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.
In the present study we have analyzed hydrogen bonding in dimer and trimer of oxalic acid, based on a recently proposed charge and energy decomposition scheme (ETS-NOCV). In the case of a dimer, two conformations, α and β, were considered. The deformation density contributions originating from NOCV’s revealed that the formation of hydrogen bonding is associated with the electronic charge deformation in both the σ—(Δρσ) and π-networks (Δρπ). It was demonstrated that σ-donation is realized by electron transfer from the lone pair of oxygen on one monomer into the empty rH - O* \rho_{H - O}^* orbital of the second oxalic acid fragment. In addition, a covalent contribution is observed by the density transfer from hydrogen of H-O group in one oxalic acid monomer to the oxygen atom of the second fragment. The resonance assisted component (Δρπ), is based on the transfer of electron density from the π—orbital localized on the oxygen of OH on one oxalic acid monomer to the oxygen atom of the other fragment. ETS-NOCV allowed to conclude that the σ(O---HO) component is roughly eight times as important as π (RAHB) contribution in terms of energetic estimation. The electrostatic factor (ΔEelstat) is equally as important as orbital interaction term (ΔEorb). Finally, comparing β-dimer of oxalic acid with trimer we found practically no difference concerning each of the O---HO bonds, neither qualitative nor quantitative.  相似文献   

2.
《Inorganica chimica acta》2006,359(11):3589-3595
Reactions between the activated cluster [Os3(CO)10(NCMe)2] and malonic acid, succinic acid and dicarboxylic acetylene, respectively, lead to the formation of the linked cluster complexes [{Os3H(CO)10}2(CO2CH2CO2)] (1), [{Os3H(CO)10}2(CO2C2H4CO2)] (2), and [{Os3H(CO)10}2(C4O4)] (3) in good yield. Cluster 3 was subsequently treated with [Co2(CO)8] and this results in the addition of a “Co2(CO)6” group giving [{Os3H(CO)10}2(C2O4){Co2(CO)6}] (4). The X-ray crystal structures are reported for 24. In each structure the two triangular triosmium units are linked by the carboxylate groups and within each complex the carboxylate groups are chelating and bridge two osmium atoms.  相似文献   

3.
The kinetics of rapid CO substitution by PPh3 in Co4(CO)12 and Rh4(CO)12 have been examined by stopped-flow and low temperature FT-IR methods. In Co4(CO)12 rapid (kobs ∼ 1.8 s−1) substitution of CO occurs after a 1–15 s induction period at 28 °C in C6H5Cl solvent by a catalytic process. Addition of PPh3 to Rh4(CO)12 yields Rh4(CO)11(PPh3) according to a predominantly second order rate law k1[Rh4- (CO)12] + k2[Rh4(CO)12][PPh3] with k1 = 25 ± 11 s−1 and k2 = 2.97 ± 0.27 X 104 M−1 s−1 at 28 °C. Substitution of a second CO ligand also occurs rapidly with k1 = 0.15 ± 0.09 s−1 and k2 = 6.54 ± 0.07 X 102 M−1 s−1 at 28 °C. The reactivity of Rh4(CO)12 toward associative substitution is 104– 1011 faster than for the Co and Ir analogues, In Rh4(CO)11(PPh3) the increase in CO substitution rates over Co and Rh analogues is 102–107. The ordering of associative substitution rates Co << Rh >>> Ir in these clusters exaggerates the trend seen in mononuclear metal complexes.  相似文献   

4.
Reaction of H2PtCl4 and K2PdCl4 with 2-hydroxyacetophenone N(4)-ethylthiosemicarbazone, H2Ap4Et, afforded [Pt(Ap4Et)(H2Ap4Et)] and [Pd(Ap4Et)(H2Ap4Et)]. Their crystal and molecular structures are reported and represent the first 1:2 thiosemicarbazone complexes with ligands having both different formal charge and denticity. The dianion, Ap4Et2−, coordinates in a planar conformation to palladium(II) or platinum(II) via the phenolato O, imine N and thiolato S atoms, while the neutral molecule exhibits monodentate coordination by the thione S atom. Intra-, intermolecular hydrogen bonds and C-H?π contacts lead to aggregation and a supramolecular assembly. Electronic, IR, and NMR spectral data, as well as electrochemical measurements, are included. The pKa values of the poorly water soluble H2Ap4Et were obtained spectrophotometrically in aqueous solutions of constant ionic strength.  相似文献   

5.
《Inorganica chimica acta》2006,359(11):3557-3564
Both Rh4(CO)12 and Rh6(CO)16 exhibit CO-fluxionality and modern, variable temperature, NMR methods allow the unambiguous assignment of the three terminal CO resonances and, for Rh4(CO)12, show that the mechanism of CO-fluxionality, which has been controversial for a long time, unambiguously involves the merry-go-round process; Rh6(CO)16, which was previously thought to be static, is also shown to be fluxional, although the rate of CO-exchange is much less than found for substituted derivatives, and possible pathways for this CO-exchange are discussed.  相似文献   

6.
Combined quantum chemical and molecular mechanics geometry optimisations have been performed on myoglobin without or with O(2) or CO bound to the haem group. The results show that the distal histidine residue is protonated on the N(epsilon 2) atom and forms a hydrogen bond to the haem ligand both in the O(2) and the CO complexes. We have also re-refined the crystal structure of CO[bond]myoglobin by a combined quantum chemical and crystallographic refinement. Thereby, we probably obtain the most accurate available structure of the active site of this complex, showing a Fe[bond]C[bond]O angle of 171 degrees, and Fe[bond]C and C[bond]O bond lengths of 170-171 and 116-117 pm. The resulting structures have been used to calculate the strength of the hydrogen bond between the distal histidine residue and O(2) or CO in the protein. This amounts to 31-33 kJ/mol for O(2) and 2-3 kJ/mol for CO. The difference in hydrogen-bond strength is 21-22 kJ/mol when corrected for entropy effects. This is slightly larger than the observed discrimination between O(2) or CO by myoglobin, 17 kJ/mol. We have also estimated the strain of the active site inside the protein. It is 2-4 kJ/mol larger for the O(2) complex than for the CO complex, independent of which crystal structure the calculations are based on. Together, these results clearly show that myoglobin discriminates between O(2) and CO mainly by electrostatic interactions, rather than by steric strain.  相似文献   

7.
Anacystis nidulans (Synechococcus) had a minimal doubling time of 5 hrs at 30 degrees C at saturating light intensity and carbon dioxide concentration. Half maximal growth rates in saturating CO2 occured at a light intensity of 0.54 mW per cm2, and there was an apparent threshold intensity of 0.13 mW per cm2 below which no growth occurred. Growth rate in saturating light was dependent on the concentration of CO2+H2CO3 in the medium, rather than on total dissolved CO2; half maximal rates were estimated at 0.1 mM CO2+H2CO3. Under saturating conditions of light and CO2, 14CO2 was fixed primarily into 3-PGA, and subsequently moved into sugar phosphates and amino acids. Incorporation into aspartate was relatively slow. CO2 fixation was strictly light-dependent. The changes in adenylate and pyridine nucleotide pools were followed in light/dark and dark/light transitions. Whereas adenylates relaxed slowly over 15-20 min to the concentrations characteristic of illuminated cells following the abrupt changes induced by darkening, the sharp drop in intracellular NADPH showed little dark recovery although rapid restoration occurred on reillumination. Other pyridine nucleotides showed no changes during these transitions. The nucleotide specificity and Km of partially purfied GAP dehydrogenase suggest a role for this enzyme in the regulation of CO2 fixation.  相似文献   

8.
C(4) photosynthesis has a number of distinct properties that enable the capture of CO(2) and its concentration in the vicinity of Rubisco, so as to reduce the oxygenase activity of Rubisco, and hence the rate of photorespiration. The aim of this review is to discuss the properties of this CO(2)-concentrating mechanism, and thus to indicate the minimum requirements of any genetically-engineered system. In particular, the Kranz leaf anatomy of C(4) photosynthesis and the division of the C(4)-cycle between two cell types involves intercellular co-operation that requires modifications in regulation and transport to make C(4) photosynthesis work. Some examples of these modifications are discussed. Comparisons are made with the C(4)-type photosynthesis found in single-celled C(4)-type CO(2)-concentrating mechanisms, such as that found in the aquatic plant, Hydrilla verticillata and the single-celled C(4) system found in the terrestrial chenopod Borszczowia aralocaspica. The outcome of recent attempts to engineer C(4) enzymes into C(3) plants is discussed.  相似文献   

9.
A mutant of the NAD-malic enzyme-type C(4) plant, Amaranthus edulis, which lacks phosphoenolpyruvate carboxylase (PEPC) in the mesophyll cells was studied. Analysis of CO(2) response curves of photosynthesis of the mutant, which has normal Kranz anatomy but lacks a functional C(4) cycle, provided a direct means of determining the liquid phase-diffusive resistance of atmospheric CO(2) to sites of ribulose 1,5-bisphosphate carboxylation inside bundle sheath (BS) chloroplasts (r(bs)) within intact plants. Comparisons were made with excised shoots of wild-type plants fed 3,3-dichloro-2-(dihydroxyphosphinoyl-methyl)-propenoate, an inhibitor of PEPC. Values of r(bs) in A. edulis were 70 to 180 m(2) s(-1) mol(-1), increasing as the leaf matured. This is about 70-fold higher than the liquid phase resistance for diffusion of CO(2) to Rubisco in mesophyll cells of C(3) plants. The values of r(bs) in A. edulis are sufficient for C(4) photosynthesis to elevate CO(2) in BS cells and to minimize photorespiration. The calculated CO(2) concentration in BS cells, which is dependent on input of r(bs), was about 2,000 microbar under maximum rates of CO(2) fixation, which is about six times the ambient level of CO(2). High re-assimilation of photorespired CO(2) was demonstrated in both mutant and wild-type plants at limiting CO(2) concentrations, which can be explained by high r(bs). Increasing O(2) from near zero up to ambient levels under low CO(2), resulted in an increase in the gross rate of O(2) evolution measured by chlorophyll fluorescence analysis in the PEPC mutant; this increase was simulated from a Rubisco kinetic model, which indicates effective refixation of photorespired CO(2) in BS cells.  相似文献   

10.
The quantum yields of C3 and C4 plants from a number of genera and families as well as from ecologically diverse habitats were measured in normal air of 21% O2 and in 2% O2. At 30 C, the quantum yields of C3 plants averaged 0.0524 ± 0.0014 mol CO2/absorbed einstein and 0.0733 ± 0.0008 mol CO2/absorbed einstein under 21 and 2% O2. At 30 C, the quantum yields of C4 plants averaged 0.0534 ± 0.0009 mol CO2/absorbed einstein and 0.0538 ± 0.0011 mol CO2/absorbed einstein under 21 and 2% O2. At 21% O2, the quantum yield of a C3 plant is shown to be strongly dependent on both the intercellular CO2 concentration and leaf temperature. The quantum yield of a C4 plant, which is independent of the intercellular CO2 concentration, is shown to be independent of leaf temperature over the ranges measured. The changes in the quantum yields of C3 plants are due to changes in the O2 inhibition. The evolutionary significance of the CO2 dependence of the quantum yield in C3 plants and the ecological significance of the temperature effects on the quantum yields of C3 and C4 plants are discussed.  相似文献   

11.
The conductance for CO2 diffusion in the mesophyll of leaves can limit photosynthesis. We have studied two methods for determining the mesophyll conductance to CO2 diffusion in leaves. We generated an ideal set of photosynthesis rates over a range of partial pressures of CO2 in the stroma and studied the effect of altering the mesophyll diffusion conductance on the measured response of photosynthesis to intercellular CO2 partial pressure. We used the ideal data set to test the sensitivity of the two methods to small errors in the parameters used to determine mesophyll conductance. The two methods were also used to determine mesophyll conductance of several leaves using measured rather than ideal data sets. It is concluded that both methods can be used to determine mesophyll conductance and each method has particular strengths. We believe both methods will prove useful in the future.  相似文献   

12.
《Inorganica chimica acta》1986,121(2):161-166
Atomic Na, K and Cs were codeposited with CO2 in excess of matrix gas at the temperature of 12 K. The IR spectra revealed the presence of ionic aggregates corresponding to the molecules M(CO)2 and M2(CO2) (M=Na, K, Cs). Both molecular species have C2v symmetry; M(CO2) species have a planar ring structure while M2(CO2) have a W-shape structure. M2(CO2) molecules with Cs symmetry were also identified. The geometrical parameters of all the molecules were determined by 12C/13C and 16O/18O isotopic shifts. Raman spectra were also recorded and the results are reported in this study. The effect of photolysis on the structure of these molecules was examined. It was determined that photolysis promotes the formation of Na(CO2) and transforms the M2(CO2) molecules with C2v symmetry into Cs symmetry isomers.  相似文献   

13.
Protease inhibitors are known to suppress basal, fluoride-, and hormone-stimulated adenylate cyclase activities. The thrombin inhibitor, dansyl-arginyl-(4'-ethyl)piperidine amide (DAPA), also specifically inhibits the binding of gonadotropins to their receptors. Our studies were undertaken to find a concentration of DAPA that would specifically inhibit gonadotropin-stimulated adenylate cyclase without significantly altering basal, fluoride-, isoproterenol-, or prostaglandin E1-stimulated cyclase. Basal adenylate cyclase activity was not inhibited by DAPA in either human chorionic gonadotropin (hCG)- or follicle-stimulating hormone (FSH)-responsive rat ovarian plasma membranes. Human chorionic gonadotropin-stimulated cyclase was completely inhibited by DAPA at a concentration of 2.96 mM; the ID50 was 1.32 mM. Follicle-stimulating hormone-stimulated cyclase was completely inhibited by a DAPA concentration of 4.44 mM, and the ID50 was 1.75 mM. Dansyl-arginyl-(4'-ethyl)piperidine amide (2.96 mM) inhibited isoproterenol-, prostaglandin E1-, and fluoride-stimulated cyclase in hCG-responsive membranes by 11%, 28%, and 35%, respectively. Dansyl-arginyl-(4'-ethyl)piperidine amide (4.44 mM) inhibited fluoride- and prostaglandin-stimulated cyclase in FSH-responsive membranes by 10% and 11%, respectively. The data show that appropriate concentrations of DAPA can antagonize gonadotropin-stimulated adenylate cyclase while only minimally affecting fluoride- and other receptor-activated cyclase activities.  相似文献   

14.
CGP 6085 A [4-(5,6-dimethyl-2-benzofuranyl) piperidine HCl], a reported serotonin uptake and MAO (16) inhibitor, is a potent hypothermic agent. The hypothermic action of CGP 6085 A is dose dependent with a maximal reduction in rectal core temperature of greater than 1 degree C within one hour after drug administration. Fluoxetine and citalopram elicit a similar response at equal doses. These results suggest that inhibition of serotonin uptake may produce the hypothermic effect. To assess the in vivo action of CGP 6085 A in inhibiting hypothalamic serotonin uptake, CGP 6085 A (10 mg/kg) was injected one hour prior to injection of 3-hydroxy-4-methyl-alpha-ethyl-phenylethylamine (H75/12), a serotonin depletor. The ability of CGP 6085 A to block the uptake of H75/12 by the 5HT uptake system was indicative of its ability to block serotonin uptake. Pretreatment with p-chlorophenylalanine (pCPA), an inhibitor of serotonin synthesis, resulted in the loss of the hypothermic response to CGP 6085 A. Thus, these data are consistent with the idea that CGP 6085 A may produce its hypothermic response by inhibiting serotonin uptake.  相似文献   

15.
Dai Z  Ku M  Edwards GE 《Plant physiology》1993,103(1):83-90
Despite previous reports of no apparent photorespiration in C4 plants based on measurements of gas exchange under 2 versus 21% O2 at varying [CO2], photosynthesis in maize (Zea mays) shows a dual response to varying [O2]. The maximum rate of photosynthesis in maize is dependent on O2 (approximately 10%). This O2 dependence is not related to stomatal conductance, because measurements were made at constant intercellular CO2 concentration (Ci); it may be linked to respiration or pseudocyclic electron flow. At a given Ci, increasing [O2] above 10% inhibits both the rate of photosynthesis, measured under high light, and the maximum quantum yield, measured under limiting light ([phi]CO2). The dual effect of O2 is masked if measurements are made under only 2 versus 21% O2. The inhibition of both photosynthesis and [phi]CO2 by O2 (measured above 10% O2) with decreasing Ci increases in a very similar manner, characteristically of O2 inhibition due to photorespiration. There is a sharp increase in O2 inhibition when the Ci decreases below 50 [mu]bar of CO2. Also, increasing temperature, which favors photorespiration, causes a decrease in [phi]CO2 under limiting CO2 and 40% O2. By comparing the degree of inhibition of photosynthesis in maize with that in the C3 species wheat (Triticum aestivum) at varying Ci, the effectiveness of C4 photosynthesis in concentrating CO2 in the leaf was evaluated. Under high light, 30[deg]C, and atmospheric levels of CO2 (340 [mu]bar), where there is little inhibition of photosynthesis in maize by O2, the estimated level of CO2 around ribulose-1,5-bisphosphate carboxylase/oxygenase (Rubisco) in the bundle sheath compartment was 900 [mu]bar, which is about 3 times higher than the value around Rubisco in mesophyll cells of wheat. A high [CO2] is maintained in the bundle sheath compartment in maize until Ci decreases below approximately 100 [mu]bar. The results from these gas exchange measurements indicate that photorespiration occurs in maize but that the rate is low unless the intercellular [CO2] is severely limited by stress.  相似文献   

16.
Wynn T 《Plant physiology》1981,68(6):1253-1256
A study was conducted on a C4 (Panicum maximum) and a C3 (Panicum bisulcatum) species to determine the nature of the dark release of 14CO2 with respect to its responses to changes in temperature and O2 tension during light CO2 uptake of 14CO2.  相似文献   

17.
18.
A series of competitive, reversible cathepsin S (CatS) inhibitors was investigated. An earlier disclosure detailed the discovery of the 4-(2-keto-1-benzimidazolinyl)-piperidin-1-yl moiety as an effective replacement for the 4-arylpiperazin-1-yl group found in our screening hit. Continued investigation into replacements for the 4-aryl piperazine resulted in the identification of potentially useful CatS inhibitors with enzymatic and cellular activity similar to that of JNJ 10329670 as disclosed in a previous publication.  相似文献   

19.
Salicylaldoxime (2 × 10−3m and less) inhibits cyclic photophosphorylation in intact Chlorella cells severely whereas photosynthetic O2-evolution and 14CO2-fixation is hardly affected. Cyclic photophosphorylation in vivo was measured by following anaerobic light dependent glucose uptake. A similar difference in susceptibility has been observed with carbonylcyanide-p-trifluoromethoxyphenylhydrazone. Various controls exclude the possibility that the difference in inhibition was caused by differing experimental conditions or, in the case of glucose assimilation, by an inhibition of a reaction other than photophosphorylation.  相似文献   

20.
The new two-breath CO(2) method was employed to test the hypotheses that small alterations in arterial P(CO(2)) had an impact on the magnitude and dynamic response time of the CO(2) effect on cerebrovascular resistance (CVRi) and the dynamic autoregulatory response to fluctuations in arterial pressure. During a 10-min protocol, eight subjects inspired two breaths from a bag with elevated P(CO(2)), four different times, while end-tidal P(CO(2)) was maintained at three levels: hypocapnia (LoCO(2), 8 mmHg below resting values), normocapnia, and hypercapnia (HiCO(2), 8 mmHg above resting values). Continuous measurements were made of mean blood pressure corrected to the level of the middle cerebral artery (BP(MCA)), P(CO(2)) (estimated from expired CO(2)), and mean flow velocity (MFV, of the middle cerebral artery by Doppler ultrasound), with CVRi = BP(MCA)/MFV. Data were processed by a system identification technique (autoregressive moving average analysis) with gain and dynamic response time of adaptation estimated from the theoretical step responses. Consistent with our hypotheses, the magnitude of the P(CO(2))-CVRi response was reduced from LoCO(2) to HiCO(2) [from -0.04 (SD 0.02) to -0.01 (SD 0.01) (mmHg x cm(-1) x s) x mmHg Pco(2)(-1)] and the time to reach 95% of the step plateau increased from 12.0 +/- 4.9 to 20.5 +/- 10.6 s. Dynamic autoregulation was impaired with elevated P(CO(2)), as indicated by a reduction in gain from LoCO(2) to HiCO(2) [from 0.021 +/- 0.012 to 0.007 +/- 0.004 (mmHg x cm(-1) x s) x mmHg BP(MCA)(-1)], and time to reach 95% increased from 3.7 +/- 2.8 to 20.0 +/- 9.6 s. The two-breath technique detected dependence of the cerebrovascular CO(2) response on P(CO(2)) and changes in dynamic autoregulation with only small deviations in estimated arterial P(CO(2)).  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