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 31 毫秒
1.
NADP-Utilizing Enzymes in the Matrix of Plant Mitochondria   总被引:9,自引:4,他引:5       下载免费PDF全文
Purified potato tuber (Solanum tuberosum L. cv Bintie) mitochondria contain soluble, highly latent NAD+- and NADP+-isocitrate dehydrogenases, NAD+- and NADP+-malate dehydrogenases, as well as an NADPH-specific glutathione reductase (160, 25, 7200, 160, and 16 nanomoles NAD(P)H per minute and milligram protein, respectively). The two isocitrate dehydrogenase activities, but not the two malate dehydrogenase activities, could be separated by ammonium sulfate precipitation. Thus, the NADP+-isocitrate dehydrogenase activity is due to a separate matrix enzyme, whereas the NADP+-malate dehydrogenase activity is probably due to unspecificity of the NAD+-malate dehydrogenase. NADP+-specific isocitrate dehydrogenase had much lower Kms for NADP+ and isocitrate (5.1 and 10.7 micromolar, respectively) than the NAD+-specific enzyme (101 micromolar for NAD+ and 184 micromolar for isocitrate). A broad activity optimum at pH 7.4 to 9.0 was found for the NADP+-specific isocitrate dehydrogenase whereas the NAD+-specific enzyme had a sharp optimum at pH 7.8. Externally added NADP+ stimulated both isocitrate and malate oxidation by intact mitochondria under conditions where external NADPH oxidation was inhibited. This shows that (a) NADP+ is taken up by the mitochondria across the inner membrane and into the matrix, and (b) NADP+-reducing activities of malate dehydrogenase and the NADP+-specific isocitrate dehydrogenase in the matrix can contribute to electron transport in intact plant mitochondria. The physiological relevance of mitochondrial NADP(H) and soluble NADP(H)-consuming enzymes is discussed in relation to other known mitochondrial NADP(H)-utilizing enzymes.  相似文献   

2.
A monomeric NADP-dependent isocitrate dehydrogenase from the multicellular prokaryote Streptomyces avermitilis MA-4680 (SaIDH) was heteroexpressed in Escherichia coli, and the His-tagged enzyme was further purified to homogeneity. The molecular weight of SaIDH was about 80 kDa which is typical for monomeric isocitrate dehydrogenases. Structure-based sequence alignment reveals that the deduced amino acid sequence of SaIDH shows high sequence identity with known momomeric isocitrate dehydrogenase, and the coenzyme, substrate and metal ion binding sites are completely conserved. The optimal pH and temperature of SaIDH were found to be pH 9.4 and 45°C, respectively. Heat-inactivation studies showed that heating for 20 min at 50°C caused a 50% loss in enzymatic activity. In addition, SaIDH was absolutely specific for NADP+ as electron acceptor. Apparent K m values were 4.98 μM for NADP+ and 6,620 μM for NAD+, respectively, using Mn2+ as divalent cation. The enzyme performed a 33,000-fold greater specificity (k cat/K m) for NADP+ than NAD+. Moreover, SaIDH activity was entirely dependent on the presence of Mn2+ or Mg2+, but was strongly inhibited by Ca2+ and Zn2+. Taken together, our findings implicate the recombinant SaIDH is a divalent cation-dependent monomeric isocitrate dehydrogenase which presents a remarkably high cofactor preference for NADP+.  相似文献   

3.
Tryptic peptides which account for all five cysteinyl residues in ribulosebisphosphate carboxylase/oxygenase from Rhodospirillum rubrum have been purified and sequenced. Collectively, these peptides contain 94 of the approximately 500 amino acid residues per molecule of subunit. Due to one incomplete cleavage at a site for trypsin and two incomplete chymotryptic-like cleavages, eight major radioactive peptides (rather than five as predicted) were recovered from tryptic digests of the enzyme that had been carboxymethylated with [3H]iodoacetate. The established sequences are: GlyTyrThrAlaPheValHisCys1Lys TyrValAspLeuAlaLeuLysGluGluAspLeuIleAla GlyGlyGluHisValLeuCys1AlaTyr AlaGlyTyrGlyTyrValAlaThrAlaAlaHisPheAla AlaGluSerSerThrGlyThrAspValGluValCys1 ThrThrAsxAsxPheThrArg AlaCys1ThrProIleIleSerGlyGlyMetAsnAla LeuArg ProPheAlaGluAlaCys1HisAlaPheTrpLeuGly GlyAsnPheIleLys In these peptides, radioactive carboxymethylcysteinyl residues are denoted with asterisks and the sites of incomplete cleavage with vertical wavy lines. None of the peptides appear homologous with either of two cysteinyl-containing, active-site peptides previously isolated from spinach ribulosebisphosphate carboxylase/oxygenase.  相似文献   

4.
Isocitrate dehydrogenase (IDH) is one of the key enzymes in the citric acid cycle, which involves in providing energy and biosynthetic precursors for metabolism. Here, we report for the first time the enzymatic characterization of a monomeric NADP+-dependent IDH from Streptomyces lividans TK54 (SlIDH). The icd gene (GenBank database accession number EU661252) encoding IDH was cloned and overexpressed in Escherichia coli. The molecular mass of SlIDH was about 80 kDa, typical of a monomeric NADP-IDH, and showed high amino acid sequence identity with known monomeric IDHs. The optimal activity of the 6His-tagged SlIDH was found at pH values 8.5 (Mn2+) and 9.0 (Mg2+), and the optimal temperature was around 46 °C. Heat-inactivation studies showed that about 50% SlIDH activity was preserved at 38 °C after 20 min of incubation. The recombinant SlIDH displayed a 62,000-fold (kcat/Km) preference for NADP+ over NAD+ with Mn2+, and a 85,000-fold greater specificity for NADP+ than NAD+ with Mg2+. Therefore, SlIDH is a divalent cation-dependent monomeric IDH with remarkably high coenzyme preference for NADP+.  相似文献   

5.
The aim of the present study was the investigation of the occurrence of NADPH-generating pathways in the endoplasmic reticulum others then hexose-6-phosphate dehydrogenase. A significant isocitrate and a moderate malate-dependent NADP+ reduction were observed in endoplasmic reticulum-derived rat liver microsomes. The isocitrate-dependent activity was very likely attributable to the appearance of the cytosolic isocitrate dehydrogenase isozyme in the lumen. The isocitrate dehydrogenase activity of microsomes was present in the luminal fraction; it showed a strong preference towards NADP+versus NAD+, and it was almost completely latent. Antibodies against the cytosolic isoform of isocitrate dehydrogenase immunorevealed a microsomal protein of identical molecular weight; the microsomal enzyme showed similar kinetic parameters and oxalomalate inhibition as the cytosolic one. Measurable luminal isocitrate dehydrogenase activity was also present in microsomes from rat epididymal fat. The results suggest that isocitrate dehydrogenase is an important NADPH-generating enzyme in the endoplasmic reticulum.  相似文献   

6.
NADP+ dependent isocitrate dehydrogenase (IDH) is an enzyme catalyzing oxidative decarboxylation of isocitrate into oxalosuccinate (intermediate) and finally the product α-ketoglutarate. The crystal structure of Thermus thermophilus isocitrate dehydrogenase (TtIDH) ternary complex with citrate and cofactor NADP+ was determined using X-ray diffraction method to a resolution of 1.80 Å. The overall fold of this protein was resolved into large domain, small domain and a clasp domain. The monomeric structure reveals a novel terminal domain involved in dimerization, very unique and novel domain when compared to other IDH’s. And, small domain and clasp domain showing significant differences when compared to other IDH’s of the same sub-family. The structure of TtIDH reveals the absence of helix at the clasp domain, which is mainly involved in oligomerization in other IDH’s. Also, helices/beta sheets are absent in the small domain, when compared to other IDH’s of the same sub family. The overall TtIDH structure exhibits closed conformation with catalytic triad residues, Tyr144-Asp248-Lys191 are conserved. Oligomerization of the protein is quantized using interface area and subunit–subunit interactions between protomers. Overall, the TtIDH structure with novel terminal domain may be categorized as a first structure of subfamily of type IV.  相似文献   

7.
The cofactor-binding site of the NAD+-dependent Arabidopsis thaliana aldehyde dehydrogenase ALDH3H1 was analyzed to understand structural features determining cofactor-specificity. Homology modeling and mutant analysis elucidated important amino acid residues. Glu149 occupies a central position in the cofactor-binding cleft, and its carboxylate group coordinates the 2′- and 3′-hydroxyl groups of the adenosyl ribose ring of NAD+ and repels the 2′-phosphate moiety of NADP+. If Glu149 is mutated to Gln, Asp, Asn or Thr the binding of NAD+ is altered and rendered the enzyme capable of using NADP+. This change is attributed to a weaker steric hindrance and elimination of the electrostatic repulsion force of the 2′-phosphate of NADP+. Simultaneous mutations of Glu149 and Ile200, which is situated opposite of the cofactor binding cleft, improved the enzyme efficiency with NADP+. The double mutant ALDH3H1Glu149Thr/Ile200Val showed a good catalysis with NADP+. Subsequently a triple mutation was generated by replacing Val178 by Arg in order to create a “closed” cofactor binding site. The cofactor specificity was shifted even further in favor of NADP+, as the mutant ALDH3H1E149T/V178R/I200V uses NADP+ with almost 7-fold higher catalytic efficiency compared to NAD+. Our experiments suggest that residues occupying positions equivalent to 149, 178 and 200 constitute a group of amino acids in the ALDH3H1 protein determining cofactor affinity.  相似文献   

8.
Enzymes from several metabolic pathways were studied quantitatively in homogenates and in homogeneous areas of frozen-dried cryostat sections of an experimental, mouse ependymoblastoma, mouse mammary carcinoma, and mouse melanoma, growing as transplants in mouse brain. Micro analyses were performed in fiveto sixfold replicates on portions of tumour with a dry weight of 0·03-0·2 μg. A close resemblance of the enzyme spectrum of the ependymoblastoma to that of immature brain was noted. Hexokinase and malate dehydrogenase were lower and lactate, glucose-6-phosphate, and NADP+-linked isocitrate dehydrogenases, and β-glucuronidase higher in the ependymoblastoma than in whole, adult mouse brain. Mouse mammary carcinoma had a higher level of hexokinase and lower levels of lactate and glucose-6-phosphate dehydrogenases and β-glucuronidase than ependymoblastoma. The concentration of malate dehydrogenase was lower and that of lactate, glucosed-6-phosphate, and NADP+-linked isocitric dehydrogenases was higher in the melanoma than in the ependymoblastoma. β-Glucuronidase levels were similar in these two neoplasms. It is suggested that the relatively high levels of several NADP+-linked enzymes in the ependymoblastoma may be related to increased capacity for lipid synthesis.  相似文献   

9.
The subcellular location of NADP+-isocitrate dehydrogenase was investigated by preparing protoplasts from leaves of pea seedlings. Washed protoplasts were gently lysed and the whole lysate separated on sucrose gradients by a rate-zonal centrifugation. Organelles were located by marker enzymes and chlorophyll analysis. Most of the NADP+-isocitrate dehydrogenase was in the soluble fraction. About 10% of the NADP+-isocitrate dehydrogenase was present in the chloroplasts as a partially latent enzyme. Less than 1% of the activity was found associated with the peroxisome fraction. NADP+-isocitrate dehydrogenase was partially characterized from highly purified chloroplasts isolated from shoot homogenates. The enzyme exhibited apparent Km values of 11 micromolar (NADP+), 35 micromolar (isocitrate), 78 micromolar (Mn2+), 0.3 millimolar (Mg2+) and showed optimum activity at pH 8 to 8.5 with Mn2+ and 8.8 to 9.2 with Mg2+. The NADP+-isocitrate dehydrogenase activity previously claimed in the peroxisomes by other workers is probably due to isolation procedures and/or nonspecific association. The NADP+-isocitrate dehydrogenase activity in the chloroplasts might help supply α-ketoglutarate for glutamate synthase action.  相似文献   

10.
Summary The utilization by yeast of two carbon sources is carried out through the operation of the glyoxylic acid cycle. Kinetic data from the isocitrate transforming enzymes suggest that the flow of isocitrate through the glyoxylic acid cycle depends upon the inhibition of the isocitrate decarboxylating enzymes. Both isocitrate dehydrogenases are inhibited by a mixture of glyoxylate + oxaloacetate, but for the reasons described in the text we consider that this inhibition is of no physiological significance. On the other hand, we have found that NADPH is a competitive inhibitor of NADP-isocitrate dehydrogenase with respect to NADP+, with a KI similar to its KM. It also produces an additive effect on the NADH-produced inhibition of NAD-isocitrate dehydrogenase. We propose NADPH as the compound that channels the utilization of isocitrate into the glyoxylic acid cycle. This is supported by the finding of an increased NADPH/NADP+ ratio in acetate grown yeast with respect to glucose grown cells.  相似文献   

11.
A thermostable homodimeric isocitrate dehydrogenase from the hyperthermophilic sulfate-reducing archaeon Archaeoglobus fulgidus was purified and characterized. The mol. mass of the isocitrate dehydrogenase subunit was 42 kDa as determined by SDS-PAGE. Following separation by SDS-PAGE, A. fulgidus isocitrate dehydrogenase could be renatured and detected in situ by activity staining. The enzyme showed dual coenzyme specificity with a high preference for NADP+. Optimal temperature for activity was 90° C or above, and a half-life of 22 min was found for the enzyme when incubated at 90° C in a 50 mM Tricine-KOH buffer (pH 8.0). Based on the N-terminal amino acid sequence, the gene encoding the isocitrate dehydrogenase was cloned. DNA sequencing identified the icd gene as an open reading frame encoding a protein of 412 amino acids with a molecular mass corresponding to that determined for the purified enzyme. The deduced amino acid sequence closely resembled that of the isocitrate dehydrogenase from the archaeon Caldococcus noboribetus (59% identity) and bacterial isocitrate dehydrogenases, with 57% identity with isocitrate dehydrogenase from Escherichia coli. All the amino acid residues directly contacting substrate and coenzyme (except Ile-320) in E. coli isocitrate dehydrogenase are conserved in the enzyme from A. fulgidus. The primary structure of A. fulgidus isocitrate dehydrogenase confirmes the presence of Bacteria-type isocitrate dehydrogenases among Archaea. Multiple alignment of all the available amino acid sequences of di- and multimeric isocitrate dehydrogenases from the three domains of life shows that they can be divided into three distinct phylogenetic groups. Received: 6 February 1997 / Accepted: 12 June 1997  相似文献   

12.
1. Aerobically grown yeast having a high activity of glyoxylate-cycle, citric acid-cycle and electron-transport enzymes was transferred to a medium containing 10% glucose. After a lag phase of 30min. the yeast grew exponentially with a mean generation time of 94min. 2. The enzymes malate dehydrogenase, isocitrate lyase, succinate–cytochrome c oxidoreductase and NADH–cytochrome c oxidoreductase lost 45%, 17%, 27% and 46% of their activity respectively during the lag phase. 3. When growth commenced pyruvate kinase, pyruvate decarboxylase, alcohol dehydrogenase, glutamate dehydrogenase (NADP+-linked) and NADPH–cytochrome c oxidoreductase increased in activity, whereas aconitase, isocitrate dehydrogenase (NAD+- and NADP+-linked), α-oxoglutarate dehydrogenase, fumarase, malate dehydrogenase, succinate–cytochrome c oxidoreductase, NADH–cytochrome c oxidoreductase, NADH oxidase, NADPH oxidase, cytochrome c oxidase, glutamate dehydrogenase (NAD+-linked), glutamate–oxaloacetate transaminase, isocitrate lyase and glucose 6-phosphate dehydrogenase decreased. 4. During the early stages of growth the loss of activity of aconitase, α-oxoglutarate dehydrogenase, fumarase and glucose 6-phosphate dehydrogenase could be accounted for by dilution by cell division. The lower rate of loss of activity of isocitrate dehydrogenase (NAD+- and NADP+-linked), glutamate dehydrogenase (NAD+-linked), glutamate–oxaloacetate transaminase, NADPH oxidase and cytochrome c oxidase implies their continued synthesis, whereas the higher rate of loss of activity of malate dehydrogenase, isocitrate lyase, succinate–cytochrome c oxidoreductase, NADH–cytochrome c oxidoreductase and NADH oxidase means that these enzymes were actively removed. 5. The mechanisms of selective removal of enzyme activity and the control of the residual metabolic pathways are discussed.  相似文献   

13.
The biotin-containing tryptic peptides of pyruvate carboxylase from sheep, chicken, and turkey liver mitochondria have been isolated and their primary structures determined. The amino acid sequences of the 19 residue peptides from chicken and turkey are identical and share a common sequence of 14 residues around biocytin with the 24-residue peptide isolated from sheep. The sequences obtained were: residue 1 → 11 Avian: Gly Ala Pro Leu Val Leu Ser Ala Met Biocytin Met Sheep: Gly Gln Pro Leu Val Leu Ser Ala Met Biocytin Met residues 12 → 19 or 24 Avian: Glu Thr Val Val Thr Ala Pro Arg Sheep: Glu Thr Val Val Thr Ser Pro Val Thr Glu Gly Val Arg A sensitive radiochemical assay for biotin was developed based on the tight binding of biotin by avidin. The ability of zinc sulfate to precipitate, without dissociating, the avidin-biotin complex provided a convenient procedure for separating free and bound biotin, and hence, for back-titrating a standard amount of avidin with [14C]biotin.  相似文献   

14.
The kinetics of a Mn2+-requiring, NADP+-specific isocitrate dehydrogenase from Salmonella typhimurium have been examined by the measurement of initial velocity rates in the presence and absence of the reaction products. The binding of each of the cosubstrates, isocitrate, and NADP+, is not independent of the other, and the isocitrate-Mn2+ complex is the kinetically important substrate species. All of the reaction products, α-ketoglutarate, CO2, and NADPH are competitive with both cosubstrates and the mechanism appears to be of the rapid equilibrium random type. The enzyme has been purified to homogeneity and has an isoelectric point at pH 4.0–4.2, and an apparent molecular weight of 102,000.  相似文献   

15.
Two isoenzymes of glucose 6-phosphate dehydrogenase (EC 1.1.1.49) have been separated from the plant fraction of soybean (Glycine max L. Merr. cv Williams) nodules by a procedure involving (NH4)2SO4 gradient fractionation, gel chromatography, chromatofocusing, and affinity chromatography. The isoenzymes, which have been termed glucose 6-phosphate dehydrogenases I and II, were specific for NADP+ and glucose 6-phosphate and had optimum activity at pH 8.5 and pH 8.1, respectively. Both isoenzymes were labile in the absence of NADP+. The apparent molecular weight of glucose 6-phosphate dehydrogenases I and II at pH 8.3 was estimated by gel chromatography to be approximately 110,000 in the absence of NADP+ and double this size in the presence of NADP+. The apparent molecular weight did not increase when glucose 6-phosphate was added with NADP+ at pH 8.3. Both isoenzymes had very similar kinetic properties, displaying positive cooperativity in their interaction with NADP+ and negative cooperativity with glucose 6-phosphate. The isoenzymes had half-maximal activity at approximately 10 micromolar NADP+ and 70 to 100 micromolar glucose 6-phosphate. NADPH was a potent inhibitor of both of the soybean nodule glucose 6-phosphate dehydrogenases.  相似文献   

16.
NADP+:isocitrate dehydrogenase has been purified to homogeneity from lactating bovine mammary gland. Purification was achieved through the use of affinity and DEAE-cellulose chromatography. The isolated enzyme gives one band when stained for protein or enzyme activity on discontinuous alkaline gel electrophoresis. The enzyme has a molecular weight of 55,000 as estimated by sodium dodecyl sulfate-gel electrophoresis and a Stokes radius of 4.1 nm as measured by gel chromatography. The enzyme will not use NAD+ in place of NADP+ and has an absolute requirement for divalent cations. The apparent Km values for dl-isocitrate, Mn2+, and NADP+ were found to be 8, 6, and 11 μm, respectively. The Mn2+-ds-isocitrate complex is the most likely substrate for the mammary enzyme with a Km of 3 μm. The properties of mammary NADP+:isocitrate dehydrogenase are compared with those of the homologous enzymes from pig heart and bovine liver, and its characteristics are discussed with respect to the function of the enzyme in lactating mammary gland.  相似文献   

17.
The subcellular distribution of NADP+ and NAD+-dependent glucose-6-phosphate and galactose-6-phosphate dehydrogenases were studied in rat liver, heart, brain, and chick brain. Only liver particulate fractions oxidized glucose-6-phosphate and galactose-6-phosphate with either NADP+ or NAD+ as cofactor. While all of the tissues examined had NADP+-dependent glucose-6-phosphate dehydrogenase activity, only rat liver and rat brain soluble fractions had NADP+-dependent galactose-6-phosphate dehydrogenase activity. Rat liver microsomal and rat brain soluble galactose-6-phosphate dehydrogenase activities were kinetically different (Km's 0.5 mm and 10 mm, respectively, for galactose-6-phosphate), although their reaction products were both 6-phosphogalactonate. Rat brain subcellular fractions did not oxidize 6-phosphogalactonate with either NADP+ or NAD+ cofactors but phosphatase activities hydrolyzing 6-phosphogalactonate, galactose-6-phosphate and galactose-1-phosphate were found in crude brain homogenates. In addition, galactose-6-phosphate and 6-phosphogalactonate were tested as inhibitors of various enzymes, with largely negative results, except that 6-phosphogalactonate was a competitive inhibitor (Ki = 0.5 mM) of rat brain 6-phosphogluconate dehydrogenase.  相似文献   

18.
The metal activator sites of NAD+-dependent and NADP+-dependent isocitrate dehydrogenases from pig heart have been probed using 113Cd- and 25Mg-NMR. In the presence of isocitrate and ADP, a broad resonance for cadmium bound to NAD-dependent isocitrate dehydrogenase was observed ( −8 ppm) arising from exchange with isocitrate (−20 ppm) and/or ADP (27 ppm) complexes. The Cd shift with ADP suggests interaction of the metal with the nucleotide ring nitrogen. Increasing shifts with excess ADP are indicative of macrochelate formation. 25Mg-NMR demonstrates that, unlike manganese, magnesium has a similar dissociation constant (1.8 mM) from NADP-dependent isocitrate dehydrogenase as from the enzyme-isocitrate complex (1.1 mM). The extrapolated line width of bound magnesium increases from 674 Hz in the binary complex to 10 200 Hz in the ternary complex. The quadrupole coupling constant, calculated from relaxation rates, is larger in the ternary complex. indicative of greater distortion in the magnesium coordination sphere. The line widths of magnesium complexed to NAD-dependent isocitrate dehydrogenase are broader, as expected for the larger octamer. 113Cd- and 25Mg-NMR both show that the metal sites have anisotropic octahedral symmetry. 25Mg relaxation rates yield correlation times corresponding to motions of a domain with motion independent of the enzyme multimers.  相似文献   

19.
The NAD+-dependent isocitrate dehydrogenase from etiolated pea (Pisum sativum L.) mitochondria was purified more than 200-fold by dye-ligand binding on Matrix Gel Blue A and gel filtration on Superose 6. The enzyme was stabilized during purification by the inclusion of 20% glycerol. In crude matrix extracts, the enzyme activity eluted from Superose 6 with apparent molecular masses of 1400 ± 200, 690 ± 90, and 300 ± 50 kD. During subsequent purification steps the larger molecular mass species disappeared and an additional peak at 94 ± 16 kD was evident. The monomer for the enzyme was tentatively identified at 47 kD by sodium dodecyl-polyacrylamide gel electrophoresis. The NADP+-specific isocitrate dehydrogenase activity from mitochondria eluted from Superose 6 at 80 ± 10 kD. About half of the NAD+ and NADP+-specific enzymes remained bound to the mitochondrial membranes and was not removed by washing. The NAD+-dependent isocitrate dehydrogenase showed sigmodial kinetics in response to isocitrate (S0.5 = 0.3 mm). When the enzyme was aged at 4°C or frozen, the isocitrate response showed less allosterism, but this was partially reversed by the addition of citrate to the reaction medium. The NAD+ isocitrate dehydrogenase showed standard Michaelis-Menten kinetics toward NAD+ (Km = 0.2 mm). NADH was a competitive inhibitor (Ki = 0.2 mm) and, unexpectedly, NADPH was a noncompetitive inhibitor (Ki = 0.3 mm). The regulation by NADPH may provide a mechanism for coordination of pyridine nucleotide pools in the mitochondria.  相似文献   

20.
A NADP+-specific isocitrate dehydrogenase (EC 1.1.1.42) was isolated and purified over 400-fold from Anacystis nidulans. The enzyme activity responded slowly to rapid changes in ligand (NADP+, isocitrate, Mg2+-ions) or enzyme concentration as well as to rapid changes in temperature. These are properties characteristic of the hysteretic enzymes. In addition, the enzyme activity was subject to product (-ketoglutarate) inhibition. ATP, ADP and CDP also inhibited the enzyme. Unlike several other cyanobacterial enzymes, the isocitrate dehydrogenase of Anacystis is not under redox control.  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