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.
To determine the stage at which H+ ions are bound during the entropy-driven polymerization of tobacco mosaic virus protein, acid-base titrations were carried out at a concentration of 5 mg/ml in 0.1 m-KCl from pH 8 to pH 5.2 and back to pH 8 at 4, 10, 15 and 20 °C. The titration was always completely reversible when the addition of acid or base was so slow that the experiment required seven hours in each direction. When the titration was started at pH 7 and performed down and up twice as rapidly, a hysteresis loop, indistinguishable from one previously published, was obtained at 20 °C.Ultracentrifugation experiments were carried out at selected pH values at the four temperatures. H+ ion uptake, as determined from the reversible titration curves, is correlated with the disappearance of the 4 S component and is independent of whether the polymerized species is in a 20 S or higher state of aggregation. At pH 7, approximately 1 mole of H+ ion is bound per mole of monomer. At pH values between 6.56 and 6.05, 1.5 moles of H+ ion are bound per mole of monomer upon polymerization. At pH 6.05, 0.5 mole of H+ ion is bound before any polymerization takes place.Tobacco mosaic virus protein at 20 °C in an unbuffered 0.1 m-KCl solution at pH 7.18 at a concentration of 41 mg/ml, largely in the 20 S state, was depolymerized entirely to the 4 S state by dilution with 0.1 m-KCl adjusted to the same pH. Under these conditions, there was no pH change, indicating that no H + ions are released.These seemingly contradictory findings can be explained by assuming that the 4 S component polymerizes to form either double discs without binding H+ ions, or, alternatively, two-turn helices accompanied by the binding of H+ ions. Both double discs and two-turn helices sediment at approximately 20 S. Whether polymerization in the neighborhood of pH 7 leads to helices or discs depends upon the availability of H+ ions.  相似文献   

2.
3.
The polymerization of tobacco mosaic virus protein has been found to proceed through metastable states under conditions where initially one of the two polymerization-linked protons is bound. These metastable polymers have been characterized and are found to be helical rods, which resemble the structure of equilibrium helical rods that form when both polymerization-linked protons are bound. At pH 6.5 and 20 °C the true equilibrium distribution of these helical rods has been shown to consist of sedimenting species that are much smaller, 24 to 34 S, than described previously, 100 to 200 S. The larger, non-equilibrium rods are produced by an overshoot in polymerization that results from the slow formation of 20 S nuclei followed by a very rapid elongation reaction. Generally, this sequence of rate processes is sensitive to the rate at which a reaction is initiated. In the present case it is the rate of heating or the rate of change of the pH that determines the reaction path and therefore the rate of attainment of equilibrium. In addition to the formation of metastable helical rods during polymerization overshoot, metastable 20 S aggregates can form when either equilibrium or non-equilibrium helical rods are depolymerized by cooling to 5 to 7 °C at pH 6.5. These 20 S aggregates are presumably two-turn disks or helices and can serve as nuclei for helical rod formation in subsequent polymerization reactions. Both helical rod and 20 S metastability are extremely sensitive to pH but, under carefully controlled conditions, the metastability is quite reproducible and reproducible nucleation-controlled polymerization kinetics can be observed even when polymerization-depolymerization cycling is carried out between branches of a hysteresis loop. Temperature- or pH-induced polymerization of tobacco mosaic virus protein can be made to proceed by the slow formation of 20 S, two-turn helix, nuclei followed by the rapid addition of one or more species comprising the 4 S protein. These results confirm a previously proposed kinetic mechanism for the non-equilibrium polymerization reaction (Scheele &; Schuster, 1974).  相似文献   

4.
The bone morphogenetic property of bone matrix is degraded at 25 ° to 37 °C within 24 hours after a bone is removed from the body. The degradation occurs in the intact undemineralized bone from the action of endogenous enzymes, presumably neutral proteinases at pH optima of 7 · 0 to 7 · 4. Degradation is: more rapid at physiologic than at acid pH; heat inactivated in the range between 40 ° and 60 °C; slow at 2 °C over a period of 7 days in EDTA at pH 7 · 4. Degradation is inhibited by iodoacetic acid at concentrations as low as 3 · 0 mmoles per liter either in phosphate buffer or EDTA. Degradative activity of endogenous enzymes, as measured by the yield of bone from implants of matrix, is comparable to those obtained from matrix treated with trypsin at 15 °C, pH 7 · 6 over a period of 12 hours. These enzymes include a neutral proteinase (BMP-ase) which degrades bone morphogenetic protein (BMP) without mobilizing bone collagen hydroxy-proline as rapidly and as selectively as a specific functional entity. Observations on carboxypeptidase A and thermolysin cleavage of phenylalanine groups and data on acetylation of tyrosyl groups reducing bone yield suggest aromatic amino acids may be necessary for the biologically active conformation of BMP.  相似文献   

5.
Adult rat heart muscle cells obtained by perfusion of the heart with collagenase have been used to characterize the insulin receptors by equilibrium binding and kinetic measurements. Binding of 125I-labelled insulin to heart cells exhibited a high degree of specificity; it was dependent on pH and temperature, binding at steady increased with decreasing temperatures. About 70% of the radioactivity bound at equilibrium at 25°C could be dissociated by addition of an excess of unlabelled insulin. 54 and 40% of 125I-labelled insulin was degraded by isolated heart cells after 2 h at 37°C and 4 h at 25°C, respectively. This degrading activity was effectively inhibited by high concentration of albumin.Equilibrium binding studies were conducted at 25°C using insulin concentrations ranging from 2.5 · 10?11 mol/l to 10?6 mol/l. Scatchard analysis of the binding data resulted in a curvilinear plot (concave upward), which was further analyzed using the average affinity profile. The empty site affinity constant was calculated to be 9.5 · 107 l/mol with a total receptor concentration of 3.4 · 106 sites per cell.The presence of site-site interactions of the negative cooperative type among the insulin receptors has been confirmed by kinetic experiments. The rate of dilution induced dissociation was enhanced in the presence of native insulin (5 · 10?9 mol/l), both, under conditions of low and high fractional saturation of receptors.  相似文献   

6.
The effect of incubation temperature (2, 4, 6, 8 and 10° C) on haddock Melanogrammus aeglefinus development and growth during the embryonic period and in subsequent ontogeny in a common post‐hatch thermal environment (6° C) was investigated. Hatching times were inversely proportional to incubation temperature and ranged from 20·3 days at 2° C to 9·1 days at 10° C. Growth rates were directly proportional to incubation temperature during both the embryonic and larval periods. There was a significant decline in growth rates following hatch in all temperature groups. Compared to the endogenously feeding embryos, growth rates in the exogenous period declined by 4·4‐fold at 4° C to 3·9‐fold at 8° C, indicative of the demarcation between the endogenous and exogenous feeding periods. Yolk utilization varied from 17 days at 2° C to 6 days at 10° C and followed a three‐stage sigmoidal pattern with the initial lag period inversely proportional to incubation temperature. Time to 50% yolk depletion varied inversely with temperature but occurred 1–1·5 days post‐hatch at all temperatures. Additionally, the period between 10 and 90% yolk depletion also decreased with increased temperature. Overall developmental rate was sequential with and directly proportional (2·3‐fold increase) to incubation temperature while the time spent in each developmental stage was inversely proportional to temperature. Larger embryos tended to be produced at lower temperatures but this pattern reversed following hatch, as larvae from higher temperature groups grew more rapidly than those from other temperature groups. Larvae from all temperatures achieved a similar length (c.total length 4·5 mm) upon complete yolk absorption. The study demonstrated the significant impact that temperature has upon developmental and growth rates in both endogenous and exogenous feeding periods. It also illustrated that temperature changes during embryogenesis had significant and persistent effects on growth in subsequent ontogeny.  相似文献   

7.
The metastable conformational states which underlie the hysteresis displayed by Escherichia coli ribosomal RNA in its pH titration in the acid range have been analyzed in terms of acid-stable RNA secondary structures. Sedimentation measurements show that the phenomenon is intramolecular, so that analysis of the hysteresis loops can, in principle, reveal details of molecular architecture. Hysteresis cycles obtained spectrophotometrically and potentiometrically were compared for RNA in solutions of different ionic strengths and ionic compositions. The effect is much smaller at lower ionic strength and disappears in the absence of magnesium ions. The curve followed upon addition of acid appears to reflect the equilibrium state of the system at each pH value. On the “base branch” of the loop, a slow absorbance change (complete in hours) was observed after the pH was raised by addition of a portion of base. This slow process is attributed to the annealing of “mismatched” multihelical regions of the ribosomal RNA. Certain regions, however, remain in metastable configurations for days and it is these long-lived non-equilibrium structures that underlie the hysteresis. Titration at 35 °C gave hysteresis loops of the same size and shape as at 20 °C; indeed, we found that the metastabilities are not removed even at 80 °C. Ultraviolet light absorbance difference spectra at 80 °C between solutions at the same pH, but on different branches of the cycle, give insight into the nature of the metastable conformation(s).Our experimental observations lead us to propose that the hysteresis is due to the formation at acidic pH of double-helical structures involving protonated guanine and adenine base pairs. The G.G pairs seem especially important to account for the very high thermal stability, as well as for the fact that the structures formed at a given pH value as acid is added dissociate only at higher pH values when the solution is titrated with base. Titrations of transfer RNA, along with literature data on 16 S rRNA primary structure, imply that the metastable regions in rRNA may consist of perhaps 10 to 15 base pairs.  相似文献   

8.
The rates of formation and dissociation of concanavalin A with some 4-methylumbelliferyl and p-nitrophenyl derivatives of α- and β-D-mannopyranosides and glucopyranosides were measured by fluorescence and spectral stopped-flow methods. All process examined were uniphasic. The second-order formation rate constants varied only from 6.8 · 104 to 12.8 · 104 M?. s?1, whereas the first-order dissociation rate constants ranged from 4.1. to 220 s?1, all at ph 5.0, I = 0.3 M, and 25°C. Dissociation rates thus controlled the value of binding constant. The effect of temperature on these reactions was examined, from which enthalpies and entropies of activation and of reaction could be calculated. The effects of pH at 25°C on the reaction rates of 4-methylumbelliferyl α-D-mannopyranoside and 4-methylumbelliferyl α-D-glucopyranoside with concanavalin A were examined. The value of the binding constant Kap (derived from the kinetics) at any pH could be related to the intrinsic binding constant K by the expression Kap = KaK(Ka + [H+])?1. The values of Ka, the ionization constant of the protein segment responsive to sugar binding, were 3 · 10?4 M and 1 · 10?4 M for 4-methylumbelliferyl α-D-mannopyranoside and 4-methylumbelliferyl α-D-glucopyranoside, respectively. The binding constant of p-nitrophenyl α-D-mannopyranoside is surprisingly much less sensitive to a pH change from 5.0 to 2.7. Ionic strength had little effect on the binding characteristics of 4-methylumbelliferyl α-D-mannopyranoside to concanavalin A at pH 5.2 and 25°C.  相似文献   

9.
Two different α‐glucosidase‐producing thermophilic E134 strains were isolated from a hot spring in Kozakli, Turkey. Based on the phenotypic, phylogenetic and chemotaxonomic evidence, the strain was proposed to be a species of G. toebii. Its thermostable exo‐α‐1,4‐glucosidases also were characterized and compared, which were purified from the intracellular and extracellular fractions with estimated molecular weights of 65 and 45 kDa. The intracellular and extracellular α‐glucosidases showed optimal activity at 65 °C, pH 7·0, and at 70 °C, pH 6·8, with 3·65 and 0·83 Km values for the pNPG substrate, respectively. Both enzymes remained active over temperature and pH ranges of 35–70 °C and 4·5–11·0. They retained 82 and 84% of their activities when incubated at 60 °C for 5 h. Their relative activities were 45–75% and 45–60% at pH 4·5 and 11·0 values for 15 h at 35 °C. They could hydrolyse the α‐1,3 and α‐1,4 bonds on substrates in addition to a high transglycosylation activity, although the intracellular enzyme had more affinity to the substrates both in hydrolysis and transglycosylation reactions. Furthermore, although sodium dodecyl sulfate behaved as an activator for both of them at 60 °C, urea and ethanol only increased the activity of the extracellular α‐glucosidase. By this study, G. toebii E134 strain was introduced, which might have a potential in biotechnological processes when the conformational stability of its enzymes to heat, pH and denaturants were considered. Copyright © 2011 John Wiley & Sons, Ltd.  相似文献   

10.
The freeze tolerant wood frog Rana sylvatica was studied to determine the impact of the freezing and thawing of this frog on the activity of γ-glutamyltranspeptidase in the liver. On exposure to ?2·5°C, for 1, 12 and 24 h, frogs were found to be cool, covered with ice crystals and frozen, respectively. Thawing for 24 h at 4°C recovered the frogs completely. A 45 per cent decrease in the liver weight: body weight ratio was notable after 1 h at ?2·5°C, suggestive of an early hepatic capacitance response. A glycemic response to freezing was observed: blood glucose levels exhibited a 55 per cent decrease after 1 h at ?2·5°C on cooling; a 10·5-fold increase after 12 h at ?2·5°C on the initiation of freezing; and a 22-fold increase after 24 h at ?2·5°C in the fully frozen state. Blood glucose levels remained elevated four-fold in the thawed state. Plasma insulin levels were increased twofold in the frozen state and 1·8-fold in the thawed state, while plasma ketone levels were increased 1·8-fold in the frozen state and 1·5-fold in the thawed state. Plasma total T3 levels were decreased by 22 per cent in the frozen state and normalized on thawing. In homogenates and plasma membranes isolated from the livers of Rana sylvatica, the activity of γ-glutamyltranspeptidase was found to be elevated at all stages of the freeze–thaw process. After 1, 12 and 24 h at ?2·5°C, activities were increased 2·5-, 2·3-, 2·4-fold respectively in the homogenates and 2·5-, 2·2-, 2·4-fold respectively in the plasma membranes. After thawing, activities were still increased 1·9-fold in both homogenates and plasma membranes. In homogenates prepared from the kidneys of Rana sylvatica, the activity of γ-glutamyltranspeptidase was increased 1·4-fold after 1 h at ?2·5°C after which it returned to normal. The role of thyroid hormone in producing the increase in γ-glutamyltranspeptidase in the liver of Rana sylvatica in response to freezing is discussed as is the significance of the enzyme increase in terms of hepatic cytoprotection and freeze tolerance.  相似文献   

11.
The pH and ionic strength dependence of conformation of the COOH-terminal fragment 206–316 (fragment FII) of thermolysin was monitored by far-uv CD and difference absorption measurements. This fragment was shown previously to possess the properties of a protein domain, i.e., able to refold into a stable nativelike structure [Fontana, A., Vita, C. & Chaiken, I. M. (1983) Biopolymers 22 , 69–78]. Analysis of the CD spectra in the pH range of 1–12 indicated that near pH 1, the conformation of fragment FII appears to be in an intermediate state (H) between the fully unfolded one (U) [the guanidine hydrochloride (Gdn · HCl)-induced unfolded state] and the nativelike state (N—that attained at neutral pH). Quantitative analysis of secondary structure from CD spectra revealed that state H at 4°C is characterized by some 30% α-helical structure, compared to 47% for state N. The heat- and Gdn · HCl-mediated unfolding transitions of state H were fully reversible and characterized by little cooperativity, which is taken as an indication that state H corresponds to several species possessing different, and low, conformational stabilities. The midpoint transition from state H to N occurs near pH 2.5, implying that the acid transition results from the titration of carboxyl groups of the fragment with anomalously low pK, as would be expected for groups involved in specific salt bridges. Fragment FII at pH 1 (state H) may be induced to exhibit nearly the same degree of helicity of state N simply by increasing the ionic strength of the solution, thus reducing the repulsive interactions between positive charges within the highly charged fragment at pH 1. The results obtained emphasize the role of electrostatic interactions in the folding and stability of fragment FII and suggest a mechanism of folding of the fragment from U to N involving an intermediate state characterized by an assembly of fluctuating α-helices.  相似文献   

12.
Psychrotrophic bacteria are known to occur in temperate, constantly cold, and artificially cooled environments. This is the first report of their occurrence in a constantly warm (ca. 24°–35°C) tropical environment. Soil samples taken from two sites along the southeastern coastal zone of Jamaica yielded growth of psychotrophic bacteria after 3–4 weeks of enrichment culture in 1/30 strength tryptic soy broth, 20 mg L−1 cycloheximide at 2°C. Growth of individual isolates at 2°C was confirmed. Isolates include aerobic and fermentative Gram-negative rods and sporeforming (Bacillus sp.) and non-sporeforming (Aureobacterium sp.) Gram-positive rods. We determined the effect of temperature on growth rate in four isolates. Strain Y1 has an unusually wide temperature range for growth, 2°–44°C, resembling that of Listeria monocytogenes. In strain R1 the optimum temperature for growth occurred unusually near the maximum temperature for growth. Strains R2 and Y2 displayed cardinal temperatures typical of known psychotrophs but appear to have evolved enhanced growth potential near the optimum temperature in response to a constantly warm environment. Received: 30 April 1997 / Accepted: 9 August 1997  相似文献   

13.
We determined the effects of metabolic inhibitors on 125I-labeled prolactin binding in monolayers of cultured rat mammary tumors. Chemical agents that blocked energy production increased binding by 8–20 fold, as did lowering the temperature from 37°C to 4°C. This difference was not due to blocking degradation of the hormone and inhibitors of degradation (lysosomotropic amines, bacitracin) did not increase binding. In the presence of a metabolic inhibitor at 37°C, binding reached a steady state within 3 h and had an apparent dissociation constant of ~6 × 10?10 M. Studies with fresh tumor slices produced comparable results. The findings indicate that the level of metabolic energy in mammary tumor cells can regulate prolactin binding.  相似文献   

14.
The critical thermal maxima (TMAX) of threadfin shad Dorosoma petenense exposed to standardized stress (30 s handling in a dip‐net), simulating stressors endured during fish loading before transport, were measured over a range of holding temperatures (15, 20 and 25° C). Dorosoma petenense TMAX showed a significant thermal effect, displaying mean ±s.d . critical thermal maxima of 26·5 ± 1·6, 30·9 ± 1·2 and 33·3 ± 1·4° C, when tested at temperatures of 15, 20 and 25° C, respectively. Dorosoma petenense TMAX levels were also affected by stress, with handled fish showing significantly lower values than control fish exposed to 15 (mean ±s.d . TMAX = 25·6 ± 2·0° C), 20 (27·6 ± 2·8° C) and 25° C (32·0 ± 2·6° C). In addition to providing basic information on D. petenense thermal tolerance, experimental results suggest that fishery managers should consider the whole suite of potential stressors, such as air exposure during handling and fish loading, when developing management criteria.  相似文献   

15.
Methods were developed for obtaining highly viable mouse hepatocytes in single cell suspension and for maintaining the hepatocytes in adherent static culture. The characteristics of transferrin binding and iron uptake into these hepatocytes was investigated. (1) After attachment to culture dishes for 18–24 h hepatocytes displayed an accelerating rate of iron uptake with time. Immediately after isolation mouse hepatocytes in suspension exhibited a linear iron uptake rate of 1.14·105molecules/cell per min in 5 μM transferrin. Iron uptake also increased with increasing transferrin concentration both in suspension and adherent culture. Pinocytosis measured in isolated hepatocytes could account only for 10–20% of the total iron uptake. Iron uptake was completely inhibited at 4°C. (2) A transferrin binding component which saturated at 0.5 μM diferric transferrin was detected. The number of specific, saturable diferric transferrin binding sites on mouse hepatocytes was 4.4·104±1.9·104 for cells in suspension and 6.6·104±2.3·104 for adherent cultured cells. The apparent association constants were 1.23·107 1·mol?1 and 3.4·106 1·mol?1 for suspension and cultured cells respectively. (3) Mouse hepatocytes also displayed a large component of non-saturable transferrin binding sites. This binding increased linearly with transferrin concentration and appeared to contribute to iron uptake in mouse hepatocytes. Assuming that only saturable transferrin binding sites donate iron, the rate of iron uptake is about 2.5 molecules iron/receptor per min at 5 μM transferrin in both suspension and adherent cells and increases to 4 molecules iron/receptor per min at 10 μM transferrin in adherent cultured cells. These rates are considerably greater than the 0.5 molcules/receptor per min observed at 0.5 μM transferrin, the concentration at which the specific transferrin binding sites are fully occupied. The data suggest that either the non-saturable binding component donates some iron or that this component stimulates the saturable component to increase the rate of iron uptake. (4) During incubations at 4°C the majority of the transferrin bound to both saturable and nonsaturable binding sites lost one or more iron atoms. Incubations including 2 mM α,α′-dipyridyl (an Fe11 chelator) decreased the cell associated 59Fe at both 4 and 37°C while completely inhibiting iron uptake within 2–3 min of exposure at 37°C. These observations suggest that most if not all iron is loosened from transferrin upon interaction of transferrin with the hepatocyte membrane. There is also greater sensitivity of 59Fe uptake compared to transferrin binding to pronase digestion, suggesting that an iron acceptor moiety on the cell surface is available to proteolysis.  相似文献   

16.
Binding of 1,N6-ethanoadenosine triphosphate to actin   总被引:3,自引:0,他引:3  
G-actin is known to bind one molecule of ATP. Its polymerization to F-actin is accompanied by the splitting off of the terminal phosphate of the bound nucleotide. We have found that the fluorescent 1,N6-ethanoadenosine triphosphate (?ATP) can substitute for ATP in G-actin and that G-actin containing bound ?ATP possesses essentially full polymerizability. The binding of this ATP analog has been studied by following the inactivation of the ?ATP·G-actin complex. The binding constant (4?5.7 × 106 M?1) obtained in the absence of EDTA is about 50% of that for ATP, while the binding constant obtained in the presence of EDTA (0.9?3.0 × 105 M?1) is comparable to those for ATP and ADP. These findings suggest that ?ATP can be used as a structural probe for actin. The fluorescence lifetime of ?ATP bound to G·actin is 36 nsec. The rotational relaxation time of ?ATP·G-actin is near 60 nsec. at 20°C.  相似文献   

17.
Effects of sodium tungstate on the steroid-binding properties of hen oviduct progesterone receptor were examined and were found to be pH-dependent. When freshly prepared hen oviduct cytosol containing progesterone receptor was heated at 37°C for 20 min, its ability to bind [3H]progesterone decreased to 20% level of unheated samples. At pH 7, presence of 2–3 mM tungstate during the above incubation period reduced this loss of binding. At higher tungstate concentrations (>5 mM), this stabilizing effect was gradually abolished. Similar results were obtained with preparations that contained [3H]progesterone-receptor complexes; 70–80% of which remained after a 20 min incubation at 37°C in the presence of 2–3 mM tungstate at pH 7. At pH 8, presence of tungstate (1–10 mM) during the 37°C incubation stabilized both the steroid-bound and the unoccupied progesterone receptor in a concentration-dependent manner. The extent of steroid binding by the receptor at 4°C remained unchanged in the presence of up to 10 mM tungstate at both pH 7 and pH 8 assay conditions while presence of 20 mM tungstate lowered this binding capacity. These results indicate that tungstate effects may be mediated via its interaction with the progesterone receptor.  相似文献   

18.
The storage stability of alcohol dehydrogenase from yeast has been considerably improved by the use of additives. Glycerol is an effective cryoprotectant at ?196 and ?20°C. At 4°C, glycerol and dl-α-glycerophosphate are stabilizers, while at 30°C dl-α-glycerophosphate, 3-phosphoglyceric acid, phosphocreatine, 6-phosphogluconic acid, phosphoarginine, phosphoserine and sucrose are examples of stabilizers. While no single mechanism can be adduced, stabilization by NAD+ and 5′-AMP may be attributed to binding at the active centre. Cryoprotection by glycerol is attributed to maintenance of the pH within the range at which the enzyme is stable. dl-α-Glycerophosphate caused a shift of 16°C in the transition temperature of the enzyme, as measured by differential scanning calorimetry.  相似文献   

19.
20.
A C Anusiem 《Biopolymers》1975,14(6):1305-1307
Measurements of the magnetic susceptibility of guinea pig, human A, and pigeon ferrihemoglobins as a function of pH at 30°C demonstrate maxima very near the extrema found for the standard enthalpies and standard entropies of ligand binding, i.e., at the so-called “characteristic pH” values, which are very different for these three species. Ligand binding is dependent on spin state, at least through the characteristic pH behavior, which is not yet understood.  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