr3 (5), PCy3 (6)], whereas in THF species believed to be either square pyramidal [Ru(acac)2L], with apical L, or the corresponding THF adducts, can be detected by 31P NMR spectroscopy. Complexes 3-6 react with CO (1 bar) giving trans-[Ru(acac)2(CO)(L)] [L = PiPr3 (trans-8), PCy3 (trans-9)], which are converted irreversibly into the cis-isomers in refluxing benzene. Complex 5 scavenges traces of dinitrogen from industrial grade dihydrogen giving a bridging dinitrogen complex, cis-[{Ru(acac)2(PiPr3)} 2(μ-N2)] (10). The structures of cis-3, trans-4, 5, 6 and 10 · C6H14 have been determined by single-crystal X-ray diffraction. Complexes trans- and cis-3, 5, 6, cis-8, and trans- and cis-9 each show fully reversible one-electron oxidation by cyclic voltammetry in CH2Cl2 at −50 °C with E1/2(Ru3+/2+) values spanning −0.14 to +0.92 V (versus Ag/AgCl), whereas for the vinylidene complexes [Ru(acac)2 (CCHR)(PiPr3)] [R = SiMe3 (11), Ph (12)] the process is irreversible at potentials of +0.75 and +0.62 V, respectively. The trend in potentials reflects the order of expected π-acceptor ability of the ligands: PiPr3, PCy3 <C 2H4 < CCHR < CO. The UV-Vis spectrum of the thermally unstable, electrogenerated RuIII-ethene cation 6+ has been observed at −50 °C. Cyclic voltammetry of the μ-dinitrogen complex 10 shows two, fully reversible processes in CH2Cl2 at −50 °C at +0.30 and +0.90 V (versus Ag/AgCl) corresponding to the formation of 10+ (RuII,III) and 102+ (RuIII,III). The former, generated electrochemically at −50 °C, shows a band in the near IR at ca. 8900 cm−1 (w1/2 ca. 3700 cm−1) consistent with the presence of a valence delocalized system. The comproportionation constant for the equilibrium 10 + 102+ ? 2 10+ at 223 K is estimated as 1013.6.  相似文献   

20.
The preparation and structural characterization of several new Ru(II) complexes in which four coordination positions are occupied by the sulfur atoms of a macrocycle, either 1,4,7,10-tetrathiacyclododecane ([12]aneS4) or 1,5,9,13-tetrathiacyclohexadecane ([16]aneS4), and the two others by relatively labile ligands (Cl, , H2O, dmso-S), are described:cis-[Ru([12]aneS4)(dmso-S)(H2O)](CF3SO3)2 (2a), cis-[Ru([12]aneS4)(dmso-S)(ONO2)](NO3) (2b), cis-[Ru([16]aneS4)Cl2] (4), and trans-[Ru([16]aneS4)(dmso-S)(H2O)](CF3SO3)2 (5).The complexes of the larger [16]aneS4 macrocycle have a flexible coordination geometry, either cis or trans, that makes them unsuited for being used as precursors in metal-driven self-assembly processes.On the contrary, the [12]aneS4 complexes cis-[Ru([12]aneS4)(dmso-S)Cl]Cl (1) and, above all, its chlorido free derivatives cis-[Ru([12]aneS4)(dmso-S)(H2O)](CF3SO3)2 (2a) and cis-[Ru([12]aneS4)(dmso-S)(ONO2)](NO3) (2b) are potential precursors of the geometrically stable 90° bis-acceptor fragment cis-[Ru([12]aneS4)]2+.Preliminary results of their reactivity towards the linear linker pyrazine (pyz) showed that the nature of the isolated product depends on that of the counter-anion.When treated with pyz 2b afforded the dinuclear complex [{Ru([12]aneS4)(ONO2)}2(μ-pyz)](NO3)2 (8), while 2a gave the molecular triangle [{cis-Ru([12]aneS4)(μ-pyz)}3](CF3SO3)6 (9), both in low yields.The X-ray structures of compounds 2a, 2b, 4, 5, [{Ru([12]aneS4)Cl}2(μ-pyz)]Cl2 (7), 9, and of the sandwich complex[Ru([12]aneS3-S)2](CF3SO3)2 (3), in which only three sulfur atoms of each macrocycle are bound to ruthenium, are also described.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号