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 31 毫秒
1.
Salts affect protein stability by multiple mechanisms (e.g., the Hofmeister effect, preferential hydration, electrostatic effects and weak ion binding). These mechanisms can affect the stability of both the native state and the unfolded state. Previous equilibrium stability studies demonstrated that KCl stabilizes dihydrofolate reductases (DHFRs) from Escherichia coli (ecDHFR, E. coli DHFR) and Haloferax volcanii (hvDHFR1, H. volcanii DHFR encoded by the hdrA gene) with similar efficacies, despite adaptation to disparate physiological ionic strengths (0.2 M versus 2 M). Kinetic studies can provide insights on whether equilibrium effects reflect native state stabilization or unfolded state destabilization. Similar kinetic mechanisms describe the folding of urea-denatured ecDHFR and hvDHFR1: a 5-ms stopped-flow burst-phase species that folds to the native state through two sequential intermediates with relaxation times of 0.1-3 s and 25-100 s. The latter kinetic step is very similar to that observed for the refolding of hvDHFR1 from low ionic strength. The unfolding of hvDHFR1 at low ionic strength is relatively slow, suggesting kinetic stabilization as observed for some thermophilic enzymes. Increased KCl concentrations slow the urea-induced unfolding of ecDHFR and hvDHFR1, but much less than expected from equilibrium studies. Unfolding rates extrapolated to 0 M denaturant, kunf(H2O), are relatively independent of ionic strength, demonstrating that the KCl-induced stabilization of ecDHFR and hvDHFR1 results predominantly from destabilization of the unfolded state. This supports the hypothesis from previous equilibrium studies that haloadaptation harnesses the effects of elevated salt concentrations on the properties of the aqueous solvent to enhance protein stability.  相似文献   

2.
The extremely halophilic bacterium Salinibacter ruber was previously shown to have a high intracellular potassium content, comparable to that of halophilic Archaea of the family Halobacteriaceae. The amino acid composition of its bulk protein showed a high content of acidic amino acids, a low abundance of basic amino acids, a low content of hydrophobic amino acids, and a high abundance of serine. We tested the level of four cytoplasmic enzymatic activities at different KCl and NaCl concentrations. Nicotinamide adenine dinucleotide (NAD)-dependent isocitrate dehydrogenase functioned optimally at 0.5-2 M KCl, with rates of 60% of the optimum value at 3.3 M. NaCl provided less activation: 70% of the optimum rates in KCl were found at 0.2-1.2 M NaCl, and above 3 M NaCl, activity was low. We also detected nicotinamide adenine dinucleotide phosphate (NADP)-dependent isocitrate activity, which remained approximately constant between 0-3.2 M NaCl and increased with increasing KCl concentration. NAD-dependent malate dehydrogenase functioned best in the absence of salt, but rates as high as 25% of the optimal values were measured in 3-3.5 M KCl or NaCl. NAD-dependent glutamate dehydrogenase, assayed by the reductive amination of 2-oxoglutarate, showed low activity in the absence of salt. NaCl was stimulatory with optimum activity at 3-3.5 M. However, no activity was found above 2.5 M KCl. Although the four activities examined all function at high salt concentrations, the behavior of individual enzymes toward salt varied considerably. The results presented show that Salinibacter enzymes are adapted to function in the presence of high salt concentrations.  相似文献   

3.
G Zaccai  S Y Xian 《Biochemistry》1988,27(4):1316-1320
Yeast tRNA(Phe) was studied in different salt-containing solvents by UV absorbance and small-angle neutron scattering (SANS). This extends results obtained previously in NaCl and KCl solutions [Li, Z.-Q., Giegé, R., Jacrot, B., Oberthür, R., Thierry, J. C., & Zaccai, G. (1983) Biochemistry 22, 4380-4388]. As expected, at low concentrations of all salts studied, the tRNA molecule is unfolded. The importance of specific counterion interactions and the flexibility of the macromolecule are emphasized by the observation that it cannot take up its folded structure in N(CH3)4Cl solvents, even when that salt concentration is increased to 1 M, in the absence of Mg ions. In CsCl solvents, on the other hand, the folded conformation is obtained in salt concentrations above about 0.2 M, similar to NaCl or KCl. By a comparison of SANS results in CsCl H2O and CsCl 2H2O solvents with the data from NaCl and KCl solvents, thermodynamic and structural parameters were derived for the solvated macromolecule. All the data are accounted for, quantitatively, by a model for the particle in NaCl, KCl, or CsCl solution made up of tRNA76-, closely associated with 76 positive hydrated counterions, surrounded by an aqueous solvent layer that excludes salt (and, therefore, of density different from that of bulk solvent). The mass of water in that layer depends on salt concentration, and the values found are consistent with those predicted by the Donnan effect.  相似文献   

4.
To examine whether dihydrofolate reductase (DHFR) from deep-sea bacteria has undergone molecular evolution to adapt to high-pressure environments, we cloned eight DHFRs from Shewanella species living in deep-sea and ambient atmospheric-pressure environments, and subsequently purified six proteins to compare their structures, stabilities, and functions. The DHFRs showed 74–90% identity in primary structure to DHFR from S. violacea, but only 55% identity to DHFR from Escherichia coli (ecDHFR). Far-ultraviolet circular dichroism and fluorescence spectra suggested that the secondary and tertiary structures of these DHFRs were similar. In addition, no significant differences were found in structural stability as monitored by urea-induced unfolding and the kinetic parameters, K m and k cat; although the DHFRs from Shewanella species were less stable and more active (2- to 4-fold increases in k cat/K m) than ecDHFR. Interestingly, the pressure effects on enzyme activity revealed that DHFRs from ambient-atmospheric species are not necessarily incompatible with high pressure, and DHFRs from deep-sea species are not necessarily tolerant of high pressure. These results suggest that the DHFR molecule itself has not evolved to adapt to high-pressure environments, but rather, those Shewanella species with enzymes capable of retaining functional activity under high pressure migrated into the deep-sea.  相似文献   

5.
The kinetic characteristics of a purified insect dihydrofolate reductase (DHFR) have been described. The Km values for the substrate dihydrofolate and the cofactor NADPH have been estimated by primary and secondary Hanes plots to be 0.3 and 5.2 microM, respectively. Drosophila melanogaster DHFR can use folate and NADH at acidic pH values, but at a much lower rate than the preferred substrate and cofactor. Folic acid is a partial competitive inhibitor of Drosophila DHFR (Ki = 0.4 microM) and trimethoprim is a complete competitive inhibitor (Ki = 5.4 microM). Methotrexate binds less tightly to the Drosophila enzyme than to many other DHFRs (Kd = 0.9 nM). Drosophila DHFR is inhibited by KCl and organic mercurials and is slightly activated by urea. These data indicate that Drosophila DHFR has some characteristics which are typical of vertebrate DHFRs and others which are typical of prokaryotic DHFRs. The study of this enzyme, therefore, should aid in the definition of the structural features that are responsible for the kinetic characteristics in different DHFRs.  相似文献   

6.
Thermophilic lactate dehydrogenases from Thermotoga maritima and Bacillus stearothermophilus are stable up to temperature limits close to the optimum growth temperature of their parent organisms. Their catalytic properties are anomalous in that Km shows a drastic increase with increasing temperature. At low temperatures, the effect levels off. Extreme halophilic malate dehydrogenase from Halobacterium marismortui exhibits a similar anomaly. Increasing salt concentration (NaCl) leads to an optimum curve for Km, oxaloacctate while Km, NADH remains constant. Previous claims that the activity of halophilic malate dehydrogenase shows a maximum at 1.25 M NaCl are caused by limiting substrate concentration; at substrate saturation, specific activity of halophilic malate dehydrogenase reaches a constant value at ionic strengths I greater than or equal to 1 M. Non-halophilic (mitochondrial) malate dehydrogenase shows Km characteristics similar to those observed for the halophilic enzyme. The drastic decrease in specific activity of the mitochondrial enzyme at elevated salt concentrations is caused by the salt-induced increase in rigidity of the enzyme, rather than gross structural changes.  相似文献   

7.
To understand the pressure-adaptation mechanism of deep-sea enzymes, we studied the effects of pressure on the enzyme activity and structural stability of dihydrofolate reductase (DHFR) of the deep-sea bacterium Moritella profunda (mpDHFR) in comparison with those of Escherichia coli (ecDHFR). mpDHFR exhibited optimal enzyme activity at 50MPa whereas ecDHFR was monotonically inactivated by pressure, suggesting inherent pressure-adaptation mechanisms in mpDHFR. The secondary structure of apo-mpDHFR was stable up to 80°C, as revealed by circular dichroism spectra. The free energy changes due to pressure and urea unfolding of apo-mpDHFR, determined by fluorescence spectroscopy, were smaller than those of ecDHFR, indicating the unstable structure of mpDHFR against pressure and urea despite the three-dimensional crystal structures of both DHFRs being almost the same. The respective volume changes due to pressure and urea unfolding were -45 and -53ml/mol at 25°C for mpDHFR, which were smaller (less negative) than the corresponding values of -77 and -85ml/mol for ecDHFR. These volume changes can be ascribed to the difference in internal cavity and surface hydration of each DHFR. From these results, we assume that the native structure of mpDHFR is loosely packed and highly hydrated compared with that of ecDHFR in solution.  相似文献   

8.
Extracts prepared from a halophilic bacterium contained a reduced nicotinamide adenine dinucleotide (NADH(2)) oxidase active at high solute concentrations. The cation requirement was nonspecific, since KCl, RbCl, and CsCl replaced NaCl with little or no loss of activity, and NH(4)Cl was only partially effective. Only LiCl failed to replace NaCl. No specific chloride requirement was observed although not all anions replaced chloride. Bromide, nitrate, and iodide were essentially ineffective, whereas acetate, formate, citrate, and sulfate proved suitable. The presence of sulfate affected the ability of a cation to satisfy the solute requirement. Sulfate enhanced the rate of NADH(2) oxidation when compared with the rate observed in the presence of chloride. Cations which were inactive as chlorides (LiCl and MgCl(2) at high concentrations) satisfied the cation requirement when added as sulfate salts. Although magnesium satisfied the cation requirement, a concentration effect, as well as an anion effect, was observed. In the presence of MgCl(2), little NADH(2) oxidation was observed at concentrations greater than 1 m. At lower concentrations, the rate of oxidation increased, reaching a maximal value at 0.1 m and remaining constant up to a concentration of 0.05 m MgCl(2). Magnesium acetate and MgSO(4) also replaced NaCl, and the maximal rate of oxidation occurred at 0.05 m with respect to magnesium. There was no change in the rate of oxidation at high magnesium acetate concentrations, whereas the rate of NADH(2) oxidation increased at higher concentrations of MgSO(4).  相似文献   

9.
To understand the pressure-adaptation mechanism of deep-sea enzymes, we studied the effects of pressure on the enzyme activity and structural stability of dihydrofolate reductase (DHFR) of the deep-sea bacterium Moritella profunda (mpDHFR) in comparison with those of Escherichia coli (ecDHFR). mpDHFR exhibited optimal enzyme activity at 50 MPa whereas ecDHFR was monotonically inactivated by pressure, suggesting inherent pressure-adaptation mechanisms in mpDHFR. The secondary structure of apo-mpDHFR was stable up to 80 °C, as revealed by circular dichroism spectra. The free energy changes due to pressure and urea unfolding of apo-mpDHFR, determined by fluorescence spectroscopy, were smaller than those of ecDHFR, indicating the unstable structure of mpDHFR against pressure and urea despite the three-dimensional crystal structures of both DHFRs being almost the same. The respective volume changes due to pressure and urea unfolding were − 45 and − 53 ml/mol at 25 °C for mpDHFR, which were smaller (less negative) than the corresponding values of − 77 and − 85 ml/mol for ecDHFR. These volume changes can be ascribed to the difference in internal cavity and surface hydration of each DHFR. From these results, we assume that the native structure of mpDHFR is loosely packed and highly hydrated compared with that of ecDHFR in solution.  相似文献   

10.
Abstract Cell-free enzyme preparations of the obligately anaerobic halophilic eubacterium Haloanaerobium praevalens synthesize fatty acids from malonyl-CoA. The reaction is stimulated by NaCl and KCl at a concentration of 1 M, and only slightly inhibited by salt concentrations as high as 3 M. Thus, the fatty acid synthetase of H. praevalens is expected to the fully active at the high intracellular salt concentrations present, and it is the first fatty acid synthetase reported to be active in the presence of high salt concentrations.  相似文献   

11.
The activity, stability and structure in solution of polypeptide elongation factor hEF-Tu from Halobacterium marismortui have been investigated. The protein is stable in aqueous solutions only at high concentrations of NaCl, KCl or ammonium sulphate, whereas it is more active in exchanging GDP at lower salt concentrations. It is more active and stable at lower pH values than is non-halophilic EF-Tu. The structure in solution of the protein was determined by complementary density, ultracentrifugation, dynamic light-scattering and neutron-scattering measurements. The protein has large hydration interactions, similar to those of other halophilic proteins: 0.4 (+/- 0.1) g of water and 0.20 (+/- 0.05) g of KCl associated with 1 g of protein, with a water/KCl mass ratio always remaining close to 2. The kinetics of inactivation at low salt concentrations showed a stabilizing effect of NaCl when compared to KCl. At low salt concentration, inactivation, protein unfolding and aggregation were strongly correlated. The results suggest that the stabilization model proposed for halophilic malate dehydrogenase by Zaccai et al., involving extensive protein interactions with hydrated salt ions, is also valid for hEF-Tu.  相似文献   

12.
The moderately halophilic bacterium Vibrio costicola accumulates α-aminoisobutyric acid (AIB) by active transport. Substantial amounts of Na+ ions are needed for this transport. This is not due to an ionic requirement for respiration; cells respire as well as KCl as in NaCl but do not transport AIB in KCl. In cells grown in the presence of 1.0 or 2.0 M NaCl, AIB transport took place in higher NaCl concentrations than in cells grown in the presence of 0.5 M NaCl. The latter cells developed salt-resistant transport when they were exposed to 1.0 M NaCl in the presence of chloramphenicol and other antibiotics that inhibit protein synthesis. Two levels of salt-resistant transport were observed. One level (resistance to 3.0 M NaCl) developed in 1.0 M NaCl without the addition of nutrients, did not seem to require an increase in internal solute concentration, and was not lost when cells grown in 1.0 M NaCl were suspended in 0.5 M NaCl. The second level (resistance to 4.0 M NaCl) developed in 1.0 M NaCl only when nutrients were added, may have required an increased internal solute concentration, and was lost when 1.0 M NaCl-grown cells were suspended in 0.5 M NaCl or KCl. Among the substances that stimulated the development of salt-resistant AIB transport, betaine was especially active. Furthermore, direct addition of betaine permitted cells to transport AIB at higher NaCl concentrations. High salt concentrations inhibited endogenous respiration to a lesser extent than AIB transport, especially in 0.5 M NaCl-grown cells. Thus, these concentrations of salt did not inhibit AIB transport by inhibiting respiration. However, oxidation of glucose and oxidation of succinate were at least as sensitive to high salt concentrations as AIB transport, suggesting that a salt-sensitive transport step(s) is involved in the oxidation of these substrates.  相似文献   

13.
Protease Formation by a Moderately Halophilic Bacillus Strain   总被引:2,自引:1,他引:1       下载免费PDF全文
A moderately halophilic strain of Bacillus, isolated from unrefined solar salt, was capable of growth in the presence of 4 M NaCl. Maximal growth was obtained in a medium containing 1 to 2 M NaCl. The organism produced protease when cultivated aerobically in media containing 0 to 3 M NaCl or 0 to 2 M KCl. The protease activity was optimal at 0.5 M NaCl and 0.75 M KCl.  相似文献   

14.
K Chan  O C Leung 《Microbios》1979,25(100):71-84
Chemically defined media have been developed for the growth of two moderately halophilic bacteria, Micrococcus morrhuae K-17 and Micrococcus luteus K-15. M. morrhuae K-17 grows well in a synthetic medium (SM-1) which contains a number of salts, 0.21 M KCl, 2 M NaCl, D-mannose, five vitamins and ten amino acids. The synthetic medium (SM-2) for M. luteus K-15 contains a number of salts, 0.21 M KCl, 1 M NaCl, D-fructose, nine vitamins and nine amino acids. Nutritional studies show that M. morrhuae K-17 can utilize a large number of organic compounds as carbon and energy source while the ability of M. luteus K-15 in utilizing the organic compounds is rather limited. The minimum salt requirement is 0.5 M NaCl for both strains when growth at the optimum temperature of 30 degrees C. However, this requirement can be lowered to 0.2 M in M. luteus K-15 when grown at a lower temperature of 25 degrees C. It is concluded that the ability to grow in a wider range of salt concentrations in response to temperature is species specific in moderate halophiles. The salt range for growth to occur can be extended when cells of both strains are grown in complex medium which might provide the amino acids and growth factors that cannot be synthesized by these strains at high salt concentrations.  相似文献   

15.
Résumé M. halodenitrificans possède la nitrate-réductase A. Cet enzyme se trouve sous la forme particulaire dans les extraits bruts. Il ne présente pas un caractère halophile: NaCl, KCl ou CsCl 1 ou 0,5 M ne l'activent pas. NaCl 1 M active cependant la réduction du nitrate en nitrite par les cellules en présence de lactate comme donneur d'électrons. Cet effet n'est pas dû à une action du sel sur la nitrate-réductase. Les cultures anaérobies avec nitrate synthétisent approximativement 7 fois plus d'enzyme que les cultures aérobies sans nitrate.
Bacterial nitrate reductasesVIII. Preliminary Study of the enzyme of Micrococcus halodenitrificans
Summary M. halodenitrificans has nitrate reductase A. This enzyme appears to be in particulate form in crude extracts. It does not present a halophilic character: 1 or 0.5 M NaCl, KCl, or CsCl does not activate it. However, 1 M NaCl activates the reduction of nitrate to nitrite by cells in the presence of lactate as electron donor. This effect is not due to an action of the salt on nitrate reductase. Anaerobic cultures containing nitrate form approximately 7 times more enzyme than aerobic cultures not containing nitrate.
  相似文献   

16.
17.
When subjected to the stress of growth in a relatively low-salt environment (1.25 M NaCl), the halophilic bacterium Halobacterium halobium induces a catalase. The protein has been purified to electrophoretic homogeneity and has an M(r) of 240,000 and a subunit size of approximately 62,000. The enzyme is active over a broad pH range of 6.5 to 10.0, with a peak in activity at pH 7.0. It has an isoelectric point of 4.0. This catalse, which is not readily reduced by dithionite, shows a Soret peak at 406 nm. Cyanide and azide inhibit the enzyme at micromolar concentrations, whereas maleimide is without effect. The addition of 20 mM 3-amino-1,2,4-triazole results in a 33% inhibition in enzymatic activity. The tetrameric protein binds NADP in a 1:1 ratio but does not peroxidize NADPH, NADH, or ascorbate. Although the enzymatic activity is maximal when assayed in a 50 mM potassium phosphate buffer with no NaCl, prolonged incubation in a buffer lacking NaCl results in inactive enzyme. Moreover, purification must be performed in the presence of 2 M NaCl. Equally as effective in retaining enzymatic function are NaCl, LiCl, KCl, CsCl, and NH4Cl, whereas divalent salts such as MgCl2 and CaCl2 result in the immediate loss of activity. The catalase is stained by pararosaniline, which is indicative of a glycosidic linkage. The Km for H2O2 is 60 mM, with inhibition observed at concentrations in excess of 90 mM. Thus, the mesohalic catalase purified from H. halobium seems to be similar to other catalases, except for the salt requirements, but differs markedly from the constitutive halobacterial hydroperoxidase.  相似文献   

18.
The sound velocities in polyelectrolyte solutions were measured at various concentrations of added salts. When aqueous solutions of tetra (n-butyl)ammonium polyacrylate were titrated with concentrated solutions of LiCl, NaCl, KCl or CsCl, the sound velocity, i.e., the adiabatic compressibility of the solution, did not change linearly with added salt concentration, but showed a breaking point. The degrees of counterion binding on polyacrylate ion estimated from the breaking points were 0.25-0.30, independent of cation species. In polystyrenesulfonate, moreover, no Na+ binding was detected from such sound velocity measurements.  相似文献   

19.
The folding and activity of halophilic enzymes are believed to require the presence of salts at high concentrations. When the inactivated nucleoside diphosphate kinase (NDK) from extremely halophilic archaea was incubated with low salt media, no activity was regained over the course of 8 days. When it was incubated with 2 M NaCl or 3 M KCl, however, it gradually regained activity. To our surprise, trimethylamine N-oxide (TMAO) also was able to induce activation at 4.0 M. The enzyme activity and secondary structure of refolded NDK in 4 M TMAO were comparable with those of the native NDK or the refolded NDK in 3.8 M NaCl. TMAO is not an electrolyte, meaning that the presence of concentrated salts is not an absolute requirement, and that charge shielding or ion binding is not a sole factor for the folding and activation of NDK. Although both NaCl and TMAO are effective in refolding NDK, the mechanism of their actions appears to be different: the effect of protein concentration and pH on refolding is qualitatively different between these two, and at pH 8.0 NDK could be refolded in the presence of 4 M TMAO only when low concentrations of NaCl are included.  相似文献   

20.
Using site-specific mutagenesis, we have constructed two mutants of Escherichia coli dihydrofolate reductase (ecDHFR) to investigate further the function of a weakly acidic side chain at position 27 in substrate protonation: Asp27-->Glu (D27E) and Asp27-->Cys (D27C). The crystal structure of D27E ecDHFR in a binary complex with methotrexate shows that the side-chain oxygen atoms of Glu27 are in almost precisely the same location as those of Asp27 in the wild-type enzyme. Kinetic evidence indicates that Glu27 can indeed function efficiently in the proton relay to dihydrofolate. Even though vertebrate DHFRs all have a glutamic acid at the structurally equivalent position, the kinetic properties of Glu27 ecDHFR more closely resemble those of wild-type bacterial DHFRs than of vertebrate DHFRs. The D27C mutation produced an enzyme still capable of relaying a proton to dihydrofolate, but with the intrinsic pKa in its pH-activity profiles shifted upward to values characteristic of the more basic thiolate group. The crystal structure of the binary complex with methotrexate reveals two unexpected features: (1) the Cys27 sulfhydryl group does not point toward the pteridine-binding site, but the side chain of this residue is instead rotated 120 degrees to interact with a tyrosine side chain projecting from a neighboring beta-strand; (2) a bound ethanol molecule occupies a cavity adjacent to methotrexate. Ethanol is a component of the crystallization medium.  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