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 921 毫秒
1.
The structural and kinetic analyses of the components of the lactate shuttle from heterotrophic Euglena gracilis were carried out. Mitochondrial membrane-bound, NAD(+)-independent d-lactate dehydrogenase (d-iLDH) was purified by solubilization with CHAPS and heat treatment. The active enzyme was a 62-kDa monomer containing non-covalently bound FAD as cofactor. d-iLDH was specific for d-lactate and it was able to reduce quinones of different redox potential values. Oxalate and l-lactate were mixed-type inhibitors of d-iLDH. Mitochondrial l-iLDH also catalyzed the reduction of quinones, but it was inactivated during the extraction with detergents. Both l-iLDH and d-iLDH were inhibited by the specific flavoprotein-inhibitor diphenyleneiodonium, suggesting that l-iLDH was also a flavoprotein. Affinity chromatography revealed that the E. gracilis cytosolic fraction contained two types of NAD(+)-dependent LDH specific for the generation of d- and l-lactate (d-nLDH and l-nLDH, respectively). These two enzymes were tetramers of 126-132 kDa and showed an ordered bi-bi kinetic mechanism. Kinetic properties were different in both enzymes. Pyruvate reduction by d-nLDH was inhibited by its two products; the d-lactate oxidation was 40-fold lower than forward reaction. l-lactate oxidation by l-nLDH was not detected, whereas pyruvate reduction was activated by fructose-1, 6-bisphosphate, K(+) or NH(4)(+). Interestingly, membrane-bound l- and d-lactate dehydrogenases with quinone reductase activity have been only detected in bacteria, whereas the activity of soluble d-nLDH has been identified in bacteria and some yeast. Also, FBP-activated l-nLDH has been found solely in lactic bacteria. Based on their similar kinetic and structural characteristics, a possible common origin among bacterial and E. gracilis lactic dehydrogenase enzymes is discussed.  相似文献   

2.
The steady-state kinetics of alcohol dehydrogenases (alcohol:NAD+ oxidoreductase, EC 1.1.1.1 and alcohol:NADP+ oxidoreductase, EC 1.1.1.2), lactate dehydrogenases (l-lactate:NAD+ oxidoreductase, EC 1.1.1.27 and d-lactate:NAD+ oxidoreductase, EC 1.1.1.28), malate dehydrogenase (l-malate:NAD+ oxidoreductase, EC 1.1.1.37), and glyceraldehyde-3-phosphate dehydrogenases [d-glyceraldehyde-3-phosphate:NAD+ oxidoreductase (phosphorylating), EC 1.2.1.12] from different sources (prokaryote and eukaryote, mesophilic and thermophilic organisms) have been studied using NAD(H), N6-(2-carboxyethyl)-NAD(H), and poly(ethylene glycol)-bound NAD(H) as coenzymes. The kinetic constants for NAD(H) were changed by carboxyethylation of the 6-amino group of the adenine ring and by conversion to macromolecular form. Enzymes from thermophilic bacteria showed especially high activities for the derivatives. The relative values of the maximum velocity (NAD = 1) of Thermus thermophilus malate dehydrogenase for N6-(2-carboxyethyl)-NAD and poly(ethylene glycol)-bound NAD were 5.7 and 1.9, respectively, and that of Bacillus stearothermophilus glyceraldehyde-3-phosphate dehydrogenase for poly(ethylene glycol)-bound NAD was 1.9.  相似文献   

3.
NAD+ has been covalently attached to dextrans having different molecular weights to give various NAD+ densities (mol NAD+ per mol d-glucosyl residue). The effects of molecular weight of dextran and of NAD+ density on the coenzyme activity of the dextran-bound NAD+ derivatives were examined for the reactions catalysed by alcohol dehydrogenase (alcohol: NAD+ oxidoreductase, EC 1.1.1.1) and lactate dehydrogenase (l-lactate:NAD+ oxidoreductase, EC 1.1.1.27). The molecular weight of dextran had little effect on coenzyme activity in the range 10 000 to 500 000. At low NAD+ density (<0.05 mol NAD+/mol d-glucosyl residue), the coenzyme activities of the derivatives were relatively low, but higher densities had little effect on the activity. Dextran-bound NAD+ derivatives were twice as stable as free NAD+.  相似文献   

4.
This study is concerned with further development of the kinetic locking-on strategy for bioaffinity purification of NAD+-dependent dehydrogenases. Specifically, the synthesis of highly substituted N6-linked immobilized NAD+ derivatives is described using a rapid solid-phase modular approach. Other modifications of the N6-linked immobilized NAD+ derivative include substitution of the hydrophobic diaminohexane spacer arm with polar spacer arms (9 and 19.5 Å) in an attempt to minimize nonbiospecific interactions. Analysis of the N6-linked NAD+ derivatives confirm (i) retention of cofactor activity upon immobilization (up to 97%); (ii) high total substitution levels and high percentage accessibility levels when compared to S6-linked immobilized NAD+ derivatives (also synthesized with polar spacer arms); (iii) short production times when compared to the preassembly approach to synthesis. Model locking-on bioaffinity chromatographic studies were carried out with bovine heart -lactate dehydrogenase ( -LDH, EC 1.1.1.27), bakers yeast alcohol dehydrogenase (YADH, EC 1.1.1.1) and Sporosarcinia sp. -phenylalanine dehydrogenase ( -PheDH, EC 1.4.1.20), using oxalate, hydroxylamine, and -phenylalanine, respectively, as locking-on ligands. Surprisingly, two of these test NAD+-dependent dehydrogenases (lactate and alcohol dehydrogenase) were found to have a greater affinity for the more lowly substituted S6-linked immobilized cofactor derivatives than for the new N6-linked derivatives. In contrast, the NAD+-dependent phenylalanine dehydrogenase showed no affinity for the S6-linked immobilized NAD+ derivative, but was locked-on strongly to the N6-linked immobilized derivative. That this locking-on is biospecific is confirmed by the observation that the enzyme failed to lock-on to an analogous N6-linked immobilized NADP+ derivative in the presence of -phenylalanine. This differential locking-on of NAD+-dependent dehydrogenases to N6-linked and S6-linked immobilized NAD+ derivatives cannot be explained in terms of final accessible substitutions levels, but suggests fundamental differences in affinity of the three test enzymes for NAD+ immobilized via N6-linkage as compared to thiol-linkage.  相似文献   

5.
The catalytic properties of the purified horseshoe crab and seaworm d-lactate dehydrogenases were determined and compared with those of several l-lactate dehydrogenases. Apparent Km's and degrees of substrate inhibition have been determined for both enzymes for pyruvate, d-lactate, NAD+ and NADH. They are similar to those found for l-lactate dehydrogenases. The Limulus “muscle”-type lactate dehydrogenase is notably different from the “heart”-type lactate dehydrogenase of this organism in a number of properties.The Limulus heart and muscle enzymes have been shown by several criteria to be stereospecific for d-lactate. They also stereospecifically transfer the 4-α hydrogen of NADH to pyruvate. The turnover number for purified Limulus muscle lactate dehydrogenase is 38,000 moles NADH oxidized per mole of enzyme, per minute. Limulus and Nereis lactate dehydrogenases are inhibited by oxamate and the reduced NAD-pyruvate adduct.Limulus muscle lactate dehydrogenase is stoichiometrically inhibited by para-hydroxymercuribenzoate. Extrapolation to two moles parahydroxymercuribenzoate bound to one mole of enzyme yields 100% inhibition. Alkylation by iodoacetamide or iodoacetate occurs even in the absence of urea or guanidine-HCl. Evidence suggests that the reactive sulfhydryl group may not be located at the coenzyme binding site.Reduced coenzyme (NADH or the 3-acetyl-pyridine analogue of NADH) stoichiometrically binds to Limulus muscle lactate dehydrogenase (two moles per mole of enzyme).Several pieces of physical and catalytic evidence suggest that the d- and l-lactate dehydrogenase are products of homologous genes. A consideration of a possible “active site” shows that as few as one or two key conservative amino acid changes could lead to a reversal of the lactate stereospecificity.  相似文献   

6.
Anaerobically grown cells of Saccharomyces cerevisiae entrapped in polyacrylamide gel have been shown to provide a stable source of alcohol dehydrogenase [(ADH) alcohol:NAD+ oxidoreductase, EC 1.1.1.1] for effective regeneration of NAD(H). This system was able to provide the coenzyme required for the operation of other dehydrogenases, such as lactate dehydrogenase [(LDH) l-lactate: NAD+ oxidoreductase, EC 1.1.1.27] and malate dehydrogenase [(MDH) l-malate:NAD+ oxidoreductase, EC 1.1.1.37]. Yeast cells coimmobilized with a dehydrogenase are capable of the reversible regeneration of the reduced or oxidized coenzyme, depending on the additions made. A two-cell system can also be constituted using the same strain of yeast, adapted differently. Cells grown anaerobically and aerobically as sources of ADH and MDH, respectively, can operate efficiently on coimmobilization. The system can be used repeatedly without measurable loss of efficiency.  相似文献   

7.
The ethanologenic bacterium Zymomonas mobilis ZM4 is of special interest because it has a high ethanol yield. This is made possible by the two alcohol dehydrogenases (ADHs) present in Z. mobilis ZM4 (zmADHs), which shift the equilibrium of the reaction toward the synthesis of ethanol. They are metal-dependent enzymes: zinc for zmADH1 and iron for zmADH2. However, zmADH2 is inactivated by oxygen, thus implicating zmADH2 as the component of the cytosolic respiratory system in Z. mobilis. Here, we show crystal structures of zmADH2 in the form of an apo-enzyme and an NAD+-cofactor complex. The overall folding of the monomeric structure is very similar to those of other functionally related ADHs with structural variations around the probable substrate and NAD+ cofactor binding region. A dimeric structure is formed by the limited interactions between the two subunits with the bound NAD+ at the cleft formed along the domain interface. The catalytic iron ion binds near to the nicotinamide ring of NAD+, which is likely to restrict and locate the ethanol to the active site together with the oxidized Cys residue and several nonpolar bulky residues. The structures of the zmADH2 from the proficient ethanologenic bacterium Z. mobilis, with and without NAD+ cofactor, and modeling ethanol in the active site imply that there is a typical metal-dependent catalytic mechanism.  相似文献   

8.
ATP and respiration (NADH)-driven NAD(P)+ transhydrogenase (EC 1.6.1.1) activities are low in membranes from Escherichia coli cultured on yeast extract medium (17 and 21 nmol/min × mg) but high on glucose (82 and 142 nmol/min × mg). The ATPase and respiratory activities in both cases appeared comparable. Growth of the bacteria in yeast extract medium followed by washing and replacement into a glucose medium showed that after 3 h the energy-linked and energy-independent NAD(P)+ transhydrogenase (reduction of acetylpyridine NAD+ by NADPH) activities had appeared simultaneously. Incorporation of chloramphenicol or omission of glucose in the induction medium resulted in no increase in these activities indicating that de novo protein synthesis is required for the induction of energy-linked and -independent NAD(P)+ transhydrogenase. It was found that the Km values for acetylpyridine NAD+ and NADPH for the energy-independent reaction in membranes from glucose grown cells (143 and 62 μm) were similar to those in membranes from cells grown on glucose-yeast extract (135 and 45 μm), respectively, but the maximum velocity at infinite acetyl pyridine NAD+ and NADPH increased from 353 to 2175 nmol/min × mg. Furthermore, the membrane-bound NAD(P)+ transhydrogenase in glucose-yeast extract grown cells showed substrate inhibition at high NADPH and low acetyl pyridine NAD+ levels. Further kinetic data demonstrate that the mechanism of the energy-independent NAD(P)+ transhydrogenase in E. coli is similar to that of the mitochondrial enzyme and exhibits similar responses to competitive inhibitors at the NAD+ and NADPH sites.  相似文献   

9.
《Process Biochemistry》2010,45(12):1912-1915
Production of pyruvate from lactate through biocatalysis is a valuable process for its simple composition of reaction system and convenience of recovery. Biocatalyst with lactate-induced NAD-independent lactate dehydrogenases (iLDHs) can effectively catalyze lactate into pyruvate. To reduce the cost of biocatalyst preparation caused by indispensable lactate addition, the mutants with constitutive iLDH of Pseudomonas sp. XP-M2 were screened. Mutant XP-LM exhibited high iLDHs activities in minimal salt medium with cheap substrate glucose as the carbon source. The biocatalyst (8.2 g dry cell weight l−1) containing 169.8 U l−1 l-iLDH was prepared with 20 g 1−1 glucose. The cost-effective biocatalyst prepared from the mutant XP-LM could efficiently catalyze lactate into pyruvate with high yield (0.961 mol mol−1). Based on the different thermostability of d-iLDH and l-iLDH in the biocatalyst, whole cells of the strain might also have the potential in production of pyruvate and d-lactate from racemic lactate.  相似文献   

10.
Sirtuin1 (SIRT1) deacetylase and poly(ADP-ribose)-polymerase-1 (PARP-1) respond to environmental cues, and both require NAD+ cofactor for their enzymatic activities. However, the functional link between environmental/oxidative stress-mediated activation of PARP-1 and SIRT1 through NAD+ cofactor availability is not known. We investigated whether NAD+ depletion by PARP-1 activation plays a role in environmental stimuli/oxidant-induced reduction in SIRT1 activity. Both H2O2 and cigarette smoke (CS) decreased intracellular NAD+ levels in vitro in lung epithelial cells and in vivo in lungs of mice exposed to CS. Pharmacological PARP-1 inhibition prevented oxidant-induced NAD+ loss and attenuated loss of SIRT1 activity. Oxidants decreased SIRT1 activity in lung epithelial cells; however increasing cellular NAD+ cofactor levels by PARP-1 inhibition or NAD+ precursors was unable to restore SIRT1 activity. SIRT1 was found to be carbonylated by CS, which was not reversed by PARP-1 inhibition or selective SIRT1 activator. Overall, these data suggest that environmental/oxidant stress-induced SIRT1 down-regulation and PARP-1 activation are independent events despite both enzymes sharing the same cofactor.  相似文献   

11.
12.
Butanedione in the presence of borate buffer reversibly inhibits Rhodospirillum rubrum chromatophore transhydrogenase complex and the separated membrane-bound and soluble factor components of the complex. NADP+ completely protected against inactivation of the membrane-bound component, whereas NAD+ was without effect. Soluble factor was maximally protected only partially by either NAD+ or NADP+, but a mixture of the substrates afforded complete protection. NADP+-dependent association of soluble factor with factor-depleted membranes was markedly decreased after incubation of membranes with butanedione in the absence, but not in the presence, of NADP+. Soluble factor was bound to agarose-NAD and was eluted by NAD+, but not by NADP+. These results demonstrate the presence of at least three nicotinamide adenine dinucleotide binding sites on R. rubrum transhydrogenase complex, including separate NADP and NAD binding sites on soluble factor and a NADP binding site on the membrane-bound component.  相似文献   

13.
The kinetic locking-on strategy utilizes soluble analogues of the target enzymes' specific substrate to promote selective adsorption of individual NAD+-dependent dehydrogenases on their complementary immobilized cofactor derivative. Application of this strategy to the purification of NAD+-dependent dehydrogenases from crude extracts has proven that it can yield bioaffinity systems capable of producing one-chromatographic-step purifications with yields approaching 100%. However, in some cases the purified enzyme preparation was found to be contaminated with other proteins weakly bound to the immobilized cofactor derivative through binary complex formation and/or nonspecific interactions, which continuously “dribbled” off the matrix during the chromatographic procedure. The fact that this problem can be overcome by including a short pulse of 5′-AMP (stripping ligand) in the irrigant a couple of column volumes prior to the discontinuation of the specific substrate analogue (locking-on ligand) is clear from the results presented in this report. The general effectiveness of this auxiliary tactic has been assessed using model studies and through incorporation into an actual purification from a crude cellular extract. The results confirm the usefulness of the stripping-ligand tactic for the resolution and purification of NAD+-dependent dehydrogenases when using the locking-on strategy. These studies have been carried out using bovine liver glutamate dehydrogenase (GDH, EC 1.4.1.3), yeast alcohol dehydrogenase (YADH, EC 1.1.1.1), porcine heart mitochondrial malate dehydrogenase (mMDH, EC 1.1.1.37), and bovine heart -lactate dehydrogenase ( -LDH, EC 1.1.1.27).  相似文献   

14.
The activities of nuclear enzymes involved in NAD+ metabolism in Saccharomyces cerevisiae strain 913a-1 and its mutant 110 previously selected as an NAD+ producer were investigated. The presence of extracellular nicotinamide increased the total NAD+ pool in the cells and increased [3H]nicotinic acid incorporation; however, NAD+ concentration in isolated nuclei decreased slightly. The stimulating effect of nicotinamide on intracellular synthesis of NAD+ correlated with increases in ADP-ribosyl transferase, NAD+-pyrophosphorylase, and NAD+ ase activities.  相似文献   

15.
The optimal concentrations of diaphorase, methyl viologen (MV2+) and NAD+ in the mediated electrocatalytic reduction of NAD+ were decided by applying cyclic voltammetry. The steady-state catalytic current was achieved under the conditions of 1.5 U diaphorase ml–1, 0.2 mM MV2+, and 4.8 mM NAD+ at the scan rate of 2 mV s–1, suggesting that the rate of the electrocatalytic reaction is the highest under the former conditions. However, NAD+ was effective at 0.3 mM as it was at 4.8 mM when the electrocatalysis is coupled with an enzymatic synthesis requiring NADH. In effect, the electrochemical procedure under the conditions of 1.5 U diaphorase ml–1, 0.2 mM MV2+, and 0.3 mM NAD+ worked satisfactorily to drive an enzymatic reduction of pyruvate to d-lactate in the presence of d-lactate dehydrogenase.  相似文献   

16.
祁肖肖  王丽敏  于波 《微生物学报》2024,64(5):1538-1549
耐热凝结芽孢杆菌因其对营养要求简单、发酵产物浓度高以及耐高温等特点,已成为乳酸发酵的主要菌种。在前期的研究中,我们发现磷酸盐可以激活凝结芽孢杆菌l-乳酸脱氢酶基因的转录,从而提高乳酸产量。然而,磷酸盐如何激活乳酸脱氢酶的基因表达,目前还不清楚,也未有类似的研究报道。【目的】对凝结芽孢杆菌响应磷酸盐的调控机制进行研究。【方法】通过RT-PCR分析磷酸盐添加时凝结芽孢杆菌乳酸脱氢酶转录水平变化,确定响应磷酸盐的关键元件区域,进一步通过分子生物学手段,分析凝结芽孢杆菌响应磷酸盐的关键基因片段。【结果】确定了响应磷酸盐的关键元件位于乳酸脱氢酶基因上游启动子区,解析了响应磷酸盐的l-乳酸脱氢酶启动子核心区,利用该启动子及核心区能够有效驱动外源d-乳酸脱氢酶基因的表达,实现在凝结芽孢杆菌中d-乳酸的合成。【结论】本研究有望获得一种新的响应磷酸盐的调控元件,为提高其他生物化学品的合成效率改造提供参考。  相似文献   

17.
B.J. Wallace  I.G. Young 《BBA》1977,461(1):84-100
A ubiA? menA? double quinone mutant of Escherichia coli K12 was constructed together with other isogenic strains lacking either ubiquinone or menaquinone. These strains were used to study the role of quinones in electron transport to oxygen and nitrate. Each of the four oxidases examined (NADH, d-lactate, α-glycerophosphate and succinate) required a quinone for activity. Ubiquinone was active in each oxidase system while menaquinone gave full activity in α-glycerophosphate oxidase, partial activity in d-lactate oxidase but was inactive in NADH and succinate oxidation. The aerobic growth rates, growth yields and products of glucose metabolism of the quinone-deficient strains were also examined. The growth rate and growth yield of the ubi+ menA? strain was the same as the wild-type strain, whereas the ubiA? men+ strain grew more slowly on glucose, had a lower growth yield (30% of wild type) and accumulated relatively large quantities of acetate and lactate. The growth of the ubiA? menA? strain was even more severely affected than that of the ubiA? men+ strain.Electron transport from formate, d-lactate, α-glycerophosphate and NADH to nitrate was also highly dependent on the presence of a quinone. Either ubiquinone or menaquinone was active in electron transport from formate and the activity of the quinones in electron transport from the other substrates was the same as for the oxidase systems. In contrast, quinones were not obligatory carriers in the anaerobic formate hydrogenlyase system. It is concluded that the quinones serve to link the various dehydrogenases with the terminal electron transport systems to oxygen and nitrate and that the dehydrogenases possess a degree of selectivity with respect to the quinone acceptors.  相似文献   

18.
AIM:To characterize phosphorylation of human glyceraldehyde 3-phosphate dehydrogenase(GAPDH),and mobility of GAPDH in cancer cells treated with chemotherapeutic agents. METHODS:We used proteomics analysis to detect and characterize phosphorylation sites within human GAPDH. Site-specific mutagenesis and alanine scanning was then performed to evaluate functional significance of phosphorylation sites in the GAPDH polypeptide chain. Enzymatic properties of mutated GAPDH variants were assessed using kinetic studies. Intranuclear dynamics parameters(diffusion coefficient and the immobile fraction) were estimated using fluorescence recovery after photobleaching(FRAP) experiments and confocal microscopy. Molecular modeling experiments were performed to estimate the effects of mutations on NAD+ cofactor binding.RESULTS:Using MALDI-TOF analysis,we identified novel phosphorylation sites within the NAD+ binding center of GAPDH at Y94,S98,and T99. Using polyclonal antibody specific to phospho-T99-containing peptide within GAPDH,we demonstrated accumulation of phospho-T99-GAPDH inthe nuclear fractions of A549,HCT116,and SW48 cancer cel s after cytotoxic stress. We performed site-mutagenesis,and estimated enzymatic properties,intranuclear distribution,and intranuclear mobility of GAPDH mutated variants. Site-mutagenesis at positions S98 and T99 in the NAD+ binding center reduced enzymatic activity of GAPDH due to decreased affinity to NAD+(Km = 741 ± 257 μmol/L in T99 I vs 57 ± 11.1 μmol/L in wild type GAPDH. Molecular modeling experiments revealed the effect of mutations on NAD+ binding with GAPDH. FRAP(fluorescence recovery after photo bleaching) analysis showed that mutations in NAD+ binding center of GAPDH abrogated its intranuclear interactions. CONCLUSION:Our results suggest an important functional role of phosphorylated amino acids in the NAD+ binding center in GAPDH interactions with its intranuclear partners.  相似文献   

19.
Several denitrifying Pseudomonas strains contained an NADP+-specific 2-oxoglutarate dehydrogenase, in contrast to an NAD+-specific pyruvate dehydrogenase, if the cells were grown anaerobically with aromatic compounds. With non-aromatic substrates or after aerobic growth the coenzyme specificity of 2-oxoglutarate dehydrogenase changed to NAD+-specificity. The reaction stoichiometry and the apparent K m-values of the enriched enzymes were determined: pyruvate 0.5 mM, coenzyme A 0.05 mM, NAD+ 0.25 mM; 2-oxoglutarate 0.6 mM, coenzyme A 0.05 mM, NADP+ 0.03 mM. Isocitrate dehydrogenase was NADP+-specific. The findings suggest that these strains contained at least two lipoamide dehydrogenases, one NAD+-specific, the other NADP+-specific.  相似文献   

20.
Leuconostoc mesenteroides increased its lactic acid production from glucose threefold when malic acid was added to the culture. This increase resulted also in a reduction of the ratio of d-lactic acid to l-lactic acid (31.5 to 1.23). Addition of malic acid increased 6.5-fold the specific activity of nicotinamide adenine dinucleotide (NAD)-linked l-lactate dehydrogenase and increased 3.2-fold that of NAD-linked d-lactate dehydrogenase. The Michaelis constant (K(m)) for NAD of the NAD-linked l-lactate dehydrogenase increased with the addition of malate, but no change was observed in the K(m) values for the respective d-enzyme. The effect of carboxylic acids on the NAD-linked l-lactate dehydrogenase activities was tested by using partially purified enzyme preparations from cells grown with glucose alone and from cells grown with glucose plus malate. Malate stimulated the l-enzyme and inhibited the d-lactate dehydrogenase. The NAD-linked l-lactate dehydrogenase exhibited the same activity bands on polyacrylamide gel electrophoresis whether the cell-free preparation originated from cells grown on glucose plus malate or on glucose as the sole carbon source. The NAD-linked d-lactate dehydrogenase, however, exhibited a different pattern of electrophoretic mobility, depending upon the source of origin of the cell-free preparation. The results suggest that malate has a stimulatory effect on the synthesis of both enzymes and may result in rearrangement of the protein structure of the d-lactate dehydrogenase. This rearrangement apparently makes the d-enzyme more susceptible to inhibition of catalytic activity. The l-lactate dehydrogenase, however, is stimulated not only in its synthesis but also in its activity. It is proposed that these effects are responsible for the regulation of lactic acid production.  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