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.
To further investigate the mechanism and function of allosteric activation by chloride in some alpha-amylases, the structure of the bacterial alpha-amylase from the psychrophilic micro-organism Pseudoalteromonas haloplanktis in complex with nitrate has been solved at 2.1 A degrees, as well as the structure of the mutants Lys300Gln (2.5 A degrees ) and Lys300Arg (2.25 A degrees ). Nitrate binds strongly to alpha-amylase but is a weak activator. Mutation of the critical chloride ligand Lys300 into Gln results in a chloride-independent enzyme, whereas the mutation into Arg mimics the binding site as is found in animal alpha-amylases with, however, a lower affinity for chloride. These structures reveal that the triangular conformation of the chloride ligands and the nearly equatorial coordination allow the perfect accommodation of planar trigonal monovalent anions such as NO3-, explaining their unusual strong binding. It is also shown that a localized negative charge such as that of Cl-, rather than a delocalized charge as in the case of nitrate, is essential for maximal activation. The chloride-free mutant Lys300Gln indicates that chloride is not mandatory for the catalytic mechanism but strongly increases the reactivity at the active site. Disappearance of the putative catalytic water molecule in this weakly active mutant supports the view that chloride helps to polarize the hydrolytic water molecule and enhances the rate of the second step in the catalytic reaction.  相似文献   

2.
Guijarro J  Engelhard M  Siebert F 《Biochemistry》2006,45(38):11578-11588
The uptake of chloride, bromide, iodide, nitrate, and azide by anion-depleted blue halorhodopsin from Natronobacterium pharaonis has been followed by FTIR difference spectroscopy using an ATR sampling device. The spectra are compared with the spectrum of the O intermediate obtained by time-resolved FTIR studies of the photocycle. It is demonstrated that anion-free blue halorhodopsin can be identified with the O intermediate and, thus, that the decay of O is due to the passive uptake of the anion. The great similarity of the anion-binding spectra and their identity in the case of the monoatomic anions indicate a rather unspecific binding site for the different anions dominated by electrostatic interactions. Comparing spectra obtained with 15N nitrate and unlabeled nitrate, the NO-stretching bands could be identified. The small splitting and the small IR intensity of those bands indicate a rather nonpolar binding site with a rather isotropic influence on the nitrate, in contrast to aqueous nitrate. In further experiments on the photocycle of blue halorhodopsin, the all-trans --> 13-cis isomerization can be clearly identified. Up to 100 micros, the isomerization-induced structural changes deduced from amide I changes are similar to those occurring during the anion-transporting photocycle. Compared to these, the molecular changes involved in the release and their reversion during the uptake of anions are considerably larger. They can be reached via two pathways: (1) by reducing the anion concentration and (2) transiently during the anion-transporting photocycle with the formation of the precursor of O with O conformation. Consequences of the anion transport mechanism are discussed.  相似文献   

3.
We report a multifaceted study of the active site region of human pancreatic alpha-amylase. Through a series of novel kinetic analyses using malto-oligosaccharides and malto-oligosaccharyl fluorides, an overall cleavage action pattern for this enzyme has been developed. The preferred binding/cleavage mode occurs when a maltose residue serves as the leaving group (aglycone sites +1 and +2) and there are three sugars in the glycon (-1, -2, -3) sites. Overall it appears that five binding subsites span the active site, although an additional glycon subsite appears to be a significant factor in the binding of longer substrates. Kinetic parameters for the cleavage of substrates modified at the 2 and 4' ' positions also highlight the importance of these hydroxyl groups for catalysis and identify the rate-determining step. Further kinetic and structural studies pinpoint Asp197 as being the likely nucleophile in catalysis, with substitution of this residue leading to an approximately 10(6)-fold drop in catalytic activity. Structural studies show that the original pseudo-tetrasaccharide structure of acarbose is modified upon binding, presumably through a series of hydrolysis and transglycosylation reactions. The end result is a pseudo-pentasaccharide moiety that spans the active site region with its N-linked "glycosidic" bond positioned at the normal site of cleavage. Interestingly, the side chains of Glu233 and Asp300, along with a water molecule, are aligned about the inhibitor N-linked glycosidic bond in a manner suggesting that these might act individually or collectively in the role of acid/base catalyst in the reaction mechanism. Indeed, kinetic analyses show that substitution of the side chains of either Glu233 or Asp300 leads to as much as a approximately 10(3)-fold decrease in catalytic activity. Structural analyses of the Asp300Asn variant of human pancreatic alpha-amylase and its complex with acarbose clearly demonstrate the importance of Asp300 to the mode of inhibitor binding.  相似文献   

4.
A novel kinetic approach was used to measure monovalent anion binding to better define the mechanistic basis for competition between stilbenedisulfonates and transportable anions on band 3. An anion-induced acceleration in the release of 4,4'-dibenzamidostilbene-2,2'-disulfonate (DBDS) from its complex with band 3 was measured using monovalent anions of various size and relative affinity for the transport site. The K1/2 values for anion binding were determined and correlated with transport site affinity constants obtained from the literature and the dehydrated radius of each anion. The results show that anions with ionic radii of 120-200 pm fall on a well-defined correlation line where the ranking of the K1/2 values matched the ranking of the transport site affinity constants (thiocyanate < nitrate approximately bromide < chloride < fluoride). The K1/2 values for the anions on this line were about 4-fold larger than expected for anion binding to inhibitor-free band 3. Such a lowered affinity can be explained in terms of allosteric site-site interactions, since the K1/2 values decreased with increasing anionic size. In contrast, iodide, with an ionic radius of about 212 pm, had a 10-fold lower affinity than predicted by the correlation line established by the smaller monovalent anions. These results indicate that smaller monovalent anions have unobstructed access to the transport site within the band 3 / DBDS binary complex, while iodide experiences significant steric hindrance when binding. The observation of steric hindrance in iodide binding to the band 3 / DBDS binary complex, but not in the binding of smaller monovalent anions, suggests that the stilbenedisulfonate binding site is located at the outer surface of an access channel leading to the transport site.  相似文献   

5.
The mechanism of allosteric activation of alpha-amylase by chloride has been studied through structural and kinetic experiments focusing on the chloride-dependent N298S variant of human pancreatic alpha-amylase (HPA) and a chloride-independent TAKA-amylase. Kinetic analysis of the HPA variant clearly demonstrates the pronounced activating effect of chloride ion binding on reaction rates and its effect on the pH-dependence of catalysis. Structural alterations observed in the N298S variant upon chloride ion binding suggest that the chloride ion plays a variety of roles that serve to promote catalysis. One of these is having a strong influence on the positioning of the acid/base catalyst residue E233. Absence of chloride ion results in multiple conformations for this residue and unexpected enzymatic products. Chloride ion and N298 also appear to stabilize a helical region of polypeptide chain from which projects the flexible substrate binding loop unique to chloride-dependent alpha-amylases. This structural feature also serves to properly orient the catalytically essential residue D300. Comparative analyses show that the chloride-independent alpha-amylases compensate for the absence of bound chloride by substituting a hydrophobic core, altering the manner in which substrate interactions are made and shifting the placement of N298. These evolutionary differences presumably arise in response to alternative operating environments or the advantage gained in a particular product profile. Attempts to engineer chloride-dependence into the chloride-independent TAKA-amylase point out the complexity of this system, and the fact that a multitude of factors play a role in binding chloride ion in the chloride-dependent alpha-amylases.  相似文献   

6.
Human pancreatic alpha-amylase (HPA) is a member of the alpha-amylase family involved in the degradation of starch. Some members of this family, including HPA, require chloride for maximal activity. To determine the mechanism of chloride activation, a series of mutants (R195A, R195Q, N298S, R337A, and R337Q) were made in which residues in the chloride ion binding site were replaced. Mutations in this binding site were found to severely affect the ability of HPA to bind chloride ions with no binding detected for the R195 and R337 mutant enzymes. X-ray crystallographic analysis revealed that these mutations did not result in significant structural changes. However, the introduction of these mutations did alter the kinetic properties of the enzyme. Mutations to residue R195 resulted in a 20-450-fold decrease in the activity of the enzyme toward starch and shifted the pH optimum to a more basic pH. Interestingly, replacement of R337 with a nonbasic amino acid resulted in an alpha-amylase that no longer required chloride for catalysis and has a pH profile similar to that of wild-type HPA. In contrast, a mutation at residue N298 resulted in an enzyme that had much lower binding affinity for chloride but still required chloride for maximal activity. We propose that the chloride is required to increase the pK(a) of the acid/base catalyst, E233, which would otherwise be lower due to the presence of R337, a positively charged residue.  相似文献   

7.
The first inhibition study of the beta-class carbonic anhydrase (CA, EC 4.2.1.1) from the methanoarchaeon Methanobacterium thermoautotrophicum (Cab) with anions is reported here. Inhibition data of the alpha-class human isozymes hCA I and hCA II (cytosolic) as well as the membrane-bound isozyme hCA IV and the gamma-class enzyme from another archaeon, Methanosarcina thermophila (Cam) with a large number of anionic species such as halides, pseudohalides, bicarbonate, carbonate, nitrate, nitrite, hydrosulfide, bisulfite, sulfate, etc., are also provided for comparison. The best Cab anion inhibitors were thiocyanate and hydrogen sulfide, with inhibition constants in the range of 0.52-0.70 mM, whereas cyanate, iodide, carbonate, and nitrate were weaker inhibitors (Ki's in the range of 7.8-13.2 mM). Fluoride, chloride, and sulfate do not inhibit this enzyme appreciably, whereas the CA substrate bicarbonate, or other anions, such as bromide, nitrite, bisulfite, or sulfamate behave as weak inhibitors (Ki in the range of 40-45 mM). It is interesting to note that the metal poison, coordinating anions cyanide and azide are also rather weak Cab inhibitors (Ki in the range of 27-55 mM), whereas sulfamide is a very weak Cab inhibitor (Ki of 103 mM), although it strongly inhibits Cam (Ki of 70 microM). Surprisingly, phenylboronic and phenylarsonic acids, which have been investigated for the inhibition of all these CAs for the first time, showed very weak activity against the alpha-CA isozymes, but were effective Cab and Cam inhibitors. The best Cab inhibitors were just these two compounds (Ki's of 0.20-0.33 mM), whereas the best Cam inhibitor was sulfamic acid (Ki of 96 nM). These major differences of behavior between the diverse CAs investigated here toward anion inhibitors can be difficultly explained considering the convergent evolution of so diverse enzymes for the binding and turnover of small molecules such as carbon dioxide and anions.  相似文献   

8.
E R Wilcox  J R Whitaker 《Biochemistry》1984,23(8):1783-1791
Bovine pancreatic alpha-amylase binds 1 mol of acarbose (a carbohydrate alpha-amylase inhibitor) per mol at the active site and also binds acarbose nonspecifically. The red kidney bean alpha-amylase inhibitor-bovine pancreatic alpha-amylase complex retained nonspecific binding for acarbose only. Binding of p-nitrophenyl alpha-D-maltoside to the final complex of red kidney bean alpha-amylase inhibitor and bovine pancreatic alpha-amylase has a beta Ks (Ks') value that is 3.4-fold greater than the Ks (16 mM) of alpha-amylase for p-nitrophenyl alpha-D-maltoside alone. The initial complex of alpha-amylase and inhibitor apparently hydrolyzes this substrate as rapidly as alpha-amylase alone. The complex retains affinity for substrates and competitive inhibitors, which, when present in high concentrations, cause dissociation of the complex. Maltose (0.5 M), a competitive inhibitor of alpha-amylase, caused dissociation of the red kidney bean alpha-amylase inhibitor--alpha-amylase complex. Interaction between red kidney bean (Phaseolus vulgaris) alpha-amylase inhibitor and porcine pancreatic alpha-amylase proceeds through two steps. The first step has a Keq of 3.1 X 10(-5) M. The second step (unimolecular; first order) has a forward rate constant of 3.05 min-1 at pH 6.9 and 30 degrees C. alpha-Amylase inhibitor combines with alpha-amylase, in the presence of p-nitrophenyl alpha-D-maltoside, noncompetitively. On the basis of the data presented, it is likely that alpha-amylase is inactivated by the alpha-amylase inhibitor through a conformational change. A kinetic model, in the presence and absence of substrate, is described for noncompetitive, slow, tight-binding inhibitors that proceed through two steps.  相似文献   

9.
A new microcalorimetric method for recording the kinetic parameters k(cat), K(m) and K(i) of alpha-amylases using polysaccharides and oligosaccharides as substrates is described. This method is based on the heat released by glycosidic bond hydrolysis. The method has been developed to study the active site properties of the cold-active alpha-amylase produced by an Antarctic psychrophilic bacterium in comparison with its closest structural homolog from pig pancreas. It is shown that the psychrophilic alpha-amylase is more active on large macromolecular substrates and that the higher rate constants k(cat) are gained at the expense of a lower affinity for the substrate. The active site is able to accommodate larger inhibitory complexes, resulting in a mixed-type inhibition of starch hydrolysis by maltose. A method for recording the binding enthalpies by isothermal titration calorimetry in a low-affinity system has been developed, allowing analysis of the energetics of weak ligand binding using the allosteric activator chloride. It is shown that the low affinity of the psychrophilic alpha-amylase for chloride is entropically driven. The high enthalpic and entropic contributions of activator binding suggest large structural fluctuations between the free and the bound states of the cold-active enzyme. The kinetic and thermodynamic data for the psychrophilic alpha-amylase indicate that the strictly conserved side-chains involved in substrate binding and catalysis possess an improved mobility, responsible for activity in the cold, and resulting from the disappearance of stabilizing interactions far from the active site.  相似文献   

10.
Qian M  Ajandouz el H  Payan F  Nahoum V 《Biochemistry》2005,44(9):3194-3201
Pig pancreatic alpha-amylase (PPA), an enzyme belonging to the alpha-amylase family, is involved in the degradation of starch. Like some other members of this family, PPA requires chloride to reach maximum activity levels. To further explain the mechanism of chloride activation, a crystal of wild-type PPA soaked with maltopentaose using a chloride-free buffer was analyzed by X-ray crystallography. A conspicuous reorientation of the acid/base catalyst Glu233 residue was found to occur. The structural results, along with kinetic data, show that the acid/base catalyst is maintained in the active site, in an optimum position, pointing toward the scissile bond-atom, due to the presence of chloride ions. The present study therefore explains the mechanism of PPA activation by chloride ions.  相似文献   

11.
Li C  Begum A  Numao S  Park KH  Withers SG  Brayer GD 《Biochemistry》2005,44(9):3347-3357
A mechanistic study of the poorly understood pathway by which the inhibitor acarbose is enzymatically rearranged by human pancreatic alpha-amylase has been conducted by structurally examining the binding modes of the related inhibitors isoacarbose and acarviosine-glucose, and by novel kinetic measurements of all three inhibitors under conditions that demonstrate this rearrangement process. Unlike acarbose, isoacarbose has a unique terminal alpha-(1-6) linkage to glucose and is found to be resistant to enzymatic rearrangement. This terminal glucose unit is found to bind in the +3 subsite and for the first time reveals the interactions that occur in this part of the active site cleft with certainty. These results also suggest that the +3 binding subsite may be sufficiently flexible to bind the alpha-(1-6) branch points in polysaccharide substrates, and therefore may play a role in allowing efficient cleavage in the direct vicinity of such junctures. Also found to be resistant to enzymatic rearrangement was acarviosine-glucose, which has one fewer glucose unit than acarbose. Collectively, structural studies of all three inhibitors and the specific cleavage pattern of HPA make it possible to outline the simplest sequence of enzymatic reactions likely involved upon acarbose binding. Prominent features incorporated into the starting structure of acarbose to facilitate the synthesis of the final tightly bound pseudo-pentasaccharide product are the restricted availability of hydrolyzable bonds and the placement of the transition state-like acarviosine group. Additional "in situ" experiments designed to elongate and thereby optimize isoacarbose and acarviosine-glucose inhibition using the activated substrate alphaG3F demonstrate the feasibility of this approach and that the principles outlined for acarbose rearrangement can be used to predict the final products that were obtained.  相似文献   

12.
An interaction between cytochrome a in oxidized cytochrome c oxidase (CcO) and anions has been characterized by EPR spectroscopy. Those anions that affect the EPR g = 3 signal of cytochrome a can be divided into two groups. One group consists of halides (Cl-, Br-, and I-) and induces an upfield shift of the g = 3 signal. Nitrogen-containing anions (CN-, NO2-, N3-, NO3-) are in the second group and shift the g = 3 signal downfield. The shifts in the EPR spectrum of CcO are unrelated to ligand binding to the binuclear center. The binding properties of one representative from each group, azide and chloride, were characterized in detail. The dependence of the shift on chloride concentration is consistent with a single binding site in the isolated oxidized enzyme with a Kd of approximately 3 mm. In mitochondria, the apparent Kd was found to be about four times larger than that of the isolated enzyme. The data indicate it is the chloride anion that is bound to CcO, and there is a hydrophilic size-selective access channel to this site from the cytosolic side of the mitochondrial membrane. An observed competition between azide and chloride is interpreted by azide binding to three sites: two that are apparent in the x-ray structure plus the chloride-binding site. It is suggested that either Mg2+ or Arg-438/Arg-439 is the chloride-binding site, and a mechanism for the ligand-induced shift of the g = 3 signal is proposed.  相似文献   

13.
The inhibition of the newly discovered cytosolic carbonic anhydrase (CA, EC 4.2.1.1) isozyme XIII of murine origin (mCA XIII) has been investigated with a series of anions, such as the physiological ones (bicarbonate, chloride), or the metal complexing anions (cyanate, cyanide, azide, hydrogen sulfide, etc), nitrate, nitrite, sulfate, sulfamate, sulfamide as well as with phenylboronic and phenylarsonic acids. The best mCA XIII inhibitors were cyanate, thiocyanate, cyanide and sulfamide, with K(I)-s in the range of 0.25microM-0.74 mM, whereas fluoride, iodide, azide, carbonate and hydrogen sulfide were less effective (K(I)-s in the range of 3.0-5.5mM). The least effective inhibitors were sulfate, chloride and bicarbonate (K(I)-s in the range of 138-267 mM). The affinity of mCA XIII for anions is very different from that of the other cytosolic isozymes (hCA I and II) or the mitochondrial isozyme hCA V. This resistance to inhibition by the physiological anions bicarbonate and chloride suggests an evolutionary adaptation of CA XIII to the presence of high concentrations of such anions (e.g., in the reproductive tract of both female and male), and the possible participation of this isozyme (similarly to CA II, CA IV and CA V) in metabolons with proteins involved in the anion exchange and transport, such as the anion exchangers (AE1-3) or the sodium bicarbonate co-transporter (NBC1 and NBC3) proteins, which remain to be identified.  相似文献   

14.
A single basic residue above the si-face of the flavin ring is the site of oxygen activation in glucose oxidase (GOX) (His516) and monomeric sarcosine oxidase (MSOX) (Lys265). Crystal structures of both flavoenzymes exhibit a small pocket at the oxygen activation site that might provide a preorganized binding site for superoxide anion, an obligatory intermediate in the two-electron reduction of oxygen. Chloride binds at these polar oxygen activation sites, as judged by solution and structural studies. First, chloride forms spectrally detectable complexes with GOX and MSOX. The protonated form of His516 is required for tight binding of chloride to oxidized GOX and for rapid reaction of reduced GOX with oxygen. Formation of a binary MSOX·chloride complex requires Lys265 and is not observed with Lys265Met. Binding of chloride to MSOX does not affect the binding of a sarcosine analogue (MTA, methylthioactetate) above the re-face of the flavin ring. Definitive evidence is provided by crystal structures determined for a binary MSOX·chloride complex and a ternary MSOX·chloride·MTA complex. Chloride binds in the small pocket at a position otherwise occupied by a water molecule and forms hydrogen bonds to four ligands that are arranged in approximate tetrahedral geometry: Lys265:NZ, Arg49:NH1, and two water molecules, one of which is hydrogen bonded to FAD:N5. The results show that chloride (i) acts as an oxygen surrogate, (ii) is an effective probe of polar oxygen activation sites, and (iii) provides a valuable complementary tool to the xenon gas method that is used to map nonpolar oxygen-binding cavities.  相似文献   

15.
Porcine pancreatic alpha-amylase (1,4-alpha-D-glucan glucanohydrolase) [EC 3.2.1.1] has both amylase activity (hydrolysis of alpha-1,4-D-glucoside bond of starch) and maltosidase activity (hydrolysis of p-nitrophenyl-alpha-D-maltoside to p-nitrophenol and maltose). By the modification of histidine residues of porcine pancreatic alpha-amylase with diethylpyrocarbonate (DEP), both amylase and maltosidase activities were decreased in the absence of chloride ion. In the presence of chloride ion, however, maltosidase activity of the modified enzyme was increased to more than 260% of that of the native enzyme, whereas amylase activity was decreased to less than 15% of the native enzyme. Since the chloride ion binding site is part of the active site loop [Buisson et al. (1987) Food Hydrocolloids 1,399-406 and Buisson et al. (1987) EMBO J. 6, 3909-3916], the special arrangements of both catalytic and modified histidine residues induced by the chloride ion binding would enhance only the maltosidase activity of the histidine-modified enzyme.  相似文献   

16.
Insulin-degrading enzyme (IDE) (insulysin) is a zinc metallopeptidase that metabolizes several bioactive peptides, including insulin and the amyloid β peptide. IDE is an unusual metallopeptidase in that it is allosterically activated by both small peptides and anions, such as ATP. Here, we report that the ATP-binding site is located on a portion of the substrate binding chamber wall arising largely from domain 4 of the four-domain IDE. Two variants having residues in this site mutated, IDEK898A,K899A,S901A and IDER429S, both show greatly decreased activation by the polyphosphate anions ATP and PPPi. IDEK898A,K899A,S901A is also deficient in activation by small peptides, suggesting a possible mechanistic link between the two types of allosteric activation. Sodium chloride at high concentrations can also activate IDE. There are no observable differences in average conformation between the IDE-ATP complex and unliganded IDE, but regions of the active site and C-terminal domain do show increased crystallographic thermal factors in the complex, suggesting an effect on dynamics. Activation by ATP is shown to be independent of the ATP hydrolysis activity reported for the enzyme. We also report that IDEK898A,K899A,S901A has reduced intracellular function relative to unmodified IDE, consistent with a possible role for anion activation of IDE activity in vivo. Together, the data suggest a model in which the binding of anions activates by reducing the electrostatic attraction between the two halves of the enzyme, shifting the partitioning between open and closed conformations of IDE toward the open form.  相似文献   

17.
Catalase isolated from thermophilic bacterium Thermus thermophilus (Mn-catalase) is composed of six subunits, each containing binuclear manganese clusters at their active site. The enzyme is active when the metal is in the completely reduced (Mn2+---Mn2+) state in the pH range 7–10, and loses activity on oxidation. ESR data suggest that the metal turns thereby into the (Mn2+---Mn4+) state. The Mn-catalase activity is inhibited by chloride, nitrate, nitrite, azide, and other singly charged anions, except cyanide. The inhibitory effect of anions increases as the pH value is reduced. The inhibition by hydroxylamine takes place through a lag-phase and is weakly dependent on pH value. The reaction mechanism is discussed in relation with current concepts of catalase reactions of heme-proteins and low-molecular binuclear manganese complexes.  相似文献   

18.
The structure of human pancreatic alpha-amylase has been determined to 1.8 A resolution using X-ray diffraction techniques. This enzyme is found to be composed of three structural domains. The largest is Domain A (residues 1-99, 169-404), which forms a central eight-stranded parallel beta-barrel, to one end of which are located the active site residues Asp 197, Glu 233, and Asp 300. Also found in this vicinity is a bound chloride ion that forms ligand interactions to Arg 195, Asn 298, and Arg 337. Domain B is the smallest (residues 100-168) and serves to form a calcium binding site against the wall of the beta-barrel of Domain A. Protein groups making ligand interactions to this calcium include Asn 100, Arg 158, Asp 167, and His 201. Domain C (residues 405-496) is made up of anti-parallel beta-structure and is only loosely associated with Domains A and B. It is notable that the N-terminal glutamine residue of human pancreatic alpha-amylase undergoes a posttranslational modification to form a stable pyrrolidone derivative that may provide protection against other digestive enzymes. Structure-based comparisons of human pancreatic alpha-amylase with functionally related enzymes serve to emphasize three points. Firstly, despite this approach facilitating primary sequence alignments with respect to the numerous insertions and deletions present, overall there is only approximately 15% sequence homology between the mammalian and fungal alpha-amylases. Secondly, in contrast, these same studies indicate that significant structural homology is present and of the order of approximately 70%. Thirdly, the positioning of Domain C can vary considerably between alpha-amylases. In terms of the more closely related porcine enzyme, there are four regions of polypeptide chain (residues 237-250, 304-310, 346-354, and 458-461) with significantly different conformations from those in human pancreatic alpha-amylase. At least two of these could play a role in observed differential substrate and cleavage pattern specificities between these enzymes. Similarly, amino acid differences between human pancreatic and salivary alpha-amylases have been localized and a number of these occur in the vicinity of the active site.  相似文献   

19.
D R Meloon  R G Wilkins 《Biochemistry》1976,15(6):1284-1290
The kinetics of anation of methemerythrin over a wide range of pH and concentration of anions have been studied at 25 degrees C. The azide and thiocyanate ions have been most intensively investigated but experiments with fluoride and chloride are also reported. The replacement of anion in methemerythrin-anionic adducts by other anions has also been studied. Except for replacement of met-fluoride by azide, all replacements can be explained by a dissociative mechanism via the aquated species. Anations are second-order and an associative mechanism is preferred. The second-order rate constant decreases with increasing anion concentrations (from 20 muM to 20 mM). This is attributed to the effect of a secondary anion binding site. The behavior of octameric and monomeric forms of the protein toward thiocyanate is identical. A comparison of results with simple Fe(III) complexes and certain metalloproteins is made.  相似文献   

20.
The anion azide, N3 -, has been previously found to be an inhibitor of oxygen evolution by Photosystem II (PS II) of higher plants. With respect to chloride activation, azide acts primarily as a competitive inhibitor but uncompetitive inhibition also occurs [Haddy A, Hatchell JA, Kimel RA and Thomas R (1999) Biochemistry 38: 6104–6110]. In this study, the effects of azide on PS II-enriched thylakoid membranes were characterized by electron paramagnetic resonance (EPR) spectroscopy. Azide showed two distinguishable effects on the S2 state EPR signals. In the presence of chloride, which prevented competitive binding, azide suppressed the formation of the multiline and g = 4.1 signals concurrently, indicating that the normal S2 state was not reached. Signal suppression showed an azide concentration dependence that correlated with the fraction of PS II centers calculated to bind azide at the uncompetitive site, based on the previously determined inhibition constant. No evidence was found for an effect of azide on the Fe(II)QA - signals at the concentrations used. This result is consistent with placement of the uncompetitive site on the donor side of PS II as suggested in the previous study. In chloride-depleted PS II-enriched membranes azide and fluoride showed similar effects on the S2 state EPR signals, including a notable increase and narrowing of the g = 4.1 signal. Comparable effects of other anions have been described previously and apparently take place through the chloride-competitive site. The two azide binding sites described here correlate with the results of other studies of Lewis base inhibitors.This revised version was published online in October 2005 with corrections to the Cover Date.  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