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 15 毫秒
1.
The kinetics of substitution reactions of [η-CpFe(CO)3]PF6 with PPh3 in the presence of R-PyOs have been studied. For all the R-PyOs (R = 4-OMe, 4-Me, 3,4-(CH)4, 4-Ph, 3-Me, 2,3-(CH)4, 2,6-Me2, 2-Me), the reactions yeild the same product [η5-CpFe(CO)2PPh3]PF6, according to a second-order rate law that is first order in concentrations of [η5-CpFe(CO)3]PF6 and of R-PyO but zero order in PPh3 concentration. These results, along with the dependence of the reaction rate on the nature of R-PyO, are consistent with an associative mechanism. Activation parameters further support the bimmolecular nature of the reactions: ΔH = 13.4 ± 0.4 kcal mol−1, ΔS = −19.1 ± 1.3 cal k−1 mol−1 for 4-PhPyO; ΔH = 12.3 ± 0.3 kcal mol−1, ΔS = 24.7 ±1.0 cal K−1 mol−1 for 2-MePyO. For the various substituted pyridine N-oxides studied in this paper, the rates of reaction increase with the increasing electron-donating abilities of the substituents on the pyridine ring or N-oxide basicities, but decrease with increasing 17O chemical shifts of the N-oxides. Electronic and steric factors contributing to the reactivity of pyridine N-oxides have been quantitatively assessed.  相似文献   

2.
The kinetics of the decomposition reactions of the CO(py)3(CO3)(H2O)+ ion have been investigated in aqueous perchloric acid solutions over a range of hydrogen ion concentrations (0.10 to 5.0 M) and at two ionic strengths (I = 1.0 and 5.0 M). At the lower ionic strength, plots of ln (AtA versus time show a nonlinearity that is consistent with that expected for consecutive first-order reactions. The rates of the faster reaction are similar to those reported for the spontaneous reduction of aquopyridine-cobalt(III) cations. At the higher ionic strength, the above noted curvature is not apparent and the decarboxylation kinetics of the title complex may be described by a pseudo-first-order rate constant: kobs = k[H3O+]. At 20°C, k = (1.75−+0.09) s−1 M−1 with activation parameters ofΔH = (97 −+ 4) kJ mol−1 and ΔS = −(54 −+ 32) J deg−1 mol−1. These kinetic parameters are compared with those previously reported for the similar complexes, Co(py)4CO3+ and Co(py)2(CO3)(H2O)2+.  相似文献   

3.
Trityl borate salts [4-RPyCPh3][B(C6F5)4] (R = H 1, tBu 2, Et 3, NMe24) and [R3PCPh3][B(C6F5)4] (R = Me 5, nBu 6, Ph[1] 7, p-MeC6H48) are readily prepared via equimolar reaction of the appropriate pyridine or phosphine and trityl borate [CPh3][B(C6F5)4]. The analogous reactions of PiPr3 affords the product [(p-iPr3P-C6H4)Ph2CH][B(C6F5)4] (9) while the corresponding reactions of Cy3P and tBu3P gave the cyclohexadienyl derivatives [(p-R3PC6H5)CPh2][B(C6F5)4] (R = Cy 10, tBu 11). X-ray structures of 5 and 9 are reported.  相似文献   

4.
The hydroxo complex [NBu4]2[Ni2(C6F5)4(μ-OH)2] reacts with ammonium O,O-dialkyldithiophosphates, O-alkyl-p-methoxyphenyldithiophosphonate acids and ammonium O-alkylferrocenyldithiophosphonates in dichloromethane under mild conditions to give, respectively, [NBu4][Ni(C6F5)2{S(S)P(OR)2}] (R=Me (1), Et (2), iPr (3)) and [NBu4][Ni(C6F5)2{S(S)P(OR)Ar}] (Ar=p-MeOC6H4, R=Me (4), Et (5), iPr (6); Ar=ferrocenyl; R=Me (7), Et (8), iPr (9)). The monothiophosphonate nickel complexes [NBu4][Ni(C6F5)2{S(S)P(OR)(ferrocenyl)}] (R=Et (10), iPr (11)) are obtained by reaction of the hydroxo complex with O-alkylferrocenyldithiophosphonate acids. Analytical (C, H, N, S), conductivity, and spectroscopic (IR, 1H, 19F and 31P NMR, and FAB-MS) data were used for structural assignments. A single-crystal X-ray diffraction study of [NBu4][Ni(C6F5)2{S(S)P(OMe)(p-MeOC6H4)}] (4) and [NBu4][Ni(C6F5)2{S(O)P(OEt)(ferrocenyl)}] (10) shows that in both cases the coordination around the nickel atom es essentially square planar with NiC2S2 and NiC2SO central cores, respectively.  相似文献   

5.
This study compared the mass-specific routine metabolic rate (RMR) of similar sized mulloway (Argyrosomus japonicus), a sedentary species, and yellowtail kingfish (Seriola lalandi), a highly active species, acclimated at one of several temperatures ranging from 10–35 °C. Respirometry was carried out in an open-top static system and RMR corrected for seawater–atmosphere O2 exchange using mass-balance equations. For both species RMR increased linearly with increasing temperature (T). RMR for mulloway was 5.78T − 29.0 mg O2 kg− 0.8 h− 1 and for yellowtail kingfish was 12.11T − 39.40 mg O2 kg− 0.8 h− 1. The factorial difference in RMR between mulloway and yellowtail kingfish ranged from 2.8 to 2.2 depending on temperature. The energetic cost of routine activity can be described as a function of temperature for mulloway as 1.93T − 9.68 kJ kg− 0.8 day− 1 and for yellowtail kingfish as 4.04T − 13.14 kJ kg− 0.8 day− 1. Over the full range of temperatures tested Q10 values were approximately 2 for both species while Q10 responses at each temperature increment varied considerably with mulloway and yellowtail kingfish displaying thermosensitivities indicative of each species respective niche habitat. RMR for mulloway was least thermally dependent at 28.5 °C and for yellowtail kingfish at 22.8 °C. Activation energies (Ea) calculated from Arrhenius plots were not significantly different between mulloway (47.6 kJ mol− 1) and yellowtail kingfish (44.1 kJ mol− 1).  相似文献   

6.
The cytoplasmic concentrations of Cl([Cl]i) and Ca2+ ([Ca2+]i) were measured with the fluorescent indicators N-(ethoxycarbonylmethyl)-6-methoxyquinilinum bromide (MQAE) and fura-2 in pancreatic β-cells isolated from ob/ob mice. Steady-state [Cl]i in unstimulated β-cells was 34 mM, which is higher than expected from a passive distribution. Increase of the glucose concentration from 3 to 20 mM resulted in an accelerated entry of Cl into β-cells depleted of this ion. The exposure to 20 mM glucose did not affect steady-state [Cl]i either in the absence or presence of furosemide inhibition of Na+, K+, 2 Cl co-transport. Glucose-induced oscillations of [Ca2+]i were transformed into sustained elevation in the presence of 4,4′ diisothiocyanato-dihydrostilbene-2,2′-disulfonic acid (H2DIDS). A similar effect was noted when replacing 25% of extracellular Cl with the more easily permeating anions SCN, I, NO3 or Br. It is concluded that glucose stimulation of the β-cells is coupled to an increase in their Cl permeability and that the oscillatory Ca2+ signalling is critically dependent on transmembrane Cl fluxes.  相似文献   

7.
Potato plants (Solanum tuberosum L. cv. Bintje) were grown to maturity in open-top chambers under three carbon dioxide (CO2; ambient and 24 h d−1 seasonal mean concentrations of 550 and 680 μmol mol−1) and two ozone levels (O3; ambient and an 8 h d−1 seasonal mean of 50 nmol mol−1). Chlorophyll content, photosynthetic characteristics, and stomatal responses were determined to test the hypothesis that elevated atmospheric CO2 may alleviate the damaging influence of O3 by reducing uptake by the leaves. Elevated O3 had no detectable effect on photosynthetic characteristics, leaf conductance, or chlorophyll content, but did reduce SPAD values for leaf 15, the youngest leaf examined. Elevated CO2 also reduced SPAD values for leaf 15, but not for older leaves; destructive analysis confirmed that chlorophyll content was decreased. Leaf conductance was generally reduced by elevated CO2, and declined with time in the youngest leaves examined, as did assimilation rate (A). A generally increased under elevated CO2, particularly in the older leaves during the latter stages of the season, thereby increasing instantaneous transpiration efficiency. Exposure to elevated CO2 and/or O3 had no detectable effect on dark-adapted fluorescence, although the values decreased with time. Analysis of the relationships between assimilation rate and intercellular CO2 concentration and photosynthetically active photon flux density showed there was initially little treatment effect on CO2-saturated assimilation rates for leaf 15. However, the values for plants grown under 550 μmol mol−1 CO2 were subsequently greater than in the ambient and 680 μmol mol−1 treatments, although the beneficial influence of the former treatment declined sharply towards the end of the season. Light-saturated assimilation was consistently greater under elevated CO2, but decreased with time in all treatments. The values decreased sharply when leaves grown under elevated CO2 were measured under ambient CO2, but increased when leaves grown under ambient CO2 were examined under elevated CO2. The results obtained indicate that, although elevated CO2 initially increased assimilation and growth, these beneficial effects were not necessarily sustained to maturity as a result of photosynthetic acclimation and the induction of earlier senescence.  相似文献   

8.
Despite a high Ca2+-permeability of the P2Z receptor in human B lymphocytes, extracellular ATP4has only a minor effect on global [Ca2+]i. The aim of this study was to reveal the mechanisms responsible for this discrepancy. We investigated the relationship between ATP4−-application, Cai2-response, membrane current and membrane potential in two human B cell lines and in human tonsillar B cells. This was achieved by a combination of FACS- and voltage clamp measurements and the usage of appropriate voltage- and Ca2-sensitive fluorescent dyes. ATP4-induced changes in whole-cell current and [Ca2]iwere blocked by extracellular as well as intracellular Na+. Under current clamp conditions, ATP4−-induced Na+-entry diminished the Ca2+entry via reduction of the driving force. A substantial increase in [Ca2+]iinduced by ATP4−was only observed in Na+-free solutions.The pathway of signal transduction activated by ATP4via P2Z receptor of human B lymphocytes under physiological conditions seems not to operate by an increase in the global intracellular Ca2-concentration, but rather by the depolarization of the cell membrane as a result of the Na+-influx.  相似文献   

9.
The reaction between nickel(II) nitrate and potassium phosphorus-1,1-dithiolates (di-sec-butyl and di-iso-butyl) in methanol yields 2:1 complexes which were characterized by FT-IR and NMR spectroscopy. 2:1 pyrazole adducts of both compounds were also obtained.The X-ray diffraction analysis of the compounds reveals square planar, four-coordination geometry for the homoleptic compounds and a six-coordinated distorted octahedral geometry for the adducts. In Ni[S2P(OBus)2]2 the molecules are associated through C-H?O hydrogen bonds (2.652 Å), and in Ni[S2P(OBui)2]2 the molecules are associated through C-H?S hydrogen bonds (2.948 Å). The pyrazole adducts are associated through N-H?O bonds and N-H?S bonds from the pyrazole nitrogen atoms, to form supramolecular assemblies. Thus, Ni[S2P(OBus)2(Pz)2]2 (Pz = pyrazole) forms bi-dimensional layers through N-H?O and N-H?S bonds (2.502 and 2.965 Å, respectively), whereas Ni[S2P(OBui)2(Pz)2]2 forms linear chains with N-H?S bonds 2.728 Å. The dithiophosphato groups behave as isobidentate chelating ligands.  相似文献   

10.
trans-[Ru(NH3)4P(OR)3(H2O)]2+ (R = Me, Pr, iPr, and Bu) reacts with isonicotinamide at second-order- specific rates k1 of 1.2, 2.3, 7.4 and 8.1 M−1 s(25 °C, μ = 0.10 NaCF3COO/CH3COOH), respectively, for R = Me, Pr, iPr and Bu. The products trans- [Ru(NH3)4P(OR)3isn](PF6)2 have been isolated and characterized by micro analysis, cyclic voltammetry, and electronic spectral data. The aquation rates k−1 for the isonicotinamide (isn) derivatives are 5.2 × 10−2, 5.9 × 10−2, 2.0 × 10−1 and 3.4 × 10−1 s−1 for R= Me, Pf, Bu and iPr, respectively. The activation parameters for the forward and backward reactions indicate the same mechanism for all of them. The substitution proceeds by a dissociative mechanism with a significant outer-sphere association of trans-[Ru(NH3)4P(OR)3(H2O)]2+ complexes with isn. Assuming k1 as indicative of the lability of the coordinated water molecule on the monophosphite complexes, the following sequence of increasing trans-effect mav be proposed: P(OMe)3 <P(OEt)3 <P(OPr)3 <P(OiPr)3 <P(OBu)3. The affinity of the monophosphite complexes for isn increases according to P(OMe)3 ⋍ P(OiPr)3 < P(OEt)3 < P(OPr)3 ⋍ P(OBu)3.  相似文献   

11.
We have previously reported that angiotensin II (ANG II) induces oscillations in the cytoplasmic calcium concentration ([Ca2+]i) of pulmonary vascular myocytes. The present work was undertaken to investigate the effect of ANG II in comparison with ATP and caffeine on membrane currents and to explore the relation between these membrane currents and [Ca2+]i. In cells clamped at −60 mV, ANG II (10 μM) or ATP (100 μM) induced an oscillatory inward current. Caffeine (5 μM) induced only one transient inward current. In control conditions, the reversal potential (Erev) of these currents was close to the equilibrium potential for Cl ions (ECl = −2.1 mV) and was shifted towards more positive values in low-Cl solutions. Niflumic acid (10–50 μM) and DIDS (0.25-1 mM) inhibited this inward current. Combined recordings of membrane current and [Ca2+]i by Indo-1 microspectrofluorimetry revealed that ANG II- and ATP-induced currents occurred simultaneously with oscillations in [Ca2+]i, whereas the caffeine-induced current was accompanied by only one transient increase in [Ca2+]i Niflumic acid (25 μM) had no effect on agonist-induced [Ca2+]i responses, whereas thapsigargin (1 μM) abolished both membrane current and the [Ca2+]i response. Heparin (5 mg/ml in the pipette solution) inhibited both [Ca2+]i responses and membrane currents induced by ANG II and ATP, but not by caffeine. In pulmonary arterial strips, ANG II-induced contraction was inhibited by niflumic acid (25 μM) or nifedipine (1 μM) to the same extent and the two substances did not have an additive effect. This study demonstrates that, in pulmonary vascular smooth muscle, ANG II, as well as ATP, activate an oscillatory calcium dependent chloride current which is triggered by cyclic increases in [Ca2+]i and that both oscillatory phenomena are primarily IP3 mediated. It is suggested that ANG II-induced oscillatory chloride current could depolarise the cell membrane leading to activation of voltage-operated Ca2+ channels. The resulting Ca2+ influx contributes to the component of ANG II-induced contraction that is equally sensitive to chloride or calcium channel blockade.  相似文献   

12.
Reactions of SbCl5 with various covalent metal halides in MeCN have been studied as a convenient and direct route to metal hexachloroantimonate salts via Sb(V) halide abstraction. The isolation and characterization (Ir, Vis-UV, 1H NMR spectroscopic and microanalytical) of the complexes [Zn(MeCN)6][SbCl6]2, [CrCl2(MeCN)4][SbCl6], [SnCl3(MeCN)3][SbCl6], [TiCl2(MeCN)4][SbCl6]2, [Cp2M(Cl)(MeCN)x][SbCl6] M = ti, x = 1; M = Zr, Hf, x = 2, and [Cp2M(MeCN)y][SbCl6]2 M = Ti, y = 2; M = Zr, Hf, y = 3, is described. The reaction of MgCl2 with SbCl5 was carried out in EtOAC as solvent and gave [Mg(EtOAc)6][SbCl6]2. 121Sb NMR, IR and UV spectroscopic measurements provide positive identification of the SbCl6 anion.  相似文献   

13.
A sensitive, selective, and rapid enzymatic method is proposed for the quantification of hydrogen peroxide (H2O2) using 3-methyl-2-benzothiazolinonehydrazone hydrochloride (MBTH) and 10,11-dihydro-5H-benz(b,f)azepine (DBZ) as chromogenic cosubstrates catalyzed by horseradish peroxidase (HRP) enzyme. MBTH traps free radical released during oxidation of H2O2 by HRP and gets oxidized to electrophilic cation, which couples with DBZ to give an intense blue-colored product with maximum absorbance at 620 nm. The linear response for H2O2 is found between 5 × 10−6 and 45 × 10−6 mol L−1 at pH 4.0 and a temperature of 25 °C. Catalytic efficiency and catalytic power of the commercial peroxidase were found to be 0.415 × 106 M−1 min−1 and 9.81 × 10−4 min−1, respectively. The catalytic constant (kcat) and specificity constant (kcat/Km) at saturated concentration of the cosubstrates were 163.2 min−1 and 4.156 × 106 L mol−1 min−1, respectively. This method can be incorporated into biochemical analysis where H2O2 undergoes catalytic oxidation by oxidase. Its applicability in the biological samples was tested for glucose quantification in human serum.  相似文献   

14.
The reaction of [ZnLI,II2] (LI = [NH2C(S)NP(O)(OiPr)2]; LII = [PhNHC(S)NP(O)(OiPr)2]) or [Cd2LIV4] (LIV = [PhC(S)NP(O)(OiPr)2]) with 2,2′-bipyridine (bpy) or 1,10-phenanthroline (phen) leads to the heteroligand complexes [Zn(bpy)LI,II2], [Zn(phen)LI,II2], [Cd(bpy)LIV2] or [Cd(phen)LIV2], respectively. The introduction of the diimine ligands into the coordination sphere of the metal cation provokes a change from 1,5-O,S- to 1,3-N,S-coordination of the anionic ligands for Zn but not for the Cd species. The reaction of [Zn(phen)LIV2] (LIV = PhC(S)NP(O)(OiPr)2) with CH2Cl2 cleaves the chlorine atoms from CH2Cl2 and leads to the formation of [Zn(phen)LIVCl] and S,S′-bis(benzimidothio-N-diisopropoxyphosphoryl)methane (LIV-CH2-LIV) in high yields. Using CHCl3 or CCl4 instead of CH2Cl2 does not lead to the formation of chlorine substituted products even under reflux conditions. The new compounds were investigated by 1H and 31P{1H} NMR, IR spectroscopy and microanalysis. Crystal structures of [ZnLII2], [Cd(phen)LIV2]·CH2Cl2, [Zn(bpy)LI2] and [Zn(phen)LIVCl] were elucidated by X-ray diffraction.  相似文献   

15.
The absolute configuration of cis-epoxyjasmone (−)-2, isolated from Trichosporum cutaneum CCT 1903 whole cells, has been unambiguously established as (7S,8R), [α]D20 −29.0° (c 1.3, CHCl3), by a new two step method, using a regioselective epoxide opening as the key step followed by Mosher acid derivatization.  相似文献   

16.
《Inorganica chimica acta》2002,327(1):169-178
New complexes [MI(CO)2(dppe){S2P(OEt)2}] (M=W, 1a; M=Mo, 1b), [MI(CO)2(dppm){S2P(OEt)2}] (M=W, 2a; M=Mo, 2b) and [W(CO)(dppe){S2P(OEt)2}2][O2dppe] (3a), were synthesised from [MI2(CO)3(NCMe)2] (M=Mo, W), after treatment with ammonium diethyldithiophosphate and phosphine under different conditions. The structure of the tungsten complexes was determined by single crystal X-ray diffraction. During the synthesis of 3a, oxidation of the phosphine took place and a molecule of oxidised phosphine occupies channels in the crystal. DFT/B3LYP calculations on models of 1a and 2a showed the capped octahedron structure, observed in most dicarbonyl complexes of this family, to be preferred by 1.4 and 2.6 kcal mol−1 for the dppm and the dppe complexes, respectively. Strong steric repulsions can reverse this trend, as happens with the rigid dppm ligand. Complex 1a adopts a pentagonal bipyramidal geometry, which is often found in related monocarbonyl complexes.  相似文献   

17.
Salts of the Fe(III) spin crossover cation [FeIII(qsal)2]+ (qsalH = N-(8-quinolyl)salicylaldimine) and monoanions [MIII(pds)2] (M = Cu, Au; pds = pirazine-2,3-diselenolate) with formula [FeIII(qsal)2][MIII(pds)2] were prepared and characterized by single crystal X-ray diffraction and magnetic measurements. These two salts present magnetic properties essentially due to the FeIII centres in the high-spin state (S = 5/2), and do not have any spin transition.  相似文献   

18.
A triangular [Zn3(μ3-OH)(OC(O)tBu)(μ2-κ1O:κ1O′-O2CtBu)4(3,5-lutidine)3] (1), a paddlewheel based dinuclear [Zn(μ2-κ1O:κ1O′-O2CtBu)2L]2 [L = 2,4-lutidine (2), 3,4-lutidine (3), and 2,3-lutidine (4)] and an hourglass based linear trinuclear [Zn3(μ2-κ1O:κ1O′-O2CtBu)6(pyridine)2] (5) complexes were synthesized to understand the role of subtle steric/basic properties of Lewis bases on the degree of aggregation of the products. The mononuclear Zn(OC(O)tBu)2·2H2O was also prepared in order to probe the origin of the μ3-OH moiety in complex 1. Complexes 1-5 and Zn(OC(O)tBu)2·2H2O were characterized by microanalytical, IR, TGA/DTA, solution (1H and 13C) NMR, solid-state cross-polarization magic angle spinning (CP-MAS) 13C NMR, mass spectral data and single crystal X-ray diffraction data. Complex 1 represents the first example of a discrete trinuclear zinc(II) carboxylate complex that contains a [Zn3(μ3-OH)]5+ core with zinc atoms in three distinct geometries namely a distorted tetrahedral, trigonal bipyramidal, and octahedral. A plausible mechanism for the formation of complexes 1-5 was explained with the aid of point zero charge (pzc) model.  相似文献   

19.
The rate constants for [1O2] [MCLA] and [1O2][NaN3] were measured by quenching the near-infrared emission (1Δg3g) in steady state with MCLA and NaN3, respectively. 1O2 was constantly generated by energy transfer to O2 from Ar laser-excited Rose Bengal. The Stern—Volmer plots yielded the second-order rate constants of 2.94 × 109 M?1 S?1 and 3.83 × 108 M?1 S?1 for quenching 1O2 with MCLA and NaN3 in water at pH 5.4, respectively. The 1O2 + MCLA reaction emitted light with maximum at 465 nm at pD 4.5 identical to the O2? + MCLA reaction.  相似文献   

20.
Described are the syntheses and structures of a phosphonium salt of the anionic ligand O-t-butyl-1,1-dithiooxalate, [PPh3Bz][i-dtotBu] ([PPh3Bz][1]), and of two Cu(I) complexes of this anion, Cu(PPh3)22-i-dtotBu) (2) and Cu(dmp)(PPh3)(η1-i-dtotBu) (3, dmp = 2,9-dimethyl-1,10-phenanthroline). In addition, it was found that the reaction of CuBr2 with i-dtotBu gives a O-t-butyl-1-perthio-1-thiooxalato complex of copper(I), [BzPh3P][Cu(Br)(S-i-dtotBu)] ([BzPh3P][4]), where [S-i-dtotBu] is a disulfide-containing anionic ligand. The electronic structure and absorption spectrum of this species were investigated by time dependent DFT methods.  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