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.
The CO2 hydration activities of cloned human carbonic anhydrase II (carbonate hydro-lyase, EC 4.2.1.1) and variants with Lys, Glu, Gln or Ala replacing His at sequence position 64 have been measured in a variety of different buffers in the pH range 6-9. The variants with Lys-64, Gln-64 and Ala-64 showed non-Michaelis-Menten behavior under some conditions, apparent substrate inhibition being prominent near pH 9. However, asymptotic Michaelis-Menten parameters could be estimated for the limit of low substrate concentrations. All variants show distinct buffer specificities, and imidazole derivatives, Ches and phosphate buffers yield higher kcat values that Bicine, Taps and Mops buffers under otherwise similar conditions. These results are interpreted in terms of different pathways for a rate-limiting proton transfer. In unmodified enzyme, the very high catalytic activity depends on His-64 functioning as an efficient proton transfer group, but this pathway is not available in the variants with Gln-64 and Ala-64. Imidazoles, Ches and phosphate are thought to participate in a metal center-to-buffer proton transfer pathway, whereas Bicine, Taps, Mops and Mes appear to lack this capacity, so that the rate-limiting proton transfer occurs in a metal center-to-bulk water pathway for these variants. The Lys-64 and Glu-64 variants give significantly higher kcat values in Taps, Mops and Mes buffers than the Ala-64 and Gln-64 variants. The pH dependencies of these kcat values are compatible with the hypothesis that Lys-64 and Glu-64 can function as proton transfer groups. Thus, at pH near 9, Lys-64 appears to be only 5-times less efficient than His-64, while Glu-64 is inefficient. At pH 6, Lys-64 is an inefficient proton transfer group, but Glu-64 is only 2-3-times less efficient than His-64. The data indicate that Lys-64 and Glu-64 have pKa values near 8 and below 6, respectively.  相似文献   

2.
Good's zwitterionic buffers are widely used in biological and biochemical research in which hydrogen peroxide is a solution component. This study was undertaken to determine whether Good's buffers exhibit reactivity toward H(2)O(2). It is found that H(2)O(2) oxidizes both morpholine ring-containing buffers (e.g., Mops, Mes) and piperazine ring-containing zwitterionic buffers (e.g., Pipes, Hepes, and Epps) to produce their corresponding N-oxide forms. The percentage of oxidized buffer increases as the concentration of H(2)O(2) increases. However, the rate of oxidation is relatively slow. For example, no oxidized Mops was detected 2h after adding 0.1M H(2)O(2) to 0.1M Mops (pH 7.0), and only 5.7% was oxidized after 24h exposure to H(2)O(2). Thus, although all of these buffers can be oxidized by H(2)O(2), their slow reaction does not significantly perturb levels of H(2)O(2) in the time frame and at the concentrations of most biochemical studies. Therefore, the previously reported rapid loss of H(2)O(2) produced from the ferroxidase reaction of ferritin is unlikely due to reaction of H(2)O(2) with buffer, a conclusion supported by the fact that H(2)O(2) is also lost rapidly when the solution pH of the ferroxidase reaction is controlled by a pH stat apparatus in the absence of buffer.  相似文献   

3.
The hydrolysis of acetylamino acids by highly purified hog kidney aminoacylase I (N-acylamino acid amidohydrolase, EC 3.5.1.14) was investigated using flow injection analysis to determine reaction rates. We show that the distinctly bell-shaped pH versus activity profiles observed in previous studies do not reflect protonic equilibria in the enzyme, but were created by buffer effects. At low pH, anions such as phosphate, nitrate or chloride markedly increase Km. These effects are reversed at higher pH. In zwitterionic 'Good' buffers (Mes, Mops, and Bicine), maximal velocities are almost independent of pH between 6.5 and 9 for all substrates studied (Ac-LAla, Ac-LGlu, Ac-LMet, Ac-LPhe). Below pH 6.5, the catalytic constants decrease with pH, apparently due to the protonation of a carboxylate with a pKa of 5.5-6. The pH dependence of Km markedly varies among different substates. We conclude that the observed profiles all result from the dissociation of an active-site residue with a pKa of 8-8.5, which we tentatively identify as an active-site cysteine residue. A working model of aminoacylase catalysis is presented that accounts for most of the known facts.  相似文献   

4.
The steady-state kinetics of CO2 hydration catalyzed by human carbonic anhydrase I (carbonate hydro-lyase, EC 4.2.1.1) has been investigated at three pH values corresponding to different parts of the pH-rate profile. Two buffer systems with similar pKa values were used at each pH. The results show that the catalyzed rates depend on the buffer concentration but also on the chemical nature of the buffer. For example, at pH 8.8 the buffer 1,2-dimethylimidazole behaves formally as a second substrate in a 'ping-pong' mechanism yielding a maximal kcat value of 2.2 x 10(5) s-1, whereas much lower rates were obtained with Taps buffers. Similarly, at pH 7.3 1-methylimidazole yields higher rates than Mops and at pH 6.3 3,5-lutidine is more efficient than Mes. Non-Michaelis-Menten kinetics were observed with all buffers except 1,2-dimethylimidazole. In addition, while the apparent buffer activation by 1,2-dimethylimidazole can be described by a single Km value of 26 mM, the Mes concentration dependence is consistent with the presence of two components of similar magnitudes with Km values of 45 mM and 0.15 mM. These results are interpreted within the framework of the 'zinc-hydroxide' mechanism in terms of multiple pathways for the rate-contributing transfer of a proton from the zinc-bound water molecule, formed during CO2/HCO3- interconversion, to the reaction medium, thus, regenerating zinc-bound OH-.  相似文献   

5.
The ionization of 4-nitroimidazole to 4-nitroimidazolate was investigated as a function of ionic strength. The apparent pKa varies from 8.99 to 9.50 between 0.001 and 1.0 M ionic strength, respectively, at 25 degrees C. The ionic strength dependence of this ionization is anomalous. The binding of 4-nitroimidazole by horse metmyoglobin was studied between pH 5.0 and 11.5 and as a function of ionic strength between 0.01 and 1.0 M. The association rate constant is pH-dependent, varying from 24 M(-1)s(-1) at pH 5 to a maximum value of 280 M(-1)s(-1) at pH 9.5 and then decreasing to 10 M(-1)s(-1) at pH 11.5 in 0.1 M ionic strength buffers. The dissociation rate constant has a much smaller pH dependence, varying from 0.082 s(-1) at low pH to 0.035 s(-1) at high pH, with an apparent pKa of 6.5. The binding affinity of 4-nitroimidazole to horse metmyoglobin is about 2.5 orders of magnitude stronger than that for imidazole and this increased affinity is attributed to the much slower dissociation rate for 4-nitroimidazole compared to that of imidazole. Although the ionic strength dependence of the binding rate is small and secondary kinetic salt effects can account for the ionic strength dependence of the association rate constant, the pH dependence of the rate constants and microscopic reversibility arguments indicate that the anionic form of the ligand binds more rapidly to all forms of metmyoglobin than does the neutral form of the ligand. However, the spectrum of the complex is similar to model complexes involving neutral imidazole and not imidazolate. The latter observation suggests that the initial metmyoglobin/4-nitroimidazolate complex rapidly binds a proton and the neutral form of the bound ligand is stabilized, probably through hydrogen binding with the distal histidine.  相似文献   

6.
Human lactotransferrin binds 2 Fe3+ tightly at two specific sites. In order to demonstrate differences between the stability of the two iron-binding sites, the removal of iron was studied in buffers in the pH range 8-3 varying the ionic strength and with or without metal chelators such as phosphate ions and EDTA.The results show that in the presence of formate and acetate buffers of ionic strength 0.1–0.4 and in a pH range of 5–3, the two Fe3+ from human lactotransferrin are removed stimultaneously.Addition of 4 mM EDTA to buffers of ionic strength 0.1 and in the pH range 8–3 shows that between pH 5–4.3 the iron from only one of the binding sites, called the ‘acid labile’ site, is removed.Addition of 0.2 M phosphate ions to buffers of ionic strength 0.2 and in pH range 8–3 containing 4 mM EDTA shows that Fe3+ from the ‘acid labile’ site may be completely removed at pH 6. Removal of Fe3+ from the ‘acid stable’ site is obtained at pH 4.The differential behavior of the two iron binding sites was also shown by saturation experiments in the presence of citrate/bicarbonate buffers at different pH values. In a pH range 6.2–4.8, 50% saturation was obtained, but at pH 6.35 complete saturation was achieved. When saturation of partially saturated samples of human lactotransferrin was performed with 59Fe it was demonstrated that in the pH range 6.2–4.8 iron is bound only to the ‘acid labile’ site.  相似文献   

7.
《Free radical research》2013,47(3):149-160
Iron autoxidation in Mops and Hepes buffers is characterized by a lag phase that becomes shorter with increasing FeCl2 concentration and pH. During iron oxidation in these buffers a yellow colour develops in the solution. When the reaction is conducted in the presence of nitro blue tetrazolium (NBT), blue formazan is formed. Of the many OH' scavengers tested, mannitol and sorbitol are most effective in inhibiting Fe2+ oxidation, yellow colour development and NBT reduction. Some inhibition was also noted with catalase. The iron product of the oxidative reaction differs from Fe3+ in its absorption spectrum and its low reactivity with thiocyanate. Similar results are obtained when iron autoxidation is studied in unbuffered solutions brought to alkaline pH with NaOH. In phosphate buffer, no lag phase is evident and the absorption spectrum of the final solution is identical to that of Fe3+ in this buffer. The iron product reacts immediately with thiocyanate. When iron oxidation is conducted in the presence of NBT the formation of formazan is almost undetectable. Of the many compounds tested only catalase inhibits iron autoxidation in this buffer. The sequence of reactions leading to iron autoxidation in Good-type buffers1 thus resembles that occurring in unbuffered solutions brought to alkaline pH with NaOH and greatly differs from that occurring in phosphate buffer. These results are in agreement with the observation that these buffers have very low affinity for iron.1 The data presented define experimental conditions where Fe2+ is substantially stable for a considerable length of time in Mops buffer.  相似文献   

8.
The reconstitution of Brome Mosaic Virus (BMV) has been studied using neutron scattering. Experiments were performed on disassembled virus without subsequent separation of components. Phase diagrams of the disassembly and subsequent reassembly of BMV were established as a function of pH and LiCl molarity by analytical centrifugation and quasi-elastic light scattering. Disassembly occurs at a pH above 6.5 and above 0.8 M LiCl. On reassembly, if the pH is lowered first, capsids are formed without subsequent incorporation of RNA. Neutron scattering was used to investigate the formation of virus particles, when the ionic strength was lowered from 1.4 to 0.1 M LiCl at pH 7.8. The reconstitution was followed continuously. As it was driven by a lowering of the ionic strength the kinetics of the process cannot be studied for short times. However the fact that at any given ionic strength no evolution of the scattering was observed with time implies that the reconstitution is complete within a few minutes. The observations in buffers with various amounts of D2O lead to the conclusion that the reassembly is achieved by co-condensation of the RNA and of the capsid proteins.  相似文献   

9.
Isolated human erythrocyte membranes crenate when suspended in isotonic medium, but can use MgATP to reduce their net positive curvature, yielding smooth discs and cup forms that eventually undergo endocytosis. An earlier report from this laboratory (Patel, V.P. and Fairbanks, G. (1981) J. Cell Biol. 88, 430-440), has described a phenomenon of ATP-independent shape change in which ghosts prepared by hemolysis and washing in synthetic zwitterionic buffers crenated at 0 degree C, but underwent conversion to smooth discs and cups when warmed in the absence of MgATP. We have further explored the effect of the hemolysis condition on the requirement for ATP in ghost shape change. 25 hemolysis buffers were applied at 10 mM (pH 7.4, 0 degree C). Eight anionic buffers with relatively high ionic strength (e.g., phosphate and diethylmalonic acid (DMA] yielded ghosts requiring ATP for shape change, while two cationic buffers (Bistris and imidazole) and ten synthetic zwitterionic buffers (e.g., Tricine and Hepes) with lower ionic strength produced ghosts that smoothed spontaneously at 30 degrees C. Hemolysis at intermediate ionic strength yielded mixed populations in which spontaneous smoothing was expressed in all-or-none fashion. Maximal ATP-independent shape change was induced by hemolysis at pH 7.3-7.7, while ATP was required after hemolysis at pH less than or equal to 7.1 even when the ionic strength at hemolysis was low. Ghosts requiring ATP could be converted to ATP independence by washing at low ionic strength, but ATP independence could not be reversed readily by washing at high ionic strength. Exposure to low ionic strength at pH greater than 7.1 presumably changes membrane organization in a way that alters the temperature dependence of tensions within the bilayer or skeleton of the composite membrane.  相似文献   

10.
The betabellin structure is a de novo designed beta-sandwich protein consisting of two 32-residue beta sheets packed against one another by hydrophobic interactions. Betabellin 16S (B16S), a 32-residue peptide chain (HSLTAKIakLTFSIAahTYTCAVakYTAKVSH, where a is DAla, h is DHis, and k is DLys), did not have beta structure in water at pH 6.5. Air oxidation of B16S furnished betabellin 16D (B16D), a 64-residue disulfide-bridged two-chain protein, which also did not fold in water at pH 6.5. However, the extent of beta structure observed for B16D increased with pH and ionic strength of the solution and the B16D concentration as observed by circular dichroism spectropolarimetry. Transmission electron microscopy showed that B16D formed narrow fibrils that associated into broad ribbons in 5.0 mM Mops and 0.25 M NaCl at pH 6.9.  相似文献   

11.
Oxidative deposition of iron in ferritin or the autoxidation of iron in the absence of protein produces radicals from Good's buffers. Radical species are formed from the piperazine ring-based buffers Hepes (4-(2-hydroxyethyl)-1-piperazineethanesulfonic acid), Epps 4-(2-hydroxyethyl)-1-piperazinepropanesulfonic acid, and Pipes 1,4-piperazinediethanesulfonic acid, but not from Mes (4-morpholineethanesulfonic acid) which contains a morpholine ring. The radicals all have half-lives around 10 min and display very similar electron paramagnetic resonance spectra consisting of at least 30 lines. The Hepes radical can be formed by the addition of potassium superoxide directly to the buffer and its production during iron(II) autoxidation is inhibited by superoxide dismutase (EC 1.15.1.1). Catalase (EC 1.11.1.6) accelerates the decay of the EPR spectrum. Thus the buffer radicals are secondary radical species produced from oxygen radicals formed during the iron catalyzed Haber-Weiss process. The deoxyribose/thiobarbituric acid assay for hydroxyl radical production shows that Hepes is an effective hydroxyl radical scavenging agent. The Hepes radical can also be formed electrolytically at a potential of +0.8 V (vs standard hydrogen electrode). Oxidation of Hepes at pH 10 during the autoxidation of iron(II) or by the addition of hydrogen peroxide produces a nitroxide radical. These results indicate that piperazine ring Good buffers should be avoided in studies of redox processes in biochemistry.  相似文献   

12.
The reaction of horse spleen ferritin (HoSF) with Fe2+ at pH 6.5 and 7.5 using O2, H2O2 and 1:1 a mixture of both showed that the iron deposition reaction using H2O2 is approximately 20- to 50-fold faster than the reaction with O2 alone. When H2O2 was added during the iron deposition reaction initiated with O2 as oxidant, Fe2+ was preferentially oxidized by H2O2, consistent with the above kinetic measurements. Both the O2 and H2O2 reactions were well defined from 15 to 40 degrees C from which activation parameters were determined. The iron deposition reaction was also studied using O2 as oxidant in the presence and absence of catalase using both stopped-flow and pumped-flow measurements. The presence of catalase decreased the rate of iron deposition by approximately 1.5-fold, and gave slightly smaller absorbance changes than in its absence. From the rate constants for the O2 (0.044 s(-1)) and H2O2 (0.67 s(-1)) iron-deposition reactions at pH 7.5, simulations of steady-state H2O2 concentrations were computed to be 0.45 microM. This low value and reported Fe2+/O2 values of 2.0-2.5 are consistent with H2O2 rapidly reacting by an alternate but unidentified pathway involving a system component such as the protein shell or the mineral core as previously postulated [Biochemistry 22 (1983) 876; Biochemistry 40 (2001) 10832].  相似文献   

13.
The pH dependence of papain catalysis was analyzed by a scheme which evaluates the kinetic contribution of both protonated and unprotonated species of functional groups involved in catalysis. Kinetic measurements were made at constant pH, without buffers, by automatic titration. The rate-determining step for papain-catalyzed hydrolysis of alpha-N-benzoyl-L-arginine ethyl ester, determined by nucleophile competition, changed from acylation below pH 6.5 to mixed acylation-deacylation above pH 6.5. Kinetic analysis indicated that three prototropic groups governed the pH-specificity of alpha-N-benzoyl-L-arginine ethyl ester hydrolysis. These prototropic groups had pKa values of 4.8, 6.5 to 6.7, and 8.7. Theoretical treatment of the kinetics provided an excellent fit with the experimentally found profile when the contribution of all three prototropic groups was considered. Analysis showed that, in acid, the pathways of papain catalysis were functional with either two or three active-site protons. In base, a single functional ionic pathway is associated with an active site with only one proton. Pathways involving an unprotonated active site are catalytically inoperative in both acid and base. These results indicate that papain exhibits several catalytically functional ionic pathways. The results are discussed in terms of pKa assignments, and the mechanism of papain catalysis.  相似文献   

14.
Osmotic pressure studies were carried on tobacco mosaic virus (TMV) and its components, protein and RNA, as well as on bis(3,3′-aminopropyl)amine, reported to be present in TMV preparations. Solvents were phosphate and barbital buffers at different values of pH and ionic strength. Measurements were made at room temperature. The Donnan effect was exhibited by TMV protein in phosphate buffer of 0.01 ionic strength at pH values ranging between 5.8 and 7.5. The observed values of the Donnan effect at pH 5.8 and 5.97 were in reasonable agreement with theoretical values calculated from the charge obtained by hydrogen ion titration. TMV-RNA in phosphate buffer at pH 7.5 and ionic strength 0.01 did not exhibit more than 1% of the expected Donnan effect. This is explained tentatively as the result of firm binding of metal ions. Negative values of osmotic pressure were observed with bis(3,3′-aminopropyl)amine. Similar anomalous osmosis was sometimes observed with TMV protein and with TMV. In agreement with earlier observations, TMV did not exhibit the Donnan effect in phosphate buffer of 0.01 ionic strength at pH values ranging from 5.5 to 8.0. However, TMV dialysed extensively in the presence of EDTA at pH 8.5 and TMV produced by reconstitution of purified protein and RNA did exhibit the Donnan effect in both phosphate and barbital buffers. The magnitude was of the same order as that calculated from the net charge determined by hydrogen ion titration. When reconstituted TMV, which did exhibit Donnan effect, was treated with calcium ions, the effect was abolished.  相似文献   

15.
S A Kuby  R N Roy 《Biochemistry》1976,15(9):1975-1987
A systematic study has been made of the pH- and temperature-dependency of the steady-state kinetic parameters of the stabilized two-subunit enzyme species of glucose-6-phosphate dehydrogenase, in the absence of superimposed association-dissociation reactions. The Vmax(app) data obtained in several buffers between pH 5 and 10 and at 18-32 degrees C lead to the postulate that at least two sets of protonic equilibria may govern the catalysis (one near pH 5.7 AT 25 DEGREES C and another near pH 9.2); furthermore, two pathways for product formation (i.e., two Vmax's) appear to be required to explain the biphasic nature of the log Vmax(app) vs. pH curves, with Vmax(basic) greater than Vmax(acidic + neutral). Of the several buffers explored, either a uniform degree of interaction or a minimal degree of buffer species interaction could be assessed from the enthalpy changes associated with the derived values for ionization constants attributed to the protonic equilibria in the enzyme-substrates ternary complexes for the case of Tris-acetate-EDTA buffers, at constant ionic strength. With the selection of this buffer at 0.1 (T/2) and at 25 and 32 degrees C, a self-consistent kinetic mechanism has emerged which allows for the random binding of the two fully ionized substrates to the enzyme via two major pathways, and product formation by both E-A--B- and HE-A--B-. As before (Kuby et al. Arch. Biochem, Biophys. 165, 153-178, 1974), a quasi-equilibrium is presumed, with rate-limiting steps (k + 5 and k + 5') at the interconversion of the ternary complexes. Values for the two sets of protonic equilibria defined by this mechanism (viz., pKk, pKH2 for the first ionizations, and pKk', pKH' for the second) could then be estimated. From their numerical values (e.g., at 25 degrees C: pKK = 5.7 PKH2 = 5.2; and pKK' = 9.1, PKH' = 8.2) and from the values for delta H degrees ioniz (e.g., delta H degrees pKK APPROXIMATELY 5.1 KCAL/MOL; DELTA H degrees pKK' APPROXIMATELY 11 KCAL/MOL), A POSTULATE IS PRESENTED WHICH ATTRIBUTES THESE Acid dissociation constants to an imidazole and epsilon-amino group, respectively.  相似文献   

16.
The binding of horse heart cytochrome c to yeast cytochrome c peroxidase in which the heme group was replaced by protoporphyrin IX was determined by a fluorescence quenching technique. The association between ferricytochrome c and cytochrome c peroxidase was investigated at pH 6.0 in cacodylate/KNO3 buffers. Ionic strength was varied between 3.5 mM and 1.0 M. No binding occurs at 1.0 M ionic strength although there was a substantial decrease in fluorescence intensity due to the inner filter effect. After correcting for the inner filter effect, significant quenching of porphyrin cytochrome c peroxidase fluorescence by ferricytochrome c was observed at 0.1 M ionic strength and below. The quenching could be described by 1:1 complex formation between the two proteins. Values of the equilibrium dissociation constant determined from the fluorescence quenching data are in excellent agreement with those determined previously for the native enzyme-ferricytochrome c complex at pH 6.0 by difference spectrophotometry (J. E. Erman and L. B. Vitello (1980) J. Biol. Chem. 225, 6224-6227). The binding of both ferri- and ferrocytochrome c to cytochrome c peroxidase was investigated at pH 7.5 as functions of ionic strength in phosphate/KNO3 buffers using the fluorescence quenching technique. The binding in independent of the redox state of cytochrome c between 10 and 20 mM ionic strength, but ferricytochrome c binds with greater affinity at 30 mM ionic strength and above.  相似文献   

17.
The buffer used during horse spleen ferritin iron loading significantly influences the mineralization process and the quantity of iron deposited in ferritin. Ferritin iron loading in imidazole shows a rapid hyperbolic curve in contrast to iron loading in 3-(N-morpholino)propanesulfonic acid (MOPS), which displays a slower sigmoidal curve. Ferritin iron loading in an equimolar mixture of imidazole and MOPS produces an iron-loading curve that is intermediate between the imidazole and MOPS curves indicating that one buffer does not dominate the reaction mechanism. The UV-visible spectrum of the ferritin mineral has a higher absorbance from 250 to 450 nm when prepared in imidazole buffer than in MOPS buffer. These results suggest that different mineral phases form in ferritin by different loading mechanisms in imidazole and MOPS buffered reactions. Samples of 1500 Fe/ferritin were prepared in MOPS or imidazole buffer and were analyzed for crystallinity and using the electron diffraction capabilities of the electron microscope. The sample prepared in imidazole was significantly more crystalline than the sample prepared in MOPS. X-ray powder diffraction studies showed that small cores (~ 500 Fe/ferritin) prepared in MOPS or imidazole possess a 2-line ferrihydrite spectrum. As the core size increases the mineral phase begins to change from 2-line to 6-line ferrihydrite with the imidazole sample favoring the 6-line ferrihydrite phase. Taken together, these results suggest that the iron deposition mechanism in ferritin can be controlled by properties of the buffer with samples prepared in imidazole forming a larger, more ordered crystalline mineral than samples prepared in MOPS.  相似文献   

18.
l-Histidine as pseudobiospecific ligand was immobilized onto poly(ethylene vinyl alcohol) hollow-fiber membranes to obtain an affinity support for immunoglobulin G (IgG) purification. The interaction of human IgG with the affinity membranes was studied by chromatography and equilibrium binding analysis. Adsorption was possible over a broad pH range and was found to depend strongly on the nature of the buffer ions rather than on ionic strength. With zwitterionic buffers like morpholinopropanesulfonic acid (Mops) and hydroxyethylpiperazineethanesulfonic acid (Hepes), much higher adsorption capacities were obtained than with other buffers like Tris-HCl and phosphate buffers. An inhibition analysis revealed that non-zwitterionic buffers competitively inhibit IgG binding, whereas Mops and Hepes in their zwitterionic form do not. By choosing the appropriate buffer system, it was possible to adsorb specifically different IgG subsets. The IgG molecules were found to adsorb on membrane immobilized histidine via their Fab part. Determination of dissociation constants at different temperatures allowed calculation of thermodynamic adsorption parameters. Decrease in KD with increasing temperature and a positive entropy value between 20 and 35°C (in Mops buffer) indicated that adsorption is partially governed by hydrophobic forces in that temperature range, whereas at lower temperatures, electrostatic forces are more important for adsorption.  相似文献   

19.
Interactions of both purified tubulin and microtubule protein (tubulin plus associated proteins) with two commonly used sulfonate buffers were examined. 1,4-Piperazineethanesulfonate (Pipes) and 4-morpholineethanesulfonate (Mes) at high concentrations induce the polymerization of purified tubulin in reactions requiring only buffer, tubulin and GTP. While both reactions were temperature-dependent, cold-reversible and inhibited by GDP, colchicine or Ca2+, there were significant differences between them. Substantially lower tubulin and buffer concentrations were required for Pipes-induced polymerization; and turbidity was much more intense in the Pipes-induced than in the Mes-induced reaction at the same protein concentration. Electron microscopy demonstrated that for the most part typical smooth-walled microtubules were formed in Mes, while aberrant forms were the predominant structures formed in Pipes. When the polymerization of microtubule protein was examined as a function of buffer concentration, biphasic patterns were observed with both Pipes and Mes: polymerization occurred at both low and high, but not intermediate, buffer concentrations. The turbidity observed at high concentrations of Pipes greatly exceeded that at low concentrations. With Mes, equivalent turbidity developed at both high and low buffer concentrations. Although associated proteins copolymerized with tubulin at low buffer concentrations, they were excluded from the polymerized material at high buffer concentrations. Pipes and Mes were compared to sodium phosphate, Tris/HCl and imidazole/HCl buffers at 0.1 M in several polymerization systems using both purified tubulin and microtubule protein. The sulfonate buffers were invariably associated with more vigorous reactions than the other buffers.  相似文献   

20.
The CuZn superoxide dismutases (SODs) from ox, sheep, pig and yeast were investigated by pulse radiolysis in order to evaluate the role of electrostatic interactions between O2.- and SOD proteins in the mechanism of action of the SOD enzymes. The protein net charge in this series varies, as evaluated by the protein pI values spanning over a large range of pH: 8.0 (sheep), 6.5 (pig), 5.2 (ox) and 4.6 (yeast). The amino acid sequences are largely conserved, with the three mammalian proteins being highly homologous and the yeast protein having some distinct variations in the region surrounding the active site. At pH 8.0 the activities of the SODs from various sources are similar, though the minor differences observed suggest that in the highly homologous mammalian series the most acidic protein is the most enzymically efficient one. The pH-dependences of the various activities in the pH range 7-12 are similar, and the related curves are best fitted by two pK values, which are approx. 9.2 and 11.0 for the mammalian enzymes and 9.1 and 11.4 for the yeast enzyme. The activities of the proteins at I 0.1 are decreased by approx. 20% when compared with the activity at I 0.02 at pH 8.5, whereas at pH above 10 the pH-dependence of the activity approaches that determined at I 0.02 and at pH 11.9 the activity is essentially independent of ionic strength. The dependence upon ionic strength also depends on the salt used, with perchlorate being more effective than phosphate or borate or Mops and still effective at pH above 10.5, where the effect of other salts becomes negligible. The dual and concerted dependence of the activities of different SODs on pH and salt concentration is explained with the encounter of O2.- with the active-site copper being governed by the protonation of two positively charged groups in the vicinity of the active site. The gradient between these localized charges and the rest of the protein may explain the different activities of the mammalian proteins at lower pH. On the basis of the sequence variation of the SODs examined it is not possible to definitely identify these groups. Likely candidates are conserved basic amino acid side chains in the vicinity (less than or equal to 1.2 nm) of the active site, i.e. Lys-134 and Arg-141, but co-ordination of OH- in the first copper co-ordination sphere may be an additional factor accounting for the higher pK.(ABSTRACT TRUNCATED AT 400 WORDS)  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