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.
The homology of Subunit III of bovine procarboxypeptidase A-S6 (EC 3.4.12.-) with Subunit II (bovine chymotrypsinogen C) of the same complex, already reported in a previous publication (Puigserver, A. and Desnuelle, P. (1975) Proc. Natl. Acad. Sci. U.S. 72, 2442-2445) has been extended to the position of the single methionine of the chains. The sequence linked by 4 disulfide bridges out of 5 are also probably homologous in the 2 proteins. The last bridge is displaced in Subunit III as a consequence of the deletion of the N-terminal half-cystine. Subunit II, which is not activatable by trypsin, due to the loss of essential residues in the N-terminal region, has conserved a weakly functional active site reacting with concentrated diisopropylfluorophosphate at exactly the same rate as that of Subunit II.  相似文献   

2.
Novel aryl derivatives of benzamidine were synthesized and tested for their inhibitory potency against bovine trypsin, rat skin tryptase, human recombinant granzyme A, human thrombin, and human plasma kallikrein. All compounds show competitive inhibition against these proteases with Ki values in the micromolar range. X-ray structures were determined to 1.8 A resolution for trypsin complexed with two of the para-substituted benzamidine derivatives, 1-(4-amidinophenyl)-3-(4-chlorophenyl)urea (ACPU) and 1-(4-amidinophenyl)-3-(4-phenoxyphenyl)urea (APPU). Although the inhibitors do not engage in direct and specific interactions outside the S1 pocket, they do form intimate indirect contacts with the active site of trypsin. The inhibitors are linked to the enzyme by a sulfate ion that forms an intricate network of three-centered hydrogen bonds. Comparison of these structures with other serine protease structures with noncovalently bound oxyanions reveals a pair of highly conserved oxyanion-binding sites in the active site. The positions of noncovalently bound oxyanions, such as the oxygen atoms of sulfate, are distinct from the positions of covalent oxyanions of tetrahedral intermediates. Noncovalent oxyanion positions are outside the "oxyanion hole." Kinetics data suggest that protonation stabilizes the ternary inhibitor/oxyanion/protease complex. In sum, both cations and anions can mediate Ki. Cation mediation of potency of competitive inhibitors of serine proteases was previously reported by Stroud and co-workers [Katz, B. A., Clark, J. M., Finer-Moore, J. S., Jenkins, T. E., Johnson, C. R., Ross, M. J., Luong, C., Moore, W. R., and Stroud, R. M. (1998) Nature 391, 608-612].  相似文献   

3.
AMP deaminase isoforms purified from endogenous sources display smaller than predicted subunit molecular masses, whereas baculoviral expression of human AMPD1 (isoform M) and AMPD3 (isoform E) cDNAs produces full-sized recombinant enzymes. However, nearly 100 N-terminal amino acid residues are cleaved from each recombinant polypeptide during storage at 4 degreesC. Expression of N-truncated cDNAs (DeltaL96AMPD1 and DeltaM90AMPD3) produces stable recombinant enzymes exhibiting subunit molecular masses and kinetic properties that are similar to those reported for purified isoforms M and E. Conversely, wild type recombinant isoforms display significantly higher Km(app) values in the absence of ATP. Gel filtration analysis demonstrates native tetrameric structures for all recombinant proteins, except the wild type AMPD1 enzyme, which forms aggregates of tetramers that disperse upon cleavage of N-terminal residues at 4 degreesC. These data: 1) confirm that available literature on AMP deaminase is likely derived from N-truncated enzymes and 2) are inconsistent with a new model proposing native trimeric structure of an N-truncated rabbit skeletal muscle AMP deaminase (Ranieri-Raggi, M., Montali, U., Ronca, F., Sabbatini, A., Brown, P. E., Moir, A. J. G., and Raggi, A. (1997) Biochem. J. 326, 641-648). N-terminal residues also influence actomyosin-binding properties of the enzyme, which are enhanced and suppressed by AMPD1 and AMPD3 sequences, respectively. Finally, co-expression of AMPD1 and AMPD3 recombinant polypeptides produces tetrameric enzymes with either isoform-specific or mixed subunits, and also reveals that tetramer assembly is driven by relative polypeptide abundance with no apparent preference for like subunits.  相似文献   

4.
The structure of (D)-glucarate dehydratase from Pseudomonas putida (GlucD) has been solved at 2.3 A resolution by multiple isomorphous replacement and refined to a final R-factor of 19.0%. The protein crystallizes in the space group I222 with one subunit in the asymmetric unit. The unit cell dimensions are a = 69.6 A, b = 108.8 A, and c = 122.6 A. The crystals were grown using the batch method where the primary precipitant was poly(ethylene glycol) 1000. The structure reveals that GlucD is a tetramer of four identical polypeptides, each containing 451 residues. The structure was determined without a bound substrate or substrate analogue. Three disordered regions are noted: the N-terminus through residue 11, a loop containing residues 99 through 110, and the C-terminus from residue 423. On the basis of primary sequence alignments, we previously concluded that GlucD is a member of the mandelate racemase (MR) subfamily of the enolase superfamily [Babbitt, P. C., Hasson, M. S., Wedekind, J. E., Palmer, D. R. J., Barrett, W. C., Reed, G. J., Rayment, I., Ringe, D., Kenyon, G. L., and Gerlt, J. A. (1996) Biochemistry 35, 16489-16501]. This prediction is now verified, since the overall fold of GlucD is strikingly similar to those of MR, muconate lactonizing enzyme I, and enolase. Also, many of the active site residues of GlucD can be superimposed on those found in the active site of MR. The implications of this structure on the evolution of catalysis in the enolase superfamily are discussed.  相似文献   

5.
Pterin-4a-carbinolamine dehydratase/dimerization cofactor for hepatocyte nuclear factor-1 alpha is a protein with two different functions. We have overexpressed and purified the human wild-type protein, and its Cys81Ser and Cys81Arg mutants. The Cys81Arg mutant has been proposed to be causative in a hyperphenylalaninaemic patient [Citron, B. A., Kaufman, S., Milstien, S., Naylor, E. W., Greene, C. L. & Davis, M. D. (1993) Am. J. Hum. Genet. 53, 768-774]. The dehydratase behaves as a tetramer on gel filtration, while cross-linking experiments showed mono-, di-, tri-, and tetrameric forms, irrespective of the presence of the single Cys81. Sulfhydryl-modifying reagents did not affect the activity, but rather showed that Cys81 is exposed. Various pterins bind and quench the tryptophan fluorescence suggesting the presence of a specific binding site. The fluorescence is destroyed upon light irradiation. Wild-type and the Cys81Ser protein enhance the rate of the phenylalanine hydroxylase assay approximately 10-fold, a value similar to that of native dehydratase from rat liver; the Cys81Arg mutant, in contrast, has significantly lower activity. This is compatible with the hypothesis that the dehydratase is a rate-limiting factor for the in vivo phenylalanine hydroxylase reaction. The three proteins enhance the spontaneous dehydration of the synthetic substrate 6,6-dimethyl-7,8-dihydropterin-4a-carbinolamine approximately 50-70-fold at 4 degrees C and pH 8.5. The results are discussed in view of the recently solved three-dimensional structure of the enzyme [Ficner, R., Sauer, U. W., Stier, G. & Suck, D. (1995) EMBO J. 14, 2032-2042].  相似文献   

6.
Replication factor C (RF-C) is a heteropentameric protein essential for DNA replication and repair. It is a molecular matchmaker required for loading of proliferating cell nuclear antigen (PCNA) onto double-stranded DNA and, thus, for PCNA-dependent DNA elongation by DNA polymerases delta and epsilon. To elucidate the mode of RF-C binding to the PCNA clamp, modified forms of human PCNA were used that could be 32P-labeled in vitro either at the C or the N terminus. Using a kinase protection assay, we show that the heteropentameric calf thymus RF-C was able to protect the C-terminal region but not the N-terminal region of human PCNA from phosphorylation, suggesting that RF-C interacts with the PCNA face at which the C termini are located (C-side). A similar protection profile was obtained with the recently identified PCNA binding region (residues 478-712), but not with the DNA binding region (residues 366-477), of the human RF-C large subunit (Fotedar, R., Mossi, R., Fitzgerald, P., Rousselle, T., Maga, G., Brickner, H., Messner, H., Khastilba, S., Hübscher, U., and Fotedar, A., (1996) EMBO J., 15, 4423-4433). Furthermore, we show that the RF-C 36 kDa subunit of human RF-C could interact independently with the C-side of PCNA. The RF-C large subunit from a third species, namely Drosophila melanogaster, interacted similarly with the modified human PCNA, indicating that the interaction between RF-C and PCNA is conserved through eukaryotic evolution.  相似文献   

7.
The flavoprotein component (SiR-FP) of the sulfite reductase from Escherichia coli is an octamer containing one FAD and one FMN as cofactors per polypeptide chain. We have constructed an expression vector containing the DNA fragment encoding for the FMN-binding domain of SiR-FP. The overexpressed protein (SiR-FP23) was purified as a partially flavin-depleted polymer. It could incorporate FMN exclusively upon flavin reconstitution to reach a maximum flavin content of 1.2 per polypeptide chain. Moreover, the protein could stabilize a neutral air-stable semiquinone radical over a wide range of pHs. During photoreduction, the flavin radical accumulated first, followed by the fully reduced state. The redox potentials, determined at room temperature [E'1 (FMNH./FMN) = -130 +/- 10 mV and E'2 (FMNH2/FMNH.) = -335 +/- 10 mV], were very close to those previously reported for Salmonella typhimurium SiR-FP [Ostrowski, J., Barber, M. J., Rueger, D. C., Miller, B. E., Siegel, L. M., & Kredich, N. M. (1989) J. Biol. Chem. 264, 15796-15808]. Both the radical and fully reduced forms of SiR-FP23 were able to transfer their electrons to cytochrome c quantitatively. Altogether, the results presented herein demonstrate that the N-terminal end of E. coli SiR-FP forms the FMN-binding domain. It folds independently, thus retaining the chemical properties of the bound FMN, and provides a good model of the FAD-depleted form of native SiR-FP. Moreover, the FMN prosthetic group in SiR-FP23 and native SiR-FP is compared to that of cytochrome P450 reductase and bacterial cytochrome P450, which also contain one FAD and one FMN per polypeptide chain.  相似文献   

8.
One popular model for protein folding, the framework model, postulates initial formation of secondary structure elements, which then assemble into the native conformation. However, short peptides that correspond to secondary structure elements in proteins are often only marginally stable in isolation. A 33-residue peptide (GCN4-p1) corresponding to the GCN4 leucine zipper folds as a parallel, two-stranded coiled coil [O'Shea, E.K., Klemm, J.D., Kim, P.S., & Alber, T.A. (1991) Science 254, 539-544]. Deletion of the first residue (Arg 1) results in local, N-terminal unfolding of the coiled coil, suggesting that a stable subdomain of GCN4-p1 can form. N- and C-terminal deletion studies result in a 23-residue peptide, corresponding to residues 8-30 of GCN4-p1, that folds as a parallel, two-stranded coil with substantial stability (the melting temperature of a 1 mM solution is 43 degrees C at pH 7). In contrast, a closely related 23-residue peptide (residues 11-33 of GCN4-p1) is predominantly unfolded, even at 0 degrees C, as observed previously for many isolated peptides of similar length. Thus, specific tertiary packing interactions between two short units of secondary structure can be energetically more important in stabilizing folded structure than secondary structure propensities. These results provide strong support for the notion that stable, cooperatively folded subdomains are the important determinants of protein folding.  相似文献   

9.
Phosphorylase kinase, a regulatory enzyme of glycogenolysis in skeletal muscle, is a hexadecameric oligomer containing four copies each of four distinct subunits: alpha, beta, gamma, and delta. By intramolecular zero-length crosslinking with transglutaminase, we have previously demonstrated that the regulatory alpha and beta subunits abut one another in the holoenzyme [Nadeau, O. W., and Carlson, G. M. (1994) J. Biol. Chem. 269, 29670-29676]. Selective partial proteolysis of the 138 kDa alpha subunit in holoenzyme that had been crosslinked by transglutaminase has revealed a high molecular weight conjugate corresponding to full-length beta subunit crosslinked to a 60 kDa N-terminal fragment of alpha (determined by SDS-PAGE, Western blotting and N-terminal sequencing). This conjugate was also observed when the enzyme was first activated by partial proteolysis of alpha and then crosslinked by transglutaminase. Both forms of the kinase, generated by either sequential crosslinking and proteolysis or the reverse, coeluted with non-crosslinked hexadecameric control enzyme in size exclusion chromatography, indicating that the crosslinking was intramolecular, i.e., within hexadecamers. This is the first demonstration of any intersubunit interaction involving the N-terminal domain of the alpha subunit and the first region of any subunit shown to interact with the beta subunit. The results are consistent with the predicted path of the polypeptide backbone of the alpha subunits within the holoenzyme and with the proposed location of the beta subunits.  相似文献   

10.
Polydisperse proteoglycan subunit from bovine proximal humeral articular cartilage has been separated into a series of relatively monodisperse fractions which have been chemically and physically characterized. The proteoglycan subunit species of the lowest molecular weight contains the least chondroitin sulfate and had an amino acid composition relatively low in serine and glycine and relatively high in cysteine, methionine, and aspartic acid, almost identical to that of the hyaluronic acid-binding region of proteoglycan subunit isolated by Heinegard and Hascall (Heinegard, D., and Hascall, V.C. (1974) J. Biol. Chem. 249, 4250-4256). The molecular weight of proteoglycan subunit increases in proportion to its chondroitin sulfate content. As the molecular weight and chondroitin sulfate content of proteoglycan subunit increase, there is a parallel increase in the serine and glycine contents, and a decrease in the cysteine, methionine, and aspartic acid contents of proteoglycan subunit core protein. The pattern of polydispersity observed strongly supports the concept that proteoglycan subunit core protein contains a hyaluronic acid-binding region of constant size and composition and a polysaccharide attachment region of variable length and composition, composed of repeating peptide sequences containing serine and glycine in equimolar amounts.  相似文献   

11.
12.
The first type III module of fibronectin (Fn) contains a cryptic site that binds Fn and its N-terminal 29 kDa fragment and is thought to be important for fibril formation (Morla, A., Zhang, Z., and Ruoslahti, E. (1994) Nature 367, 193-196; Hocking, D. C., Sottile, J. , and McKeown-Longo, P. J. (1994) J. Biol. Chem. 269, 19183-19191). A synthetic 31-mer peptide (NAPQ ... TIPG) derived from the middle of domain III1 was also shown to bind Fn, but the site of its interaction was not determined (Morla, A., and Ruoslahti, E. (1992) J. Cell Biol. 118, 421-429). By affinity chromatography on peptide-agarose, we tested a set of fragments representing the entire light chain of plasma Fn. Only 40-kDa Hep-2 (III12-15) failed to bind. The concentration of urea required for peak elution of Fn and the other fragments decreased in the order Fn > 42-kDa GBF (I6II1-2I7-9) > 19-kDa Fib-2 (I10-12) > 110-kDa CBF(III2-10) > 29-kDa Fib-1 (I1-I5). Neither Fn nor any of the fragments bound immobilized intact III1, confirming the cryptic nature of this activity. In an effort to detect interactions between other Fn domains, all fragments were coupled to Sepharose, and each fragment was tested on each affinity matrix before and after denaturation. The only interaction detected was that of fluid phase III1 with immobilized denatured 110-kDa CBF and 40-kDa Hep-2, both of which contain type III domains. Analysis of subfragments revealed this activity to be dominated by domains III7 and III15. Fn itself did not bind to the denatured fragments. Thus, domain III1 contains two cryptic "self-association sites," one that is buried in the core of the fold but recognizes many Fn fragments when presented as a peptide and another that is concealed in Fn but exposed in the native isolated domain and recognizes cryptic sites in two other type III domains. These interactions between type III domains could play an important role in assembly of Fn multimers in the extracellular matrix.  相似文献   

13.
The QCR8 genes encoding subunit VIII of the bc1 complex from Kluyveromyces lactis and Schizosaccharomyces pombe partially complement the respiratory-deficient phenotype of a S. cerevisiae QCR8-null mutant. This implies that the heterologous Qcr8 subunits can be imported by S. cerevisiae mitochondria and that they assemble to form a hybrid bc1 complex that is sufficiently active to support growth. In contrast, the QCR8 gene from bovine heart, encoding the 9.5-kDa subunit, is not able to restore respiratory function to the S. cerevisiae null mutant. This lack of functional complementation is directly attributable to the inability of S. cerevisiae mitochondria to import this protein as shown by in vitro assays. However, a hybrid gene encoding the N-terminal 26 residues of S. cerevisiae subunit VIII and the rest of the 9.5-kDa bovine heart homologue, was able to functionally complement the QCR8-null mutant, albeit to a very low extent. Successful import into S. cerevisiae mitochondria was confirmed by in vitro import experiments. Surprisingly, although assembly of these hybrid complexes is reduced to an extent that is proportional to the evolutionary distance of the homologue to S. cerevisiae, the specific activities of the assembled complexes are the same as for the wild-type bc1 complex. After solubilisation of the mitochondrial membranes with the mild detergent dodecyl maltoside, the wild-type enzyme can be inactivated by incubation at increased temperature, independent of protease activity. The rate of inactivation can be significantly increased by the addition of o-phenanthroline [Boumans, H., Grivell, L. A. & Berden, J. A. (1997) J. Biol. Chem. 272, 16753-16760]. The hybrid complexes are much more sensitive to both types of treatment. We conclude that substitution of subunit VIII by a homologous counterpart results in a loosening of the structure of the bc1 complex on the intermembrane space side, resulting in a less stable insertion of the Rieske Fe-S protein in vivo and therefore a lower stability of the assembled enzyme under certain in vitro conditions, but without an effect on catalytic activity.  相似文献   

14.
Diaminopimelate dehydrogenase catalyzes the NADPH-dependent reduction of ammonia and L-2-amino-6-ketopimelate to form meso-diaminopimelate, the direct precursor of L-lysine in the bacterial lysine biosynthetic pathway. Since mammals lack this metabolic pathway inhibitors of enzymes in this pathway may be useful as antibiotics or herbicides. Diaminopimelate dehydrogenase catalyzes the only oxidative deamination of an amino acid of D configuration and must additionally distinguish between two chiral amino acid centers on the same symmetric substrate. The Corynebacterium glutamicum enzyme has been cloned, expressed in Escherichia coli, and purified to homogeneity using standard biochemical procedures [Reddy, S. G., Scapin, G., & Blanchard, J. S. (1996) Proteins: Structure, Funct. Genet. 25, 514-516]. The three-dimensional structure of the binary complex of diaminopimelate dehydrogenase with NADP+ has been solved using multiple isomorphous replacement procedures and noncrystallographic symmetry averaging. The resulting model has been refined against 2.2 A diffraction data to a conventional crystallographic R-factor of 17.0%. Diaminopimelate dehydrogenase is a homodimer of structurally not identical subunits. Each subunit is composed of three domains. The N-terminal domain contains a modified dinucleotide binding domain, or Rossman fold (six central beta-strands in a 213456 topology surrounded by five alpha-helices). The second domain contains two alpha-helices and three beta-strands. This domain is referred to as the dimerization domain, since it is involved in forming the monomer--monomer interface of the dimer. The third or C-terminal domain is composed of six beta-strands and five alpha-helices. The relative position of the N- and C-terminal domain in the two monomers is different, defining an open and a closed conformation that may represent the enzyme's binding and active state, respectively. In both monomers the nucleotide is bound in an extended conformation across the C-terminal portion of the beta-sheet of the Rossman fold, with its C4 facing the C-terminal domain. In the closed conformer two molecules of acetate have been refined in this region, and we postulate that they define the DAP binding site. The structure of diaminopimelate dehydrogenase shows interesting similarities to the structure of glutamate dehydrogenase [Baker, P. J., Britton, K. L., Rice, D. W., Rob, A., & Stillmann, T.J. (1992a) J. Mol. Biol. 228, 662-671] and leucine dehydrogenase [Baker, P.J., Turnbull, A.P., Sedelnikova, S.E., Stillman, T. J., & Rice, D. W. (1995) Structure 3, 693-705] and also resembles the structure of dihydrodipicolinate reductase [Scapin, G., Blanchard, J. S., & Sacchettini, J. C. (1995) Biochemistry 34, 3502-3512], the enzyme immediately preceding it in the diaminopimelic acid/lysine biosynthetic pathway.  相似文献   

15.
The zinc metalloenzyme beta-lactamase II (betaLII) from Bacillus cereus has been overexpressed in Escherichia coli as a fusion protein with glutathione-S-transferase, and the metal binding properties of recombinant betaLII toward Zn(II) and Co(II) have been studied by fluorescence and activity measurements. The apoenzyme is able to bind two metal ion equivalents, which confer on betaLII its maximum enzymatic efficiency. The enzyme is partially active with one metal ion equivalent. The diCo(II) and a mixed Zn(II)Co(II) derivative of betaLII were obtained and probed by electronic and paramagnetic NMR spectroscopy. In the high-affinity site, the metal is bound to three His residues and a solvent molecule, adopting a tetrahedral geometry. A Cys, a His, and an Asp residue are coordinated to the low-affinity metal site, together with two or three solvent molecules. This coordination polyhedron resembles the binuclear metal site of the Bacteroides fragilis beta-lactamase [Concha, N., Rasmussen, B. A., Bush, K., and Herzberg, O. (1996) Structure 4, 823-836; Carfi, A., Duée, E., Paul-Soto, R., Galleni, M., Frère, J. M., and Dideberg, O. (1998) Acta Crystallogr. D54, 47-57] but differs from that resulting from the X-ray study of betaLII [Carfi, A., Pares, S., Duée, E., Galleni, M., Duez, C., Frère, J. M., and Dideberg, O. (1995) EMBO J. 14, 4914-4921]. These results suggest that this binuclear metal site may be a general feature of metallo-beta-lactamases.  相似文献   

16.
The betagamma subunits of the heterotrimeric GTP-binding proteins (G proteins) that couple heptahelical, plasma membrane-bound receptors to intracellular effector enzymes or ion channels directly regulate several types of effectors, including phospholipase Cbeta and adenylyl cyclase. The beta subunit is made up of two structurally different regions: an N-terminal alpha helix followed by a toroidal structure made up of 7 blades, each of which is a twisted beta sheet composed of four anti-parallel beta strands (Wall, M. A., Coleman, D. E., Lee, E., I?iguez-Lluhi, J. A., Posner, B. A., Gilman, A. G., and Sprang, S. R. (1995) Cell 83, 1047-1058; Lambright, D. G., Sondek, J., Bohm, A., Skiba, N. P., Hamm, H. E., and Sigler, P. B. (1996) Nature 379, 311-319). We have previously shown that sites for activation of PLCbeta2, PLCbeta3, and adenylyl cyclase II overlap on the "top" surface of the propeller, where Galpha also binds (Li, Y., Sternweis, P. M., Charnecki, S., Smith, T. F., Gilman, A. G., Neer, E. J., and Kozasa, T. (1998) J. Biol. Chem. 273, 16265-16272). The present study was undertaken to identify the regions on the side of the torus that might be important for effector interactions. We made mutations in each of the outer beta strands of the G protein beta1 propeller, as well as mutations in the loops that connect the outer strands to the adjacent beta strands. Our results suggest that activation of PLCbeta2 involves residues in the outer strands of blades 2, 6, and 7 of the propeller. We tested three of the mutations that most severely affected PLCbeta2 activity against two forms of adenylyl cyclase (ACI and ACII). Both inhibition of ACI and activation of ACII were unaffected by these mutations, suggesting that if ACI and ACII contact the outer strands, the sites of contact are different from those for PLCbeta2. We propose that distinct sets of contacts along the sides of the propeller will define the specificity of the interaction of betagamma with effectors.  相似文献   

17.
The nucleotide sequence of the Clostridium thermocellum F7 cbhA gene, coding for the cellobiohydrolase CbhA, has been determined. An open reading frame encoding a protein of 1,230 amino acids was identified. Removal of a putative signal peptide yields a mature protein of 1,203 amino acids with a molecular weight of 135,139. Sequence analysis of CbhA reveals a multidomain structure of unusual complexity consisting of an N-terminal cellulose binding domain (CBD) homologous to CBD family IV, an immunoglobulin-like beta-barrel domain, a catalytic domain homologous to cellulase family E1, a duplicated domain similar to fibronectin type III (Fn3) modules, a CBD homologous to family III, a highly acidic linker region, and a C-terminal dockerin domain. The cellulosomal localization of CbhA was confirmed by Western blot analysis employing polyclonal antibodies raised against a truncated enzymatically active version of CbhA. CbhA was identified as cellulosomal subunit S3 by partial amino acid sequence analysis. Comparison of the multidomain structures indicates striking similarities between CbhA and a group of cellulases from actinomycetes. Average linkage cluster analysis suggests a coevolution of the N-terminal CBD and the catalytic domain and its spread by horizontal gene transfer among gram-positive cellulolytic bacteria.  相似文献   

18.
H-Ras, the protein product of the cellular homologue of the Harvey ras oncogene, undergoes a complex series of post-translational modifications that include C-terminal isoprenylation, proteolysis, methylation, and palmitoylation. Palmitoylation has been shown to enhance the transformation efficiency of H-Ras about 10-fold in vivo. A recent study (Magee, A. I., Gutierrez, L., McKay, I. A., Marshall, C. J., and Hall, A. (1987) EMBO J. 6, 3353-3357) has provided strong evidence that the palmitate undergoes a dynamic acylation-deacylation cycle, but details concerning the enzymology of this process and its regulation are lacking. To begin to dissect this event, we have developed an assay for the enzymatic removal of palmitate from [3H]palmitate-labeled H-Ras. This substrate was produced in a baculovirus expression system and was used to purify to homogeneity a novel 37-kDa enzyme from bovine brain cytosol that removes the radiolabeled palmitate. The purified enzyme is sensitive to diethyl pyrocarbonate and insensitive to phenylmethylsulfonyl fluoride and N-ethylmaleimide. Interestingly, the thioesterase recognizes H-Ras as a substrate only when H-Ras is in its native conformation (bound to Mg2+ and guanine nucleotide). The palmitoylated alpha subunits of the heterotrimeric G proteins are also substrates for the enzyme.  相似文献   

19.
Analysis of mutational effects in the human immunodeficiency virus type-1 (HIV-1) provirus has revealed that as few as four amino acid side-chain substitutions in the HIV-1 protease (M46I/L63P/V82T/I84V) suffice to yield viral variants cross-resistant to a panel of protease inhibitors either in or being considered for clinical trials (Condra, J. H., Schleif, W. A., Blahy, O. M., Gadryelski, L. J., Graham, D. J., Quintero, J. C., Rhodes, A., Robbins, H. L., Roth, E., Shivaprakash, M., Titus, D., Yang, T., Teppler, H., Squires, K. E., Deutsch, P. J., and Emini, E. A. (1995) Nature 374, 569-571). As an initial effort toward elucidation of the molecular mechanism of drug resistance in AIDS therapies, the three-dimensional structure of the HIV-1 protease mutant containing the four substitutions has been determined to 2.4-A resolution with an R factor of 17.1%. The structure of its complex with MK639, a protease inhibitor of the hydroxyaminopentane amide class of peptidomimetics currently in Phase III clinical trials, has been resolved at 2.0 A with an R factor of 17.0%. These structures are compared with those of the wild-type enzyme and its complex with MK639 (Chen, Z., Li, Y., Chen, E., Hall, D. L., Darke, P. L., Culberson, C., Shafer, J., and Kuo, L. C. (1994) J. Biol. Chem. 269, 26344-26348). There is no gross structural alteration of the protease due to the site-specific mutations. The C alpha tracings of the two native structures are identical with a root-mean-square deviation of 0.5 A, and the four substituted side chains are clearly revealed in the electron density map. In the MK639-bound form, the V82T substitution introduces an unfavorable hydrophilic moiety for binding in the active site and the I84V substitution creates a cavity (unoccupied by water) that should lead to a decrease in van der Waals contacts with the inhibitor. These changes are consistent with the observed 70-fold increase in the Ki value (approximately 2.5 kcal/mol) for MK639 as a result of the mutations in the HIV-1 protease. The role of the M46I and L63P substitutions in drug resistance is not obvious from the crystallographic data, but they induce conformational perturbations (0.9-1.1 A) in the flap domain of the native enzyme and may affect the stability and/or activity of the enzyme unrelated directly to binding.  相似文献   

20.
Time-resolved circular dichroism spectroscopy in the far-UV spectral region was used to examine the intermediates of the phytochrome photoreversion reaction (Pfr --> Pr). Three intermediates, lumi-F (tau = 320 ns), meta-Fa (tau = 265 micros) and meta-Fb (tau = 5.5 ms), have been identified in a simple sequential kinetic photoreversion mechanism by absorption spectroscopy [Linschitz, H., Kasche, V., Butler, W. L., & Siegelman, H. W. (1966) J. Biol. Chem. 241, 3395-3403; Pratt, L. H., & Butler, W. L. (1968) Photochem. Photobiol. 8, 477-485; Burke, M., Pratt, D. C., & Moscowitz, A. (1972) Biochemistry 11, 4025-4031; Spruit, C. J. P., Kendrick, R. E., & Cooke, R. J. (1975) Planta (Berlin) 127, 121-132; Eilfeld, P., & Rüdiger, W. (1985) Z. Naturforsch. 40c, 109-114; Chen, E., Lapko, V. N., Lewis, J. W., Song, P.-S., & Kliger, D. S. (1996) Biochemistry 35, 843-850]. In order to correlate the unfolding of the N-terminal alpha-helical segment with one or more of the intermediate species, time-resolved methods were coupled with the structurally sensitive probe of CD in the far-UV spectral region. Analysis of the TRCD data associates the decrease in alpha-helical content that occurs upon formation of Pr with decay of the meta-Fa intermediate. This unfolding process occurs with a time constant of 310 +/- 125 micros, which is consistent with the 265-micros lifetime for meta-Fa.  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