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.
We performed a detailed bioinformatic study of the catalytic step of fructose-6-phosphate phosphorylation in glycolysis based on the raw genomic draft of Propionibacterium freudenreichii subsp. shermanii (P. shermanii) ATCC9614 [Meurice et al., 2004]. Our results provide the first in silico evidence of the coexistence of genes coding for an ATP-dependent phosphofructokinase (ATP-PFK) and a PPi-dependent phosphofructokinase (PPi-PFK), whereas the fructose-1,6-bisphosphatase (FBP) and ADP-dependent phosphofructokinase (ADP-PFK) are absent. The deduced amino acid sequence corresponding to the PPi-PFK (AJ508922) shares 100% similarity with the already characterised propionibacterial protein (P29495; Ladror et al., 1991]. The unexpected ATP-PFK gene (AJ509827) encodes a protein of 373 aa which is highly similar (50% positive residues) along at least 95% of its sequence length to different well-characterised ATP-PFKs. The characteristic PROSITE pattern important for the enzyme function of ATP-PFKs (PS00433) was conserved in the putative ATP-PFK sequence: 8 out of 9 amino acid residues. According to the recent evolutionary study of PFK proteins with different phosphate donors [Bapteste et al., 2003], the propionibacterial ATP-PFK harbours a G104-K124 residue combination, which strongly suggested that this enzyme belongs to the group of atypical ATP-PFKs. According to our phylogenetic analyses the amino acid sequence of the ATP-PFK is clustered with the atypical ATP-PFKs from group III of the Siebers classification [Siebers et al., 1998], whereas the expected PPi-PFK protein is closer to the PPi-PFKs from clade P [Müller et al., 2001]. The possible significance of the co-existence of these two PFKs and their importance for the regulation of glycolytic pathway flux in P. shermanii is discussed.  相似文献   

2.
According to the X-ray structure analysis of an EcoRI-oligodeoxynucleotide complex [McClarin et al. (1986) Science 234, 1526], sequence specificity is mediated by 12 hydrogen bonds, 6 from each of the two identical subunits of the dimeric enzyme to the recognition site -GAATTC-: Arg200 forms two hydrogen bonds with guanine, while Glu144 and Arg145 form four hydrogen bonds to adjacent adenine residues. Changing the hydrogen-bonding potential at the recognition site without perturbing the rest of the interface should lead to the recognition of degenerate sequences [Rosenberg et al. (1987) in Protein Engineering (Oxender, D. L., & Fox, C. F., Eds.) pp 237-250, Liss, New York]. We have shown previously that replacing Glu144 by Gln and Arg145 by Lys affects the activity of the enzyme, not, however, its specificity [Wolfes et al. (1986) Nucleic Acids Res. 14, 9063]. We show now that also the mutation of Arg200 to Lys, the double mutation Glu144Arg145 to GlnLys, and the triple mutation Glu144Arg145Arg200 to GlnLysLys do not lead to a detectable degeneracy of the specificity of cleavage by EcoRI but significantly impair the catalytic activity of this enzyme. A detailed analysis of the steady-state kinetics of cleavage of pUC8 DNA and a tridecadeoxynucleotide substrate demonstrates that the reduction in activity for all DNA binding site mutants investigated so far is mainly due to a decrease in kcat, with the exception of the Arg200 to Lys mutant, which is only impaired in its KM.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

3.
A Hussain  D Lewis  M Yu  P W Melera 《Gene》1992,112(2):179-188
Simian virus 40 promoter-enhancer-based mammalian expression plasmids using dihydrofolate reductase (DHFR)-encoding cDNA sequences originally isolated from two methotrexate (MTX)-resistant, DHFR-overproducing Chinese hamster lung cell lines were constructed. One, designated pSVA75, contains a DHFR cDNA that encodes leucine (Leu22) and corresponds to the wild type (wt), MTX-sensitive form of the enzyme [Melera et al., J. Biol. Chem. 263 (1988) 1978-1990]. The other plasmid, pSVA3, contains a cDNA that encodes a novel mutant form of the enzyme in which Leu22 has been changed to Phe [Melera et al., Mol. Cell Biol. 4 (1984) 38-48]. The resulting DHFR displays a 20-fold-enhanced resistance to inhibition by MTX, but maintains the catalytic activity of the wt enzyme [Albrecht et al., Cancer Res. 32 (1972) 1539-1546]. Transfection of DHFR- Chinese hamster ovary cells with either plasmid demonstrated that both were able to reconstitute the DHFR+ phenotype with equal efficiency (i.e., greater than 2.5 x 10(-3), indicating that both the wt and mutant enzymes were catalytically active in transfected cells. In addition, the mutant form of the enzyme was found to act as a dominant selectable marker when transfected into diploid DHFR+ cells, and to allow selection of resistant clones at low MTX concentrations (125 nM MTX) with a frequency of greater than 8 x 10(-4). Moreover, transfected clones were found to amplify their exogenous DHFR sequences to reasonably high levels (42-fold) at relatively low (888 nM) MTX concentrations, suggesting that substantial amplification of DHFR DNA and cotransfected sequences as well, can be achieved with this vector.  相似文献   

4.
Caveolins are scaffolding proteins able to collect on caveolae a large number of signalling proteins bearing a caveolin-binding motif. The proteins of the striatin family, striatin, SG2NA, and zinedin, are composed of several conserved, collinearly aligned, protein-protein association domains, among which a putative caveolin-binding domain [Castets et al. (2000) J. Biol. Chem. 275, 19970-19977]. They are associated in part with membranes. These proteins are mainly expressed within neurons and thought to act both as scaffolds and as Ca(2+)-dependent signalling proteins [Bartoli et al. (1999) J. Neurobiol. 40, 234-243]. Here, we show that (1) rat brain striatin, SG2NA and zinedin co-immunoprecipitate with caveolin-1; (2) all are pulled down by glutathione-S-transferase (GST)-caveolin-1; (3) a fragment of recombinant striatin containing the putative caveolin-binding domain binds GST-caveolin-1. Hence, it is likely that the proteins of the striatin family are addressed to membrane microdomains by their binding to caveolin, in accordance with their putative role in membrane trafficking [Baillat et al. (2001) Mol. Biol. Cell 12, 663-673].  相似文献   

5.
Li SX  Vaccaro JA  Sweasy JB 《Biochemistry》1999,38(15):4800-4808
DNA polymerase beta is a small monomeric polymerase that participates in base excision repair and meiosis [Sobol, R., et al. (1996) Nature 379, 183-186; Plug, A., et al. (1997) Proc. Natl. Acad. Sci. U.S.A. 94, 1327-1331]. A DNA polymerase beta mutator mutant, F272L, was identified by an in vivo genetic screen [Washington, S., et al. (1997) Proc. Natl. Acad. Sci. U.S.A. 94, 1321-1326]. Residue 272 is located within the deoxynucleoside triphosphate (dNTP) binding pocket of DNA polymerase beta according to the known DNA polymerase beta crystal structures [Pelletier, H., et al. (1994) Science 264, 1891-1893; Sawaya, M., et al. (1997) Biochemistry 36, 11205-11215]. The F272L mutant produces errors at a frequency 10-fold higher than that of wild type in vivo and in the in vitro HSV-tk gap-filling assay. F272L shows an increase in the frequency of both base substitution mutations and frameshift mutations. Single-enzyme turnover studies of misincorporation by wild type and F272L DNA polymerase beta demonstrate that there is a 4-fold decrease in fidelity of the mutant as compared to that of the wild type enzyme for a G:A mismatch. The decreased fidelity is due primarily to decreased discrimination between the correct and incorrect dNTP during ground-state binding. These results suggest that the phenylalanine 272 residue is critical for maintaining fidelity during the binding of the dNTP.  相似文献   

6.
Joseph E  Wei W  Tittmann K  Jordan F 《Biochemistry》2006,45(45):13517-13527
The X-ray crystal structure of pyruvamide-activated yeast pyruvate decarboxylase (YPDC) revealed a flexible loop spanning residues 290 to 304 on the beta-domain of the enzyme, not seen in the absence of pyruvamide, a substrate activator surrogate. Site-directed mutagenesis studies revealed that residues on the loop affect the activity, with some residues reducing k(cat)/K(m) by at least 1000-fold. In the pyruvamide-activated form, the loop located on the beta domain can transfer information to the active center thiamin diphosphate (ThDP) located at the interface of the alpha and gamma domains. The sigmoidal v(0)-[S] curve with wild-type YPDC attributed to substrate activation is modulated for most variants, but is not abolished. Pre-steady-state stopped-flow studies for product formation on these loop variants provided evidence for three enzyme conformations connected by two transitions, as already noted for the wild-type YPDC at pH 5.0 [Sergienko, E. A., and Jordan, F. (2002) Biochemistry 41, 3952-3967]. (1)H NMR analysis of the intermediate distribution resulting from acid quench [Tittmann et al. (2003) Biochemistry 42, 7885-7891] with all YPDC variants indicated that product release is rate limiting in the steady state. Apparently, the loop is not solely responsible for the substrate activation behavior, rather it may affect the behavior of residue C221 identified as the trigger for substrate activation. The most important function of the loop is to control the conformational equilibrium between the "open" and "closed" conformations of the enzyme identified in the pyruvamide-activated structure [Lu et al. (2000) Eur. J. Biochem. 267, 861-868].  相似文献   

7.
A deoxyribonuclease I was purified from the urine of a 46-year-old male (a single individual) by using a series of column chromatographies to a homogeneous state as judged by sodium dodecyl sulfate-polyacrylamide gel electrophoresis. The enzyme was found to be a glycoprotein, containing 1 fucose, 7 galactose, 10 mannose, 6 glucosamine, and 2 sialic acid residues per molecule. The N-terminal amino acid sequence up to the 27th residue of the enzyme was similar to that of pancreatic deoxyribonuclease I from bovine and other species. The catalytic properties of the enzyme derived from a single individual closely resembled those of deoxyribonuclease I purified from human urine collected from several volunteers [Ito, K. et al. (1984) J. Biochem. 95, 1399-1406]. The purified enzyme was found to consist of multiple forms with different pI values. These findings are compatible with the existence of genetic polymorphism of deoxyribonuclease I in human urine previously reported [Kishi, K. et al. (1989) Hum. Genet. 81, 295-297]. This multiplicity of the urine enzyme might be due to variations in the primary structure and/or differences in the content of sialic acid.  相似文献   

8.
Böhmer C  Wehner F 《FEBS letters》2001,494(1-2):125-128
The epithelial Na(+) channel (ENaC) is composed of the subunits alpha, beta, and gamma [Canessa et al., Nature 367 (1994) 463-467] and typically exhibits a high affinity to amiloride [Canessa et al., Nature 361 (1993) 467-470]. When expressed in Xenopus oocytes, conflicting results were reported concerning the osmo-sensitivity of the channel [Ji et al., Am. J. Physiol. 275 (1998) C1182-C1190; Hawayda and Subramanyam, J. Gen. Physiol. 112 (1998) 97-111; Rossier, J. Gen. Physiol. 112 (1998) 95-96]. Rat hepatocytes were the first system in which amiloride-sensitive sodium currents in response to hypertonic stress were reported [Wehner et al., J. Gen. Physiol. 105 (1995) 507-535; Wehner et al., Physiologist 40 (1997) A-4]. Moreover, all three ENaC subunits are expressed in these cells [B?hmer et al., Cell. Physiol. Biochem. 10 (2000) 187-194]. Here, we injected specific antisense oligonucleotides directed against alpha-rENaC into single rat hepatocytes in confluent primary culture and found an inhibition of hypertonicity-induced Na(+) currents by 70%. This is the first direct evidence for a role of the ENaC in cell volume regulation.  相似文献   

9.
Fatty acid synthase: A metabolic oncogene in prostate cancer?   总被引:13,自引:0,他引:13  
In 1920, Warburg suggested that tumors consistently rely on anaerobic pathways to convert glucose to ATP even in the presence of abundant oxygen [Warberg, 1956] despite the fact that it is less efficient for energy supply than aerobic glycolysis. The reasons for this remain obscure to date. More often than not, the microenvironment of solid tumors contains regions of poor oxygenation and high acidity. In this context hypoxia can act in an epigenetic fashion, inducing changes in gene expression and in metabolism for survival. It is reasonable to assume that only the tumor cells capable of developing an unusual tolerance to limiting oxygen availability and to the acidosis resulting from excessive lactate production, can survive. In addition to the striking changes that occur in glucose metabolism, studies in human cancer patients suggest that there is often also an increase in free fatty acid turnover, oxidation and clearance [Legaspi et al., 1987; Hyltander et al., 1991]. For instance, a lipid mobilizing factor produced by tumor cells appears to be responsible for the increase in whole body fatty acid oxidation [Russell and Tisdale, 2002]. Fatty acids synthesis in tumor tissues also occurs at very high rates, as first demonstrated more than half a century ago [Medes et al., 1953]. Importantly, (14)C glucose studies have shown that in tumor cells almost all fatty acids derive from de novo synthesis despite adequate nutritional supply [Sabine and Abraham, 1967; Ookhtens et al., 1984; Weiss et al., 1986]. In addition, tumors overexpressing fatty acid synthase (FAS), the enzyme responsible for de novo synthesis of fatty acids, display aggressive biologic behavior compared to those tumors with normal FAS levels, suggesting that FAS overexpression confers a selective growth advantage. Here, we will review the roles that FAS plays in important cellular processes such as apoptosis and proliferation. In addition, speculations on the putative role of FAS in the altered metabolic pathways of prostate cancer cells will be explored. Because of the frequent overexpression of this enzyme prostate cancer, FAS constitutes a therapeutic target in this disease.  相似文献   

10.
The gene PA4866 from Pseudomonas aeruginosa is documented in the Pseudomonas genome database as encoding a 172 amino acid hypothetical acetyltransferase. We and others have described the 3D structure of this protein (termed pita) [Davies et al. (2005) Proteins: Struct., Funct., Bioinf. 61, 677-679; Nocek et al., unpublished results], and structures have also been reported for homologues from Agrobacterium tumefaciens (Rajashankar et al., unpublished results) and Bacillus subtilis [Badger et al. (2005) Proteins: Struct., Funct., Bioinf. 60, 787-796]. Pita homologues are found in a large number of bacterial genomes, and while the majority of these have been assigned putative phosphinothricin acetyltransferase activity, their true function is unknown. In this paper we report that pita has no activity toward phosphinothricin. Instead, we demonstrate that pita acts as an acetyltransferase using the glutamate analogues l-methionine sulfoximine and l-methionine sulfone as substrates, with Km(app) values of 1.3 +/- 0.21 and 1.3 +/- 0.13 mM and kcat(app) values of 505 +/- 43 and 610 +/- 23 s-1 for l-methionine sulfoximine and l-methionine sulfone, respectively. A high-resolution (1.55 A) crystal structure of pita in complex with one of these substrates (l-methionine sulfoximine) has been solved, revealing the mode of its interaction with the enzyme. Comparison with the apoenzyme structure has also revealed how certain active site residues undergo a conformational change upon substrate binding. To investigate the role of pita in P. aeruginosa, a mutant strain, Depp4, in which pita was inactivated through an in-frame deletion, was constructed by allelic exchange. Growth of strain Depp4 in the absence of glutamine was inhibited by l-methionine sulfoximine, suggesting a role for pita in protecting glutamine synthetase from inhibition.  相似文献   

11.
Crystal structure of human dihydrofolate reductase complexed with folate   总被引:8,自引:0,他引:8  
The crystal structure of recombinant human dihydrofolate reductase with folate bound in the active site has been determined and the structural model refined at 0.2-nm resolution. Preliminary studies of the binding of the inhibitors methotrexate and trimethoprim to the human apoenzyme have been performed at 0.35-nm resolution. The conformations of the chemically very similar ligands folate and methotrexate, one a substrate the other a potent inhibitor, differ substantially in that their pteridine rings are in inverse orientations relative to their p-aminobenzoyl-L-glutamate moieties. Methotrexate binding is similar to that previously observed in two bacterial enzymes but is quite different from that observed in the enzyme from a mouse lymphoma cell line [Stammers et al. (1987) FEBS Lett. 218, 178-184]. The geometry of the polypeptide chain around the folate binding site in the human enzyme is not consistent with conclusions previously drawn with regard to the species selectivity of the inhibitor trimethoprim [Matthews et al. (1985) J. Biol. Chem. 260, 392-399].  相似文献   

12.
In most tissues, ribonucleases (RNases) are found in a latent form complexed with ribonuclease inhibitor (RI). To examine whether these so-called cytoplasmic RNases belong to the same superfamily as pancreatic RNases, we have purified from porcine liver two such RNases (PL1 and PL3) and examined their primary structures. It was found that RNase PL1 belonged to the same family as human RNase Us [Beintema et al. (1988) Biochemistry 27, 4530-4538] and bovine RNase K2 [Irie et al. (1988) J. Biochem. (Tokyo) 104, 289-296]. RNase PL3 was found to be a hitherto structurally uncharacterized type of RNase. Its polypeptide chain of 119 amino acid residues was N-terminally blocked with pyroglutamic acid, and its sequence differed at 63 positions with that of the pancreatic enzyme. All residues important for catalysis and substrate binding have been conserved. Comparison of the primary structure of RNase PL3 with that of its bovine counterpart (RNase BL4; M. Irie, personal communication) revealed an unusual conservation for this class of enzymes; the 2 enzymes were identical at 112 positions. Moreover, comparison of the amino acid compositions of these RNases with that of a human colon carcinoma-derived RNase, RNase HT-29 [Shapiro et al. (1986) Biochemistry 25, 7255-7264], suggested that these three proteins are orthologous gene products. The structural characteristics of RNases PL1 and PL3 were typical of secreted RNases, and this observation questions the proposed cytoplasmic origin of these RI-associated enzymes.  相似文献   

13.
R67 dihydrofolate reductase (DHFR) shares no sequence or structural homology with chromosomal DHFRs. This enzyme arose recently in response to the clinical use of the antibacterial drug trimethoprim. R67 DHFR is a homotetramer possessing a single active site pore. A high-resolution crystal structure shows the homotetramer possesses exact 222 symmetry [Narayana, N., et al. (1995) Nat. Struct. Biol. 2, 1018-1025]. This symmetry dictates four symmetry-related binding sites must exist for each substrate as well as each cofactor. Isothermal titration calorimetry studies, however, indicate only two molecules bind: either two dihydrofolate molecules, two NADPH molecules, or one substrate and one cofactor [Bradrick, T. D., et al. (1996) Biochemistry 35, 11414-11424]. The latter is the productive ternary complex. To evaluate the role of S65, Q67, I68, and Y69 residues, located near the center of the active site pore, site-directed mutagenesis was performed. One mutation in the gene creates four mutations per active site pore which typically result in large cumulative effects. Steady state kinetic data indicate the mutants have altered K(m) values for both cofactor and substrate. For example, the Y69F R67 DHFR displays an 8-fold increase in the K(m) for dihydrofolate and a 20-fold increase in the K(m) for NADPH. Residues involved in ligand binding in R67 DHFR display very little, if any, specificity, consistent with their possessing dual roles in binding. These results support a model where R67 DHFR utilizes an unusual "hot spot" binding surface capable of binding both ligands and indicate this enzyme has adopted a novel yet simple approach to catalysis.  相似文献   

14.
Fenton chemistry [Fenton (1894) J. Chem. Soc. 65, 899-910] techniques were employed to identify the residues involved in metal binding located at the active sites of restriction endonucleases. This process uses transition metals to catalytically oxidize the peptide linkage that is in close proximity to the amino acid residues involved in metal ligation. Fe2+ was used as the redox-active transition metal. It was expected that Fe2+ would bind to the endonucleases at the Mg2+-binding site [Liaw et al. (1993) Biochemistry 32, 7999-4003; Ermácora et al. (1992) Proc. Natl. Acad. Sci. U.S.A. 89, 6383-6387; Soundar and Colman (1993) J. Biol. Chem. 268, 5264-5271; Wei et al. (1994) Biochemistry 33, 7931-7936; Ettner et al. (1995) Biochemistry 34, 22-31; Hlavaty and Nowak (1997) Biochemistry 36, 15515-15525). Fe2+-mediated oxidation was successfully performed on TaqI endonulease, suggesting that this approach could be applied to a wide array of endonucleases [Cao and Barany (1998) J. Biol. Chem. 273, 33002-33010]. The restriction endonucleases BamHI, FokI, BglI, BglII, PvuII, SfiI, BssSI, BsoBI, EcoRI, EcoRV, MspI, and HinP1I were subjected to oxidizing conditions in the presence of Fe2+ and ascorbate. All proteins were inactivated upon treatment with Fe2+ and ascorbate. BamHI, FokI, BglI, BglII, PvuII, SfiI, BssSI, and BsoBI were specifically cleaved upon treatment with Fe2+/ascorbate. The site of Fe2+/ascorbate-induced protein cleavage for each enzyme was determined. The Fe2+-mediated oxidative cleavage of BamHI occurs between residues Glu77 and Lys78. Glu77 has been shown by structural and mutational studies to be involved in both metal ligation and catalysis [Newman et al. (1995) Science 269, 656-663; Viadiu and Aggarwal (1998) Nat. Struct. Biol. 5, 910-916; Xu and Schildkraut (1991) J. Biol. Chem. 266, 4425-4429]. The sites of Fe2+/ascorbate-induced cleavage for PvuII, FokI, BglI, and BsoBI agree with the metal-binding sites identified in their corresponding three-dimensional structures or from mutational studies [Cheng et al. (1994) EMBO J. 13, 3297-3935; Wah et al. (1997) Nature 388, 97-100; Newman et al. (1998) EMBO J. 17, 5466-5476; Ruan et al. (1997) Gene 188, 35-39]. The metal-binding residues of BglII, SfiI, and BssSI are proposed based on amino acid sequencing of their Fe2+/ascorbate-generated cleavage fragments. These results suggest that Fenton chemistry may be a useful methodology in identifying amino acids involved in metal binding in endonucleases.  相似文献   

15.
Bovine milk is by far the most commonly consumed milk in the western world. The protein composition in milk consists of casein and whey proteins, of which β-lactoglobulin (BLG) is the principal constituent of the latter. Here we provide biochemical evidence that this milk protein, in purified form and in pasteurized store-bought milk, promotes the formation of cycloretinal (all-trans retinal dimer), and a variety of other cycloterpenals of biological relevance [Fishkin et al., Proc. Natl. Acad. Sci. U. S. A., 2005, 102, 7091-7096; Fishkin et al., Chirality, 2004, 16, 637-641; Kim et al., Proc. Natl. Acad. Sci. U. S. A., 2007, 104, 19273-19278]. Cycloretinal is an eye metabolite and among several toxic byproducts of the visual cycle firmly established to cause age-related macular degeneration. Experiments in rabbits further demonstrate that BLG/milk can survive the digestive system and promote this reaction in vivo [Caillard et al., Am. J. Physiol., 1994, 266(6), G1053-G1059]. Proteomic studies on age-related macular degeneration patients have detected BLG in the eye of these patients further suggesting that this milk protein could contribute to disease progression [Crabb et al., Proc. Natl. Acad. Sci. U. S. A., 2002, 99(23), 14682-14687].  相似文献   

16.
Tubulin folding cofactor D is a microtubule destabilizing protein   总被引:1,自引:0,他引:1  
A rapid switch between growth and shrinkage at microtubule ends is fundamental for many cellular processes. The main structural components of microtubules, the alphabeta-tubulin heterodimers, are generated through a complex folding process where GTP hydrolysis [Fontalba et al. (1993) J. Cell Sci. 106, 627-632] and a series of molecular chaperones are required [Sternlicht et al. (1993) Proc. Natl. Acad. Sci. USA 90, 9422-9426; Campo et al. (1994) FEBS Lett. 353, 162-166; Lewis et al. (1996) J. Cell Biol. 132, 1-4; Lewis et al. (1997) Trends Cell Biol. 7, 479-484; Tian et al. (1997) J. Cell Biol. 138, 821-823]. Although the participation of the cofactor proteins along the tubulin folding route has been well established in vitro, there is also evidence that these protein cofactors might contribute to diverse microtubule processes in vivo [Schwahn et al. (1998) Nature Genet. 19, 327-332; Hirata et al. (1998) EMBO J. 17, 658-666; Fanarraga et al. (1999) Cell Motil. Cytoskel. 43, 243-254]. Microtubule dynamics, crucial during mitosis, cellular motility and intracellular transport processes, are known to be regulated by at least four known microtubule-destabilizing proteins. OP18/Stathmin and XKCM1 are microtubule catastrophe-inducing factors operating through different mechanisms [Waters and Salmon (1996) Curr. Biol. 6, 361-363; McNally (1999) Curr. Biol. 9, R274-R276]. Here we show that the tubulin folding cofactor D, although it does not co-polymerize with microtubules either in vivo or in vitro, modulates microtubule dynamics by sequestering beta-tubulin from GTP-bound alphabeta-heterodimers.  相似文献   

17.
A Zurlinden  M E Schweingruber 《Gene》1992,117(1):141-143
Biosyntheses of the pyrimidine and thiazole moieties of the thiamine molecule occur by separate pathways. In Schizosaccharomyces pombe, a gene, thi2, is responsible for thiazole synthesis [Schweingruber et al., Curr. Genet. 19 (1991) 249-254]. We have cloned a 3.1-kb genomic S. pombe fragment which can functionally complement a thi2 mutant. The fragment maps genetically at the thi2 site, indicating that it carries thi2. As shown by Northern hybridization analysis, the appearance of thi2 mRNA levels is repressed when cells are grown in the presence of thiamine and 5-(2-hydroxyethyl)-4-methylthiazole. The thi3 gene involved in the biosynthesis of the pyrimidine moiety, is also regulated by thiamine [Maundrell, J. Biol. Chem. 265 (1990) 10857-10864; Schweingruber et al., Curr. Genet. 19 (1991) 249-254]. We previously identified and analyzed four regulatory genes (tnr1, tnr2, tnr3, and thi1) that are responsible for the regulation of thi3 [Schweingruber et al., Genetics (1992) in press]. Mutants defective in these regulatory genes affect expression of thi2 in a similar way to thi3. This indicates that biosynthesis of the pyrimidine and thiazole moieties are under common genetic control in S. pombe.  相似文献   

18.
A structural gene encoding bovine (b) tryptophanyl-tRNA synthetase (WRS) has recently been cloned and sequenced [Garret et al., Biochemistry 30 (1991) 7809-7817]. Using part of this sequence as a hybridisation probe we have cloned and sequenced a structural gene encoding human polypeptide highly homologous with two mammalian proteins, bWRS [Garret et al., Biochemistry 30 (1991) 7809-7817; EMBL accession No. X52113] and rabbit peptide chain release factor [Lee et al., Proc. Natl. Acad. Sci. USA 87 (1990) 3508-3512]. Identification of the sequence encoding a human WRS is based on (i) the presence of 'HIGH' and 'KMSKS' structural motifs typical for class-I aminoacyl-tRNA synthetases [Eriani et al., Nature 347 (1990) 203-206]; (ii) coincidence of the number of SH groups per subunit estimated experimentally [Muench et al., Science 187 (1975) 1089-1091] and deduced from the cDNA sequence (six in both cases); (iii) close resemblance of two WRS polypeptides sequenced earlier [Muench et al., Science 187 (1975) 1089-1091] and the predicted structure in two different regions.  相似文献   

19.
Contradicting early suggestions, the sequencing of the gene encoding the Desulfovibrio desulfuricans (ATCC 27774) nine-heme cytochrome c proves that this cytochrome is not the product of the degradation of the 16-heme containing cytochrome c [Coelho et al. (1996) Acta Cryst. D52, 1202-1208]. However, preliminary data indicate that the cytochrome gene is part of an operon similar to the DvH hmc operon, which contains the gene coding for the 16-heme cytochrome c [Rossi et al. (1993) J. Bacteriol. 175, 4699-4711]. Also, the amino acid sequence deduced from the DNA sequence shows four residues in the C-terminal not predicted in the amino acid sequence obtained by X-ray methods [Matias et al. (1999) Structure 7, 119-130].  相似文献   

20.
Lancaster CR 《FEBS letters》2001,504(3):133-141
The structure of Wolinella succinogenes quinol:fumarate reductase by X-ray crystallography has been determined at 2.2-A resolution [Lancaster et al. (1999), Nature 402, 377-385]. Based on the structure of the three protein subunits A, B, and C and the arrangement of the six prosthetic groups (a covalently bound FAD, three iron-sulphur clusters, and two haem b groups) a pathway of electron transfer from the quinol-oxidising dihaem cytochrome b in the membrane to the site of fumarate reduction in the hydrophilic subunit A has been proposed. By combining the results from site-directed mutagenesis, functional and electrochemical characterisation, and X-ray crystallography, a residue was identified which is essential for menaquinol oxidation. [Lancaster et al. (2000), Proc. Natl. Acad. Sci. USA 97, 13051-13056]. The location of this residue in the structure suggests that the coupling of the oxidation of menaquinol to the reduction of fumarate in dihaem-containing succinate:quinone oxidoreductases could be associated with the generation of a transmembrane electrochemical potential. Based on crystallographic analysis of three different crystal forms of the enzyme and the results from site-directed mutagenesis, we have derived a mechanism of fumarate reduction and succinate oxidation [Lancaster et al. (2001) Eur. J. Biochem. 268, 1820-1827], which should be generally relevant throughout the superfamily of succinate:quinone oxidoreductases.  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