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 109 毫秒
1.
Kinetic properties of cyanase   总被引:6,自引:0,他引:6  
P M Anderson  R M Little 《Biochemistry》1986,25(7):1621-1626
Cyanase is an inducible enzyme in Escherichia coli that catalyzes the hydrolysis of cyanate. Bicarbonate is required for activity, perhaps as a substrate, and the initial product of the reaction is carbamate, which spontaneously breaks down to ammonia and bicarbonate [Anderson, P. M. (1980) Biochemistry 19, 2882]. The purpose of this study was to characterize the kinetic properties of cyanase. Initial velocity studies showed that both cyanate and bicarbonate act as competitive substrate inhibitors. A number of monovalent anions act as inhibitors. Azide and acetate appear to act as competitive inhibitors with respect to cyanate and bicarbonate, respectively. Chloride, bromide, nitrate, nitrite, and formate also inhibit, apparently as the result of binding at either substrate site. Malonate and several other dicarboxylic dianions at very low concentrations display "slow-binding", reversible inhibition which can be prevented by saturating concentrations of either substrate. The results are consistent with a rapid equilibrium random mechanism in which bicarbonate acts as a substrate, bicarbonate and cyanate bind at adjacent anion-binding sites, and both substrates can bind at the other substrate anion binding site to give a dead-end complex.  相似文献   

2.
Determination of whether CO2 or HCO3- is the substrate for an enzymatic carboxylation has generally been accomplished by taking advantage of the fact that equilibration of these two compounds requires more than a minute at temperatures below 15 degrees C; thus different kinetics of carboxylation are obtained depending on whether CO2 or HCO3- is used to initiate the reaction. We report a new method using 13C18O2 as substrate for determining the CO2/HCO3- specificity of carboxylases. If CO2 is the substrate, then the 18O content of the 13C-containing product is the same as that of the 13CO2 used, whereas if HCO3- is the substrate, the 18O content is 2/3 that of the starting material. The method is independent of the detailed kinetics of the CO2/HCO3- interconversion and independent of the presence of contaminating unlabeled CO2 or HCO3-. Isotopic analysis is accomplished by 13C NMR. The method has been used to confirm that HCO3- is the substrate for phosphoenolpyruvate carboxylase. Studies of oxygen-18 isotope shifts in phosphorus NMR spectra have permitted confirmation of the observation that label is transferred from HC18O3- into Pi during the carboxylation of phosphoenolpyruvate.  相似文献   

3.
With physiological portal HCO3- and CO2 concentrations of 25mM and 1.2mM in the perfusate, respectively, acetazolamide inhibited urea synthesis from NH4Cl in isolated perfused rat liver by 50-60%, whereas urea synthesis from glutamine was inhibited by only 10-15%. A decreased sensitivity of urea synthesis from glutamine to acetazolamide inhibition was also observed when the extracellular HCO3- and CO2 concentrations were varied from 0-50mM and 0-2.4mM, respectively. Stimulation of intramitochondrial CO2 formation at pyruvate dehydrogenase with high pyruvate concentrations (7mM) was without effect on the acetazolamide sensitivity of urea synthesis from NH4Cl. Urea synthesis was studied under conditions of a limiting HCO3- supply for carbamoyl-phosphate synthesis. In the absence of externally added HCO3- or CO2, when 14CO2 was provided intracellularly by [U-14C]glutamine or [1-14C]-glutamine oxidation, acetazolamide had almost no effect on label incorporation into urea, whereas label incorporation from an added tracer H14CO3- dose was inhibited by about 70%. 14CO2 production from [U-14C]glutamine was about twice as high as from [1-14C]glutamine, indicating that about 50% of the CO2 produced from glutamine is formed at 2-oxoglutarate dehydrogenase. The fractional incorporation of 14CO2 into urea was about 13% with [1-14C]-as well as with [U-14C]glutamine. Addition of small concentrations of HCO3- (1.2mM) to the perfusate increased urea synthesis from glutamine by about 70%. This stimulation of urea synthesis was fully abolished by acetazolamide. The carbonate-dehydratase inhibitor prevented the incorporation of added HCO3- into urea, whereas incorporation of CO2 derived from glutamine degradation was unaffected. Without HCO3- and CO2 in the perfusion medium, when 14CO2 was provided by [1-14C]-pyruvate oxidation, acetazolamide inhibited urea synthesis from NH4Cl as well as 14C incorporation into urea by about 50%. Therefore carbonate-dehydratase activity is required for the utilization of extracellular CO2 or pyruvate-dehydrogenase-derived CO2 for urea synthesis, but not for CO2 derived from glutamine oxidation. This is further evidence for a special role of glutamine as substrate for urea synthesis.  相似文献   

4.
C K Tu  D N Silverman 《Biochemistry》1985,24(21):5881-5887
We have measured the catalysis by Co(II)-substituted bovine carbonic anhydrase II from red cells of the exchange of 18O between CO2 and H2O using membrane-inlet mass spectrometry. We chose Co(II)-substituted carbonic anhydrase II because the apparent equilibrium dissociation constant of HCO3- and enzyme at pH 7.4, KHCO3-eff approximately equal to 55 mM, was within a practicable range of substrate concentrations for the 18O method. For the native, zinc-containing enzyme KHCO3-eff is close to 500 mM at this pH. The rate constant for the release from the active site of water bearing substrate oxygen kH2O was dependent on the fraction of enzyme that was free, not bound by substrate HCO3- or anions. The pH dependence of kH2O in the pH range 6.0-9.0 can be explained entirely by a rate-limiting, intramolecular proton transfer between cobalt-bound hydroxide and a nearby group, probably His-64. The rate constant for this proton transfer was found to be 7 X 10(5) S-1 for the Co(II)-substituted enzyme and 2 X 10(6) S-1 for the native enzyme. These results are applied to models derived from proton-relaxation enhancement of water exchanging from the inner coordination shell of the cobalt in carbonic anhydrase. The anions iodide, cyanate, and thiocyanate inhibited catalysis of 18O exchange by Co(II)-substituted carbonic anhydrase II in a manner competitive with total substrate (CO2 and HCO3-) at chemical equilibrium and pH 7.4. These results are discussed in terms of observed steady-state inhibition patterns and suggest that there is no significant contribution of a ternary complex between substrate, inhibitor, and enzyme.  相似文献   

5.
Pseudomonas fluorescens NCIB 11764 was capable of utilizing cyanate (OCN-) as a sole nitrogen source for growth. Crude cell extracts from cells grown on cyanate, but not on ammonium sulfate, were induced for an enzyme catalyzing cyanate conversion to ammonia. Enzymatic activity was shown to be bicarbonate dependent and specific for cyanate as a substrate, suggesting that cyanate utilization in this organism is facilitated by an enzyme resembling cyanase (cyanate amidohydrolase; EC 3.5.5.3), as described previously in Escherichia coli and Flavobacterium sp.  相似文献   

6.
Cyanase is an inducible enzyme in Escherichia coli that catalyzes the reaction of cyanate with bicarbonate to give two CO2 molecules. The gene for cyanase is part of the cyn operon, which includes cynT and cynS, encoding carbonic anhydrase and cyanase, respectively. Carbonic anhydrase functions to prevent depletion of cellular bicarbonate during cyanate decomposition (the product CO2 can diffuse out of the cell faster than noncatalyzed hydration back to bicarbonate). Addition of cyanate to the culture medium of a delta cynT mutant strain of E. coli (having a nonfunctional carbonic anhydrase) results in depletion of cellular bicarbonate, which leads to inhibition of growth and an inability to catalyze cyanate degradation. These effects can be overcome by aeration with a higher partial CO2 pressure (M. B. Guilloton, A. F. Lamblin, E. I. Kozliak, M. Gerami-Nejad, C. Tu, D. Silverman, P. M. Anderson, and J. A. Fuchs, J. Bacteriol. 175:1443-1451, 1993). The question considered here is why depletion of bicarbonate/CO2 due to the action of cyanase on cyanate in a delta cynT strain has such an inhibitory effect. Growth of wild-type E. coli in minimal medium under conditions of limited CO2 was severely inhibited, and this inhibition could be overcome by adding certain Krebs cycle intermediates, indicating that one consequence of limiting CO2 is inhibition of carboxylation reactions. However, supplementation of the growth medium with metabolites whose syntheses are known to depend on a carboxylation reaction was not effective in overcoming inhibition related to the bicarbonate deficiency induced in the delta cynT strain by addition of cyanate. Similar results were obtained with a deltacyn strain (since cyanase is absent, this strain does not develop a bicarbonate deficiency when cyanate is added); however, as with the deltacynT strain, a higher partial CO(2) pressure in the aerating gas or expression of carbonic anhydrase activity (which contributes to a higher intercellular concentration of bicarbonate/CO(2)) significantly reduced inhibition of growth. There appears to be competition between cyanate and bicarbonate/CO(2) at some unknown but very important site such that cyanate binding inhibits growth. These results suggest that bicarbonate/CO(2) plays a significant role in the growth of E. coli other than simply as a substrate for carboxylation reactions and that strains with mutations in the cyn operon provide a unique model system for studying aspects of the metabolism of bicarbonate/CO(2) and its regulation in bacteria.  相似文献   

7.
The effects of ethoxyzolamide (EZ), a carbonic anhydrase inhibitor, on the active CO2 and Na+-independent and Na+-dependent HCO3- transport systems of the unicellular cyanobacterium Synechococcus sp. UTEX 625 were examined. Measurements of transport and accumulation using radiochemical, fluorometric, and mass spectrometric assays indicated that active CO2 transport and active Na+-independent HCO3- transport were inhibited by EZ. However, Na+-independent HCO3- transport was about 1 order of magnitude more sensitive to EZ inhibition than was CO2 transport (50% inhibition = 12 [mu]M versus 80 [mu]M). The data suggest that both the active CO2 (G.D. Price, M.R. Badger [1989] Plant Physiol 89: 37-43) and the Na+ -independent HCO3 - transport systems possessed carbonic anhydrase-like activity as part of their mechanism of action. In contrast, Na+-dependent HCO3- transport was only partially (50% inhibition = 230 [mu]M) and noncompetitively inhibited by EZ. The collective evidence suggested that EZ inhibition of Na+ -dependent HCO3- transport was an indirect consequence of the action of EZ on the CO2 transport system, rather than a direct effect on HCO3- transport. A model is presented in which the core of the inorganic carbon translocating system is formed by Na+-dependent HCO3- transport and the CO2 transport system. It is argued that the Na+-independent HCO3 - utilizing system was not directly involved in translocation, but converted HCO3- to CO2 for use in CO2 transport.  相似文献   

8.
The 13C NMR spectra of the pheophorbide of bacteriochlorophyll c, formed in the presence of L-[1-13C]glutamate and [2-13C]glycine and [13C]bicarbonate in Prosthecochloris aestaurii, were analysed. The isotope in the glutamate was specifically incorporated into the eight carbon atoms in the tetrapyrrole macrocycle derived from the C-5 of 5-aminolevulinic acid, while no specific enrichment of these eight carbon atoms was observed in the spectrum of the pigment formed in the presence of [2-13C]glycine. These labelling patterns provide evidence for the operation of the C5 pathway of 5-aminolevulinic acid synthesis for bacteriochlorophyll c formation in the bacterium. The labelling of bacteriochlorophyll c by [13C]bicarbonate is consistent with its formation from 5-[1,4,5-13C]aminolevulinic acid formed by the C5 pathway from [1,2,5-13C]glutamic acid. It is proposed that this glutamate is the transamination product of 2-[1,2,5-13C]oxoglutaric acid, arising by carboxylation of [1,4-13C]succinyl-CoA with 13CO2 catalysed by 2-oxoglutaric acid synthase activity, and that the labelled succinyl-CoA is, in turn, derived by the fixation of 13CO2 by the reductive tricarboxylic acid cycle. The 13C chemical shifts of two sp2 quaternary carbons of bacteriopheophorbide c methyl ester (C-2 and C-4) were reassigned.  相似文献   

9.
The cyanobacteria Synechococcus elongatus strain PCC7942 and Synechococcus sp. strain UTEX625 decomposed exogenously supplied cyanate (NCO-) to CO2 and NH3 through the action of a cytosolic cyanase which required HCO3- as a second substrate. The ability to metabolize NCO- relied on three essential elements: proteins encoded by the cynABDS operon, the biophysical activity of the CO2-concentrating mechanism (CCM), and light. Inactivation of cynS, encoding cyanase, and cynA yielded mutants unable to decompose cyanate. Furthermore, loss of CynA, the periplasmic binding protein of a multicomponent ABC-type transporter, resulted in loss of active cyanate transport. Competition experiments revealed that native transport systems for CO2, HCO3-, NO3-, NO2-, Cl-, PO4(2-), and SO4(2-) did not contribute to the cellular flux of NCO- and that CynABD did not contribute to the flux of these nutrients, implicating CynABD as a novel primary active NCO- transporter. In the S. elongatus strain PCC7942 DeltachpX DeltachpY mutant that is defective in the full expression of the CCM, mass spectrometry revealed that the cellular rate of cyanate decomposition depended upon the size of the internal inorganic carbon (Ci) (HCO3- + CO2) pool. Unlike wild-type cells, the rate of NCO- decomposition by the DeltachpX DeltachpY mutant was severely depressed at low external Ci concentrations, indicating that the CCM was essential in providing HCO3- for cyanase under typical growth conditions. Light was required to activate and/or energize the active transport of both NCO- and Ci. Putative cynABDS operons were identified in the genomes of diverse Proteobacteria, suggesting that CynABDS-mediated cyanate metabolism is not restricted to cyanobacteria.  相似文献   

10.
In this study, it is shown how to transfer tared aliquots of (HCO3 + CO2)-containing luminal fluids directly into the mercury-sealed chamber of a modified Van Slyke apparatus and how to obtain direct as well as indirect manometric determinations of dissolved CO2 ([CO2]f) in each aliquot of such fluids. It is next shown that the pattern of in vitro luminal acidification in an isolated turtle bladder sac depends upon the prior in vivo ambient temperature to which the donor turtle had become adapted. Under in vivo conditions, the food intake, physical activity, and acid excretion of 32 degrees C-adapted turtles are greater than those of 21 degrees C or 26 degrees C-adapted turtles. Under in vitro conditions of incubating isolated bladder sacs (from 21, 26, and 32 degrees C turtles) in (HCO3 + CO2)-containing Ringer media at a single temperature (21 degrees C), the patterns of luminal acidification are as follows: (a) The rate of depletion of luminal [HCO3] is greatest in bladders from the 32 degrees C-adapted turtles. (b) Concomitant decreases in luminal [CO2]f, [HCO3], and pH (the 'CO2-decreasing patterns' of luminal acidification) develop in all bladders from 32 degrees C turtles, in half of those from 26 degrees C turtles, but in less than one-fifth of those from 21 degrees C-adapted turtles: and (c) a CO2-increasing pattern of luminal acidification is found in most of the bladders from 21 degrees C-adapted turtles. A postulated bicarbonate ion-reabsorbing pump is consistent with all of these patterns of luminal acidification.  相似文献   

11.
At pH 7.4, CO2, rather than HCO3-, markedly enhances the oxidation of diverse substrates by SOD1 plus H2O2. Since the concentration of CO2 would fall with rising pH in HCO3- buffers, it was of interest to explore the effects of pH on the peroxidase activity of SOD1 in the presence and in the absence of HCO3-. The rate of NADPH peroxidation in the HCO3- buffer was minimally affected by pH in the range of 8-10.5; in a pyrophosphate buffer, the rate increased markedly, such that at pH 10.5 the rates in the two buffers were nearly identical. Similar results were obtained when urate was used as the peroxidizeable substrate. These results are explicable on the basis of an increase in the rate with pH due to the ionization of H2O2 to the effective HO2- coupled with a decrease in [CO2] due to the ionizations of H2CO3, which displaces the hydration equilibrium to the right. These two opposing effects counteract in the HCO3(-)-buffered reaction mixtures; in the pyrophosphate buffer, only the effect of increasing [H02-] was seen.  相似文献   

12.
This study addresses the mechanism of covalent aggregation of human Cu,Zn-superoxide dismutase (hSOD1WT) induced by bicarbonate (HCO3-)-mediated peroxidase activity. Higher molecular weight species (apparent dimers and trimers) of hSOD1WT were formed from incubation mixtures containing hSOD1WT, H2O2, and HCO3-. HCO3--dependent peroxidase activity and covalent aggregation of hSOD1WT were mimicked by UV photolysis of hSOD1-WT in the presence of a [Co(NH3)5CO3]+ complex that generates the carbonate radical anion (CO3.). Human SOD1WT has but one aromatic residue, a tryptophan residue (Trp-32) on the surface of the protein. Substitution of Trp-32 with phenylalanine produced a mutant (hSOD1W32F) that exhibits HCO3--dependent peroxidase activity similar to wild-type enzyme. However, unlike hSOD1WT, incubations containing hSOD1W32F,H2O2, and HCO3-did not result in covalent aggregation of SOD1. These findings indicate that Trp-32 is crucial for CO3.-induced covalent aggregation of hSOD1WT. Spin-trapping results revealed the formation of the Trp-32 radical from hSOD1WT, but not from hSOD1W32F. Spin traps also inhibited the covalent aggregation of hSOD1WT. Fluorescence experiments revealed that Trp-32 was further oxidized by CO3., forming kynurenine-type products in the presence of oxygen. Molecular oxygen was needed for HCO3-/H2O2-dependent aggregation of hSOD1WT, implicating a role for a Trp-32-dependent peroxidative reaction in the covalent aggregation of hSOD1WT. Taken together, these results indicate that Trp-32 oxidation is crucial for covalent aggregation of hSOD1. Implications of HCO3--dependent SOD1 peroxidase activity in amyotrophic lateral sclerosis disease are discussed.  相似文献   

13.
R W Clark  M Volpi  R D Berlin 《Biochemistry》1988,27(3):1025-1033
Carbamoylation and reductive methylation of tubulin have been shown previously to inhibit microtubule assembly, probably by attack on essential internal lysine residues [Mellado, W., Slebe, J., & Maccioni, R.B. (1982) Biochem. J. 203, 675-681; Szasz, J., Burns, R., & Sternlicht, H. (1982) J. Biol. Chem. 257, 3697-3704]. We show first that this inhibition is blocked by the presence of HCO3-/CO2 buffer at physiological concentrations during the carbamoylation or reductive methylation. Under conditions that block assembly, the amount of radiolabeled cyanate or formaldehyde incorporated by these reactions in the absence of HCO3-/CO2 was approximately four carbamoyl or five methyl groups in a ratio of approximately 1.7 alpha chain/beta chain. In the presence of HCO3-/CO2, the formaldehyde incorporation is decreased roughly 0.5 mol in each of the alpha and beta chains, and cyanate incorporation, roughly 1.0 mol/mol of alpha or beta monomer. These results are consistent with the hypothesis that CO2 competed with formaldehyde or cyanate for uncharged amino groups and led to the reversible formation of carbamates. The complete antagonism of the inhibition of microtubule assembly by reductive methylation by CO2, even though the number of methyl groups incorporated was reduced by only 0.5 mol/tubulin monomer, was consistent with the possibility that reductive methylation opened up additional residues for attack. Indeed, using an adaptation of the method of Gros et al. for measurement of carbamates [Gros, G., Forster, R.E., & Lin, L. (1976) J. Biol. Chem. 251, 4398-4407], we found that reductive methylation with 2 mM formaldehyde (assembly blocked) did not decrease carbamate formation (carbamate formation was inhibited at higher formaldehyde concentrations).(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

14.
Products formed from the lactoperoxidase (LPO) catalyzed oxidation of thiocyanate ion (SCN-) with hydrogen peroxide (H2O2) have been studied by 13C-NMR at pH 6 and pH 7. Ultimate formation of hypothiocyanite ion (OSCN-) as the major product correlates well with the known optical studies. The oxidation rate of SCN- appears to be greater at pH < or = 6.0. At [H2O2]/[SCN-] ratios of < or = 0.5, OSCN- is not formed immediately, but an unidentified intermediate is produced. At [H2O2]/[SCN-] > 0.5, SCN- appears to be directly oxidized to OSCN-. Once formed, OSCN- slowly degrades over a period of days to carbon dioxide (CO2), bicarbonate ion (HCO3-), and hydrogen cyanide (HCN). An additional, previously unrecognized product also appears after formation of OSCN-. On the basis of carbon-13 chemical shift information this new species is suggested to result from rearrangement of OSCN- to yield the thiooxime isomer, SCNO- or SCNOH.  相似文献   

15.
The Cu,Zn SOD catalyzes the bicarbonate-dependent oxidation of a wide range of substrates by H2O2. A mechanism in accord with this activity has been described. It involves the generation of a strong oxidant (Cu(I)O, Cu(II)OH, or Cu(III)) by reaction of the active site Cu with H2O2, followed by oxidation of bicarbonate to CO3-* that in turn diffuses from the active site to oxidize the various substrates in free solution. Recently, an alternative mechanism, entailing firmly bound HCO3- and CO3-*, has been proposed [J. Biol. Chem. 278 (2003) 21032-21039]. We present data supporting the diffusible CO3-* and discuss the properties of this system that can be accommodated in this way and that preclude bound intermediates.  相似文献   

16.
J J Led  E Neesgaard 《Biochemistry》1987,26(1):183-192
A detailed analysis of the stability and activity of Mn(II) human carbonic anhydrase I and the kinetics and mechanism of its catalysis of the HCO3-/CO2 exchange have been performed at pH 8.5. The analysis was based on the paramagnetic relaxation rates R1p and R2p of the 13C atom of HCO3- in the Mn2+/apoenzyme/HCO3-/CO2 system and the HCO3(-)----CO2 interconversion rate obtained by the magnetization-transfer technique. The R1p and R2p rates were measured as functions of the temperature, magnetic field strength, and substrate and apoenzyme concentrations and were interpreted on the basis of the Solomon-Bloembergen-Morgan theories and general equations for the ligand exchange [Led, J. J., & Grant, D. M. (1977) J. Am. Chem. Soc. 99, 5845-5858]. From the analysis of the data, a formation constant for the Mn(II) enzyme of log KMAM = 5.8 +/- 0.4 was obtained while the activity of the Mn(II) enzyme, measured as the HCO3-/CO2 interconversion rate at [HCO3-] = 0.100 M and pH 8.5, was found to be about 4% of that of the native Zn(II) enzyme. However, an effective dissociation constant KeffHCO3- less than or approximately 12 mM and a maximal exchange rate constant kcatexch approximately equal to 400 s-1, also derived by the analysis, result in an apparent second-order rate constant kcatexch/KeffHCO3- only a factor of 4 smaller than the corresponding rate constant for the native Zn(II) isoenzyme I. Most conspicuously, the resulting distance of only 2.71 +/- 0.03 A between the Mn2+ ion of the enzyme and the 13C atom of HCO3- in the enzyme-bicarbonate complex indicates that the bicarbonate is bound to the metal ion by two of its oxygen atoms in the central catalytic step, thereby supporting the modified Zn(II)-OH mechanism [Lindskog, S., Engberg, P., Forsman, C., Ibrahim, S. A., Jonsson, B.-H., Simonsson, I., & Tibell, L. (1984) Ann. N.Y. Acad. Sci. 429, 61-75 (and references cited therein)]. In contrast, this binding mode differs from the structure of the complexes suggested in the rapid-equilibrium kinetic model [Pocker, Y., & Deits, T. L. (1983) J. Am. Chem. Soc. 105, 980-986; Pocker, Y., & Deits, T. L. (1984) Ann. N.Y. Acad. Sci. 429, 76-83].  相似文献   

17.
J W Janc  W W Cleland  M H O'Leary 《Biochemistry》1992,31(28):6441-6446
Formate is an alternate substrate for bicarbonate in the reaction with PEP catalyzed by phosphoenolpyruvate carboxylase from Zea mays, producing formyl phosphate and pyruvate. The Km for formate is 25 +/- 2 mM, and the maximum velocity is 1% of that for bicarbonate at pH 8.0. Use of [18O]formate produces inorganic phosphate containing 1 equiv of 18O, but no label is incorporated into residual phosphoenolpyruvate. PEP carboxylase catalyzes the hydrolysis of phosphoglycolate or L-phospholactate 2000 times more slowly and D-phospholactate 4000 times more slowly than the reaction between bicarbonate and PEP.  相似文献   

18.
It is widely accepted that the oxygen produced by photosystem II of cyanobacteria, algae, and plants is derived from water. Earlier proposals that bicarbonate may serve as substrate or catalytic intermediate are almost forgotten, though not rigorously disproved. These latter proposals imply that CO2 is an intermediate product of oxygen production in addition to O2. In this work, we investigated this possible role of exchangeable HCO3- in oxygen evolution in two independent ways. (1) We studied a possible product inhibition of the electron transfer into the catalytic Mn4Ca complex during the oxygen-evolving reaction by greatly increasing the pressure of CO2. This was monitored by absorption transients in the near UV. We found that a 3,000-fold increase of the CO2 pressure over ambient conditions did not affect the UV transient, whereas the S(3) --> S(4) --> S(0) transition was half-inhibited by raising the O2 pressure only 10-fold over ambient, as previously established. (2) The flash-induced O2 and CO2 production by photosystem II was followed simultaneously with membrane inlet mass spectrometry under approximately 15% H2(18)O enrichment. Light flashes that revealed the known oscillatory O2 release failed to produce any oscillatory CO2 signal. Both types of results exclude that exchangeable bicarbonate is the substrate for (and CO2 an intermediate product of) oxygen evolution by photosynthesis. The possibility that a tightly bound carbonate or bicarbonate is a cofactor of photosynthetic water oxidation has remained.  相似文献   

19.
The composition and catalase-like activity of Mn2+ complexes with bicarbonate were investigated with voltammetry and kinetic methods (by the rate of O2 production from H2O2). Three linear sections were revealed on the dependence of the reduction potential of Mn2+ on logarithm of bicarbonate concentration (logC(NaHCO3)) having slopes equal to 0 mV/logC(NaHCO3), -14 mV/logC(NaHCO3), and -59 mV/logC(NaHCO3), corresponding to Mn2+ aqua complex (Mn2+(aq)) and to Mn2+-bicarbonate complexes of the composition [Mn2+(HCO3(-))]+ (at concentration of HCO3(-) 10-100 mM) and [Mn2+(HCO3(-))2]0 (at concentration of HCO3(-) 100-600 mM). Comparison of HCO3(-) concentration needed for the catalase-like activity of Mn2+ with the electrochemical data showed that only electroneutral complex Mn2+(HCO3(-))2 catalyzed decomposition of H2O2, whereas positively charged Mn2+(aq) complex and [Mn2+(HCO3(-))]+ were not active. The catalase-like activity of Mn2+ did not appear upon substitution of anions of carbonic acids (acetate and formate) for HCO3(-). The rate of O2 production in the system Mn2+-HCO3(-)-H2O2 (pH 7.4) is proportional to the second power of Mn2+ concentration and to the fourth power of HCO3(-) concentration that indicates simultaneous involvement of two Mn2+(HCO3(-))2 complexes in the reaction of H2O2 decomposition.  相似文献   

20.
Spinach-leaf ribulose-5-phosphate kinase catalyzes the reaction of (Rp)-[beta, gamma-18O, gamma-18O]adenosine 5'-(3-thiotriphosphate) with ribulose 5-phosphate to form ribulose 1-[18O]phosphorothioate 5-phosphate. This product is incubated with CO2, Mg2+, and ribulose-bisphosphate carboxylase to form the [18O]phosphorothioate of D-glycerate. Reduction of this material using phosphoglycerate kinase/ATP, glyceraldehyde-3-phosphate dehydrogenase/NADH, triose-phosphate isomerase, and glycerol-phosphate dehydrogenase/NADH produces glycerol 3-[18O]phosphorothioate, which is subjected to ring closure using diethylphosphorochloridate. This in-line reaction produces a diastereoisomeric mixture of glycerol 2,3-cyclic phosphorothioates. 31P NMR spectroscopy was used to analyze the 18O content of the products. The anti-diastereoisomer, which is the major isomer formed and corresponds to the downfield 31P NMR signal (Pliura, D.H., Schomburg, D., Richard, J.P., Frey, P.A., and Knowles, J.R. (1980) Biochemistry 19, 325-329), retains the 18O label. This observation indicates that the ribulose-5-phosphate kinase reaction proceeds with inversion of configuration at phosphorus. The reaction is, therefore, unlikely to involve the participation of a covalent phosphoryl-enzyme intermediate.  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