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 187 毫秒
1.
Catabolic dehydroquinase which functions in the inducible quinic acid catabolic pathway in Neurospora crassa has been purified 8000-fold. The enzyme was purified by two methods. One used heat denaturation of contaminating proteins; the other used antibody affinity chromatography. The preparations obtained by these two methods were identical by all criteria. The purified enzyme is extremely resistant to thermal denaturation as well as denaturation 0y urea and guanidine hydrochloride at 25 degrees. It is irreversibly inactivated, although not efficiently dissociated, by sodium dodecyl sulfate and guanidine hydrochloride at 55 degrees. At pH 3.0, the enzyme is reversibly dissociated into inactive subunits. At high concentrations catabolic dehydroquinase aggregates into an inactive, high molecular weight complex. The native enzyme, which has a very high specific activity, has a molecular weight of approximately 220,000 and is composed of identical subunits of 8,000 to 12,000 molecular weight each. The native enzyme and the subunit are both asymmetric.  相似文献   

2.
The effects of various concentrations of urea and guanidine hydrochloride on enzyme activity and on subunit association were determined. Incubation of thymidylate synthetase with buffered solutions of 3M to 3.5M guanidine hydrochloride or 5 M to 6 M urea resulted in the loss of about 90% of the enzyme activity. Under these denaturing conditions a red shift of the fluorescence emission maximum from 340 nm to 351 nm was observed together with a significant decrease in the relative fluorescence intensity of the protein. Studies at both 4 degrees C and 25 degrees C indicated that the enzyme was in the dimer form in 2 M guanidine hydrochloride but was dissociated into monomers in concentrations of this denaturant of 3 M and above. Although only monomeric species were evident at 4 degrees C in 6 M urea, at 25 25 degrees C this denaturant caused protein aggregation which increased with decreasing phosphate buffer concentration. Enzyme (5 mg/ml) in 0.5 M potassium phosphate buffer, pH 6.8, containing 4 M guanidine hydrochloride gave a minimum S20, w value of 1.22S at 25 degrees C. Sedimentation behavior of the native enzyme in the range of 5 to 20 mg/ml was only slightly concentration-dependent (4.28 S to 4.86 S) but extensive aggregation occurred above 20 mg/ml.  相似文献   

3.
The pH dependence of the reversible guanidine hydrochloride denaturation of the major fraction of ovalbumin (ovalbumin A1) was studied by a viscometric method in the pH range 1-7, at 25 degrees C and at six different denaturant concentrations (1.5-2.6 M). At any denaturant concentrationa reduction in pH favoured the transition from the native to the denatured state. The latter was essentially 'structureless', as revealed by the fact that the reduced viscosity of the acid and guanidine hydrochloride denatured state of ovalbumin A1 (obtained at different denaturant concentrations in acidic solutions) was measured (at a protein concentration of 3.8 mg/ml) to be 29.2 ml/g which is identical to that found in 6 M guanidine hydrochloride wherein the protein behaves as a cross-linked random coil. A quantitative analysis of the results on the pH dependence of the equilibrium constant for the denaturation process showed that on denaturation the intrinsic pK of two carboxyl groups in ovalbumin A1 went up from 3.1 in the native state to 4.4 in the denatured state of the protein.  相似文献   

4.
Effect of pH, urea, and guanidine hydrochloride on the activity and structure of buffalo spleen cathepsin B was investigated. At alkaline pH, there was an irreversible loss of the structure as well as the activity of the buffalo enzyme. At acidic pH, however, the inactivation of the enzyme was reversible. The enzyme reversibly lost most of its activity at denaturant concentrations which did not cause a significant change in its secondary structure. The inactivation could be attributed to minor perturbations in the environment of the amino acid residue(s) at and/or around the active site of the enzyme. High urea/guanidine hydrochloride concentrations leading to the structural changes in cathepsin B made the inactivation process irreversible.  相似文献   

5.
The cell surface glycoprotein of Halobacterium halobium is soluble in dilute buffer at neutral pH. At low counterion concentrations, the protein is monomeric (Ms,D = 209 kDa) and exhibits the characteristics of a highly charged polyelectrolyte. Evidence obtained from intrinsic fluorescence and far-UV circular dichroism shows that the monomer at low salt loses both its native conformation and its inherent tendency to form high molecular mass assemblies. In 4M NaCl, 25 mM KCl, and in the presence of divalent ions (greater than or equal to 50mM Mg2+ or Ca2+), association to well-defined assemblies of up to approximately 4 X 10(6) Da occurs. At low Mg2+ concentration and in the presence of Ba2+, a wide size-distribution of aggregates is observed. The assembly pattern of the protein may be correlated with salt-dependent alterations in the morphology of the bacterium.  相似文献   

6.
Protein stabilization by urea and guanidine hydrochloride   总被引:7,自引:0,他引:7  
Bhuyan AK 《Biochemistry》2002,41(45):13386-13394
The urea, guanidine hydrochloride, salt, and temperature dependence of the rate of dissociation of CO from a nonequilibrium state of CO-bound native ferrocytochrome c has been studied at pH 7. The heme iron of ferrocytochrome c in the presence of denaturing concentrations of guanidine hydrochloride (GdnHCl) and urea prepared in 0.1 M phosphate, pH 7, binds CO. When the unfolded protein solution is diluted 101-fold into CO-free folding buffer, the protein chain refolds completely, leaving the CO molecule bonded to the heme iron. Subsequently, slow thermal dissociation of the CO molecule yields to the heme coordination of the native M80 ligand. Thus, the reaction monitors the rate of thermal conversion of the CO-liganded native ferrocytochrome c to the M80-liganded native protein. The rate of this reaction, k(diss), shows a characteristic dependence on the presence of nondenaturing concentrations of the denaturants in the reaction medium. The rate decreases by approximately 1.9-3-fold as the concentration of GdnHCl in the refolding medium increases from nearly 0 to approximately 2.1 M. Similarly, the rate decreases by 1.8-fold as the urea concentration is raised from 0.l to approximately 5 M. At still higher concentrations of the denaturants the denaturing effect sets in, the protein is destabilized, and hence the CO dissociation rate increases sharply. The activation energy of the reaction, E(a), increases when the denaturant concentration in the reaction medium is raised: from 24.1 to 28.3 kcal mol(-1) for a 0.05-2.1 M rise in GdnHCl and from 25.2 to 26.9 kcal mol(-1) for a 0.1-26.9 M increase in urea. Corresponding to these increases in denaturant concentrations are also increases in the activation entropy, S(diss)/R, where R is the gas constant of the reaction. The denaturant dependence of these kinetic and thermodynamic parameters of the CO dissociation reaction suggests that binding interactions with GdnHCl and urea can increase the structural and energetic stability of ferrocytochrome c up to the limit of the subdenaturing concentrations of the additives. NaCl and Na(2)SO(4), which stabilize proteins through their salting-in effect, also decrease the rate with a corresponding increase in activation entropy of CO dissociation from CO-bound native ferrocytochrome c, lending support to the view that low concentrations of GdnHCl and urea stabilize proteins. These results have direct relevance to the understanding and interpretation of the free energy-denaturant relationship and protein folding chevrons.  相似文献   

7.
The denaturation of ribonuclease A by guanidine hydrochloride, lithium bromide, and lithium chloride and by mixed denaturants consisting of guanidine hydrochloride and one of the denaturants lithium chloride, lithium bromide, and sodium bromide was followed by difference spectral measurements at pH 4.8 and 25 degrees C. Both components of mixed denaturant systems enhance each other's effect in unfolding the protein. The effect of lithium bromide on the midpoint of guanidine hydrochloride denaturation transition is approximately the sum of the effects of the constituent ions. For all the mixed denaturants tested, the dependence of the free energy change on denaturation is linear. The conformational free energy associated with the guanidine hydrochloride denaturation transition in water is 7.5 +/- 0.1 kcal mol-1, and it is unchanged in the presence of low concentrations of lithium bromide, lithium chloride, and sodium bromide which by themselves are not concentrated enough to unfold the protein. The conformational free energy associated with the lithium bromide denaturation transition in water is 11.7 +/- 0.3 kcal mol-1, and it is not affected by the presence of low concentrations of guanidine hydrochloride which by themselves do not disrupt the structure of native ribonuclease A.  相似文献   

8.
Phenolic titration of ovalbumin was performed in the pH range 7-12 at 30 degrees C and at three ionic strengths viz. 0.033, 0.133 and 0.200. The conformational integrity of ovalbumin was studied by viscosity measurements at different pH values in the pH range 7-12.4. At ionic strength 0.133 two phenolic groups titrated reversibly with pKint = 10.31, and w = 0.032 up to pH 11.25 under native conditions. The value of w expectedly decreased with increase in ionic strength. Two additional phenolic groups became available for reversible titration between pH 11.25 and 11.95 after some conformational change. Above pH 12, the phenolic titration became irreversible and all of the nine tyrosine residues were titrated at pH 13.3 Exposure of ovalbumin to alkaline pH (12.4) caused considerable disruption of the native protein conformation. The reduced viscosity increased from 4.2 ml/g at pH 7.0 to 16.8 ml/g at pH 12.4 under identical conditions of the protein concentration. All of the nine tyrosyl groups of ovalbumin were titrated normally (pKint = 9.9) in a mixture of 5 M guanidine hydrochloride and 1.2 M urea. However, even in this mixture electrostatic interaction, as measured by w was not completely abolished.  相似文献   

9.
Sasahara K  Yagi H  Sakai M  Naiki H  Goto Y 《Biochemistry》2008,47(8):2650-2660
Amyloid nucleation through agitation was studied with beta2-microglobulin, which is responsible for dialysis-related amyloidosis, in the presence of salt under acid and neutral pH conditions. First, the aggregation of beta2-microglobulin in NaCl solutions was achieved by mildly agitating for 24 h at 37 degrees C protein solutions in three different states: acid-unfolded, salt-induced protofibrillar, and native. The formation of aggregates was confirmed by an increase in light scattering intensity of the solutions. Then, the aggregated samples were incubated without agitation at 37 degrees C for up to 25-45 days. The structural changes in the aggregated state during the incubation period were examined by means of fluorescence spectroscopy with thioflavin T, circular dichroism spectroscopy, and electron microscopy. The results revealed that all the samples in the different states produced a mature amyloid nucleus upon agitation, after which the fibrils elongated without any detectable lag phase during the incubation, with the acid-unfolded protein better suited to undergoing the structural rearrangements necessary to form amyloid fibrils than the more structured forms. The amount of aggregate including the amyloid nucleus produced by agitation from the native conformation at neutral pH was estimated to be about 9% of all the protein by an analysis using ultracentrifugation. Additionally, amyloid nucleation by agitation was similarly achieved for a different protein, hen egg-white lysozyme, in 0.5 M NaCl solution at neutral pH. Taken together, the agitation-treated aggregates of both proteins have a high propensity to produce an amyloid nucleus even at neutral pH, providing evidence that the aggregation pathway involves amyloid nucleation under entirely native conditions.  相似文献   

10.
Two homologous fungal short-chain dehydrogenase/reductase (SDR) proteins have been cloned from the fungus Curvularia lunata (teleomorph: Cochliobolus lunatus) and expressed in Escherichia coli: trihydroxynaphthalene reductase (3HNR), an enzyme of the melanin biosynthetic pathway that catalyzes the conversion of 1,3,8-trihydroxynaphthalene to vermelone, and 17beta-hydroxysteroid dehydrogenase (17beta-HSDcl), which acts on androgens and estrogens, although its physiological substrate remains to be defined. In the present study, we have compared the structures, specificities to substrates and inhibitors, temperature and pH optima of 3HNR and 17beta-HSDcl. Sequence analysis and homology-built models revealed that these enzymes are highly similar. Both of these enzymes are NADP(H)-preferring reductases and act on steroids at position 17; however, 17beta-HSDcl presented considerably higher initial rates than 3HNR. In vitro, 17beta-HSDcl preferably catalyzed the reduction of 4-estrene-3,17-dione, while the best steroid substrate for 3HNR was 5alpha-androstane-3,17-dione. On the other hand, 2,3-dihydro-2,5-dihydroxy-4H-benzopyran-4-one (DDBO), an artificial substrate of 3HNR, was oxidized rapidly by 3HNR, while it was not a substrate for 17beta-HSDcl. Additionally, our data show that tricyclazole, a specific inhibitor of 3HNR, is 100-fold less effective for 17beta-HSDcl inhibition, while flavonoids can inhibit both 3HNR and 17beta-HSDcl. We have also examined the effects of temperature and pH on the oxidation of DDBO by 3HNR and the oxidation of 4-estrene-17beta-ol-3-one by 17beta-HSDcl. The apparent optimal temperature for 3HNR activity was between 25 and 30 degrees C, while it was between 40 and 45 degrees C for 17beta-HSDcl activity. The pH optimum of 3HNR activity was between 8 and 9, and for 17beta-HSDcl, between 7 and 8. Our data show that in spite of high homology and similar backbone structure, differences between 3HNR and 17beta-HSDcl were not only in substrate specificities, but also in temperature and pH optima.  相似文献   

11.
Thermal unfolding of chicken pancreatic polypeptide at two different concentrations was studied at various pH values. The thermal stability was higher at higher protein concentrations. The transition temperatures at two different protein concentrations changed with pH in parallel and decreased by about 30 degrees C on lowering pH from 5 to 2. The results on the thermal unfolding were analyzed by assuming that the dimerization constant is independent of pH, that the thermal unfolding occurs only after the pancreatic polypeptide dimers dissociated into the monomers, and that one ionizable group participates in the acid unfolding of the monomer. The free energy change for the unfolding of the pancreatic polypeptide monomer was estimated to be 1.4 kcal/mol. The unfolding of pancreatic polypeptide by guanidine hydrochloride at pH 6.0 and 25 degrees C was also studied. The stability to guanidine hydrochloride was higher at higher protein concentrations.  相似文献   

12.
The effect of pH on the conformation of ricin and its A- and B-chains has been studied by measuring their intrinsic fluorescence. At pH 5.0 and 7.5, the structural stability of toxin and subunits was estimated according to the denaturing action of guanidine hydrochloride. It was demonstrated that the fluorescence of native toxin and catalytic A-subunit does not depend significantly on pH in the range pH 3-8, whereas ricin B-chain undergoes a structural transition at pH less than 5.0. The structural stability of ricin and isolated chains differs significantly at pH 7.5 and 5.0; the structural stability of ricin and the A-chain increases, whereas that of the B-chain decreases.  相似文献   

13.
Sridevi K  Udgaonkar JB 《Biochemistry》2002,41(5):1568-1578
The folding and unfolding rates of the small protein, barstar, have been monitored using stopped-flow measurements of intrinsic tryptophan fluorescence at 25 degrees C, pH 8.5, and have been compared over a wide range of urea and guanidine hydrochloride (GdnHCl) concentrations. When the logarithms of the rates of folding from urea and from GdnHCl unfolded forms are extrapolated linearly with denaturant concentration, the same rate is obtained for folding in zero denaturant. Similar linear extrapolations of rates of unfolding in urea and GdnHCl yield, however, different unfolding rates in zero denaturant, indicating that such linear extrapolations are not valid. It has been difficult, for any protein, to determine unfolding rates under nativelike conditions in direct kinetic experiments. Using a novel strategy of coupling the reactivity of a buried cysteine residue with 5,5'-dithiobis(2-nitrobenzoic acid) (DTNB) to the unfolding reaction of barstar, the global unfolding and refolding rates have now been determined in low denaturant concentrations. The logarithms of unfolding rates obtained at low urea and GdnHCl concentrations show a markedly nonlinear dependence on denaturant concentration and converge to the same unfolding rate in the absence of denaturant. It is shown that the native protein can sample the fully unfolded conformation even in the absence of denaturant. The observed nonlinear dependences of the logarithms of the refolding and unfolding rates observed for both denaturants are shown to be due to the presence of (un)folding intermediates and not due to movements in the position of the transition state with a change in denaturant concentration.  相似文献   

14.
The destabilizing effect of electrostatic repulsions on protein stability has been studied by using synthetic two-stranded alpha-helical coiled-coils as a model system. The native coiled-coil consists of two identical 35-residue polypeptide chains with a heptad repeat QgVaGbAcLdQeKf and a Cys residue at position 2 to allow formation of an interchain disulfide bridge. This peptide, designed to contain no intrahelical or interhelical electrostatic interactions, forms a stable coiled-coil structure at 20 degrees C in benign medium (50 mM KCl, 25 mM PO4, pH 7) with a [urea]1/2 value of 6.1 M. Four mutant coiled-coils were designed to contain one or two Glu substitutions for Gln per polypeptide chain. The resulting coiled-coils contained potential i to i' + 5 Glu-Glu interchain repulsions (denoted as peptide E2(15,20)), i to i' + 2 Glu-Glu interchain repulsions (denoted E2(20,22)), or no interchain ionic interactions (denoted E2(13,22) and E1(20)). The stabilities of the coiled-coils were determined by measuring the ellipticities at 222 nm as a function of urea or guanidine hydrochloride concentration at 20 degrees C in the presence and absence of an interchain disulfide bridge. At pH 7, in the presence of urea, the stabilities of E2(13,22) and E2(20,22) were identical suggesting that the potential i to i' + 2 interchain Glu-Glu repulsion in the E2(20,22) coiled-coil does not occur. In contrast, the mutant E2(15,20) is substantially less stable than E2(13,22) or E2(15,20) by 0.9 kcal/mol due to the presence of two i to i' + 5 interchain Glu-Glu repulsions, which destabilize the coiled-coil by 0.45 kcal/mol each. At pH 3 the coiled-coils were found to increase in stability as the number of Glu substitutions were increased. This, combined with reversed-phase HPLC results at pH 7 and pH 2, supports the conclusion that the protonated Glu side chains present at low pH are significantly more hydrophobic than Gln side chains which are in turn more hydrophobic than the ionized Glu side chains present at neutral pH. The protonated Glu residues increase the hydrophobicity of the coiled-coil interface leading to higher coiled-coil stability. The guanidine hydrochloride results at pH 7 show similar stabilities between the native and mutant coiled-coils indicating that guanidine hydrochloride masks electrostatic repulsions due to its ionic nature and that Glu and Gln in the e and g positions of the heptad repeat have very similar effects on coiled-coil stability in the presence of GdnHCl.  相似文献   

15.
H Kikuchi  Y Goto  K Hamaguchi 《Biochemistry》1986,25(8):2009-2013
The constant (CL) fragment of the immunoglobulin light chain contains only one intrachain disulfide bond buried in the interior of the molecule. The kinetics of reduction with dithiothreitol of the disulfide bond were studied at various concentrations of guanidine hydrochloride at pH 8.0 and 25 degrees C. It was found that the disulfide bond is reduced even in the absence of guanidine hydrochloride. The results of the reduction kinetics were compared with those of the unfolding and refolding kinetics of the CL fragment previously reported [Goto, Y., & Hamaguchi, K. (1982) J. Mol. Biol. 156, 891-910]. It was shown that the reduction of the disulfide bond proceeds through a species with a conformation very similar to that of the fully unfolded one and that the CL fragment undergoes global unfolding transition even in water.  相似文献   

16.
Studies were carried out of viscous-elastic properties of monomolecular layers of human immunoglobulin IgG formed at the interface water solutions of NaCl--air within 20 degrees C to 50 degrees C at NaCl concentration in sublayer from 0.3 to 1.0 M. It has been shown that at concentrations from 0.3 to 0.5 M of NaCl IgG macromolecules keep the tertiary structure, and in the region 35 +/- 5 degrees C a conformation transition is observed. The recorded conformation transition is reversible and of entropy nature. At NaCl concentration 0.75 in the sublayer and temperature close to 35 degrees C IgG macromolecules undergo irreversible structural changes due to the destruction of hydrogen and disulfide bonds in IgG molecules. Macromolecules dissociate to fragments with molecular mass 49,000 +/- 2000. At NaCl concentration 1.0 M and temperatures 30-50 degrees C IgG macromolecules of the monolayer are in a dissociated state. Changes in entropy, enthalpy and heat capacity as well as areas occupied by the macromolecules at dense packing, molecular mass and efficient electric dipole moment of the monolayer are calculated.  相似文献   

17.
The structural and functional aspects of ervatamin B were studied in solution. Ervatamin B belongs to the alpha + beta class of proteins. The intrinsic fluorescence emission maximum of the enzyme was at 350 nm under neutral conditions, and at 355 nm under denaturing conditions. Between pH 1.0- 2.5 the enzyme exists in a partially unfolded state with minimum or no tertiary structure, and no proteolytic activity. At still lower pH, the enzyme regains substantial secondary structure, which is predominantly a beta-sheet conformation and shows a strong binding to 8-anilino-1- napthalene-sulfonic acid (ANS). In the presence of salt, the enzyme attains a similar state directly from the native state. Under neutral conditions, the enzyme was stable in urea, while the guanidine hydrochloride (GuHCl) induced equilibrium unfolding was cooperative. The GuHCl induced unfolding transition curves at pH 3.0 and 4.0 were non-coincidental, indicating the presence of intermediates in the unfolding pathway. This was substantiated by strong ANS binding that was observed at low concentrations of GuHCl at both pH 3.0 and 4.0. The urea induced transition curves at pH 3.0 were, however, coincidental, but non-cooperative. This indicates that the different structural units of the enzyme unfold in steps through intermediates. This observation is further supported by two emission maxima in ANS binding assay during urea denaturation. Hence, denaturant induced equilibrium unfolding pathway of ervatamin B, which differs from the acid induced unfolding pathway, is not a simple two-state transition but involves intermediates which probably accumulate at different stages of protein folding and hence adds a new dimension to the unfolding pathway of plant proteases of the papain superfamily.  相似文献   

18.
The unfolding thermodynamics of the circular enterocin protein AS-48, produced by Enterococcus faecalis, has been studied. The native structure of the 70-amino-acid-long protein turned out to be extremely stable against heat and denaturant-induced unfolding. At pH 2.5 and low ionic strength, it denatures at 102 degrees C, while at 25 degrees C, the structure only unfolds in 6.3 M guanidinium hydrochloride (GuHCl) and does not unfold even in 8 M urea. A comparison of its thermal unfolding in water and in the presence of urea shows a good correspondence between the two deltaGw(298) values, which are about 30 kJ mol(-1) at pH 2.5 and low ionic strength. The stability of the structure is highly dependent upon ionic strength and so GuHCl acts both as a denaturant and a stabilising agent. This seems to be why the deltaGw(298) value calculated from the unfolding data in GuHCl is twice as high as in the absence of this salt. At least part of the high stability of native AS-48 can almost certainly be put down to its circular organization since other structural features are quite normal for a protein of this size.  相似文献   

19.
The formation of amyloid fibrils is an intractable problem in which normally soluble protein polymerizes and forms insoluble ordered aggregates. Such aggregates can range from being a nuisance in vitro to being toxic in vivo. The latter is true for lysozyme, which has been shown to form toxic deposits in humans. In the present study, the effects of partial denaturation of hen egg-white lysozyme via incubation in a concentrated solution of the denaturant guanidine hydrochloride are investigated. Results show that when lysozyme is incubated under moderate guanidine hydrochloride concentrations (i.e., 2-5 M), where lysozyme is partially unfolded, fibrils form rapidly. Thioflavin T, Congo red, X-ray diffraction, transmission electron microscopy, atomic force microscopy, and circular dichroism spectroscopy are all used to verify the production of fibrils under these conditions. Incubation at very low or very high guanidine hydrochloride concentrations fails to produce fibrils. At very low denaturant concentrations, the structure of lysozyme is fully native and very stable. On the other hand, at very high denaturant concentrations, guanidine hydrochloride is capable of dissolving and dis-aggregating fibrils that are formed. Raising the temperature and/or concentration of lysozyme accelerates fibril formation by further adding to the concentration of partially unfolded species. The addition of preformed fibrils also accelerates fibril formation but only under partially unfolding conditions. The results presented here provide further evidence that partial unfolding is a prerequisite to fibril formation. Partial denaturation can accelerate fibril formation in much the same way that mutations have been shown to accelerate fibril formation.  相似文献   

20.
The effects of urea and several methylamine solutes on the catalytic stability and aggregation properties of rabbit muscle phosphofructokinase were assessed at physiologically realistic concentrations of the solutes under several pH and temperature regimes. The loss of catalytic activity observed under conditions of pH-induced cold lability was significantly reduced in the presence of trimethylamine-N-oxide, N-trimethylglycine and N-methylglycine (order of decreasing effectiveness). The concentration-dependent methylamine stabilization of the enzyme, seen with as little as 50 mM trimethylamine-N-oxide, was accompanied by increased aggregation of the enzyme to molecular weights greater than the tetramer (polytetramer) as solute concentration was raised to 400 mM. At pH 6.5-6.7 and 25 degrees C, concentrations of urea greater than 25 mM promoted a time-dependent inactivation of the enzyme which was enhanced at lower temperatures. The urea sensitivity of the enzyme exhibited with 0.8 M urea for 1 h at pH 8.0 did not result in measurable inactivation. The fluorescence emission wavelength maximum of the enzyme was shifted to longer wavelengths and the fluorescence intensity was increased as pH was lowered to 7.0, suggesting the occurrence of a protein conformation change as specific amino acid residues of the tetramer became protonated. Measurements of enzyme light scattering indicated that perturbation by urea was correlated with tetramer dissociation, which was irreversible by dialysis at 25 degrees C. The urea and methylamine influences on phosphofructokinase activity and structure were not counteracting. The synergistic interactions among pH, temperature, and solutes observed with phosphofructokinase are compared to effects on other associating-dissociating protein systems in order to evaluate possible mechanisms of action of these low molecular weight solutes.  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