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.

Background

Vibrio carchariae chitinase A (EC3.2.1.14) is a family-18 glycosyl hydrolase and comprises three distinct structural domains: i) the amino terminal chitin binding domain (ChBD); ii) the (α/β)8 TIM barrel catalytic domain (CatD); and iii) the α + β insertion domain. The predicted tertiary structure of V. carchariae chitinase A has located the residues Ser33 & Trp70 at the end of ChBD and Trp231 & Tyr245 at the exterior of the catalytic cleft. These residues are surface-exposed and presumably play an important role in chitin hydrolysis.

Results

Point mutations of the target residues of V. carchariae chitinase A were generated by site-directed mutagenesis. With respect to their binding activity towards crystalline α-chitin and colloidal chitin, chitin binding assays demonstrated a considerable decrease for mutants W70A and Y245W, and a notable increase for S33W and W231A. When the specific hydrolyzing activity was determined, mutant W231A displayed reduced hydrolytic activity, whilst Y245W showed enhanced activity. This suggested that an alteration in the hydrolytic activity was not correlated with a change in the ability of the enzyme to bind to chitin polymer. A mutation of Trp70 to Ala caused the most severe loss in both the binding and hydrolytic activities, which suggested that it is essential for crystalline chitin binding and hydrolysis. Mutations varied neither the specific hydrolyzing activity against pNP-[GlcNAc]2, nor the catalytic efficiency against chitohexaose, implying that the mutated residues are not important in oligosaccharide hydrolysis.

Conclusion

Our data provide direct evidence that the binding as well as hydrolytic activities of V. carchariae chitinase A to insoluble chitin are greatly influenced by Trp70 and less influenced by Ser33. Though Trp231 and Tyr245 are involved in chitin hydrolysis, they do not play a major role in the binding process of crystalline chitin and the guidance of the chitin chain into the substrate binding cleft of the enzyme.  相似文献   

2.
The lactam 16 , the hydroximolactams 8 , 20 , 23 , and 27 , and the imidazole 32 were prepared following known methods. They were tested together with the known tetrazole 35 and the hydroximolactams 2 and 36 as inhibitors of the cellobiohydrolases Cel7A and Cel6A from Trichoderma reesei. Cel7A is only weakly inhibited by these compounds. Comparing their inhibitory activity evidences the importance of occupying subsites +1 and +2. The results strongly suggest that the shape of none of the variants of the lactone-type inhibitor motif embodied by these inhibitors is complementary to the subsite −1, i. e., analogous to the transition state. Cel6A is rather strongly inhibited by the cellobiose analogues 20 , 23 , and 32 , and by the cellotriose analogue 27 . Their relative inhibitory activities evidence that binding at subsite −2 depends upon the shape of the moiety occupying subsite −1. There is only a small difference between the inhibition by the hydroximolactams 20 and 23 , which may be (partially) protonated by the catalytic acid of either anti- or syn-protonating glycosidases, and the imidazole 32 , which can only be protonated by anti-protonating glycosidases. The results strongly suggest that shape requirements must be met by glycosidase inhibitors before they can be used to characterize the proton trajectory of glycosidases.  相似文献   

3.
The binding properties of O2 and CO to recombinant human serum albumin (rHSA) mutants with a prosthetic heme group have been physicochemically and kinetically characterized. Iron(III) protoporphyrin IX (hemin) is bound in subdomain IB of wild-type rHSA [rHSA(wt)] with weak axial coordination by Tyr-161. The reduced ferrous rHSA(wt)-heme under an Ar atmosphere exists in an unusual mixture of four- and five-coordinate complexes and is immediately autoxidized by O2. To confer O2 binding capability on this naturally occurring hemoprotein, a proximal histidine was introduced into position Ile-142 or Leu-185 by site-directed mutagenesis. A single mutant (I142H) and three double mutants (I142H/Y161L, I142H/Y161F, and Y161L/L185H) were prepared. Both rHSA(I142H/Y161L)-heme and rHSA(I142H/Y161F)-heme formed ferrous five-N-coordinate high-spin complexes with axial ligation of His-142 under an Ar atmosphere. These artificial hemoproteins bind O2 at room temperature. Mutation at the other side of the porphyrin, Y161L/L185H, also allowed O2 binding to the heme. In contrast, the single mutant rHSA(I142H)-heme could not bind O2, suggesting that removal of Y161 is necessary to confer reversible O2 binding. Laser flash photolysis experiments showed that the kinetics of CO recombination with the rHSA(mutant)-heme were biphasic, whereas O2 rebinding exhibited monophasic kinetics. This could be due to the two different geometries of the axial imidazole coordination arising from the two orientations of the porphyrin plane in the heme pocket. The O2 binding affinities of the rHSA(mutant)-heme were significantly lower than those of hemoglobin and myoglobin, principally due to the high O2 dissociation rates. Changing Leu-161 to Phe-161 at the distal side increased the association rates of both O2 and CO, which resulted in enhanced binding affinity.  相似文献   

4.
The synthesis of four novel 3′‐C‐branched and 4′‐C‐branched nucleosides and their transformation into the corresponding 3′‐O‐phosphoramidite building blocks for automated oligonucleotide synthesis is reported. The 4′‐C‐branched key intermediate 11 was synthesized by a convergent strategy and converted to its 2′‐O‐methyl and 2′‐deoxy‐2′‐fluoro derivatives, leading to the preparation of novel oligonucleotide analogues containing 4′‐C‐(aminomethyl)‐2′‐O‐methyl monomer X and 4′‐C‐(aminomethyl)‐2′‐deoxy‐2′‐fluoro monomer Y (Schemes 2 and 3). In general, increased binding affinity towards complementary single‐stranded DNA and RNA was obtained with these analogues compared to the unmodified references (Table 1). The presence of monomer X or monomer Y in a 2′‐O‐methyl‐RNA oligonucleotide had a negative effect on the binding affinity of the 2′‐O‐methyl‐RNA oligonucleotide towards DNA and RNA. Starting from the 3′‐C‐allyl derivative 28 , 3′‐C‐(3‐aminopropyl)‐protected nucleosides and 3′‐O‐phosphoramidite derivatives were synthesized, leading to novel oligonucleotide analogues containing 3′‐C‐(3‐aminopropyl)thymidine monomer Z or the corresponding 3′‐C‐(3‐aminopropyl)‐2′‐O,5‐dimethyluridine monomer W (Schemes 4 and 5). Incorporation of the 2′‐deoxy monomer Z induced no significant changes in the binding affinity towards DNA but decreased binding affinity towards RNA, while the 2′‐O‐methyl monomer Z induced decreased binding affinity towards DNA as well as RNA complements (Table 2).  相似文献   

5.
Lysozyme is a well‐studied enzyme that hydrolyzes the β‐(1,4)‐glycosidic linkage of N‐acetyl‐β‐glucosamine (NAG)n oligomers. The active site of hen egg‐white lysozyme (HEWL) is believed to consist of six subsites, A‐F that can accommodate six sugar residues. We present studies exploring the use of polarizable force fields in conjunction with all‐atom molecular dynamics (MD) simulations to analyze binding structures of complexes of lysozyme and NAG trisaccharide, (NAG)3. MD trajectories are applied to analyze structures and conformation of the complex as well as protein–ligand interactions, including the hydrogen‐bonding network in the binding pocket. Two binding modes (ABC and BCD) of (NAG)3 are investigated independently based on a fixed‐charge model and a polarizable model. We also apply molecular mechanics with generalized born and surface area (MM‐GBSA) methods based on MD using both nonpolarizable and polarizable force fields to compute binding free energies. We also study the correlation between root‐mean‐squared deviation and binding free energies of the wildtype and W62Y mutant; we find that for this prototypical system, approaches using the MD trajectories coupled with implicit solvent models are equivalent for polarizable and fixed‐charge models. © 2012 Wiley Periodicals, Inc.  相似文献   

6.
The interaction of CrII with taurine/alpha-ketoglutarate (alphaKG) dioxygenase (TauD) was examined. CrII replaces FeII and binds stoichiometrically with alphaKG to the FeII/alphaKG binding site of the protein, with additional CrII used to generate a chromophore attributed to a CrIII-semiquinone in a small percentage of the sample. Formation of the latter oxygen-sensitive species requires the dihydroxyphenylalanine (DOPA) quinone form of Tyr-73. This preformed side chain is generated by intracellular self-hydroxylation of Tyr-73 to form DOPA, which is subsequently oxidized to the quinone. No chromophore is generated when using NaBH4-treated sample, protein isolated from anaerobically grown cells, inactive TauD variants that are incapable of self-hydroxylation, or the Y73F active mutant of TauD. A CrIII-DOPA semiquinone also was observed in the herbicide hydroxylase SdpA.  相似文献   

7.
Ginsenosides are major bioactive constituents of ginseng (Panax spp.; Araliaceae), a traditional Chinese medicinal herb. In order to increase the molecular diversity and broaden the potential usage of ginsenosides, ginsenosides Rd ( 1 ), Rg3 ( 2 ), (20R)‐Rg3 ( 3 ), Rh2 ( 4 ), Re ( 5 ), Rh1 ( 8 ), Rg2 ( 9 ), gypenoside XVII ( 6 ), and pseudoginsenoside F11 ( 7 ) were regioselectively acylated with vinyl acetate, catalyzed by Novozyme 435 (lipase B from Candida antarctica), in organic solvents to afford different mono‐acetyl ginsenosides. Ginsenoside Rd ( 1 ) was also acylated with vinyl decanoate or vinyl cinnamate to generate 1b and 1c , respectively. Acylation of glucosylated ginsenosides ( 1 – 4, 6, 8 ) occurred at the primary 6‐OH function of the terminal glucose (Glc) moiety of the sugar at C(3) or C(20) of the dammarane‐type aglycone. In contrast, ginsenosides 5, 7 , and 9 , containing mixed sugar moieties, resulted in acylation of both the rhamnose (Rha) and the glucose (Glc) moieties. In the case of ginsenoside Re ( 5 ) and pseudoginsenoside F11 ( 7 ), acylation at the secondary 4‐OH function of the terminal Rha moiety, attached at C(3) of the aglycone, is preferred. The structures of all acylated products were determined by extensive MALDI‐TOF‐MS and NMR analyses.  相似文献   

8.
ABSTRACT

Mutans streptococcus glucosyltransferases are the significant virulent factors in causing dental caries. The binding specificity of acarbose was probed with glucosyl and fructosyl sub-site binding ligands using multiple inhibition kinetics. The results indicate that acarbose and a glucosyl subsite binding ligand (1-deoxynojirimycin) are mutually or partially exclusive. On the other hand, acarbose with a fructosyl subsite ligand (fructose) might induce a conformational change leading to enhanced binding at the adjacent subsite.  相似文献   

9.
Chitinase was purified from the culture medium of Bacillus licheniformis SK-1 by colloidal chitin affinity adsorption followed by diethylamino ethanol-cellulose column chromatography. The purified enzyme showed a single band on sodium dodecyl sulfate polyacrylamide gel electrophoresis. The molecular size and pI of chitinase 72 (Chi72) were 72 kDa and 4.62 (Chi72) kDa, respectively. The purified chitinase revealed two activity optima at pH 6 and 8 when colloidal chitin was used as substrate. The enzyme exhibited activity in broad temperature range, from 40 to 70°C, with optimum at 55°C. It was stable for 2 h at temperatures below 60°C and stable over a broad pH range of 4.0–9.0 for 24 h. The apparent K m and V max of Chi72 for colloidal chitin were 0.23 mg ml−1 and 7.03 U/mg, respectively. The chitinase activity was high on colloidal chitin, regenerated chitin, partially N-acetylated chitin, and chitosan. N-bromosuccinamide completely inhibited the enzyme activity. This enzyme should be a good candidate for applications in the recycling of chitin waste.  相似文献   

10.
Cytochrome P450 BM-3 with the mutations A74G, F87V, and L188Q could catalyze indole to produce indigo and indirubin. To further enhance this capability, site-directed and random mutageneses on the monooxygenase domain of P450 BM-3 mutant (A74G/F87V/L188Q; 3X) were performed. The mutant libraries created by error-prone polymerase chain reaction were screened using a colorimetric colony-based method on agar plates followed by a spectroscopic assay involving in absorption of indigo at 670 nm and NADPH at 340 nm in microtiter plate. Three mutants (K434R/3X, E435D/3X, and D168N/A225V/K440N/3X) exhibited higher hydroxylation activity toward indole in comparison to parent enzyme. Moreover, using saturation site-directed mutagenesis at amino acid positions 168, 225, 434, 435, and 440, two P450 BM-3 variants (D168H/3X, E435T/3X) with an up to sixfold increase in catalytic efficiency (k cat/K m) were identified, and the mutant D168H/3X acquired higher regioselectivity resulting in more indigo (dimerized 3-hydroxy-indole) compared to parent mutant (93 vs72%).  相似文献   

11.
Summary: A chitin‐xylan hybrid polysaccharide having β(1 → 4)‐linked alternating structure of N‐acetyl‐D ‐glucosamine and D ‐xylose was synthesized via chitinase‐catalyzed polymerization. An oxazoline derivative of D ‐xylosyl‐β(1 → 4)‐N‐acetyl‐D ‐glucosamine ( 1 ) was effectively polymerized by the catalysis of chitinase from Bacillus sp., giving rise to a water‐soluble chitin‐xylan hybrid polysaccharide ( 2 ) in good yields. Molecular weights ( ) of 2 reached 1 500, which corresponds to 8–10 saccharide units.

A chitin‐xylan hybrid polysaccharide ( 2 ) synthesized via chitinase‐catalyzed polymerization.  相似文献   


12.
A macrocyclic mechanism‐based inhibitor for neuraminidases (NAs) bearing a 2‐difluoromethylphenyl aglycone and a linker between the aglycone and C‐9 positions of sialic acid was synthesized and evaluated. The macrocyclic structure was designed to keep the aglycone moiety in the active site of the neuraminidase after cleavage of the glycoside bond. When Vibrio chorelae neuraminidase (VCNA) was treated with a similar acyclic derivative in the presence of detergent, the irreversible inhibition property was disabled. In contrast, this macrocyclic compound acted as an irreversible inhibitor for VCNA in the presence of detergent. Inhibition assay for various NAs using this macrocyclic compound revealed that the irreversible inhibition property depends on the kcat of the neuraminidase treated. NAs having small kcat values, such as Influenza viruses, Clostridium, Trypanosoma cruzi, and Human, were also inhibited irreversibly. However, Salmonella typhimurium NA, which has an extremely high kcat, was not affected irreversibly by the inhibitor. Interestingly, in contrast to common kcat inhibitors, the irreversibility of inhibition by this macrocyclic compound is inversely proportional to the kcat of the target neuraminidase.  相似文献   

13.
Summary : We present the O2 binding properties of recombinant human serum albumin (rHSA) mutants complexed with an iron(II) protoporphyrin IX as a prosthetic heme group. Iron(III) protoporphyrin IX (hemin) is bound within subdomain IB of HSA with weak axial coordination by Tyr-161. In order to confer O2 binding capability to this naturally occurring hemoprotein: (i) a proximal histidine was introduced into position Ile-142; and (ii) the coordinated Tyr-161 was replaced with hydrophobic Leu using site-directed mutagenesis. It provided a recombinant HSA double-mutant [rHSA(I142H/Y161L) = rHSA(HL)]. The rHSA(HL)–heme formed a ferrous five-coordinate high-spin complex with axial ligation of His-142 under an Ar atmosphere. This artificial hemoprotein binds O2 at room temperature. Laser flash photolysis experiments demonstrated that O2 rebinidng to rHSA(HL)–heme displays monophasic kinetics, whereas the CO recombination process obeyed a double-exponential pattern. This might be attributable to the two different geometries of the axial imidazole coordination arising from the two orientations of the porphyrin plane in the heme pocket. The O2 binding affinity of rHSA(HL)–heme was considerably lower than those of R-state hemoglobin (Hb) and myoglobin (Mb), principally because of the high O2 dissociation rate constant. The third mutations have been introduced into the distal side of the heme (at position Leu-185 or Arg-186) to increase the O2 binidng affinity. The rHSA(HL/L185N)–heme showed high O2 binding affinity ( : 1 Torr), which is 18-fold greater than that of the original double mutant rHSA(HL)–heme and which is rather close to those of Hb (R-state) and Mb. Furthermore, replacement of polar Arg-186 with Leu or Phe adjusted the O2 binding affinity ( ) to 10 Torr, which is almost equivalent to value for human red blood cells.  相似文献   

14.
Glycoside hydrolase family 99 (GH99) was created to categorize sequence‐related glycosidases possessing endo‐α‐mannosidase activity: the cleavage of mannosidic linkages within eukaryotic N‐glycan precursors (Glc1–3Man9GlcNAc2), releasing mono‐, di‐ and triglucosylated‐mannose (Glc1–3‐1,3‐Man). GH99 family members have recently been implicated in the ability of Bacteroides spp., present within the gut microbiota, to metabolize fungal cell wall α‐mannans, releasing α‐1,3‐mannobiose by hydrolysing αMan‐1,3‐αMan→1,2‐αMan‐1,2‐αMan sequences within branches off the main α‐1,6‐mannan backbone. We report the development of a series of substrates and inhibitors, which we use to kinetically and structurally characterise this novel endo‐α‐1,2‐mannanase activity of bacterial GH99 enzymes from Bacteroides thetaiotaomicron and xylanisolvens. These data reveal an approximate 5 kJ mol?1 preference for mannose‐configured substrates in the ?2 subsite (relative to glucose), which inspired the development of a new inhibitor, α‐mannopyranosyl‐1,3‐isofagomine (ManIFG), the most potent (bacterial) GH99 inhibitor reported to date. X‐ray structures of ManIFG or a substrate in complex with wild‐type or inactive mutants, respectively, of B. xylanisolvens GH99 reveal the structural basis for binding to D ‐mannose‐ rather than D ‐glucose‐configured substrates.  相似文献   

15.
GDP-mannose mannosyl hydrolase (GDPMH) from E. coli catalyzes the hydrolysis of GDP-α-d-sugars to GDP and β-d-sugars by nucleophilic substitution with inversion at the anomeric C1 of the sugar, with general base catalysis by His-124. The 1.3 Å X-ray structure of the GDPMH-Mg2+-GDP complex was used to model the complete substrate, GDP-mannose into the active site. The substrate is linked to the enzyme by 12 hydrogen bonds, as well as by the essential Mg2+. In addition, His-124 was found to participate in a hydrogen bonded triad: His-124-NδHTyr-127-OHPro-120(CO). The contributions of these hydrogen bonds to substrate binding and to catalysis were investigated by site-directed mutagenesis. The hydrogen bonded triad detected in the X-ray structure was found to contribute little to catalysis since the Y127F mutation of the central residue shows only 2-fold decreases in both kcat and Km. The GDP leaving group is activated by the essential Mg2+ which contributes at least 105-fold to kcat, and by nine hydrogen bonds, including those from Tyr-103, Arg-37, Arg-52, and Arg-65 (via an intervening water), each of which contribute factors to kcat ranging from 24- to 309-fold. Both Arg-37 and Tyr-103 bind the β-phosphate of the leaving GDP and are only 5.0 Å apart. Accordingly, the R37Q/Y103F double mutant shows partially additive effects of the two single mutants on kcat, indicating cooperativity of Arg-37 and Tyr-103 in promoting catalysis. The extensive activation of the GDP leaving group suggests a mechanism with dissociative character with a cationic oxocarbenium-like transition state and a half-chair conformation of the sugar ring, as found with glycosidase enzymes. Accordingly, Asp-22 which contributes 102.1- to 102.6-fold to kcat, is positioned to both stabilize a developing cationic center at C1 and to accept a hydrogen bond from the C2–OH of the mannosyl group, and His-88, which contributes 102.3-fold to kcat, is positioned to accept a hydrogen bond from the C3–OH of the mannose facilitating its distortion to a half-chair conformation. Also, the fluorinated substrate GDP-2-fluoro-α-d-mannose, for which the oxocarbenium ion-like transition state centered at C1 would be destabilized by electron withdrawal, shows a 16-fold lower kcat and a 2.5-fold greater Km than does GDP-α-d-mannose. The product of the contributions to catalysis of Arg-37 and Tyr-103 (taking their cooperativity into account), Arg-52, Arg-65, Mg2+, Asp-22, His-124, and His-88 is ≥1019, which exceeds the 1012-fold rate acceleration produced by GDPMH by a factor ≥107. Hence, additional pairs or groups of catalytic residues must act cooperatively to promote catalysis.  相似文献   

16.
Berkecz  R.  T&#;r&#;k  R.  Ilisz  I.  Forr&#;  E.  F&#;l&#;p  F.  Armstrong  D. W.  P&#;ter  A. 《Chromatographia》2006,63(13):S37-S43

Direct reversed-phase high-performance liquid chromatographic methods were developed for the separation of the enantiomers of tricyclic β-lactams, cis-3,4-benzo-6-azabicyclo[3.2.0]heptan-7-one, cis-4,5-benzo-7-azabicyclo[4.2.0]-octan-8-one, cis-5,6-benzo-8-azabicyclo[5.2.0]nonan-9-one and new bicyclic β-amino acids, the six- and seven-membered homologues of cis-1-amino-4,5-benzocyclopentane-2-carboxylic acid (benzocispentacin), cis-1-amino-5,6-benzocyclohexane-2-carboxylic acid and cis-1-amino-6,7-benzocycloheptane-2-carboxylic acid. The direct separations of the analytes were performed on chiral stationary phase (CSP) columns containing the macrocyclic glycopeptide antibiotic teicoplanin (Chirobiotic T), teicoplanin aglycone (Chirobiotic TAG), vancomycin (Chirobiotic V), vancomycin aglycone (Chirobiotic VAG), ristocetin A (Chirobiotic R) or a new dimethylphenyl carbamate-derivatized β-cyclodextrin-based Cyclobond DMP. The results achieved with the different methods were compared in systematic chromatographic examinations. The effects of an organic modifier and of the mobile phase composition on the separation and the separation efficiency of different columns were investigated. The difference in enantioselective free energy between the aglycone CSP and the teicoplanin CSP for these β-lactams and β-amino acids ranged between 0.3 and −1.1 kJmol−1. Better enantioseparations were attained in most cases on the aglycone CSP.

  相似文献   

17.
A furostanol saponin mixture and a known phytoecdysteroid were isolated from the roots of Helleborus orientalis Lam. Their structures were established as 26-[(β-D-glucopyranosyl)oxy]-22α-hidroxyfurosta-5,25(27)-dien-1β,3β,11α-triol (1a), 26-[(β-D-glucopyranosyl)oxy]-22α-methoxyfurosta-5,25(27)-dien-1β,3β,11α-triol (1b), and 20-hydroxy-β-ecdyson-3-O-β-D-glycoside (2). Acid hydrolysis of 1a,b gave (1β,3β,11α,22α)-22,26-dimethoxyfurosta-5,25(27)-dien-1,3,11-triol (aglycone 1) and of 2 gave 20-hydroxy-β-ecdyson (aglycone 2). Their structures were elucidated by spectral analysis. Published in Khimiya Prirodnykh Soedinenii, No. 1, pp. 75–77, January–February, 2007.  相似文献   

18.
19.
A Y‐shaped diblock copolymer with a functional block poly(glycidyl methacrylate) was synthesized via the combination of enzymatic ring‐opening polymerization (eROP) and atom transfer radical polymerization (ATRP). The synthetic procedure involved eROP of ε‐caprolactone (ε‐CL) in the presence of biocatalyst Novozyme 435 and initiator 1H,1H,2H,2H‐perfluoro‐1‐octaoxy, subsequently the resulting poly(ε‐caprolactone) (PCL) was converted to a macroinitiator by esterification of it with 2,2‐dichloro acetyl chloride, and finally the ATRP of glycidyl methacrylate (GMA) was conducted at 60 °C with CuCl/2,2′‐bipyridine as the catalyst system. By this process, we obtained copolymers with a controlled molecular weight and a low polydispersity. The structure and composition of the obtained polymers were characterized by H NMR, GPC, and IR. Linear first‐order kinetics, linearly increased molecular weight with conversion, and low polydispersities were observed for the ATRP of GMA. The thermal properties of the copolymer were characterized by differential scanning calorimetry. The self‐assembly behavior of the Y‐shaped block copolymer was also investigated in different solvents and at different concentrations. The aggregates of various morphologies (spheres, worm‐like patterns, nanowell patterns, and dendritic patterns) were observed. It was found that solvents remarkably influenced the morphologies of the films spin‐coated from the corresponding solutions. © 2009 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 47: 5509–5526, 2009  相似文献   

20.
The mutation sites of the four mutants F35Y, P40V, V45E and V45Y of cytochrome b5 are located at the edge of the heme-binding pocket. The solvent accessible areas of the “pocket inte-rior“ of the four mutants and the wild-type cytochrome b5 have been calculated based on their crystal structures at high resolu-tion. The change in the hydrophobicity of the heme-binding pocket resulting from the mutation can be quantitatively de-scribed using the difference of the solvent accessible area of the “pocket interior“ of each mutant from that of the wild-type cy-tochrome b5. The influences of the hydrophobicity of the heme-binding pocket on the protein stability and redox potential are discussed.  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