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.
The substrate specificity of Escherichia coli N-acetylneuraminic acid lyase was previously switched from the natural condensation of pyruvate with N-acetylmannosamine, yielding N-acetylneuraminic acid, to the aldol condensation generating N-alkylcarboxamide analogues of N-acetylneuraminic acid. This was achieved by a single mutation of Glu192 to Asn. In order to analyze the structural changes involved and to more fully understand the basis of this switch in specificity, we have isolated all 20 variants of the enzyme at position 192 and determined the activities with a range of substrates. We have also determined five high-resolution crystal structures: the structures of wild-type E. coli N-acetylneuraminic acid lyase in the presence and in the absence of pyruvate, the structures of the E192N variant in the presence and in the absence of pyruvate, and the structure of the E192N variant in the presence of pyruvate and a competitive inhibitor (2R,3R)-2,3,4-trihydroxy-N,N-dipropylbutanamide. All structures were solved in space group P21 at resolutions ranging from 1.65 Å to 2.2 Å. A comparison of these structures, in combination with the specificity profiles of the variants, reveals subtle differences that explain the details of the specificity changes. This work demonstrates the subtleties of enzyme-substrate interactions and the importance of determining the structures of enzymes produced by directed evolution, where the specificity determinants may change from one substrate to another.  相似文献   

2.
3.
N-Acetyl-l-glutamate kinase (NAGK), the paradigm enzyme of the amino acid kinase family, catalyzes the second step of arginine biosynthesis. Although substrate binding and catalysis were clarified by the determination of four crystal structures of the homodimeric Escherichia coli enzyme (EcNAGK), we now determine 2 Å resolution crystal structures of EcNAGK free from substrates or complexed with the product N-acetyl-l-glutamyl-5-phosphate (NAGP) and with sulfate, which reveal a novel, very open NAGK conformation to which substrates would associate and from which products would dissociate. In this conformation, the C-domain, which hosts most of the nucleotide site, rotates ∼ 24°-28° away from the N-domain, which hosts the acetylglutamate site, whereas the empty ATP site also exhibits some changes. One sulfate is found binding in the region where the β-phosphate of ATP normally binds, suggesting that ATP is first anchored to the β-phosphate site, before perfect binding by induced fit, triggering the shift to the closed conformation. In contrast, the acetylglutamate site is always well formed, although its β-hairpin lid is found here to be mobile, being closed only in the subunit of the EcNAGK-NAGP complex that binds NAGP most strongly. Lid closure appears to increase the affinity for acetylglutamate/NAGP and to stabilize the closed enzyme conformation via lid-C-domain contacts. Our finding of NAGP bound to the open conformation confirms that this product dissociates from the open enzyme form and allows reconstruction of the active center in the ternary complex with both products, delineating the final steps of the reaction, which is shown here by site-directed mutagenesis to involve centrally the invariant residue Gly11.  相似文献   

4.
The crystal structure of poly(3-hydroxybutyrate) (PHB) depolymerase PhaZ7 purified from Paucimonas lemoignei was determined at 1.90 Å resolution. The structure consists of a single domain with an α/β hydrolase fold in its core. The active site is analogous to that of serine esterases/lipases and is characterized by the presence of a catalytic triad comprising Ser136, Asp242, and His306. Comparison with other structures in the Protein Data Bank showed a high level of similarity with the Bacillus subtilis lipase LipA (RMSD, 1.55 Å). Structural comparison with Penicillium funiculosum PHB depolymerase, the only PHB depolymerase whose structure is already known, revealed significant differences, resulting in an RMSD of 2.80-3.58 Å. The two enzymes appear to utilize different types of solvent-exposed residues for biopolymer binding, with aliphatic and hydroxyl residues used in P. funiculosum PHB depolymerase and aromatic residues in PhaZ7. Moreover, the active site of P. funiculosum PHB depolymerase is accessible to the substrate in contrast to the active site of PhaZ7, which is buried. Hence, considerable conformational changes are required in PhaZ7 for the creation of a channel leading to the active site. Taken together, the structural data suggest that PhaZ7 and P. funiculosum PHB depolymerase have adopted different strategies for effective substrate binding in response to their diverse substrate specificity and the lack of a substrate-binding domain.  相似文献   

5.
Electron transfer reactions between optically-active RuII/III complexes incorporating (S)-/(R)-amino acids, and the two azurins, azurin-1 (az-1Cu) and azurin-2 (az-2Cu) isolated from Alcaligenes xylosoxidans GIFU 1051, have been studied to probe molecular recognition sites on the two azurins. The RuII/III complexes are K[RuII(L)(bpy)] and [RuIII(L)(bpy)], and have a tripodal ligand (L) derived from the (S)-/(R)-amino acids, which are in turn exchanged for other functional substituent groups, such as (S)-/(R)-phenylalanine, -leucine, -valine, -alanine, and -glutamic acid (L = (S)-/(R)-BCMPA, -BCMLE, -BCMVA, -BCMAL, and -BCMGA). In the oxidation reaction of az-1CuI promoted by the RuIII complexes, the kinetic parameters exhibited enantio- and stereo-selectivities, while the same reaction of az-2CuI was less enantio- and stereo-selective. These differences suggest that the processes of formation of the activated states are different for the two azurins. On the other hand, such a difference has not been observed for az-1 and az-2 with respect to the reduction reactions promoted by both azurins CuII by the RuII complexes within the experimental error. This suggests that the neutrality of the Ru complexes is important for precise molecular recognition of azurins. His117 has been proposed as the electron transfer site. The local structures in the vicinity of the His117 side chain in the two azurins, are essentially identical with the exception of the 43rd residue, Val43 and Ala43 for az-1 and az-2, respectively. Electron transfer reactions between RuIII complexes and a mutant azurin, V43A-az-1, were also carried out. Interestingly, the activation parameters estimated were very similar to those of az-2, indicating that the 43rd residue acts as the electron transfer site in azurins and provides rationalization for the different mechanisms of az-1 and az-2 in redox reactions.  相似文献   

6.
Aminoaldehyde dehydrogenases (AMADHs, EC 1.2.1.19) belong to the large aldehyde dehydrogenase (ALDH) superfamily, namely, the ALDH9 family. They oxidize polyamine-derived ω-aminoaldehydes to the corresponding ω-amino acids. Here, we report the first X-ray structures of plant AMADHs: two isoenzymes, PsAMADH1 and PsAMADH2, from Pisum sativum in complex with β-nicotinamide adenine dinucleotide (NAD+) at 2.4 and 2.15 Å resolution, respectively. Both recombinant proteins are dimeric and, similarly to other ALDHs, each monomer is composed of an oligomerization domain, a coenzyme binding domain and a catalytic domain. Each subunit binds NAD+ as a coenzyme, contains a solvent-accessible C-terminal peroxisomal targeting signal (type 1) and a cation bound in the cavity close to the NAD+ binding site. While the NAD+ binding mode is classical for PsAMADH2, that for PsAMADH1 is unusual among ALDHs. A glycerol molecule occupies the substrate binding site and mimics a bound substrate. Structural analysis and substrate specificity study of both isoenzymes in combination with data published previously on other ALDH9 family members show that the established categorization of such enzymes into distinct groups based on substrate specificity is no more appropriate, because many of them seem capable of oxidizing a large spectrum of aminoaldehyde substrates. PsAMADH1 and PsAMADH2 can oxidize N,N,N-trimethyl-4-aminobutyraldehyde into γ-butyrobetaine, which is the carnitine precursor in animal cells. This activity highly suggests that in addition to their contribution to the formation of compatible osmolytes such as glycine betaine, β-alanine betaine and γ-aminobutyric acid, AMADHs might participate in carnitine biosynthesis in plants.  相似文献   

7.
The emergence of resistance to vancomycin and related glycopeptide antibiotics is spurring efforts to develop new antimicrobial therapeutics. High-resolution structural information about antibiotic-ligand recognition should prove valuable in the rational design of improved drugs. We have determined the X-ray crystal structure of the complex of vancomycin with N-acetyl-d-Ala-d-Ala, a mimic of the natural muramyl peptide target, and refined this structure at a resolution of 1.3 Å to R and Rfree values of 0.172 and 0.195, respectively. The crystal asymmetric unit contains three back-back vancomycin dimers; two of these dimers participate in ligand-mediated face-face interactions that produce an infinite chain of molecules running throughout the crystal. The third dimer packs against the side of a face-face interface in a tight “side-side” interaction that involves both polar contacts and burial of hydrophobic surface. The trimer of dimers found in the asymmetric unit is essentially identical to complexes seen in three other crystal structures of glycopeptide antibiotics complexed with peptide ligands. These four structures are derived from crystals belonging to different space groups, suggesting that the trimer of dimers may not be simply a crystal packing artifact and prompting us to ask if ligand-mediated oligomerization could be observed in solution. Using size-exclusion chromatography, dynamic light scattering, and small-angle X-ray scattering, we demonstrate that vancomycin forms discrete supramolecular complexes in the presence of tripeptide ligands. Size estimates for these complexes are consistent with assemblies containing four to six vancomycin monomers.  相似文献   

8.
The structures of rice BGlu1 β-glucosidase, a plant β-glucosidase active in hydrolyzing cell wall-derived oligosaccharides, and its covalent intermediate with 2-deoxy-2-fluoroglucoside have been solved at 2.2 Å and 1.55 Å resolution, respectively. The structures were similar to the known structures of other glycosyl hydrolase family 1 (GH1) β-glucosidases, but showed several differences in the loops around the active site, which lead to an open active site with a narrow slot at the bottom, compatible with the hydrolysis of long β-1,4-linked oligosaccharides. Though this active site structure is somewhat similar to that of the Paenibacillus polymyxa β-glucosidase B, which hydrolyzes similar oligosaccharides, molecular docking studies indicate that the residues interacting with the substrate beyond the conserved -1 site are completely different, reflecting the independent evolution of plant and microbial GH1 exo-β-glucanase/β-glucosidases. The complex with the 2-fluoroglucoside included a glycerol molecule, which appears to be in a position to make a nucleophilic attack on the anomeric carbon in a transglycosylation reaction. The coordination of the hydroxyl groups suggests that sugars are positioned as acceptors for transglycosylation by their interactions with E176, the catalytic acid/base, and Y131, which is conserved in barley BGQ60/β-II β-glucosidase, that has oligosaccharide hydrolysis and transglycosylation activity similar to rice BGlu1. As the rice and barley enzymes have different preferences for cellobiose and cellotriose, residues that appeared to interact with docked oligosaccharides were mutated to those of the barley enzyme to see if the relative activities of rice BGlu1 toward these substrates could be changed to those of BGQ60. Although no single residue appeared to be responsible for these differences, I179, N190 and N245 did appear to interact with the substrates.  相似文献   

9.
Carbamate kinase (CK) makes ATP from ADP and carbamoyl phosphate (CP) in the final step of the microbial fermentative catabolism of arginine, agmatine, and oxalurate/allantoin. Two previously reported CK structures failed to clarify CP binding and catalysis and to reveal the significance of the protruding subdomain (PSD) that hangs over the CK active center as an exclusive and characteristic CK feature. We clarify now these three questions by determining two crystal structures of Enterococcus faecalis CK (one at 1.5 Å resolution and containing bound MgADP, and the other at 2.1 Å resolution and having in the active center one sulfate and two fixed water molecules that mimic one bound CP molecule) and by mutating active-center residues, determining the consequences of these mutations on enzyme functionality. Superimposition of the present crystal structures reconstructs the filled active center in the ternary complex, immediately suggesting in-line associative phosphoryl group transfer and a mechanism for enzyme catalysis involving N51, K209, K271, D210, and the PSD residue K128. The large respective increases and decreases in KmCP and kcat triggered by the mutations N51A, K128A, K209A, and D210N corroborate the ternary complex active-site architecture and the catalytic mechanism proposed. The extreme negative effects of K128A demonstrate a key role of the PSD in substrate binding and catalysis. The crystal structures reveal large rigid-body movements of the PSD towards the enzyme body that place K128 next to CP and bury the CP site. A mechanism that connects CP site occupation with the PSD approach, involving V206-I207 in the CP site and P162-S163 in the PSD stem, is identified. The effects of the V206A and V206L mutations support this mechanism. It is concluded that the PSD movement allows CK to select against the abundant CP/carbamate analogues acetylphosphate/acetate and bicarbonate, rendering CK highly selective for CP/carbamate.  相似文献   

10.
The specific insertion of a divalent metal ion into tetrapyrrole macrocycles is catalyzed by a group of enzymes called chelatases. Distortion of the tetrapyrrole has been proposed to be an important component of the mechanism of metallation. We present the structures of two different inhibitor complexes: (1) N-methylmesoporphyrin (N-MeMP) with the His183Ala variant of Bacillus subtilis ferrochelatase; (2) the wild-type form of the same enzyme with deuteroporphyrin IX 2,4-disulfonic acid dihydrochloride (dSDP). Analysis of the structures showed that only one N-MeMP isomer out of the eight possible was bound to the protein and it was different from the isomer that was earlier found to bind to the wild-type enzyme. A comparison of the distortion of this porphyrin with other porphyrin complexes of ferrochelatase and a catalytic antibody with ferrochelatase activity using normal-coordinate structural decomposition reveals that certain types of distortion are predominant in all these complexes. On the other hand, dSDP, which binds closer to the protein surface compared to N-MeMP, does not undergo any distortion upon binding to the protein, underscoring that the position of the porphyrin within the active site pocket is crucial for generating the distortion required for metal insertion. In addition, in contrast to the wild-type enzyme, Cu2+-soaking of the His183Ala variant complex did not show any traces of porphyrin metallation. Collectively, these results provide new insights into the role of the active site residues of ferrochelatase in controlling stereospecificity, distortion and metallation.  相似文献   

11.
The structures of the solvated iron(II) and iron(III) ions have been studied in solution and solid state by extended X-ray absorption fine structure (EXAFS) in three oxygen donor solvents, water, dimethylsulfoxide (Me2SO), N,N′-dimethylpropyleneurea (DMPU), and one sulfur donor solvent, N,N-dimethylthioformamide (DMTF); these solvents have different coordination and solvation properties. In addition, the structure of hexakis(dimethylsulfoxide)iron(III) perchlorate has been determined crystallographically to support the determination of the corresponding solvate in solution. The hydrated, the dimethylsulfoxide and N,N-dimethylthioformamide solvated iron(II) ions show regular octahedral coordination in both solution and solid state with mean Fe-O, Fe-O, and Fe-S bond distances of 2.10, 2.10, and 2.52 Å, respectively, whereas the N,N′-dimethylpropyleneurea iron(II) solvate is five-coordinated, d(Fe-O) = 2.06 Å. The compounds vary in color from light green (hydrate) to dark orange or red (DMPU). The hydrated iron(III) ion in aqueous solution and the dimethylsulfoxide solvated iron(III) ions in solution and solid state show the expected octahedral coordination, the Fe-O bond distances are 2.00 Å for both, whereas the N,N′-dimethylpropyleneurea iron(III) solvate is found to be five-coordinated with a mean Fe-O bond distance of 1.99 Å. The N,N-dimethylthioformamide solvated iron(III) ion in the solid perchlorate salt is tetrahedrally four-coordinated, the mean Fe-S bond distance is 2.20 Å. Iron(III) is reduced with time to iron(II) in N,N-dimethylthioformamide solution. The compounds vary in color from pale yellow (hydrate) to blackish red (DMPU).  相似文献   

12.
Two isomers (R,S,R,S- and R,R,S,S-) of five coordinate complex [Cu(L)Cl]+ have been separated and characterised. These two isomers have significantly different spectrochemical and electrochemical properties. Absorption maximum of R,S,R,S-[Cu(L)Cl]+ shifts to longer wavelength and its reduction potential shifts to more positive direction comparing those of R,R,S,S-[Cu(L)Cl]+. R,S,R,S-[Cu(L)Cl]+ is significantly distorted to trigonal-bipyramidal structure, whereas R,R,S,S-[Cu(L)Cl]+ retains almost square-planar geometry. The average bond distance of Cu-N in basal plane of R,S,R,S-[Cu(L)Cl]+ is longer by 0.024 Å than that of R,R,S,S-[Cu(L)Cl]+, whereas the bond distance of Cu-Cl in former is shorter by 0.200 Å than that in latter. The isolated square-planar complexes of R,R,S,S- and R,S,R,S-[Cu(L)](ClO4)2 are converted to the R,R,S,S- and R,S,R,S-[Cu(L)Cl]+ by the addition of Cl in nitromethane solution with the rate constants, k=1.70 (±0.02) and 8.31 (±0.07) M−1 s−1, respectively.  相似文献   

13.
As part of an effort to inhibit S100B, structures of pentamidine (Pnt) bound to Ca2+-loaded and Zn2+,Ca2+-loaded S100B were determined by X-ray crystallography at 2.15 Å (Rfree = 0.266) and 1.85 Å (Rfree = 0.243) resolution, respectively. These data were compared to X-ray structures solved in the absence of Pnt, including Ca2+-loaded S100B and Zn2+,Ca2+-loaded S100B determined here (1.88 Å; Rfree = 0.267). In the presence and absence of Zn2+, electron density corresponding to two Pnt molecules per S100B subunit was mapped for both drug-bound structures. One Pnt binding site (site 1) was adjacent to a p53 peptide binding site on S100B (± Zn2+), and the second Pnt molecule was mapped to the dimer interface (site 2; ± Zn2+) and in a pocket near residues that define the Zn2+ binding site on S100B. In addition, a conformational change in S100B was observed upon the addition of Zn2+ to Ca2+-S100B, which changed the conformation and orientation of Pnt bound to sites 1 and 2 of Pnt-Zn2+,Ca2+-S100B when compared to Pnt-Ca2+-S100B. That Pnt can adapt to this Zn2+-dependent conformational change was unexpected and provides a new mode for S100B inhibition by this drug. These data will be useful for developing novel inhibitors of both Ca2+- and Ca2+,Zn2+-bound S100B.  相似文献   

14.
ALKBH5, a member of AlkB family proteins, has been reported as a mammalian N6-methyladenosine (m6A) RNA demethylase. Here we report the crystal structure of zebrafish ALKBH5 (fALKBH5) with the resolution of 1.65 Å. Structural superimposition shows that fALKBH5 is comprised of a conserved jelly-roll motif. However, it possesses a loop that interferes potential binding of a duplex nucleic acid substrate, suggesting an important role in substrate selection. In addition, several active site residues are different between the two known m6A RNA demethylases, ALKBH5 and FTO, which may result in their slightly different pathways of m6A demethylation.  相似文献   

15.
Enzymes from the de novo purine biosynthetic pathway have been exploited for the development of anti-cancer drugs, and represent novel targets for anti-bacterial drug development. In Mycobacterium tuberculosis, the cause of tuberculosis, this pathway has been identified as essential for growth and survival. The structure of M. tuberculosis PurN (MtPurN) has been determined in complex with magnesium and iodide at 1.30 Å resolution, and with cofactor analogue, 5-methyltetrahydrofolate (5MTHF) at 2.2 Å resolution. The structure shows a Rossmann-type fold that is very similar to the known structures of the human and E. coli PurN proteins. In contrast, MtPurN forms a dimer that is quite different from that formed by the Escherichia coli PurN, and which suggests a mechanism whereby communication could take place between the two active sites. Differences are seen in two active site loops and in the binding mode of the 5MTHF cofactor analogue between the two MtPurN molecules of the dimer. A binding site for halide ions is found in the dimer interface, and bound magnesium and iodide ions in the active site suggest sites that might be exploited in potential drug discovery strategies.  相似文献   

16.
Maricaulis maris N-acetylglutamate synthase/kinase (mmNAGS/K) catalyzes the first two steps in l-arginine biosynthesis and has a high degree of sequence and structural homology to human N-acetylglutamate synthase, a regulator of the urea cycle. The synthase activity of both mmNAGS/K and human NAGS are regulated by l-arginine, although l-arginine is an allosteric inhibitor of mmNAGS/K, but an activator of human NAGS. To investigate the mechanism of allosteric inhibition of mmNAGS/K by l-arginine, we have determined the structure of the mmNAGS/K complexed with l-arginine at 2.8 Å resolution. In contrast to the structure of mmNAGS/K in the absence of l-arginine where there are conformational differences between the four subunits in the asymmetric unit, all four subunits in the l-arginine liganded structure have very similar conformations. In this conformation, the AcCoA binding site in the N-acetyltransferase (NAT) domain is blocked by a loop from the amino acid kinase (AAK) domain, as a result of a domain rotation that occurs when l-arginine binds. This structural change provides an explanation for the allosteric inhibition of mmNAGS/K and related enzymes by l-arginine. The allosterically regulated mechanism for mmNAGS/K differs significantly from that for Neisseria gonorrhoeae NAGS (ngNAGS). To define the active site, several residues near the putative active site were mutated and their activities determined. These experiments identify roles for Lys356, Arg386, Asn391 and Tyr397 in the catalytic mechanism.  相似文献   

17.
We report the molecular cloning and characterization of two novel β-N-acetylhexosaminidases (β-HEX, EC 3.2.1.52) from Paenibacillus sp. strain TS12. The two β-HEXs (Hex1 and Hex2) were 70% identical in primary structure, and the N-terminal region of both enzymes showed significant similarity with β-HEXs belonging to glycoside hydrolase family 20 (GH20). Interestingly, however, the C-terminal region of Hex1 and Hex2 shared no sequence similarity with the GH20 β-HEXs or other known proteins. Both recombinant enzymes, expressed in Escherichia coli BL21(DE3), hydrolyzed the β-N-acetylhexosamine linkage of chitooligosaccharides and glycosphingolipids such as asialo GM2 and Gb4Cer in the absence of detergent. However, the enzyme was not able to hydrolyze GM2 ganglioside in the presence or in the absence of detergent. We determined three crystal structures of Hex1; the Hex1 deletion mutant Hex1-ΔC at a resolution of 1.8 Å; Hex1-ΔC in complex with β-N-acetylglucosamine at 1.6 Å; and Hex1-ΔC in complex with β-N-acetylgalactosamine at 1.9 Å. We made a docking model of Hex1-ΔC with GM2 oligosaccharide, revealing that the sialic acid residue of GM2 could hinder access of the substrate to the active site cavity. This is the first report describing the molecular cloning, characterization and X-ray structure of a procaryotic β-HEX capable of hydrolyzing glycosphingolipids.  相似文献   

18.
Free methionine-R-sulfoxide reductase (fRMsr) reduces free methionine R-sulfoxide back to methionine, but its catalytic mechanism is poorly understood. Here, we have determined the crystal structures of the reduced, substrate-bound, and oxidized forms of fRMsr from Staphylococcus aureus. Our structural and biochemical analyses suggest the catalytic mechanism of fRMsr in which Cys102 functions as the catalytic residue and Cys68 as the resolving Cys that forms a disulfide bond with Cys102. Cys78, previously thought to be a catalytic Cys, is a non-essential residue for catalytic function. Additionally, our structures provide insights into the enzyme-substrate interaction and the role of active site residues in substrate binding. Structural comparison reveals that conformational changes occur in the active site during catalysis, particularly in the loop of residues 97–106 containing the catalytic Cys102. We have also crystallized a complex between fRMsr and isopropyl alcohol, which acts as a competitive inhibitor for the enzyme. This isopropyl alcohol-bound structure helps us to understand the inhibitory mechanism of fRMsr. Our structural and enzymatic analyses suggest that a branched methyl group in alcohol seems important for competitive inhibition of the fRMsr due to its ability to bind to the active site.  相似文献   

19.
20.
We used small-angle X-ray solution scattering (SAXS) technique to investigate the nucleotide-mediated conformational changes of the head domains [subfragment 1 (S1)] of myosin V and VI processive motors that govern their directional preference for motility on actin. Recombinant myosin V-S1 with two IQ motifs (MV-S1IQ2) and myosin VI-S1 (MVI-S1) were engineered from Sf9 cells using a baculovirus expression system. The radii of gyration (Rg) of nucleotide-free MV-S1IQ2 and MVI-S1 were 48.6 and 48.8 Å, respectively. In the presence of ATP, the Rg value of MV-S1IQ2 decreased to 46.7 Å, while that of MVI-S1 increased to 51.7 Å, and the maximum chord length of the molecule decreased by ca 9% for MV-S1IQ2 and increased by ca 6% for MVI-S1. These opposite directional changes were consistent with those occurring in S1s with ADP and Vi or AlF4− 2 bound (i.e., in states mimicking the ADP/Pi-bound state of ATP hydrolysis). Binding of AMPPNP induced Rg changes of both constructs similar to those in the presence of ATP, suggesting that the timing of the structural changes for their motion on actin is upon binding of ATP (the pre-hydrolysis state) during the ATPase cycle. Binding of ADP to MV-S1IQ2 and MVI-S1 caused their Rg values to drop below those in the nucleotide-free state. Thus, upon the release of Pi, the reverse conformational change could occur, coupling to drive the directional motion on actin. The amount of Rg change upon the release of Pi was ca 6.4 times greater in MVI-S1 than in MV-S1IQ2, relating to the production of the large stroke of the MVI motor during its translocation on actin. Atomic structural models for these S1s based upon the ab initio shape reconstruction from X-ray scattering data were constructed, showing that MVI-S1 has the light-chain-binding domain positioned in the opposite direction to MV-S1IQ2 in both the pre- and post-powerstroke transition. The angular change between the light chain-binding domains of MV-S1IQ2 in the pre- to post-powerstroke transition was ∼ 50°, comparable to that of MII-S1. On the other hand, that of MVI-S1 was ∼ 100° or ∼ 130° much less than the currently postulated changes to allow the maximal stroke size of myosin VI-S1 but still significantly larger than those of other myosins reported so far. The results suggest that some additional alterations or elements are required for MVI-S1 to take maximal working strokes along the actin filament.  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