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.
Acid–base reactions that are exceedingly unfavorable under standard conditions can be catalytically important at enzyme active sites. For example, in triose phosphate isomerase, a glutamate side chain (nominal pK a ≈ 4 in solution) can in fact deprotonate a CH group that is vicinal to a carbonyl (pK a ≈ 18 in solution). This is true because of three distinct interactions: (a) ground state pK a shifts due to environment polarity and electrostatics; (b) dramatic increases in effective molarity due to optimization of proximity and orientation; and (c) transition state pK a shifts due to binding interactions and the formation of strong low barrier hydrogen bonds. In this report, we review the literature showing that the sum of these three effects supplies more than enough free energy to push forward proton transfer reactions that under standard conditions are exceedingly nonspontaneous and slow.  相似文献   

2.
3.
The side chains of Lys66, Asp66, and Glu66 in staphylococcal nuclease are fully buried and surrounded mainly by hydrophobic matter, except for internal water molecules associated with carboxylic oxygen atoms. These ionizable side chains titrate with pKa values of 5.7, 8.8, and 8.9, respectively. To reproduce these pKa values with continuum electrostatics calculations, we treated the protein with high dielectric constants. We have examined the structural origins of these high apparent dielectric constants by using NMR spectroscopy to characterize the structural response to the ionization of these internal side chains. Substitution of Val66 with Lys66 and Asp66 led to increased conformational fluctuations of the microenvironments surrounding these groups, even under pH conditions where Lys66 and Asp66 are neutral. When Lys66, Asp66, and Glu66 are charged, the proteins remain almost fully folded, but resonances for a few backbone amides adjacent to the internal ionizable residues are broadened. This suggests that the ionization of the internal groups promotes a local increase in dynamics on the intermediate timescale, consistent with either partial unfolding or increased backbone fluctuations of helix 1 near residue 66, or, less likely, with increased fluctuations of the charged side chains at position 66. These experiments confirm that the high apparent dielectric constants reported by internal Lys66, Asp66, and Glu66 reflect localized changes in conformational fluctuations without incurring detectable global structural reorganization. To improve structure-based pKa calculations in proteins, we will need to learn how to treat this coupling between ionization of internal groups and local changes in conformational fluctuations explicitly.  相似文献   

4.
Two glutamic acid-rich fusion peptide analogs of influenza hemagglutinin were synthesized to study the organization of the charged peptides in the membranous media. Fluorescence and gel electrophoresis experiments suggested a loose association between the monomers in the vesicles. A model was built which showed that a positional difference of 3, 7 and 4, 8 results in the exposure of Glu3 and Glu7 side chains to the apolar lipidic core. Supportive results include: first, pKa values of two pH units higher than reference value in aqueous medium for Glu3 and Glu7 CγH, whereas the deviation of pKa from the reference value for Glu4 and Glu8 CγH is substantially smaller; second, Hill coefficients of titration shift of these protons indicate anti-cooperativity for Glu3 and Glu7 side chain protons but less so for Glu4 and Glu8, implying a strong electrostatic interaction between Glu3 and Glu7 possibly resulting from their localization in an apolar environment; third, positive and larger titration shift for NH of Glu3 is observed compared to that of Glu4, suggesting stronger hydrogen bond between the NH and the carboxylic group of Glu3 than that of Glu4, consistent with higher degree of exposure to hydrophobic medium for the side chain of Glu3.  相似文献   

5.
Many proteins fibrillate at low pH despite a high population of charged side chains. Therefore exchange of protons between the fibrillating peptide and its surroundings may play an important role in fibrillation. Here, we use isothermal titration calorimetry to measure exchange of protons between buffer and the peptide hormone glucagon during fibrillation. Glucagon absorbs or releases protons to an extent which allows it to attain a net charge of zero in the fibrillar state, both at acidic and basic pH. Similar results are obtained for lysozyme. This suggests that side chain pKa values change dramatically in the fibrillar state.  相似文献   

6.
Ion pairs are ubiquitous in X-ray structures of coiled coils, and mutagenesis of charged residues can result in large stability losses. By contrast, pKa values determined by NMR in solution often predict only small contributions to stability from charge interactions. To help reconcile these results we used triple-resonance NMR to determine pKa values for all groups that ionize between pH 1 and 13 in the 33 residue leucine zipper fragment, GCN4p. In addition to the native state we also determined comprehensive pKa values for two models of the GCN4p denatured state: the protein in 6 M urea, and unfolded peptide fragments of the protein in water. Only residues that form ion pairs in multiple X-ray structures of GCN4p gave large pKa differences between the native and denatured states. Moreover, electrostatic contributions to stability were not equivalent for oppositely charged partners in ion pairs, suggesting that the interactions between a charge and its environment are as important as those within the ion pair. The pH dependence of protein stability calculated from NMR-derived pKa values agreed with the stability profile measured from equilibrium urea-unfolding experiments as a function of pH. The stability profile was also reproduced with structure-based continuum electrostatic calculations, although contributions to stability were overestimated at the extremes of pH. We consider potential sources of errors in the calculations, and how pKa predictions could be improved. Our results show that although hydrophobic packing and hydrogen bonding have dominant roles, electrostatic interactions also make significant contributions to the stability of the coiled coil.  相似文献   

7.
Improvement of enzyme function by engineering pH dependence of enzymatic activity is of importance for industrial application of Bacillus circulans xylanases. Target mutation sites were selected by structural alignment between B. circulans xylanase and other xylanases having different pH optima. We selected non-conserved mutant sites within 8 Å from the catalytic residues, to see whether these residues have some role in modulating pKas of the catalytic residues. We hypothesized that the non-conserved residues which may not have any role in enzyme catalysis might perturb pKas of the catalytic residues. Change in pKa of a titratable group due to change in electrostatic potential of a mutation was calculated and the change in pH optimum was predicted from the change in pKa of the catalytic residues. Our strategy is proved to be useful in selection of promising mutants to shift the pH optimum of the xylanases towards desired side.  相似文献   

8.
Hiroshi Ishikita 《FEBS letters》2010,584(15):3464-3468
The pKa value of Lys115, the catalytic residue in acetoacetate decarboxylate, was calculated using atomic coordinates of the X-ray crystal structure with consideration of the protonation states of all titratable sites in the protein. The calculated pKa value of Lys115 (pKa(Lys115)) was unusually low (≈6) in agreement with the experimentally measured value. Although charged residues impact pKa(Lys115) considerably in the native protein, the significant pKa(Lys115) downshift in the protein with respect to aqueous solution was mainly due to loss of the solvation energy in the catalytic active site relative to bulk water.  相似文献   

9.
The visible absorption of crystals of Co(II)-substituted human carbonic anhydrase II (Co(II)-HCA II) were measured over a pH range of 6.0-11.0 giving an estimate of pKa 8.4 for the ionization of the metal-bound water in the crystal. This is higher by about 1.2 pKa units than the pKa near 7.2 for Co(II)-CA II in solution. This effect is attributed to a nonspecific ionic strength effect of 1.4 M citrate in the precipitant solution used in the crystal growth. A pKa of 8.3 for the aqueous ligand of the cobalt was measured for Co(II)-HCA II in solution containing 0.8 M citrate. Citrate is not an inhibitor of the catalytic activity of Co(II)-HCA II and was not observed in crystal structures. The X-ray structures at 1.5-1.6 Å resolution of Co(II)-HCA II were determined for crystals prepared at pH 6.0, 8.5 and 11.0 and revealed no conformational changes of amino-acid side chains as a result of the use of citrate. However, the studies of Co(II)-HCA II did reveal a change in metal coordination from tetrahedral at pH 11 to a coordination consistent with a mixed population of both tetrahedral and penta-coordinate at pH 8.5 to an octahedral geometry characteristic of the oxidized enzyme Co(III)-HCA II at pH 6.0.  相似文献   

10.
The high-resolution structure of bovine trypsin inhibited with DFP2 was determined by Stroud et al. (1971 and R. M. Stroud, L. M. Kay, A. Cooper &; R. E. Dickerson, Abstr. 8th Int. Congr. Biochem. 1970). The experiments reported here were designed to study the specific side-chain binding pocket of trypsin using benzamidine, which is a competitive, specific inhibitor of trypsin. High-resolution electron density syntheses and difference syntheses unambiguously identify the side-chain binding pocket, which normally recognizes and binds the side chains of arginine or lysine during proteolysis. Several important conformational differences in the protein structure are apparent between DIP- and BA-trypsins, and these are discussed with particular reference to inhibition, the binding of lysine and arginine, subsequent orientation of the target at the active site, and the enhancement of tryptic activity towards non-specific substrates seen on binding small alkyl amines or guanidines in the specific binding pocket.The BA-trypsin structure provides a good model for the binding of real substrate side chains to trypsin during catalysis, explaining the sharp trypsin specificity for lysine or arginine side chains (Weinstein &; Doolittle, 1972) and the lack of specificity for stereochemically different basic side chains. Benzamidine is shown to inhibit trypsin by steric interference with the inferred position of good substrates, even when they do not carry any side chain.Apart from the substitution of benzamidine and DIP, the most significant differences between DIP-trypsin and BA-trypsin involve complete repositioning of the side chain of Gln192, alterations in the side chains of Asp102, His57 and Ser195 at the active site, and changes in the solvent structure around this region. The carboxyl group of Asp189, which is responsible for trypsin specificity, shows no movement on binding benzamidine. The amidinium cation of benzamidine forms a salt bridge with Asp189 in BA-trypsin; a similar salt bridge can be constructed between the side chains of model substrates with lysyl or arginyl side chains and Aspl89. The γ-oxygen of Ser190 is displaced by a 120 ° rotation about its αβ bond on binding benzamidine and the binding pocket closes to sandwich the inhibitor ring between the peptide planes of 190–191 and 215–216. These contacts are presumably found in the enzyme-substrate complex with specific substrates.The active site structure at pH 8.0 is discussed with particular reference to the microscopic pKa values of Asp102 and His57, the pKa of the Asp-His system, and the mechanistic consequences of these assignments.  相似文献   

11.
Protein disulfide isomerase (PDI) plays a central role in disulfide bond formation in the endoplasmic reticulum. It is implicated both in disulfide bond formation and in disulfide bond reduction and isomerization. To be an efficient catalyst of all three reactions requires complex mechanisms. These include mechanisms to modulate the pKa values of the active-site cysteines of PDI. Here, we examined the role of arginine 120 in modulating the pKa values of these cysteines. We find that arginine 120 plays a significant role in modulating the pKa of the C-terminal active-site cysteine in the a domain of PDI and plays a role in determining the reactivity of the N-terminal active-site cysteine but not via direct modulation of its pKa. Mutation of arginine 120 and the corresponding residue, arginine 461, in the a′ domain severely reduces the ability of PDI to catalyze disulfide bond formation and reduction but enhances the ability to catalyze disulfide bond isomerization due to the formation of more stable PDI-substrate mixed disulfides. These results suggest that the modulation of pKa of the C-terminal active cysteine by the movement of the side chain of these arginine residues into the active-site locales has evolved to allow PDI to efficiently catalyze both oxidation and isomerization reactions.  相似文献   

12.
Multi-conformation continuum electrostatics (MCCE) was used to analyze various structures of the NS3 RNA helicase from the hepatitis C virus in order to determine the ionization state of amino acid side chains and their pKas. In MCCE analyses of HCV helicase structures that lacked ligands, several active site residues were identified to have perturbed pKas in both the nucleic acid binding site and in the distant ATP-binding site, which regulates helicase movement. In all HCV helicase structures, Glu493 was unusually basic and His369 was abnormally acidic. Both these residues are part of the HCV helicase nucleic acid binding site, and their roles were analyzed by examining the pH profiles of site-directed mutants. Data support the accuracy of MCCE predicted pKa values, and reveal that Glu493 is critical for low pH enzyme activation. Several key residues, which were previously shown to be involved in helicase-catalyzed ATP hydrolysis, were also identified to have perturbed pKas including Lys210 in the Walker-A motif and the DExD/H-box motif residues Asp290 and His293. When DNA was present in the structure, the calculated pKas shifted for both Lys210 and Asp290, demonstrating how DNA binding might lead to electrostatic changes that stimulate ATP hydrolysis.  相似文献   

13.
1. The absorption spectra of deutero- and proto-ferrihaem in aqueous solution at 25°C show marked changes with concentration and pH in the Soret band region. Quantitative studies of these phenomena imply that they are associated with ferrihaem dimerization and with protolytic equilibria involving monomeric (M) and dimeric (D) ferrihaem species according to the scheme: [Formula: see text] 2. For deuteroferrihaem we obtain K=1.9×10−2, pKa(M)=7.1, pKa(D)=7.4. Protoferrihaem has a much higher dimerization constant, K=4.5 and pKa(D)=7.5 (pKa(M) is not accessible). 3. Possible structural relationships between monomeric and dimeric ferrihaem species in solution are discussed in relation to recent work on the oxo-bridged nature of crystalline ferrihaem dimers.  相似文献   

14.
Using guanidinium and n-butylammonium cations (C+) as models for the positively charged side chains in arginine and lysine, we have determined the association constants with various oxyanions by potentiometric titration. For a dibasic acid, H2A, three association complexes may exist: K1M = [CHA][C+] [HA?]; K1D = [CA?][C+] [A2?]; K2D = [C2A][C+] [CA?]. For guanidinium ion and phosphate, K1M = 1.4, K1D = 2.6, and K2D = 5.1. The data for carboxylates indicate that the basicity of the oxyanion does not affect the association constant: acetate, pKa = 4.8, K1M = 0.37; formate, pKa = 3.8, K1M = 0.32; and chloroacetate, pKa = 2.9, K1M = 0.43, all with guanidinium ion. Association constants are also reported for carbonate, dimethylphosphinate, benzylphosphonate, and adenylate anions.  相似文献   

15.
A series of 4-anilinoquinoline derivatives related to the known inhibitor SGI-1027, containing side chains of varying pKa, were prepared by acid-catalysed coupling of the pre-formed side chains with 4-chloroquinolines. The compounds were evaluated for their ability to reduce the level of DNMT1 protein in HCT116 human colon carcinoma cells by Western blotting. With a very strongly basic N-methylpyridinium side chain, only NHCO-linked compounds were effective, whereas less strongly basic ((diaminomethylene)hydrazono)ethyl or 3-methylpyrimidine-2,4-diamine side chains allowed both NHCO- and CONH-linked compounds to show activity. In contrast, the pKa of the quinoline unit had little apparent influence on activity.  相似文献   

16.
Siderophore A (SidA) from Aspergillus fumigatus is a flavin‐containing monooxygenase that hydroxylates ornithine (Orn) at the amino group of the side chain. Lysine (Lys) also binds to the active site of SidA; however, hydroxylation is not efficient and H2O2 is the main product. The effect of pH on steady‐state kinetic parameters was measured and the results were consistent with Orn binding with the side chain amino group in the neutral form. From the pH dependence on flavin oxidation in the absence of Orn, a pKa value >9 was determined and assigned to the FAD‐N5 atom. In the presence of Orn, the pH dependence displayed a pKa value of 6.7 ±0.1 and of 7.70 ±0.10 in the presence of Lys. Q102 interacts with NADPH and, upon mutation to alanine, leads to destabilization of the C4a‐hydroperoxyflavin (FADOOH). Flavin oxidation with Q102A showed a pKa value of ~8.0. The data are consistent with the pKa of the FAD N5‐atom being modulated to a value >9 in the absence of Orn, which aids in the stabilization of FADOOH. Changes in the FAD‐N5 environment lead to a decrease in the pKa value, which facilitates elimination of H2O2 or H2O. These findings are supported by solvent kinetic isotope effect experiments, which show that proton transfer from the FAD N5‐atom is rate limiting in the absence of a substrate, however, is significantly less rate limiting in the presence of Orn and or Lys.  相似文献   

17.
The depolarization ratios of heme protein Raman lines arising from vibrations of the heme group exhibit significant dependence on the excitation wavelength. From the analysis of this depolarization ratio dispersion, one obtains information about symmetry-lowering distortions δQΓ of the heme group that can be classified in terms of the symmetry races Γ = A1g, B1g, B2g, and A2g in D4h symmetry. The heme-protein interaction can be changed by the protonation of distinct amino acid side chains (i.e., for instance the Bohr groups in hemoglobin derivates), which gives rise to specific static heme distortions for each protonation state. From the Raman dispersion data, it is possible to obtain parameters by fitting to a theoretical expression of the Raman tensor, which provide information on these static distortions and also about the pK values of the involved titrable side chains. We have applied this method to the ν4 (1,355 cm-1) and ν10 (1,620 cm-1) lines of deoxygenated hemoglobin of the fourth component of trout and have measured their depolarization ratio dispersion as a function of pH between 6 and 9. From the pH dependence of the thus derived parameters, we obtain pK values identical to those of the Bohr groups, which were earlier derived from the corresponding O2-binding isotherms. These are pKα1 = pKα2 = 8.5 for the α and pKβ1 = 7.5, pKβ2 = 7.4 for the β chains. We also obtain the specific distortion parameters for each protonation state. As shown in earlier studies, the ν4 mode mainly probes distortions from interactions between the proximal histidine and atoms of the heme core (i.e., the nitrogens and the Cα atoms of the pyrroles). Group theoretical argumentation allows us to relate specific changes of the imidazole geometry as determined by its tilt and azimuthal angle and the iron-out-of-plane displacement to distinct variations of the normal distortions δQΓ derived from the Raman dispersion data. Thus, we found that the pH dependence of the heme distortions δQA1g (totally symmetric) and δQB1g (asymmetric) is caused by variations of the azimuthal rather than the tilt angle of the Fe-His (F8) bond. In contrast to this, the ν10 line mainly monitors changes resulting from the interaction between peripheral substituents of the porphyrin macrocycle (vinyl). From the pH dependence of the parameters, it is possible to separately identify distortions δQΓ affecting the hemes in the α and β chains, respectively. From this, we find that in the α subunit structural changes induced on protonation of the corresponding Bohr groups are mainly transferred via the Fe—Nε bond and give rise to changes in the azimuthal angle. In the β subunit, however, in addition, structural changes of the heme pocket arise, which most probably result from protonation of the imidazole of the COOH-terminal His (HC3 β). This rearranges the net of H bonds between His HC3 β, Ser (F9 β), and Glu (F7 β).  相似文献   

18.
Several small ribozymes employ general acid–base catalysis as a mechanism to enhance site-specific RNA cleavage, even though the functional groups on the ribonucleoside building blocks of RNA have pKa values far removed from physiological pH. The rate of the cleavage reaction is strongly affected by the identity of the metal cation present in the reaction solution; however, the mechanism(s) by which different cations contribute to rate enhancement has not been determined. Using the Neurospora VS ribozyme, we provide evidence that different cations confer particular shifts in the apparent pKa values of the catalytic nucleobases, which in turn determines the fraction of RNA in the protonation state competent for general acid–base catalysis at a given pH, which determines the observed rate of the cleavage reaction. Despite large differences in observed rates of cleavage in different cations, mathematical models of general acid–base catalysis indicate that k1, the intrinsic rate of the bond-breaking step, is essentially constant irrespective of the identity of the cation(s) in the reaction solution. Thus, in contrast to models that invoke unique roles for metal ions in ribozyme chemical mechanisms, we find that most, and possibly all, of the ion-specific rate enhancement in the VS ribozyme can be explained solely by the effect of the ions on nucleobase pKa. The inference that k1 is essentially constant suggests a resolution of the problem of kinetic ambiguity in favor of a model in which the lower pKa is that of the general acid and the higher pKa is that of the general base.  相似文献   

19.
The signaling state of the photoreceptor photoactive yellow protein is the long-lived intermediate I2′. The pH dependence of the equilibrium between the transient photocycle intermediates I2 and I2′ was investigated. The formation of I2′ from I2 is accompanied by a major conformational change. The kinetics and intermediates of the photocycle and of the photoreversal were measured by transient absorption spectroscopy from pH 4.6 to 8.4. Singular value decomposition (SVD) analysis of the data at pH 7 showed the presence of three spectrally distinguishable species: I1, I2, and I2′. Their spectra were determined using the extrapolated difference method. I2 and I2′ have electronic absorption spectra, with maxima at 370 ± 5 and 350 ± 5 nm, respectively. Formation of the signaling state is thus associated with a change in the environment of the protonated chromophore. The time courses of the I1, I2, and I2′ intermediates were determined from the wavelength-dependent transient absorbance changes at each pH, assuming that their spectra are pH-independent. After the formation of I2′ (~2 ms), these three intermediates are in equilibrium and decay together to the initial dark state. The equilibrium between I2 and I2′ is pH dependent with a pKa of 6.4 and with I2′ the main species above this pKa. Measurements of the pH dependence of the photoreversal kinetics with a second flash of 355 nm at a delay of 20 ms confirm this pKa value. I2 and I2′ are photoreversed with reversal times of ~55 μs and several hundred microseconds, respectively. The corresponding signal amplitudes are pH dependent with a pKa of ~6.1. Photoreversal from I2′ dominates above the pKa. The transient accumulation of I2′, the active state of photoactive yellow protein, is thus controlled by the proton concentration. The rate constant k3 for the recovery to the initial dark state also has a pKa of ~6.3. This equality of the equilibrium and kinetic pKa values is not accidental and suggests that k3 is proportional to [I2′].  相似文献   

20.
Dioxygenases catalyze a diverse range of chemical reactions that involve the incorporation of oxygen into a substrate and typically use a transition metal or organic cofactor for reaction. Bacterial (1H)-3-hydroxy-4-oxoquinaldine 2,4-dioxygenase (HOD) belongs to a class of oxygenases able to catalyze this energetically unfavorable reaction without any cofactor. In the quinaldine metabolic pathway, HOD breaks down its natural N-heteroaromatic substrate using a mechanism that is still incompletely understood. Experimental and computational approaches were combined to study the initial step of the catalytic cycle. We have investigated the role of the active site His-251/Asp-126 dyad, proposed to be involved in substrate hydroxyl group deprotonation, a critical requirement for subsequent oxygen reaction. The pH profiles obtained under steady-state conditions for the H251A and D126A variants show a strong pH effect on their kcat and kcat/Km constants, with a decrease in kcat/Km of 5500- and 9-fold at pH 10.5, respectively. Substrate deprotonation studies under transient-state conditions show that this step is not rate-limiting and yield a pKa value of ∼7.2 for WT HOD. A large solvent isotope effect was found, and the pKa value was shifted to ∼8.3 in D2O. Crystallographic and computational studies reveal that the mutations have a minor effect on substrate positioning. Computational work shows that both His-251 and Asp-126 are essential for the proton transfer driving force of the initial reaction. This multidisciplinary study offers unambiguous support to the view that substrate deprotonation, driven by the His/Asp dyad, is an essential requirement for its activation.  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