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.
Bicarbonate is a recycling substrate for cyanase   总被引:1,自引:0,他引:1  
Cyanase is an inducible enzyme in Escherichia coli that catalyzes bicarbonate-dependent decomposition of cyanate to ammonia and bicarbonate. Previous studies provided evidence that carbamate is an initial product and that the kinetic mechanism is rapid equilibrium random (bicarbonate serving as substrate as opposed to activator); the following mechanism was proposed (Anderson, P. M. (1980) Biochemistry 19, 2282-2888; Anderson, P. M., and Little, R. M. (1986) Biochemistry 25, 1621-1626). (formula; see text) Direct evidence for this mechanism was obtained in this study by 1) determining whether CO2 or HCO3- serves as substrate and is formed as product, 2) identifying the products formed from [14C]HCO3- and [14C] OCN-, 3) identifying the products formed from [13C] HCO3- and [12C]OCN- in the presence of [18O]H2O, and 4) determining whether 18O from [18O]HCO3- is incorporated into CO2 derived from OCN-. Bicarbonate (not CO2) is the substrate. Carbon dioxide (not HCO3-) is produced in stoichiometric amounts from both HCO3- and OCN-. 18O from [18O]H2O is not incorporated into CO2 formed from either HCO3- or OCN-. Oxygen-18 from [18O]HCO3- is incorporated into CO2 derived from OCN-. These results support the above mechanism, indicating that decomposition of cyanate catalyzed by cyanase is not a hydrolysis reaction and that bicarbonate functions as a recycling substrate.  相似文献   

2.
1. The steady-state kinetics of the interconversion of CO2 and HCO3 catalyzed by human carbonic anhydrase C was studied using 1H2O and 2H2O as solvents. The pH-independent parts of the parameters k(cat) and Km are 3-4 times larger in 1H2O than in 2H2O for both directions of the reaction, while the ratios k(cat)/Km show much smaller isotope effects. With either CO2 or HCO3 as substrate the major pH dependence is observed in k(cat), while Km appears independent of pH. The pKa value characterizing the pH-rate profiles is approximately 0.5 unit larger in 2H2O than in 1H2O. 2. The hydrolysis of p-nitrophenyl acetate catalyzed by human carbonic anhudrase C is approximately 35% faster in 2H2O than in 1H2O. In both solvents the pKa values of the pH-rate profiles are similar to those observed for the CO2-HCO3 interconversion. 3. It is tentatively proposed that the rate-limiting step at saturating concentrations of CO2 or HCO3 is an intramolecular proton transfer between two ionizing groups in the active site. It cannot be decided whether the transformation between enzyme-bound CO2 and HCO3 involves a proton trnasfer or not.  相似文献   

3.
S H Hwang  T Nowak 《Biochemistry》1986,25(19):5590-5595
The stereochemistry of the carboxylation of phosphoenolpyruvate to yield oxalacetate, catalyzed by chicken liver phosphoenolpyruvate carboxykinase and by Ascaris muscle phosphoenolpyruvate carboxykinase, was determined. The substrate (Z)-3-fluorophosphoenolpyruvate was used for the stereochemical analysis. The carboxylation reaction was coupled to malate dehydrogenase to yield 3-fluoromalate, and the stereochemistry of the products was identified by 19F NMR. In separate experiments, the enantiomeric tautomers of 3-fluorooxalacetate were shown to be utilized by malate dehydrogenase to yield (2R,3R)- and (2R,3S)-3-fluoromalate in nearly identical amounts. The products were identified by 19F NMR. When (Z)-3-fluorophosphoenolpyruvate was used as a substrate for phosphoenolpyruvate carboxykinase from avian liver and from Ascaris, and malate dehydrogenase was used to trap the product, only a single diastereomer was observed. This product was shown to be (2R,3R)-3-fluoromalate in each case. The assignments were based on coupling constants taken from Keck et al. [Keck, R., Hess, H., & Rétey, J. (1980) FEBS Lett. 114, 287]. These results indicate that the stereochemistry of carboxylation, catalyzed by chicken phosphoenolpyruvate carboxykinase and by Ascaris phosphoenolpyruvate carboxykinase, is identical and takes place from the si side of the enzyme-bound phosphoenolpyruvate. The carboxylation reaction was run both in H2O and in D2O. No deuterium incorporation into fluoromalate was shown to occur. The product 3-fluorooxalacetate is thus released from phosphoenolpyruvate carboxykinase as the keto form and is reduced more rapidly by reduced nicotinamide adenine dinucleotide with malate dehydrogenase than by the occurrence of tautomerization.  相似文献   

4.
The catalytic mechanism of phosphoenolpyruvate (PEP) carboxylase from Zea mays has been studied using (Z)- and (E)-3-fluorophosphoenolpyruvate (F-PEP) as substrates. Both (Z)- and (E)-F-PEP partition between carboxylation to produce 3-fluorooxalacetate and hydrolysis to produce 3-fluoropyruvate. Carboxylation accounts for 3% of the reaction observed with (Z)-F-PEP, resulting in the formation of (R)-3-fluorooxalacetate, and for 86% of the reaction of (E)-F-PEP forming (S)-3-fluorooxalacetate. Carboxylation of F-PEP occurs on the 2-re face, which corresponds to the 2-si face of PEP. The partitioning of F-PEP between carboxylation and hydrolysis is insensitive to pH but varies with metal ion. Use of 18O-labeled bicarbonate produces phosphate that is multiply labeled with 18O; in addition, 18O is also incorporated into residual (Z)- and (E)-F-PEP. The 13(V/K) isotope effect on the carboxylation of F-PEP catalyzed by PEP carboxylase at pH 8.0, 25 degrees C, is 1.049 +/- 0.003 for (Z)-F-PEP and 1.009 +/- 0.006 for (E)-F-PEP. These results are consistent with a mechanism in which carboxylation of PEP occurs via attack of the enolate of pyruvate on CO2 rather than carboxy phosphate. In this mechanism phosphorylation of bicarbonate to give carboxy phosphate and decarboxylation of the latter are reversible steps. An irreversible step, however, precedes partitioning between carboxylation to give oxalacetate and release of CO2, which results in hydrolysis of PEP.  相似文献   

5.
1. A simple kinetic method was devised to show whether dissolved CO(2) or HCO(3)- ion is the substrate in enzyme-catalysed carboxylation reactions. 2. The time-course of the reductive carboxylation of 2-oxoglutarate by NADPH, catalysed by isocitrate dehydrogenase, was studied by a sensitive fluorimetric method at pH7.3 and pH6.4, with large concentrations of substrate and coenzyme and small carbon dioxide concentrations. 3. Reaction was initiated by the addition of carbon dioxide in one of three forms: (i) as the dissolved gas in equilibrium with bicarbonate; (ii) as unbuffered bicarbonate solution; (iii) as the gas or as an unbuffered solution of the gas in water. Different progress curves were obtained in the three cases. 4. The results show that dissolved CO(2) is the primary substrate of the enzyme, and that HCO(3)- ion is at best a very poor substrate. The progress curves are in quantitative agreement with this conclusion and with the known rates of the reversible hydration of CO(2) under the conditions of the experiments. The effects of carbonic anhydrase confirm the conclusions. 5. Similar experiments on the reductive carboxylation of pyruvate catalysed by the ;malic' enzyme show that dissolved CO(2) is the primary substrate of this enzyme also. 6. The results are discussed in relation to the mechanisms of these enzymes, and the effects of pH on the reactions. 7. The advantages of the method and its possible applications to other enzymes involved in carbon dioxide metabolism are discussed.  相似文献   

6.
The cmpABCD operon of the cyanobacterium Synechococcus sp. strain PCC 7942 encodes an ATP-binding cassette transporter involved in HCO(3)(-) uptake. The three genes, cmpBCD, encode membrane components of an ATP-binding cassette transporter, whereas cmpA encodes a 42-kDa cytoplasmic membrane protein, which is 46.5% identical to the membrane-anchored substrate-binding protein of the nitrate/nitrite transporter. Equilibrium dialysis analysis using H(14)CO(3)(-) showed that a truncated CmpA protein lacking the N-terminal 31 amino acids, expressed in Escherichia coli cells as a histidine-tagged soluble protein, specifically binds inorganic carbon (CO(2) or HCO(3)(-)). The addition of the recombinant CmpA protein to a buffer caused a decrease in the concentration of dissolved CO(2) because of the binding of inorganic carbon to the protein. The decrease in CO(2) concentration was accelerated by the addition of carbonic anhydrase, indicating that HCO(3)(-), but not CO(2), binds to the protein. Mass spectrometric measurements of the amounts of unbound and bound HCO(3)(-) in CmpA solutions containing low concentrations of inorganic carbon revealed that CmpA binds HCO(3)(-) with high affinity (K(d) = 5 microm). A similar dissociation constant was obtained by analysis of the competitive inhibition of the CmpA protein on the carboxylation of phosphoenolpyruvate by phosphoenolpyruvate carboxylase at limiting concentrations of HCO(3)(-). These findings showed that the cmpA gene encodes the substrate-binding protein of the HCO(3)(-) transporter.  相似文献   

7.
1. The carboxylation of pyruvate to oxaloacetate by pyruvate carboxylase in guinea-pig liver mitochondria was determined by measuring the amount of (14)C from H(14)CO(3) (-) fixed into organic acids in the presence of pyruvate, ATP, Mg(2+) and P(i). The main products of pyruvate carboxylation were malate, fumarate and citrate. Pyruvate utilization, metabolite formation and incorporation of (14)C from H(14)CO(3) (-) into these metabolites in the presence and the absence of ATP were examined. The synthesis of phosphoenolpyruvate from pyruvate and bicarbonate is minimal during continued oxidation of pyruvate. Larger amounts of phosphoenolpyruvate are formed from alpha-oxoglutarate than from pyruvate. Addition of glutamate, alpha-oxoglutarate or fumarate did not appreciably increase formation of phosphoenolpyruvate when pyruvate was used as substrate. With alpha-oxoglutarate as substrate addition of fumarate resulted in increased formation of phosphoenolpyruvate, whereas addition of succinate inhibited phosphoenolpyruvate formation. In the presence of added oxaloacetate guinea-pig liver mitochondria synthesized phosphoenolpyruvate in amount sufficiently high to play an appreciable role in gluconeogenesis. 2. Addition of fatty acids of increasing carbon chain length caused a strong inhibition of pyruvate oxidation and phosphoenolpyruvate formation, and greatly promoted carbon dioxide fixation and malate, citrate and acetoacetate accumulation. The incorporation of (14)C from H(14)CO(3) (-), [1-(14)C]pyruvate and [2-(14)C]pyruvate into organic acids formed was examined. 3. It is concluded that guinea-pig liver pyruvate carboxylase contributes significantly to gluconeogenesis and that fatty acids and metabolites play an important role in its regulation.  相似文献   

8.
The non-invasive technique of 13C nuclear magnetic resonance was applied to study glucose metabolism in vivo in the insect parasite Crithidia fasciculata. It was found that under anaerobic conditions [1-13C]glucose underwent a glycolytic pathway whose main metabolic products were identified as [2-13C]ethanol, [2-13C]succinate and [1,3-13C2]glycerol. These metabolites were excreted by C. fasciculata into the incubation medium, while in the cells [3-13C]phosphoenolpyruvate was also detected in addition to the aforementioned compounds. The C3 acid is apparently the acceptor of the primary CO2 fixation reaction, which leads in Trypanosomatids to the synthesis of succinate. By addition of sodium bicarbonate to the incubation mixture L-[3-13C]malate was detected among the excretion products, while the ethanol:succinate ratio of 2.0 in the absence of bicarbonate changed to a ratio of 0.6 in the presence of the latter. This was due to a shift of the balance between carboxylation of phosphoenolpyruvate, leading to succinate, and pyruvate decarboxylation leading to ethanol. The addition of 25% 2H2O to the incubation mixture led to the formation of [2-13C, 2-2H]ethanol derived from the prior incorporation of 2H+ into pyruvate in the reactions mediated by either pyruvate kinase or malic enzyme. However, no 2H+ incorporation into L-malate was detected, excluding the possibility that the latter was formed by carboxylation of pyruvate, and lending support to the idea that L-malate results from the carboxylation of phosphoenolpyruvate to oxaloacetate by phosphoenolpyruvate carboxykinase. The formation of [2-13C, 2-2H]-succinate under the same conditions reflected the uptake of 2H+ during the reduction of fumarate. When the incubations were carried out in the presence of 100% 2H2O, several [1-13C, 1-2H]ethanol species were detected, as well as [2-13C, 2-2H]malate and [1,3-13C2, 1,3-2H2]glycerol. The former deuterated compounds reflect the existence of NAD2H species when the incubations were carried out in 100% 2H2O, while the incorporation of 2H+ into [1,3-13C2]glycerol must be attributed to the phosphoglucose-isomerase-mediated reaction during glycolysis.  相似文献   

9.
The active species of "CO(2)" and the amount of fractionation of stable carbon isotopes have been determined for a partially purified preparation of phosphoenolpyruvate (PEP) carboxylase (EC 4.1.1.31) from corn (Zea mays) leaves. The rates of the enzyme reactions, using substrate amounts of HCO(3) (-), CO(2) or CO(2) plus carbonic anhydrase, show that HCO(3) (-) is the active species of "CO(2)" utilized by PEP carboxylase. The K(m) values for CO(2) and HCO(3) (-) are 1.25 mm and 0.11 mm, respectively, which further suggest the preferential utilization of HCO(3) (-) by PEP carboxylase. The amount of fractionation of stable carbon isotopes by PEP carboxylase from an infinite pool of H(12)CO(3) (-) and H(13)CO(3) (-) was -2.03 per thousand. This enzyme fractionation (delta), together with the fractionation associated with absorption of CO(2) into plant cells and the equilibrium fractionation associated with atmospheric CO(2) and dissolved HCO(3) (-) are discussed in relation to the fractionation of stable carbon isotopes of atmospheric CO(2) during photosynthesis in C(4) plants.  相似文献   

10.
13C-nuclear magnetic resonance (NMR) spectroscopy was used to investigate the products of glycerol and acetate metabolism released by Leishmania braziliensis panamensis promastigotes and also to examine the interaction of each of these substrates with glucose or alanine. The NMR data were supplemented by measurements of the rates of oxygen consumption and substrate utilization, and of 14CO2 production from 14C-labeled substrate. Cells incubated with [2-13C]glycerol released acetate, succinate and D-lactate in addition to CO2. Cells incubated with acetate released only CO2. More succinate C-2/C-3 than C-1/C-4 was released from both [2-13C]glycerol and [2-13C]glucose, indicating that succinate was formed predominantly by CO2 fixation followed by reverse flux through part of the Krebs cycle. Some redistribution of the position of labeling was also seen in alanine and pyruvate, suggesting cycling through pyruvate/oxaloacetate/phosphoenolpyruvate. Cells incubated with combinations of 2 substrates consumed oxygen at the same rate as cells incubated with 1 or no substrate, even though the total substrate utilization had increased. When promastigotes were incubated with both glycerol and glucose, the rate of glucose consumption was unchanged but glycerol consumption decreased about 50%, and the rate of 14CO2 production from [1,(3)-14C]glycerol decreased about 60%. Alanine did not affect the rates of consumption of glucose or glycerol, but decreased 14CO2 production from these substrates by increasing flow of label into alanine. Although glucose decreased alanine consumption by 70%, it increased the rate of 14CO2 production from [U-14C]- and [l-14C]alanine by about 20%.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

11.
Phosphoenolpyruvate carboxykinase (PCK) reversibly catalyzes the carboxylation of phosphoenolpyruvate to oxaloacetate. Carbon dioxide, and not bicarbonate ion, is the substrate utilized. Assays of the carboxylation reaction show that initial velocities are 7.6-fold higher when CO(2) is used instead of HCO(3)(-). Two Escherichia coli PCK-CO(2) crystal structures are presented here. The location of CO(2) is the same for both structures; however the orientation of CO(2) is significantly different, likely from the presence of a manganese ion in one of the structures. PCK and the other three known protein-CO(2) crystal structure complexes have been compared; all have CO(2) hydrogen bonding with a basic amino acid side chain (Arg65 or Lys213 in PCK), likely to polarize CO(2) to make the central carbon atom more electrophilic and thus more reactive. Kinetic studies found that the PCK mutant Arg65Gln increased the K(M) for substrates PEP and oxaloacetate but not for CO(2). The unchanged K(M) for CO(2) can be explained since the Arg65Gln mutant likely maintains a hydrogen bond to one of the oxygen atoms of carbon dioxide.  相似文献   

12.
S L Ausenhus  M H O'Leary 《Biochemistry》1992,31(28):6427-6431
In addition to the normal carboxylation reaction, phosphoenolpyruvate carboxylase from Zea mays catalyzes a HCO3(-)-dependent hydrolysis of phosphoenolpyruvate to pyruvate and Pi. Two independent methods were used to establish this reaction. First, the formation of pyruvate was coupled to lactate dehydrogenase in assay solutions containing high concentrations of L-glutamate and aspartate aminotransferase. Under these conditions, oxalacetic acid produced in the carboxylation reaction was efficiently transaminated, and decarboxylation to form spurious pyruvate was negligible. Second, sequential reduction of oxalacetate and pyruvate was achieved by initially running the reaction in the presence of malate dehydrogenase with NADH in excess over phosphoenolpyruvate. After the reaction was complete, lactate dehydrogenase was added, thus giving a measure of pyruvate concentration. At pH 8.0 in the presence of Mg2+, the rate of phosphoenolpyruvate hydrolysis was 3-7% of the total reaction rate. The hydrolysis reaction catalyzed by phosphoenolpyruvate carboxylase was strongly metal dependent, with rates decreasing in the order Ni2+ greater than Co2+ greater than Mn2+ greater than Mg2+ greater than Ca2+. These results suggest that the active site metal ion binds to the enolate oxygen, thus stabilizing the proposed enolate intermediate. The more stable the enolate, the less reactive it is toward carboxylation and the greater the opportunity for hydrolysis.  相似文献   

13.
Glucagon administration to the intact rat has been shown to stimulate pyruvate metabolism in liver mitochondria, presumably by increasing pyruvate transport into the organelle. In this report, we used alanine in place of pyruvate to examine the possibility that glucagon might stimulate pyruvate carboxylation per se independent of its postulated action on pyruvate transport. In agreement with previous reports, injection of a low dose of glucagon (50 micrograms/kg of rat) increased respiration, ATP synthesis, pyruvate decarboxylation, and CO2 fixation in liver mitochondria subsequently isolated. When alanine was used as a substrate, CO2 fixation, but not decarboxylation, was increased in liver mitochondria isolated from glucagon-treated rats. Pyruvate accumulation under these conditions was significantly lower in the glucagon-treated rat preparation. When mitochondria were incubated in a HCO3- -deficient buffer, pyruvate accumulation was identical in both preparations. The addition of a pyruvate transport inhibitor, alpha-cyanohydroxycinnamate (0.5 mM), inhibited CO2 fixation with pyruvate by 70%, but had no effect when alanine was used. Our data therefore suggest that glucagon stimluates mitochondrial pyruvate carboxylation independent of its possible action on pyruvate transport.  相似文献   

14.
Laisk A  Sumberg A 《Plant physiology》1994,106(2):689-695
Photorespiration was calculated from chlorophyll fluorescence and ribulose-1,5-bisphosphate carboxylase/oxygenase (Rubisco) kinetics and compared with CO2 evolution rate in the light, measured by three gas-exchange methods in mature sunflower (Helianthus annuus L.) leaves. The gas-exchange methods were (a) postillumination CO2 burst at unchanged CO2 concentration, (b) postillumination CO2 burst with simultaneous transfer into CO2-free air, and (c) extrapolation of the CO2 uptake to zero CO2 concentration at Rubisco active sites. The steady-state CO2 compensation point was proportional to O2 concentration, revealing the Rubisco specificity coefficient (Ksp) of 86. Electron transport rate (ETR) was calculated from fluorescence, and photorespiration rate was calculated from ETR using CO2 and O2 concentrations, Ksp, and diffusion resistances. The values of the best-fit mesophyll diffusion resistance for CO2 ranged between 0.3 and 0.8 s cm-1. Comparison of the gas-exchange and fluorescence data showed that only ribulose-1,5-bisphosphate (RuBP) carboxylation and photorespiratory CO2 evolution were present at limiting CO2 concentrations. Carboxylation of a substrate other than RuBP, in addition to RuBP carboxylation, was detected at high CO2 concentrations. A simultaneous decarboxylation process not related to RuBP oxygenation was also detected at high CO2 concentrations in the light. We propose that these processes reflect carboxylation of phosphoenolpyruvate, formed from phosphoglyceric acid and the subsequent decarboxylation of malate.  相似文献   

15.
Cytosol PEP carboxykinase has been purified to electrophoretic homogeneity from bullfrog liver homogenate. The enzyme is a single polypeptide chain with a molecular weight of approximately 72,000-75,000. The purified enzyme catalyzed oxaloacetate decarboxylation (nucleoside triphosphate-supported), phosphoenolpyruvate carboxylation, and an exchange reaction between oxaloacetate and [14C]HCO3-in the presence of ITP or CTP. Manganese is absolutely required for the enzyme-catalyzed phosphoenolpyruvate carboxylation, whereas it can be replaced by Mg2+ for the oxaloacetate decarboxylation and the exchange reaction. The optimal pH of each reaction is dependent on the divalent metal ion used. The dependence of the enzyme activity on Mn2+ is markedly different in the phosphoenolpyuvate carboxylation and the oxaloacetate decarboxylation reactions.  相似文献   

16.
T J Williams  R W Henkens 《Biochemistry》1985,24(10):2459-2462
Using 13C NMR spectroscopy, we have further investigated the binding of HCO3- in the active site of an artificial form of human carbonic anhydrase I in which the native zinc is replaced by Co(II). The Co(II) enzyme, unlike all other metal-substituted derivatives, has functional properties closely similar to those of the native zinc enzyme. By measuring the spin-lattice relaxation rate and the line width for both the CO2 and HCO3- at two field strengths, we have determined both the paramagnetic effects that reflect substrate binding and the exchange effects due to catalysis at chemical equilibrium. The following are the results at 14 degrees C and pH 6.3 (1) HCO3- is bound in the active site of the catalytically competent enzyme with the 13C of the HCO3- located 3.22 +/- 0.02 A from the Co(II); (2) the apparent equilibrium dissociation constant for the bound HCO3- is 7.6 +/- 1.5 mM, determined by using the paramagnetic effects on the line widths, and 10 +/- 2 mM, determined by using the exchange effects; (3) the lifetime of HCO3- bound to the metal is (4.4 +/- 0.4) X 10(-5) s; (4) the overall catalyzed CO2 in equilibrium HCO3- exchange rate constant of the Co(II) enzyme is (9.6 +/- 0.4) X 10(3) s-1; (5) the electron spin relaxation time of the Co(II), determined by using paramagnetic effects on the bound HCO3-, is (1.1 +/- 0.1) X 10(-11) s. The data did not provide any direct information on the binding of CO2.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

17.
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   

18.
Light- and CO(2)-saturated photosynthetic rates of the submersed aquatic plants Hydrilla verticillata, Ceratophyllum demersum, and Myriophyllum spicatum were 50 to 60 mumol O(2)/mg Chl.hr at 30 C. At air levels of CO(2), the rates were less than 5% of those achieved by terrestrial C(3) plants. The low photosynthetic rates correlated with low activities of the carboxylation enzymes. In each species, ribulose 1,5-diphosphate carboxylase was the predominant carboxylation enzyme. The apparent K(m)(CO(2)) values for photosynthesis were 150 to 170 mum at pH 4, and 75 to 95 mum at pH 8. The K(m)(CO(2)) of Hydrilla ribulose 1,5-diphosphate carboxylase was 45 mum at pH 8. Optimum temperatures for the photosynthesis of Hydrilla, Myriophyllum, and Ceratophyllum were 36.5, 35.0, and 28.5 C, respectively. The apparent ability of each species to use HCO(3) (-) ions for photosynthesis was similar, but at saturating free CO(2) levels, there was no indication of HCO(3) (-) use. Increasing the pH from 3.1 to 9.2 affected the photosynthetic rate indirectly, by decreasing the free CO(2). With saturating free CO(2) (0.5 mm), the maximum photosynthetic rates were similar at pH 4 and 8. Carbonic anhydrase activity, although much lower than in terrestrial C(3) plants, was still in excess of that required to support HCO(3) (-) utilization.Hydrilla and Ceratophyllum had CO(2) compensation points of 44 and 41 mul/l, respectively, whereas the value for Myriophyllum was 19. Relatively high CO(2) compensation points under 1% O(2) indicated that some "dark" respiration occurred in the light. The inhibition of photosynthesis by O(2) was less than with terrestrial C(3) plants. Glycolate oxidase activity was 12.3 to 27.5 mumol O(2)/mg Chl.hr, as compared to 78.4 for spinach. Light saturation of photosynthesis occurred at 600 to 700 mueinsteins/m(2).sec in each species grown under full sunlight. Hydrilla had the lowest light compensation point, and required the least irradiance to achieve the half-maximal photosynthetic rate.Field measurements in a Hydrilla mat indicated that in the afternoon, free CO(2) dropped to zero, and O(2) rose to over 200% air saturation. Most photosynthetic activity occurred in the morning when the free CO(2) was highest and O(2) and solar radiation lowest. The low light requirement of Hydrilla probably provides a competitive advantage under these field conditions.  相似文献   

19.
When we incubated biotin carboxylase from Escherichia coli with ATP in absence of biotin we observed HCO3- -dependent ATP hydrolysis, which was activated by 10% ethanol in the same proportion as the activity of D-biotin carboxylation assayed in the presence of biotin. The two activities exhibited identical heat stability and were protected equally by glycerol; both required Mg2+ and K+ and showed similar dependency on the concentration of ATP. Biotin assay excluded potential contamination by traces of biotin as a cause of the observed ATP hydrolysis, and this was confirmed by the findings that carboxybiotin did not accumulate and that avidin was uninhibitory. Therefore we concluded that this HCO3- -dependent ATPase was genuinely a partial activity of biotin carboxylase. This partial activity supports a sequential mechanism for enzymatic carboxylation of biotin in which HCO3- is activated by ATP in a first step. It is consistent with the initial formation of the carbonic-phosphoric anhydride (HOCO2PO3(2-)), and it does not agree with models where biotin is phosphorylated by ATP prior to reaction with HCO3-. It appears that enzymes that use HCO3- for carboxylation, including biotin-dependent carboxylases, phosphoenolpyruvate carboxylase, and carbamoyl phosphate synthetase, activate HCO3- by a common mechanism involving the initial formation of the carbonic-phosphoric anhydride.  相似文献   

20.
1. It was shown that dissolved CO(2) and not HCO(3) (-) or H(2)CO(3) is the primary substrate for reductive carboxylation with 6-phosphogluconate dehydrogenase from sheep liver. 2. The equilibrium constant of the reaction was measured in solutions of various ionic strengths and at several temperatures, and the free energy and heat of reaction were determined.  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