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 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   

2.
Hydrogen peroxide can interact with the active site of copper-zinc superoxide dismutase (SOD1) to generate a powerful oxidant. This oxidant can either damage amino acid residues at the active site, inactivating the enzyme (the self-oxidative pathway), or oxidize substrates exogenous to the active site, preventing inactivation (the external oxidative pathway). It is well established that the presence of bicarbonate anion dramatically enhances the rate of oxidation of exogenous substrates. Here, we show that bicarbonate also substantially enhances the rate of self-inactivation of human wild type SOD1. Together, these observations suggest that the strong oxidant formed by hydrogen peroxide and SOD1 in the presence of bicarbonate arises from a pathway mechanistically distinct from that producing the oxidant in its absence. Self-inactivation rates are further enhanced in a mutant SOD1 protein (L38V) linked to the fatal neurodegenerative disorder, familial amyotrophic lateral sclerosis. The 1.4 A resolution crystal structure of pathogenic SOD1 mutant D125H reveals the mode of oxyanion binding in the active site channel and implies that phosphate anion attenuates the bicarbonate effect by competing for binding to this site. The orientation of the enzyme-associated oxyanion suggests that both the self-oxidative and external oxidative pathways can proceed through an enzyme-associated peroxycarbonate intermediate.  相似文献   

3.
Carbonate radicals (CO3-) can be formed biologically by the reaction of OH with bicarbonate, the decomposition of the peroxynitrite-carbon dioxide adduct (ONOOCO2-), and enzymatic activities, i.e., peroxidase activity of CuZnSOD and xanthine oxidase turnover in the presence of bicarbonate. It has been reported that the spin-trap DMPO reacts with CO3(-) to yield transient species to yield finally the DMPO-OH spin adduct. In this study, the kinetics of reaction of CO3(-) with DMPO were studied by pulse radiolysis, yielding a second-order rate constant of 2.5 x 10(6) M(-1) s(-1). A Fenton system, composed of Fe(II)-DTPA plus H2O2, generated OH that was trapped by DMPO; the presence of 50-500 mM bicarbonate, expected to convert OH to CO3(-), markedly inhibited DMPO-OH formation. This was demonstrated to be due mainly to a fast reaction of CO3(-) with FeII-DTPA (k=6.1 x 10(8) M(-1) s(-1)), supported by kinetic analysis. Generation of CO3(-) by the Fenton system was further proved by analysis of tyrosine oxidation products: the presence of bicarbonate caused a dose-dependent inhibition of 3,4-dihydroxiphenylalanine with a concomitant increase of 3,3'-dityrosine yields, and the presence of DMPO inhibited tyrosine oxidation, in agreement with the rate constants with OH or CO3(-). Similarly, the formation of CO3(-) by CuZnSOD/H(2)O(2)/bicarbonate and peroxynitrite-carbon dioxide was supported by DMPO hydroxylation and kinetic competition data. Finally, the reaction of CO3(-) with DMPO to yield DMPO-OH was shown in peroxynitrite-forming macrophages. In conclusion, CO3(-) reacts quite rapidly with DMPO and may contribute to DMPO-OH yields in chemical and cellular systems; in turn, the extent of oxidation of other target molecules (such as tyrosine) by CO3(-) will be sensitive to the presence of DMPO.  相似文献   

4.
Free radicals are produced continuously by skeletal muscle fibers. Extracellular release of reactive oxygen species (ROS) and nitric oxide (NO) derivatives has been demonstrated, but little is known about intracellular oxidant regulation. We used a fluorescent oxidant probe, 2',7'-dichlorofluorescin (DCFH), to assess net oxidant activity in passive muscle fiber bundles isolated from mouse diaphragm and studied in vitro. We tested the following three hypotheses. 1) Net oxidant activity is decreased by muscle cooling. 2) CO(2) exposure depresses intracellular oxidant activity. 3) Muscle-derived ROS and NO both contribute to overall oxidant activity. Our results indicate that DCFH oxidation was diminished by cooling muscle fibers from 37 degrees C to 23 degrees C (P < 0.001). The rate of DCFH oxidation correlated positively with CO(2) exposure (0-10%; P < 0.05) and negatively with concurrent changes in pH (7.0-8.5; P < 0.05). Separate exposures to anti-ROS enzymes (superoxide dismutase, 1 kU/ml; catalase, 1 kU/ml), a glutathione peroxidase mimetic (ebselen, 30 microM), NO synthase inhibitors (N(omega)-nitro-l-arginine methyl ester, 1 mM; N(omega)-monomethyl-l-arginine, 1 mM), or an NO scavenger (hemoglobin, 1 microM) each inhibited DCFH oxidation (P < 0.05). Oxidation was increased by hydrogen peroxide, 100 microM, an NO donor (NOC-22, 400 microM), or the substrate for NO synthase (l-arginine, 5 mM). We conclude that net oxidant activity in resting muscle fibers is 1) decreased at subphysiological temperatures, 2) increased by CO(2) exposure, and 3) influenced by muscle-derived ROS and NO derivatives to similar degrees.  相似文献   

5.
The peroxidase activity of Cu,Zn-superoxide dismutase (Cu,Zn-SOD) has been extensively studied in recent years due to its potential relationship to familial amyotrophic lateral sclerosis. The mechanism by which Cu,Zn-SOD/hydrogen peroxide/bicarbonate is able to oxidize substrates has been proposed to be dependent on an oxidant whose nature, diffusible carbonate radical anion or enzyme-bound peroxycarbonate, remains debatable. One possibility to distinguish these species is to examine whether protein targets are oxidized to protein radicals. Here, we used EPR methodologies to study bovine serum albumin (BSA) oxidation by Cu,Zn-SOD/hydrogen peroxide in the absence and presence of bicarbonate or nitrite. The results showed that BSA oxidation in the presence of bicarbonate or nitrite at pH 7.4 produced mainly solvent-exposed and -unexposed BSA-tyrosyl radicals, respectively. Production of the latter was shown to be preceded by BSA-cysteinyl radical formation. The results also showed that hydrogen peroxide/bicarbonate extensively oxidized BSA-cysteine to the corresponding sulfenic acid even in the absence of Cu,Zn-SOD. Thus, our studies support the idea that peroxycarbonate acts as a two-electron oxidant and may be an important biological mediator. Overall, the results prove the diffusible and radical nature of the oxidants produced during the peroxidase activity of Cu,Zn-SOD in the presence of bicarbonate or nitrite.  相似文献   

6.
The steady-state kinetic mechanism of vitamin K-dependent carboxylase from calf liver has been investigated by initial-velocity measurements with varying concentrations of two carboxylase substrates and constant, nonsaturating concentrations of the other two substrates. With all combinations of the varied substrates tested linear kinetics were obtained with lines intersecting on the left side of the 1/v axis in double-reciprocal plots. Thus the carboxylase has a sequential reaction mechanism which includes the quinternary complex of the enzyme with its four substrates. A mechanism with the ordered steady-state addition of all substrates to the enzyme accords well with the results. A totally random mechanism was excluded but the alternative possibility remained that part of the substrates are added in a rapid-equilibrium random reaction. Experiments with saturating constant concentrations of sodium bicarbonate and varying concentrations of the other substrates suggest that bicarbonate (CO2) is either the first or, more probably, the last substrate bound to the enzyme.  相似文献   

7.
In 16 experiments male subjects, age 22.4 +/- 0.5 (SE) yr, inspired CO2 for 15 min (8% end-tidal CO2) or hyperventilated for 30 min (2.5% end-tidal CO2). Osmolality (Osm) and acid-base status of arterialized venous blood were determined at short intervals until 30 min after hypo- and hypercapnia, respectively. During hypocapnia [CO2 partial pressure (PCO2) -2.31 +/- 0.32 kPa (-17.4 Torr), pH + 0.19 units], Osm decreased by 3.9 +/- 0.3 mosmol/kg H2O; during hypercapnia [PCO2 + 2.10 +/- 0.28 kPa (+15.8 Torr), pH -0.12 units], Osm increased by 5.8 +/- 0.7 mosmol/kg H2O. Presentation of the data in Osm-PCO2 or Osm-pH diagrams yields hysteresis loops probably caused by exchange between blood and tissues. The dependence of Osm on PCO2 must result mainly from CO2 buffering and therefore from the formation of bicarbonate. In spite of the different buffer capacities in various body compartments, water exchange allows rapid restoration of osmotic equilibrium throughout the organism. Thus delta Osm/delta pH during a PCO2 jump largely depends on the mean buffer capacity of the whole body. The high estimated buffer value during hypercapnia (38 mmol/kg H2O) compared with hypocapnia (19 mmol/kg H2O) seems to result from very strong muscle buffering during moderate acidosis.  相似文献   

8.
R Ahmad  Z Wu  D A Armstrong 《Biochemistry》1983,22(8):1806-1810
The kinetics of reaction of oxidized lumiflavin (F0) with the radicals .CO2(-), CH3CHOH, and (CH3)2COH have been investigated at pH 7 and 24 +/- 1 degree C by the pulse radiolysis technique. The radicals have been shown to react with lumiflavin with second-order rate constants of 36 +/- 4, 26 +/- 3, and 20 +/- 3 in units of 10(8) M-1 s-1, respectively. These rate constants are close to the diffusion limit. The main product in each case was the lumiflavin semiquinone radical FH.. By utilization of long pulses (approximately 100 mus), it was shown that the reaction FH. + .AH(alpha) leads to FH- + A(alpha) + H+ [.AH(alpha) = .CO2(-), CH3CHOH, or (CH3)2COH] proceeded for all three types of .AH(alpha) radical with second-order rate constants of 17 (+4,-3), 9 (+5,-3), and 9 (+4,-3), respectively, in the above units. The beta-carbon radical .CH2CH(OH)CH3 added to .FH, forming an alkylated flavin, while the .CH2CH2OH radical appeared to be capable of addition or hydrogen atom donation to .FH.  相似文献   

9.
Copper,zinc superoxide dismutase (Cu,Zn-SOD) catalyzes the HCO(3)(-)-dependent oxidation of diverse substrates. The mechanism of these oxidations involves the generation of a strong oxidant, derived from H(2)O(2), at the active site copper. This bound oxidant then oxidizes HCO(3)(-) to a strong and diffusible oxidant, presumably the carbonate anion radical that leaves the active site and then oxidizes the diverse substrates. Cu,Zn-SOD is also subject to inactivation by H(2)O(2). It is now demonstrated that the rates of HCO(3)(-)-dependent oxidations of NADPH and urate exceed the rate of inactivation of the enzyme by approximately 100-fold. Cu,Zn-SOD is also seen to catalyze a HCO(3)(-)-dependent consumption of the H(2)O(2) and that HCO(3)(-) does not protect Cu,Zn-SOD against inactivation by H(2)O(2). A scheme of reactions is offered in explanation of these observations.  相似文献   

10.
The visualization at near atomic resolution of transient substrates in the active site of enzymes is fundamental to fully understanding their mechanism of action. Here we show the application of using CO(2)-pressurized, cryo-cooled crystals to capture the first step of CO(2) hydration catalyzed by the zinc-metalloenzyme human carbonic anhydrase II, the binding of substrate CO(2), for both the holo and the apo (without zinc) enzyme to 1.1A resolution. Until now, the feasibility of such a study was thought to be technically too challenging because of the low solubility of CO(2) and the fast turnover to bicarbonate by the enzyme (Liang, J. Y., and Lipscomb, W. N. (1990) Proc. Natl. Acad. Sci. U. S. A. 87, 3675-3679). These structures provide insight into the long hypothesized binding of CO(2) in a hydrophobic pocket at the active site and demonstrate that the zinc does not play a critical role in the binding or orientation of CO(2). This method may also have a much broader implication for the study of other enzymes for which CO(2) is a substrate or product and for the capturing of transient substrates and revealing hydrophobic pockets in proteins.  相似文献   

11.
The Rhinelander free-air CO(2) enrichment (FACE) experiment is designed to understand ecosystem response to elevated atmospheric carbon dioxide (+CO(2)) and elevated tropospheric ozone (+O(3)). The objectives of this study were: to understand how soil respiration responded to the experimental treatments; to determine whether fine-root biomass was correlated to rates of soil respiration; and to measure rates of fine-root turnover in aspen (Populus tremuloides) forests and determine whether root turnover might be driving patterns in soil respiration. Soil respiration was measured, root biomass was determined, and estimates of root production, mortality and biomass turnover were made. Soil respiration was greatest in the +CO(2) and +CO(2) +O(3) treatments across all three plant communities. Soil respiration was correlated with increases in fine-root biomass. In the aspen community, annual fine-root production and mortality (g m(-2)) were positively affected by +O(3). After 10 yr of exposure, +CO(2) +O(3)-induced increases in belowground carbon allocation suggest that the positive effects of elevated CO(2) on belowground net primary productivity (NPP) may not be offset by negative effects of O(3). For the aspen community, fine-root biomass is actually stimulated by +O(3), and especially +CO(2) +O(3).  相似文献   

12.
The threonine dehydrogenase (TDG) pathway is a significant route of threonine degradation, yielding glycine in experimental animals, but has not been accurately quantitated in humans. Therefore, the effect of a large excess of dietary threonine, given either as free amino acid (+Thr) or as a constituent of protein (+P-Thr), on threonine catabolism to CO(2) and to glycine was studied in six healthy adult males using a 4-h constant infusion of L-[1-(13)C]threonine and [(15)N]glycine. Gas chromatography-combustion isotope ratio mass spectrometry was used to determine [(13)C]glycine produced from labeled threonine. Threonine intakes were higher on +Thr and +P-Thr diets compared with control (126, 126, and 50 micromol x kg(-1) x h(-1), SD 8, P < 0.0001). Threonine oxidation to CO(2) increased threefold in subjects on +Thr and +P-Thr vs. control (49, 45, and 15 micromol x kg(-1) x h(-1), SD 6, P < 0.0001). Threonine conversion to glycine tended to be higher on +Thr and +P-Thr vs. control (3.5, 3.4, and 1.6 micromol x kg(-1) x h(-1), SD 1.3, P = 0.06). The TDG pathway accounted for only 7-11% of total threonine catabolism and therefore is a minor pathway in the human adult.  相似文献   

13.
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   

14.
Nuclear maturation of domestic cat ovarian oocytes in vitro.   总被引:3,自引:0,他引:3  
Using the domestic cat as a model for salvaging genetic material from rare Felidae, we collected oocytes from ovarian tissue and placed them in 1 of 3 treatments to observe time-related, meiotic changes of in vitro oocyte maturation. Oocytes obtained from ovaries collected at ovario-hysterectomy were assigned to 1 of 3 treatment groups: 1) modified Krebs-Ringer bicarbonate buffer (mKRB) + 4% BSA and 5 micrograms/ml FSH (+FSH, n = 499); 2) mKRB + 4% BSA (-FSH, n = 502); or 3) mKRB + 5% natural estrus cat serum (NE, n = 873). They were placed in the respective media in a 5% CO2 humidified environment at 38 degrees C. Beginning at 16 h, oocytes were removed at 4-h intervals through 48 h, and the meiotic status was evaluated by means of cytogenetic analysis. On the basis of chromosomal analysis, each cell was placed into one of the following categories: metaphase II (MII); metaphase I (MI); pre-MI (germinal vesicle [GV], GV breakdown, or diakinesis); degenerate or unidentifiable. The percentage of oocytes with degenerate chromatin increased over time in all culture treatments, but was always greatest (p less than 0.05) in the NE group. In the +FSH and -FSH treatments, the proportion of oocytes with nuclear material reaching MII increased with time in culture to 32 h and was equal to or greater than the proportion of oocytes with pre-MI + MI chromatin at this time interval (-FSH, 55%; +FSH, 38%).(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

15.
Hydrolytic mechanisms of family 18 chitinases from rice (Oryza sativa L.) and Bacillus circulans WL-12 were comparatively studied by a combination of HPLC analysis of the reaction products and theoretical calculation of reaction time-courses. All of the enzymes tested produced beta-anomers from chitin hexasaccharide [(GlcNAc)(6)], indicating that they catalyze the hydrolysis through a retaining mechanism. The rice chitinases hydrolyzed predominantly the fourth and fifth glycosidic linkages from the nonreducing end of (GlcNAc)(6), whereas B. circulans chitinase A1 hydrolyzed the second linkage from the nonreducing end. In addition, the Bacillus enzyme efficiently catalyzed transglycosylation, producing significant amounts of chitin oligomers larger than the initial substrate, but the rice chitinases did not. The time-courses of (GlcNAc)(6) degradation obtained by HPLC were analyzed by theoretical calculation, and the subsite structures of the rice chitinases were identified to be (-4)(-3)(-2)(-1)(+1)(+2). From the HPLC profile of the reaction products previously reported [Terwisscha van Scheltinga et al. (1995) Biochemistry 34, 15619-15623], family 18 chitinase from rubber tree (Hevea brasiliensis) was estimated to have the same type of subsite structure. Theoretical analysis of the reaction time-course for the Bacillus enzyme revealed that the enzyme has (-2)(-1) (+1)(+2)(+3)(+4)-type subsite structure, which is identical to that of fungal chitinase from Coccidioides immitis [Fukamizo et al. (2001) Biochemistry 40, 2448-2454]. The Bacillus enzyme also resembled the fungal chitinase in its transglycosylation activity. Minor structural differences between plant and microbial enzymes appear to result in such functional variations, even though all of these chitinases are classified into the identical family of glycosyl hydrolases.  相似文献   

16.
The effects of acid--base alterations produced by changing bicarbonate (metabolic type), carbon dioxide tension (respiratory type), or both bicarbonate and carbon dioxide tension (compensated type) on skeletal muscle twitch tension, intracellular pH, and intracellular potassium were studied in vitro. Hemidiaphragm muscles from normal rats and rats fed a potassium-deficient diet were used. Decreasing the extracellular pH by decreasing bicarbonate or increasing CO2 in the bathing fluid produced a decrease in intracellular pH, intracellular K+, and muscle twitch tension. However, at a constant extracellular pH, an increase in CO2 (compensated by an increase in bicarbonate) produced an increase in intracellular K+ and twitch tension in spite of a decrease in intracellular pH. The effect on twitch tension of the hemidiaphragms showed a rapid onset, was reversible, persisted until the buffer composition was changed, and was independent of synaptic transmission. It is concluded that the twitch tension of the skeletal muscle decrease with a decrease in intracellular K+. The muscle tension also decreases with an increase in the ratio of intracellular and extracellular H+ concentration. However, there is no consistent relationship between muscle tension and extracellular or intracellular pH. The muscle tension of the diaphragms taken from K+-deficient rats is more sensitive to variations in CO2, PH, and bicarbonate concentration of the medium than that of the control rat diaphragms.  相似文献   

17.
Kopper BJ  Lindroth RL 《Oecologia》2003,134(1):95-103
The purpose of this study was to assess the independent and interactive effects of CO(2), O(3), and plant genotype on the foliar quality of a deciduous tree and the performance of a herbivorous insect. Two trembling aspen (Populus tremuloides Michaux) genotypes differing in response to CO(2) and O(3) were grown at the Aspen FACE (Free Air CO(2) Enrichment) site located in northern Wisconsin, USA. Trees were exposed to one of four atmospheric treatments: ambient air (control), elevated carbon dioxide (+CO(2); 560 microl/l), elevated ozone (+O(3); ambient x1.5), and elevated CO(2)+O(3). We measured the effects of CO(2) and O(3) on aspen phytochemistry and on performance of forest tent caterpillar (Malacosoma disstria Hübner) larvae. CO(2) and O(3) treatments influenced foliar quality for both genotypes, with the most notable effects being that elevated CO(2) reduced nitrogen and increased tremulacin levels, whereas elevated O(3) increased early season nitrogen and reduced tremulacin levels, relative to controls. With respect to insects, the +CO(2) treatment had little or no effect on larval performance. Larval performance improved in the +O(3) treatment, but this response was negated by the addition of elevated CO(2) (i.e., +CO(2)+O(3) treatment). We conclude that tent caterpillars will have the greatest impact on aspen under current CO(2) and high O(3) levels, due to increases in insect performance and decreases in tree growth, whereas tent caterpillars will have the least impact on aspen under high CO(2) and low O(3) levels, due to moderate changes in insect performance and increases in tree growth.  相似文献   

18.
(14)CO(2) production by perfused rat livers was simulated by infusing NaH(14)CO(3) into the perfusate. Recovery of label as (14)CO(2) gas + perfusate bicarbonate was 45-85%. Rates of (14)CO(2) exchange in the liver are 3-70 times greater than net rates of CO(2) production. Therefore (14)CO(2) reincorporation can lead to significant underestimations of rates of oxidation of (14)C-labelled substrates in liver.  相似文献   

19.
The interaction of Cu,ZnSOD with H2O2 generates an oxidant at the active site that can then cause either the inactivation of this enzyme or the oxidation of a variety of exogenous substrates. We show that the rate of inactivation, imposed by 10-mM H2O2 at 25 degrees C and pH 7.2, is not influenced by 10-mM HCO3-; whereas the oxidation of 2,2'-azino-bis-[3-ethylbenzothiazoline sulfonate] (ABTS=) is virtually completely dependent upon HCO3-. The reduction of the active site Cu(II) by H2O2, which precedes inactivation of the enzyme, occurred at the same rate in phosphate buffer with or without bicarbonate added. These results indicate that HCO3- does not play a role in facilitating the interaction of H2O2 with the active site copper, but they can be accommodated by the proposal that HCO3- is oxidized to HCO3*, which then diffuses from that site and causes the oxidation of substrates, such as ABTS=, that are too large to traverse the solvent access channel to the Cu(II).  相似文献   

20.
The enzymes of the Calvin reductive pentose phosphate cycle and the hexose monophosphate pathway have been demonstrated in cell-free extracts of Thiobacillus ferrooxidans. This, together with analyses of the products of CO(2) fixation in cell-free systems, suggests that these pathways are operative in whole cells of this microorganism. Nevertheless, the amount of CO(2) fixed in these cell-free systems was limited by the type and amount of compound added as substrate. The inability of cell extracts to regenerate pentose phosphates and to perpetuate the cyclic fixation of CO(2) is partially attributable to low activity of triose phosphate dehydrogenase under the experimental conditions found to be optimal for the enzymes involved in the utilization of ribose-5-phosphate or ribulose-1,5-diphosphate as substrate for CO(2) incorporation. With the exception of ribulose-1,5-diphosphate, all substrates required the addition of adenosine triphosphate (ATP) or adenosine diphosphate (ADP) for CO(2) fixation. Under optimal conditions, with ribose-5-phosphate serving as substrate, each micromole of ATP added resulted in the fixation of 1.5 mumoles of CO(2), whereas each micromole of ADP resulted in 0.5 mumole of CO(2) fixed. These values reflect the activity of adenylate kinase in the extract preparations. The K(m) for ATP in the phosphoribulokinase reaction was 0.91 x 10(-3)m. Kinetic studies conducted with carboxydismutase showed K(m) values of 1.15 x 10(-4)m and 5 x 10(-2)m for ribulose-1,5-diphosphate and bicarbonate, respectively.  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