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.
Detailed knowledge of the pH-dependence in both folded and unfolded states of proteins is essential to understand the role of electrostatics in protein stability. The increasing number of natively disordered proteins constitutes an excellent source for the NMR analysis of pKa values in the unfolded state of proteins. However, the tendency of many natively disordered proteins to aggregate via intermolecular hydrophobic clusters limits their NMR analysis over a wide pH range. To assess whether the pKa values in natively disordered polypeptides can be extrapolated from NMR measurements in the presence of denaturants, the natively disordered backbone of the C-terminal fragment 75 to 105 of Human Thioredoxin was studied. First, assignments using triple resonance experiments were performed to confirm lack of secondary structure. Then the pH-dependence of the amides and carboxylate side chains of Glu residues (Glu88, Glu95, Glu98, and Glu103) in the pH range from 2.0 to 7.0 was monitored using 2D 1H15N HSQC and 3D C(CO)NH experiments, and the behavior of their amides and corresponding carboxyl groups was compared to confirm the absence of nonlocal interactions. Lastly, the effect of increasing dimethyl urea concentration on the pKa values of these Glu residues was monitored. The results indicate that: (i) the dispersion in the pKa of carboxyl groups and the pH midpoints of amides in Glu residues is about 0.5 pH units and 0.6 pH units, respectively; (ii) the backbone amides of the Glu residues exhibit pH midpoints which are within 0.2 pH units from those of their carboxylates; (iii) the addition of denaturant produces upshifts in the pKa values of Glu residues that are nearly independent of their position in the sequence; and (iv) these upshifts show a nonlinear behavior in denaturant concentration, complicating the extrapolation to zero denaturant. Nevertheless, the relative ordering of the pKa values of Glu residues is preserved over the whole range of denaturant concentrations indicating that measurements at high denaturant concentration (e.g. 4 M dimethyl urea) can yield a qualitatively correct ranking of the pKa of these residues in natively disordered proteins whose pH-dependence cannot be monitored directly by NMR.  相似文献   

2.
A growing number of natively disordered proteins undergo a folding/binding process that is essential for their biological function. An interesting question is whether these proteins have incompletely solvated regions that drive the folding/binding process. Although the presence of predominantly hydrophobic buried regions can be easily ascertained by high-sensitivity differential scanning calorimetry analysis, the identification of those residues implicated in the burial requires NMR analysis. We have selected a partially solvated natively disordered fragment of Escherichia coli, thioredoxin, C37 (38-108), for full NMR spectral assignment. The secondary chemical shifts, temperature coefficients, and relaxation rates (R(1) and R(2)) of this fragment indicate the presence of a flexible backbone without a stable hydrogen bond network near neutral pH. (1)H-(15)N heteronuclear single quantum coherence analysis of the pH dependence of amide chemical shifts in fragment C37 within pH 2.0 and 7.0 suggests the presence of interactions between nonionizable residues and the carboxylate groups of four Asp and four Glu residues. The pH midpoints (pH(m)) of the amides in the ionizable residues (Asp or Glu) and, consequently, the shifts in the pH(m) (DeltapH(m)) of these residues with respect to model tetrapeptides, are sequence-dependent; and the nonionizable residues that show pH dependence cluster around the ionizable ones. The same pH dependence has been observed in two fragments: M37 (38-73) and C73 (74-108), ruling out the participation of long-range interactions. Our studies indicate the presence of a 15-residue pH-dependent segment with the highest density of ionizable sites in the disordered ensembles of fragments C37 and M37. The observed correlations between ionizable and nonionizable residues in this segment suggest the organization of the backbone and side chains through local and medium-range interactions up to nine residues apart, in contrast to only a few interactions in fragment C73. These results agree qualitatively with the predominantly hydrophobic buried surface detected only in fragments C37 and M37 by highly sensitive differential scanning calorimetry analysis. This work offers a sensitive and rapid new tool to obtain clues about local and nonlocal interactions between ionizable and nonionizable residues in the growing family of natively disordered small proteins with full NMR assignments.  相似文献   

3.
The interactions of the substrate analogs beta-methyl-GlcNAc, (GlcNAc)2, and (GlcNAc)3 with hen egg-white lysozyme [EC 3.2.1.17] in which an ester linkage had been formed between Glu 35 and Trp 108 (108 ester lysozyme), were studied by the circular dichroic and fluorescence techniques, and were compared with those for intact lysozyme. The binding constants of beta-methyl-GlcNAc and (GlcNAc)2 to 108 ester lysozyme were essentially the same as those for intact lysozyme in the pH range of 1 to 5. Above pH 5, the binding constants of these saccharides to 108 ester lysozyme did not change with pH, while the binding constants to intact lysozyme decreased. This indicates that Glu 35 (pK 6.0 in intact lysozyme) participates in the binding of these saccharides. The extent and direction of the pK shifts of Asp 52 (pK 3.5), Asp 48 (pK 4.4), and Asp 66 (pK 1.3) observed when beta-methyl-GlcNAc is bound to 108 ester lysozyme were the same as those for intact lysozyme. The participation of Asp 101 and Asp 66 in the binding of (GlcNAc)2 to 108 ester lysozyme was also the same as that for intact lysozyme. These findings indicate that the conformations of subsites B and C are not changed by the formation of the ester linkage. On the other hand, the binding constants of (GlcNAc)3 to 108 ester lysozyme were higher than those for intact lysozyme at all pH values studied. This result is interpreted in terms of an increase in the affinity for a GlcNAc residue of subsite D, which is situated near the esterified Glu 35.  相似文献   

4.
In green fluorescent protein (GFP), chromophore biosynthesis is initiated by a spontaneous main-chain condensation reaction. Nucleophilic addition of the Gly67 amide nitrogen to the Ser65 carbonyl carbon is catalyzed by the protein fold and leads to a heterocyclic intermediate. To investigate this mechanism, we substituted the highly conserved residues Arg96 and Glu222 in enhanced GFP (EGFP). In the R96M variant, the rate of chromophore formation is greatly reduced (time constant = 7.5 x 10(3) h, pH 7) and exhibits pH dependence. In the E222Q variant, the rate is also attenuated at physiological pH (32 h, pH 7) but is accelerated severalfold beyond that of EGFP at pH 9-10. In contrast, EGFP maturation is pH-independent and proceeds with a time constant of 1 h (pH 7-10). Mass spectrometric results for R96M and E222Q indicate accumulation of the pre-cyclization state, consistent with rate-limiting backbone condensation. The pH-rate profile implies that the Glu222 carboxylate titrates with an apparent pK(a) of 6.5 in R96M and that the Gly67 amide nitrogen titrates with an apparent pK(a) of 9.2 in E222Q. These data suggest a model for GFP chromophore synthesis in which the carboxylate of Glu222 plays the role of a general base, facilitating proton abstraction from the Gly67 amide nitrogen or the Tyr66 alpha-carbon. Arg96 fulfills the role of an electrophile by lowering the respective pK values and stabilizing the alpha-enolate. Modulating the base strength of the proton-abstracting group may aid in the design of fast-maturing GFPs with improved characteristics for real-time monitoring of cellular events.  相似文献   

5.
Sac7d and Sso7d are homologous, hyperthermophile proteins with a high density of charged surface residues and potential ion pairs. To determine the relative importance of specific amino acid side-chains in defining the stability and function of these Archaeal chromatin proteins, pK(a) values were measured for the acidic residues in both proteins using (13)C NMR chemical shifts. The stability of Sso7d enabled titrations to pH 1 under low-salt conditions. Two aspartate residues in Sso7d (D16 and D35) and a single glutamate residue (G54) showed significantly perturbed pK(a) values in low salt, indicating that the observed pH-dependence of stability was primarily due to these three residues. The pH-dependence of backbone amide NMR resonances demonstrated that perturbation of all three pK(a) values was primarily the result of side-chain to backbone amide hydrogen bonds. Few of the significantly perturbed acidic pK(a) values in Sac7d and Sso7d could be attributed to primarily ion pair or electrostatic interactions. A smaller perturbation of E48 (E47 in Sac7d) was ascribed to an ion pair interaction that may be important in defining the DNA binding surface. The small number (three) of significantly altered pK(a) values was in good agreement with a linkage analysis of the temperature, pH, and salt-dependence of folding. The linkage of the ionization of two or more side-chains to protein folding led to apparent cooperativity in the pH-dependence of folding, although each group titrated independently with a Hill coefficient near unity. These results demonstrate that the acid pH-dependence of protein stability in these hyperthermophile proteins is due to independent titration of acidic residues with pK(a) values perturbed primarily by hydrogen bonding of the side-chain to the backbone. This work demonstrates the need for caution in using structural data alone to argue the importance of ion pairs in stabilizing hyperthermophile proteins.  相似文献   

6.
Site-specific pK(a) values measured by NMR spectroscopy provide essential information on protein electrostatics, the pH-dependence of protein structure, dynamics and function, and constitute an important benchmark for protein pK(a) calculation algorithms. Titration curves can be measured by tracking the NMR chemical shifts of several reporter nuclei versus sample pH. However, careful analysis of these curves is needed to extract residue-specific pK(a) values since pH-dependent chemical shift changes can arise from many sources, including through-bond inductive effects, through-space electric field effects, and conformational changes. We have re-measured titration curves for all carboxylates and His 15 in Hen Egg White Lysozyme (HEWL) by recording the pH-dependent chemical shifts of all backbone amide nitrogens and protons, Asp/Glu side chain protons and carboxyl carbons, and imidazole protonated carbons and protons in this protein. We extracted pK(a) values from the resulting titration curves using standard fitting methods, and compared these values to each other, and with those measured previously by 1H NMR (Bartik et al., Biophys J 1994;66:1180–1184). This analysis gives insights into the true accuracy associated with experimentally measured pK(a) values. We find that apparent pK(a) values frequently differ by 0.5–1.0 units depending upon the nuclei monitored, and that larger differences occasionally can be observed. The variation in measured pK(a) values, which reflects the difficulty in fitting and assigning pH-dependent chemical shifts to specific ionization equilibria, has significant implications for the experimental procedures used for measuring protein pK(a) values, for the benchmarking of protein pK(a) calculation algorithms, and for the understanding of protein electrostatics in general.  相似文献   

7.
Li H  Robertson AD  Jensen JH 《Proteins》2004,55(3):689-704
A computational methodology for protein pK(a) predictions, based on ab initio quantum mechanical treatment of part of the protein and linear Poisson-Boltzmann equation treatment of the bulk solvent, is presented. The method is used to predict and interpret the pK(a) values of the five carboxyl residues (Asp7, Glu10, Glu19, Asp27, and Glu43) in the serine protease inhibitor turkey ovomucoid third domain. All the predicted pK(a) values are within 0.5 pH units of experiment, with a root-mean-square deviation of 0.31 pH units. We show that the decreased pK(a) values observed for some of the residues are primarily due to hydrogen bonds to the carboxyl oxygens. Hydrogen bonds involving amide protons are shown to be particularly important, and the effect of hydrogen bonding is shown to be nonadditive. Hydrophobic effects are also shown to be important in raising the pK(a). Interactions with charged residues are shown to have relatively little effect on the carboxyl pK(a) values in this protein, in general agreement with experiment.  相似文献   

8.
Bacterial L-asparaginases (E.C. 3.5.1.1) have been used as therapeutic agents in the treatment of acute childhood lymphoblastic leukaemia. L-asparaginase from Erwinia carotovora NCYC 1526 (ErA) was cloned and expressed in E. coli. The enzyme was purified to homogeneity by a two-step procedure comprising cation-exchange chromatography and affinity chromatography on immobilised L-asparagine. The enzymatic properties of the recombinant enzyme were investigated and the kinetic parameters (K(m), k(cat)) for a number of substrates were determined. Molecular modelling studies were also employed to create a model of ErA, based on the known structure of the Erwinia chrysanthemi enzyme. The molecular model was used to help interpret biochemical data concerning substrate specificity and catalytic mechanism of the enzyme. The kinetic parameters of selected substrates were determined at various pH values, and the pH-dependence profiles of V(max) and V(max)/K(m) were analyzed. The pH-dependence of V(max) shows one transition in the acidic pH range with pK(a)=5.4, and the pH-dependence of V(max)/K(m) exhibits two transitions with pK(a)=5.4 and 8.5. Based on analysis of alternative substrates and molecular modelling studies, it was concluded that the pK(a) at the acidic pH range corresponds to the active site residues Asp115 or Glu82, whereas the pK(a) observed at the alkaline pH range is not due to substrate amino group ionisation, but rather is the result of enzyme ionisation. The effect of temperature and viscosity on the catalytic activity of the enzyme was also investigated and it was concluded that the rate-limiting step of the catalytic reaction is relevant to structural transitions of the protein. Thermodynamic analysis of the activity data showed that the activation energies are dependent on the substrate, and entropy changes appear to be the main determinant contributing to substrate specificity.  相似文献   

9.
A A Ismail  H H Mantsch 《Biopolymers》1992,32(9):1181-1186
The carboxylate-containing homopolypeptides poly(L-glutamate) [poly(Glu)] and poly(L-aspartate) [poly(Asp)] were found to form different types of ordered structures in the presence of poly(L-lysine) [poly(Lys)]. Mixing poly(Glu) with poly(Lys) in aqueous solution at neutral pH results in the instantaneous formation of a gel-like precipitate. The secondary structure of the gel precipitate can be best described as intermolecular antiparallel beta-strands, involving the backbone amide groups, as evidenced by the presence of characteristic amide I bands in the ir spectrum at 1684 and 1612 cm-1. Mixing poly(Asp) with poly(Lys) under identical conditions results in the formation of a fine precipitate with a different morphology. Examination of the ir spectrum of the precipitate revealed that unlike poly(Glu), poly(Asp) did not yield any discrete secondary structure upon precipitation with poly(Lys). Addition of solutions containing Ca2+ or Mg2+ to the poly(Glu)/poly(Lys) aggregates resulted in complete dissolution of the gel, with the disappearance of the ir bands characteristic of the intermolecular hydrogen-bonded network. The results demonstrate the importance of salt bridges in establishing strong hydrogen bonds between the backbone amide groups. Reaggregation occurred upon heating the poly(Glu)/poly(Lys) mixture in the presence of Ca2+, but not in the presence of Mg2+ ions. In the presence of Ca2+ ions, aggregation and formation of an extended hydrogen-bonded network occurred upon heating. The aggregates formed upon heating poly(Glu)/poly(Lys) in the presence of Ca2+ were attributed solely to complexation of Ca2+ to the carboxylate groups of poly(Glu) with poly(Lys) remaining free in solution. Dissolution of the aggregate could be accomplished through addition of Mg2+ at room temperature.  相似文献   

10.
The binding constants of N-acetylglucosamine (G1cNAc) and its methyl alpha- and beta- glycosides to hen and turkey egg-white lysozymes [EC 3.2.1.17], in the latter of which Asp 101 is replaced by Gly, were determined at various pH values by measuring changes in the circular dichroic (DC) band at 295 nm. The binding of beta-methyl-G1cNAc to turkey and hen lysozymes perturbed the pK value of Glu 35 from 6.0 to 6.5, the pK value of Asp 52 from 3.5 to 3.9, and the pK value of Asp 66 from 1.3 to 0.7. In addition, perturbation of the pK value of Asp 101 from 4.4 to 4.0 was observed in the binding of this saccharide to hen lysozyme. The binding of alpha-methyl-GlcNAc to hen and turkey lysozymes perturbed the pK value of Glu 35 to the alkaline side by about 0.5 pH unit, the pK value of Asp 66 to the acidic side by about 0.5 pH unit, and the pK value (4.4) of an ionizable group to the acidic side by about 0.6 pH unit. The last ionizable group was tentatively assigned to Asp 48. The pK value of Asp 52 was not perturbed by the binding of this saccharide. The pH dependence curves for the binding of GlcNAc to hen and turkey lysozymes were very similar and it was suggested that Asp 48, in addition to Asp 66, Asp 52, and Glu 35, is perturbed by the binding of GlcNAc.  相似文献   

11.
Asn112 is located at the active site of thermolysin, 5-8 A from the catalytic Zn2+ and catalytic residues Glu143 and His231. When Asn112 was replaced with Ala, Asp, Glu, Lys, His, and Arg by site-directed mutagenesis, the mutant enzymes N112D and N112E, in which Asn112 is replaced with Asp and Glu, respectively, were secreted as an active form into Escherichia coli culture medium, while the other four were not. In the hydrolysis of a neutral substrate N-[3-(2-furyl)acryloyl]-Gly-L-Leu amide, the kcat/Km values of N112D and N112E exhibited bell-shaped pH-dependence, as did the wild-type thermolysin (WT). The acidic pKa of N112D was 5.7 +/- 0.1, higher by 0.4 +/- 0.2 units than that of WT, suggesting that the introduced negative charge suppressed the protonation of Glu143 or Zn2+-OH. In the hydrolysis of a negatively charged substrate, N-carbobenzoxy-l-Asp-l-Phe methyl ester (ZDFM), the pH-dependence of kcat/Km of the mutants decreased with increase in pH from 5.5 to 8.5, while that of WT was bell-shaped. This difference might be explained by the electrostatic repulsion between the introduced Asp/Glu and ZDFM, suggesting that introducing ionizing residues into the active site of thermolysin might be an effective means of modifying its pH-activity profile.  相似文献   

12.
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   

13.
Lysozymes from family 22 of glycoside hydrolases are usually part of the defense system against bacteria. However in ruminant artiodactyls and saprophagous insects, lysozymes are involved in the digestion of bacteria. Here, we report the first crystallographic structure of a digestive lysozyme in its native and complexed forms, the structure of lysozyme 1 from Musca domestica larvae midgut (MdL1). Structural and biochemical data presented for MdL1 are analyzed in light of digestive lysozymes' traits. The structural core is similar, but a careful analysis of a structural alignment generated with other lysozymes c reveals that significant differences occur in coil regions. The loop from MdL1 defined by residues 98-100 has one deletion previous to residue Gln100, which leads to a less exposed conformation and might justify the resistance to proteolysis observed for MdL1. In addition, Gln100 is directly involved in a few hydrogen bonds to the ligand in a yet unobserved substrate binding mode. The pK(a)s of the MdL1 catalytic residues (Glu32 and Asp50) are lower (6.40 and 3.09, respectively) than those from Gallus gallus egg lysozyme (GgL, hen egg white lysozyme-HEWL) (6.61 and 3.85, respectively). A unique feature of MdL1 is a hydrogen bond between Thr107 Ogamma and Glu32 carboxylate group, which combined with the presence of Ser106 contributes to decrease the pK(a) of Glu32. Furthermore, in MdL1 the presence of Asn46 preventing the occurrence of an electrostatic repulsion with Asp50 and the increment in the solvent exposition of Asp50 due to Pro42 insertion contribute to reduce the pK(a) of Asp50. These structural elements affecting the pK(a)s of the catalytic residues should contribute to the acidic pH optimum presented by MdL1.  相似文献   

14.
The pH optima of family 11 xylanases are well correlated with the nature of the residue adjacent to the acid/base catalyst. In xylanases that function optimally under acidic conditions, this residue is aspartic acid, whereas it is asparagine in those that function under more alkaline conditions. Previous studies of wild-type (WT) Bacillus circulans xylanase (BCX), with an asparagine residue at position 35, demonstrated that its pH-dependent activity follows the ionization states of the nucleophile Glu78 (pKa 4.6) and the acid/base catalyst Glu172 (pKa 6.7). As predicted from sequence comparisons, substitution of this asparagine residue with an aspartic acid residue (N35D BCX) shifts its pH optimum from 5.7 to 4.6, with an approximately 20% increase in activity. The bell-shaped pH-activity profile of this mutant enzyme follows apparent pKa values of 3.5 and 5.8. Based on 13C-NMR titrations, the predominant pKa values of its active-site carboxyl groups are 3.7 (Asp35), 5.7 (Glu78) and 8.4 (Glu172). Thus, in contrast to the WT enzyme, the pH-activity profile of N35D BCX appears to be set by Asp35 and Glu78. Mutational, kinetic, and structural studies of N35D BCX, both in its native and covalently modified 2-fluoro-xylobiosyl glycosyl-enzyme intermediate states, reveal that the xylanase still follows a double-displacement mechanism with Glu78 serving as the nucleophile. We therefore propose that Asp35 and Glu172 function together as the general acid/base catalyst, and that N35D BCX exhibits a "reverse protonation" mechanism in which it is catalytically active when Asp35, with the lower pKa, is protonated, while Glu78, with the higher pKa, is deprotonated. This implies that the mutant enzyme must have an inherent catalytic efficiency at least 100-fold higher than that of the parental WT, because only approximately 1% of its population is in the correct ionization state for catalysis at its pH optimum. The increased efficiency of N35D BCX, and by inference all "acidic" family 11 xylanases, is attributed to the formation of a short (2.7 A) hydrogen bond between Asp35 and Glu172, observed in the crystal structure of the glycosyl-enzyme intermediate of this enzyme, that will substantially stabilize the transition state for glycosyl transfer. Such a mechanism may be much more commonly employed than is generally realized, necessitating careful analysis of the pH-dependence of enzymatic catalysis.  相似文献   

15.
The binding constants of alpha- and beta-GlcNAc to hen and turkey lysozymes [EC 3.2.1.17] were determined at various pH's using the method proposed by Ikeda and Hamaguchi (1975) J. Biochem. 77, 1-16). The pH dependence of the binding of beta-GlcNAc to hen lysozyme was essentially the same as that for turkey lysozyme. The pH dependence curves of the binding constants of beta-GlcNAc to hen and turkey lysozymes were interpreted in terms of the participation of Glu 35 (pK 6.0), Asp 52 (pK 3.5), Asp 48 (pK 4.5), and Asp 66 (pK 1.5). The binding constants of alpha-GlcNAc to hen and turkey lysozymes were the same below pH 3.5 but were different above this pH. The main participant residues in the binding of alpha-GlcNAc were Glu 35, Asp 48, and Asp 66 for hen lysozyme and Glu 35 and Asp 66 for turkey lysozyme. The results obtained here were well explained by the following assumptions: (1) above about pH 4, alpha-GlcNAc binds to hen lysozyme in both alpha- and beta-modes, which correspond to the binding orientation of alpha-GlcNAc and that of beta-GlcNAc, respectively, as determined by X-ray crystallographic studies, but it binds predominantly in the beta-mode below about pH 4, (2) beta-GlcNAc binds to hen and turkey lysozymes predominantly in the beta-mode above about pH 4 and in both alpha- and beta-modes below pH 4, and (3) alpha-GlcNAc binds to turkey lysozyme predominantly in the beta-mode over the whole pH range studied.  相似文献   

16.
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   

17.
Betz M  Löhr F  Wienk H  Rüterjans H 《Biochemistry》2004,43(19):5820-5831
Xylanase from Bacillus agaradhaerens belongs to a large group of glycosyl hydrolases which catalyze the degradation of xylan. The protonation behavior of titratable groups of the uniformly (15)N- and (13)C-labeled xylanase was investigated by multinuclear NMR spectroscopy. A total of 224 chemical shift titration curves corresponding to (1)H, (13)C, and (15)N resonances revealed pK(a) values for all aspartic and glutamic acid residues, as well as for the C-terminal carboxylate and histidine residues. Most of the titratable groups exhibit a complex titration behavior, which is most likely due to the mutual interactions with other neighboring groups or due to an unusual local microenvironment. Subsite -1 containing the catalytic dyad shows a long-range interaction over 9 A with Asp21 via two hydrogen bonds with Asn45 as the mediator. This result illuminates the pivotal role of the conserved position 45 among family 11 endoxylanases, determining an alkaline pH optimum by asparagine residues or an acidic pH optimum by an aspartate. The asymmetric interactions of neighboring tryptophan side chains with respect to the catalytic dyad can be comprehended as a result of hydrogen bonding and aromatic stacking. Most of the chemical shift-pH profiles of the backbone amides exhibit biphasic behavior with two distinct inflection points, which correspond to the pK(a) values of the nearby acidic side chains. However, the alternation of both positive and negative slopes of individual amide titration curves is interpreted as a consequence of a simultaneous reorganization of side chain conformational space at pH approximately 6 and/or an overall change in the hydrogen network in the substrate binding cleft.  相似文献   

18.
Isoaspartyl dipeptidase (IAD) is a member of the amidohydrolase superfamily and catalyzes the hydrolytic cleavage of beta-aspartyl dipeptides. Structural studies of the wild-type enzyme have demonstrated that the active site consists of a binuclear metal center positioned at the C-terminal end of a (beta/alpha)(8)-barrel domain. Steady-state kinetic parameters for the hydrolysis of beta-aspartyl dipeptides were obtained at pH 8.1. The pH-rate profiles for the hydrolysis of beta-Asp-Leu were obtained for the Zn/Zn-, Co/Co-, Ni/Ni-, and Cd/Cd-substituted forms of IAD. Bell-shaped profiles were observed for k(cat) and k(cat)/K(m) as a function of pH for all four metal-substituted forms. The pK(a) of the group that must be unprotonated for catalytic activity varied according to the specific metal ion bound in the active site, whereas the pK(a) of the group that must be protonated for catalytic activity was relatively independent of the specific metal ion present. The identity of the group that must be unprotonated for catalytic activity was consistent with the hydroxide that bridges the two divalent cations of the binuclear metal center. The identity of the group that must be protonated for activity was consistent with the free alpha-amino group of the dipeptide substrate. Kinetic constants were obtained for the mutant enzymes at conserved residues Glu77, Tyr137, Arg169, Arg233, Asp285, and Ser289. The catalytic properties of the wild-type and mutant enzymes, coupled with the X-ray crystal structure of the D285N mutant complexed with beta-Asp-His, are consistent with a chemical reaction mechanism for the hydrolysis of dipeptides that is initiated by the polarization of the amide bond via complexation to the beta-metal ion of the binuclear metal center. Nucleophilic attack by the bridging hydroxide is facilitated by abstraction of its proton by the side chain carboxylate of Asp285. Collapse of the tetrahedral intermediate and cleavage of the carbon-nitrogen bond occur with donation of a proton from the protonated form of Asp285.  相似文献   

19.
The difference spectra of hen and turkey egg-white lysozymes [EC 3.2.1.17] produced by acidification were measured. The difference spectra of both lysozymes had peaks at 295 and 301 nm which are characteristic of tryptophyl residues. The pH dependence curves of the extinction differences (delta eplision) at 301 nm and 295 nm for hen lysozyme were identical with the corresponding curves for turkey lysozyme. The pH dependence of delta eplision at 301 nm was analyzed assuming that the extinction at 301 nm is due to Trp 108 only, which interacts with the catalytic carboxyls, Glu 35 and Asp 52. The macroscopic pK values of Glu 35 and Asp 52 in both lysozymes thus determined were 6.0 and 3.3, respectively. These values were in excellent agreement with those determined by measuring the pH dependence of the circular dichroic band at 305 nm (Kuramitsu et al. (1974) J. Biochem, 76, 671-683; (1975) ibid. 77, 291-301). The pH dependence of delta eplision at 295 nm could not be completely explained in terms of the electrostatic effects of the catalytic groups on Trp 108.  相似文献   

20.
Koide A  Jordan MR  Horner SR  Batori V  Koide S 《Biochemistry》2001,40(34):10326-10333
It is generally considered that electrostatic interactions on the protein surface, such as ion pairs, contribute little to protein stability, although they may play important roles in conformational specificity. We found that the tenth fibronectin type III domain of human fibronectin (FNfn10) is more stable at acidic pH than neutral pH, with an apparent midpoint of transition near pH 4. Determination of pK(a)'s for all the side chain carboxyl groups of Asp and Glu residues revealed that Asp 23 and Glu 9 have an upshifted pK(a). These residues and Asp 7 form a negatively charged patch on the surface of FNfn10, with Asp 7 centrally located between Asp 23 and Glu 9, suggesting repulsive electrostatic interactions among these residues at neutral pH. Mutant proteins, D7N and D7K, in which Asp 7 was replaced with Asn and Lys, respectively, exhibited a modest but significant increase in stability at neutral pH, compared to the wild type, and they no longer showed pH dependence of stability. The pK(a)'s of Asp 23 and Glu 9 in these mutant proteins shifted closer to their respective unperturbed values, indicating that the unfavorable electrostatic interactions have been reduced in the mutant proteins. Interestingly, the wild-type and mutant proteins were all stabilized to a similar degree by the addition of 1 M sodium chloride at both neutral and acidic pH, suggesting that the repulsive interactions between the carboxyl groups cannot be effectively shielded by 1 M sodium chloride. These results indicate that repulsive interactions between like charges on the protein surface can destabilize a protein, and protein stability can be significantly improved by relieving these interactions.  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