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.
l-beta-lysine and beta-glutamate are produced by the actions of lysine 2,3-aminomutase and glutamate 2,3-aminomutase, respectively. The pK(a) values have been titrimetrically measured and are for l-beta-lysine: pK(1)=3.25 (carboxyl), pK(2)=9.30 (beta-aminium), and pK(3)=10.5 (epsilon-aminium). For beta-glutamate the values are pK(1)=3.13 (carboxyl), pK(2)=3.73 (carboxyl), and pK(3)=10.1 (beta-aminium). The equilibrium constants for reactions of 2,3-aminomutases favor the beta-isomers. The pH and temperature dependencies of K(eq) have been measured for the reaction of lysine 2,3-aminomutase to determine the basis for preferential formation of beta-lysine. The value of K(eq) (8.5 at 37 degrees C) is independent of pH between pH 6 and pH 11; ruling out differences in pK-values as the basis for the equilibrium constant. The K(eq)-value is temperature-dependent and ranges from 10.9 at 4 degrees C to 6.8 at 65 degrees C. The linear van't Hoff plot shows the reaction to be enthalpy-driven, with DeltaH degrees =-1.4 kcal mol(-1) and DeltaS degrees =-0.25 cal deg(-1) mol(-1). Exothermicity is attributed to the greater strength of the bond C(beta)-N(beta) in l-beta-lysine than C(alpha)-N(alpha) in l-lysine, and this should hold for other amino acids.  相似文献   

2.
The pH dependence of the binding constant of (GlcNAc)3 to Asp 52-esterified lysozyme was determined by the fluorescence technique. The pK values of Asp 101 in the modified lysozyme and its complex with (GlcNAc)3 were determined to be 4.5 and 3.6, respectively, at 25 degrees C and 0.1 ionic strength. This result is different from that obtained by Parsons and Raftery ((1972) Biochemistry 11, 1633--1638), who observed no pK shift of Asp 101. The macroscopic pK value of Asp 52 in intact lysozyme determined by them using the pH difference titration data of Asp 52-esterified lysozyme relative to intact lysozyme ((1972) Biochemistry 11, 1623--1629) was 4.5, which is higher by about one pH unit than the pK value determined by our group (Kuramitsu et al. (1974) J. Biochem. 76, 671--683; (1977) ibid. 82, 585--597; (1978) ibid. 83, 159--170. We found that their pH difference titration data in the absence and presence of saccharides can be consistently interpreted in terms of our pK values of Asp 52, Glu 35, and Asp 101, if we assume that the pK value of another ionizable group (probably Asp 48) is perturbed on esterification of Asp 52.  相似文献   

3.
Investigation of the active site of papain with fluorescent probes   总被引:6,自引:5,他引:1       下载免费PDF全文
7-Chloro-4-nitrobenzo-2-oxa-1,3-diazole (NBD chloride) and 7-(2'-hydroxyethylthio)-NBD (obtained from NBD chloride and mercaptoethanol) undergo a reversible spectral change in alkaline solution that depends respectively on a single apparent pK(a) 9.76 (at 25 degrees C) and 8.81 (at 32 degrees C). In acid solution however no spectral change was observed. NBD chloride reacts slowly with papain at pH7, but the rate of inhibition increases at lower pH and depends on an apparent pK(a) of 3.7 (at 35 degrees C), which has been tentatively assigned to the carboxyl group of aspartic acid-158. The spectral properties of NBD-papain indicate that the thiol group of cysteine-25 is the site of reaction. The intensity of the fluorescence-emission spectrum of NBD-papain depends on a single pK(a) of 4.2 (at 26.7 degrees C). The intensity of the fluorescence-emission spectrum of the mixed disulphide formed from papain and 7-(2'-mercaptoethylamino)-NBD (obtained from NBD chloride and cysteamine) depended on a single pK(a) of 3.94 in water and 3.89 in aq. 19.2% (v/v) dioxan (at 27 degrees C). This small change to lower pK(a) value in a medium of lower dielectric constant is characteristic of a cationic acid. The only acid of this type in the active-site region is the conjugate acid of histidine-159.  相似文献   

4.
The difference absorption spectra of hen and turkey lysozymes in the alkaline pH region had three maxima at around 245, 292, and 300 nm and had no isosbestic points. The ratio of the extinction difference at 245 nm to that at 295 nm changed with pH. These spectral features are quite different from those observed when only tyrosyl residues are ionized, and it was impossible to determine precisely the pK values of the tyrosyl residues in lysozyme by spectrophotometric titration. A time-dependent spectral change was observed above about pH 12. This is not due to exposure of a buried tyrosyl residue on alkali denaturation. The disulfide bonds and the peptide bonds in the lysozyme molecule were cleaved by alkali above about pH 11. The intrinsic pK value of Tyr 23 of hen lysozyme was determined to be 10.24 (apparent pK 9.8) at 0.1 ionic strength and 25 degrees C from the CD titration data. Comparison of the CD titration of turkey lysozyme with that of hen lysozyme suggested that Tyr 3 and Tyr 23 in turkey lysozyme have apparent pK values of 11.9 and 9.8, respectively.  相似文献   

5.
The ionization and phase behavior of 1,2-dipalmitoyl-sn-glycero-3-phosphoserine have been investigated under a variety of condtions by several different methods. As measured by turbidity changes, the temperature of the crystal-liquid crystal phase transition of this lipid is influenced by pH and mono- and divalent cation concentrations. The pH-transition temperature curve is congruent with the curve relating temperature to the degree of ionization of the carboxyl group of the crystalline form. The transition temperature falls from an upper plateau of 72 degrees C at low pH values, where the carboxyl group is fully protonated, to a lower plateau of 55 degrees C at high pH values, where this group is fully ionized. The apparent pK (pH at 50% ionization) of the crystalline form shifts from 6.0 to 4.6 to 3.7 with an increase of NaCl concentration from 10(-3) to 0.1 to l.0 M, respectively. These observations are in accord with a simple theoretical analysis that utilizes diffuse double layer theory and the influence of surface potential on surface concentration of protons. In qualitative terms, an increase in electrolyte concentration reduces the surface potential, the result of which is a diminution of the surface-bulk pH difference and a lowering of the apparent pK. Assuming an area of 50 A2/molecule, the intrinsic pKa (apparent pK corrected for surface pH) of the carboxyl group is 2.7. A 1000-fold change of NaCl concentration produces a very large change in surface potential without influencing the transition temperature of the ionized form of the lipid.  相似文献   

6.
Earlier studies have reported that trimethylamine N-oxide (TMAO), a naturally occurring osmolyte, is a universal stabilizer of proteins because it folds unstructured proteins and counteracts the deleterious effects of urea and salts on the structure and function of proteins. This conclusion has been reached from the studies of the effect of TMAO on proteins in the pH range 6.0-8.0. In this pH range TMAO is almost neutral (zwitterionic form), for it has a pK(a) of 4.66 +/- 0.10. We have asked the question of whether the effect of TMAO on protein stability is pH-dependent. To answer this question we have carried out thermal denaturation studies of lysozyme, ribonuclease-A, and apo-alpha-lactalbumin in the presence of various TMAO concentrations at different pH values above and below the pK(a) of TMAO. The main conclusion of this study is that near room temperature TMAO destabilizes proteins at pH values below its pK(a), whereas it stabilizes proteins at pH values above its pK(a). This conclusion was reached by determining the T(m) (midpoint of denaturation), delta H(m) (denaturational enthalpy change at T(m)), delta C(p) (constant pressure heat capacity change), and delta G(D) degrees (denaturational Gibbs energy change at 25 degrees C) of proteins in the presence of different TMAO concentrations. Other conclusions of this study are that T(m) and delta G(D) degrees depend on TMAO concentration at each pH value and that delta H(m) and the delta C(p) are not significantly changed in presence of TMAO.  相似文献   

7.
The influence of pH within the range 6.9–10.0 on the kinetic parameters of Micrococcus lysodeicticus cell lysis catalyzed by hen egg lysozyme has been studied at 25°C and 37°C. The effective pK b values have been calculated for the group determining lysozyme catalytic activity. The ΔH ion value indicates that this group is a carboxyl, although its pK (9.15 at 25°C) is far beyond the range characteristic of carboxylic groups. The cause of this abnormal pK b value is supposed to be the strong negative charge of the bacterial cell wall. As a result, the enzyme, which catalyzes the hydrolysis of N-acetylglucosamine-N-acetylmuramic acid copolymer, operates in a highly acidic microenvironment.  相似文献   

8.
Unlike mammalian oxymyoglobins, Aplysia MbO2 is extremely susceptible to autoxidation, and its pH dependence is also unusual. Kinetic formulation has revealed that two kinds of dissociable group with pK1 = 4.3 and pK2 = 6.1, respectively, at 25 degrees C are involved in the stability property of Aplysia MbO2. In order to characterize thermodynamically these dissociation processes involved, the effect of temperature on K1 and K2 was studied by analyzing the pH dependence for the autoxidation rate of Aplysia MbO2 in 0.1 M buffer over the pH range of 4-11, and at 15, 25 and 35 degrees C. The resulting thermodynamic parameters for each group were both those to be expected for the ionization of a carboxyl group; the delta H degrees value being numerically much less than 1 kcal.mol-1, or zero in practice, but being associated with a large negative value of delta S degrees of the order of -20 cal.mol-1.K-1. Taking into account the fact that Aplysia myoglobin contains only a single histidine residue corresponding to the heme-binding proximal one, we can unequivocally conclude that the two kinds of the dissociable group involved in the unusual stability of Aplysia MbO2 must both be carboxyl groups, the protonation of these groups being responsible for an increase in its autoxidation rate in the acidic pH range.  相似文献   

9.
G Cevc  A Watts  D Marsh 《Biochemistry》1981,20(17):4955-4965
The dependence of the gel-to-fluid phase transition temperature of dimyristoyl- and dipalmitoylphosphatidylserine bilayers on pH, NaCl concentration, and degree of hydration has been studied with differential scanning calorimetry and with spin-labels. On protonation of the carboxyl group (pK2app = 5.5), the transition temperature increases from 36 to 44 degrees C in the fully hydrated state of dimyristoylphosphatidylserine (from 54 to 62 degrees C for dipalmitoylphosphatidylserine), at ionic strength J = 0.1. In addition, at least two less hydrated states, differing progressively by 1 H2O/PS, are observed at low pH with transition temperatures of 48 and 52 degrees C for dimyristoyl- and 65 and 68.5 degrees C for dipalmitoylphosphatidylserine. On deprotonation of the amino group (pK3app = 11.55) the transition temperature decreases to approximately 15 degrees C for dimyristoyl- and 32 degrees C for dipalmitoylphosphatidylserine, and a pretransition is observed at approximately 6 degrees C (dimyristoylphosphatidylserine) and 21.5 degrees C (dipalmitoylphosphatidylserine), at J = 0.1. No titration of the transition is observed for the fully hydrated phosphate group down to pH less than or equal to 0.5, but it affinity for water binding decreases steeply at pH greater than or equal to 2.6. Increasing the NaCl concentration from 0.1 to 2.0 M increases the transition temperature of dimyristoyphosphatidylserine by approximately 8 degrees C at pH 7, by approximately 5 degrees at pH 13, and by approximately 0 degrees C at pH 1. These increases are attributed to the screening of the electrostatic titration-induced shifts in transition temperature. On a further increase of the NaCl concentration to 5.5 M, the transition temperature increases by an additional 9 degree C at pH 7, 13 degree C at pH 13, approximately 7 degree C in the fully hydrated state at pH 1, and approximately 4 and approximately 0 degree C in the two less hydrated states. These shifts are attributed to displacement of water of hydration by ion binding. From the salt dependence it is deduced that the transition temperature shift at the carboxyl titration can be accounted for completely by the surface charge and change in hydration of approximately 1 H2O/lipid, whereas that of the amino group titration arises mostly from other sources, probably hydrogen bonding. The shifts in pK (delta pK2 = 2.85, delta pK3 = 1.56) are consistent with a reduced polarity in the head-group region, corresponding to an effective dielectric constant epsilon approximately or equal to 30, together with surface potentials of psi congruent to -100 and -150 mV at the carboxyl and amino group pKs, respectively. The transition temperature of dimyristoylphosphatidylserine-water mixtures decreases by approximately 4 degree C each water/lipid molecule added, reaching a limiting value at a water content of approximately 9-10 H2O/lipid molecule.  相似文献   

10.
A systematic study of the H+ titration curve of purified ovomucoid was made at three temperatures (15, 25 and 35 degrees C) and three ionic strengths (0.05, 0.15 and 1.0). In all, 49 protons were dissociated reversibly in the pH range, 2.0-12.0. From the analysis of the results up to pH 12.0, the numbers of different dissociable groups per 28 300 g protein, together with their intrinsic pK values in parentheses were found tp be' 27 sode-chain carboxyl (pKint=4.0), four imidazole (pKint=6.5), one alpha-amino (pKint=7.5), 12 epsilon-amino (pKint=9.6), one guanidino (pKint=11.8) and one alpha-carboxyl group with abnormally low pK. The total number of basic nitrogens per mole of the protein was 22 so that four guanidino groups remained untitrated up to pH 12.0. Spectrophotometric titration showed that three out of five phenolic groups were titrated reversibly up to pH 11.9 with an intrinsic pK of 10.25; the remaining two groups became accessible only on protein denaturation. Viscosity results suggested absence of conformational change in the pH range 2.0-11.2. This explains the constancy of the pK values of carboxyl groups in the pH range 2.0-5.0. The empirical value of the electrostatic interaction factor, w, was 0.04, both in the carboxyl and phenolic regions.  相似文献   

11.
To study the interaction between carboxyl groups and amino groups in native lysozyme [EC 3.2.1.17], and to identify the positions and the pK values of the abnormal carboxyl groups, N-acetylated lysozyme was prepared. The acetylation did not affect the molecular shape of the enzyme, but changed six amino groups to a non-ionizable form, leaving one amino group free; this was determined to be Lys 33. In addition, pH titration of the acetylated lysozyme in 0.2 or 0.02 M KCl aqueous solution indicated fewer titratable groups with pK(int) of 7.8 or 10.4 compared with the native protein, though the number of titratable carboxyl groups was not affected by the acetylation. From the pH titration results and structural considerations, the unititratable carboxyl groups were suggested to be Asp 48, Asp 66, and Asp 87. On the other hand, spectrophotometric titration in 0.2 M KCl showed that all three tyrosine residues are titratable in the acetylated protein, although an abnormal tyrosine residue exists in the native state. Tyr 20 was suggested to be untitratable in the pH range of 8-12.6.  相似文献   

12.
The interaction of N-acetyl-chitotriose ((GlcNAc)3) with human lysozyme [EC 3.2.1.17] was studied at various pH values by measuring changes in the circular dichroic (CD) band at 294 or 255 nm and the data were compared with the results for hen and turkey lysozymes reported previously (Kuramitsu et al. (1974) J. Biochem.76, 671-683; Kuramitsu et al. (1975) J. Biochem. 77, 291-301). The pH dependence of the binding constant of (GlcNAc)3 to human lysozyme was different from those for hen and turkey lysozymes. The catalytic carboxyls of human lysozyme, Asp 52 and Glu 35, were not perturbed on binding of (GlcNAc)3. This is consistent with the previous findings that the macroscopic pK values of Asp 52 and Glu 35 of human lysozyme are 3.4 and 6.8 at 0.1 ionic strength and 25 degrees and were unchanged on complexing with (GlcNAc)3. An ionizable group with pK 4.5, which participates in the binding of (GlcNAc)3 to hen lysozyme and was assigned as Asp 101, did not participate in the binding of the saccharide to human lysozyme. Between pH 9 and 11, the binding constants of (GlcNAc)3 to hen lysozyme remained unchanged, whereas perturbation of an ionizable group with pK 10.5 to 10.0 was observed for human lysozyme. This group may be Tyr 62 in the active-site cleft. The binding constants of (GlcNAc)3 to human lysozyme molecules having different microscopic protonation forms, with respect to the catalytic carboxyls, were estimated using the binding constants obtained in the present experiments and the microscopic ionization constants of the catalytic carboxyls obtained previously. All four species of human lysozyme had similar binding constants to (GlcNAc)3. This result is different from those for hen and turkey lysozymes.  相似文献   

13.
Secretin has a single histidine residue located at the amino terminus which plays a crucial role in its biological activity. The chemical properties, viz. pK and reactivity, of the alpha-amino and imidazole groups of this residue were determined at a secretin concentration of 10(-6) M in 0.1 M KCl at 37 degrees C. Competitive labelling using tritiated 1-fluoro-2,4-dinitrobenzene (DNP-F) as the labelling reagent was the experimental approach employed. The alpha-amino group was found to have a pK value of 8.83 and a reactivity 5-times that of the alpha-amino group in the model compound, histidylglycine. For the imidazole function a pK value of 8.24 and a reactivity 26-times that of the imidazole function in histidylglycine was found. Both these groups in secretin had pK values which were shifted one pK unit higher than in histidylglycine, but like the model compound the reactivity of the imidazole function was still linked to the state of ionization of the alpha-amino group. These observations are interpreted as evidence for the existence of a major conformational state in dilute aqueous solution in which the amino-terminal histidine of secretion is interacting with a negatively charged carboxyl group.  相似文献   

14.
1. The reaction of exo-cis-3,6-endoxo-Delta(4)-tetrahydrophthalic anhydride with amino groups of model compounds and lysozyme is described. 2. Reaction with the in-amino group of N(alpha)-acetyl-l-lysine amide gives rise to two diastereoisomeric products; at acid pH the free amino group is liberated with anchimeric assistance by the neighbouring protonated carboxyl group with a half-time of 4-5h at pH3.0 and 25 degrees C. 3. The amino groups of lysozyme can be completely blocked, with total loss of enzymic activity. Dialysis at pH3.0 results in complete recovery of the native primary and tertiary structure of lysozyme and complete return of catalytic activity. 4. The specificity of reaction of this and other anhydrides with amino groups in proteins is discussed.  相似文献   

15.
To explore electrostatic interactions in ubiquitin, pK(a) values have been determined by NMR for all 12 carboxyl groups in wild-type ubiquitin and in variants where single lysines have been replaced by neutral residues. Aspartate pK(a) values in ubiquitin range from 3.1 to 3.8 and are generally less than model compound values. Most aspartate pK(a) values are within 0.2 pH unit of those predicted with a simple Tanford-Kirkwood model. Glutamate pK(a) values range from 3.8 to 4.5, close to model compound values and differing by 0.1-0.8 pH unit from calculated values. To determine the role of positive charges in modulating carboxyl pK(a) values, we mutated lysines at positions 11, 29, and 33 to glutamine and threonine. NMR studies with these six single-site mutants reveal significant interactions of Lys 11 and Lys 29 with Glu 34 and Asp 21, respectively: pK(a) values for Glu 34 and Asp 21 increase by approximately 0.5-0.8 pH unit, similar to predicted values, when the lysines are replaced by neutral residues. In contrast, the predicted interaction between Lys 33 and Glu 34 is not observed experimentally. In some instances, substitution of lysine by glutamine and threonine did not lead to the same changes in carboxyl pK(a) values. These may reflect new short-range interactions between the mutated residues and the carboxyl groups. Carboxyl pK(a) shifts > 0.5 pH unit result from mutations at groups that are <5 A from the carboxyl group. No interactions are observed at >10 A.  相似文献   

16.
The enzymic properties of urokinase (EC 3.4.21.31) were studied. The kinetic parameters of hydrolysis of 5-oxo-Pro-Gly-Arg-NA were determined in the pH range 5-9, at 25 degrees C and 37 degrees C. The reaction is affected by only one ionizing group of urokinase with pK 7.15 (25 degrees C) and pK 6.82 (37 degrees C). The results indicate that 5-oxo-Pro-Gly-Arg-NA is a good model substrate for studies of the conversion of plasminogen to plasmin. The Km values of the urokinase-catalysed hydrolysis of plasminogen and 5-oxo-Pro-Gly-Arg-NA are of the same order of magnitude. Plasmin catalyses the hydrolysis of 5-oxo-Pro-Gly-Arg-NA, but the Km value is several hundred times that of urokinase. Urokinase is shown not to react with good plasmin substrates, such as Bz-Arg-OEt and D-Val-Leu-Lys-NA, but is linearly competitively inhibited by 6-amino-hexanoic acid and trans-4-aminomethylcyclohexane-1-carboxylic acid.  相似文献   

17.
The pK values of the titratable groups in ribonuclease Sa (RNase Sa) (pI=3.5), and a charge-reversed variant with five carboxyl to lysine substitutions, 5K RNase Sa (pI=10.2), have been determined by NMR at 20 degrees C in 0.1M NaCl. In RNase Sa, 18 pK values and in 5K, 11 pK values were measured. The carboxyl group of Asp33, which is buried and forms three intramolecular hydrogen bonds in RNase Sa, has the lowest pK (2.4), whereas Asp79, which is also buried but does not form hydrogen bonds, has the most elevated pK (7.4). These results highlight the importance of desolvation and charge-dipole interactions in perturbing pK values of buried groups. Alkaline titration revealed that the terminal amine of RNase Sa and all eight tyrosine residues have significantly increased pK values relative to model compounds.A primary objective in this study was to investigate the influence of charge-charge interactions on the pK values by comparing results from RNase Sa with those from the 5K variant. The solution structures of the two proteins are very similar as revealed by NMR and other spectroscopic data, with only small changes at the N terminus and in the alpha-helix. Consequently, the ionizable groups will have similar environments in the two variants and desolvation and charge-dipole interactions will have comparable effects on the pK values of both. Their pK differences, therefore, are expected to be chiefly due to the different charge-charge interactions. As anticipated from its higher net charge, all measured pK values in 5K RNase are lowered relative to wild-type RNase Sa, with the largest decrease being 2.2 pH units for Glu14. The pK differences (pK(Sa)-pK(5K)) calculated using a simple model based on Coulomb's Law and a dielectric constant of 45 agree well with the experimental values. This demonstrates that the pK differences between wild-type and 5K RNase Sa are mainly due to changes in the electrostatic interactions between the ionizable groups. pK values calculated using Coulomb's Law also showed a good correlation (R=0.83) with experimental values. The more complex model based on a finite-difference solution to the Poisson-Boltzmann equation, which considers desolvation and charge-dipole interactions in addition to charge-charge interactions, was also used to calculate pK values. Surprisingly, these values are more poorly correlated (R=0.65) with the values from experiment. Taken together, the results are evidence that charge-charge interactions are the chief perturbant of the pK values of ionizable groups on the protein surface, which is where the majority of the ionizable groups are positioned in proteins.  相似文献   

18.
The pH dependence of the binding constant of Ca2+ to a phospholipase A2 of Agkistrodon halys blomhoffii, in which the alpha-amino group had been selectively modified to an alpha-keto group, was studied at 25 degrees C and ionic strength 0.1 by the tryptophyl fluorescence method. The dependence was compared with the results for the intact enzyme (Ikeda et al. (1981) J. Biochem. 90, 1125-1130). The pH-dependence curve could be well interpreted in terms of the participation of the two ionizable groups Asp 49 and His 48, with pK values of 4.70 and 6.69, respectively. These values were slightly different from the respective pK values for the intact enzyme, 5.15 and 6.45. Ca2+ binding to the intact enzyme involves the participation of an additional ionizable group with a pK value of 7.30, which was thus assigned as alpha-amino group. The pH dependence of the binding constant of monodispersed n-dodecylphosphorylcholine (n-C12PC) to the alpha-NH2-modified enzyme was studied at 25 degrees C and ionic strength 0.1 by the aromatic circular dichroism (CD) method. The pH-dependence curve for the modified apoenzyme was interpreted as reflecting the participation of a single ionizable group with a pK value of 4.7, which was assigned to Asp 49 (to which a Ca2+ ion can coordinate) since the curve for the Ca2+ complex lacked this transition: the binding constant was independent of pH. The pH-dependence curves for the intact apoenzyme and its Ca2+ complex involve the participation of an additional ionizable group with pK values of 7.30 and 6.30, respectively (Ikeda & Samejima (1981) J. Biochem. 90, 799-804), which was assigned as the alpha-amino group. The hydrolysis of monodispersed 1,2-dihexanoyl-sn-glycero-3-phosphorylcholine (diC6PC), catalyzed by the intact and the alpha-NH2-modified enzymes was studied by the pH stat method at 25 degrees C, pH 8.2, and ionic strength 0.1 in the presence of 3 mM Ca2+. The Km value for the modified enzyme was found to be very similar to that for the intact enzyme: this was compatible with the results of the direct binding study on the monodispersed n-C12PC under the same conditions. However, the kcat value was about 43% of the value for the intact enzyme, suggesting that the alpha-keto group introduced by the chemical modification perturbed the network of hydrogen bonds in the active site.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

19.
Electrostatic binding sites of extracellular polymeric substances (EPS) were characterized from titration data using linear programming analysis. Test results for three synthetic solutions of given solutes comprising amino, carboxyl, and phenolic groups indicated that this method was able to identify the electrostatic binding sites. For the six sites with pK(a) between 3 and 10, the estimated pK(a) deviated 0.11 +/- 0.09 from the theoretical values, and the estimated concentrations deviated 3.0% +/- 0.9% from the actual concentrations. Two EPS samples were then extracted from a hydrogen-producing sludge (HPS) and a sulfate-reducing biofilm (SRB). Analysis of charge excess data in titration from pH 3 to 11 indicated that the EPS of HPS comprised of five electrostatic binding sites with pK(a) ranging from 3 to 11. The pK(a) values of these binding sites and the possible corresponding functional groups were pK(a) 4.8 (carboxyl), pK(a) 6.0 (carboxyl/phosphoric), pK(a) 7.0 (phosphoric), pK(a) 9.8 (amine/phenolic), and pK(a) 11.0 (hydroxyl). EPS of the SRB comprised five of similar binding sites (with corresponding pK(a) values of 4.4, 6.0, 7.4, 9.4, and 11.0), plus one extra site at pK(a) 8.2, which was likely corresponding to the sulfhydryl group. The total electrostatic binding site concentration of EPS extracted from HPS were 10.88 mmol/g-EPS, of which the highest concentration was from the site of pK(a) 11.0. The corresponding values for the EPS extracted from SRB were 16.44 mmol/g-EPS and pK(a) 4.4. The total concentrations of electrostatic binding sites found in this study were 20- to 30-fold of those reported for bacterial cell surface, implying that EPS might be more crucial in biosorption of metals than bacterial cell surface in wastewater treatment and in bioremediation.  相似文献   

20.
Experimentally determined mean pK(a) values of carboxyl residues located at the N-termini of alpha-helices are lower than their overall mean values. Here, we perform three types of analyses to account for this phenomenon. We estimate the magnitude of the helix macrodipole to determine its potential role in lowering carboxyl pK(a) values at the N-termini. No correlation between the magnitude of the macrodipole and the pK(a) values is observed. Using the pK(a) program propKa we compare the molecular surroundings of 18 N-termini carboxyl residues versus 233 protein carboxyl groups from a previously studied database. Although pK(a) lowering interactions at the N-termini are similar in nature to those encountered in other protein regions, pK(a) lowering backbone and side-chain hydrogen bonds appear in greater number at the N-termini. For both Asp and Glu, there are about 0.5 more hydrogen bonds per residue at the N-termini than in other protein regions, which can be used to explain their lower than average pK(a) values. Using a QM-based pK(a) prediction model, we investigate the chemical environment of the two lowest Asp and the two lowest Glu pK(a) values at the N-termini so as to quantify the effect of various pK(a) determinants. We show that local interactions suffice to account for the acidity of carboxyl residues at the N-termini. The effect of the helix dipole on carboxyl pK(a) values, if any, is marginal. Backbone amide hydrogen bonds constitute the single biggest contributor to the lowest carboxyl pK(a) values at the N-termini. Their estimated pK(a) lowering effects range from about 1.0 to 1.9 pK(a) units.  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