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 62 毫秒
1.
The primary structures for several members of both the vicilin and legumin families of storage proteins were examined using a computer routine based on amino acid physical characteristics. The comparison algorithm revealed that sequences from the two families could be aligned and share a number of predicted secondary structural features. The COOH-terminal half of the subunits in both families displayed a highly conserved core region that was largely hydrophobic and in which a high proportion of the residues were predicted to be in beta-sheet conformations. The central region of the molecules which contained mixed areas of predicted helical and sheet conformations showed more variability in residue selection than the COOH-terminal regions. The NH2-terminal segments of subunits from the two different families could not be aligned though they characteristically had a high proportion of residues predicted to be in helical conformations. The feature which most clearly distinguished subunits between the two families was an inserted span in the legumin group with a high proportion of acidic amino acids located between the central and COOH-terminal domains. Residues in this insertion were predicted to exist mainly in helical conformation. Since considerable size variation occurs in this area amongst the legumin subunits, alterations in this region may have a minimal detrimental effect on the structure of the proteins.  相似文献   

2.
Recognition of Ras by its downstream target Raf is mediated by a Ras-recognition region in the Ras-binding domain (RBD) of Raf. Residues 78–89 in this region occupy two different conformations in the ensemble of NMR solution structures of the RBD: a fully α-helical one, and one where 87–90 form a type IV β-turn. Molecular dynamics simulations of the RBD in solution were performed to explore the stability of these and other possible conformations of both the wild-type RBD and the R89K mutant, which does not bind Ras. The simulations sample a fully helical conformation for residues 78–89 similar to the NMR helical structures, a conformation where 85–89 form a 310-helical turn, and a conformation where 87–90 form a type I |iB-turn, whose free energies are all within 0.3 kcal/mol of each other. NOE patterns and Hα chemical shifts from the simulations are in reasonable agreement with experiment. The NMR turn structure is calculated to be 3 kcal/mol higher than the three above conformations. In a simulation with the same implicit solvent model used in the NMR structure generation, the turn conformation relaxes into the fully helical conformation, illustrating possible structural artifacts introduced by the implicit solvent model. With the Raf R89K mutant, simulations sample a fully helical and a turn conformation, the turn being 0.9 kcal/mol more stable. Thus, the mutation affects the population of RBD conformations, and this is expected to affect Ras binding. For example, if the fully helical conformation of residues 78–89 is required for binding, its free energy increase in R89K will increase the binding free energy by about 0.6 kcal/mol. Proteins 31:186–200, 1998. © 1998 Wiley-Liss, Inc.  相似文献   

3.
4.
Structural studies of the streptavidin binding loop.   总被引:7,自引:5,他引:2       下载免费PDF全文
The streptavidin-biotin complex provides the basis for many important biotechnological applications and is an interesting model system for studying high-affinity protein-ligand interactions. We report here crystallographic studies elucidating the conformation of the flexible binding loop of streptavidin (residues 45 to 52) in the unbound and bound forms. The crystal structures of unbound streptavidin have been determined in two monoclinic crystal forms. The binding loop generally adopts an open conformation in the unbound species. In one subunit of one crystal form, the flexible loop adopts the closed conformation and an analysis of packing interactions suggests that protein-protein contacts stabilize the closed loop conformation. In the other crystal form all loops adopt an open conformation. Co-crystallization of streptavidin and biotin resulted in two additional, different crystal forms, with ligand bound in all four binding sites of the first crystal form and biotin bound in only two subunits in a second. The major change associated with binding of biotin is the closure of the surface loop incorporating residues 45 to 52. Residues 49 to 52 display a 3(10) helical conformation in unbound subunits of our structures as opposed to the disordered loops observed in other structure determinations of streptavidin. In addition, the open conformation is stabilized by a beta-sheet hydrogen bond between residues 45 and 52, which cannot occur in the closed conformation. The 3(10) helix is observed in nearly all unbound subunits of both the co-crystallized and ligand-free structures. An analysis of the temperature factors of the binding loop regions suggests that the mobility of the closed loops in the complexed structures is lower than in the open loops of the ligand-free structures. The two biotin bound subunits in the tetramer found in the MONO-b1 crystal form are those that contribute Trp 120 across their respective binding pockets, suggesting a structural link between these binding sites in the tetramer. However, there are no obvious signatures of binding site communication observed upon ligand binding, such as quaternary structure changes or shifts in the region of Trp 120. These studies demonstrate that while crystallographic packing interactions can stabilize both the open and closed forms of the flexible loop, in their absence the loop is open in the unbound state and closed in the presence of biotin. If present in solution, the helical structure in the open loop conformation could moderate the entropic penalty associated with biotin binding by contributing an order-to-disorder component to the loop closure.  相似文献   

5.
A series of copolymers of β-p-nitrobenzyl L -aspartate with β-benzyl L -aspartate and with β-mcthyl L -aspartatc in helix-supporting and helix-breaking conditions have been reexamined by using ultraviolet isotropic, absorption, optical rotatory dispersion, and circular dichroism techniques. Many different conformations are apparent, depending on solvent and temperature. Chloroform, trifluoroethanol, and methylene dichloride support the left-handed helical conformation of the copolymers containing less than about 20 mole-% nitroaromatic residues and the right-handed helical conformation of the copolymers containing more than approximately 30 mole-% nitroaromatic residues. In trifluoroacetic acid all the copolymers are in a random-coil conformation. In hexa-fluoroacetone trihydrate and in trimethyl phosphate, the copolypeptides with low nitroaromatic residues content are predominantly in a disordered conformation, while those with high nitroaromatic residues content show a right-handed helical array. Reversible helix-ramlom-coil transitions are observed with increasing temperature in trimethyl phosphate. An example of right-handed-left-handed helix reversible transition with temperature is reported in a chloroform-trimethyl phosphate (2:1) mixture. Nitrobenzyl-nilrobenzyl side-chain interactions in chloroform, but not in trifluoroacetic acid or in trimethyl phosphate, have been confirmed. For the first time we report the circular dichroism spectra in which the n-π* peptide band of a left-handed helical conformation is almost completely evident.  相似文献   

6.
Conformational studies of -glucans   总被引:1,自引:0,他引:1  
A study of the effect of linkage on the possible conformations of di-and polysaccharides of α-D -glucose and also the probable intramolecular hydrogen bonds has been made. The differences in the nature of linkage is shown to effect the energetically preferred conformations; (1 → 2) linkages lead only to righthanded helical conformations, (1 → 3) linkages lead to extended as well as both left and righthanded helical conformations; (1 → 4) linkages lead to both right-and lefthanded wide helical conformations. The possible hydrogen bonds between adjacent residues are also dependent on the nature of the linkage. A comparison of the conformational data of α-D -glucans with those of β-D -glucans has indicated that the favored conformations and hydrogen bonds between contiguous residues in the chain are influenced by the configuration at the anomeric carbon atom in all the three types of polysaccharides. From the energy calculations a probable conformation (?M = ?10°, ψM = ?30°, ?N = ?23°, ψN = ?19°) has also been proposed for crystalline mycodextran in conformity with x-ray data. This conformation contains two types of hydrogen bonds between contiguous residues one between 0–2 and 0–3 atoms at (1 → 4) linkage and the other between 0?2 and 0–4 atoms at (1 → 3) linkage in the chain. The conformation of maltose unit (?10°,?30°) that is likely to occur in the crystalline mycodextran coincides with the minimum energy conformation of maltose.  相似文献   

7.
The LAH4 family of histidine-rich peptides exhibits potent antimicrobial and DNA transfection activities, both of which require interactions with cellular membranes. The bilayer association of the peptides has been shown to be strongly pH-dependent, with in-planar alignments under acidic conditions and transmembrane orientations when the histidines are discharged. Therefore, we investigated the pH- and temperature-dependent conformations of LAH4 in DPC micellar solutions and in a TFE/PBS solvent mixture. In the presence of detergent and at pH 4.1, LAH4 adopts helical conformations between residues 9 and 24 concomitantly with a high hydrophobic moment. At pH 6.1, a helix-loop-helix structure forms with a hinge encompassing residues His10-Ala13. The data suggest that the high density of histidine residues and the resulting electrostatic repulsion lead to both a decrease in the pK values of the histidines and a less stable α-helical conformation of this region. The hinged structure at pH 6.1 facilitates membrane anchoring and insertion. At pH 7.8, the histidines are uncharged and an extended helical conformation including residues 4-21 is again obtained. LAH4 thus exhibits a high degree of conformational plasticity. The structures provide a stroboscopic view of the conformational changes that occur during membrane insertion, and are discussed in the context of antimicrobial activity and DNA transfection.  相似文献   

8.
BACKGROUND: Mitochondrial processing peptidase (MPP) is a metalloendopeptidase that cleaves the N-terminal signal sequences of nuclear-encoded proteins targeted for transport from the cytosol to the mitochondria. Mitochondrial signal sequences vary in length and sequence, but each is cleaved at a single specific site by MPP. The cleavage sites typically contain an arginine at position -2 (in the N-terminal portion) from the scissile peptide bond in addition to other distal basic residues, and an aromatic residue at position +1. Mitochondrial import machinery recognizes amphiphilic helical conformations in signal sequences. However, it is unclear how MPP specifically recognizes diverse presequence substrates. RESULTS: The crystal structures of recombinant yeast MPP and a cleavage-deficient mutant of MPP complexed with synthetic signal peptides have been determined. MPP is a heterodimer; its alpha and beta subunits are homologous to the core II and core I proteins, respectively, of the ubiquinol-cytochrome c oxidoreductase complex. Crystal structures of two different synthetic substrate peptides cocrystallized with the mutant MPP each show the peptide bound in an extended conformation at the active site. Recognition sites for the arginine at position -2 and the +1 aromatic residue are observed. CONCLUSIONS: MPP bound two mitochondrial import presequence peptides in extended conformations in a large polar cavity. The presequence conformations differ from the amphiphilic helical conformation recognized by mitochondrial import components. Our findings suggest that the presequences adopt context-dependent conformations through mitochondrial import and processing, helical for recognition by mitochondrial import machinery and extended for cleavage by the main processing component.  相似文献   

9.
Models for the 3(10)-helix/coil and pi-helix/coil equilibria have been derived. The theory is based on classifying residues into helical or nonhelical (coil) conformations. Statistical weights are assigned to residues in a helical conformation with an associated helical hydrogen bond, a helical conformation with no hydrogen bond, an N-cap position, a C-cap position, or the reference coil conformation. The models for alpha-helix formation and 3(10)-helix formation have also been combined to describe a three-state equilibrium in which alpha-helical, 3(10)-helical, and coil conformations are populated. The results are compared with the modified Lifson-Roig theory for the alpha-helix/coil equilibrium. The comparison accounts for the experimental observations that 3(10)-helices tend to be short and pi-helices are not favored for any length. This work may provide a framework for quantitatively rationalizing experimental work on isolated 3(10)-helices and mixed 3(10)-/alpha-helices.  相似文献   

10.
We report the solution structure of the chemotactic cytokine (chemokine) vMIP-II. This protein has unique biological activities in that it blocks infection by several different human immunodeficiency virus type 1 (HIV-1) strains. This occurs because vMIP-II binds to a wide range of chemokine receptors, some of which are used by HJV to gain cell entry. vMIP-II is a monomeric protein, unlike most members of the chemokine family, and its structure consists of a disordered N-terminus, followed by a helical turn (Gln25-Leu27), which leads into the first strand of a three-stranded antiparallel beta-sheet (Ser29-Thr34; Gly42-Thr47; Gln52-Asp56). Following the sheet is a C-terminal alpha-helix, which extends from residue Asp60 until Gln68. The final five residues beyond the C-terminal helix (Pro70-Arg74) are in an extended conformation, but several of these C-terminal residues contact the first beta-strand. The structure of vMIP-II is compared to other chemokines that also block infection by HIV-1, and the structural basis of its lack of ability to form a dimer is discussed.  相似文献   

11.
Yuko Okamoto 《Biopolymers》1994,34(4):529-539
Monte Carlo simulated annealing is applied to the tertiary structure prediction of a 17-residue synthetic peptide, which is known by experiment to exhibit high helical content at low pH. Two dielectric models are considered: sigmoidal distance-dependent dielectric function and a constant dielectric function (? = 2). Starting from completely random initial conformations, our simulations for both dielectric models at low pH gave many helical conformations. The obtained low-energy conformations are compared with the nuclear Overhauser effect spectroscopy cross-peak data for both main chain and side chains, and it is shown that the results for the sigmoidal dielectric function are in remarkable agreement with the experimental data. The results predict the existence of two disjoint helices around residues 5–9 and 11–16, while nmr experiments imply significant α-helix content between residues 5 and 14. Simulations with high pH, on the other hand, hardly gave a helical conformation, which is also in accord with the experiment. These findings indicate that when side chains are charged, electrostatic interactions due to these charges play a major role in the helix stability. Our results are compared with the previous 500 ps molecular dynamics simulations of the same peptide. It is argued that simulated annealing is superior to molecular dynamics in two respects: (1) direct folding of α-helix from completely random initial conformations is possible for the former, whereas only unfolding of an α-helix can be studied by the latter; (2) while both methods predict high helix content for low pH, the results for high pH agree with experiment (low helix content) only for the former method. © 1994 John Wiley & Sons, Inc.  相似文献   

12.
Cathelicidins comprise a major family of host-defense antimicrobial peptides in vertebrates. The C-terminal part of the cathelicidins is bestowed with antimicrobial and lipopolysaccharide (LPS) neutralizing activities. In this work, we repot high resolution solution structures of two nontoxic active fragments, residues 1-16 or RG16 and residues 8-26 or LK19, of fowlicidin-1, a cathelicidin family of peptide from chicken, as a complex with LPS using two-dimensional transferred nuclear Overhauser effect (Tr-NOE) spectroscopy. Both peptides are highly flexible and do not assume any preferred conformations in their free states. Upon complexation with endotoxin or LPS, peptides undergo structural transitions towards folded conformations. Structure calculations reveal that the LK19 peptide adopts a well defined helical structure with a bend at the middle. By contrast, the first seven amino acids of RG16 are found to be flexible followed by a helical conformation for the residues L8-A15. In addition, a truncated version of LK19 encompassing residues A15-K26 or AK12 displays an amphipathic helical structure in LPS. Saturation transfer difference (STD) NMR studies demonstrate that all peptides, RG16, LK19, and AK12, are in close proximity with LPS, whereby the aromatic residues showed the strongest STD effects. Fluorescence studies with fluorescein isothiocyanate (FITC) labeled LPS in the presence of full-length fowlicidin-1, LK19, RG16, and AK12 indicated that LPS-neutralization property of these peptides may result from plausible dissociation of LPS aggregates. The helical structures of peptide fragments derived from fowlicidin-1 in LPS could be utilized to develop nontoxic antiendotoxic compounds.  相似文献   

13.
Conformational energy calculations have been used to explore the conformations which may be realized for the sugar moiety of murein and pseudomurein. For the building blocks of the pseudomurein sugar strands, i.e. for the monosaccharides beta-D-N-acetylglucosamine (NAG) and alpha-L-N-acetyltalosaminuronic acid (NAT), both in C1 ring conformation, as well as for their 1,3 and 1,4 linked disaccharides, the favoured conformations were obtained. The helical parameters of sugar strands of both linkage types, which describe the regular structure of the corresponding polysaccharides, poly-(1,3-NAT-NAG) and poly-(1,4-NAT-NAG), were calculated. Both types of polysaccharides poly-(NAG-NAT) considered in this study favoured extended conformations, which in the case of 1,3 linked polymers showed less gain of length per saccharide unit compared to 1,4 linked poly-(NAG-NAT) residues. For a 1,3 linked sugar moiety of pseudomurein every pair of neighbouring peptides attached to glycan chain pointed in favoured conformations approximately to opposite sides of the strands, whereas in a 1,4 linked poly-(NAG-NAT) the peptides protruded approximately to the same side of the glycan moiety. A comparison between pseudomurein and murein revealed that the sugar moieties of both peptidoglycans have similar features in respect to their overall structure, i.e. both favoured more or less extended structures. In contrast to these data the shapes of the resulting polysaccharide moieties were remarkably different. In poly-(1,3-NAG-NAT) the glycan chains possessed a zig-zag-like arrangement, whereas for glycan chains of the murein type relatively flat structures were preferred. These remaining differences in the conformational arrangement between both peptidoglycans depend strongly on the C1 chair conformation of NAT. It is, therefore, attractive to speculate about an hypothetical pseudomurein sugar chain configuration comprising beta-L-N-acetyltalosaminuronic acid in its 1C conformation.  相似文献   

14.
Grace CR  Cervini L  Gulyas J  Rivier J  Riek R 《Biopolymers》2007,87(2-3):196-205
The C-terminally amidated CRF antagonist astressin binds to CRF-R1 or CRF-R2 receptors with low nanomolar affinity while the corresponding astressin-acid has >100 times less affinity. To understand the role of the amide group in binding, the conformations of astressin-amide and astressin-acid were studied in DMSO using NMR techniques. The 3D NMR structures show that the backbones of both analogs prefer an alpha-helical conformation, with a small kink around Gln(26). However, astressin-amide has a well-defined helical structure from Leu(27) to Ile(41) and a conformation very similar to the bioactive conformation reported by our group (Grace et al., Proc Natl Acad Sci USA 2007, 104, 4858-4863). In contrast, astressin-acid has an irregular helical conformation from Arg(35) onward, including a rearrangement of the side chains in that region. This structural difference highlights the crucial role of the C-terminal amidation for stabilization of astressin's bioactive conformation.  相似文献   

15.
H Mizuno 《Proteins》1989,5(1):47-65
The active conformation of an ice-nucleation protein, whose major portion consists of a long polypeptide segment of nearly repetitive octapeptides, is predicted by the analyses of conformational energy and the mechanism of crystal growth. The protein ideally has an exact octapeptide repetition and is assumed to have a helical conformation. The present study searched for low-energy helical conformations and each of the obtained low-energy conformations examined as to whether it has a surface structure that can promote crystal formation. Two conformations obtained were good candidates for an ice nucleus. Both were found to have on their surfaces an arrangement of hydrogen-bonding sites, which fits well with those of hydrogen bonds in hexagonal ice crystal. Further, one of the two conformations had a hexagonal conformational symmetry consistent with the hexagonal ice crystal structure. The other conformation had a pentagonal conformational symmetry that could enable the growth of an ice crystal--dendritic polycrystalline snow crystal--which grows on metastable cubic ice.  相似文献   

16.
The conformation of pituitary adenylate cyclase activating polypeptide with 27 residues (PACAP27) has been determined by two-dimensional NMR and CD spectroscopies and distance geometry in 25% methanol. Residues 9-20 and 22-25 have well-defined conformations but other residues do not show ordered conformations. The conformation of residues 9-20 is composed of three distinct regions of beta turn-like conformation (residues 9-12), alpha helix (residues 12-14) and the looser helical conformation (residues 15-20), while residues 22-24 form alpha helix. PACAP27 has a 2 helices separated by a disordered region similar to a VIP analog reported by Fry et al. but is distinct from the VIP analog in the position of the first helix, which is shifted by 2 residues toward the C-terminus, and in the form of the second helix [Fry, D.C., Madison, V.S., Bolin, D.R., Greeley, D.N., Toome, V. and Wegrzynski, B.B. (1989) Biochemistry 28, 2399-2409].  相似文献   

17.
Synthetic peptides with defined secondary structure scaffolds, namely hairpins and helices, containing tryptophan residues, have been investigated in this study to probe the influence of a large number of aromatic amino acids on backbone conformations. Solution NMR investigations of Boc-W-L-W-(D)P-G-W-L-W-OMe (peptide 1), designed to form a well-folded hairpin, clearly indicates the influence of flanking aromatic residues at the (D)Pro-Gly region on both turn nucleation and strand propagation. Indole-pyrrolidine interactions in this peptide lead to the formation of the less-frequent type I' turn at the (D)Pro-Gly segment and frayed strand regions, with the strand residues adopting local helical conformations. An analog of peptide 1 with an Aib-Gly turn-nucleated hairpin (Boc-W-L-W-U-G-W-L-W-OMe (peptide 2)) shows a preference for helical structures in solution, in both chloroform and methanol. Peptides with either one (Boc-W-L-W-U-W-L-W-OMe (peptide 3)) or two (Boc-U-W-L-W-U-W-L-W-OMe (peptide 4)) helix-nucleating Aib residues give rise to the well-folded helical conformations in the chloroform solution. The results are indicative of a preference for helical folding in peptides containing a large number of Trp residues. Investigation of a tetrapeptide analog of peptide 2, Boc-W-U-G-W-OMe (peptide 5), in solution and in the crystal state (by X-ray diffraction), also indicates a preference for a helical fold. Additionally, peptide 5 is stabilized in crystals by both aromatic interactions and an array of weak interactions. Examination of Trp-rich sequences in protein structures, however, reveals no secondary structure preference, suggesting that other stabilizing interactions in a well-folded protein may offset the influence of indole rings on backbone conformations.  相似文献   

18.
Gorshkova IN  Liu T  Zannis VI  Atkinson D 《Biochemistry》2002,41(33):10529-10539
To probe the structure and stability of the central region of lipid-free apolipoprotein (apo) A-I (residues 123-165), we studied the effects of four mutations made in this region on the conformation, stability, dimyristoylphosphatidylcholine (DMPC) binding kinetics, and size of discoidal reconstituted high-density lipoprotein (rHDL) particles. The apoA-I deletion delta(144-165) leads to a red shift in the wavelength of maximum fluorescence and a reduction in the alpha-helical content, the stability, the initial rate of association with DMPC liposomes, and the size of the discoidal particles. The data are consistent with the helical structure of residues 144-165, and the deletion appears to perturb the tertiary organization of the N-terminal half of apoA-I. In contrast, the deletion of the adjacent region, delta(136-143), leads to stabilization without altering the number of residues in the helical conformation or the initial rate of association with DMPC liposomes. The quadruple substitution E125K/E128K/K133E/E139K leads to approximately 17 additional residues in the helical conformation and an increase in the stability, the initial rate of association with DMPC liposomes, and the size of the rHDL particles. The findings are consistent with the disordered structure of the segment of residues 123-142, which becomes helical as a result of the quadruple mutation or upon lipid binding. The naturally occurring mutation L141R (also associated with coronary heart disease) that is located in this segment does not change the protein conformation but leads to a reduced stability and a decreased rate of association with DMPC liposomes that may relate to the observed altered functions of this mutant.  相似文献   

19.
Diethylglycine (Deg) residues incorporated into peptides can stabilize fully extended (C5) or helical conformations. The conformations of three tetrapeptides Boc-Xxx-Deg-Xxx-Deg-OMe (Xxx=Gly, GD4; Leu, LD4 and Pro, PD4) have been investigated by NMR. In the Gly and Leu peptides, NOE data suggest that the local conformations at the Deg residues are fully extended. Low temperature coefficients for the Deg(2) and Deg(4) NH groups are consistent with their inaccessibility to solvent, in a C5 conformation. NMR evidence supports a folded beta-turn conformation involving Deg(2)-Gly(3), stabilized by a 4-->1 intramolecular hydrogen bond between Pro(1) CO and Deg(4) NH in the proline containing peptide (PD4). The crystal structure of GD4 reveals a hydrated multiple turn conformation with Gly(1)-Deg(2) adopting a distorted type II/II' conformation, while the Deg(2)-Pro(3) segment adopts a type III/III' structure. A lone water molecule is inserted into the potential 4-->1 hydrogen bond of the Gly(1)-Deg(2) beta-turn.  相似文献   

20.
The active site loop of triosephosphate isomerase (TIM) exhibits a hinged-lid motion, alternating between the two well defined "open" and "closed" conformations. Until now the closed conformation had only been observed in protein complexes with substrate analogues. Here, we present the first rabbit muscle apo TIM structure, refined to 1.5A resolution, in which the active site loop is either in the open or in the closed conformation in different subunits of the enzyme. In the closed conformation described here, the lid loop residues participate in stabilizing hydrogen bonds characteristic of holo TIM structures, whereas chemical interactions observed in the open loop conformation are similar to those found in the apo structures of TIM. In the closed conformation, a number of water molecules are observed at the projected ligand atom positions that are hydrogen bonded to the active site residues. Additives used during crystallization (DMSO and Tris molecules and magnesium atoms) were modeled in the electron density maps. However, no specific binding of these molecules is observed at, or close to, the active site and the lid loop. To further investigate this unusual closed conformation of the apo enzyme, two more rabbit muscle TIM structures, one in the same and another in a different crystal form, were determined. These structures present the open lid conformation only, indicating that the closed conformation cannot be explained by crystal contact effects. To rationalize why the active site loop is closed in the absence of ligand in one of the subunits, extensive comparison with previously solved TIM structures was carried out, supported by the bulk of available experimental information about enzyme kinetics and reaction mechanism of TIM. The observation of both open and closed lid conformations in TIM crystals might be related to a persistent conformational heterogeneity of this protein in solution.  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