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 46 毫秒
1.
The preponderance of nonpolar contacts between CoA and chloramphenicol acetyltransferase in the high resolution structure of the binary complex prompted a study of selected hydrophobic residues by site-directed mutagenesis and steady-state kinetic analysis. Substitutions of three aromatic residues were used to evaluate binding contacts with the adenine moiety of CoA (Tyr-178), the pantetheine arm of the coenzyme (Tyr-56), and the S-acyl substituent (Phe-33). For those substitutions at residues 56 and 178 that cannot promote alternative polar interactions there is a correlation between residue hydrophobicity and the free energy of formation of the binary and ternary complexes of acetyl-CoA and chloramphenicol acetyltransferase and of the transition-state complex. Substitutions at Tyr-178 destabilize all such complexes to approximately the same extent (uniform binding changes), whereas those at Tyr-56 and Phe-33 cause differential binding changes, having a greater effect on the transition state than on either of the other complexes with acetyl-CoA.  相似文献   

2.
Phosphorylation and inactivation of acetyl-coenzyme A (CoA) carboxylase by acetyl-CoA carboxylase kinase in the presence of ATP and Mg2+ requires coenzyme A. Coenzyme A did not enhance the phosphorylation of alternative substrates of the carboxylase kinase such as protamine or histones. Analogs of coenzyme A were also effective in stimulating the inactivation of carboxylase. The KA of CoA for stimulated carboxylase inactivation was 25 microM. The presence of coenzyme A did not alter the Km of the carboxylase kinase for its substrates, ATP and acetyl-CoA carboxylase. Fluorescence binding studies showed that CoA binds to carboxylase but not to the kinase. The KD of CoA binding to carboxylase is 27 microM. These results indicate that coenzyme A, acting on acetyl-CoA carboxylase, may play an important role in the regulation of the covalent modification mechanism for acetyl-CoA carboxylase.  相似文献   

3.
Analogues of coenzyme A (CoA) and of CoA thioesters have been prepared in which the amide bond nearest the thiol group has been modified. An analogue of acetyl-CoA in which this amide bond is replaced with an ester linkage was a good substrate for the enzymes carnitine acetyltransferase, chloramphenicol acetyltransferase, and citrate synthase, with K(m) values 2- to 8-fold higher than those of acetyl-CoA and V(max) values from 14 to >80% those of the natural substrate. An analogue in which an extra methylene group was inserted between the amide bond and the thiol group showed less than 4-fold diminished binding to the three enzymes but exhibited less than 1% activity relative to acetyl-CoA with carnitine acetyltransferase and no measurable activity with the other two enzymes. Analogues of several CoA thioesters in which the amide bond was replaced with a hemithioacetal linkage exhibited no measurable activity with the appropriate enzymes. The results indicate that some aspects of the amide bond and proper distance between this amide and the thiol/thioester moiety are critical for activity of CoA ester-utilizing enzymes.  相似文献   

4.
Derivatives of ribonuclease A (RNase A) with modifications in positions 1 and/or 7 were prepared by subtilisin-catalyzed semisynthesis starting from synthetic RNase 1-20 peptides and S-protein (RNase 21-124). The lysyl residue at position 1 was replaced by alanine, whereas Lys-7 was replaced by cysteine that was specifically modified prior to semisynthesis. The enzymes obtained were characterized by protein chemical methods and were active toward uridylyl-3',5'-adenosine and yeast RNA. When Lys-7 was replaced by S-methyl-cysteine or S-carboxamido-contrast, the catalytic properties were only slightly altered. The dissociation constant for the RNase A-RI complex increased from 74 fM (RNase A) to 4.5 pM (Lys-1, Cys-7-methyl RNase), corresponding to a decrease in binding energy of 10 kJ mol-1. Modifications that introduced a positive charge in position 7 (S-aminoethyl- or S-ethylpyridyl-cysteine) led to much smaller losses. The replacement of Lys-1 resulted in a 4-kJ mol-1 loss in binding energy. S-protein bound to RI with Ki = 63.4 pM, 800-fold weaker than RNase A. This corresponded to a 16-kJ mol-1 difference in binding energy. The results show that the N-terminal portion of RNase A contributes significantly to binding of ribonuclease inhibitor and that ionic interactions of Lys-7 and to a smaller extent of Lys-1 provide most of the binding energy.  相似文献   

5.
The apparent binding energy for the interaction of the 3-hydroxyl group of chloramphenicol (CM) with the proposed general base (His-195) in chloramphenicol acetyltransferase (CAT) was determined by comparison of the dissociation constants of CM and 3-deoxyCM with CAT. The delta Gapp for this hydrogen bond to the N-3 of the imidazole ring is 1.5 kcal mol-1. Extending the use of modified ligands, in an approach which is complementary to that of site-directed mutagenesis, the binding affinity of each of a family of 3-halo-3-deoxychloramphenicol derivatives was observed to increase in the series F less than Cl less than Br less than I and is dominated by hydrophobic considerations. There is a linear free energy relationship between the dissociation constants for binding to CAT and an empirical hydrophobicity scale derived from reverse-phase HPLC retention times. The existence of such a relationship allows a true estimate of the total energetic contribution of interactions between the 3-hydroxyl group of CM and its contacts at the active site of CAT to be made on the basis of a regression analysis. The calculated value of delta Gbind (2.7 kcal mol-1) must include not only the hydrogen bond but also some favorable van der Waals interactions. The results demonstrate some of the advantages of an analysis of the energetics of ligand binding using modified ligands, in an approach that is formally analogous with and complementary to the use of site-directed mutations.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

6.
The imidazole of His-195 plays an essential role in the proposed general base mechanism of chloramphenicol acetyltransferase (CAT). The structure of the binary complex of CATIII and chloramphenicol suggests that two unusual interactions might determine the conformation of the side chain of His-195: (i) an intraresidue hydrogen bond between its main chain carbonyl and the protonated N delta 1 of the imidazole ring and (ii) face-to-face van der Waals contact between the His-195 imidazole group and the aromatic side chain of Tyr-25. Tyr-25 also makes a hydrogen bond, via its phenolic hydroxyl, to the carbonyl oxygen of the substrate chloramphenicol. Replacement of Tyr-25 of CATIII by phenylalanine results in a modest increase in the Km for chloramphenicol (from 11.6 to 14.6 microM) and a 2-fold fall in kcat (599 to 258 s-1), indicative of a free energy contribution to transition state binding of 0.6 kcal mol-1 for the hydrogen bond between Tyr-25 and chloramphenicol. In contrast, substitution of Tyr-25 by alanine yields an enzyme that is dramatically impaired in its ability to bind chloramphenicol (Km = 173 microM). As kcat for Ala-25 CAT is also reduced (130 s-1), the loss of the aryl group results in a 69-fold decrease in kcat/Km, corresponding to a free energy contribution to binding and catalysis of 2.5 kcal mol-1. In addition to the loss of the hydrogen bond between Tyr-25 and chloramphenicol, the loss of substrate affinity in Ala-25 CAT may be a direct consequence of reduced hydrophobicity of the chloramphenicol-binding site and/or the loss of critical constraints on the precise conformation of the catalytic imidazole. However, as with wild type CAT, inactivation of Ala-25 CAT by the affinity reagent 3-(bromoacetyl) chloramphenicol is accompanied by modification solely at N epsilon 2 of His-195. Hence, the results demonstrate that tautomeric stabilization of the imidazole ring persists in the absence of van der Waals interactions with the side chain of Tyr-25, probably as a consequence of hydrogen bonding between the protonated N delta 1 and the carbonyl oxygen of His-195.  相似文献   

7.
An analogue 2 of coenzyme A (CoA) has been prepared in which the geminal methyl groups are replaced with hydrogens. An NMR titration study was conducted and shifts in frequency of protons in the pantetheine portion of the molecule upon titration of the adenine base were observed as has been previously reported with CoA. These studies indicate that the geminal dimethyl groups are not essential for adoption of a partially folded conformation in solution. Based on 1H-1H coupling constants, the distribution of conformations about the carbon-carbon bonds in the region of the methyl deletion were estimated. The results suggest that the conformer distribution is similar to that of CoA, but with small increases in population of the anti conformers. A simple model compound containing the didemethyl pantoamide moiety was prepared and subjected to similar conformational analysis. The coupling constants and predicted conformer distribution were almost identical to that of the CoA analogue, indicating that the conformer distribution is controlled by local interactions and not influenced by interactions between distant parts of the CoA molecule. The acetyl derivative of 2 was a fairly good substrate for the acetyl-CoA utilizing enzymes carnitine acetyltransferase, chloramphenicol acetyltransferase, and citrate synthase, with 1.3- to 10-fold increased Km values and 2.5- to 11-fold decreases in Vmax. The combined results indicate that the geminal dimethyl groups of CoA have modest effects on function and minimal effects on conformation.  相似文献   

8.
The carbon monoxide dehydrogenase (CODH) complex from Methanosarcina thermophila catalyzed the synthesis of acetyl coenzyme A (acetyl-CoA) from CH3I, CO, and coenzyme A (CoA) at a rate of 65 nmol/min/mg at 55 degrees C. The reaction ended after 5 min with the synthesis of 52 nmol of acetyl-CoA per nmol of CODH complex. The optimum temperature for acetyl-CoA synthesis in the assay was between 55 and 60 degrees C; the rate of synthesis at 55 degrees C was not significantly different between pHs 5.5 and 8.0. The rate of acetyl-CoA synthesis was independent of CoA concentrations between 20 microM and 1 mM; however, activity was inhibited 50% with 5 mM CoA. Methylcobalamin did not substitute for CH3I in acetyl-CoA synthesis; no acetyl-CoA or propionyl coenzyme A was detected when sodium acetate or CH3CH2I replaced CH3I in the assay mixture. CO could be replaced with CO2 and titanium(III) citrate. When CO2 and 14CO were present in the assay, the specific activity of the acetyl-CoA synthesized was 87% of the specific activity of 14CO, indicating that CO was preferentially incorporated into acetyl-CoA without prior oxidation to free CO2. Greater than 100 microM potassium cyanide was required to significantly inhibit acetyl-CoA synthesis, and 500 microM was required for 50% inhibition; in contrast, oxidation of CO by the CODH complex was inhibited 50% by approximately 10 microM potassium cyanide.  相似文献   

9.
Studies of the reactivity of succinyl-CoA:3-keto acid CoA transferase with a small coenzyme A analog, methylmercaptopropionate, have shown that noncovalent interactions between the enzyme and the side chain of CoA are responsible for a rate acceleration of approximately 10(12), which is close to the total rate acceleration brought about by the enzyme (Moore, S. A., and Jencks, W. P. (1982) J. Biol. Chem. 257, 10893-10907). We report here that interaction between the enzyme and the pantetheine moiety of CoA provides the majority of the rate acceleration and destabilization of the enzyme-thiol ester intermediate that is observed with CoA substrates. The role of the adenosine 3'-phosphate 5'-diphosphate moiety of CoA is to provide 6.9 kcal/mol of binding energy in order to pull the pantetheine moiety into the active site. The enzyme-thiol ester intermediate, E-pantetheine, was generated by reaction of pantetheine with the thiol ester of enzyme and methylmercaptopropionate. E-Pantetheine undergoes hydrolysis with khyd = 2 min-1, 140-fold faster than E-CoA, and reacts with acetoacetate with kAcAc = 3 X 10(6) M-1 min-1, only 10-fold slower than E-CoA. However, in the reverse direction acetoacetylpantetheine reacts with CoA transferase (kAcAc-SP = 220 M-1 min-1) 1.6 X 10(6) times slower than acetoacetyl-CoA. The equilibrium constant for the reaction of pantetheine with E-CoA is approximately 8 X 10(-6).  相似文献   

10.
Pantothenate kinase catalyzes a key regulatory step in coenzyme A biosynthesis, and there are four mammalian genes that encode isoforms of this enzyme. Pantothenate kinase isoform PanK3 is highly related to the previously characterized PanK1beta isoform (79% identical, 91% similar), and these two almost identical proteins are expressed most highly in the same tissues. PanK1beta and PanK3 had very similar molecular sizes, oligomeric form, cytoplasmic cellular location, and kinetic constants for ATP and pantothenate. However, these two PanK isoforms possessed distinct regulatory properties. PanK3 was significantly more sensitive to feedback regulation by acetyl-CoA (IC50 = 1 microm) than PanK1beta (IC50 = 10 microm), and PanK3 was stringently regulated by long-chain acyl-CoA (IC50 = 2 microm), whereas PanK1beta was not. Domain swapping experiments localized the difference in the two proteins to a 48-amino-acid domain, where they are the most divergent. Consistent with these more stringent regulatory properties, metabolic labeling experiments showed that coenzyme A (CoA) levels in cells overexpressing PanK3 were lower than in cells overexpressing an equivalent amount of PanK1beta. Thus, the distinct regulatory properties exhibited by the family of the pantothenate kinases allowed the rate of CoA biosynthesis to be controlled by regulatory signals from CoA thioesters involved in different branches of intermediary metabolism.  相似文献   

11.
A new coenzyme A (CoA)-transferase from the anaerobe Clostridium aminobutyricum catalyzing the formation of 4-hydroxybutyryl-CoA from 4-hydroxybutyrate and acetyl-CoA is described. The enzyme was purified to homogeneity by standard techniques, including fast protein liquid chromatography under aerobic conditions. Its molecular mass was determined to be 110 kDa, and that of the only subunit was determined to be 54 kDa, indicating a homodimeric structure. Besides acetate and acetyl-CoA, the following substrates were detected (in order of decreasing kcat/Km): 4-hydroxybutyryl-CoA, butyryl-CoA and propionyl-CoA, vinyl-acetyl-CoA (3-butenoyl-CoA), and 5-hydroxyvaleryl-CoA. In an indirect assay the corresponding acids were also found to be substrates; however, DL-lactate, DL-2-hydroxybutyrate, DL-3-hydroxybutyrate, crotonate, and various dicarboxylates were not.  相似文献   

12.
A new coenzyme A (CoA)-transferase from the anaerobe Clostridium aminobutyricum catalyzing the formation of 4-hydroxybutyryl-CoA from 4-hydroxybutyrate and acetyl-CoA is described. The enzyme was purified to homogeneity by standard techniques, including fast protein liquid chromatography under aerobic conditions. Its molecular mass was determined to be 110 kDa, and that of the only subunit was determined to be 54 kDa, indicating a homodimeric structure. Besides acetate and acetyl-CoA, the following substrates were detected (in order of decreasing kcat/Km): 4-hydroxybutyryl-CoA, butyryl-CoA and propionyl-CoA, vinyl-acetyl-CoA (3-butenoyl-CoA), and 5-hydroxyvaleryl-CoA. In an indirect assay the corresponding acids were also found to be substrates; however, DL-lactate, DL-2-hydroxybutyrate, DL-3-hydroxybutyrate, crotonate, and various dicarboxylates were not.  相似文献   

13.
Intracellular levels of three coenzyme A (CoA) molecular species, i.e., nonesterified CoA (CoASH), acetyl-CoA, and malonyl-CoA, in a variety of aerobic and facultatively anaerobic bacteria were analyzed by the acyl-CoA cycling method developed by us. It was demonstrated that there was an intrinsic difference between aerobes and facultative anaerobes in the changes in the size and composition of CoA pools. The CoA pools in the aerobic bacteria hardly changed and were significantly smaller than those of the facultatively anaerobic bacteria. On the other hand, in the facultatively anaerobic bacteria, the size and composition of the CoA pool drastically changed within minutes in response to the carbon and energy source provided. Acetyl-CoA was the major component of the CoA pool in the facultative anaerobes grown on sufficient glucose, although CoASH was dominant in the aerobes. Therefore, the acetyl-CoA/CoASH ratios in facultatively anaerobic bacteria were 10 times higher than those in aerobic bacteria. In Escherichia coli K-12 cells, the addition of reagents to inhibit the respiratory system led to a rapid decrease in the amount of acetyl-CoA with a concomitant increase in the amount of CoASH, whereas the addition of cerulenin, a specific inhibitor of fatty acid synthase, triggered the intracellular accumulation of malonyl-CoA. The acylation and deacylation of the three CoA molecular species coordinated with the energy-yielding systems and the restriction of the fatty acid-synthesizing system of cells. These data suggest that neither the accumulation of acetyl-CoA nor that of malonyl-CoA exerts negative feedback on pyruvate dehydrogenase and acetyl-CoA carboxylase, respectively.  相似文献   

14.
A Basu  S Subramanian  C SivaRaman 《Biochemistry》1982,21(18):4434-4437
p-Azidobenzoyl coenzyme A functions as a linear competitive inhibitor for (3S)-citryl-CoA in the citryl-CoA oxaloacetate-lyase reaction catalyzed by the Klebsiella aerogenes deacetylcitrate lyase complex (Ki = 80 microM; (3S)-citryl-CoA Km = 67 microM). Inactivation is irreversible on photolysis of p-azidobenzoyl-CoA in the presence of the deacetylcitrate lyase complex. Mg2+ is not required for the inactivation. Inactivation is blocked by (3S)-citryl-CoA in the presence of ethylenediaminetetraacetic acid. p-Azidobenzoyl-CoA has no effect on the acetyl-CoA:citrate CoA transferase activity of both the deacetylcitrate lyase complex and its isolated transferase subunit. The stoichiometry of the CoA ester binding has been investigated by the use of p-azido[14C]benzoyl-CoA as a photoaffinity reagent. The labeling is exclusively on the lyase beta subunit of the citrate lyase complex.  相似文献   

15.
Escherichia coli mutants [coaA16(Fr); Fr indicates feedback resistance] were isolated which possessed a pantothenate kinase activity that was refractory to feedback inhibition by coenzyme A (CoA). Strains harboring this mutation had CoA levels that were significantly elevated compared with strains containing the wild-type kinase and also overproduced both intra- and extracellular 4'-phosphopantetheine. The origin of 4'-phosphopantetheine was investigated by using strain SJ135 [panD delta(aroP-aceEF)], in which synthesis of acetyl-CoA was dependent on the addition of an acetate growth supplement. Rapid degradation of CoA to 4'-phosphopantetheine was triggered by the conversion of acetyl-CoA to CoA following the removal of acetate from the media. CoA hydrolysis under these conditions appeared not to involve acyl carrier protein prosthetic group turnover since [acyl carrier protein] phosphodiesterase was inhibited equally well by acetyl-CoA or CoA. These data support the view that the total cellular CoA content is controlled by modulation of biosynthesis at the pantothenate kinase step and by degradation of CoA to 4'-phosphopantetheine.  相似文献   

16.
The phototrophic bacterium Chloroflexus aurantiacus uses the 3-hydroxypropionate cycle for autotrophic CO(2) fixation. This cycle starts with acetyl-coenzyme A (CoA) and produces glyoxylate. Glyoxylate is an unconventional cell carbon precursor that needs special enzymes for assimilation. Glyoxylate is combined with propionyl-CoA to beta-methylmalyl-CoA, which is converted to citramalate. Cell extracts catalyzed the succinyl-CoA-dependent conversion of citramalate to acetyl-CoA and pyruvate, the central cell carbon precursor. This reaction is due to the combined action of enzymes that were upregulated during autotrophic growth, a coenzyme A transferase with the use of succinyl-CoA as the CoA donor and a lyase cleaving citramalyl-CoA to acetyl-CoA and pyruvate. Genomic analysis identified a gene coding for a putative coenzyme A transferase. The gene was heterologously expressed in Escherichia coli and shown to code for succinyl-CoA:d-citramalate coenzyme A transferase. This enzyme, which catalyzes the reaction d-citramalate + succinyl-CoA --> d-citramalyl-CoA + succinate, was purified and studied. It belongs to class III of the coenzyme A transferase enzyme family, with an aspartate residue in the active site. The homodimeric enzyme composed of 44-kDa subunits was specific for succinyl-CoA as a CoA donor but also accepted d-malate and itaconate instead of d-citramalate. The CoA transferase gene is part of a cluster of genes which are cotranscribed, including the gene for d-citramalyl-CoA lyase. It is proposed that the CoA transferase and the lyase catalyze the last two steps in the glyoxylate assimilation route.  相似文献   

17.
We describe in vitro selection of two novel ribozymes that mediate coenzyme reactions. The first is a trans-capping ribozyme that attaches coenzyme A (CoA) at the 5' end of any RNA with the proper short terminal sequence, including RNAs with randomized internal sequences. From such a trans-capped CoA-RNA pool, we derive ribozymes that attack biotinyl-AMP using the SH group of CoA. These ribozymes, selected to acylate CoA with the valeryl side chain of biotin, also produce the crucial metabolic intermediates acetyl-CoA and butyryl-CoA with substantial velocities. Thus, we argue that RNAs might have used the chemical functionality offered by coenzymes to support an RNA world metabolism. In particular, we can combine our results with those of other labs to argue that simple chemistry and RNA catalysis suffice to proceed from simple chemicals to catalysis with acyl-CoAs. The trans-capping method can be generalized for production of varied coenzyme ribozymes using a single catalytic RNA subunit. Finally, the long-suggested RNA origin for CoA itself appears to be chemically feasible.  相似文献   

18.
Regulation of pantothenate kinase by coenzyme A and its thioesters   总被引:17,自引:0,他引:17  
Pantothenate kinase catalyzes the rate-controlling step in the coenzyme A (CoA) biosynthetic pathway, and its activity is modulated by the size of the CoA pool. The effect of nonesterified CoA (CoASH) and CoA thioesters on the activity of pantothenate kinase was examined to determine which component of the CoA pool is the most effective regulator of the enzyme from Escherichia coli. CoASH was five times more potent than acetyl-CoA or other CoA thioesters as an inhibitor of pantothenate kinase activity in vitro. Inhibition by CoA thioesters was not due to their hydrolysis to CoASH. CoASH inhibition was competitive with respect to ATP, thus providing a mechanism to coordinate CoA production with the energy state of the cell. There were considerable differences in the size and composition of the CoA pool in cells grown on different carbon sources, and a carbon source shift experiment was used to test the inhibitory effect of the different CoA species in vivo. A shift from glucose to acetate as the carbon source resulted in an increase in the CoASH:acetyl-CoA ratio from 0.7 to 4.3. The alteration in the CoA pool composition was associated with the selective inhibition of pantothenate phosphorylation, consistent with CoASH being a more potent regulator of pantothenate kinase activity in vivo. These results demonstrate that CoA biosynthesis is regulated through feedback inhibition of pantothenate kinase primarily by the concentration of CoASH and secondarily by the size of the CoA thioester pool.  相似文献   

19.
The steady state mitochondrial content of coenzyme A-SH (CoA), acetyl-CoA, succinyl-CoA, and long chain acyl-CoA has been determined during the oxidation of palmitoylcarnitine by rabbit heart mitochondria. Variation of the substrate concentration during ADP-stimulated (state 3) respiration varies the mitochondrial content of long chain acyl-CoA and the rate of O2 uptake, and permits the conclusion that the Km of beta oxidation for intramitochondrial long chain acyl-CoA is approximately 1 nmol/mg of mitochondrial protein. At near saturating concentrations of palmitoylcarnitine, plus L-malate, the addition of ADP causes a decrease in acetyl-CoA, an increase in CoA and succinyl-CoA, and no clear change in long chain acyl-CoA content. These changes reverse upon the depletion of ADP (state 3 leads to 4 transition). Similar changes in CoA, acetyl-CoA, and succinyl-CoA are seen during state 4 leads to 3 leads to 4 transitions with pyruvate plus L-malate and octanoate plus L-malate as substrates. These results suggest a limitation of flux by citrate synthase during the controlled oxidation of these three substrates. The ratio acetyl-CoA/succinyl-CoA was determined not only during state 3 and state 4 oxidation of palmitoylcarnitine plus L-malate and pyruvate plus L-malate, but also during intermediate respiratory states (state 3 1/2) generated by adding glucose and varying amounts of hexokinase. These intermediate states are characterized by a high succinyl-CoA content, relative to either state 3 or state 4, and a suboptimal flux through citrate synthase, estimated either by pyruvate disappearance or by O2 uptake.  相似文献   

20.
Dormant spores of Bacillus megaterium were found to contain approximately 850 pmol of coenzyme A (CoA) per milligram of dry weight. Of this total, less than 1.5% was acetyl-CoA, 25% was CoA-disulfide, 43% was in disulfide linkage to protein, and the remainder was the free thiol. Dormand spores of Bacillus cereus and Clostridium bifermentans contained 700 and 600 pmol of CoA per milligram of dry weight, respectively; in both species approximately 45% of the CoA 45% of the CoA was in disulfide linkage to protein. During germination of spores of all three species, greater than 75% of the CoA-protein disulfides were cleaved. In B. megaterium, cleavage of these disulfides during spore germination did not require exogenous metabolites and occurred at about the same time as the initiation of germination. Much of the CoA was converted to acetyl-CoA at this time. Dormant spores also contained reduced nicotinamide adenine dinucleotide-dependent CoA-disulfide reductase at levels higher than those in other stages of growth. The level of total CoA in the growing cells was two- to three-fold higher than in spores. This level remained constant throughout growth and sporulation, but less than 2% of the total cellular CoA was in disulfide linkage to protein until late in sporulation. The CoA-protein disulfides accumulated exclusively within the developing spore at about the time when dipicolinic acid was accumulated.  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