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 15 毫秒
1.
We report the X-ray crystal structures and rate constants for proton transfer in site-specific mutants of human carbonic anhydrase III (HCA III) that place a histidine residue in the active-site cavity: K64H, R67H, and K64H-R67N HCA III. Prior evidence from the exchange of 18O between CO2 and water measured by mass spectrometry shows each mutant to have enhanced proton transfer in catalysis compared with wild-type HCA III. However, His64 in K64H and K64H-R67N HCA III have at most a capacity for proton transfer that is only 13% that of His64 in HCA II. This reduced rate in mutants of HCA III is associated with a constrained side-chain conformation of His64, which is oriented outward, away from the active-site zinc in the crystal structures. This conformation appears stabilized by a prominent pi stacking interaction of the imidazole ring of His64 with the indole ring of Trp5 in mutants of HCA III. This single orientation of His64 in K64H HCA III predominates also in a double mutant K64H-R67N HCA III, indicating that the positive charge of Arg67 does not influence the observed conformation of His64 in the crystal structure. Hence, the structures and catalytic activity of these mutants of HCA III containing His64 account only in small part for the lower activity of this isozyme compared with HCA II. His67 in R67H HCA III was also shown to be a proton shuttle residue, having a capacity for proton transfer that was approximately four times that of His64 in K64H HCA III. This is most likely due to its proximity and orientation inward towards the zinc-bound solvent. These results emphasize the significance of side chain orientation and range of available conformational states as characteristics of an efficient proton shuttle in carbonic anhydrase.  相似文献   

2.
The undisputed role of His64 in proton transfer during catalysis by carbonic anhydrases in the α class has raised questions concerning the details of its mechanism. The highly conserved residues Tyr7, Asn62, and Asn67 in the active-site cavity function to fine tune the properties of proton transfer by human carbonic anhydrase II (HCA II). For example, hydrophobic residues at these positions favor an inward orientation of His64 and a low pKa for its imidazole side chain. It appears that the predominant manner in which this fine tuning is achieved in rate constants for proton transfer is through the difference in pKa between His64 and the zinc-bound solvent molecule. Other properties of the active-site cavity, such as inward and outward conformers of His64, appear associated with the change in ΔpKa; however, there is no strong evidence to date that the inward and outward orientations of His64 are in themselves requirements for facile proton transfer in carbonic anhydrase.  相似文献   

3.
The binding of bovine oxyhemoglobin to bovine carbonic anhydrase with a dissociation constant between 10(-5) and 10(-7) M has been determined by countercurrent distribution using aqueous, biphasic polymer systems. This result provides an explanation for the very efficient proton transfer between hemoglobin and carbonic anhydrase, a transfer which enhances the catalytic activity of carbonic anhydrase as measured by 18O exchange between bicarbonate and water at chemical equilibrium (Silverman, D. N., Tu, C. K., and Wynns, G. C. (1978) J. Biol. Chem, 253, 2563-2567). Two rate constants describing 18O exchange activity of carbonic anhydrase at pH 7.5 show saturation behavior when plotted against hemoglobin concentration consistent with a dissociation constant of 2.5 X 10(-6) M between bovine hemoglobin and carbonic anhydrase. Interpretation of these rate constants in terms of a two-step model for 18O exchange indicates that hemoglobin enhances the rate of exchange from carbonic anhydrase of water containing the oxygen abstracted from bicarbonate, but does not affect the catalytic interconversion of CO2 and HCO3- at chemical equilibrium.  相似文献   

4.
The rate limiting step in catalysis of bicarbonate dehydration by human carbonic anhydrase II (HCA II) is an intramolecular proton transfer from His64 to the zinc-bound hydroxide. We have examined the role of Tyr7 using site-specific mutagenesis and measuring catalysis by the 18O exchange method using membrane inlet mass spectrometry. The side chain of Tyr7 in HCA II extends into the active-site cavity about 7 Å from the catalytic zinc atom. Replacement of Tyr7 with eight other amino acids had no effect on the interconversion of bicarbonate and CO2, but in some cases caused enhancements in the rate constant of proton transfer by nearly 10-fold. The variant Y7I HCA II enhanced intramolecular proton transfer approximately twofold; its structure was determined by X-ray crystallography at 1.5 Å resolution. No changes were observed in the ordered solvent structure in the active-site cavity or in the conformation of the side chain of the proton shuttle His64. However, the first 11 residues of the amino-terminal chain in Y7I HCA II assumed an alternate conformation compared with the wild type. Differential scanning calorimetry showed variants at position 7 had a melting temperature approximately 8 °C lower than that of the wild type.  相似文献   

5.
6.
7.
Cobalt(III)bovine carbonic anhydrase B was prepared by the oxidation of the cobalt(II) enzyme with hydrogen peroxide and was purified by affinity chromatography. The oxidation reaction is inhibited by specific inhibitors of carbonic anhydrase. The inhibition is explained by the fact that the Co(II)-enzyme . inhibitor complex cannot be directly oxidized by hydrogen peroxide, but has to dissociate to give free Co(II) enzyme which is then oxidized. The Co(III) ion in Co(III) carbonic anhydrase cannot be directly substituted by zinc ions. It can be reduced by either dithionite or BH-4 ions to give, first, their complexes with the Co(II) enzyme, and upon their removal, a fully active Co(II) enzyme. Cyanide and azide bind to cobalt(III) carbonic anhydrase with similar rate constants of 0.060 +/- 0.005 and 0.070 +/- 0.007 M-1 S-1 respectively. These rates are faster than those found for Co(III) inorganic complexes. The Co(III) ion in both Co(III) carbonic anhydrase and Co(III) carboxypeptidase A was found to be diamagnetic, indicating a near octahedral symmetry.  相似文献   

8.
The visualization at near atomic resolution of transient substrates in the active site of enzymes is fundamental to fully understanding their mechanism of action. Here we show the application of using CO(2)-pressurized, cryo-cooled crystals to capture the first step of CO(2) hydration catalyzed by the zinc-metalloenzyme human carbonic anhydrase II, the binding of substrate CO(2), for both the holo and the apo (without zinc) enzyme to 1.1A resolution. Until now, the feasibility of such a study was thought to be technically too challenging because of the low solubility of CO(2) and the fast turnover to bicarbonate by the enzyme (Liang, J. Y., and Lipscomb, W. N. (1990) Proc. Natl. Acad. Sci. U. S. A. 87, 3675-3679). These structures provide insight into the long hypothesized binding of CO(2) in a hydrophobic pocket at the active site and demonstrate that the zinc does not play a critical role in the binding or orientation of CO(2). This method may also have a much broader implication for the study of other enzymes for which CO(2) is a substrate or product and for the capturing of transient substrates and revealing hydrophobic pockets in proteins.  相似文献   

9.
Investigation of some pH-dependent properties of human erythrocyte carbonic anhydrase B indicate that the active site is influenced by at least two charged groups. The properties studied include the pH dependence of inhibition of native, monocarboxamidomethyl, and monocarboxymethyl enzymes by iodide ion and the pH dependence of the visible spectra of the cobalt derivatives of these enzymes. One ionizing group has a pKa of about 7.3 in the native enzyme, 8.2 in the carboxyamidomethyl enzyme, and 9.0 in the carboxymethyl enzyme. It has a major influence on activity and anion inhibition, and on the visible spectra of the cobalt enzymes. A second group has a pKa of about 6.1 in native and modified enzymes. When zinc is at the active site, the secondary group in its acidic form decreases the Ki for I-. With the carboxyamidomethyl and carboxymethyl enzymes, the Ki decreases by about an order of magnitude. However, if cobalt is substituted for zinc in the modified enzymes, this group does not influence the Ki for I- and the binding of I- does not influence the pKa of the spectral transitions caused by ionization of this secondary group. In the case of nonalkylated Co2+-enzyme, another ionizing group with a pK of about 6.2 prevents the binging of I- at low pH. These results show that the active site is altered when cobalt is substituted for zinc in carbonic anhydrase B.  相似文献   

10.
X-ray absorption spectroscopy at the Zn K-edge indicates that the active site of the marine diatom Thalassiosira weissflogii carbonic anhydrase is strikingly similar to that of mammalian alpha-carbonic anhydrase enzymes. The zinc has three histidine ligands and a single water at 1.98 A. This is quite different from the beta-carbonic anhydrases of higher plants in which zinc is coordinated by two cysteine thiolates, one histidine, and a water molecule. The diatom carbonic anhydrase shows no significant sequence similarity with other carbonic anhydrases and may represent an example of convergent evolution at the molecular level.  相似文献   

11.
12.
13.
The cobalt(II)-substituted proinsulin and insulin hexamers have been studied in solution via electronic absorption spectroscopy. Hexameric proinsulin is shown to undergo the phenol-induced T6 to R6 conformational transition in a manner analogous to that previously established for insulin. In the absence of coordinating anions, the coordination spheres of the Co(II) ions in the proinsulin and insulin R6 hexamers comprise identical pseudotetrahedral arrangements of 3 histidine residues and 1 hydroxide ion. At alkaline pH, the visible absorption spectrum of the phenol-induced R6 Co(II) center is strikingly similar to the distinctive spectrum of the alkaline form of Co(II)-carbonic anhydrase. Exogenous ligands may coordinate to the Co(II) ions of the R6 proinsulin and insulin hexamers via replacement of the hydroxide ion, forming pseudotetrahedral adducts possessing characteristic spectra. The binding affinity of such ligands is shown to be strongly pH-dependent. The data presented establish that, although the Co(II)-substituted proinsulin and insulin R6 hexamers lack enzyme-like activity, these species duplicate spectrochemical characteristics of the Co(II)-carbonic anhydrase active site that are believed to be important signatures of carbonic anhydrase catalytic function.  相似文献   

14.
Carbonic anhydrase (CA) exists in three forms: the low-pH form (L); the high-pH form (H); and the anion-inhibited from (A). The latter includes the bicarbonate complex. All three forms have been demonstrated in CA I and, when sulfate is removed, in CA II. The L-form of CA III has not yet been seen, even at pH 5. Equilibrium among the three forms in a sample of CA can be established, in principle, by kinetic pathways connecting any two forms; which pathway dominates is as yet an open question. By invoking the usual ping-pong mechanism of CA, during which hydration of CO2 causes the enzyme to go from H to L, the kinetic pathway connecting A and H is ignored, essentially by definition. Rarely has the AH pathway been considered (cf. Koenig et al., 1980). Though there are few data to demonstrate the relative kinetics of the AL and AH pathways, it can be argued that the latter is buffer-mediated, which could distinguish the two. In this case, the lifetime of a bound anion would be buffer-dependent. We have investigated this point by measuring the nuclear relaxation rates of fluorine of trifluoroacetate in Co2+CA II solutions. The fluorine linewidth, and thus the anion exchange rate, is independent of buffer concentration up to ~50 mM, which argues for the AL pathway predominating.  相似文献   

15.
Between pH 7 and pH 10.5, p-nitrophenyl p-sulfamyl benzoate (PNP-SAB) binds very strongly to human carbonic anhydrase B (dissociation constant on the order of 10?9 M or less at pH 7.5), but is not hydrolyzed by the enzyme. Because the binding is essentially stoichiometric under readily accessible conditions, this ester may be used as an active site titrant, by measuring the rapid hydrolysis of excess unbound PNP-SAB catalyzed by an added nucleophile (“reverse burst”).  相似文献   

16.
Among the isozymes of carbonic anhydrase, isozyme III is the least efficient in the catalysis of the hydration of CO2 and was previously thought to be unaffected by proton transfer from buffers to the active site. We report that buffers of small size, especially imidazole, increase the rate of catalysis by human carbonic anhydrase III (HCA III) of (1) 18O exchange between HCO3- and water measured by membrane-inlet mass spectrometry and (2) the dehydration of HCO3- measured by stopped-flow spectrophotometry. Imidazole enhanced the rate of release of 18O-labeled water from the active site of wild-type carbonic anhydrase III and caused a much greater enhancement, up to 20-fold, for the K64H, R67H, and R67N mutants of this isozyme. Imidazole had no effect on the rate of interconversion of CO2 and HCO3- at chemical equilibrium. Steady-state measurements showed that the addition of imidazole resulted in increases in the turnover number (kcat) for the hydration of CO2 catalyzed by HCA III and for the dehydration of HCO3- catalyzed by R67N HCA III. These results are consistent with the transfer of a proton from the imidazolium cation to the zinc-bound hydroxide at the active site, a step required to regenerate the active form of enzyme in the catalytic cycle. Like isozyme II of carbonic anhydrase, isozyme III can be enhanced in catalytic rate by the presence of small molecule buffers in solution.  相似文献   

17.
Due to prevalence of negative charges on the protein surface, opposite to the active site pocket of human carbonic anhydrase XII (hCA XII), both positively charged CdTe quantum dots (Qds+) and polylysine electrostatically interact with the enzyme, and such interaction does not influence the catalytic activity of the enzyme. However, both these cationic macromolecules differently modulate the active site environment of the enzyme. The steady-state kinetic data revealed that whereas polylysine exhibited no influence on dansylamide (DNSA) dependent inhibition of the enzyme, Qds+ overcame such an inhibitory effect, leading to almost 70% restoration of the catalytic activity of the enzyme. We provide evidence that DNSA remains bound to the enzyme upon interaction with both polylysine and Qds+. Arguments are presented that the above differential feature of polylysine and Qds+ on hCA XII is encoded in the “rigidity” versus “flexibility” of these cationic macromolecules.  相似文献   

18.
The initial water proteolysis step in the proton transfer "half-reaction" of human carbonic anhydrase I is simulated using the empirical valence bond method in combination with free energy perturbation molecular dynamics calculations. A free energy profile for the enzyme catalysed reaction and the corresponding pKa associated with ionization of the zinc-bound water is calculated. The obtained pKa value of 7 to 8 appears to be in good agreement with experimental observations and the calculated rate constant for this step is also compatible with kinetic data. The simulations clearly emphasize the important electrostatic effect associated with the catalytic zinc ion.  相似文献   

19.
We investigated a series of derivatized fullerenes possessing alcohol, amine, and amino acid pendant groups as inhibitors of the zinc enzymes carbonic anhydrases (CAs, EC 4.2.1.1). We discovered that fullerenes bind CAs with submicromolar—low micromolar affinity, despite the fact that these compounds do not possess moieties normally associated with CA inhibitors such as the sulfonamides and their isosteres, or the coumarins. The 13 different mammalian CA isoforms showed a diverse inhibition profile with these compounds. By means of computational methods we assessed the inhibition mechanism as being due to occlusion of the active site entrance by means of the fullerene cage (possessing dimension of the same order of magnitude as the opening of the enzyme cavity, of 1 nm). The pendant moieties to the fullerene cage make interactions with amino acid residues from the active site, among which His64, His94, His96, Val121, and Thr200. Fullerenes thus represent a totally new class of nanoscale CA inhibitors which may show applications for targeting physiologically relevant isoforms, such as the dominant CA II and the tumor-associated CA IX.  相似文献   

20.
We report here a theoretical study on the formation of long-range proton transfer pathways in proteins due to side chain conformational fluctuations of amino acid residues and reorganization of interior hydration positions. The proton transfer pathways in such systems may be modeled as fluctuating hydrogen-bonded networks with both short- and long-lived connections between the networked nodes, the latter being formed by polar protein atoms and water molecules. It is known that these fluctuations may extend over several decades of time ranging from a few femtoseconds to a few milliseconds. We have shown in this article how the use of a variety of theoretical methods may be utilized to detect a generic set of pathways and assess the feasibility of forming one or more transient connections. We demonstrate the application of these methods to the enzyme human carbonic anhydrase II and its mutants. Our results reveal several alternative pathways in addition to the one mediated by His-64. We also probe at length the mechanism of key conformational fluctuations contributing to the formation of the detected pathways.  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