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.
2.
Horse liver alcohol dehydrogenase occurs as isoenzymes: E is active on ethanol but not steroids; S is active on ethanol and steroids. The cDNAs for these isoenzymes were cloned; both were 1.8-kilobase long and contained complete coding sequences. Both enzymes were expressed in Escherichia coli, and the purified proteins had properties similar to those of the natural enzymes. The amino acid sequence deduced from the open reading frame of the E-type cDNA agreed with the amino acid sequence of the E isoenzyme determined by protein sequencing and x-ray crystallography. When compared with the E-type cDNA, the coding region of the S-type cDNA contains 24 substitutions and 3 deletions, giving rise to an amino acid sequence for the S. isoenzyme that differs from that of the E isoenzyme at 10 positions: nine conservative substitutions and one deletion, of Asp-115. These changes can be accommodated in the three-dimensional structure of the E isoenzyme, and models of the E and S isoenzymes complexed with a 3 beta-hydroxy-5 beta-steroid were built. The modeling shows that Leu-116 apparently sterically hinders binding of steroids in the E isoenzyme, and deletion in the S isoenzyme of Asp-115 moves Leu-116 and relieves the hindrance. The human gamma and rat liver enzymes are also active on steroids, but they have a different constellation of amino acid residues in the substrate pocket. Thus, there are multiple bases for the activity on steroids.  相似文献   

3.
The family 3 beta-glucosidase from Thermotoga maritima is a highly thermostable enzyme (85 degrees C) that displays transglycosylation activity. In contrast, the beta-glucosidase from Cellvibrio gilvus is mesophilic (35 degrees C) and displays no such transglycosylation activity. Both enzymes consist of two domains, an N-terminal and a C-terminal domain, and the amino acid identities between the two enzymes in these domains are 32.4 and 36.4%, respectively. In an attempt to identify the molecular basis underpinning the display of transglycosylation activity and the requirements for thermal stability, eight chimeric genes were constructed by shuffling the two parental beta-glucosidase genes at four selected borders, two in the N-terminal domain and two in the C-terminal domain. Of the eight chimeric genes constructed, only two chimeric enzymes (Tm578/606Cg and Tm638/666Cg) gave catalytically active forms and these were the ones shuffled in the C-terminal domain. For these active chimeric enzymes, 80% (Tm578/606Cg) and 88% (Tm638/666Cg) of their amino acid sequences originated from T. maritima. With regard to their thermal profiles, the two active chimeric enzymes, Tm578/606Cg and Tm638/666Cg, displayed profiles intermediate to those of the two parental enzymes as they were optimally active at 65 and 70 degrees C, respectively. These two chimeric enzymes were optimally active at pH 4.1 and 3.9, which is closer to that observed for the T. maritima enzyme (pH 3.2-3.5) than that for the C. gilvus enzyme (pH 6.2-6.5). Kinetic parameters for the chimeric enzymes were investigated with five different substrates including pNP-beta-D-glucopyranoside. The kinetic parameters obtained for the chimeric enzymes were closer to those of the T. maritima enzyme than to those of the C. gilvus enzyme. Transglycosylation activity was observed for both chimeric enzymes and the activity of the Tm578/606Cg chimera was at a level twice that observed with the T. maritima enzyme. This study is an effective demonstration of the usefulness of chimeric enzymes in altering the characteristics of an enzyme.  相似文献   

4.
The tyrosine (eTATase) and aspartate (eAATase) aminotransferases of Escherichia coli transaminate diacarboxylic amino acids with similar rate constants. However, eTATase exhibits approximately 10(2)-10(4)-fold higher second-order rate constants for the transamination of aromatic amino acids than does eAATase. A series of natural and unnatural amino acid substrates was used to quantitate specificity differences for these two highly related enzymes. A general trend toward lower transamination activity with increasing side-chain length (extending from aspartate to glutamate to alpha-aminoadipate) is observed for both enzymes. This result suggests that dicarboxylate ligands associate with the two highly related enzymes in a similar manner. The high reactivity of the enzymes with L-Asp and L-Glu can be attributed to an ion pair interaction between the side-chain carboxylate of the amino acid substrate and the guanidino group of the active site residue Arg 292 that is common to both enzymes. A strong linear correlation between side-chain hydrophobicity and transamination rate constants obtains for n-alkyl side-chain amino substrates with eTATase, but not for eAATase. The present kinetic data support a model in which eAATase contains one binding mode for all classes of substrate, whereas the active site of eTATase allows an additional mode that has increased affinity for hydrophobic amino acid.  相似文献   

5.
Molybdenum enzymes containing the pterin cofactor are a diverse group of enzymes that catalyse in general oxygen atom transfer reactions. Aiming at studying the amino acid residues, which are important for the enzymatic specificity, we used nitrate reductase from Ralstonia eutropha (R.e.NAP) as a model system for mutational studies at the active site. We mutated amino acids at the Mo active site (Cys181 and Arg421) as well as amino acids in the funnel leading to it (Met182, Asp196, Glu197, and the double mutant Glu197-Asp196). The mutations were made on the basis of the structural comparison of nitrate reductases with formate dehydrogenases (FDH), which show very similar three-dimensional structures, but clear differences in amino acids surrounding the active site. For mutations Arg421Lys and Glu197Ala we found a reduced nitrate activity while the other mutations resulted in complete loss of activity. In spite of the partial of total loss of nitrate reductase activity, these mutants do not, however, display FDH activity.  相似文献   

6.
Two conjugated polyketone reductases (CPRs) were isolated from Candida parapsilosis IFO 0708. The primary structures of CPRs (C1 and C2) were analyzed by amino acid sequencing. The amino acid sequences of both enzymes had high similarity to those of several proteins of the aldo-keto-reductase (AKR) superfamily. However, several amino acid residues in the putative active sites of AKRs were not conserved in CPRs-C1 and -C2.  相似文献   

7.
Using the MolScript version 2.1 computer program for protein molecule modeling and X-ray structure analysis data the spatial structures of several hydrolytic enzymes have been compared. These include glucoamylase from Aspergillus awamori and Saccharomycopsis fibuligera and lipases from Rhizopus japonicus. Results on homology of amino acid sequences and topology of secondary structure elements were obtained. 3D models of these enzymes with positioning of functionally important groups in the active site cavity were built.  相似文献   

8.
Rittschof D  Cohen JH 《Peptides》2004,25(9):1503-1516
Crustacean peptide pheromones, kairomones, and substituted amino sugar kairomones are reviewed from a historical perspective. These crustacean information molecules are secondary functions of structural polymers. They are partial hydrolysis products, generated usually by the action of trypsin-like enzymes on proteins, and glycosidase enzymes on glycoproteins and proteoglycans. Structure-function studies based upon synthetic mimics of peptide information molecules show neutral amino acids with a basic carboxyl terminal are active in modifying physiological and or behavioral responses. Behaviorally active substituted amino sugar mimics are disaccharide hydrolysis products of heparin and chondroitin sulfate. Similar molecules are also used as information molecules by a variety of other marine organisms indicating they are a common biological theme.  相似文献   

9.
The importance of five amino acids at the active site of the multicomponent naphthalene dioxygenase (NDO) system was determined by generating site-directed mutations in various combinations. The substrate specificities of the mutant enzymes were tested with the substrates indole, indoline, 2-nitrotoluene (2NT), naphthalene, biphenyl, and phenanthrene. Transformation of these substrates measured the ability of the mutant enzymes to catalyze dioxygenation, monooxygenation, and desaturation reactions. In addition, the position of oxidation and the enantiomeric composition of products were characterized. All enzymes with up to three amino acid substitutions were able to catalyze dioxygenation reactions. A subset of these enzymes could also catalyze the monooxygenation of 2NT and desaturation of indoline. Single amino acid substitutions at positions 352 and 206 had the most profound effects on product formation. Of the single mutations made, only changes at position 352 affected the stereochemistry of naphthalene cis-dihydrodiol formed from naphthalene, but in the presence of the F352I mutation, changes at positions 206 and 295 also affected enantioselectivity. Major shifts in regioselectivity with biphenyl and phenanthrene resulted with several of the singly, doubly, and triply mutated enzymes. A new product not formed by the wild-type enzyme, phenanthrene cis-9,10-dihydrodiol, was formed as a major product from phenanthrene by enzymes with two (A206I/F352I) or three amino acid substitutions (A206I/F352I/H295I). The results indicate that a variety of amino acid substitutions are tolerated at the active site of NDO. Journal of Industrial Microbiology & Biotechnology (2001) 27, 94–103. Received 25 September 2000/ Accepted in revised form 29 June 2001  相似文献   

10.
Monochlorotrifluoro-p-benzoquinone (CFQ) was used for investigating the state of the amino groups of acid-stable α-amylase and acid-unstable α-amylase. About half of the total amino groups in both enzyme molecules were reacted with the reagent. The unreactive amino groups seemed to exist in a different state from the reactive ones. Both enzymes whose amino groups were modified by CFQ still maintained the α-phenylmaltosidase activity in spite of losing or decreasing the amylase activity. These facts suggest that the amino groups of both enzymes were not in the active site but the modification of them caused steric hindrance.

The pH-stability of the acid-unstable α-amylase whose one or two amino groups were modified with succinic anhydride or 2,4,6-trinitrobenzene-l-sulfonate (TNBS) increased on the acidic side and decreased on the alkaline side, but further modification of them led to decrease the stability on both sides.  相似文献   

11.
Comparison and multiple alignments of amino acid sequences of a representative number of related enzymes demonstrate the existence of certain positions of amino acid residues which are permanently reproducible in all members of the whole family. The use of the bioinformatic approach revealed conservative residues in each of the related enzymes and ranked amino acid conservatism for the overall enzymatic catalysis. Glycine and aspartic acid residues were shown to be the most essential for structure and catalytic activity of enzymes. Amino acid residues forming catalytic subsite of the active site of enzymes are always highly conservative. Analysis revealed that aspartic acid carboxyl group is the most frequently employed nucleophilic (in deprotonated form) and electrophilic (in protonated form) agent involved in activation of molecules by the mechanism of general base and acidic catalyses in the catalytic sites of enzymes. Glycine is a unique amino acid possessing the highest possibilities for rotation along C–C and C–N bonds of the polypeptide chain. The conservative fixation of the glycine residue in polypeptide chains of related enzymes provides a possibility for directed assembly of amino acid residues into the catalytic subsite structure. It is possible that the conservative glycines provide known conformational mobility of the protein and the active site. Methods of molecular modeling were used for analysis of structural substitutions of conservative and non-conservative glycines and their effects on geometry of catalytic site of typical hydrolases. The substitution of glycine(s) for alanine significantly altered the catalytic site structures.  相似文献   

12.
alpha-Amylase cDNA was cloned and sequenced from Aspergillus shirousamii RIB2504. The putative protein deduced from the cDNA open reading frame (ORF) consisted of 499 amino acids with a molecular weight of 55,000. The amino acid sequence was identical to that of the ORF of the Taka-amylase A gene of Aspergillus oryzae, while the nucleotide sequence was different at two and six positions in the cDNA ORF and 3' non-coding regions, respectively, so far determined. The alpha-amylase cDNA was expressed in Saccharomyces cerevisiae under the control of the yeast ADH1 promoter using a YEp-type plasmid, pYcDE1. The cDNA of glucoamylase, which was previously cloned from the same organism, was also expressed under the same conditions. Consequently, active alpha-amylase and glucoamylase were efficiently secreted into the culture medium. The amino acid sequence of the N-terminal regions of these enzymes purified from the yeast culture medium confirmed that the signal sequences of these enzymes were cleaved off at the same positions as those of the native enzymes of A. shirousamii.  相似文献   

13.
Plant aromatic amino acid decarboxylase (AAAD) enzymes are capable of catalyzing either decarboxylation or decarboxylation-deamination on various combinations of aromatic amino acid substrates. These two different activities result in the production of arylalkylamines and the formation of aromatic acetaldehydes, respectively. Variations in product formation enable individual enzymes to play different physiological functions. Despite these catalytic variations, arylalkylamine and aldehyde synthesizing AAADs are indistinguishable without protein expression and characterization. In this study, extensive biochemical characterization of plant AAADs was performed to identify residues responsible for differentiating decarboxylation AAADs from aldehyde synthase AAADs. Results demonstrated that a tyrosine residue located on a catalytic loop proximal to the active site of plant AAADs is primarily responsible for dictating typical decarboxylase activity, whereas a phenylalanine at the same position is primarily liable for aldehyde synthase activity. Mutagenesis of the active site phenylalanine to tyrosine in Arabidopsis thaliana and Petroselinum crispum aromatic acetaldehyde synthases primarily converts the enzymes activity from decarboxylation-deamination to decarboxylation. The mutation of the active site tyrosine to phenylalanine in the Catharanthus roseus and Papaver somniferum aromatic amino acid decarboxylases changes the enzymes decarboxylation activity to a primarily decarboxylation-deamination activity. Generation of these mutant enzymes enables the production of unusual AAAD enzyme products including indole-3-acetaldehyde, 4-hydroxyphenylacetaldehyde, and phenylethylamine. Our data indicates that the tyrosine and phenylalanine in the catalytic loop region could serve as a signature residue to reliably distinguish plant arylalkylamine and aldehyde synthesizing AAADs. Additionally, the resulting data enables further insights into the mechanistic roles of active site residues.  相似文献   

14.
Isomalto-dextranase, from Arthrobacter globiformis T6, is a member of the glycoside hydrolase family 27. However, the alignments of the whole amino acid sequence are distinct from other members of this family. The enzymes cleave the glycosidic bond of the substrate in two different manners: either retaining or inverting the anomeric configuration. We believe that a retaining enzyme is involved in a two-step, double-displacement mechanism utilizing active site carboxylic acids as the nucleophile and general acid/base catalysts in the hydrolytic reaction. The critical amino acid residues at the isomalto-dextranase active site that catalyzes the hydrolysis reaction of dextran have been identified and the roles of nine amino acid residues (D107, D163, D227, D295, D340, D342, D373, D396, and E420) in the isomalto-dextranase from A. globiformis analyzed by site-directed mutagenesis. Of 15 mutant enzymes that were prepared, eight had reduced activities for dextran hydrolysis. Aspartic acids-227 and -342, which are part of the apparent catalytic dyad, were essential for hydrolase activity toward dextran.  相似文献   

15.
Haloalkaliphilic microorganisms isolated from soda lakes were compared in terms of the amino acid composition of total cellular protein and the reaction of a number of key enzymes to salts and pH of the medium. In the extremely halophilic bacterium Natroniella acetigena (salt-inside osmoadaptation strategy), acidic amino acids (glutamic and aspartic) made up 30.91 mol % of the total of cellular protein amino acids. In the moderate haloalkaliphiles Tindallia magadiensis, Halomonas campisalis, and Halomonas sp. AIR-1 (compatible-solutes osmoadaptation strategy), the proportion of acidic amino acids (24.36, 23.15, and 23.58 mol %, respectively) was lower than in N. acetigena but higher than in the freshwater Acetobacterium paludosum (20.77 mol %). The excess of acidic amino acids over basic amino acids (lysine and arginine) increased with the degree of halophily. The enzymes of haloalkaliphiles proved to be tolerant to salts and high pH values, although the degree of tolerance varied. The activity of N. acetigena CO dehydrogenase was maximum in the presence of 0.7 M NaCl, but it was virtually independent of the NaHCO3 concentration. The hydrogenase and CO dehydrogenase of T. magadiensis exhibited maximum activity in the absence of NaCl; the CO dehydrogenase was most active at 0.25 M NaHCO3, and hydrogenase activity was only weakly dependent on NaHCO3 in the concentration range of 0-1.2 M. The nitrate reductases of H. campisalis and Halomonas sp. AIR-2 were active in broad ranges of NaCl and KCl concentrations; the activity maxima were recorded at moderate concentrations of these salts. The pH optima of most of the studied enzymes of haloalkaliphiles were in the alkaline zone. Thus, it was shown that the amino acid composition of total cellular protein is determined by the osmoadaptation strategy employed by the bacterium. A correlation was found between the salt tolerance of enzymes and the proportion of acidic amino acids in the total cellular protein. The ability of enzymes to function at high pH values is one of the mechanisms of adaptation of microorganisms to high pH values.  相似文献   

16.
The gene (nplT) for a new type of pullulan-hydrolysing enzyme, neopullulanase, from Bacillus stearothermophilus TRS40 was sequenced. The DNA sequence revealed only one large open reading frame, composed of 1764 bases and 588 amino acid residues (Mr 69144). Although the thermostable neopullulanase contained eight cysteine residues, they did not provide conformational stability by disulphide bonds. A comparison was made of the amino acid sequences of alpha-amylase, neopullulanase, isoamylase, pullulanase and cyclodextrin glucanotransferase. All the enzymes examined contained four highly conserved regions which probably constitute the active centres of the enzymes. The amino acid residues required for the specificity of neopullulanase are compared with those of alpha-amylase and other amylolytic enzymes.  相似文献   

17.
We identified the critical amino acid residues for substrate recognition using two firefly luciferases from Pylocoeria miyako (PmL) and Hotaria parvura (HpL), as these two luciferase enzymes exhibit different activities toward ketoprofen. Specifically, PmL can catalyze the apparent enantioselective thioesterification reaction, while HpL cannot. By comparing the amino acid sequences around the active site, we identified two residues (I350 and M397 in PmL and F351 and S398 in HpL) that were different between the two enzymes, and the replacement of these amino acids resulted in changing the ketoprofen recognition pattern. The inactive HpL was converted to the active enzyme toward ketoprofen and vice versa for PmL. These residues also affected the enantioselectivity toward ketoprofen; however, the bioluminescent color was not affected. In addition, using molecular dynamics calculations, the replacement of these two amino acids induced changes in the state of hydrogen bonding between ketoprofen and the S349 side chain through the active site water. As S349 is not considered to influence color tuning, these changes specifically caused the differences in ketoprofen recognition in the enzyme.  相似文献   

18.
The genes of the BanI restriction-modification system specific for GGPyPuCC were cloned from the chromosomal DNA of Bacillus aneurinolyticus IAM1077, and the coding regions were assigned on the nucleotide sequence on the basis of the N-terminal amino acid sequences and molecular weights of the enzymes. The restriction and modification genes coded for polypeptides with calculated molecular weights of 39,841 and 42,637, respectively. Both the enzymes were coded by the same DNA strand. The restriction gene was located upstream of the methylase gene, separated by 21 bp. The cloned genes were significantly expressed in E. coli cells, so that the respective enzymes could be purified to homogeneity. Analysis by sodium dodecyl sulfate-polyacrylamide gel electrophoresis and gel filtration indicated that the catalytically active form of the endonuclease was dimeric and that of the methylase was monomeric. Comparison of the amino acid sequences revealed no significant homology between the endonuclease and methylase, though both enzymes recognize the same target sequence. Sequence comparison with other related enzymes indicated that BanI methylase contains sequences common to cytosine-specific methylases.  相似文献   

19.
1. Specific proteases which inactivate the apo-proteins of many pyridoxal enzymes were found in skeletal muscle, liver and small intestine of rats. The protease from these three organs were purified and their properties were compared. 2. The purified proteases from liver and skeletal muscle appeared homogeneous on acrylamide gel electrophoresis. Two different proteases were separated from small intestine. A homogeneous, crystalline enzyme was obtained from the muscle layer while enzyme from the mucosa was partially purified. 3. They showed substrate specificity for pyridoxal enzymes. Their pH optima were in an alkaline region. They showed activity with the substrate of chymotrypsin, N-acetyl-L-tyrosine ethyl ester, but not with that of trypsin, p-toluenesulfonyl-L-arginine ethyl ester. They were inhibited by pyridoxal phosphate or pyridoxamine phosphate and seryl residues were involved in their active center. 4. The four enzymes differed in the following characters: (a) molecular weights; (b) patterns of elution from a CM-Sephadex column; (c) rates of inactivation of substrate enzymes; (d) rates of cleavage of N-acetyl-L-tyrosine ethyl ester; (e) reactivities with antiserum against the enzyme from the muscle layer of small intestine; (f) specific activities. 5. The amino acid composition and effect of chemical modifications of the crystalline enzyme from the muscle layer of small intestine were examined to elucidate its active sites and mode of action. Serine and histidine residues were found to be essential for protease activity. A tyrosine residue was also necessary for activity. Modifications of its sulfhydryl group, amino residues and carboxyl group had no effect on its activity.  相似文献   

20.
The molecular basis for the specificity of aldehyde dehydrogenases (ALDHs) for retinal, the precursor of the morphogen retinoic acid, is still poorly understood. We have expressed in Escherichia coli both retinal dehydrogenase (RALDH), a cytosolic aldehyde dehydrogenase originally isolated from rat kidney, and the highly homologous phenobarbital-induced aldehyde dehydrogenase (PB-ALDH). Oxidation of propanal was observed with both enzymes. On the other hand, recombinant RALDH efficiently catalyzed oxidation of 9-cis- and all-trans-retinal, whereas PB-ALDH was inactive with all-trans-retinal and poorly active with 9-cis-retinal. A striking difference between PB-ALDH and all other class I ALDHs is the identity of the amino acid immediately preceding the active nucleophile Cys(302) (Ile(301) instead of Cys(301)). Nevertheless, these amino acids could be exchanged in either RALDH or PB-ALDH without affecting substrate specificity. Characterization of chimeric enzymes demonstrates that distinct groups of amino acids control the differential activity of RALDH and PB-ALDH with all-trans- and 9-cis-retinal. Of 52 divergent amino acids, the first 17 are crucial for activity with all-trans-retinal, whereas the next 25 are important for catalysis of 9-cis-retinal oxidation. Recombinant enzymes with specificity for all-trans- or 9-cis-retinal were obtained, which should provide useful tools to study the relative importance of local production of all-trans- versus 9-cis-retinoic acid in development and tissue differentiation.  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