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 515 毫秒
1.
The denaturation and reconstitution of Erwinia carotovora and Escherichia coli L-asparaginases has been followed by optical rotatory dispersion, circular dichroism and analytical ultracentrifugation. Denaturation in urea results in dissociation of the native enzyme (mol. wt. 140 000 approx.) to produce unfolded subunits (mol. wt. 35 000 approx.); the Erwinia L-asparaginase subunits can be refolded by dilution or dialysis in alkaline conditions, pH 10.5, without aggregation to the active tetramer, to give a rather unstable solution of a monomer possibly in equilibrium with dimer. These alkaline-reconstituted subunits undergo a conformational change to a more ordered state in the presence of sodium dodecylsulphate, similar to those produced by the action of sodium dodecylsulphate on the native enzyme. If the denatured subunits are reconstituted in the pH range 5.0-7.5, the enzymically active tetramer is reformed in up to 80% yield, depending upon the conditions of temperature and concentration. Kinetic data for these various transitions suggest that dissociation is a rate-limiting step while conformational changes of the polypeptide chains are relatively much more rapid. The possible significance of these different rates of change to therapeutic considerations is discussed.  相似文献   

2.
Experimental conditions favouring the dissociation of tetrameric rabbit muscle D-glyceraldehyde-3-phosphate dehydrogenase into active monomers were elaborated. The urea-induced dissociation of the tetramer was shown to be a stepwise process (in 2 M urea only dimers are formed; an increase in urea concentration up to 3 M causes the splitting of the dimers into monomers). The specific activity of immobilized monomers in the glyceraldehyde-3-phosphate oxidation reaction does not differ from that of the parent immobilized tetrameric form. The tetrameric enzyme molecule binds the coenzyme with a negative cooperativity (the first two NAD+ molecules bind with KD below 0.1 microM; for the third and fourth molecules the dissociation constant was determined to be equal to 5.5 +/- 1.5 microM (50 mM medinal buffer, 10 mM sodium phosphate, pH 8.2). The cooperativity of NAD+ binding is preserved in the immobilized preparation of tetrameric dehydrogenase. The immobilized monomers bind NAD+ with KD of 1.6 +/- 1.0 microM. The experimental results are consistent with the hypothesis according to which the association of catalytically active subunits into a tetramer changes their coenzyme-binding properties in such a way that the first two NAD+ molecules bind more firmly to a tetramer than to a monomer, whereas the third and the fourth NAD+ molecules bind less firmly.  相似文献   

3.
Purified L-asparaginase from Serratia marcescens had an apparent-weight average molecular weight of 171,000 to 180,000 as determined by electrophoresis on polyacrylamide gels and by sedimentation equilibrium at low speed in an analytical ultracentrifuge. A subunit molecular weight of 31,500 +/- 1,500 was estimated for the enzyme after treatment with sodium dodecyl sulfate and urea and electrophoresis on polyacrylamide gels; a similar value was obtained by high-speed sedimentation equilibrium in the presence of guanidine hydrochloride. Our data indicate that the Serratia enzyme could have five or six subunits of 32,000 daltons, compared to four subunits of 32,000 daltons in the Escherichia coli enzyme. The Serratia L-asparaginase also appears to be a larger molecule than the enzyme from Erwinia carotovora, Proteus vulgaris, Acinetobacter glutaminasificans, and Alcaligenes eutrophus. The Serratia enzyme, like that from E. caratovora, was more resistant than the E. coli enzyme to dissociation by sodium dodecyl sulfate. This resistance could be due to the finding that the Serratia enzyme had a relatively high hydrophobicity, similar to the enzyme from E. caratovora, when compared with the hydrophobicity of the E. coli enzyme. The isoelectric point of the Serratia enzyme was approximately 5.2. The influence of certain physical characteristics of the enzyme on the biological properties is discussed.  相似文献   

4.
L-Asparaginase (L-asparagine amidohydrolase, EC 3.5.1.1) from Erwinia carotovora undergoes extensive dissociation from active tetramer to inactive monomers when freeze-dried. The monomeric state is stabilized by reconstitution of the freeze-dried enzyme with buffers of high pH and high ionic strength. Some compounds, particularly sugars and sugar derivatives, prevent dissociation on freeze-drying, whereas others, such as urea and chaotropic ions, increase dissociation. The effects of additives are not related to water retention. The dissociation is completely reversible on reconstitution at neutral pH, but the alkali-stabilized monomer only partially reassociates when the pH is brought back to neutrality.  相似文献   

5.
The binding of a spin-labeled AMP analog to tetrameric glyceraldehyde-3-phosphate dehydrogenase from rabbit muscle is described. The spin label, perdeuterated and 15N-substituted 4-amino-2,2,6,6-tetramethylpiperidine-1-oxyl, was attached to C-8 of AMP (C8-SL-AMP). Up to 8 equivalents of C8-SL-AMP bind per enzyme tetramer, i.e., 2 per monomer. Combining sites are the adenine subsite of the coenzyme-binding domain and the phosphate site. Glyceraldehyde 3-phosphate causes a conformational change in the enzyme that brings C8-SL-AMP molecules bound to adjacent R-axis-related subunits closer to one another by 0.2-0.3 nm and allows for spin-spin interaction between the nitroxide radicals. Similar, but less pronounced structural changes take place upon lowering the pH from 8 to 7. Addition of a single equivalent of NAD+ to a complex of the enzyme with 7.6 equivalents of C8-SL-AMP leads to the release of almost 4 C8-SL-AMP molecules. This supports our previous findings that binding of just one NAD+ molecule induces conformational changes in all four subunits.  相似文献   

6.
7.
Analysis of the native enzyme and of the subunits produced upon its denaturation shows that pyruvate carboxylase from a thermophilic Bacillus is a tetramer with a molecular weight (mean value) of 558,000 and that the four polypeptide subunits are probably identical. The three functions (carboxyl carrier, carboxylation, and carboxyl transfer) in the pyruvate carboxylation reaction must therefore reside in this quarter-molecular polypeptide. The enzyme molecule contains four atoms of zinc and four molecules of D-biotin, and in the electron microscope the disposition of its four subunits presents a rhombic appearance. Reaction of the denatured enzyme with 5,5'-dithiobis (2-nitrobenzoic acid) (DTNB) reveals 10 sulfhydryl groups/subunit. In the native enzyme less than one of these groups reacts with DTNB. By contrast, all of these groups (11/subunit) of the native chicken liver pyruvate carboxylase are accessible to DTNB. The thermophile enzyme is also more resistant to other sulfhydryl reagents and to denaturation under certain conditions than the avian enzyme.  相似文献   

8.
d-Arabinose isomerase (EC 5.3.1.3) has been isolated from l-fucose-induced cultures of Escherichia coli K-12 and d-arabinose-induced cultures of E. coli B/r. Both enzymes were homogeneous in an ultracentrifuge and migrated as single bands upon disc electrophoresis in acrylamide gels. The s(20,w) was 14.5 x 10(-13) sec for the E. coli K-12 enzyme and 14.3 x 10(-13) sec for the E. coli B/r enzyme. The molecular weight, determined by high-speed sedimentation equilibrium, was 3.55 +/- 0.06 x 10(5) for the E. coli K-12 enzyme and 3.42 +/- 0.04 x 10(5) for the enzyme isolated from E. coli B/r. Both enzyme preparations were active wth l-fucose or d-arabinose as substrates and showed no activity on any of the other aldopentoses or aldohexoses tested. With the E. coli K-12 enzyme, the K(m) was 2.8 x 10(-1)m for d-arabinose and 4.5 x 10(-2)m for l-fucose; with the E. coli B/r enzyme, the K(m) was 1.7 x 10(-1)m for d-arabinose and 4.2 x 10(-2)m for l-fucose. Both enzymes were inhibited by several of the polyalcohols tested, ribitol, l-arabitol, and dulcitol being the strongest. Both enzymes exhibited a broad plateau of optimal catalytic activity in the alkaline range. Both enzymes were stimulated by the presence of Mn(2+) or Co(2+) ions, but were strongly inhibited by the presence of Cd(2+) ions. Both enzymes were precipitated by antisera prepared against either enzyme preparation. The amino acid composition for both proteins has been determined; a striking similarity has been detected. Both enzymes could be dissociated, by protonation at pH 2 or by dialysis against buffer containing 8 m urea, into subunits that were homogeneous in an ultracentrifuge and migrated as single bands on disc electrophoresis in acrylamide gels containing urea. The molecular weight of the subunit, determined by high-speed sedimentation equilibrium, was 9.09 +/- 0.2 x 10(4) for the enzyme from E. coli K-12 and 8.46 +/- 0.1 x 10(4) for the enzyme from E. coli B/r. On the basis of biophysical studies, both isomerases appear to be oligomeric proteins consisting of four identical subunits.  相似文献   

9.
Some properties of homogeneous transketolase from pig liver were studied. It was shown that the pH optimum of the transketolase reaction lies within the range of 7.8--8.2. The isoelectric point is at pH 7.6--7.8. The molecular weight of transketolase is 138,000 +/- 3,000 as determined by the sedimentation equilibrium method and about 152,000 according to the data from gel filtration through Sephadex G-200. The enzyme molecule is a tetramer of the alpha 2 beta 2 type. The molecular weights of the alpha- and beta- subunits determined by polyacrylamide gel in the presence of sodium dodecyl sulfate are 52,000--56,000 and 27,000--29,000, respectively. Transketolase contains about two moles of TPP per mole of protein and does not require metal ions for its catalytic activity.  相似文献   

10.
Succinyl-CoA synthetase catalyzes the substrate-level phosphorylation step of the tricarboxylic acid cycle. The enzyme, as isolated from Escherichia coli, has an alpha 2 beta 2 subunit structure. It is known that substrate-binding sites are distributed between both subunit types and that the active enzyme is the nondissociating tetramer. This paper describes a study of the process of assembly of the enzyme from its denatured constituent subunits. Starting with equimolar mixtures of the subunits that are prepared in denaturing conditions (6 M urea, 5% acetic acid), rapid renaturation to produce virtually a fully active enzyme occurs after neutralization and dilution under suitable conditions. This process occurs most efficiently in the presence of either ATP or Pi, indicating that occupation of the phosphoryl-binding site on the refolding alpha subunit facilitates productive intrasubunit interactions. We have determined conditions of protein concentration, pH, temperature, final urea concentration, and buffer compositions that optimize both the rate and extent of production of active enzyme. The final refolded product is indistinguishable from the native species with respect to its specific catalytic activity, size, and other physical properties. To probe further the mechanism and route of renaturation, we have shown that the rate of appearance of activity has first-order dependence on each of the two subunits. The step that determines the rate of assembly is thus bimolecular, such as the association of structural monomers to form a dimeric transient species. The highly specific mutual interactions between the refolding transient species of subunits must be essential for the correct assembly of this enzyme from the two gene products in vivo.  相似文献   

11.
The dissociation of porcine erythrocyte catalase [EC 1.11.1.6] into subunits on denaturation with alkali, GuHCl and urea was investigated by following the changes in hydrodynamic properties, absorption and CD spectra in the Soret region and inactivation of the enzyme. It was found that dissociation proceeded in an "all or none" manner from the native tetramer (molecular weight, ca. 250,000) into identical 1/4-sized monomers (molecular weight, ca. 54,000 with alkali, 65,000 with urea and 71,000 with GuHCl) as estimated by ultracentrifugal analyses. On this dissociation, the sedimentation coefficient decreased from about 11S to 5.1 - 3.7S, and absorption spectra in the Soret region decreased to about 40% of the native level and showed a broad band around 365-375 nm and a shoulder around 415-420 nm; these changes were accompanied by complete loss of enzyme activity. The change in enzyme activity correlated well with that of absorption and CD spectra in the Soret region, depending on denaturation time, alkaline pH used and concentration of both denaturants. The reassociated catalase obtained by removing urea by dialysis was characterized by recovery of distinct CD bands in the Soret and near ultraviolet regions, although the partial refolding of alpha-helical conformation occurred without recovery of enzyme activity. These results indicate that the conformational changes and dissociation process of catalase into subunits can be monitored spectrophotometrically in relation to enzyme activity, and that subtle conformations near the heme groups and polypeptide backbone play an important role in maintaining full enzyme activity of the catalase molecule.  相似文献   

12.
A crystalline tryptophanase can be obtained from extracts of Spaerophorus funduliformis using a heat treatment, hydroxyapatite chromatography and solubility in solutions of (NH4)2SO4 as a function of pH and temperature. The purified enzyme is homogeneous by several criteria. S. funduliformis tryptophanase has a specific activity of 11.5-13.5 and requires pyridoxal 5'-phosphate for enzymatic activity. Like other tryptophanases that have been studied, the S. funduliformis enzyme is a tetramer protein consisting of four apparently identical subunits. The native enzyme has a sedimentation coefficient of 11.2 S and a molecular weight of 244 000. In solutions of 5 M guanidine - HCl, 8 M urea, and sodium dodecylsulfate, at high pH or in the presence of thiols, the enzyme dissociates to 59 000 molecular weight species which are homogeneous by the criterion of weight. Peptide maps of the reduced holo-tryptophanase show one pyridoxal-containing peptide and, lacking agreement with the determined amino acid composition, suggest that the subunits of the enzyme contain a high degree of internal sequence homology.  相似文献   

13.
Conformational changes of duck ovostatin (ovomacroglobulin) upon complexing with thermolysin have been studied by electron microscopy. Both free and thermolysin-bound ovostatin preparations were negatively stained with uranyl acetate, a series of three pictures were taken at 10 degrees specimen tilt intervals (+10 degrees, 0 degrees, and -10 degrees), and images of the inhibitor molecules were observed in three dimensions. Four approximately cylindrical subunits were observed in free ovostatin. Two subunits associated approximately midway from both ends to form a dimer of four arms. Two dimers associated with each other at the midpoint to form a tetramer. The proteinase susceptible "bait" regions were located near the center of the molecule. Eight arms of the tetramer take various configurations. The orthogonal extent of free tetrameric ovostatin in a two-dimensional micrograph averages 26.0 +/- 4.7 x 34.0 +/- 5.0 nm. Upon complexing with thermolysin, all eight arms curl toward the center of the molecule, having four arms upward and the other four downward. Thus, proteinase-bound ovostatin has a uniform structure with a 2-fold axis of symmetry. The overall structure of the complex is more compact with average dimensions of 16.9 +/- 0.6 x 16.9 +/- 0.6 x 19.9 +/- 0.4 nm. From these electron microscopic studies we propose that a proteinase reaches to the center of the free ovostatin molecule and attacks the bait region. Subsequent to proteolysis the subunit arms curl and entrap the enzyme within the ovostatin molecule. The results support the unique mechanism of inhibition of proteinases by alpha 2-macroglobulin and ovostatin postulated from biochemical observations (Barrett, A. J., and Starkey, P. M. (1973) Biochem. J. 133, 709-724; Nagase, H., and Harris, E. D., Jr. (1983) J. Biol. Chem. 258, 7490-7498).  相似文献   

14.
Liver glucosidase II from control and ethanol-treated rats was purified and its physical and catalytic properties studied. No significant variation was found in the purification and properties of the enzyme from either source (ethanol-treated and control rats), except for activity after storage. Glucosidase II was purified to electrophoretic homogeneity from liver microsomes by solubilization, protamine sulphate precipitation, and anion exchange (DEAE-Sephacel) and affinity (Con A-Sepharose-4B) chromatographies. The native enzyme molecule is a tetramer (Mr 425 +/- 10 kDa) with identical subunits (Mr 106 +/- 3 kDa). Km values, determined at pH 6.8 for the p-nitrophenyl-glucosidase activity of glucosidase II from control and ethanol-treated rats, were 1.20 +/- 0.12 and 1.14 +/- 0.13 mM, respectively. The Arrhenius plot was linear, and the value for the apparent activation energy, calculated from this plot, was 56.64 kJ/mol. The p-nitrophenyl-glucosidase activity of glucosidase II from control and ethanol-treated rats was inhibited to the same extent by NH4+, by the divalent cations Ca2+ and Mg2+, and by methanol, ethanol, 2-propanol, n-propanol, isobutanol and n-butanol.  相似文献   

15.
Catalase was purified to an electrophoretically homogeneous state from the facultative alkalophilic bacterium, Bacillus YN-2000, and some of its properties were studied. Its molecular weight was 282,000 and its molecule was composed of four identical subunits. The enzyme contained two protoheme molecules per tetramer. The enzyme showed an absorption spectrum of typical high-spin ferric heme with a peak at 406 nm in the oxidized form and peaks at 440, 559, and 592 nm in the reduced form. In contrast to the typical catalases, the enzyme was reduced with sodium dithionite, like peroxidases. The enzyme showed an appreciable peroxidase activity in addition to high catalase activity. The amino acid composition of Bacillus YN-2000 catalase was very similar to those of catalase from Neurospora crassa and peroxidase from Halobacterium halobium. The catalase content in the soluble fraction from the bacterium was higher with the cells grown at pH 10 than with the cells grown at lower pHs (pH 7-9).  相似文献   

16.
DHDPS (dihydrodipicolinate synthase) catalyses the branch point in lysine biosynthesis in bacteria and plants and is feedback inhibited by lysine. DHDPS from the thermophilic bacterium Thermotoga maritima shows a high level of heat and chemical stability. When incubated at 90 degrees C or in 8 M urea, the enzyme showed little or no loss of activity, unlike the Escherichia coli enzyme. The active site is very similar to that of the E. coli enzyme, and at mesophilic temperatures the two enzymes have similar kinetic constants. Like other forms of the enzyme, T. maritima DHDPS is a tetramer in solution, with a sedimentation coefficient of 7.2 S and molar mass of 133 kDa. However, the residues involved in the interface between different subunits in the tetramer differ from those of E. coli and include two cysteine residues poised to form a disulfide bond. Thus the increased heat and chemical stability of the T. maritima DHDPS enzyme is, at least in part, explained by an increased number of inter-subunit contacts. Unlike the plant or E. coli enzyme, the thermophilic DHDPS enzyme is not inhibited by (S)-lysine, suggesting that feedback control of the lysine biosynthetic pathway evolved later in the bacterial lineage.  相似文献   

17.
Rabbit antisera against L-asparaginase preparations from Escherichia coli, Erwinia carotovora, Citrobacter sp. and Chromobacterium violaceum showed on immunoelectrophoresis that only the enzymes from E. coli and Citrobacter are immunologically related. Purified preparations had to be used to determine the immunological cross-reactions. Immunoelectrophoresis at different pH values yielded the zero mobility points of the enzymes. The activity of the Er. carotovora preparation was enhanced up to fourfold by homologous antiserum but not by normal sera. Heterologous antisera also enhanced, but only at a higher concentration. Less enhancement was observed for the other enzymes with antisera as well as with bovine serum albumin. Inhibition was not observed.  相似文献   

18.
The proton-translocating NADH-quinone oxidoreductase (NDH-1) of Paracoccus denitrificans is composed of at least 14 subunits (NQO1-14) and is located in the cytoplasmic membrane. In the present study, topological properties and stoichiometry of the 7 subunits (NQO1-6 and NQO9) of the P. denitrificans NDH-1 in the membranes were investigated using immunological techniques. Treatments with chaotropic reagents (urea, NaI, or NaBr) or with alkaline buffer (pH 10-12) resulted in partial or complete extraction of all the subunits from the membranes. Of interest is that when NaBr or urea were used, the NQO6 and NQO9 subunits remained in the membranes, whereas the other subunits were completely extracted, suggesting their direct association with the membrane part of the enzyme complex. Both deletion study and homologous expression study of the NQO9 subunit provided a clue that its hydrophobic N-terminal stretch plays an important role in such an association. In light of this observation and others, topological properties of the subunits in the NDH-1 enzyme complex are discussed. In addition, determination of stoichiometry of the peripheral subunits of the P. denitrificans NDH-1 was completed by radioimmunological methods. All the peripheral subunits are present as one molecule each in the enzyme complex. These results estimated the total number of cofactors in the P. denitrificans NDH-1; the enzyme complex contains one molecule of FMN and up to eight iron-sulfur clusters, 2x[2Fe-2S] and 6x[4Fe-4S], provided that the NQO6 subunit bears one [4Fe-4S] cluster.  相似文献   

19.
The Escherichia coli single strand binding (SSB) protein is an essential protein required for DNA replication and involved in recombination and a number of repair processes. It is a stable homotetramer in solution; however the ssb-1 mutation (His-55 to Tyr) destabilizes the tetramer with respect to monomers and this defect seems to explain the observed phenotype (Williams, K. R., Murphy, J. B., and Chase, J. W. (1984) J. Biol. Chem. 259, 11804-11811). We report a quantitative study of the SSB-1 monomer-tetramer equilibrium in vitro as a function of temperature, pH, NaCl, MgCl2, urea, and guanidine hydrochloride concentrations. The self-assembly equilibrium was monitored by the increase in intrinsic protein fluorescence anisotropy accompanying the formation of the tetramer. The experimental isotherms indicate that SSB-1 dimers are not highly populated at equilibrium, hence the formation of the tetramer is well-described as a one-step association of four monomers. At 25 degrees C, pH 8.1, the monomer concentration for 50% tetramer dissociation is (MT)1/2 = 0.87 microM, corresponding to a monomer-tetramer equilibrium constant, KT = 3 +/- 1 x 10(18) M-3. The tetramerization constant, KT, is highly dependent upon temperature and pH, with delta H0 = -51 +/- 7 kcal/mol (pH 8.1) and delta H0 = -37 +/- 5 kcal/mol (pH 6.9). There is no effect of NaCl on the monomer-tetramer association in the range from 0.20 to 1.0 M; however, MgCl2 decreases the stability of the SSB-1 tetramer. In the presence of high concentrations of the single-stranded oligonucleotide, dT(pT)15, the tetramerization constant is slightly increased indicating that binding of the oligonucleotide to the SSB-1 monomer promotes the assembly process, although not dramatically. The large negative delta H0 that is associated with formation of the tetramer provides a likely explanation for the temperature sensitivity of the ssb-1 mutation.  相似文献   

20.
The collagen prolyl 4-hydroxylases (C-P4Hs), enzymes residing within the lumen of the endoplasmic reticulum, play a central role in the synthesis of all collagens. The vertebrate enzymes are alpha(2)beta(2) tetramers in which the two catalytic sites are located in the alpha subunits, and protein disulfide isomerase serves as the beta subunit. All attempts to assemble an active C-P4H tetramer from its subunits in in vitro cell-free systems have been unsuccessful, but assembly of a recombinant enzyme has been reported in several cell types by coexpression of the two types of subunit. An active type I C-P4H tetramer was obtained here by periplasmic expression in Escherichia coli strains BL21 and RB791. Further optimization for production by stepwise regulated coexpression of its subunits in the cytoplasm of a thioredoxin reductase and glutathione reductase mutant E. coli strain resulted in large amounts of human type I C-P4H tetramer. The specific activity of the C-P4H tetramer purified from the cytoplasmic expression was within the range of values reported for human type I C-P4H isolated as a nonrecombinant enzyme or produced in the endoplasmic reticulum of insect cells, but the expression level, about 25 mg/l in a fermenter, is about 5-10 times that obtained in insect cells. The enzyme expressed in E. coli differed from those present in vivo and those produced in other hosts in that it lacked the N glycosylation of its alpha subunits, which may be advantageous in crystallization experiments.  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