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 46 毫秒
1.
To understand the effect of visible light on the stability of photoactive yellow protein (PYP), urea denaturation experiments were performed with PYP in the dark and with PYP(M) under continuous illumination. The urea concentrations at the midpoint of denaturation were 5.26 +/- 0.29 and 3.77 +/- 0.19 M for PYP and PYP(M), respectively, in 100 mM acetate buffer, and 5.26 +/- 0.24 and 4.11 +/- 0.12 M for PYP and PYP(M), respectively, in 100 mM citrate buffer. The free energy change upon denaturation (DeltaG(D)(H2O)), obtained from the denaturation curve, was 11.0 +/- 0.4 and 7.6 +/- 0.2 kcal/mol for PYP and PYP(M), respectively, in acetate buffer, and 11.5 +/- 0.3 and 7.8 +/- 0.1 kcal/mol for PYP and PYP(M), respectively, in citrate buffer. Even though the DeltaG(D)(H2O) value for PYP(M) is almost identical in the two buffer systems, the urea concentration at the midpoint of denaturation is lower in acetate buffer than in citrate buffer. Although their CD spectra indicate that the protein conformations of the denatured states of PYP and PYP(M) are indistinguishable, the configurations of the chromophores in their denatured structures are not necessarily identical. Both denatured states are interconvertible through PYP and PYP(M). Therefore, the free energy difference between PYP and PYP(M) is 3.4-3.7 kcal/mol for the protein moiety, plus the additional contribution from the difference in configuration of the chromophore.  相似文献   

2.
The long lived intermediate (signaling state) of photoactive yellow protein (PYP(M)), which is formed in the photocycle, was characterized at various pHs. PYP(M) at neutral pH was in equilibrium between two spectroscopically distinct states. Absorption maxima of the acidic form (PYP(M)(acid)) and alkaline form (PYP(M)(alkali)) were located at 367 and 356 nm, respectively. Equilibrium was represented by the Henderson-Hasselbalch equation, in which apparent pK(a) was 6.4. Content of alpha- and/or beta-structure of PYP(M)(acid) was significantly greater than PYP(M)(alkali) as demonstrated by the molar ellipticity at 222 nm. In addition, changes in amide I and II modes of beta-structure in the difference Fourier transform infrared spectra for formation of PYP(M)(acid) was smaller than that of PYP(M)(alkali). The vibrational mode at 1747 cm(-1) of protonated Glu-46 was found as a small band for PYP(M)(acid) but not for PYP(M)(alkali), suggesting that Glu-46 remains partially protonated in PYP(M)(acid), whereas it is fully deprotonated in PYP(M)(alkali). Small angle x-ray scattering measurements demonstrated that the radius of gyration of PYP(M)(acid) was 15.7 Angstroms, whereas for PYP(M)(alkali) it was 16.2 Angstroms. These results indicate that PYP(M)(acid) assumes a more ordered and compact structure than PYP(M)(alkali). Binding of citrate shifts this equilibrium toward PYP(M)(alkali). UV-visible absorption spectra and difference infrared spectra of the long lived intermediate formed from E46Q mutant was consistent with those of PYP(M)(acid), indicating that the mutation shifts this equilibrium toward PYP(M)(acid). Alterations in the nature of PYP(M) by pH, citrate, and mutation of Glu-46 are consistently explained by the shift of the equilibrium between PYP(M)(acid) and PYP(M)(alkali).  相似文献   

3.
It is widely accepted that PYP undergoes global structural changes during the formation of the biologically active intermediate PYP(M). High-angle solution x-ray scattering experiments were performed using PYP variants that lacked the N-terminal 6-, 15-, or 23-amino-acid residues (T6, T15, and T23, respectively) to clarify these structural changes. The scattering profile of the dark state of intact PYP exhibited two broad peaks in the high-angle region (0.3 A(-1) < Q < 0.8 A(-1)). The intensities and positions of the peaks were systematically changed as a result of the N-terminal truncations. These observations and the agreement between the observed scattering profiles and the calculated profiles based on the crystal structure confirm that the high-angle scattering profiles were caused by intramolecular interference and that the structure of the chromophore-binding domain was not affected by the N-terminal truncations. The profiles of the PYP(M) intermediates of the N-terminally truncated PYP variants were significantly different from the profiles of the dark states of these proteins, indicating that substantial conformational rearrangements occur within the chromophore-binding domain during the formation of PYP(M). By use of molecular fluctuation analysis, structural models of the chromophore-binding region of PYP(M) were constructed to reproduce the observed profile of T23. The structure obtained by averaging 51 potential models revealed the displacement of the loop connecting beta4 and beta5, and the deformation of the alpha4 helix. High-angle x-ray scattering with molecular fluctuation simulation allows us to derive the structural properties of the transient state of a protein in solution.  相似文献   

4.
Light-activation of the PAS domain protein photoactive yellow protein (PYP) is believed to trigger a negative phototactic response in the phototropic bacterium Halorhodospira halophila. To investigate transient conformational changes of the PYP photocycle, we utilized the PYP mutant M100L that displays an increased lifetime of the putative signaling-state photointermediate PYP(M) by 3 orders of magnitude, as previously reported for the M100A mutant [Devanathan, S., Genick, U. K., Canestrelli, I. L., Meyer, T. E., Cusanovich, M. A., Getzoff, E. D., and Tollin, G. Biochemistry (1998) 37, 11563-11568]. The FTIR difference spectrum of PYP(M) and the ground state of M100L demonstrated extensive peptide-backbone structural changes as observed in the FTIR difference spectrum of the wild-type protein and PYP(M). The conformational change investigated by CD spectroscopy in the far-UV region showed reduction of the alpha-helical content by approximately 40%, indicating a considerable amount of changes in the secondary structure. The optical activity of the p-coumaric acid chromophore completely vanished upon PYP(M) in contrast to the dark state, indicating deformation of the binding pocket structure in PYP(M). The tertiary structural changes were further monitored by small-angle X-ray scattering measurements, which demonstrated a significant increase of the radius of gyration of the molecule by approximately 5% in PYP(M). These structural changes were reversed concomitantly with the chromophore anionization upon the dark state recovery. The observed changes of the quantities provided a more vivid view of the structural changes of the mutant PYP in going from PYP(M) to PYP(dark), which can be regarded as a process of folding of the secondary and the tertiary structures of the "PAS" domain structure, coupled with the p-coumaric acid chromophore deprotonation and isomerization.  相似文献   

5.
The spectroscopic properties of photoactive yellow protein (PYP) partially digested by chymotrypsin were studied. Chymotrypsin yielded three major products that were yellow but distinguishable by SDS-PAGE. They were readily separated by DEAE-Sepharose column chromatography. Protein sequencing and mass spectrometry demonstrated that chymotrypsin cleaved the N-terminal 6, 15, or 23 amino acids (T6, T15, and T23). The blue-shifts of the absorption maxima and the increases in the apparent pK(a) of the chromophores relative to those of intact PYP were less than 4 nm and 0.2, respectively. The absorption spectra of the near-UV intermediates produced from T6, T15, and T23 were identical to that of intact PYP, but with lifetimes that were 140, 2,300, and 4,500 times longer, respectively. These observations suggest that the recovery of the dark state of PYP from the near-UV intermediate is accelerated by the N-terminal region, and that this region acts as a regulatory factor for the photocycle of PYP.  相似文献   

6.
PYP (photoactive yellow protein) is a photoreceptor protein, which is activated upon photo-isomerization of the p-coumaric acid chromophore and is inactivated as the chromophore is thermally back-isomerized within a second (in PYP(M)-to-PYP(dark) conversion). Here we have examined the mechanism of the rapid thermal isomerization by analyzing mutant PYPs of Met100, which was previously shown to play a major role in facilitating the reaction [Devanathan, S. et al. (1998) Biochemistry 37, 11563-11568]. The mutation to Lys, Leu, Ala, or Glu decelerated the dark state recovery by one to three orders of magnitude. By evaluating temperature-dependence of the kinetics, it was found that the retardation resulted unequivocally from elevations of activation enthalpy (DeltaH( double dagger )) but not the other parameters such as activation entropy or heat capacity changes. Another effect exerted by the mutations was an up-shift of the apparent pK(a) of the chromophore [the pK(a) of a titratable group (X) that controls the pK(a) of the chromophore] in the PYP(M)-decay process. The pK(a) up-shift and the DeltaH( double dagger ) elevation show an approximately linear correlation. We, therefore, postulate that the role of Met100 is to reduce the energy barrier of the PYP(M)-decay process by an indirect interaction through X and that the process is thereby facilitated.  相似文献   

7.
Cyanide binding to prostaglandin H (PGH) synthase results in a spectral shift in the Soret region. This shift was exploited to determine equilibrium and kinetic parameters of the cyanide binding process. At pH 8.0, ionic strength 0.22 M, 4 degrees C, the cyanide dissociation constant, determined from equilibrium experiments, is (65 +/- 10) microM. The binding rate constant is (2.8 +/- 0.2) x 10(3) M-1 s-1, and the dissociation rate constant is zero within experimental error. Through a kinetic study of the binding process as a function of pH, from pH 3.96 to 8.00, it was possible to determine the pKa of a heme-linked acid group on the enzyme of 4.15 +/- 0.10 with citrate buffer. An apparent pKa of 4.75 +/- 0.03 was determined with acetate buffer; this different value is attributed to complexation of the enzyme with one of the components of the acetate buffer.  相似文献   

8.
Purification of cysteine proteinases from adult Schistosoma mansoni   总被引:5,自引:0,他引:5  
Proteolytic activity against hemoglobin and low molecular weight synthetic substrates has been previously found in homogenates and excretion/secretion products of adult Schistosoma mansoni worms. This activity is stimulated in the presence of thiol compounds and is maximally active at acidic pH. To characterize further this proteolytic activity, lyophilized adult worms were extracted, and proteinases were isolated and purified. From extracts prepared in 0.2 M citrate buffer, pH 4.9, two proteinase species were purified to homogeneity by centrifugation, gel filtration, dialysis, and chromatofocusing chromatography. The proteinases, designated SMw32 and SMw28, have apparent molecular weights (SDS-PAGE) of 31,700 +/- 1400 and 27,800 +/- 1700, respectively. Both are thiol-dependent, acidic endopeptidases that cleave hemoglobin and a synthetic substrate, CBZ-arg-arg-AFC. A statistical comparison of amino acid compositions reveals that the proteinases are highly related.  相似文献   

9.
A sodium dodecyl sulfate (SDS)-urea polyacrylamide gel system was used to investigate certain properties of the subunits of the beef heart mitochondrial ATPase, (native F1, nF1). By examining the affects of urea concentration and acrylamide concentration upon the electrophoretic mobilities of the polypeptides comprising the nF1 enzyme, we have obtained conditions under which all five subunits are simultaneously resolved when the discontinuous buffer system of Laemmli is used (U. K. Laemmli (1970) Nature (London) 277, 680-685). The determination of the apparent molecular weights by analysis of Ferguson plots (K. A. Ferguson (1964) Metabolism 13, 985-1002) revealed that the addition of urea to the SDS gels resulted in a decrease in the apparent molecular weight of the beta subunit. A dramatic increase in the apparent molecular weight of the delta subunit was also brought about by the presence of urea in the SDS gels. In addition, the apparent molecular weight of both the alpha and the beta subunits was dependent upon the acrylamide concentration used, indicating that these subunits contain either areas highly resistant to denaturation by the combined action of urea and SDS, or covalent modifications leading to anomalous electrophoretic mobility. The results of experiments in which urea analogs were used indicate that the interactions of urea with the beta subunit involve the formation of hydrogen bonds between urea and regions of this subunit. On the other hand, the interactions of urea with the delta subunit are primarily of a hydrophobic nature, suggesting that these interactions could involve domains of the delta subunit required for binding of the coupling factor to the mitochondrial membrane.  相似文献   

10.
A hemagglutinin, with a molecular weight of 30,000 and expressing hemagglutinating activity which could not be inhibited by simple sugars and glycoproteins, was isolated from fresh fruiting bodies of the edible mushroom Lyophyllum shimeiji. The protein was adsorbed on CM-Sepharose even in 20 mM ammonium acetate (pH 5.5) containing 1 M NaCl and was desorbed by 20 mM ammonium bicarbonate (pH 9). The hemagglutinating activity was subsequently adsorbed on Mono S in 20 mM ammonium acetate (pH 5.5) and was desorbed by a linear gradient of 0.2-0.5 M NaCl in ammonium acetate buffer. The hemagglutinin exhibited a novel N-terminal sequence not found in any lectin and hemagglutinin reported so far. It was devoid of antifungal activity.  相似文献   

11.
D A Yphantis  T Arakawa 《Biochemistry》1987,26(17):5422-5427
Recombinant DNA derived human interferon gamma (IFN-gamma) from Escherichia coli was examined by equilibrium ultracentrifugation. Short-column equilibrium experiments at pH 6.9 in 0.1 M ammonium acetate buffer gave a z-average molecular weight of 33,500 +/- 1400 at infinite dilution, corresponding to 1.98 +/- 0.08 times the formula weight. Long- (2.6 mm) column experiments at pH 7.5 in 0.04 M imidazole buffer gave a molecular weight of 33,400 +/- 500. Under the latter conditions IFN-gamma behaves somewhat nonideally, with the departure from ideality accounted for by an effective (Donnan) charge of about 6+. No association of this dimer to form tetramer or higher polymers was observed, with the association constant for formation of tetramer from dimer K24 found to be less than 34 L mol-1. Similarly, no dissociation to monomers was observable, with the dissociation constant to monomer K21 being less than 5 X 10(-8) mol L-1. At pH 3.55 in 0.02 M buffer (acetate plus acetic acid), there was virtually complete dissociation of the dimer to monomer. Extreme nonideality was seen in this low ionic strength system, and the effective charge on the protein was estimated to be about 11+. The reduced molecular weight M(1 -upsilon rho) of the monomer was found to be about 4.09 +/- 0.20 kg mol-1; this corresponds to a molecular weight of 16,410 +/- 820, with the Scatchard definition of components. A small amount of a polymer with a molecular weight of about 0.5 X 10(6) was detected under these conditions.  相似文献   

12.
Soluble succinate dehydrogenase prepared by butanol extraction reacts with N-ethylmaleimide according to first-order kinetics with respect to both remaining active enzyme and the inhibitor concentration. Binding of the sulfhydryl groups of the enzyme prevents its alkylation by N-ethylmaleimide and inhibition by oxaloacetate. A kinetic analysis of the inactivation of alkylating reagent in the presence of succinate or malonate suggests that N-ethylmaleimide acts as a site-directed inhibitor. The apparent first-order rate constant of alkylation increases between pH 5.8 and 7.8 indicating a pKa value for the enzyme sulfhydryl group equal to 7.0 at 22 degrees C in 50 mM Tris-sufate buffer. Certain anions (phosphate, citrate, maleate and acetate) decrease the reactivity of the enzyme towards the alkylating reagent. Succinate/phenazine methosulfate reductase activity measured in the presence of a saturating concentration of succinate shows the same pH-dependence as the alkylation rate by N-ethylmaleimide. The mechanism of the first step of succinate oxidation, including a nucleophilic attack of substrate by the active-site sulfhydryl group, is discussed.  相似文献   

13.
A gene for photoactive yellow protein (PYP) was identified from the genome sequence of the extremely halophilic aerobic bacterium Salinibacter ruber (Sr). The sequence is distantly related to the prototypic PYP from Halorhodospira halophila (Hh) (37% identity) and contains most of the amino acid residues identified as necessary for function. However, the Sr pyp gene is not flanked by its two biosynthetic genes as in other species. To determine as to whether the Sr pyp gene encodes a functional protein, we cloned and expressed it in Escherichia coli, along with the genes for chromophore biosynthesis from Rhodobacter capsulatus. The Sr PYP has a 31-residue N-terminal extension as compared to other PYPs that appears to be important for dimerization; however, truncation of these extra residues did not change the spectral and photokinetic properties. Sr PYP has an absorption maximum at 431 nm, which is at shorter wavelengths than the prototypical Hh PYP (at 446 nm). It is also photoactive, being reversibly bleached by either blue or white light. The kinetics of dark recovery is slower than any of the PYPs reported to date (4.27 x 10(-4) s(-1) at pH 7.5). Sr PYP appears to have a normal photocycle with the I1 and I2 intermediates. The presence of the I2' intermediate is also inferred on the basis of the effects of temperature and alchohol on recovery. Sr PYP has an intermediate spectral form in equilibrium with the 431 nm form, similar to R. capsulatus PYP and the Y42F mutant of Hh PYP. Increasing ionic strength stabilizes the 431 nm form at the expense of the intermediate spectral form, and the kinetics of recovery is accelerated 6.4-fold between 0 and 3.5 M salt. This is observed with ions from both the chaotropic and the kosmotropic series. Ionic strength also stabilizes PYP against thermal denaturation, as the melting temperature is increased from 74 degrees C in buffer alone to 92 degrees C in 2 M KCl. Sr accumulates KCl in the cytoplasm, like Halobacterium, to balance osmotic pressure and has very acidic proteins. We thus believe that Sr PYP is an example of a halophilic protein that requires KCl to electrostatically screen the excess negative charge and stabilize the tertiary structure.  相似文献   

14.
15.
The reproducibility of the determination of the molecular weight of chitosans in the 90–210 kDa range (Mn) by analytical size exclusion chromatography with multi-angle laser light scattering (SEC-MALLS) was improved by reducing the salt concentration in the mobile phase from (0.3 M acetic acid, 0.2 M sodium acetate, and 0.8 mM sodium azide) to (0.15 M acetic acid, 0.1 M sodium acetate, and 0.4 mM sodium azide) using Tosoh TSKgel G6000PWXL and G5000PWXL columns in series. The variability of measured molecular weight was significantly reduced by lowering the acetate concentration in the mobile phase, while the average molecular weight did not change significantly. The coefficient of variation of the number-average molecular weight, CV(Mn), decreased from 7–12% to 3–6% upon mobile phase dilution. This reduced variability in molecular weight of chitosans obtained from SEC is a significant improvement when precise values of chitosan molecular weight are required, for example in stability studies where viscosity changes in concentrated chitosan solutions are assessed, and in gene delivery applications.  相似文献   

16.
Pressey R 《Plant physiology》1990,93(2):798-804
Anionic peroxidase from tomato (Lycopersicon esculentum) fruit oxidized indoleacetic acid (IAA) slowly in the presence of Mn2+ and dichlorophenol in acetate buffers. The addition of certain anions to the reaction mixture increased the rate of oxidation. Phosphate was one of the effective anions and exerted maximal activation at 0.1 molar. The most effective activator of tomato peroxidase was nitrilotriacetate (NTA) at an optimum concentration of 60 micromolar. Only 0.17 nanomolar peroxidase was needed to oxidize 0.1 micromole IAA/5 minutes in the presence of NTA compared to 650 nanomolar peroxidase for the same rate in the absence of NTA. Other effective anions were oxalate, pyrophosphate, malate, and citrate. Each activator exhibited an optimum concentration and higher concentrations were inhibitory. Anionic peroxidase from horseradish was activated by the same anions. A cationic peroxidase from horseradish and lactoperoxidase oxidized IAA in acetate buffer although anions activated these enzymes severalfold. Microperoxidase and other hematoporphrins also catalyzed IAA oxidation in the presence of anions. It is proposed that IAA oxidation by peroxidase may be important when vacuolar contents mix with peroxidase as during plant injury.  相似文献   

17.
Substitution of any of a variety of organic anions, including acetate, propionate, lactate, gluconate, and succinate, for chloride in the reactivation medium improves the motility of demembranated sperm of Tripneustes gratilla. At the optimum concentration of 0.20 N, all of these anions improve the duration of motility, with lactate and gluconate being the best. The Michaelis constant for beat frequency (Kmf) is lower (0.11-0.14 mM at 22 degrees C) in most of the organic anions than it is in Cl- (0.20 mM), and the minimum ATP concentration required to support oscillatory beating is reduced from 10 microM in chloride to 2 microM in acetate, which together indicate a greater affinity of the axonemal ATPase for MgATP2- in the organic anions media. The maximal beat frequency, fmax, is as high as 42 Hz in 0.2 N succinate compared to 31 Hz in Cl-, whereas the mean bend angle averages 2.8 rad in acetate compared to 2.4 rad in Cl-; these values give a calculated average velocity of tubule sliding of approximately 15 micron/s in acetate and succinate, which is approximately 30% greater than the value of 11 micron/s observed in chloride. The reactivated sperm are sixfold more sensitive to vanadate inhibition in 0.2 M acetate than they are in 0.15 M Cl-. The specific ATPase activity of soluble dynein 1, which increases more than 15-fold between 0 and 1.0 N Cl-, undergoes only a twofold activation over the same range of organic anion concentration, and, like the reactivated motility, is up to 50-fold more sensitive to vanadate. This greater apparent mechanochemical efficiency and the increased sensitivity to vanadate inhibition in the organic anions suggest that they, unlike chloride, do not promote the spontaneous dissociation of ADP and PO4(3-) from the dynein-ADP-PO4 kinetic intermediate in the dynein crossbridge cycle. The use of organic anion media may lead to significant improvements in reactivation of other motile and transport systems.  相似文献   

18.
The study of the acidic lipid requirement of human spleen glucocerebrosidase was extended to include two new series of acidic lipids, namely, monoacylglycol sulfates and diacylglycerol sulfates. Lysosomal glucocerebrosidase was extracted with sodium cholate and 1-butanol to render its beta-glucosidase activity dependent upon exogenous lipids. Maximum reactivation of control glucocerebrosidase was obtained with nonanoylglycol sulfate (NGS) and diheptanoylglycerol sulfate (DHGS). However, the effects of these lipids were markedly dependent on the nature of buffer used in the assay medium; specifically, 0.2 M sodium citrate-phosphate (pH 5.5) was much more effective than 0.2 M sodium acetate (pH 5.5) in permitting these lipids to reactivate glucocerebrosidase. In contrast, the marked activation of glucocerebrosidase by phosphatidylserine and galactocerebroside 3-sulfate (sulfatide) that was achievable in the sodium acetate buffer was totally inhibited by citrate or phosphate ions. The effects of NGS and DHGS on the kinetic parameters of control glucocerebrosidase were to lower the Km for the substrate, 4-methylumbelliferyl-beta-D-glucoside from 5.5 mM to approximately 2 mM (in sodium citrate-phosphate buffer) and markedly increase the Vmax. Furthermore, with DHGS, significant activation was achieved at concentrations below the lipid's critical micellar concentration. None of the monoacylglycol- or diacylglycerol sulfates were capable of stimulating mutant glucocerebrosidases from either type 1 (Ashkenazi-Jewish) or type 2 Gaucher's disease patients. Like control glucocerebrosidase, the type 1 glucocerebrosidase was unresponsive to phosphatidylserine and sulfatide when the beta-glucosidase assay was conducted in 0.2 M sodium citrate-phosphate buffer. Based on the differential action of these lipid activators in the two buffers and their effects on the mutant enzymes, we propose that, with regard to the lipid requirement of glucocerebrosidase, there are two classes of acidic lipids--one comprised of phosphatidylserine and sulfatide and the other comprised of the likes of NGS, DHGS, or sodium taurodeoxycholate. It appears that control glucocerebrosidase and the mutant enzyme of the patient with type 1 Gaucher's disease is reconstitutable with the first class of lipids whereas the glucocerebrosidase of the type 2 patient is not. The observations in this report are interpreted in terms of a model which postulates that normal glucocerebrosidase possesses at least two distinct lipid binding domains.  相似文献   

19.
We report a comparative study of the isomerization reaction in native and denatured photoactive yellow protein (PYP) and in various chromophore analogues in their trans deprotonated form. The excited-state relaxation dynamics was followed by subpicosecond transient absorption and gain spectroscopy. The free p-hydroxycinnamate (pCA(2-)) and its amide analogue (pCM(-)) are found to display a quite different transient spectroscopy from that of PYP. The excited-state deactivation leads to the formation of the ground-state cis isomer without any detectable intermediate with a mechanism comparable to trans-stilbene photoisomerization. On the contrary, the early stage of the excited-state deactivation of the free thiophenyl-p-hydroxycinnamate (pCT(-)) and of the denatured PYP is similar to that of the native protein. It involves the formation of an intermediate absorbing in the spectral region located between the bleaching and gain bands in less than 2 ps. However, in these two cases, the formation of the cis isomer has not been proved yet. This difference with pCA(-) and pCM(-) might result from the fact that, in the thioester substituted chromophore, simultaneous population of two quasi-degenerate excited states occurs upon excitation. This comparative study highlights the determining role of the chromophore structure and of its intrinsic properties in the primary molecular events in native PYP.  相似文献   

20.
Concentration dependence of NaCl salting of 0-1.5 mM lysozyme solution in 0.1 M sodium acetate buffer, pH 4.25, was investigated for NaCl concentration varying up to 0.9 M. Calorimetric experiments demonstrated that depending on the salt concentration the estimated number of the binding sites on the lysozyme surface varied in the range of 5 up to 13, and the increase of salt concentration caused the decrease of the number of accessible sites. The small, but significant, local maximum centered at 0.63 M NaCl concentration indicated the specific salting-out of the lysozyme accompanied by binding of approximately 2-3 chloride anions. Generalized McMillan and Mayer's approach reduced to the third-order virial coefficients demonstrates the domination of lysozyme aggregation upon salt addition (a(21)-h(xxy)) and salt organization on the lysozyme surface (a(12)-h(xyy)) processes.  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