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 703 毫秒
1.
Chung HW  Tu SC 《Biochemistry》2012,51(24):4880-4887
Vibrio harveyi NADPH-FMN oxidoreductase (FRP) catalyzes flavin reduction by NADPH. In comparing amino acid sequence and crystal structure with Escherichia coli NfsA, residues N134, R225, R133, K167, and R15 were targeted for investigation of their possible roles in the binding and utilization of the NADPH substrate. By mutation of each of these five residues to an alanine, steady-state rate analyses showed that the variants K167A and R15A had apparently greatly increased K(m,NADPH) and reduced k(cat)/K(m,NADPH), whereas little or much more modest changes were found for the other variants. The deuterium isotope effects (D)(V/K) for (4R)-[4-(2)H]-NADPH were markedly increased to 6.3 and 7.4 for K167A and R15A, respectively, indicating that the rate constants for NADPH and NADP(+) dissociation were greatly enhanced relative to the hydride transfer steps. Also, anaerobic stopped-flow analyses revealed that the equilibrium dissociation constant for NADPH binding (K(d)) to be 2.5-3.9 and 1.1 mM for K167A and R15A, respectively, much higher than the 0.4 μM K(d) for the native FRP, whereas the k(cat) of these two variants were similar to that of the wild-type enzyme. Moreover, the K167 to alanine mutation led to even a slight increase in k(cat)/K(m) for NADH. These results, taken together, provide a strong support to the conclusion that K167 and R15 each was critical in the binding of NADPH by FRP. Such a functional role may also exist for other FRP homologous proteins.  相似文献   

2.
The beta-subunit of the voltage-sensitive K(+) (K(v)) channels belongs to the aldo-keto reductase superfamily, and the crystal structure of K(v)beta2 shows NADP bound in its active site. Here we report that K(v)beta2 displays a high affinity for NADPH (K(d) = 0.1 micrometer) and NADP(+) (K(d) = 0.3 micrometer), as determined by fluorometric titrations of the recombinant protein. The K(v)beta2 also bound NAD(H) but with 10-fold lower affinity. The site-directed mutants R264E and N333W did not bind NADPH, whereas, the K(d)(NADPH) of Q214R was 10-fold greater than the wild-type protein. The K(d)(NADPH) was unaffected by the R189M, W243Y, W243A, or Y255F mutation. The tetrameric structure of the wild-type protein was retained by the R264E mutant, indicating that NADPH binding is not a prerequisite for multimer formation. A C248S mutation caused a 5-fold decrease in K(d)(NADPH), shifted the pK(a) of K(d)(NADPH) from 6.9 to 7.4, and decreased the ionic strength dependence of NADPH binding. These results indicate that Arg-264 and Asn-333 are critical for coenzyme binding, which is regulated in part by Cys-248. The binding of both NADP(H) and NAD(H) to the protein suggests that several types of K(v)beta2-nucleotide complexes may be formed in vivo.  相似文献   

3.
Mouse thymidylate synthase R209K (a mutation corresponding to R218K in Lactobacillus casei), overexpressed in thymidylate synthase-deficient Escherichia coli strain, was poorly soluble and with only feeble enzyme activity. The mutated protein, incubated with FdUMP and N(5,10)-methylenetetrahydrofolate, did not form a complex stable under conditions of SDS/polyacrylamide gel electrophoresis. The reaction catalyzed by the R209K enzyme (studied in a crude extract), compared to that catalyzed by purified wild-type recombinant mouse thymidylate synthase, showed the K(m) value for dUMP 571-fold higher and V(max) value over 50-fold (assuming that the mutated enzyme constituted 20% of total crude extract protein) lower. Thus the ratios k(cat, R209K)/k(cat, 'wild') and (k(cat, R209K)/K(m, R209K)(dUMP))/( k(cat, 'wild')/K(m, 'wild')(dUMP)) were 0.019 and 0.000032, respectively, documenting that mouse thymidylate synthase R209, similar to the corresponding L. casei R218, is essential for both dUMP binding and enzyme reaction.  相似文献   

4.
Sellers VM  Wu CK  Dailey TA  Dailey HA 《Biochemistry》2001,40(33):9821-9827
The terminal step in heme biosynthesis, the insertion of ferrous iron into protoporphyrin IX to form protoheme, is catalyzed by the enzyme ferrochelatase (EC 4.99.1.1). A number of highly conserved residues identified from the crystal structure of human ferrochelatase as being in the active site were examined by site-directed mutagenesis. The mutants Y123F, Y165F, Y191H, and R164L each had an increased K(m) for iron without an altered K(m) for porphyrin. The double mutant R164L/Y165F had a 6-fold increased K(m) for iron and a 10-fold decreased V(max). The double mutant Y123F/Y191F had low activity with an elevated K(m) for iron, and Y123F/Y165F had no measurable activity. The mutants H263A/C/N, D340N, E343Q, E343H, and E343K had no measurable enzyme activity, while E343D, E347Q, and H341C had decreased V(max)s without significant alteration of the K(m)s for either substrate. D340E had near-normal kinetic parameters, while D383A and H231A had increased K(m)s for iron. On the basis of these data and the crystal structure of human ferrochelatase, it is proposed that residues E343, H341, and D340 form a conduit from H263 in the active site to the protein exterior and function in proton extraction from the porphyrin macrocycle. The role of H263 as the porphyrin proton-accepting residue is central to catalysis since metalation only occurs in conjunction with proton abstraction. It is suggested that iron is transported from the exterior of the enzyme at D383/H231 via residues W227 and Y191 to the site of metalation at residues R164 and Y165 which are on the opposite side of the active site pocket from H263. This model should be general for mitochondrial membrane-associated eucaryotic ferrochelatases but may differ for bacterial ferrochelatases since the spatial orientation of the enzyme within prokaryotic cells may differ.  相似文献   

5.
NAD(P)H regeneration is important for biocatalytic reactions that require these costly cofactors. A mutant phosphite dehydrogenase (PTDH-E175A/A176R) that utilizes both NAD and NADP efficiently is a very promising system for NAD(P)H regeneration. In this work, both the kinetic mechanism and practical application of PTDH-E175A/A176R were investigated for better understanding of the enzyme and to provide a basis for future optimization. Kinetic isotope effect studies with PTDH-E175A/A176R showed that the hydride transfer step is (partially) rate determining with both NAD and NADP giving (D)V values of 2.2 and 1.7, respectively, and (D)V/K(m,phosphite) values of 1.9 and 1.7, respectively. To better comprehend the relaxed cofactor specificity, the cofactor dissociation constants were determined utilizing tryptophan intrinsic fluorescence quenching. The dissociation constants of NAD and NADP with PTDH-E175A/A176R were 53 and 1.9 microm, respectively, while those of the products NADH and NADPH were 17.4 and 1.22 microm, respectively. Using sulfite as a substrate mimic, the binding order was established, with the cofactor binding first and sulfite binding second. The low dissociation constant for the cofactor product NADPH combined with the reduced values for (D)V and k(cat) implies that product release may become partially rate determining. However, product inhibition does not prevent efficient in situ NADPH regeneration by PTDH-E175A/A176R in a model system in which xylose was converted into xylitol by NADP-dependent xylose reductase. The in situ regeneration proceeded at a rate approximately fourfold faster with PTDH-E175A/A176R than with either WT PTDH or a NADP-specific Pseudomonas sp.101 formate dehydrogenase mutant with a total turnover number for NADPH of 2500.  相似文献   

6.
Ruth A  Stein WD  Rose E  Roninson IB 《Biochemistry》2001,40(14):4332-4339
The MDR1 P-glycoprotein (Pgp), responsible for a clinically important form of multidrug resistance in cancer, is an ATPase efflux pump for multiple lipophilic drugs. The G185V mutation near transmembrane domain 3 of human Pgp increases its relative ability to transport several drugs, including etoposide, but decreases the transport of other substrates. MDR1 cDNA with the G185V substitution was used in a function-based selection to identify mutations that would further increase Pgp-mediated resistance to etoposide. This selection yielded the I186N substitution, adjacent to G185V. Pgps with G185V, I186N, or both mutations were compared to the wild-type Pgp for their ability to confer resistance to different drugs in NIH 3T3 cells. In contrast to the differential effects of G185V, I186N mutation increased resistance to all the tested drugs and augmented the effect of G185V on etoposide resistance. The effects of the mutations on conformational transitions of Pgp induced by different drugs were investigated using a conformation-sensitive antibody UIC2. Ligand-binding analysis of the drug-induced increase in UIC2 reactivity was used to determine the K(m) value that reflects the apparent affinity of drugs for Pgp, and the Hill number reflecting the apparent number of drug-binding sites. Both mutations altered the magnitude of drug-induced increases in UIC2 immunoreactivity, the K(m) values, and the Hill numbers for individual drugs. Mutation-induced changes in the magnitude of UIC2 reactivity shift did not correlate with the effects of the mutations on resistance to the corresponding drugs. In contrast, an increase or a decrease in drug resistance relative to that of the wild type was accompanied by a corresponding increase or decrease in the K(m) or in both the K(m) and the Hill number. These results suggest that mutations that alter the ability of Pgp to transport individual drugs change the apparent affinity and the apparent number of drug-binding sites in Pgp.  相似文献   

7.
Vagin O  Denevich S  Munson K  Sachs G 《Biochemistry》2002,41(42):12755-12762
Inhibition of the gastric H,K-ATPase by the imidazo[1,2-alpha]pyridine, SCH28080, is strictly competitive with respect to K+ or its surrogate, NH4+. The inhibitory kinetics [V(max), K(m,app)(NH4+), K(i)(SCH28080), and competitive, mixed, or noncompetitive] of mutants can define the inhibitor binding domain and the route to the ion binding region within M4-6. While mutations Y799F, Y802F, I803L, S806N, V807I (M5), L811V (M5-6), Y928H (M8), and Q905N (M7-8) had no effect on inhibitor kinetics, mutations P798C, Y802L, P810A, P810G, C813A or -S, I814V or -F, F818C, T823V (M5, M5-6, and M6), E914Q, F917Y, G918E, T929L, and F932L (M7-8 and M8) reduced the affinity for SCH28080 up to 10-fold without affecting the nature of the kinetics. In contrast, the L809F substitution in the loop between M5 and M6 resulted in an approximately 100-fold decrease in inhibitor affinity, and substitutions L809V, I816L, Y925F, and M937V (M5-6, M6, and M8) reduced the inhibitor affinity by 10-fold, all resulting in noncompetitive kinetics. The mutants L811F, Y922I, and I940A also reduced the inhibitor affinity up to 10-fold but resulted in mixed inhibition. The mutations I819L, Q923V, and Y925A also gave mixed inhibition but without a change in inhibitor affinity. These data, and the 9-fold loss of SCH28080 affinity in the C813T mutant, suggest that the binding domain for SCH28080 contains the surface between L809 in the M5-6 loop and C813 at the luminal end of M6, approximately two helical turns down from the ion binding region, where it blocks the normal ion access pathway. On the basis of a model of the Ca-ATPase in the E2 conformation (PDB entry 1kju), the mutants that change the nature of the kinetics are arranged on one side of M8 and on the adjacent side of the M5-6 loop and M6 itself. This suggests that mutations in this region modify the enzyme structure so that K+ can access the ion binding domain even with SCH28080 bound.  相似文献   

8.
Wang H  Lei B  Tu SC 《Biochemistry》2000,39(26):7813-7819
Luminous bacteria contain three types of NAD(P)H-FMN oxidoreductases (flavin reductases) with different pyridine nucleotide specificities. Among them, the NADPH-specific flavin reductase from Vibrio harveyi exhibits a uniquely high preference for NADPH. In comparing the substrate specificity, crystal structure, and primary sequence of this flavin reductase with other structurally related proteins, we hypothesize that the conserved Arg203 residue of this reductase is critical to the specific recognition of NADPH. The mutation of this residue to an alanine resulted in only small changes in the binding and reduction potential of the FMN cofactor, the K(m) for the FMN substrate, and the k(cat). In contrast, the K(m) for NADPH was increased 36-fold by such a mutation. The characteristic perturbation of the FMN cofactor absorption spectrum upon NADP(+) binding by the wild-type reductase was abolished by the same mutation. While the k(cat)/K(m,NADPH) was reduced from 1990 x 10(5) to 46 x 10(5) M(-1) min(-1) by the mutation, the mutated variant showed a k(cat)/K(m,NADH) of 4 x 10(5) M(-1) min(-1), closely resembling that of the wild-type reductase. The deuterium isotope effects (D)V and (D)(V/K) for (4R)-[4-(2)H]-NADPH were 1.7 and 1.4, respectively, for the wild-type reductase but were increased to 3.8 and 4.0, respectively, for the mutated variant. Such a finding indicates that the rates of NADPH and NADP(+) dissociation in relation to the isotope-sensitive redox steps were both increased as a result of the mutation. These results all provide support to the critical role of the Arg203 in the specific recognition and binding of NADPH.  相似文献   

9.
Glucose-6-phosphate dehydrogenase (G6PD; EC 1.1.1.49) is the key regulatory enzyme of the pentose phosphate pathway and produces NADPH and riboses. In this study, the kinetic properties of G6PD activity were determined in situ in chemically induced hepatocellular carcinomas, and extralesional and control parenchyma in rat livers and were directly compared with those of the second NADPH-producing enzyme of the pentose phosphate pathway, phosphogluconate dehydrogenase (PGD). Distribution patterns of G6PD activity, protein, and mRNA levels were also compared to establish the regulation mechanisms of G6PD activity. In (pre)neoplastic lesions, the V(max) of G6PD was 150-fold higher and the K(m) for G6P was 10-fold higher than in control liver parenchyma, whereas in extralesional parenchyma, the V(max) was similar to that in normal parenchyma but the K(m) was fivefold lower. This means that virtual fluxes at physiological substrate concentrations are 20-fold higher in lesions and twofold higher in extralesional parenchyma than in normal parenchyma. The V(max) of PGD was fivefold higher in lesions than in normal and extralesional liver parenchyma, whereas the K(m) was not affected. Amounts of G6PD protein and mRNA were similar in lesions and in extralesional liver parenchyma. These results demonstrate that G6PD is strongly activated post-translationally in (pre)neoplastic lesions to produce NADPH.  相似文献   

10.
Nitrilases from Aspergillus niger CBS 513.88, A. niger K10, Gibberella moniliformis, Neurospora crassa OR74A, and Penicillium marneffei ATCC 18224 were expressed in Escherichia coli BL21-Gold (DE3) after IPTG induction. N. crassa nitrilase exhibited the highest yield of 69,000 U L(-1) culture. Co-expression of chaperones (GroEL/ES in G. moniliformis and P. marneffei; GroEL/ES and trigger factor in N. crassa and A. niger CBS 513.88) enhanced the enzyme solubility. Specific activities of strains expressing the former two enzymes increased approximately fourfold upon co-expression of GroEL/ES. The enzyme from G. moniliformis (co-purified with GroEL) preferred benzonitrile as substrate (K(m) of 0.41 mM, V(max) of 9.7 μmol min(-1) mg(-1) protein). The P. marneffei enzyme (unstable in its purified state) exhibited the highest V(max) of 7.3 μmol min(-1) mg(-1) protein in cell-free extract, but also a high K(m) of 15.4 mM, for 4-cyanopyridine. The purified nitrilases from A. niger CBS 513.88 and N. crassa acted preferentially on phenylacetonitrile (K(m) of 3.4 and 2.0 mM, respectively; V(max) of 10.6 and 17.5 μmol min(-1) mg(-1) protein, respectively), and hydrolyzed also (R,S)-mandelonitrile with higher K(m) values. Significant amounts of amides were only formed by the G. moniliformis nitrilase from phenylacetonitrile and 4-cyanopyridine.  相似文献   

11.
The complete amino acid sequence of coagulogen purified from the hemocytes of the horseshoe crab Carcinoscorpius rotundicauda was determined by characterization of the NH2-terminal sequence and the peptides generated after digestion of the protein with lysyl endopeptidase, Staphylococcal aureus protease V8 and trypsin. Upon sequencing the peptides by the automated Edman method, the following sequence was obtained: A D T N A P L C L C D E P G I L G R N Q L V T P E V K E K I E K A V E A V A E E S G V S G R G F S L F S H H P V F R E C G K Y E C R T V R P E H T R C Y N F P P F V H F T S E C P V S T R D C E P V F G Y T V A G E F R V I V Q A P R A G F R Q C V W Q H K C R Y G S N N C G F S G R C T Q Q R S V V R L V T Y N L E K D G F L C E S F R T C C G C P C R N Y Carcinoscorpius coagulogen consists of a single polypeptide chain with a total of 175 amino acid residues and a calculated molecular weight of 19,675. The secondary structure calculated by the method of Chou and Fasman reveals the presence of an alpha-helix region in the peptide C segment (residue Nos. 19 to 46), which is released during the proteolytic conversion of coagulogen to coagulin gel. The beta-sheet structure and the 16 half-cystines found in the molecule appear to yield a compact protein stable to acid and heat. The amino acid sequences of coagulogen of four species of limulus have been compared and the interspecies evolutionary differences are discussed.  相似文献   

12.
The highly conserved motif +(534)DPPR of Saccharomyces cerevisiae H(+)-ATPase, located in the putative ATP binding site, has been mutagenized and the resulting 23 mutant genes conditionally expressed in secretory vesicles. Fourteen mutant ATPases (D534A, D534V, D534L, D534N, D534G, D534T, P535A, P535V, P535L, P535G, P535T, P535E, P535K and R537T) failed to reach the secretory vesicles. Of these mutants, nine (D534N, D534T, P535A, P535V, P535L, P535G, P535T, P535E and P535K) were not detected in total cellular membranes, and five (D534A, D534V, D534G, D534L and R537T) were retained at the endoplasmic reticulum and exhibited a dominant lethal phenotype. The remaining mutants (D534E, R537A, R537V, R537L, R537N, R537G, R537E, R537K and R537H) reached the secretory vesicles at levels similar to that of the wild type. Of these, six (R537A, R537V, R537L, R537N, R537G, and R537E) showed severely decreased ATPase activity compared to the wild type enzyme, and three (D534E, R537K and R537H) rendered an enzyme with an altered K(m) for ATP.  相似文献   

13.
Bovine papillomavirus type 1 (BPV-1) requires viral proteins E1 and E2 for efficient DNA replication in host cells. E1 functions at the BPV origin as an ATP-dependent helicase during replication initiation. Previously, we used alanine mutagenesis to identify two hydrophilic regions of the E1 DNA binding domain (E1DBD), HR1 (E1(179-191)) and HR3 (E1(241-252)), which are critical for sequence-specific recognition of the papillomavirus origin. Based on sequence and structure, these regions are similar in spacing and location to DNA binding regions A and B2 of T antigen, the DNA replication initiator of simian virus 40 (SV40). HR1 and A are both part of extended loops which are supported by residues from the HR3 and B2 alpha-helices. Both elements contain basic residues which may contact DNA, although lack of cocrystal structures for both E1 and T antigen make this uncertain. To better understand how E1 interacts with origin DNA, we used random mutagenesis and a yeast one-hybrid screen to select mutations of the E1DBD which disrupt sequence-specific DNA interactions. From the screen we selected seven single point mutants and one double point mutant (F175S, N184Y/K288R, D185G, V193M, F237L, K241E, R243K, and V246D) for in vitro analysis. All mutants tested in electrophoretic mobility shift assays displayed reduced sequence-specific DNA binding compared to the wild-type E1DBD. Mutants D185G, F237L, and R243K were rescued in vitro for DNA binding by the replication enhancer protein E2. We also tested the eight mutations in full-length E1 for the ability to support DNA replication in Chinese hamster ovary cells. Only mutants D185G, F237L, and R243K supported significant DNA replication in vivo which highlights the importance of E1DBD-E2 interactions for papillomavirus DNA replication. Based on the specific point mutations examined, we also assigned putative roles to individual residues in DNA binding. Finally, we discuss sequence and spacing similarities between E1 HR1 and HR3 and short regions of two other DNA tumor virus origin-binding proteins, SV40 T antigen and Epstein-Barr virus nuclear antigen 1 (EBNA1). We propose that all three proteins use a similar DNA recognition mechanism consisting of a loop structure which makes base-specific contacts (HR1) and a helix which primarily contacts the DNA backbone (HR3).  相似文献   

14.
Lysine decarboxylase (LDC; EC 4.1.1.18) from Selenomonas ruminantium comprises two identical monomeric subunits of 43 kDa and has decarboxylating activities toward both L-lysine and L-ornithine with similar K(m) and V(max) values (Y. Takatsuka, M. Onoda, T. Sugiyama, K. Muramoto, T. Tomita, and Y. Kamio, Biosci. Biotechnol. Biochem. 62:1063-1069, 1999). Here, the LDC-encoding gene (ldc) of this bacterium was cloned and characterized. DNA sequencing analysis revealed that the amino acid sequence of S. ruminantium LDC is 35% identical to those of eukaryotic ornithine decarboxylases (ODCs; EC 4.1.1.17), including the mouse, Saccharomyces cerevisiae, Neurospora crassa, Trypanosoma brucei, and Caenorhabditis elegans enzymes. In addition, 26 amino acid residues, K69, D88, E94, D134, R154, K169, H197, D233, G235, G236, G237, F238, E274, G276, R277, Y278, K294, Y323, Y331, D332, C360, D361, D364, G387, Y389, and F397 (mouse ODC numbering), all of which are implicated in the formation of the pyridoxal phosphate-binding domain and the substrate-binding domain and in dimer stabilization with the eukaryotic ODCs, were also conserved in S. ruminantium LDC. Computer analysis of the putative secondary structure of S. ruminantium LDC showed that it is approximately 70% identical to that of mouse ODC. We identified five amino acid residues, A44, G45, V46, P54, and S322, within the LDC catalytic domain that confer decarboxylase activities toward both L-lysine and L-ornithine with a substrate specificity ratio of 0.83 (defined as the k(cat)/K(m) ratio obtained with L-ornithine relative to that obtained with L-lysine). We have succeeded in converting S. ruminantium LDC to form with a substrate specificity ratio of 58 (70 times that of wild-type LDC) by constructing a mutant protein, A44V/G45T/V46P/P54D/S322A. In this study, we also showed that G350 is a crucial residue for stabilization of the dimer in S. ruminantium LDC.  相似文献   

15.
Short chain acyl-CoA dehydrogenase (SCAD) is a homotetrameric flavoenzyme that catalyzes the first intramitochondrial step in the beta-oxidation of fatty acids. Two polymorphisms in the coding region of the SCAD gene, 511C>T (R147W) and 625G>A (G185S), have been shown to be associated with an increased level of ethylmalonic acid excretion in urine, a clinical characteristic of SCAD deficiency. To characterize the biochemical consequences of these variations, in vitro site-directed mutagenesis and prokaryotic expression were used to produce the corresponding SCAD variant proteins. Both variant proteins were unstable when produced in Escherichia coli, but could be rescued and subsequently purified by coexpressing them with the bacterial chaperonin GroEL/ES. The k(cat)/K(m) values of the green wild-type, R147W, and G185S SCAD enzymes coexpressed with GroEL/ES were 33, 30, and 10 microM(-)(1) s(-)(1), respectively. There were minimal differences in the kinetic parameters measured for the green, degreened, and wild-type enzymes coexpressed with GroEL/ES, and the R147W variant when butyryl-CoA was used as a substrate. The catalytic efficiency of the G185S variant enzyme, however, was reduced compared to that of the wild-type enzyme. The thermal and guanidine HCl stability of the purified enzymes as determined by fluorescence, far-UV CD spectroscopy, and incubation-induced rest activity showed the following order of relative stability: wild-type enzyme > R147W > G185S. Near-UV CD spectroscopy indicated that these impairments are caused by decreased flexibility in the tertiary conformation of the two mutant enzymes. The common SCAD polymorphisms may lead to clinically relevant alterations in enzyme function.  相似文献   

16.
Omote H  Figler RA  Polar MK  Al-Shawi MK 《Biochemistry》2004,43(13):3917-3928
A glycine 185 to valine mutation of human P-glycoprotein (ABCB1, MDR1) has been previously isolated from high colchicine resistance cell lines. We have employed purified and reconstituted P-glycoproteins expressed in Saccharomyces cerevisiae [Figler et al. (2000) Arch. Biochem. Biophys. 376, 34-46] and devised a set of thermodynamic analyses to reveal the mechanism of improved resistance. Purified G185V enzyme shows altered basal ATPase activity but a strong stimulation of colchicine- and etoposide-dependent activities, suggesting a tight regulation of ATPase activity by these drugs. The mutant enzyme has a higher apparent K(m) for colchicine and a lower K(m) for etoposide than that of wild type. Kinetic constants for other transported drugs were not significantly modified by this mutation. Systematic thermodynamic analyses indicate that the G185V enzyme has modified thermodynamic properties of colchicine- and etoposide-dependent activities. To improve the rate of colchicine or etoposide transport, the G185V enzyme has lowered the Arrhenius activation energy of the transport rate-limiting step. The high transition state energies of wild-type P-glycoprotein, when transporting etoposide or colchicine, increase the probability of nonproductive degradation of the transition state without transport. G185V P-glycoprotein transports etoposide or colchicine in an energetically more efficient way with decreased enthalpic and entropic components of the activation energy. Our new data fully reconcile the apparently conflicting results of previous studies. EPR analysis of the spin-labeled G185C enzyme in a cysteine-less background and kinetic parameters of the G185C enzyme indicate that position 185 is surrounded by other residues and is volume sensitive. These results and atomic detail structural modeling suggest that residue 185 is a pivotal point in transmitting conformational changes between the catalytic sites and the colchicine drug binding domain. Replacement of this residue with a bulky valine alters this communication and results in more efficient transport of etoposide or colchicine.  相似文献   

17.
Xiang J  Jung JY  Sampson NS 《Biochemistry》2004,43(36):11436-11445
Many proteins utilize segmental motions to catalyze a specific reaction. The Omega loop of triosephosphate isomerase (TIM) is important for preventing the loss of the reactive enediol(ate) intermediate. The loop opens and closes even in the absence of the ligand, and the loop itself does not change conformation during movement. The conformational changes are localized to two hinges at the loop termini. Glycine is never observed in native TIM hinge sequences. In this paper, the hypothesis that limited access to conformational space is a requirement for protein hinges involved in catalysis was tested. The N-terminal hinge was mutated to P166/V167G/W168G (PGG), and the C-terminal hinge was mutated to K174G/T175G/A176G (GGG) in chicken TIM. The single-hinge mutants PGG and GGG had k(cat) values 200-fold lower than that of the wild type and K(m) values 10-fold higher. The k(cat) of double-hinge mutant P166/V167G/W168G/K174G/T175G/A176G was reduced 2500-fold; the K(m) was 10-fold higher. A combination of primary kinetic isotope effect measurements, isothermal calorimetric measurements, and (31)P NMR spectroscopic titration with the inhibitor 2-phosphoglycolate revealed that the mutants have a different ligand-binding mode than that of the wild-type enzyme. The predominant conformations of the mutants even in the presence of the inhibitor are loop-open conformations. In conclusion, mutation of the hinge residues to glycine resulted in the sampling of many more hinge conformations with the consequence that the population of the active-closed conformation is reduced. This reduced population results in a reduced catalytic activity.  相似文献   

18.
The most common mutation in the alpha subunit of the pyruvate dehydrogenase (E1) component of the human pyruvate dehydrogenase complex (PDC) is arginine-234 to glycine and glutamine in 12 and 3 patients, respectively. Interestingly, these two mutations at the same amino acid position cause E1 (and hence PDC) deficiency by apparently different mechanisms. Recombinant human R234Q E1 had similar V(max) (25.7 +/- 4.4 units/mg E1) and apparent K(m) (101 +/- 4 nM) values for TPP as recombinant wild-type human E1, while R234G E1 had no significant change in V(max) (33.6 +/- 4.7 units/mg E1) but had a 7-fold increase in its apparent K(m) value for TPP (497 +/- 25 nM). Both of the R234 mutant proteins had similar apparent K(m) values for pyruvate. Both R234Q and R234G mutant proteins displayed similar phosphorylation rates of sites 1 and 2 by pyruvate dehydrogenase kinase 2 (PDK2) and site 3 by PDK1 compared to wild-type E1. Phosphorylated R234Q E1, R234G E1, and wild-type E1 also had similar dephosphorylation rates of sites 1 and 2 by phosphopyruvate dehydrogenase phosphatase 1. The rate of dephosphorylation of site 3 was about 50% for R234Q E1 and without a significant change for R234G E1 compared to the wild type. The data indicate that the patients with the R234G E1 mutation are symptomatic due to a decreased ability of this mutant protein to bind TPP, whereas the patients with the R234Q E1 mutation are symptomatic due to a decreased rate of dephosphorylation of site 3, hence keeping the enzyme in a phosphorylated/inactivated form.  相似文献   

19.
20.
D A Toke  M L McClintick  G M Carman 《Biochemistry》1999,38(44):14606-14613
Diacylglycerol pyrophosphate (DGPP) phosphatase, encoded by the DPP1 gene, is a membrane-associated enzyme in the yeast Saccharomyces cerevisiae. The enzyme removes the beta phosphate from DGPP to form phosphatidate. The substrate and product of the DGPP phosphatase reaction play roles in lipid signaling and in cell metabolism. The deduced primary structure of the DGPP phosphatase protein contains a three-domain phosphatase sequence motif. In this work, we examined the hypothesis that the phosphatase sequence motif in the enzyme is involved in the DGPP phosphatase reaction. The amino acid residues Arg(125), His(169), and His(223) in domains 1, 2, and 3, respectively, of the phosphatase sequence motif were changed to alanine residues by site-directed mutagenesis. The mutant DPP1(R125A), DPP1(H169A), and DPP1(H223A) alleles were cloned into a yeast shuttle vector and then expressed in a dpp1Delta lpp1Delta double mutant that lacks DGPP phosphatase activity. Northern blot and immunoblot analyses showed that the mutations in the phosphatase sequence motif did not affect the expression of the enzyme. The DGPP phosphatase activities of the R125A, the H169A, and the H223A mutant enzymes were 0.05, 9, and 0.03%, respectively, of the DGPP phosphatase activity of the wild-type enzyme. Enzymes with mutations in more than one domain of the phosphatase sequence motif had no measurable DGPP phosphatase activity. The R125A and H233A mutant DGPP phosphatase enzymes had reduced V(max) and elevated K(m) values for DGPP when compared with the wild-type enzyme. The H169A mutant enzyme had reduced V(max) and K(m) values when compared with the control. The specificity constants (V(max)/K(m)()) for DGPP of the R125A mutant and H233A mutant enzymes were 4610-fold and 15 367-fold lower, respectively, when compared to the wild-type enzyme. The studies reported here indicated that the phosphatase sequence motif played an important role in the reaction catalyzed by the S. cerevisiae DGPP phosphatase.  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