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.
An NAD+-linked, reduced glutathione-dependent formaldehyde dehydrogenase was purified to homogeneity from soluble extracts of methanol-grown yeast, Pichia sp. Formaldehyde and methylglyoxal are oxidized in the presence of NAD+ as an electron acceptor. NADP+ could not replace NAD+. Other straight chain aldehydes (C2–C6 tested), branched-chain aldehydes (e.g., isobutyaldehyde), aromatic aldehydes (e.g., salicylal-dehyde, benzaldehyde), glutyraldehyde, glyceraldehyde, glycoaldehyde, and glyoxal-dehyde tested were not oxidized by the purified formaldehyde dehydrogenase. The product of formaldehyde oxidation by purified enzyme was demonstrated to be S-for-mylglutathione by measuring the absorption at 240 nm due to the formation of thioester of formaldehyde and reduced glutathione. The Km values for NAD+, formaldehyde, and reduced glutathione were 0.12, 0.31, and 0.16 mm, respectively, for the forward reaction at pH 8.0. The purified formaldehyde dehydrogenase also catalyzed the reduction of S-formylglutathione in the presence of NADH. Formate was not reduced by the purified enzyme. The Km values for S-formylglutathione and NADH were 0.60 and 0.25 mm, respectively, for the reverse reaction at pH 6.0. Formaldehyde dehydrogenase has a molecular weight of 84,000 as determined by gel filtration and subunit molecular weight of 41,000 as determined by sodium dodecyl sulfate-gel electrophoresis. S-Formylglutathione, a product of formaldehyde oxidation, was oxidized by the partially purified formate dehydrogenase from Pichia sp. Formate dehydrogenase has a higher affinity toward S-formylglutathione (Km value 1.8 mm) than toward formate (Km value 25 mm). Antiserum prepared against the purified formaldehyde dehydrogenase from Pichia sp. NRRL-Y-11328 forms strong precipitin bands with isofunctional enzymes from methanol-grown Pichia pastoris NRRL-Y-7556 and Torulopsis candida Y-11419 and weak precipitin bands with Hansenula polymorpha NRRL-Y-2214. No cross-reaction was observed with isofunctional enzyme derived from methanol-grown Kloeckera sp.  相似文献   

2.
Alcohol dehydrogenase (ADH) isozymes in annual sunflowers (Helianthus annuus) are dimers whose subunits are produced by two genes, Adh 1 and Adh 2 .The codominant F and S alleles of Adh 1 produce the slower-migrating set of three isozymes. The faster-migrating set of three isozymes is controlled by Adh 2 , which also has at least two alleles, F and S. Hybridization experiments indicated that the Adh 2 alleles segregate in expected Mendelian fashion and that Adh 1 and Adh 2 are not linked. A third common 1-locus allele is designated early (E) because when homozygous it results in a blank at the 1FF isozyme position in mature seeds, but in developing seeds produces a normal-appearing band at the 1FF position. Hybridization studies showed that the early alleles segregated normally. Correlation between genotype and presence or absence of isozymes electrophoretically intermediate between those of Adh 1 and Adh 2 suggests that four intergenic isozymes may be formed as a result of dimerization of the four basic subunits. Studies of zymograms of developing seeds suggest that the remaining but inconstant zymogram bands are mature seed isozymes which have altered charges during early morphogenesis and thus are developmental artifacts.  相似文献   

3.
C3 phenotypes were examined in species of Papio and Macaca. Baboons (P. cynocephalus) showed extensive polymorphism, with estimated gene frequencies of 0.815, 0.174, and 0.011 for C3*S, C3*F, and C3*F 1 alleles, respectively. Clear segregation patterns showing codominant inheritance were evident in family studies. Such extensive polymorphism was not observed in M. nemestrina or M. fascicularis. The S gene is the most common allele in all the species studied. The F gene is relatively common in baboons.  相似文献   

4.
Glucose dehydrogenase (E.C. 1.1.1.47) from B. megaterium M 1286 was immobilized together with mutarotase (E.C. 5.1.3.3) on several organic carriers and by different methods. The storage stability of the enzyme at pH-values > 6 is slightly improved by immobilization and the pH-optimum is shifted from 8.3 to 8.0. Kinetic constants of the immobilized enzyme are: KM(NAD+) = 5.36 × 10?4 mol/l KM(glucose) = 3.76 · 10?2 mol/l and Vmax = 5.54 · 10?5 mol/(l min g carrier) for the most active preparation (2.16 mg enzyme/g carrier). In reactor experiments the immobilized glucose dehydrogenase was used with glucose to regenerate NADPH in NADPH-dependent iron-III-protoporphyrin-IX-imidazole catalyzed hydroxylation and demethylation of model substrates of cytochrome P-450. The advantages of the coupling of both reactions with cofactor recycling are shown and discussed.  相似文献   

5.
Chambers  G. K. 《Biochemical genetics》1984,22(5-6):529-549
Alcohol dehydrogenase has been purified from Drosophila melanogaster lines bearing the Adh F, AdhS, and Adh FCh.D. alleles. Biochemical investigations show that the properties of the purified enzymes are very similar to those of crude enzyme extracts except that the pure enzymes are more heat stable. ADH-FCh.D. resembles ADH-S very closely in specific activity, substrate specificity, and a number of kinetic parameters including limiting values for K m(app.) for ethanol. However, it is considerably more heat stable than either of the two common variants. ADH-F differs from ADH-S and ADH-FCh.D. particularly with regard to the rate of oxidation of secondary alcohols. Atomic absorbtion spectroscopy shows that all three allozymes lack zine or other divalent cations as active-site components. Peptide mapping experiments identify one very active cysteinyl residue; and amide residues in the NAD+ binding domain.  相似文献   

6.
Glucose-6-phosphate dehydrogenase from Leuconostoc mesenteroides utilizes either NAD+ or NADP+ as coenzyme. Kinetic studies showed that NAD+ and NADP+ interact with different enzyme forms (Olive, C., Geroch, M. E., and Levy, H. R. (1971) J. Biol. Chem.246, 2047–2057). In the present study the techniques of fluorescence quenching and fluorescence enhancement were used to investigate the interaction between Leuconostoc mesenteroides glucose-6-phosphate dehydrogenase and coenzymes. In addition, kinetic studies were performed to examine interaction between the enzyme and various coenzyme analogs. The maximum quenching of protein fluorescence is 5% for NADP+ and 50% for NAD+. The dissociation constant for NADP+, determined from fluorescence quenching measurements, is 3 μm, which is similar to the previously determined Km of 5.7 μm and Ki of 5 μm. The dissociation constant for NAD+ is 2.5 mm, which is 24 times larger than the previously determined Km of 0.106 mm. Glucose 1-phosphate, a substrate-competitive inhibitor, lowers the dissociation constant and maximum fluorescence quenching for NAD+ but not for NADP+. This suggests that glucose 6-phosphate may act similarly and thus play a role in enabling the enzyme to utilize NAD+ under physiological conditions. When NADPH binds to the enzyme its fluorescence is enhanced 2.3-fold. The enzyme was titrated with NADPH in the absence and presence of NAD+; binding of these two coenzymes is competitive. The dissociation constant for NADPH from these measurements is 24 μm; the previously determined Ki is 37.6 μm. The dissociation constant for NAD′ is 2.8 mm, in satisfactory agreement with the value obtained from protein fluorescence quenching measurements. Various compounds which resemble either the adenosine or the nicotinamide portion of the coenzyme structure are coenzyme-competitive inhibitors; 2′,5′-ADP, the most inhibitory analog tested, gives NADP+-competitive and NAD+-noncompetitive inhibition, consistent with the kinetic mechanism previously proposed. By using pairs of coenzyme-competitive inhibitors it was shown in kinetic studies that the two portions of the NAD+ structure cannot be accommodated on the enzyme simultaneously unies they are covalently linked. Fluorescence studies showed that there are both “buried” and “exposed” tryptophan residues in the enzyme structure.  相似文献   

7.
We have isolated and purified two parental homodimers and a unique heterodimer of acid phosphatase [coded byAcph-1 1.05(F) andAcph-1 0.95(S)] from isogenic homozygotes and heterozygotes ofDrosophila malerkotliana. F andS produce qualitatively different allozymes and the two alleles are expressed equally within and across all three genotypes andF andS play an equal role in the epigenetics of dominance. Subunit interaction in the heterodimer over a wide range of H+ concentrations accounts for the epigenetics of dominance for enzyme activity  相似文献   

8.
Isopropylmalate dehydrogenase (IPMDH) is the third enzyme specific to leucine biosynthesis in microorganisms and plants, and catalyzes the oxidative decarboxylation of (2R,3S)-3-isopropylmalate to α-ketoisocaproate using NAD+ as an oxidizing agent. In this study, a thia-analogue of the substrate was designed and synthesized as an inhibitor for IPMDH. The analogue showed strong competitive inhibitory activity with Ki = 62 nM toward IPMDH derived from Thermus thermophilus. Moreover, the crystal structure of T. thermophilus IPMDH in a ternary complex with NAD+ and the inhibitor has been determined at 2.8 Å resolution. The inhibitor exists as a decarboxylated product with an enol/enolate form in the active site. The product interacts with Arg 94, Asn 102, Ser 259, Glu 270, and a water molecule hydrogen-bonding with Arg 132. All interactions between the product and the enzyme were observed in the position associated with keto-enol tautomerization. This result implies that the tautomerization step of the thia-analogue during the IPMDH reaction is involved in the inhibition.  相似文献   

9.
Apparent selection affecting starch gel electrophoretic alleles at the Esterase-2 locus of Drosophila buzzatii has been detected in laboratory and natural populations. Perturbation-reperturbation of allele frequencies in replicated laboratory populations attempts to test direct selective effects at the locus versus effects of linked loci. Sequential gel electrophoresis has identified more alleles within starch classes, and three of these alleles (within the a, b and c starch alleles) were used in cage population experiments. Allele a/1.00/1.00/1.00 was set up in 10 replicate populations with allele c/1.00/1.00/1.00, and in an independent 10 replicate populations with allele b/0.99/1.01/1.00. For each set, three reperturbations were done. Replicate populations generally showed similar patterns of allele frequency change and clear directionality: effects of selection, not drift. However, four populations deviated from their replicates, indicating dissipation of linkage disequilibrium. Estimates of pre-adult viability in the F2 of pair-wise crosses among 12 sequential gel electrophoretic alleles showed very variable modes of inheritance and relative viability fitnesses. Together with the diversity of patterns of allele frequency change in the cage populations, these results suggest a gene complex, with selection acting on an interacting set of loci which may include Esterase-2.  相似文献   

10.
Malate dehydrogenase (l-malate:NAD+ oxidoreductase, EC 1.1.1.37) has been purified about 480-fold from crude extract of the facultative phototrophic bacterium, Rhodopseudomonas capsulata by only two purification steps, involving Red-Sepharose affinity chromatography. The enzyme has a molecular mass of about 80 kDa and consists of two subunits with identical molecular mass (35 kDa). The enzyme is susceptible to heat inactivation and loses its activity completely upon incubation at 40°C for 10 min. Addition of NAD+, NADH and oxaloacetate, but not l-malate, to the enzyme solution stabilized the enzyme. The enzyme catalyzes exclusively the oxidation of l-malate, and the reduction of oxaloacetate and ketomalonate in the presence of NAD+ and NADH, respectively, as the coenzyme. The pH optima are around 9.5 for the l-malate oxidation, and 7.75–8.5 and 4.3–7.0 for the reduction of oxaloacetate and ketomalonate, respectively. The Km values were determined to be 2.1 mM for l-malate, 48 μM for NAD+, 85 μM for oxaloacetate, 25 μM for NADH and 2.2 mM for ketomalonate. Initial velocity and product inhibition patterns of the enzyme reactions indicate a random binding of the substrates, NAD+ and l-malate, to the enzyme and a sequential release of the products: NADH is the last product released from the enzyme in the l-malate oxidation.  相似文献   

11.
Schwartz D 《Genetics》1973,74(4):615-617
The method of high-resolution electrophoresis was employed to compare the relative amounts of enzyme produced by the Adh1F and Adh1S alleles in heterozygotes at different stages of development. The results are in complete agreement with those obtained from enzymatic analyses and support the competition hypothesis for the regulation of the alcohol dehydrogenase gene.  相似文献   

12.
A spontaneously occurring self-compatible mutant has been identified in Ipomoea trifida, a species possessing sporophytic self-incompatibility controlled by a single multiallelic S locus. Analysis of the segregation of compatibility/incompatibility phenotypes in selfed and crossed progenies of the self-compatible mutant plant indicated that the self-compatibility trait was caused by a mutation at the S locus; the mutated S allele was therefore designated Sc. RFLP analysis of progeny plants segregating for the Sc allele using the SSP gene (a gene linked closely to the S locus of I. trifida) as a probe confirmed that the mutation was present at the S locus. Self-incompatibility responses were examined in F1 progenies obtained from crosses between the self-compatible mutant and self-incompatible plants homozygous for one of three S alleles, S 1 , S 3 and S 22 , where the dominance relationship is S 22 >S 1 >S 3 . All F1 progeny plants from crosses with S 22 and S 1 homozygotes were self-incompatible and exhibited the respective phenotypes of each self-incompatible parent (either S 22 or S 1 ) in both stigma and pollen. However, of the F1 progeny plants from the cross with the S 3 homozygote, those carrying the genotype ScS 3 were all self-compatible and cross-compatible as both female and male parents with the S 3 homozygote. These results indicate that the dominance relationship between the four S alleles is: S 22 >S 1 >Sc>S 3 and so reveal the unexpected finding that the mutated Sc allele is dominant over a functional S 3 allele. A possible explanation for this observation is that the gene product encoded by the Sc allele confers a dominant negative effect on the S 3 gene product. Received: 21 June 2000 / Accepted: 18 July 2000  相似文献   

13.
FST is frequently used as a summary of genetic differentiation among groups. It has been suggested that FST depends on the allele frequencies at a locus, as it exhibits a variety of peculiar properties related to genetic diversity: higher values for biallelic single-nucleotide polymorphisms (SNPs) than for multiallelic microsatellites, low values among high-diversity populations viewed as substantially distinct, and low values for populations that differ primarily in their profiles of rare alleles. A full mathematical understanding of the dependence of FST on allele frequencies, however, has been elusive. Here, we examine the relationship between FST and the frequency of the most frequent allele, demonstrating that the range of values that FST can take is restricted considerably by the allele-frequency distribution. For a two-population model, we derive strict bounds on FST as a function of the frequency M of the allele with highest mean frequency between the pair of populations. Using these bounds, we show that for a value of M chosen uniformly between 0 and 1 at a multiallelic locus whose number of alleles is left unspecified, the mean maximum FST is ∼0.3585. Further, FST is restricted to values much less than 1 when M is low or high, and the contribution to the maximum FST made by the most frequent allele is on average ∼0.4485. Using bounds on homozygosity that we have previously derived as functions of M, we describe strict bounds on FST in terms of the homozygosity of the total population, finding that the mean maximum FST given this homozygosity is 1 − ln 2 ≈ 0.3069. Our results provide a conceptual basis for understanding the dependence of FST on allele frequencies and genetic diversity and for interpreting the roles of these quantities in computations of FST from population-genetic data. Further, our analysis suggests that many unusual observations of FST, including the relatively low FST values in high-diversity human populations from Africa and the relatively low estimates of FST for microsatellites compared to SNPs, can be understood not as biological phenomena associated with different groups of populations or classes of markers but rather as consequences of the intrinsic mathematical dependence of FST on the properties of allele-frequency distributions.DIFFERENTIATION among groups is one of the fundamental subjects of the field of population genetics. Comparisons of the level of variation among subpopulations with the level of variation in the total population have been employed frequently in population-genetic theory, in statistical methods for data analysis, and in empirical studies of distributions of genetic variation. Wright’s (Wright 1951) fixation indices, and FST in particular, have been central to this effort.Wright’s FST was originally defined as the correlation between two randomly sampled gametes from the same subpopulation when the correlation of two randomly sampled gametes from the total population is set to zero. Several definitions of FST or FST-like quantities are now available, relying on a variety of different conceptual formulations but all measuring some aspect of population differentiation (e.g., Charlesworth 1998; Holsinger and Weir 2009). Many authors have claimed that one or another formulation of FST is affected by levels of genetic diversity or by allele frequencies, either because the range of FST is restricted by these quantities or because these quantities affect the degree to which FST reflects population differentiation (e.g., Charlesworth 1998; Nagylaki 1998; Hedrick 1999, 2005; Long and Kittles 2003; Jost 2008; Ryman and Leimar 2008; Long 2009; Meirmans and Hedrick 2011). For example, Nagylaki (1998) and Hedrick (1999) argued that measures of FST may be poor measures of genetic differentiation when the level of diversity is high. Charlesworth (1998) suggested that FST can be inflated when diversity is low, arguing that FST might not be appropriate for comparing loci with substantially different levels of variation. In a provocative recent article, Jost (2008) used the diversity dependence of forms of FST to question their utility as differentiation measures at all.One definition that is convenient for mathematical assessment of the relationship of an FST-like quantity and allele frequencies is the quantity labeled GST by Nei (1973), which for a given locus measures the difference between the heterozygosity of the total (pooled) population, hT, and the mean heterozygosity across subpopulations, hS, divided by the heterozygosity of the total population:GST=hThShT.(1)In terms of the homozygosity of the total population, HT = 1 − hT, and the mean homozygosity across subpopulations, HS = 1 − hS, we can writeGST=HSHT1HT.(2)The Wahlund (1928) principle guarantees that HSHT and, therefore, because HS ≤ 1 and for a polymorphic locus with finitely many alleles, 0 < HT < 1, GST lies in the interval [0,1].Using GST for their definition of FST, Hedrick (1999, 2005) and Long and Kittles (2003) pointed out that because hT < 1, FST cannot exceed the mean homozygosity across subpopulations, HS:FST = 1 ? hS/hT < 1 ? hSHS.(3)Hedrick (2005) obtained this result by considering a set of K equal-sized subpopulations, in which each allele is private to a single subpopulation. In the limit as K → ∞, a stronger upper bound on FST as a function of HS and K reduces to Equation 3 (see also Jin and Chakraborty 1995 and Long and Kittles 2003).While Hedrick (1999, 2005) and Long and Kittles (2003) have clarified the relationship between FST and the mean homozygosity HS across subpopulations, their approaches do not easily illuminate the connection between FST and allele frequencies themselves. A formal understanding of the relationship between FST and allele frequencies would make it possible to more fully understand the behavior of FST in situations where markers of interest differ substantially in allele frequencies or levels of genetic diversity. Our recent work on the relationship between homozygosity and the frequency of the most frequent allele (Rosenberg and Jakobsson 2008; Reddy and Rosenberg 2012) provides a mathematical approach for formal investigation of bounds on population-genetic statistics in terms of allele frequencies. In this article, we therefore seek to thoroughly examine the dependence of FST on allele frequencies by investigating the upper bound on FST in terms of the frequency M of the most frequent allele across a pair of populations. We derive bounds on FST given the frequency of the most frequent allele and bounds on the frequency of the most frequent allele given FST. We consider loci with arbitrarily many alleles in a pair of subpopulations. Using theory for the bounds on homozygosity given the frequency of the most frequent allele, we obtain strict bounds on FST given the homozygosity of the total population. Our analysis clarifies the relationships among FST, allele frequencies, and homozygosity, providing explanations for peculiar observations of FST that can be attributed to allele-frequency dependence.  相似文献   

14.
The electrochemistry for the reduction of tetrakis(dialkyl- and diphenyldithiocarbamato)-μ-oxodioxodirhenium complexes trans-Re2O3(S2CNR2)4 (R = methyl, ethyl, propyl, butyl and phenyl) was investigated in seven nonaqueous solvents. The complexes underwent a reversible reduction involving one- electron at a platinum electrode to [Re2O3(S2CNR2)4], which decomposed with the cleavage of the μ-oxo bridge to form ReO(S2CNR2)2, R2CNS2 and other rhenium complexes. The redox potential Eo′ of [Re2O3(S2CNR2)4]0/− couples and the stability of the reduction product [Re2O3(S2CNR2)4] depend on the R group. The Eo′ are appreciably solvent-dependent. The difference in Eo′ with solvents could be interpreted in terms of the solubility parameters.  相似文献   

15.
Using whole cell invertase of Saccharomyces pastorianus, entrapped in spherical agar pellets, sucrose hydrolysis was carried out in a continuously fed fluidized bed reactor. The effective rate of reaction determined experimentally for the catalytic pellet was correlated with particle radius (R), intraparticle concentration of enzyme (Ep) and external concentration of substrate (S R). The results were elucidated by theoretical analysis incorporating internal mass transfer resistance. At high degrees of diffusional resistance, the effectiveness factor was successfully estimted from Bischoff's equation. A dimensionless number, mA ? R(k2Ep/KmD)0.5(Km/(Km + S R)), was used conveniently to predict the effectiveness factor in those cases wher the intraparticle diffusional effect was less significant. This number was employed to determine critical pellet size for an optimal reaction. The relationship between the properties of the pellet (size and intraparticle enzyme activity) and its apparent kinetic constants (k2 and Km), estimated according to Lineweaver-Burk, are discussed.  相似文献   

16.
The nicotinamide adenine dinucleotide-specific glutamate dehydrogenase (l-glutamate:NAD+ oxidoreductase, EC 1.4.1.2) of Chlorella sorokiniana was purified 1,000-fold to electrophoretic homogeneity. The native enzyme was shown to have a molecular weight of 180,000 and to be composed of four identical subunits with a molecular weight of 45,000. The N-terminal amino acid was determined to be lysine. The pH optima for the aminating and deaminating reactions were approximately 8 and 9, respectively. The Km values for α-ketoglutarate, NADH, NH4+, NAD+, and l-glutamate were 2 mm, 0.15 mm, 40 mm, 0.15 mm, and 60 mm, respectively. Whereas the Km for α-ketoglutarate and l-glutamate increased 10-fold, 1 pH unit above or below the pH optima for the aminating or deaminating reactions, respectively, the Km values for NADH and NAD+ were independent of change in pH from 7 to 9.6. By initial velocity, product inhibition, and equilibrium substrate exchange studies, the kinetic mechanism of enzyme was shown to be consistent with a bi uni uni uni ping-pong addition sequence. Although this kinetic mechanism differs from that reported for any other glutamate dehydrogenase, the chemical mechanism still appears to involve the formation of a Schiff base between α-ketoglutarate and an ε-amino group of a lysine residue in the enzyme. The physical, chemical, and kinetic properties of this enzyme differ greatly from those reported for the NH4+-inducible glutamate dehydrogenase in this organism.  相似文献   

17.
15-Hydroxyprostaglandin dehydrogenase from bovine lung was purified 7.4 times to a specific activity of 1.4 mU/mg of protein. The isoelectric point was estimated to 5.4 and the molecular weight by gelfiltration to 40,000. Km for prostaglandin E1 and for NAD+ were found to be 3.4 μM and 1.1 × 10?4M respectively. The enzyme was inhibited by indomethacin and aspirin. The indomethacin inhibition was found to be non-competitive to prostaglandin E1 having a Ki=1.4 × 10?4M and a Ki=1.6 × 10?5M.  相似文献   

18.
In order to investigate the enzymatic mechanism of tomato alcohol dehydrogenase, kinetic studies were carried out at pH 5.8 and 9.4 for the forward and reverse reactions, respectively. Primary double reciprocal plots for several fixed concentrations of the associated substrate in all cases intersect, suggesting a sequential mechanism. Exploitation of secondary plots (slope-intercept values on the primary plots versus the reciprocals of the non-varied substrates) gives the following values: Kms 500 μM for MeCHO, 30 μM for NADH, 2700 μM for EtOH, 12 μM for NAD+; Kis 40 μM for MeCHO, 3 μM for NADH, 104 μM for EtOH and 45 μM for NAD+. The results obtained in product inhibition studies agree with an ordered bi-bi mechanism for both forward and reverse reactions. Application of Cleland's rules shows that the coenzyme was the first substrate to complex with the enzyme in both cases.  相似文献   

19.
Cytosolic malate dehydrogenase from human liver was isolated and its physical and kinetic properties were determined. The enzyme had a molecular weight of 72,000 ± 2000 and an amino acid composition similar to those of malate dehydrogenases from other species. The kinetic behaviour of the enzyme was consistent with an Ordered Bi Bi mechanism. The following values (μm) of the kinetic parameters were obtained at pH 7.4 and 37 °C: Ka, 17; Kia, 3.6; Kb, 51; Kib, 68; Kp, 770; Kip, 10,700; Kq, 42; Kiq, 500, where a, b, p, and q refer to NADH, oxalacetate, malate, and NAD+, respectively. The maximum velocity of the enzyme in human liver homogenates was 102 μmol/min/g wet wt of liver for oxalacetate reduction and 11.2 μmol/min/g liver for malate oxidation at pH 7.4 and 37 °C. Calculations using these parameters showed that, under conditions in vivo, the rate of NADH oxidation by the enzyme would be much less than the maximum velocity and could be comparable to the rate of NADH production during ethanol oxidation in human liver. The rate of NADH oxidation would be sensitive to the concentrations of NADH and oxalacetate; this sensitivity can explain the change in cytosolic NAD+NADH redox state during ethanol metabolism in human liver.  相似文献   

20.
Self-incompatibility in Brassica campestris c.v. Arlo is controlled by a single locus sporophytic system. The identity and expression of the S alleles were determined in eight inbred and two hybrid families. It was found that co-dominance of alleles is more frequent in the stigma, whereas dominance relations between pairs of alleles predominate in the pollen. A linear order of dominance was established between six S alleles and alleles high, intermediate and low in the dominance series were recognized.In considering the variation in the expression of compatibility and the segregation ratios in inbred, F1, F2 and backcross progenies, the presence of a specific S allele conditioning self-fertility, or a single dominant self-compatibility factor independent of the S locus could not be established. Instead, self-compatibility in this cultivar was ascribed to the segregation of a polygenic complex which is capable of modifying the incompatibility reaction to the point of self-fertility, or to a reduction in the strength of the reaction due to the presence of S alleles low in the dominance series.  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