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.
2.
Wise EL  Yew WS  Gerlt JA  Rayment I 《Biochemistry》2004,43(21):6438-6446
3-Keto-L-gulonate 6-phosphate decarboxylase (KGPDC), a member of the orotidine monophosphate decarboxylase (OMPDC) suprafamily, catalyzes the Mg(2+)-dependent decarboxylation of 3-keto-L-gulonate 6-phosphate to L-xylulose 5-phosphate. Structural and biochemical evidence suggests that the KGPDC reaction proceeds via a Mg(2+)-stabilized 1,2-cis-enediolate intermediate. Protonation of the enediolate intermediate occurs in a nonstereospecific manner to form L-xylulose 5-phosphate. Although the exact mechanism of proton delivery is not known, Glu112, His136, and Arg139 have been implicated in this process [Yew, W. S., Wise, E., Rayment, I., and Gerlt, J. A. (2004) Biochemistry 43, 6427-6437]. Surprisingly, single amino acid substitutions of these positions do not substantially reduce catalytic activity but rather alter the stereochemical course of the reaction. Here, we report the X-ray crystal structures of four mutants, K64A, H136A, E112Q, and E112Q/H136A, each determined in the presence of L-threonohydroxamate 4-phosphate, an analogue of the enediolate intermediate, to 1.7, 1.9, 1.8, and 1.9 A resolution, respectively. These structures reveal that substitutions of Lys64, Glu112, and His136 cause changes in the positions of the intermediate analogue and two active site water molecules that were previously identified as possible proton donors. These changes correlate with the observed alterations in the reaction stereochemistry for these mutants, thereby supporting a reaction mechanism in which water molecules competitively shuttle protons from the side chains of His136 and Arg139 to alternate faces of the cis-enediolate intermediate. These studies further underscore the wide variation in the reaction mechanisms in the OMPDC suprafamily.  相似文献   

3.
4.
5.
The "ribulose phosphate binding" superfamily defined by the Structural Classification of Proteins (SCOP) database is considered the result of divergent evolution from a common (beta/alpha)(8)-barrel ancestor. The superfamily includes d-ribulose 5-phosphate 3-epimerase (RPE), orotidine 5'-monophosphate decarboxylase (OMPDC), and 3-keto-l-gulonate 6-phosphate decarboxylase (KGPDC), members of the OMPDC suprafamily, as well as enzymes involved in histidine and tryptophan biosynthesis that utilize phosphorylated metabolites as substrates. We now report studies of the functional and structural relationships of RPE to the members of the superfamily. As suggested by the results of crystallographic studies of the RPEs from rice [Jelakovic, S., Kopriva, S., Suss, K. H., and Schulz, G. E. (2003) J. Mol. Biol. 326, 127-35] and Plasmodium falciparum [Caruthers, J., Bosch, J., Bucker, F., Van Voorhis, W., Myler, P., Worthey, E., Mehlin, C., Boni, E., De Titta, G., Luft, J., Kalyuzhniy, O., Anderson, L., Zucker, F., Soltis, M., and Hol, W. G. J. (2006) Proteins 62, 338-42], the RPE from Streptococcus pyogenes is activated by Zn(2+) which binds with a stoichiometry of one ion per polypeptide. Although wild type RPE has a high affinity for Zn(2+) and inactive apoenzyme cannot be prepared, the affinity for Zn(2+) is decreased by alanine substitutions for the two histidine residues that coordinate the Zn(2+) ion (H34A and H67A); these mutant proteins can be prepared in an inactive, metal-free form and activated by exogenous Zn(2+). The crystal structure of the RPE was solved at 1.8 A resolution in the presence of d-xylitol 5-phosphate, an inert analogue of the d-xylulose 5-phosphate substrate. This structure suggests that the 2,3-enediolate intermediate in the 1,1-proton transfer reaction is stabilized by bidentate coordination to the Zn(2+) that also is liganded to His 34, Asp 36, His 67, and Asp 176; the carboxylate groups of the Asp residues are positioned also to function as the acid/base catalysts. Although the conformation of the bound analogue resembles those of ligands bound in the active sites of OMPDC and KGPDC, the identities of the active site residues that coordinate the essential Zn(2+) and participate as acid/base catalysts are not conserved. We conclude that only the phosphate binding motif located at the ends of the seventh and eighth beta-strands of the (beta/alpha)(8)-barrel is functionally conserved among RPE, OMPDC, and KGPDC, consistent with the hypothesis that the members of the "ribulose phosphate binding" (beta/alpha)(8)-barrel "superfamily" as defined by SCOP have not evolved by evolutionary processes involving the intact (beta/alpha)(8)-barrel. Instead, this "superfamily" may result from assembly from smaller modules, including the conserved phosphate binding motif associated with the C-terminal (beta/alpha)(2)-quarter barrel.  相似文献   

6.
Wise EL  Yew WS  Gerlt JA  Rayment I 《Biochemistry》2003,42(42):12133-12142
3-Keto-L-gulonate 6-phosphate decarboxylase (KGPDC) and orotidine 5'-phosphate decarboxylase (OMPDC) are members of an enzyme suprafamily, the OMPDC suprafamily, because they are homologous enzymes that catalyze mechanistically distinct reactions using different substrates. KGPDC catalyzes the Mg(2+) ion-dependent decarboxylation of 3-keto-L-gulonate 6-phosphate to yield L-xylulose 5-phosphate and CO(2); OMPDC catalyzes the metal ion-independent decarboxylation of OMP to UMP and CO(2). Structural studies have shown that KGPDC and OMPDC share several strictly conserved active site residues that are used differently by each enzyme to catalyze their mechanistically distinct reactions. Although the mechanism of the KGPDC-catalyzed reaction has yet to be elucidated, it is thought to proceed via a Mg(2+) ion-stabilized 1,2-enediolate intermediate. Here we report the crystal structures of KGPDC complexed with L-gulonate 6-phosphate, L-threonohydroxamate 4-phosphate, and L-xylitol 5-phosphate, analogues of the substrate, enediolate intermediate, and product, as well as with the product, L-xylulose 5-phosphate, at 1.2, 1.8, 1.7, and 1.8 A resolution, respectively. These structures support a mechanism that involves the formation of a cis-1,2-enediolate intermediate. Contrary to expectations, the geometry of the intermediate does not involve bidentate coordination of both enediolate oxygen atoms to the Mg(2+) ion but rather involves only the coordination of the oxygen on C2 to the Mg(2+) ion. The oxygen atom on C1 instead forms hydrogen bonds to both Lys64 and Asp67, two strictly conserved active site residues. Lys64 also interacts with the oxygen on C2 and may serve to stabilize a cis conformation of the 1,2-enediolate. These structures also implicate His136 to be the general acid that protonates the 1,2-enediolate intermediate. This study further demonstrates that multiple unrelated enzyme functions can evolve from a single active site architecture without regard for substrate binding affinity or mechanism.  相似文献   

7.
The two preceding papers [Powers, V. M., Koo, C. W., Kenyon, G. L., Gerlt, J. A., & Kozarich, J. W. (1991) Biochemistry (first paper of three in this issue); Neidhart, D. J., Howell, P. L., Petsko, G. A., Powers, V. M., Li, R., Kenyon, G. L., & Gerlt, J. A. (1991) Biochemistry (second paper of three in this issue)] suggest that the active site of mandelate racemase (MR) contains two distinct general acid/base catalysts: Lys 166, which abstracts the alpha-proton from (S)-mandelate, and His 297, which abstracts the alpha-proton from (R)-mandelate. In this paper we report on the properties of the mutant of MR in which His 297 has been converted to asparagine by site-directed mutagenesis (H297N). The structure of H297N, solved by molecular replacement at 2.2-A resolution, reveals that no conformational alterations accompany the substitution. As expected, H297N has no detectable MR activity. However, H297N catalyzes the stereospecific elimination of bromide ion from racemic p-(bromomethyl)mandelate to give p-(methyl)-benzoylformate in 45% yield at a rate equal to that measured for wild-type enzyme; the unreacted p-(bromomethyl)mandelate is recovered as (R)-p-(hydroxymethyl)mandelate. At pD 7.5, H297N catalyzes the stereospecific exchange of the alpha-proton of (S)- but not (R)-mandelate with D2O solvent at a rate 3.3-fold less than that observed for incorporation of solvent deuterium into (S)-mandelate catalyzed by wild-type enzyme. The pD dependence of the rate of the exchange reaction catalyzed by H297N reveals a pKa of 6.4 in D2O, which is assigned to Lys 166.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

8.
Divergent evolution of enzyme function is commonly explained by a gene duplication event followed by mutational changes that allow the protein encoded by the copy to acquire a new function. An alternate hypothesis is that this process is facilitated when the progenitor enzyme acquires a second function while maintaining the original activity. This phenomenon has been suggested to occur in the o-succinylbenzoate synthase (OSBS) from a species of Amycolatopsis that catalyzes not only the physiological syn-dehydration reaction of 2-succinyl-6-hydroxy-2,4-cyclohexadiene-1-carboxylate but also an accidental racemization of N-acylamino acids [Palmer, D. R., Garrett, J. B., Sharma, V., Meganathan, R., Babbitt, P. C., and Gerlt, J. A. (1999) Biochemistry 38, 4252-4258]. To understand the molecular basis of this promiscuity, three-dimensional structures of liganded complexes of this enzyme have been determined, including the product of the OSBS reaction and three N-acylamino acid substrates for the N-acylamino acid racemase (NAAAR) reaction, N-acetylmethionine, N-succinylmethionine, and N-succinylphenylglycine, to 2.2, 2.3, 2.1, and 1.9 A resolution, respectively. These structures show how the active-site cavity can accommodate both the hydrophobic substrate for the OSBS reaction and the substrates for the accidental NAAAR reaction. As expected, the N-acylamino acid is sandwiched between lysines 163 and 263, which function as the catalytic bases for the abstraction of the alpha-proton in the (R)- and (S)-racemization reactions, respectively [Taylor Ringia, E. A., Garrett, J. B, Thoden, J. B., Holden, H. M., Rayment, I., and Gerlt, J. A. (2004) Biochemistry 42, 224-229]. Importantly, the protein forms specific favorable interactions with the hydrophobic amino acid side chain, alpha-carbon, carboxylate, and the polar components of the N-acyl linkage. Accommodation of the components of the N-acyl linkage appears to be the reason that this enzyme is capable of a racemization reaction on these substrates, whereas the orthologous OSBS from Escherichia coli lacks this functionality.  相似文献   

9.
Vick JE  Gerlt JA 《Biochemistry》2007,46(50):14589-14597
The molecular details of the processes involved in divergent evolution of "new" enzymatic functions are ill-defined. Likely starting points are either a progenitor promiscuous for the new reaction or a progenitor capable of catalyzing the new reaction following a single substitution that results from a single base change. However, the molecular (sequence) pathway by which the selective advantage provided by this protein can be improved and ultimately optimized is unclear. In the mechanistically diverse enolase superfamily, we discovered that a monofunctional progenitor could acquire the ability to catalyze a "new" reaction by a single base change: the D297G mutant of the monofunctional l-Ala-d/l-Glu epimerase (AEE) from Escherichia coli catalyzed a low level of the o-succinylbenzoate synthase (OSBS) reaction as well as a reduced level of the AEE reaction [Schmidt, D. M. Z., Mundorff, E. C., Dojka, M., Bermudez, E., Ness, J. E., Govindarajan, S., Babbitt, P. C., Minshull, J., and Gerlt, J. A. (2003) Biochemistry 42, 8387-8393]. We then discovered that the selective advantage and OSBS activity of the D297G mutant are both enhanced by the I19F substitution [Vick, J. E., Schmidt, D. M. Z., and Gerlt, J. A. (2005) Biochemistry 44, 11722-11729]. Both the D297G and I19F substitutions are positioned to alter the substrate specificity so that the substrate for the OSBS reaction is more productively positioned vis a vis the active site catalytic groups. We now report that both the selective advantage and OSBS activity of the D297G/I19F double mutant are enhanced by the R24C (one base change from the wild type Arg codon), R24W (two base changes from the wild type Arg codon and one base change from the R24C codon), and L277W (one base change from the wild type Leu codon) substitutions. The effects of the R24C and L277W mutants are "additive" in the D297G/I19F/R24C/L277W mutant. The greatest selective advantage and OSBS activity are associated with the D297G/I19F/R24W mutant. These "new" substitutions that enhance both the selective advantage and kinetic constants are positioned in the active site where they can alter the specificity, highlighting that the evolution of the "new" OSBS function can be accomplished by changes in substrate specificity.  相似文献   

10.
Geeganage S  Ling VW  Frey PA 《Biochemistry》2000,39(18):5397-5404
Galactose-1-phosphate uridylyltransferase (GalT) catalyzes the reversible transformation of uridine 5'-diphosphate glucose (UDPGlc) and galactose-1-phosphate into uridine 5'-diphosphate galactose (UDPGal) and glucose-1-phosphate through a double displacement mechanism, with the intermediate formation of a covalent uridylyl-enzyme (UMP-enzyme). The covalent linkage is a phosphoramidate formed between the UMP moiety and the His 166 N(epsilon)(2) of GalT, with His 166 N(delta1) retaining a proton throughout the catalytic cycle. Cys 160 and Ser 161 in Escherichia coli GalT are engaged in hydrogen bonding with the peripheral phosphoryl oxygen atoms of the substrate in the crystalline UMP-enzyme and in the crystalline complex of H166G-GalT with UDPGlc [Wedekind, J. E., Frey, P. A., and Rayment, I. (1996) Biochemistry 35, 11560-11569; Thoden, J. B., Ruzicka, F. J., Frey, P. A., Rayment, I., and Holden, H. M. (1997) Biochemistry 36, 1212-1222]. Site-directed mutagenesis, thermodynamic, transient kinetic, and steady-state kinetic studies have been performed to investigate the roles of Cys 160 and Ser 161 in catalysis. The absence of the thiol group of Cys 160 in the variants C160S and C160A did not seriously alter the enzymatic activity. However, the variant S161A displayed 7000-fold less activity than wild-type GalT. The low activity of S161A was directly related to impaired uridylylation rate constant (3.7 x 10(-)(2) s(-)(1)) and de-uridylylation rate constant (0.5 x 10(-)(2) s(-)(1)) resulting from a higher kinetic barrier for uridylyl-group transfer by the variant S161A as compared with the wild-type GalT. Equilibrium uridylylation studies showed that neither Cys 160 nor Ser 161 was involved in stabilizing the uridylyl-enzyme intermediate. The results lead to the conclusion that the conserved Cys 160 does not play a critical role in catalysis. Ser 161 is most likely involved in donating a hydrogen bond to the beta-phosphoryl group of a substrate, thereby providing proper orientation for nucleophilic catalysis.  相似文献   

11.
The high-resolution X-ray crystal structure of staphylococcal nuclease (SNase) suggests that the guanidinium groups of Arg 35 and Arg 87 participate as electrophilic catalysts in the attack of water on the substrate phosphodiester. Both arginine residues have been replaced with "conservative" lysine residues so that both the importance of these residues in catalysis and the effect of changes in electrostatic interactions on active site conformation can be assessed. The catalytic efficiencies of R35K and R87K are decreased by factors of 10(4) and 10(5) relative to wild-type SNase, with R87K showing a very significant reduction in its affinity for both DNA substrate and the competitive inhibitor thymidine 3',5'-bisphosphate (pdTp). The thermal denaturation behavior of both mutant enzymes differs from that of wild type both in the absence and in the presence of the active site ligands Ca2+ and pdTp. Both the 1H NMR chemical shifts and interresidue nuclear Overhauser effects (NOEs) of residues previously assigned to be in the hydrophobic core of SNase are altered in R35K and R87K. These observations, similar to those recently reported by our laboratories for substitutions for Glu 43 [Hibler, D. W., Stolowich, N. J., Reynolds, M. A., Gerlt, J. A., Wilde, J. A., & Bolton, P. H. (1987) Biochemistry 26, 6278; Wilde, J. A., Bolton, P. H., Dell'Acqua, M., Hibler, D. W., Pourmotabbed, T., & Gerlt, J. A. (1988) Biochemistry 27, 4127], suggest that lysine substitutions are not conservative in SNase and disrupt the conformation of the active site.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

12.
The reaction catalyzed by orotidine 5'-monophosphate decarboxylase (OMPDC) is accompanied by exceptional values for rate enhancement (k(cat)/k(non) = 7.1 × 10(16)) and catalytic proficiency [(k(cat)/K(M))/k(non) = 4.8 × 10(22) M(-1)]. Although a stabilized vinyl carbanion/carbene intermediate is located on the reaction coordinate, the structural strategies by which the reduction in the activation energy barrier is realized remain incompletely understood. This laboratory recently reported that "substrate destabilization" by Asp 70 in the OMPDC from Methanothermobacter thermoautotrophicus (MtOMPDC) lowers the activation energy barrier by ~5 kcal/mol (contributing ~2.7 × 10(3) to the rate enhancement) [Chan, K. K., Wood, B. M., Fedorov, A. A., Fedorov, E. V., Imker, H. J., Amyes, T. L., Richard, J. P., Almo, S. C., and Gerlt, J. A. (2009) Biochemistry 48, 5518-5531]. We now report that substitutions of hydrophobic residues in a pocket proximal to the carboxylate group of the substrate (Ile 96, Leu 123, and Val 155) with neutral hydrophilic residues decrease the value of k(cat) by as much as 400-fold but have a minimal effect on the value of k(ex) for exchange of H6 of the FUMP product analogue with solvent deuterium; we hypothesize that this pocket destabilizes the substrate by preventing hydration of the substrate carboxylate group. We also report that substitutions of Ser 127 that is proximal to O4 of the orotate ring decrease the value of k(cat)/K(M), with the S127P substitution that eliminates hydrogen bonding interactions with O4 producing a 2.5 × 10(6)-fold reduction; this effect is consistent with delocalization of the negative charge of the carbanionic intermediate on O4 that produces an anionic carbene intermediate and thereby provides a structural strategy for stabilization of the intermediate. These observations provide additional information about the identities of the active site residues that contribute to the rate enhancement and, therefore, insights into the structural strategies for catalysis.  相似文献   

13.
Chan KK  Fedorov AA  Fedorov EV  Almo SC  Gerlt JA 《Biochemistry》2008,47(36):9608-9617
Enzymes that share the (beta/alpha) 8-barrel fold catalyze a diverse range of reactions. Many utilize phosphorylated substrates and share a conserved C-terminal (beta/alpha) 2-quarter barrel subdomain that provides a binding motif for the dianionic phosphate group. We recently reported functional and structural studies of d-ribulose 5-phosphate 3-epimerase (RPE) from Streptococcus pyogenes that catalyzes the equilibration of the pentulose 5-phosphates d-ribulose 5-phosphate and d-xylulose 5-phosphate in the pentose phosphate pathway [J. Akana, A. A. Fedorov, E. Fedorov, W. R. P. Novack, P. C. Babbitt, S. C. Almo, and J. A. Gerlt (2006) Biochemistry 45, 2493-2503]. We now report functional and structural studies of d-allulose 6-phosphate 3-epimerase (ALSE) from Escherichia coli K-12 that catalyzes the equilibration of the hexulose 6-phosphates d-allulose 6-phosphate and d-fructose 6-phosphate in a catabolic pathway for d-allose. ALSE and RPE prefer their physiological substrates but are promiscuous for each other's substrate. The active sites (RPE complexed with d-xylitol 5-phosphate and ALSE complexed with d-glucitol 6-phosphate) are superimposable (as expected from their 39% sequence identity), with the exception of the phosphate binding motif. The loop following the eighth beta-strand in ALSE is one residue longer than the homologous loop in RPE, so the binding site for the hexulose 6-phosphate substrate/product in ALSE is elongated relative to that for the pentulose 5-phosphate substrate/product in RPE. We constructed three single-residue deletion mutants of the loop in ALSE, DeltaT196, DeltaS197 and DeltaG198, to investigate the structural bases for the differing substrate specificities; for each, the promiscuity is altered so that d-ribulose 5-phosphate is the preferred substrate. The changes in k cat/ K m are dominated by changes in k cat, suggesting that substrate discrimination results from differential transition state stabilization. In both ALSE and RPE, the phosphate group hydrogen bonds not only with the conserved motif but also with an active site loop following the sixth beta-strand, providing a potential structural mechanism for coupling substrate binding with catalysis.  相似文献   

14.
St Maurice M  Bearne SL 《Biochemistry》2000,39(44):13324-13335
Mandelate racemase (MR) catalyzes the interconversion of the enantiomers of mandelic acid, stabilizing the altered substrate in the transition state by 26 kcal/mol relative to the substrate in the ground state. To understand the origins of this binding discrimination, carboxylate-, phosphonate-, and hydroxamate-containing substrate and intermediate analogues were examined for their ability to inhibit MR. Comparison of the competitive inhibition constants revealed that an alpha-hydroxyl function is required for recognition of the ligand as an intermediate analogue. Two intermediate analogues, alpha-hydroxybenzylphosphonate (alpha-HBP) and benzohydroxamate, were bound with affinities approximately 100-fold greater than that observed for the substrate. Furthermore, MR bound alpha-HBP enantioselectively, displaying a 35-fold higher affinity for the (S)-enantiomer relative to the (R)-enantiomer. In the X-ray structure of mandelate racemase [Landro, J. A., Gerlt, J. A., Kozarich, J. W., Koo, C. W., Shah, V. J., Kenyon, G. L., Neidhart, D. J., Fujita, J., and Petsko, G. A. (1994) Biochemistry 33, 635-643], the alpha-hydroxyl function of the competitive inhibitor (S)-atrolactate is within hydrogen bonding distance of Asn 197. To demonstrate the importance of the alpha-hydroxyl function in intermediate binding, the N197A mutant was constructed. The values of k(cat) for N197A were reduced 30-fold for (R)-mandelate and 179-fold for (S)-mandelate relative to wild-type MR; the values of k(cat)/K(m) were reduced 208-fold for (R)-mandelate and 556-fold for (S)-mandelate. N197A shows only a 3.5-fold reduction in its affinity for the substrate analogue (R)-atrolactate but a 51- and 18-fold reduction in affinity for alpha-HBP and benzohydroxamate, respectively. Thus, interaction between Asn 197 and the substrate's alpha-hydroxyl function provides approximately 3.5 kcal/mol of transition-state stabilization free energy to differentially stabilize the transition state relative to the ground state.  相似文献   

15.
o-Succinylbenzoate synthase (OSBS) from Escherichia coli, a member of the enolase superfamily, catalyzes an exergonic dehydration reaction in the menaquinone biosynthetic pathway in which 2-succinyl-6-hydroxy-2,4-cyclohexadiene-1-carboxylate (SHCHC) is converted to 4-(2'-carboxyphenyl)-4-oxobutyrate (o-succinylbenzoate or OSB). Our previous structural studies of the Mg(2+).OSB complex established that OSBS is a member of the muconate lactonizing enzyme subgroup of the superfamily: the essential Mg(2+) is coordinated to carboxylate ligands at the ends of the third, fourth, and fifth beta-strands of the (beta/alpha)(7)beta-barrel catalytic domain, and the OSB product is located between the Lys 133 at the end of the second beta-strand and the Lys 235 at the end of the sixth beta-strand [Thompson, T. B., Garrett, J. B., Taylor, E. A, Meganathan, R., Gerlt, J. A., and Rayment, I. (2000) Biochemistry 39, 10662-76]. Both Lys 133 and Lys 235 were separately replaced with Ala, Ser, and Arg residues; all six mutants displayed no detectable catalytic activity. The structure of the Mg(2+).SHCHC complex of the K133R mutant has been solved at 1.62 A resolution by molecular replacement starting from the structure of the Mg(2+).OSB complex. This establishes the absolute configuration of SHCHC: the C1-carboxylate and the C6-OH leaving group are in a trans orientation, requiring that the dehydration proceed via a syn stereochemical course. The side chain of Arg 133 is pointed out of the active site so that it cannot function as a general base, whereas in the wild-type enzyme complexed with Mg(2+).OSB, the side chain of Lys 133 is appropriately positioned to function as the only acid/base catalyst in the syn dehydration. The epsilon-ammonium group of Lys 235 forms a cation-pi interaction with the cyclohexadienyl moiety of SHCHC, suggesting that Lys 235 also stabilizes the enediolate anion intermediate in the syn dehydration via a similar interaction.  相似文献   

16.
17.
E I Hyde  B R Reid 《Biochemistry》1985,24(16):4315-4325
The effects of magnesium, spermine, and temperature on the conformation of Escherichia coli tRNAPhe have been examined by proton and phosphorus nuclear magnetic resonance spectroscopy. In the low-field proton NMR spectra we have characterized two slowly interconverting conformations of this tRNA at low magnesium ion concentrations. The relative proportion of the conformers is ion dependent but not ion specific. Magnesium affects protons in all the stems of tRNA while spermine effects are localized near the s4U-8-A-14 and G-15-C-48 tertiary bonds. The effects seen in the proton NMR spectra are compared and correlated with those observed in the phosphorus spectra to give assignments of some of the resolved signals from the phosphate groups. The phosphorus spectra are compared with those of yeast tRNAPhe [Gorenstein, D. G., Goldfield, E. M., Chen, R., Kovar, K., & Luxon, B. A. (1981) Biochemistry 20, 2141; Salemink, P. J. M., Reijerse, E. J., Mollevanger, L., & Hilbers, C. W. (1981) Eur. J. Biochem. 115, 635], and the ion effects are discussed with reference to the magnesium and spermine sites found in the crystal structures of yeast tRNAPhe [Holbrook, S. R., Sussman, J. L., Warrant, R. W., Church, G. M., & Kim, S.-H. (1977) Nucleic Acids Res. 4, 2811; Quigley, G. J., Teeter, M. M., & Rich, A. (1978) Proc. Natl. Acad. Sci. U.S.A. 75, 64; Jack, A., Ladner, J. E., Rhodes, D., Brown, R. S., & Klug, A. (1977) J. Mol. Biol. 111, 315].  相似文献   

18.
The genes that encode the five known enzymes of the mandelate pathway of Pseudomonas putida (ATCC 12633), mandelate racemase (mdlA), (S)-mandelate dehydrogenase (mdlB), benzoylformate decarboxylase (mdlC), NAD(+)-dependent benzaldehyde dehydrogenase (mdlD), and NADP(+)-dependent benzaldehyde dehydrogenase (mdlE), have been cloned. The genes for (S)-mandelate dehydrogenase and benzoylformate decarboxylase have been sequenced; these genes and that for mandelate racemase [Ransom, S. C., Gerlt, J. A., Powers, V. M., & Kenyon, G. L. (1988) Biochemistry 27, 540] are organized in an operon (mdlCBA). Mandelate racemase has regions of sequence similarity to muconate lactonizing enzymes I and II from P. putida. (S)-Mandelate dehydrogenase is predicted to be 393 amino acids in length and to have a molecular weight of 43,352; it has regions of sequence similarity to glycolate oxidase from spinach and ferricytochrome b2 lactate dehydrogenase from yeast. Benzoylformate decarboxylase is predicted to be 499 amino acids in length and to have a molecular weight of 53,621; it has regions of sequence similarity to enzymes that decarboxylate pyruvate with thiamin pyrophosphate as cofactor. These observations support the hypothesis that the mandelate pathway evolved by recruitment of enzymes from preexisting metabolic pathways. The gene for benzoylformate decarboxylase has been expressed in Escherichia coli with the trc promoter, and homogeneous enzyme has been isolated from induced cells.  相似文献   

19.
To evaluate their role in the active site of the UDP-N-acetylmuramoyl-L-alanine:D-glutamate ligase (MurD) from Escherichia coli, 12 residues conserved either in the Mur superfamily [Eveland, S. S., Pompliano, D. L., and Anderson, M. S. (1997) Biochemistry 36, 6223-6229; Bouhss, A., Mengin-Lecreulx, D., Blanot, D., van Heijenoort, J., and Parquet, C. (1997) Biochemistry 36, 11556-11563] or in the sequences of 26 MurD orthologs were submitted to site-directed mutagenesis. All these residues lay within the cleft of the active site of MurD as defined by its 3D structure [Bertrand, J. A., Auger, D., Fanchon, E., Martin, L., Blanot, D., van Heijenoort, J., and Dideberg, O. (1997) EMBO J. 16, 3416-3425]. Fourteen mutant proteins (D35A, K115A, E157A/K, H183A, Y194F, K198A/F, N268A, N271A, H301A, R302A, D317A, and R425A) containing a C-terminal (His)(6) extension were prepared and their steady-state kinetic parameters determined. All had a reduced enzymatic activity, which in many cases was very low, but no mutation led to a total loss of activity. Examination of the specificity constants k(cat)/K(m) for the three MurD substrates indicated that most mutations affected both the binding of one substrate and the catalytic process. These kinetic results correlated with the assigned function of the residues based on the X-ray structures.  相似文献   

20.
The evolution of biosynthetic pathways is difficult to reconstruct in hindsight; however, the structures of the enzymes that are involved may provide insight into their development. One enzyme in the cobalamin biosynthetic pathway that appears to have evolved from a protein with different function is L-threonine-O-3-phosphate decarboxylase (CobD) from Salmonella enterica, which is structurally similar to histidinol phosphate aminotransferase [Cheong, C. G., Bauer, C. B., Brushaber, K. R., Escalante-Semerena, J. C., and Rayment, I. (2002) Biochemistry 41, 4798-4808]. This enzyme is responsible for synthesizing (R)-1-amino-2-propanol phosphate which is the precursor for the linkage between the nucleotide loop and the corrin ring in cobalamin. To understand the relationship between this decarboxylase and the aspartate aminotransferase family to which it belongs, the structures of CobD in its apo state, the apo state complexed with the substrate, and its product external aldimine complex have been determined at 1.46, 1.8, and 1.8 A resolution, respectively. These structures show that the enzyme steers the breakdown of the external aldimine toward decarboxylation instead of amino transfer by positioning the carboxylate moiety of the substrate out of the plane of the pyridoxal ring and by placing the alpha-hydrogen out of reach of the catalytic base provided by the lysine that forms the internal aldimine. It would appear that CobD evolved from a primordial PLP-dependent aminotransferase, where the selection was based on similarities between the stereochemical properties of the substrates rather than preservation of the fate of the external aldimine. These structures provide a sequence signature for distinguishing between L-threonine-O-3-phosphate decarboxylase and histidinol phosphate aminotransferases, many of which appear to have been misannotated.  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