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.
The ATP-dependent 6-phosphofructokinase (ATP-PFK) of the hyperthermophilic bacterium Thermotoga maritimawas purified 730-fold to homogeneity. The enzyme is a 140-kDa homotetramer composed of 34 kDa subunits. Kinetic constants were determined for all substrates in both reaction directions at pH 7 and at 75 degrees C. Rate dependence (forward reaction) on fructose 6-phosphate (F-6-P) showed sigmoidal kinetics with a half-maximal saturation constant ( S(0.5)) of 0.7 mM and a Hill coefficient of 2.2. The apparent K(m) for ATP was 0.2 mM and the apparent V(max) value was about 360 U/mg. The enzyme also catalyzed in vitro the reverse reaction with an apparent K(m) for fructose 1,6-bisphosphate and ADP of 7.6 mM and 1.4 mM, respectively, and an apparent V(max) of about 13 U/mg. Divalent cations were required for maximal activity; Mg(2+), which was most effective, could partially be replaced by Mn(2+) and Fe(2+). Enzyme activity was allosterically regulated by classical effectors of ATP-PFKs of Eukarya and Bacteria; it was activated by ADP and inhibited by PEP. The enzyme had a temperature optimum of 93 degrees C and showed a significant thermostability up to 100 degrees C. Using the N-terminal amino acid sequence of the subunit, the pfk gene coding for ATP-PFK was identified and functionally overexpressed in Escherichia coli. The purified recombinant ATP-PFK had identical kinetic and allosteric properties as the native enzyme purified from T. maritima. The deduced amino acid sequence showed high sequence similarity to members of the PFK-A family. In accordance with its allosteric properties, ATP-PFK of T. maritima contained the conserved allosteric effector-binding sites for ADP and PEP.  相似文献   

2.
The actinomycete Amycolatopsis methanolica was found to employ the normal bacterial set of glycolytic and pentose phosphate pathway enzymes, except for the presence of a PPi-dependent phosphofructokinase (PPi-PFK) and a 3-phosphoglycerate mutase that is stimulated by 2,3-bisphosphoglycerate. Screening of a number of actinomycetes revealed PPi-PFK activity only in members of the family Pseudonocardiaceae. The A. methanolica PPi-PFK and 3-phosphoglycerate mutase enzymes were purified to homogeneity. PPi-PFK appeared to be insensitive to the typical effectors of ATP-dependent PFK enzymes. Nevertheless, strong N-terminal amino acid sequence homology was found with ATP-PFK enzymes from other bacteria. The A. methanolica pyruvate kinase was purified over 250-fold and characterized as an allosteric enzyme, sensitive to inhibition by P(i) and ATP but stimulated by AMP. By using mutants, evidence was obtained for the presence of transketolase isoenzymes functioning in the pentose phosphate pathway and ribulose monophosphate cycle during growth on glucose and methanol, respectively.  相似文献   

3.
Cells of the actinomycete Amycolatopsis methanolica grown on glucose possess only a single, exclusively PP(i)-dependent phosphofructokinase (PP(i)-PFK) (A. M. C. R. Alves, G. J. W. Euverink, H. J. Hektor, J. van der Vlag, W. Vrijbloed, D.H.A. Hondmann, J. Visser, and L. Dijkhuizen, J. Bacteriol. 176:6827-6835, 1994). When this methylotrophic bacterium is grown on one-carbon (C(1)) compounds (e.g., methanol), an ATP-dependent phosphofructokinase (ATP-PFK) activity is specifically induced, completely replacing the PP(i)-PFK. The two A. methanolica PFK isoenzymes have very distinct functions, namely, in the metabolism of C(6) and C(1) carbon substrates. This is the first report providing biochemical evidence for the presence and physiological roles of PP(i)-PFK and ATP-PFK isoenzymes in a bacterium. The novel ATP-PFK enzyme was purified to homogeneity and characterized in detail at the biochemical and molecular levels. The A. methanolica ATP-PFK and PP(i)-PFK proteins possess a low level of amino acid sequence similarity (24%), clearly showing that the two proteins are not the result of a gene duplication event. PP(i)-PFK is closely related to other (putative) actinomycete PFK enzymes. Surprisingly, the A. methanolica ATP-PFK is most similar to ATP-PFK from the protozoon Trypanosoma brucei and PP(i)-PFK proteins from the bacteria Borrelia burgdorferi and Treponema pallidum, both spirochetes, very distinct from actinomycetes. The data thus suggest that A. methanolica obtained the ATP-PFK-encoding gene via a lateral gene transfer event.  相似文献   

4.
An extracellular lipase from Streptomyces rimosus R6-554W has been recently purified and biochemically characterized. In this report the cloning, sequencing, and high-level expression of its gene is described. The cloned DNA contained an ORF of 804 bp encoding a 268-amino-acid polypeptide with 34 amino acid residues at the amino terminus of the sequence that were not found in the mature protein. The theoretical molecular mass (24.172 kDa) deduced from the amino acid sequence of the mature enzyme was experimentally confirmed. This lipase showed no overall amino acid sequence similarity to other lipases in the databases. However, two hypothetical proteins, i. e. putative hydrolases, derived from the genome sequencing data of Streptomyces coelicolor A3(2), showed 66% and 33% identity. In addition, a significant similarity to esterases from Streptomyces diastatochromogenes and Aspergillus terreus was found. Sequence analysis revealed that our novel S. rimosus lipase containing a GDS(L)-like consensus motif belongs to family II of lipolytic enzymes, previously unrecognized in Streptomyces. When the lipase gene was expressed in a S. rimosus lipase-deficient strain harboring the lipase gene on a high-copy-number vector, lipase activity was 22-fold higher than in the original strain.  相似文献   

5.
6.
Some strains of Bacillus sphaericus are entomopathogenic to mosquito larvae, which transmit diseases, such as filariasis and malaria, affecting millions of people worldwide. This species is unable to use hexoses and pentoses as unique carbon sources, which was proposed to be due to the lack of glycolytic enzymes, such as 6-phosphofructokinase (PFK). In this study, PFK activity was detected and the pfk gene was cloned and sequenced. Furthermore, this gene was shown to be present in strains belonging to all the homology groups of this heterogeneous species, in which PFK activity was also detected. A careful sequence analysis revealed the conservation of different catalytic and regulatory residues, as well as the enzyme's phylogenetic affiliation with the family of allosteric ATP-PFK enzymes.  相似文献   

7.
Yeast NAD(+)-specific isocitrate dehydrogenase (IDH) is an octameric enzyme composed of four heterodimers of regulatory IDH1 and catalytic IDH2 subunits. The crystal structure suggested that the interactions between tetramers in the octamer are restricted to defined regions in IDH1 subunits from each tetramer. Using truncation and mutagenesis, we constructed three tetrameric forms of IDH. Truncation of five residues from the amino terminus of IDH1 did not alter the octameric form of the enzyme, but this truncation with an IDH1 G15D or IDH1 D168K residue substitution produced tetrameric enzymes as assessed by sedimentation velocity ultracentrifugation. The IDH1 G15D substitution in the absence of any truncation of IDH1 was subsequently found to be sufficient for production of a tetrameric enzyme. The tetrameric forms of IDH exhibited ~50% reductions in V(max) and in cooperativity with respect to isocitrate relative to those of the wild-type enzyme, but they retained the property of allosteric activation by AMP. The truncated (-5)IDH1/IDH2 and tetrameric enzymes were much more sensitive than the wild-type enzyme to inhibition by the oxidant diamide and concomitant formation of a disulfide bond between IDH2 Cys-150 residues. Binding of ligands reduced the sensitivity of the wild-type enzyme to diamide but had no effect on inhibition of the truncated or tetrameric enzymes. These results suggest that the octameric structure of IDH has in part evolved for regulation of disulfide bond formation and activity by ensuring the proximity of the amino terminus of an IDH1 subunit of one tetramer to the IDH2 Cys-150 residues in the other tetramer.  相似文献   

8.
We have achieved the purification of an alpha-keto ester reductase (SCKER) from S. coelicolor A3(2) whole cells. SCKER proved to be a homotetramer of 132 kDa containing one equivalent of zinc ion per subunit. The enzyme differed from other alpha-keto ester reductases from microorganisms with regard to subunit structure and metal ion dependency. From a computer search using the protein data banks, the N-terminal amino acid sequence of SCKER was consistent with that of a possible zinc containing alcohol dehydrogenase in S. coelicolor A3(2). None of three hypothetical proteins of S. coelocor A3(2) having a high homology sequence with those of already purified alpha-keto ester reductases from S. thermocyaneoviolaceus [Yamaguchi, H., et al., Biosci. Biotechnol. Biochem., 66, 588-597 (2002)] was identical with that of SCKER.  相似文献   

9.
We performed a detailed bioinformatic study of the catalytic step of fructose-6-phosphate phosphorylation in glycolysis based on the raw genomic draft of Propionibacterium freudenreichii subsp. shermanii (P. shermanii) ATCC9614 [Meurice et al., 2004]. Our results provide the first in silico evidence of the coexistence of genes coding for an ATP-dependent phosphofructokinase (ATP-PFK) and a PPi-dependent phosphofructokinase (PPi-PFK), whereas the fructose-1,6-bisphosphatase (FBP) and ADP-dependent phosphofructokinase (ADP-PFK) are absent. The deduced amino acid sequence corresponding to the PPi-PFK (AJ508922) shares 100% similarity with the already characterised propionibacterial protein (P29495; Ladror et al., 1991]. The unexpected ATP-PFK gene (AJ509827) encodes a protein of 373 aa which is highly similar (50% positive residues) along at least 95% of its sequence length to different well-characterised ATP-PFKs. The characteristic PROSITE pattern important for the enzyme function of ATP-PFKs (PS00433) was conserved in the putative ATP-PFK sequence: 8 out of 9 amino acid residues. According to the recent evolutionary study of PFK proteins with different phosphate donors [Bapteste et al., 2003], the propionibacterial ATP-PFK harbours a G104-K124 residue combination, which strongly suggested that this enzyme belongs to the group of atypical ATP-PFKs. According to our phylogenetic analyses the amino acid sequence of the ATP-PFK is clustered with the atypical ATP-PFKs from group III of the Siebers classification [Siebers et al., 1998], whereas the expected PPi-PFK protein is closer to the PPi-PFKs from clade P [Müller et al., 2001]. The possible significance of the co-existence of these two PFKs and their importance for the regulation of glycolytic pathway flux in P. shermanii is discussed.  相似文献   

10.
11.
The greatly increased tetramer strength of liganded fetal hemoglobin compared with adult hemoglobin is shown by its 70-fold smaller tetramer-dimer dissociation constant. This property has been shown previously to be only partially caused by the 5-amino-acid differences at both types of interfaces in each hemoglobin. A major contributor to tetramer strengthening is the 18-amino-acid N-terminal A helix of the gamma-subunit of fetal hemoglobin, which differs from the beta-subunit of adult hemoglobin at eight amino acid residues. This long-distance communication between the A helix and the distant C helix and FG helical corner comprising the subunit contacts at the allosteric interface represents internal signaling. Physiologically, its greater tetramer strength endows fetal hemoglobin with the capacity to abstract oxygen from maternal adult hemoglobin. It also leads to resistance of fetal red cells to the malaria parasite because the HbF tetramer does not dissociate to dimers as readily as HbA; dimers are digested by malaria proteases but tetramers are not. In this communication, we report which sites on the A helix of the gamma-subunit are important for tetramer strengthening in HbF by substituting certain amino acids in the beta-subunit by the corresponding residues in the gamma-subunit. The recombinant hemoglobins containing up to five replacements together have been extensively characterized. Mass values were within 1 unit of theory. Gly 1 (gamma) of HbF with its high pK(a) of 8.1 compared with a 7.1 value for Val 1 (beta) of HbA creates a highly electropositive N terminus that may couple with the electronegative sequence just after it on the gamma-subunit. The Leu 3 to Phe replacement has no apparent role; however, position 5 is important because replacement of Pro 5 (beta) by Glu 5 (gamma) promotes tetramer strengthening. The Glu --> Asp replacement at position 7 enhances this effect because of the lower pK(a) of Asp but the Val --> Ile substitution at position 11 has no effect. Thus, the three positive/negative sites at positions 1, 5, and 7 account for practically all of the tetramer strength of HbF, as illustrated by an electrostatic surface potential analysis. The pathway by which information is transmitted to the distant allosteric subunit interfaces is currently under study. Oxygen-binding properties of the hemoglobins with charged substitutions more closely resemble those of HbA rather than those of HbF. Thus, whereas the A helix has a major role in controlling the strength of interactions at the tetramer-dimer allosteric interface, oxygen-binding properties of HbA and HbF are influenced by sequences in the C helix and at the FG helical corner constituting the allosteric interface.  相似文献   

12.
The sco6545 gene of Streptomyces coelicolor A3(2) was nominated as a putative cellulase with 863 mature-form amino acids (90.58 kDa). We overexpressed and purified Sco6545 and demonstrated that the protein is not a cellulase but a xyloglucan-specific glycosyl hydrolase which cleaves xyloglucan at unbranched glucose residues.  相似文献   

13.
The native Escherichia coli aspartate transcarbamoylase (ATCase, E.C. 2.1.3.2) provides a classic allosteric model for the feedback inhibition of a biosynthetic pathway by its end products. Both E. coli and Erwinia herbicola possess ATCase holoenzymes which are dodecameric (2(c3):3(r2)) with 311 amino acid residues per catalytic monomer and 153 and 154 amino acid residues per regulatory (r) monomer, respectively. While the quaternary structures of the two enzymes are identical, the primary amino acid sequences have diverged by 14 % in the catalytic polypeptide and 20 % in the regulatory polypeptide. The amino acids proposed to be directly involved in the active site and nucleotide binding site are strictly conserved between the two enzymes; nonetheless, the two enzymes differ in their catalytic and regulatory characteristics. The E. coli enzyme has sigmoidal substrate binding with activation by ATP, and inhibition by CTP, while the E. herbicola enzyme has apparent first order kinetics at low substrate concentrations in the absence of allosteric ligands, no ATP activation and only slight CTP inhibition. In an apparently important and highly conserved characteristic, CTP and UTP impose strong synergistic inhibition on both enzymes. The co-operative binding of aspartate in the E. coli enzyme is correlated with a T-to-R conformational transition which appears to be greatly reduced in the E. herbicola enzyme, although the addition of inhibitory heterotropic ligands (CTP or CTP+UTP) re-establishes co-operative saturation kinetics. Hybrid holoenzymes assembled in vivo with catalytic subunits from E. herbicola and regulatory subunits from E. coli mimick the allosteric response of the native E. coli holoenzyme and exhibit ATP activation. The reverse hybrid, regulatory subunits from E. herbicola and catalytic subunits from E. coli, exhibited no response to ATP. The conserved structure and diverged functional characteristics of the E. herbicola enzyme provides an opportunity for a new evaluation of the common paradigm involving allosteric control of ATCase.  相似文献   

14.
A pyrophosphate-dependent phosphofructokinase (PPi-PFK) and an ATP-dependent phosphofructokinase (ATP-PFK) from Thermotoga maritima have been cloned and characterized. The PPi-PFK is unique in that the Km and Vmax values indicate that polyphosphate is the preferred substrate over pyrophosphate; the enzyme in reality is a polyphosphate-dependent PFK. The ATP-PFK was not significantly affected by common allosteric effectors (e.g., phosphoenolpyruvate) but was strongly inhibited by PPi and polyphosphate. The results suggest that the control of the Embden-Meyerhof pathway in this organism is likely to be modulated by pyrophosphate and/or polyphosphate.  相似文献   

15.
The polypeptide chain of the allosteric L-lactate dehydrogenase (EC 1.1.1.27) of Lactobacillus casei consists of 325 amino acid residues. Despite the strikingly different enzymatic characteristics of the allosteric L-lactate dehydrogenase of L. casei and of the non-allosteric vertebrate enzymes, the sequence of the allosteric enzyme shows a distinct homology with that of the non-allosteric vertebrate enzymes (average identity: 37%). An especially high sequence homology can be identified within the active center (average identity: 70%). A clear deviation of the L. casei enzyme from the vertebrate enzyme is the lack of the first 12 amino acid residues at the N terminus and an additional 7 amino acid residues at the C terminus. The localization of the binding site of the allosteric effector D-fructose 1,6-bisphosphate and pH and effector-induced changes of the spectroscopic properties are discussed on the basis of the primary structure.  相似文献   

16.
A chitinase gene (pCHi58) encoding a 58 kDa chitinase was isolated from theSerratia marcescens KCTC 2172 cosmid library. The chitinase gene consisted of a 1686 bp open reading frame that encoded 562 amino acids.Escherichia coil harboring the pChi58 gene secreted a 58 kDa chitinase into the culture supernatant. The 58 kDa chitinase was purified using a chitin affinity column and mono-S column. A nucleotide andN-terminal amino acid sequence analysis showed that the 58 kDa chitinase had a leader peptide consisting of 23 amino acids which was cleaved prior to the 24th alanine. The 58 KDa chitinase exhibited a 98% similarity to that ofS. marcescens QMB 1466 in its nuclotide sequence. The chitinolytic patterns of the 58 kDa chitinase released N,N′-diacetyl chitobiose (NAG2) as the major hydrolysis end-product with a trace amount ofN-acetylglucosamine. When a 4-methylumbellyferyl-N-acetylglucosamin monomer, dimmer, and tetramer were used as substrates, the 58 kDa chitinase did not digest the 4-Mu-NAG monomer (analogue of NAG2), thereby indicating that the 58 kDa chitinase was likely an endochitinase. The optimum reaction temperature and pH of the enzyme were 50°C and 5.0, respectively.  相似文献   

17.
The structural gene for glutamine synthetase, glnA, from Amycolatopsis mediterranei U32 was cloned via screening a genomic library using the analog gene from Streptomyces coelicolor. The clone was functionally verified by complementing for glutamine requirement of an Escherichia coli glnA null mutant under the control of a lac promoter. Sequence analysis showed an open reading frame encoding a protein of 466 amino acid residues. The deduced amino acid sequence bears significant homologies to other bacterial type I glutamine synthetases, specifically, 71% and 72% identical to the enzymes of S. coelicolor and Mycobacterium tuberculosis, respectively. Disruption of this glnA gene in A. mediterranei U32 led to glutamine auxotrophy with no detectable glutamine synthetase activity in vivo. In contrast, the cloned glnA^+ gene can complement for both phenotypes in trans. It thus suggested that in A. mediterranei U32, the glnA gene encoding glutamine synthetase is uniquely responsible for in vivo glutamine synthesis under our laboratory defined physiological conditions.  相似文献   

18.
The aspartate transcarbamoylases (ATCase, EC 2.1.3.2) of Escherichia coli and Serratia marcescens have similar dodecameric enzyme structures (2(c3):3(r2] but differ in both regulatory and catalytic characteristics. The catalytic cistrons (pyrB) of the ATCases from E. coli and S. marcescens encode polypeptides of 311 and 306 amino acids, respectively; there is a 76% identity between the DNA sequences and an overall amino acid homology of 88% (38 differences). The regulatory cistrons (pyrI) of these ATCases encode polypeptides of 153 and 154 amino acids, respectively, and there is a 75% identity between the DNA sequences and an overall amino acid homology of 77% (36 differences). In both species, the two genes are arranged as a bicistronic operon, with pyrB promoter proximal. A comparison of the deduced amino acid sequences reveals that the active site and the allosteric binding sites, as well as most of the intrasubunit interactions and intersubunit associations, are conserved in the E. coli and the S. marcescens enzymes; however, there are specific differences which undoubtedly contribute to the catalytic and regulatory differences between the enzymes of the two species. These differences include residues that have been implicated in the T-R transition, c1:r1 interface interactions, and the CTP binding site. A hybrid ATCase assembled in vivo with catalytic subunits from E. coli and regulatory subunits from S. marcescens has a 6 mM requirement for aspartate at half-maximal saturation, similar to the 5.5 mM aspartate requirement of the native E. coli holoenzyme at half-maximal saturation. However, the heterotropic response of this hybrid enzyme is characteristic of the heterotropic response of the native S. marcescens holoenzyme: ATP activation and CTP activation. Activation by both allosteric effectors indicates that the heterotropic response of this hybrid holoenzyme (Cec:Rsm) is determined by the associated S. marcescens regulatory subunits.  相似文献   

19.
The quaternary protein structure of two methanol:N,N'-dimethyl-4-nitrosoaniline (NDMA) oxidoreductases purified from Amycolatopsis methanolica and Mycobacterium gastri MB19 was analyzed by electron microscopy and image processing. The enzymes are decameric proteins (displaying fivefold symmetry) with estimated molecular masses of 490 to 500 kDa based on their subunit molecular masses of 49 to 50 kDa. Both methanol:NDMA oxidoreductases possess a tightly but noncovalently bound NADP(H) cofactor at an NADPH-to-subunit molar ratio of 0.7. These cofactors are redox active toward alcohol and aldehyde substrates. Both enzymes contain significant amounts of Zn2+ and Mg2+ ions. The primary amino acid sequences of the A. methanolica and M. gastri MB19 methanol:NDMA oxidoreductases share a high degree of identity, as indicated by N-terminal sequence analysis (63% identity among the first 27 N-terminal amino acids), internal peptide sequence analysis, and overall amino acid composition. The amino acid sequence analysis also revealed significant similarity to a decameric methanol dehydrogenase of Bacillus methanolicus C1.  相似文献   

20.
During growth with maltotriose or amylose, Streptomyces lividans and Streptomyces coelicolor A3(2) synthesize a maltodextrin uptake system with highest specificity for maltotriose. The transport activity is absent in mutants of S. coelicolor A3(2) lacking a functional MalE binding protein. Cloning and sequencing data suggest that the mal operon of S. coelicolor A3(2) corresponds to the one of S. lividans and that the deduced S. lividans Reg1 amino acid sequence is identical to that of MalR from S. coelicolor A3(2). It can be concluded that both strains have the same ABC transport system for maltodextrins. The S. lividans malR was cloned in Escherichia coli in frame with six histidine-encoding codons. The resulting, purified 6HisMalR(SI) was shown to bind to two motifs within the S. lividans malR-malE intergenic region and to dissociate in the presence of maltopentaose.  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