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.
The meso-2,3-butanediol dehydrogenase (meso-BDH) from S. marcescens H30 is responsible for converting acetoin into 2,3-butanediol during sugar fermentation. Inactivation of the meso-BDH encoded by budC gene does not completely abolish 2,3-butanediol production, which suggests that another similar enzyme involved in 2,3-butanediol formation exists in S. marcescens H30. In the present study, a glycerol dehydrogenase (GDH) encoded by gldA gene from S. marcescens H30 was expressed in Escherichia coli BL21(DE3), purified and characterized for its properties. In vitro conversion indicated that the purified GDH could catalyze the interconversion of (3S)-acetoin/meso-2,3-butanediol and (3R)-acetoin/(2R,3R)-2,3-butanediol. (2S,3S)-2,3-Butanediol was not a substrate for the GDH at all. Kinetic parameters of the GDH enzyme showed lower K m value and higher catalytic efficiency for (3S/3R)-acetoin in comparison to those for (2R,3R)-2,3-butanediol and meso-2,3-butanediol, implying its physiological role in favor of 2,3-butanediol formation. Maximum activity for reduction of (3S/3R)-acetoin and oxidations of meso-2,3-butanediol and glycerol was observed at pH 8.0, while it was pH 7.0 for diacetyl reduction. The enzyme exhibited relative high thermotolerance with optimum temperature of 60 °C in the oxidation–reduction reactions. Over 60 % of maximum activity was retained at 70 °C. Additionally, the GDH activity was significantly enhanced for meso-2,3-BD oxidation in the presence of Fe2+ and for (3S/3R)-acetoin reduction in the presence of Mn2+, while several cations inhibited its activity, particularly Fe2+ and Fe3+ for (3S/3R)-acetoin reduction. The properties provided potential application for single configuration production of acetoin and 2,3-butanediol .  相似文献   

2.
Using a conserved sequence motif, a new gene (YAL060W) of the MDR family has been identified in Saccharomyces cerevisiae. The expressed protein was a stereoespecific (2R,3R)-2,3-butanediol dehydrogenase (BDH). The best substrates were (2R,3R)-2,3-butanediol for the oxidation and (3R/3S)-acetoin and 1-hydroxy-2-propanone for the reduction reactions. The enzyme is extremely specific for NAD(H) as cofactor, probably because the presence of Glu223 in the cofactor binding site, instead of the highly conserved Asp223. BDH is inhibited competitively by 4-methylpyrazole with a Ki of 34 μM. Yeast could grow on 2,3-butanediol or acetoin as a sole energy and carbon sources, and a 3.6-fold increase in BDH activity was observed when cells were grown in 2,3-butanediol, suggesting a role of the enzyme in 2,3-butanediol metabolism. However, the disruption of the YAL060W gene was not lethal for the yeast under laboratory conditions, and the disrupted strain could also grow in 2,3-butanediol and acetoin. This suggests that other enzymes, in addition to BDH, can also metabolize 2,3-butanediol in yeast.  相似文献   

3.
The chemical similarity of cellulose and chitin supports the idea that their corresponding hydrolytic enzymes would bind β-1,4-linked glucose residues in a similar manner. A structural and mutational analysis was performed for the plant cellulolytic enzyme BGlu1 from Oryza sativa and the insect chitinolytic enzyme OfHex1 from Ostrinia furnacalis. Although BGlu1 shows little amino-acid sequence or topological similarity with OfHex1, three residues (Trp490, Glu328, Val327 in OfHex1, and Trp358, Tyr131 and Ile179 in BGlu1) were identified as being conserved in the +1 sugar binding site. OfHex1 Glu328 together with Trp490 was confirmed to be necessary for substrate binding. The mutant E328A exhibited a 8-fold increment in K m for (GlcNAc)2 and a 42-fold increment in K i for TMG-chitotriomycin. A crystal structure of E328A in complex with TMG-chitotriomycin was resolved at 2.5 Å, revealing the obvious conformational changes of the catalytic residues (Glu368 and Asp367) and the absence of the hydrogen bond between E328A and the C3-OH of the +1 sugar. V327G exhibited the same activity as the wild-type, but acquired the ability to efficiently hydrolyse β-1,2-linked GlcNAc in contrast to the wild-type. Thus, Glu328 and Val327 were identified as important for substrate-binding and as glycosidic-bond determinants. A structure-based sequence alignment confirmed the spatial conservation of these three residues in most plant cellulolytic, insect and bacterial chitinolytic enzymes.  相似文献   

4.
NAD-dependent butanediol dehydrogenase (Bdh1p) from Saccharomyces cerevisiae reversibly transforms acetoin to 2,3-butanediol in a stereospecific manner. Deletion of BDH1 resulted in an accumulation of acetoin and a diminution of 2,3-butanediol in two S. cerevisiae strains under two different growth conditions. The concentrations of (2R,3R)-2,3-butanediol are mostly dependent on Bdh1p activity, while those of (meso)-2,3-butanediol are also influenced by the activity of NADP(H)-dependent oxidoreductases. One of them has been purified and shown to be d-arabinose dehydrogenase (Ara1p), which converts (R/S)-acetoin to meso-2,3-butanediol and (2S,3S)-2,3-butanediol. Deletion of BDH2, a gene adjacent to BDH1, whose encoded protein is 51% identical to Bdh1p, does not significantly alter the levels of acetoin or 2,3-butanediol in comparison to the wild-type strain. Furthermore, we have expressed Bdh2p with a histidine tag and have shown it to be inactive toward 2,3-butanediol. A whole-genome expression analysis with microarrays demonstrates that BDH1 and BDH2 are reciprocally regulated.Acetoin and 2,3-butanediol are minor products generated by Saccharomyces cerevisiae during alcohol fermentation. Their sensory impacts on wine are poorly documented. Acetoin may affect the wine bouquet, although its perception threshold in wine is relatively high, around 150 mg/liter (21, 31). On the other hand, 2,3-butanediol is odorless (33) and cannot be expected to appreciably affect the sensory quality of wine. However, the compound may contribute to the wine body (28).Acetaldehyde, pyruvate, and α-acetolactate are the main precursors of acetoin in S. cerevisiae. Acetoin can be formed from acetaldehyde and/or pyruvate through an anomalous reaction of pyruvate decarboxylase. Thus, although its main activity is to irreversibly decarboxylate pyruvate to acetaldehyde, it can also catalyze carbon-carbon bond formation, yielding acetoin from pyruvate and/or acetaldehyde (2, 4). In addition, α-acetolactate would produce acetoin through its nonenzymatic decarboxylation to diacetyl and subsequent reduction to acetoin through the action of several NADH- and NADPH-dependent oxidoreductases (12). However, the situation is more complex in wine fermentation, where other yeasts and bacteria display supplementary enzymatic activities capable of producing both acetoin and 2,3-butanediol (1, 27).We have previously characterized a butanediol dehydrogenase (Bdh1p) as a medium-chain dehydrogenase/reductase (MDR) that can reversibly transform R-acetoin and S-acetoin to (2R,3R)-2,3-butanediol and meso-2,3-butanediol, respectively, in a NAD(H)-dependent reaction (10). BDH2 is a gene adjacent to BDH1 whose uncharacterized protein product (Bdh2p) shares 51% sequence identity with Bdh1p. To evaluate the in vivo roles of Bdh1p and Bdh2p, we compared the levels of several extracellular metabolites in cultures of wild-type and deficient strains. The results show that, although Bdh1p is the main enzyme in 2,3-butanediol production [essentially the (2R,3R)-2,3-butanediol stereoisomer], some meso-2,3-butanediol is still produced by the bdh1Δ strains. We have characterized Ara1p as an oxidoreductase that can reduce racemic acetoin to meso-2,3-butanediol and (2S,3S)-2,3-butanediol in the presence of NADPH.Furthermore, we have overexpressed Bdh2p with a histidine tag at its carboxyl terminus and have shown it to be inactive toward acetoin and 2,3-butanediol. A microarray study indicated that BDH1 and BDH2 are reciprocally regulated under the conditions studied.  相似文献   

5.
α-l-Arabinofuranosidase, which belongs to the glycoside hydrolase family 62 (GH62), hydrolyzes arabinoxylan but not arabinan or arabinogalactan. The crystal structures of several α-l-arabinofuranosidases have been determined, although the structures, catalytic mechanisms, and substrate specificities of GH62 enzymes remain unclear. To evaluate the substrate specificity of a GH62 enzyme, we determined the crystal structure of α-l-arabinofuranosidase, which comprises a carbohydrate-binding module family 13 domain at its N terminus and a catalytic domain at its C terminus, from Streptomyces coelicolor. The catalytic domain was a five-bladed β-propeller consisting of five radially oriented anti-parallel β-sheets. Sugar complex structures with l-arabinose, xylotriose, and xylohexaose revealed five subsites in the catalytic cleft and an l-arabinose-binding pocket at the bottom of the cleft. The entire structure of this GH62 family enzyme was very similar to that of glycoside hydrolase 43 family enzymes, and the catalytically important acidic residues found in family 43 enzymes were conserved in GH62. Mutagenesis studies revealed that Asp202 and Glu361 were catalytic residues, and Trp270, Tyr461, and Asn462 were involved in the substrate-binding site for discriminating the substrate structures. In particular, hydrogen bonding between Asn462 and xylose at the nonreducing end subsite +2 was important for the higher activity of substituted arabinofuranosyl residues than that for terminal arabinofuranoses.  相似文献   

6.
7.
To put forward BDH from Pseudomonas aeruginosa’s enzymatic properties, we report a two-step purification of BDH and its gene sequencing allowing the investigation of its structural properties. Purification of BDH was achieved, using ammonium sulfate fractionation and Blue Sepharose CL-6B affinity chromatography. SDS–PAGE analysis reveals a MM of 29 kDa, whereas the native enzyme showed a MM of 120 kDa suggesting a homotetrameric structure. BDH encoding gene sequence shows a nucleotide open reading frame sequence of 771 bp encoding a 265 amino acid residues polypeptide chain. The modeling analysis of the three dimensional structure fits with the importance of amino acids in the catalysis reaction especially a strictly conserved tetrad. Amino-acid residues in interaction with the coenzyme NAD+ were also identified.  相似文献   

8.
9.
Erv1 is an FAD-dependent thiol oxidase of the ERV (essential for respiration and viability)/ALR (augmenter of liver regeneration) sub-family and an essential component of the mitochondrial import and assembly pathway. Erv1 contains six tryptophan residues, which are all located in the highly conserved C-terminal FAD-binding domain. Though important structural roles were predicted for the invariable Trp95, no experimental study has been reported. In the present study, we investigated the structural and functional roles of individual tryptophan residues of Erv1. Six single tryptophan-to-phenylalanine yeast mutant strains were generated and their effects on cell viability were tested at various temperatures. Then, the mutants were purified from Escherichia coli. Their effects on folding, FAD-binding and Erv1 activity were characterized. Our results showed that Erv1W95F has the strongest effect on the stability and function of Erv1 and followed by Erv1W183F. Erv1W95F results in a decrease in the Tm of Erv1 by 23°C, a significant loss of the oxidase activity and thus causing cell growth defects at both 30°C and 37°C. Erv1W183F induces changes in the oligomerization state of Erv1, along with a pronounced effect on the stability of Erv1 and its function at 37°C, whereas the other mutants had no clear effect on the function of Erv1 including the highly conserved Trp157 mutant. Finally, computational analysis indicates that Trp95 plays a key role in stabilizing the isoalloxazine ring to interact with Cys133. Taken together, the present study provided important insights into the molecular mechanism of how thiol oxidases use FAD in catalysing disulfide bond formation.  相似文献   

10.
The 6-phospho-β-glucosidase BglA-2 (EC 3.2.1.86) from glycoside hydrolase family 1 (GH-1) catalyzes the hydrolysis of β-1,4-linked cellobiose 6-phosphate (cellobiose-6′P) to yield glucose and glucose 6-phosphate. Both reaction products are further metabolized by the energy-generating glycolytic pathway. Here, we present the first crystal structures of the apo and complex forms of BglA-2 with thiocellobiose-6′P (a non-metabolizable analog of cellobiose-6′P) at 2.0 and 2.4 Å resolution, respectively. Similar to other GH-1 enzymes, the overall structure of BglA-2 from Streptococcus pneumoniae adopts a typical (β/α)8 TIM-barrel, with the active site located at the center of the convex surface of the β-barrel. Structural analyses, in combination with enzymatic data obtained from site-directed mutant proteins, suggest that three aromatic residues, Tyr126, Tyr303, and Trp338, at subsite +1 of BglA-2 determine substrate specificity with respect to 1,4-linked 6-phospho-β-glucosides. Moreover, three additional residues, Ser424, Lys430, and Tyr432 of BglA-2, were found to play important roles in the hydrolytic selectivity toward phosphorylated rather than non-phosphorylated compounds. Comparative structural analysis suggests that a tryptophan versus a methionine/alanine residue at subsite −1 may contribute to the catalytic and substrate selectivity with respect to structurally similar 6-phospho-β-galactosidases and 6-phospho-β-glucosidases assigned to the GH-1 family.  相似文献   

11.
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   

12.
Enzymic and structural studies on Drosophila alcohol dehydrogenases and other short-chain dehydrogenases/reductases (SDRs) are presented. Like alcohol dehydrogenases from other Drosophila species, the enzyme from D. simulans is more active on secondary than on primary alcohols, although ethanol is its only known physiological substrate. Several secondary alcohols were used to determine the kinetic parameters kcat and Km. The results of these experiments indicate that the substrate-binding region of the enzyme allows optimal binding of a short ethyl side-chain in a small binding pocket, and of a propyl or butyl side-chain in large binding pocket, with stereospecificity for R(−) alcohols. At a high concentration of R(−) alcohols substrate activation occurs. The kcat and Km values determined under these conditions are about two-fold, and two orders of magnitude, respectively, higher than those at low substrate concentrations. Sequence alignment of several SDRs of known, and unknown three-dimensional structures, indicate the presence of several conserved residues in addition to those involved in the catalyzed reactions. Structural roles of these conserved residues could be derived from observations made on superpositioned structures of several SDRs with known structures. Several residues are conserved in tetrameric SDRs, but not in dimeric ones. Two halohydrin-halide-lyases show significant homology with SDRs in the catalytic domains of these enzymes, but they do not have the structural features required for binding NAD+. Probably these lyases descend from an SDR, which has lost the capability to bind NAD+, but the enzyme reaction mechanisms may still be similar. Received: 23 May 2000 / Accepted: 4 January 2001  相似文献   

13.
Two 2,3-butanediol dehydrogenases (enzymes 1 and 2; molecular weight of each, 170,000) have been partially purified from Lactococcus lactis subsp. lactis (Streptococcus diacetylactis) D10 and shown to have reductase activity with either diacetyl or acetoin as the substrate. However, the reductase activity with 10 mM diacetyl was far greater for both enzymes (7.0- and 4.7-fold for enzymes 1 and 2, respectively) than with 10 mM acetoin as the substrate. In contrast, when acetoin and diacetyl were present together, acetoin was the preferred substrate for both enzymes, with enzyme 1 showing the more marked preference for acetoin. meso-2,3-Butanediol was the only isomeric product, with enzyme 1 independent of the substrate combinations. For enzyme 2, both the meso and optical isomers of 2,3-butanediol were formed with acetoin as the substrate, but only the optical isomers were produced with diacetyl as the substrate. With batch cultures of strain D10 at or near the point of citrate exhaustion, the main isomers of 2,3-butanediol present were the optical forms. If the pH was sufficiently high (>pH 5), acetoin reduction occurred over time and was followed by diacetyl reduction, and meso-2,3-butanediol became the predominant isomer. Interconversion of the optical isomers into the meso isomer did occur. The properties of 2,3-butanediol dehydrogenases are consistent with diacetyl and acetoin removal and the appearance of the isomers of 2,3-butanediol.  相似文献   

14.
The catalytic mechanism of 6-phosphogluconate dehydrogenase requires the inversion of a Lys/Glu couple from its natural ionization state. The pKa of these residues in free and substrate bound enzymes has been determined measuring by ITC the proton release/uptake induced by substrate binding at different pH values. Wt 6-phosphogluconate dehydrogenase from Trypanosoma brucei and two active site enzyme mutants, K185H and E192Q were investigated. Substrate binding was accompanied by proton release and was dependent on the ionization of a group with pKa 7.07 which was absent in the E192Q mutant. Kinetic data highlighted two pKa, 7.17 and 9.64, in the enzyme–substrate complex, the latter being absent in the E192Q mutant, suggesting that the substrate binding shifts Glu192 pKa from 7.07 to 9.64. A comparison of wt and E192Q mutant appears to show that the substrate binding shifts Lys185 pKa from 9.9 to 7.17. By comparing differences in proton release and the binding enthalpy of wt and mutant enzymes, the enthalpic cost of the change in the protonation state of Lys185 and Glu192 was estimated at ≈ 6.1 kcal/mol. The change in protonation state of Lys185 and Glu192 has little effect on Gibbs free energy, 240–325 cal/mol. However proton balance evidences the dissociation of other group(s) that can be collectively described by a single pKa shift from 9.1 to 7.54. This further change in ionization state of the enzyme causes an increase of free energy with a total cost of 1.2–2.3 kcal/mol to set the enzyme into a catalytically competent form.  相似文献   

15.
Pyridoxal kinase (PLK; EC 2.7.1.35) is a key enzyme for vitamin B6 metabolism in animals. It catalyzes the ATP-dependent phosphorylation of pyridoxal, generating pyridoxal 5′-phosphate, an important cofactor for many enzymatic reactions. Bombyx mori PLK (BmPLK) is 10 or more residues shorter than mammalian PLKs, and some amino acid residues conserved in the PLKs from mammals are not maintained in the protein. Multiple sequence alignment suggested that amino acid residues Thr47, Ile54, Arg88, Asn121 and Glu230 might play important roles in BmPLK. In this study, we used a site-directed specific mutagenesis approach to determine the functional significance of these particular amino acid residues in BmPLK. Our results demonstrated that the mutation of Asn121 to Glu did not affect the catalytic function of BmPLK. The corresponding site-directed mutants of Thr47 to Asn, Ile54 to Phe, and Arg88 to Ile displayed a decreased catalytic efficiency and an elevated Km value for substrate relative to the wild-type value, and no enzyme activity could be detected in mutant of Trp230 to Glu. Circular dichroism analysis revealed that the mutation of Trp230 to Glu resulted in mis-folding of the protein. Our results provided direct evidence that residue Trp230 is crucial to maintain the structural and functional integrity of BmPLK. This study will add to the existing understanding of the characteristic of structure and function of BmPLK.  相似文献   

16.
ADP-glucose pyrophosphorylase is the enzyme responsible for the regulation of glycogen synthesis in bacteria. The enzyme N-terminal domain has a Rossmann-like fold with three neighbor loops facing the substrate ATP. In the Escherichia coli enzyme, one of those loops also faces the regulatory site containing Lys39, a residue involved in binding of the allosteric activator fructose-1,6-bisphosphate and its analog pyridoxal-phosphate. The other two loops contain Trp113 and Gln74, respectively, which are highly conserved among all the ADP-glucose pyrophosphorylases. Molecular modeling of the E. coli enzyme showed that binding of ATP correlates with conformational changes of the latter two loops, going from an open to a closed (substrate-bound) form. Alanine mutants of Trp113 or Gln74 did not change apparent affinities for the substrates, but they became insensitive to activation by fructose-1,6-bisphosphate. By capillary electrophoresis we found that the mutant enzymes still bind fructose-1,6-bisphosphate, with similar affinity as the wild type enzyme. Since the mutations did not alter binding of the activator, they must have disrupted the communication between the regulatory and the substrate sites. This agrees with a regulatory mechanism where the interaction with the allosteric activator triggers conformational changes at the level of loops containing residues Trp113 and Gln74.  相似文献   

17.
The depolymerization of complex glycans is an important biological process that is of considerable interest to environmentally relevant industries. β-Mannose is a major component of plant structural polysaccharides and eukaryotic N-glycans. These linkages are primarily cleaved by glycoside hydrolases, although recently, a family of glycoside phosphorylases, GH130, have also been shown to target β-1,2- and β-1,4-mannosidic linkages. In these phosphorylases, bond cleavage was mediated by a single displacement reaction in which phosphate functions as the catalytic nucleophile. A cohort of GH130 enzymes, however, lack the conserved basic residues that bind the phosphate nucleophile, and it was proposed that these enzymes function as glycoside hydrolases. Here we show that two Bacteroides enzymes, BT3780 and BACOVA_03624, which lack the phosphate binding residues, are indeed β-mannosidases that hydrolyze β-1,2-mannosidic linkages through an inverting mechanism. Because the genes encoding these enzymes are located in genetic loci that orchestrate the depolymerization of yeast α-mannans, it is likely that the two enzymes target the β-1,2-mannose residues that cap the glycan produced by Candida albicans. The crystal structure of BT3780 in complex with mannose bound in the −1 and +1 subsites showed that a pair of glutamates, Glu227 and Glu268, hydrogen bond to O1 of α-mannose, and either of these residues may function as the catalytic base. The candidate catalytic acid and the other residues that interact with the active site mannose are conserved in both GH130 mannoside phosphorylases and β-1,2-mannosidases. Functional phylogeny identified a conserved lysine, Lys199 in BT3780, as a key specificity determinant for β-1,2-mannosidic linkages.  相似文献   

18.
Despite extensive investigation of the irreversible oxidations undergone by proteins in vitro and in vivo, the products formed from the oxidation of Trp residues remain incompletely understood. Recently, we characterized a ditryptophan cross-link produced by the recombination of hSOD1-tryptophanyl radicals generated from attack of the carbonate radical produced during the bicarbonate-dependent peroxidase activity of the enzyme. Here, we examine whether the ditryptophan cross-link is produced by the attack of the carbonate radical on proteins other than hSOD1. To this end, we treated hen egg white lysozyme with photolytically and enzymatically generated carbonate radical. The radical yields were estimated and the lysozyme modifications were analyzed by SDS-PAGE, western blot, enzymatic activity and MS/MS analysis. Lysozyme oxidation by both systems resulted in its inactivation and dimerization. Lysozyme treated with the photolytic system presented monomers oxidized to hydroxy-tryptophan at Trp28 and Trp123 and N-formylkynurenine at Trp28, Trp62 and Trp123. Lysozyme treated with the enzymatic system rendered monomers oxidized to N-formylkynurenine at Trp28. The dimers were characterized as lysozyme-Trp28-Trp28-lysozyme and lysozyme-Trp28-Trp32-hSOD1. The results further demonstrate that the carbonate radical is prone to causing biomolecule cross-linking and hence, may be a relevant player in pathological mechanisms. The possibility of exploring the formation of ditryptophan cross-links as a carbonate radical biomarker is discussed.  相似文献   

19.
Deleting the box C/D RNA-containing intron in the Haloferax volcanii tRNATrp gene abolishes RNA-guided 2′-O methylations of C34 and U39 residues of tRNATrp. However, this deletion does not affect growth under standard conditions.  相似文献   

20.
STUDIES on the primary structure of aldolases isolated from ox, pig and rabbit muscle show that the amino-acid sequence of fructose 1,6-diphosphate aldolase [EC 4.1.2.13] has been highly conserved throughout mammalian evolution1. But comparison of the primary structure of the enzyme from these species with that from the muscle of a single North Sea sturgeon, presumably Acipenser sturio, indicated that although the proteins were homologous, a number of amino-acid replacements occurred between sturgeon aldolase and the aldolases of the phylogenetically distant mammalian species1. As a knowledge of the nature and number of amino-acid replacements between homologous proteins caft provide information both about the functional role of individual residues and about evolution, further comparative studies of rabbit and sturgeon aldolases were undertaken and an account of the sequence homology around the active-site-lysine residue of aldolases from rabbit muscle, rabbit liver and the muscle of the river sturgeon of Eastern Canada, Acipenser fulvescens, has been given2,3.  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