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 437 毫秒
1.
Prokaryotic release factor RF3 is a stimulatory protein that increases the rate of translational termination by the decoding release factors RF1 and RF2. The favoured model for RF3 function is the recycling of RF1 and RF2 after polypeptide release by displacing the factors from the ribosome. In this study, we have demonstrated that RF3 also plays an indirect role in the decoding of stop signals of highly expressed genes and recoding sites by accentuating the influence of the base following the stop codon (+4 base) on termination signal strength. The efficiency of decoding strong stop signals (e.g. UAAU and UAAG) in vivo is markedly improved with increased RF3 activity, while weak signals (UGAC and UAGC) are only modestly affected. However, RF3 is not responsible for the +4 base influence on termination signal strength, since prfC- strains lacking the protein still exhibit the same qualitative effect. The differential effect of RF3 at stop signals can be mimicked by modest overexpression of decoding RF. These findings can be interpreted according to current views of RF3 as a recycling factor, which functions to maintain the concentration of free decoding RF at stop signals, some of which are highly responsive to changes in RF levels.  相似文献   

2.
Prokaryotic translational release factors, RF1 and RF2, catalyze polypeptide release at UAG/UAA and UGA/UAA stop codons, respectively. In this study, we isolated a bacterial RF2 mutant (RF2*) containing an E167K substitution that restored the growth of a temperature-sensitive RF1 strain of Escherichia coli and the viability of a chromosomal RF1/RF2 double knockout. In both in vivo and in vitro polypeptide termination assays, RF2* catalyzed UAG/UAA termination, as does RF1, as well as UGA termination, showing that RF2* acquired omnipotent release activity. This result suggests that the E167K mutation abolished the putative third-base discriminator function of RF2. These findings are interpreted as indicating that prokaryotic and eukaryotic release factors share the same anticodon moiety and that only one omnipotent release factor is sufficient for bacterial growth, similar to the eukaryotic single omnipotent factor.  相似文献   

3.
RF3 was initially characterized as a factor that stimulates translational termination in an in vitro assay. The factor has a GTP binding site and shows sequence similarity to elongation factors EF-Tu and EF-G. Paradoxically, addition of GTP abolishes RF3 stimulation in the classical termination assay, using stop triplets. We here show GTP hydrolysis, which is only dependent on the simultaneous presence of RF3 and ribosomes. Applying a new termination assay, which uses a minimessenger RNA instead of separate triplets, we show that GTP in the presence of RF3 stimulates termination at rate-limiting concentrations of RF1. We show that RF3 can substitute for EF-G in RRF-dependent ribosome recycling reactions in vitro. This activity is GTP-dependent. In addition, excess RF3 and RRF in the presence of GTP caused release of nonhydrolyzed fmet-tRNA. This supports previous genetic experiments, showing that RF3 might be involved in ribosomal drop off of peptidyl-tRNA. In contrast to GTP involvement of the above reactions, stimulation of termination with RF2 by RF3 was independent of the presence of GTP. This is consistent with previous studies, indicating that RF3 enhances the affinity of RF2 for the termination complex without GTP hydrolysis. Based on our results, we propose a model of how RF3 might function in translational termination and ribosome recycling.  相似文献   

4.
The VanC phenotype for clinical resistance of enterococci to vancomycin is exhibited by Enterococcus gallinarum and Enterococcus casseliflavus. Based on the detection of the cell precursor UDP-N-acetylmuramic acid pentapeptide intermediate terminating in D-Ala-D-Ser instead of D-Ala-D-Ala, it has been predicted that the VanC ligase would be a D-Ala-D-Ser rather than a D-Ala-D-Ala ligase. Overproduction of the E. casseliflavus ATCC 25788 vanC2 gene in Escherichia coli and its purification to homogeneity allowed demonstration of ATP-dependent D-Ala-D-Ser ligase activity. The kcat/Km2 (Km2 = Km for D-Ser or C-terminal D-Ala) ratio for D-Ala-D-Ser/D-Ala-D-Ala dipeptide formation is 270/0.69 for a 400-fold selection against D-Ala in the C-terminal position. VanC2 also has substantial D-Ala-D-Asn ligase activity (kcat/Km2 = 74 mM-1min-1).  相似文献   

5.
The binding properties of substrates and competitive inhibitors of Escherichia coli cytidine deaminase are compared with those of the fragments obtained by cutting these ligands at several positions including the glycosidic bond. In contrast with the normal substrate cytidine (kcat/Km = 2.6 x 10(6) M-1 s-1), cytosine is found to serve as an extremely slow substrate (kcat/Km = 1.8 x 10(-3) M-1 s-1), despite the ability of cytosine to enter any active site that can accommodate the normal substrate cytidine. Spontaneous nonenzymatic deamination proceeds at similar rates for cytosine and cytidine at pH 7 and 25 degrees C, indicating that substituent ribose exerts little effect on the intrinsic reactivity of cytidine in solution. Dividing knon by kcat/Km, the maximal Kd value of the enzyme's complex with the altered substrate in the transition state is estimated as 6.1 x 10(-8) M for cytosine, very much higher than the value (1.2 x 10(-16) M) estimated for cytidine. The Kd value of ribofuranose, the missing substituent, is roughly 1.8 x 10(-2) M, as indicated by the Ki values of D-ribose and 1-methyl-D-ribofuranoside as competitive inhibitors. Thus, the free energy of binding of the altered substrate in the transition state is 9.5 kcal/mol more favorable for the whole molecule cytidine than for the sum of those of its parts, cytosine plus ribofuranose. As a separate molecule, however, ribose shows no detectable effect on the enzyme's activity on cytosine. Connectivity effects of similar magnitude are indicated by the equilibrium binding affinities of inhibitors. Thus, the Ki value of the transition state analogue inhibitor zebularine hydrate (1.2 x 10(-12) M) is very much lower than the combined affinities of N-ribofuranosylurea (1.6 x 10(-4) M) and allyl alcohol (0.14 M), indicating that the glycoside bond, by its presence, exerts a connectivity effect of 9.9 kcal/mol on the observed free energy of binding.  相似文献   

6.
Ricin A-chain (RTA) catalyzes the depurination of a single adenine at position 4324 of 28S rRNA in a N-ribohydrolase reaction. The mechanism and specificity for RTA are examined using RNA stem-loop structures of 10-18 nucleotides which contain the required substrate motif, a GAGA tetraloop. At the optimal pH near 4.0, the preferred substrate is a 14-base stem-loop RNA which is hydrolyzed at 219 min-1 with a kcat/Km of 4.5 x 10(5) M-1 s-1 under conditions of steady-state catalysis. Smaller or larger stem-loop RNAs have lower kcat values, but all have Km values of approximately 5 microM. Both the 10- and 18-base substrates have kcat/Km near 10(4) M-1 s-1. Covalent cross-linking of the stem has a small effect on the kinetic parameters. Stem-loop DNA (10 bases) of the same sequence is also a substrate with a kcat/Km of 0.1 that for RNA. Chemical mechanisms for enzymatic RNA depurination reactions include leaving group activation, stabilization of a ribooxocarbenium transition state, a covalent enzyme-ribosyl intermediate, and ionization of the 2'-hydroxyl. A stem-loop RNA with p-nitrophenyl O-riboside at the depurination site is not a substrate, but binds tightly to the enzyme (Ki = 0.34 microM), consistent with a catalytic mechanism of leaving group activation. The substrate activity of stem-loop DNA eliminates ionization of the 2'-hydroxyl as a mechanism. Incorporation of the C-riboside formycin A at the depurination site provides an increased pKa of the adenine analogue at N7. Binding of this analogue (Ki = 9.4 microM) is weaker than substrate which indicates that the altered pKa at this position is not an important feature of transition state recognition. Stem-loop RNA with phenyliminoribitol at the depurination site increases the affinity substantially (Ki = 0.18 microM). The results are consistent with catalysis occurring by leaving group protonation at ring position(s) other than N7 leading to a ribooxocarbenium ion transition state. Small stem-loop RNAs have been identified with substrate activity within an order of magnitude of that reported for intact ribosomes.  相似文献   

7.
Leukotriene-A4 hydrolase (EC 3.3.2.6) cleaved the NH2-terminal amino acid from several tripeptides, typified by arginyl-glycyl-aspartic acid, arginyl-glycyl-glycine, and arginyl-histidyl-phenylalanine, with catalytic efficiencies (kcat/Km) > or = 1 x 10(6) M-1 s-1. This exceeds by 10-fold the kcat/Km for its lipid substrate leukotriene A4. Catalytic efficiency declined for dipeptides which had kcat/Km ratios 10-100-fold lower than tripeptides. Tetrapeptides and pentapeptides were even poorer substrates with catalytic efficiencies below 10(3) M-1 s-1. The enzyme preferentially hydrolyzed tripeptide substrates and single amino acid p-nitroanilides with L-arginine at the NH2 terminus. Peptides with proline at the second position were not hydrolyzed, suggesting a requirement for an N-hydrogen at the peptide bond cleaved. Peptides with a blocked NH2 terminus were not hydrolyzed. The specificity constant (kcat/Km) was optimal at pH 7.2 with pK values at 6.8 and 7.9; binding was maximal at pH 8.0. Serum albumins activated the peptidase, increasing tripeptide affinities (Km) by 3-10-fold and specificities (kcat/Km) by 4-13-fold. Two known inhibitors of arginine peptidases, arphamenine A and B, inhibited hydrolysis of L-arginine p-nitroanilide with dissociation constants = 2.0 and 2.5 microM, respectively. Although the primary role of LTA4 hydrolase is widely regarded as the conversion of the lipid substrate leukotriene A4 into the inflammatory lipid mediator leukotriene B4, our data are the first showing that tripeptides are "better" substrates. This is compatible with a biological role for the peptidase activity of the enzyme and may be relevant to the distribution of the enzyme in organs like the ileum, liver, lung, and brain. We present a model which accommodates the available data on the interaction of substrates and inhibitors with the enzyme. This model can account for overlap in the active site for hydrolysis of leukotriene A4 and peptide or p-nitroanilide substrates.  相似文献   

8.
The biologically active isomer of 5-fluoro-5,6-dihydrouracil [(R)-5-fluoro-5,6-dihydrouracil, R-FUH2] was synthesized to study the kinetics of its enzymatic oxidation and hydrolysis by homogeneous dihydropyrimidine dehydrogenase (DPDase) and dihydropyrimidine aminohydrolase (DPHase), respectively. DPDase catalyzed the slow oxidation of R-FUH2 at pH 8 and 37 degrees with a Km of 210 microM and a kcat of 0.026 sec-1 at a saturating concentration of NADP+. The catalytic efficiency (kcat/Km) of DPDase for R-FUH2 was 1/14th of that for 5,6-dihydrouracil (UH2). In the opposite direction, DPDase catalyzed the reduction of 5-fluorouracil (FU) with a Km of 0.70 microM and a kcat of 3 sec-1 at a saturating concentration of NADPH. Thus, DPDase catalyzed the reduction of FU 30,000-fold more efficiently than the oxidation of R-FUH2. In contrast to the slow oxidation of R-FUH2 by DPDase, R-FUH2 was hydrolyzed very efficiently by DPHase with a Km of 130 microM and a kcat of 126 sec-1. The catalytic efficiency of DPHase for the hydrolysis of R-FUH2 was approximately twice that for the hydrolysis of UH2. Because R-FUH2 is hydrolysis of R-FUH2 was approximately twice that for the hydrolysis of UH2. Because R-FUH2 is hydrolyzed considerably more efficiently than it is oxidized and because the activity of DPHase was 250- to 500-fold greater than that of DPDase in bovine and rat liver, the hydrolytic pathway should predominate in vivo.  相似文献   

9.
Pseudomonas carboxyl proteinase (PCP), isolated from Pseudomonas sp. 101, and Xanthomonas carboxyl proteinase (XCP), isolated from Xanthomonas sp. T-22, are the first and second examples of unique carboxyl proteinases [EC 3.4.23.33] which are insensitive to aspartic proteinase inhibitors, such as pepstatin, diazoacetyl-DL-norleucine methylester, and 1,2-epoxy-3(p-nitrophenoxy)propane. The substrate specificities of PCP and XCP were studied using a series of synthetic chromogenic peptide substrates with the general structure, P5-P4-P3-P2-Phe-Nph-P2'-P3' (P5, P4, P3, P2, P2', P3': a variety of amino acids, Nph is p-nitro-L-phenylalanine, and the Phe-Nph bond is cleaved). PCP and XCP were shown to hydrolyze a synthetic substrate, Lys-Pro-Ala-Leu-Phe-Nph-Arg-Leu, most effectively among 28 substrates. The kinetic parameters of this peptide for PCP were Km = 6.3 microM, Kcat = 51.4 s-1, and kcat/Km = 8.16 microM-1.s-1. The kinetic parameters for XCP were Km = 3.6 microM, kcat = 52.2 s-1, and kcat/Km = 14.5 microM-1.s-1. PCP showed a stricter substrate specificity than XCP. That is, the specificity constant (kcat/Km) of each substrate for PCP was in general < 0.5 microM-1.s-1, but was drastically improved by the replacement of Lys by Leu at the P2 position. On the other hand, XCP showed a less stringent substrate specificity, with most of the peptides exhibiting reasonable kcat/Km values (> 1.0 microM-1.s-1). Thus it was found that the substrate specificities of PCP and XCP differ considerably, in spite of the high similarity in their primary structures. In addition, tyrostatin was found to be a competitive inhibitor for XCP, with a Ki value of 2.1 nM, as well as for PCP (Ki = 2.6 nM).  相似文献   

10.
Translational stop signals are defined in the genetic code as UAA, UAG and UGA, although the mechanism of their decoding via protein factors is clearly different from that of the other codons. There are strong biases in the upstream and downstream nucleotides surrounding stop codons. Experimental tests have shown that termination-signal strength is strongly influenced by the identity of the nucleotide immediately downstream of the codon (+4), with a correlation between the strength of this four-base signal and its occurrence at termination sites. The +4 nucleotide and other biases downstream of the stop codon may reflect sites of contact between the release factor and the mRNA, whereas upstream biases may be due to coding restrictions, with the release factor perhaps recognizing the final tRNA and the last two amino acids of the polypeptide undergoing synthesis. This means that the translational stop signal is probably larger than the triplet codon, but its exact length will be clearer when it is known which nucleotides are in direct contact with the release factor. Ultimately it will be defined exactly when a crystal structure of the release factor with its recognition substrate becomes available.  相似文献   

11.
The transmembrane PTPase HPTP beta differs from its related family members in having a single rather than a tandemly duplicated cytosolic catalytic domain. We have expressed the 354-amino acid, 41-kDa human PTP beta catalytic fragment in Escherichia coli, purified it, and assessed catalytic specificity with a series of pY peptides. HPTP beta shows distinctions from the related LAR PTPase and T cell CD45 PTPase domains: it recognizes phosphotyrosyl peptides of 9-11 residues from lck, src, and PLC gamma with Km values of 2, 4, and 1 microM, some 40-200-fold lower than the other two PTPases. With kcat values of 30-205 s-1, the catalytic efficiency, kcat/Km, of the HPTP beta 41-kDa catalytic domain is very high, up to 5.7 x 10(7) M-1 s-1. The peptides corresponding to PLC gamma (766-776) and EGFR (1,167-1,177) phosphorylation sites were used for structural variation to assess pY sequence context recognition by HPTP beta catalytic domain. While exchange of the alanine residue at the +2 position of the PLC gamma (Km of 1 microM) peptide to lysine or aspartic acid showed little or no effect on substrate affinity, replacement by arginine increased the Km 35-fold. Similarly, the high Km value of the EGFR pY peptide (Km of 104 microM) derives largely from the arginine residue at the +2 position of the peptide, since arginine to alanine single mutation at the -2 position of the EGFR peptide decreased the Km value 34-fold to 3 microM. Three thiophosphotyrosyl peptides have been prepared and act as substrates and competitive inhibitors of these PTPase catalytic domains.  相似文献   

12.
A number of cytochrome P450 (P450) 2E1 substrates are known to show kinetic deuterium isotope effects of approximately 5 on Km (DK = DKm/HKm), but not on kcat, in rat liver microsomes (e.g. N-nitrosodimethylamine, ethanol, and CH2Cl2). We observed DKm values of 3-5 for recombinant human P450 2E1-catalyzed ethanol oxidation. Replacing NADPH and O2 with the oxygen surrogate cumene hydroperoxide yielded similar results. Ferric P450 2E1 reduction was fast (k >1000 min-1) even in the absence of substrate. These results indicate that the basis for the increase in Km is in the latter portion of the catalytic cycle. The intrinsic isotope effect (Dk) for ethanol oxidation was determined (competitively) to be 3.8, indicating that C-H bond cleavage is isotopically sensitive. Pre-steady-state studies showed a burst of product formation (k = 410 min-1), with the burst amplitude corresponding to the P450 concentration. Deuteration of ethanol resulted in an isotope effect of 3.2 on the rate of the burst. We conclude that product release is rate-limiting in the oxidation of ethanol to acetaldehyde by P450 2E1. The steady-state kinetics can be described by a paradigm in which the kcat approximates the rate of product release, and Km is an expression in which the denominator is dominated by the rate of C-H bond breaking.  相似文献   

13.
In order to study the basis of resistance of human immunodeficiency virus, type 1 (HIV-1), to HIV-1 protease inhibitor saquinavir, the catalytic and inhibition properties of the wild-type HIV-1 protease and three saquinavir resistant mutants, G48V, L90M, and G48V/L90M, were compared. The kinetic parameter kcat/Km was determined for these proteases using eight peptide substrates whose sequences were derived from the natural processing site sequences of HIV-1. The kcat/Km values were determined using conventional steady-state kinetics as well as initial velocities of mixed substrate cleavages under the condition where the substrate concentrations [S]o < Km. The independently determined kcat and Km values for some of the substrates confirmed the accuracy of the mixed-substrate method and also permitted the calculation in all cases of true rather than relative kcat/Km values. The Ki values were also determined. Using a previously described kinetic model [Tang, J., & Hartsuck, J. A. (1995) FEBS Lett. 367, 112-116], the relative processing activities of HIV-1 protease variants were estimated in the saquinavir concentration range of 0-10(-7) M. Although the protease activity of G48V, L90M, and G48V/L90M are only about 10, 7, and 3% of that of the wild-type HIV-1 protease in the absence of inhibitor, the resistance tendencies of the three mutants are clearly manifest by relatively less activity loss as inhibitor concentration becomes higher. Also, the ratios of the activities of the four protease species at certain saquinavir concentrations appear to correlate with the population ratios of the four protease species at different time points of clinical trials. This correlation suggests that the population ratio of the protease species is driven by in vivo saquinavir concentration, which appears to be in the range 10(-10)-10(-9) M during the clinical trials.  相似文献   

14.
An in vivo translation assay system has been designed to measure, in one and the same assay, the three alternatives for a ribosome poised at a stop codon (termination, read-through and frameshift). A quantitative analysis of the competition has been done in the presence and absence of release factor (RF) mutants, nonsense suppressors and an upstream Shine-Dalgarno-like sequence. The ribosomal +1 frameshift product is measurable when the stop codon is decoded by wild-type or mutant RF (prf A1 or prf B2) and also in the presence of competing suppressor tRNAs. Frameshift frequency appears to be influenced by RF activity. The amount of frameshift product decreases in the presence of competing suppressor tRNAs, however, this decrease is not in proportion to the corresponding increase in the suppression product. Instead, there is an increase in the total amount of protein expressed from the gene, perhaps due to the purging of queued ribosomes. Mutated RFs reduce the total output of the reporter gene by reducing the amount of all three protein products. The nascent peptide has earlier been shown to influence the translation termination process by interacting with the RFs. At 42 degrees C in a temperature-sensitive RF mutant strain, protein measurements indicate that the nascent peptide seems to influence the binding efficiencies of the RFs.  相似文献   

15.
Two regions of the 16S rRNA, helix 34, and the aminoacyl site component of the decoding site at the base of helix 44, have been implicated in decoding of translational stop signals during the termination of protein synthesis. Antibiotics specific for these regions have been tested to see how they discriminate the decoding of UAA, UAG, and UGA by the two polypeptide chain release factors (RF-1 and RF-2). Spectinomycin, which interacts with helix 34, stimulated RF-1 dependent binding to the ribosome and termination. It also stimulated UGA dependent RF-2 termination at micromolar concentrations but inhibited UGA dependent RF-2 binding at higher concentrations. Alterations at position C1192 of helix 34, known to confer spectinomycin resistance, reduced the binding of f[3H]Met-tRNA to the peptidyl-tRNA site. They also impaired termination in vitro, with both factors and all three stop codons, although the effect was greater with RF-2 mediated reactions. These alterations had previously been shown to inhibit EF-G mediated translocation. As perturbations in helix 34 effect both termination and elongation reactions, these results indicate that helix 34 is close to the decoding site on the bacterial ribosome. Several antibiotics, hygromycin, neomycin and tetracycline, specific for the aminoacyl site, were shown to inhibit the binding and function of both RFs in termination with all three stop codons in vitro. These studies indicate that decoding of all stop signals is likely to occur at a similar site on the ribosome to the decoding of sense codons, the aminoacyl site, and are consistent with a location for helix 34 near this site.  相似文献   

16.
The exoglucanase/xylanase from Cellulomonas fimi (Cex) has been subjected to a detailed kinetic investigation with a range of aryl beta-D-glycoside substrates. This enzyme hydrolyzes its substrates with net retention of anomeric configuration, and thus it presumably follows a double-displacement mechanism. Values of kcat are found to be invariant with pH whereas kcat/Km is dependent upon two ionizations of pKa = 4.1 and 7.7. The substrate preference of the enzyme increases in the order glucosides < cellobiosides < xylobiosides, and kinetic studies with a range of aryl glucosides and cellobiosides have allowed construction of Broensted relationships for these substrate types. A strong dependence of both kcat (beta 1g = -1) and kcat/Km (beta 1g = -1) upon leaving group ability is observed for the glucosides, indicating that formation of the intermediate is rate-limiting. For the cellobiosides a biphasic, concave downward plot is seenj for kcat, indicating a change in rate-determining step across the series. Pre-steady-state kinetic experiments allowed construction of linear Broensted plots of log k2 and log (k2/Kd) for the cellobiosides of modest (beta 1g = -0.3) slope. These results are consistent with a double-displacement mechanism in which a glycosyl-enzyme intermediate is formed and hydrolyzed via oxocarbonium ion-like transition states. Secondary deuterium kinetic isotope effects and inactivation experiments provide further insight into transition-state structures and, in concert with beta 1g values, reveal that the presence of the distal sugar moiety in cellobiosides results in a less highly charged transition state.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

17.
Papaya proteinase omega (pp omega) has been purified from dried latex both by immunoaffinity and traditional methods. Kinetic analysis revealed that (1), the pp omega-catalysed hydrolysis of N-benzoyl-L-arginine p-nitroanilide (BApNA) has a lower specificity (kcat/Km) than the same reaction catalysed by papain; (2), the pp omega-catalysed hydrolysis of a tripeptide substrate having phenylalanine at the second position (S2-site) showed a more similar specificity to that catalysed by papain; (3), the significant difference between the two enzymes is that steady state kinetics with both L-BApNA and a tripeptide enables the identification in pp omega of other ionizations affecting binding. The active sites of papain and pp omega can therefore be distinguished by pH-dependence of kcat/Km.  相似文献   

18.
S-Ribosylhomocysteinase(LuxS) is the key enzyme in the synthetic pathway of a quorum sensing autoinducer AI-2. LuxS from a 2,3-butanediol produced strain Klebisella pneumoniae CICC 10011 was cloned and characterized. The luxS gene is composed of 540 bp with 172 amino acids encoded. The Km value for S-ribosylhomocysteine(SRH) was (27+1) μmol/L, kcat was (0.112+0.004) s-1 and kcat/Km was 4.4×103 L·mol-1· s-1 at 25℃. LuxS was activated by divalent metal ions, the highest activity was detected with Co2+ form, followed by Mg2+, Ba2+, Mn2+, Fe2+ and Ca2+, and activation constant for Co2+ is (16+2) μmol/L.  相似文献   

19.
Using a significantly simplified modification procedure, four charged analogues of the coenzyme NAD, N(1)- and N6-(2-hydroxy-3-trimethylammoniumpropyl)-NAD, N(1)- and N6-(3-sulfopropyl)-NAD were prepared. The kinetic parameters of these derivatives and N(1)-(2-aminoethyl)-NAD, N6-(2-aminoethyl)-NAD and tricyclic 1,N6-ethanoadenine-NAD, all with alterations to the adenine moiety, were determined for porcine heart lactate dehydrogenase isoenzyme H4. The coenzyme activity depends on both position and charge of the introduced groups. Modification of the N6-position leads to a 25-250-fold increase of the kcat/Km value compared to the related N(1) derivative. The kcat/Km value for 1,N6-ethanoadenine-NAD is in the range between that of N(1)-(2-aminoethyl)-NAD and N6-(2-aminoethyl)-NAD. In the case of both N(1) and N6 functionalization, the Km values increase from (3-sulfopropyl)-NAD, with a negatively charged substituent at the adenine, over (2-amino-ethyl)-NAD to (2-hydroxy-3-trimethylammoniumpropyl)-NAD with an uncharged and positively charged substituent, respectively, at the adenine. All N6 derivatives are analogues like NAD with respect to Km and/or Vmax and kcat/Km. The conformation of NAD and its derivatives was calculated and their interaction in the active site of lactate dehydrogenase was simulated using the molecular mechanics program AMBER. The significant differences in activity in correlation to porcine heart lactate dehydrogenase isoenzyme H4 could be rationalized by modelling the three-dimensional structure of the NAD site.  相似文献   

20.
The kinetic constants for horse urinary kallikrein and trypsin hydrolysis of BAEE, TAME, bradykinin methyl ester and bradykinyl-Ser-Val-Gin-Val-Ser were determined. The values of the ratio kcat/Km show that (1) kallikrein is catalytically less efficient than trypsin for all the substrates (2) the three esters are equally good substrates for trypsin while horse urinary kallikrein is 100-fold more effective on bradykinin methyl ester than on the other substrates (3) for both enzymes the ester of bradykinin is a better substrate than the tetradecapeptide.  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