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.
Structural and biochemical studies of Cys(2)His(2) zinc finger proteins initially led several groups to propose a "recognition code" involving a simple set of rules relating key amino acid residues in the zinc finger protein to bases in its DNA site. One recent study from our group, involving geometric analysis of protein-DNA interactions, has discussed limitations of this idea and has shown how the spatial relationship between the polypeptide backbone and the DNA helps to determine what contacts are possible at any given position in a protein-DNA complex. Here we report a study of a zinc finger variant that highlights yet another source of complexity inherent in protein-DNA recognition. In particular, we find that mutations can cause key side-chains to rearrange at the protein-DNA interface without fundamental changes in the spatial relationship between the polypeptide backbone and the DNA. This is clear from a simple analysis of the binding site preferences and co-crystal structures for the Asp20-->Ala point mutant of Zif268. This point mutation in finger one changes the specificity of the protein from GCG TGG GCG to GCG TGG GC(G/T), and we have solved crystal structures of the D20A mutant bound to both types of sites. The structure of the D20A mutant bound to the GCG site reveals that contacts from key residues in the recognition helix are coupled in complex ways. The structure of the complex with the GCT site also shows an important new water molecule at the protein-DNA interface. These side-chain/side-chain interactions, and resultant changes in hydration at the interface, affect binding specificity in ways that cannot be predicted either from a simple recognition code or from analysis of spatial relationships at the protein-DNA interface. Accurate computer modeling of protein-DNA interfaces remains a challenging problem and will require systematic strategies for modeling side-chain rearrangements and change in hydration.  相似文献   

2.
De novo protein design provides a tool for testing the principles that stabilize the structures of proteins. Recently, we described the design and structure determination of alpha(3)D, a three-helix bundle protein with a well-packed hydrophobic core. Here, we test the malleability and adaptability of this protein's structure by mutating a small, Ala residue (A60) in its core to larger, hydrophobic side-chains, Leu and Ile. Such changes introduce strain into the structures of natural proteins, and therefore generally destabilize the native state. By contrast, these mutations were slightly stabilizing ( approximately 1.5 kcal mol(-1)) to the tertiary structure of alpha(3)D. The value of DeltaC(p) for unfolding of these mutants was not greatly affected relative to wild-type, indicating that the change in solvent accessibility for unfolding was similar. However, two-dimensional heteronuclear single quantum coherence spectra indicate that the protein adjusts to the introduction of steric bulk in different ways. A60L-alpha(3)D showed serious erosion in the dispersion of both the amide backbone as well as the side-chain methyl chemical shifts. By contrast, A60I-alpha(3)D showed excellent dispersion of the backbone resonances, and selective changes in dispersion of the aliphatic side-chains proximal to the site of mutation. Together, these data suggest that alpha(3)D, although folded into a unique three-dimensional structure, is nevertheless more malleable and flexible than most natural, native proteins.  相似文献   

3.
How is the native structure encoded in the amino acid sequence? For the traditional backbone centric view, the dominant forces are hydrogen bonds (backbone) and phi-psi propensity. The role of hydrophobicity is non-specific. For the side-chain centric view, the dominant force of protein folding is hydrophobicity. In order to understand the balance between backbone and side-chain forces, we have studied the contributions of three components of a beta-hairpin peptide: turn, backbone hydrogen bonding and side-chain interactions, of a 16-residue fragment of protein G. The peptide folds rapidly and cooperatively to a conformation with a defined secondary structure and a packed hydrophobic cluster of aromatic side-chains. Our strategy is to observe the structural stability of the beta-hairpin under systematic perturbations of the turn region, backbone hydrogen bonds and the hydrophobic core formed by the side-chains, respectively. In our molecular dynamics simulations, the peptides are solvated. with explicit water molecules, and an all-atom force field (CFF91) is used. Starting from the original peptide (G41EWTYDDATKTFTVTE56), we carried out the following MD simulations. (1) unfolding at 350 K; (2) forcing the distance between the C(alpha) atoms of ASP47 and LYS50 to be 8 A; (3) deleting two turn residues (Ala48 and Thr49) to form a beta-sheet complex of two short peptides, GEWTYDD and KTFTVTE; (4) four hydrophobic residues (W43, Y45, F52 and T53) are replaced by a glycine residue step-by-step; and (5) most importantly, four amide hydrogen atoms (T44, D46, T53, and T55, which are crucial for backbone hydrogen bonding), are substituted by fluorine atoms. The fluorination not only makes it impossible to form attractive hydrogen bonding between the two beta-hairpin strands, but also introduces a repulsive force between the two strands due to the negative charges on the fluorine and oxygen atoms. Throughout all simulations, we observe that backbone hydrogen bonds are very sensitive to the perturbations and are easily broken. In contrast, the hydrophobic core survives most perturbations. In the decisive test of fluorination, the fluorinated peptide remains folded under our simulation conditions (5 ns, 278 K). Hydrophobic interactions keep the peptide folded, even with a repulsive force between the beta-strands. Thus, our results strongly support a side-chain centric view for protein folding.  相似文献   

4.
X-linked lymphoproliferative disease is caused by mutations in the protein SAP, which consists almost entirely of a single SH2 domain. SAP interacts with the Tyr281 site of the T<-->B cell signaling protein SLAM via its SH2 domain. Interestingly, binding is not dependent on phosphorylation but does involve interactions with residues N-terminal to the Tyr. We have used 15N and 2H NMR relaxation experiments to investigate the motional properties of the SAP SH2 domain backbone amides and side-chain methyl groups in the free protein and complexes with phosphorylated and non-phosphorylated peptides derived from the Tyr281 site of SLAM. The most mobile methyl groups are in side-chains with large RMSD values between the three crystal structures of SAP, suggesting that fast time-scale dynamics in side-chains is associated with conformational plasticity. The backbone amides of two residues which interact with the C-terminal part of the peptides experience fast time-scale motions in the free SH2 domain that are quenched upon binding of either the phosphorylated or non-phosphorylated peptide. Of most importance, the mobility of methyl groups in and around the binding site for residues in the N-terminus of the peptide is significantly restricted in the complexes, underscoring the dominance of this interaction with SAP and demonstrating a correlation between changes in rapid side-chain motion upon binding with local binding energy.  相似文献   

5.
DigA16 is an artificial digoxigenin-binding protein, which was derived from the bilin-binding protein, a lipocalin of Pieris brassicae, via reshaping of its natural ligand pocket. Here we report the crystal structures of DigA16 in the presence of either digoxigenin or digitoxigenin and for the apo-protein at resolutions below 1.9A. As a consequence of the altogether 17 amino acid substitutions within the binding site significant structural changes have occurred in the four loops that form the entrance to the ligand pocket on top of the structurally conserved beta-barrel framework. For example, one loop adopts a new alpha-helical backbone structure, which seems to be induced by few critical side-chain contacts. Digoxigenin becomes almost fully buried (by 95%) upon complexation, whereby specificity for the hydrophilic steroid is maintained through hydrogen-bonding networks and shape complementarity. The differential binding of the related steroid digitoxigenin is mainly governed by an internal histidine residue, whose side-chain undergoes significant induced fit. Among those amino acids that line the ligand pocket two tyrosine and one tryptophan residue provide the largest contacts. Interestingly, corresponding three side-chains are found with the same mutual orientation in the anti-digoxigenin antibody 26-10, even though the hapten orientation is quite different there and only 66% of the steroid surface is buried in the combining site. Hence, in the case of the engineered lipocalin DigA16 an example of convergent in vitro evolution is observed. Generally, the remarkable structural plasticity of the loop region and the role of polar residues in the binding site illustrate the potential of the lipocalin scaffold for the generation of specific receptor proteins towards a variety of ligands.  相似文献   

6.
The three-dimensional solution structure of reduced (CuI) plastocyanin from French bean leaves has been determined by distance geometry and restrained molecular dynamics methods using constraints obtained from 1H n.m.r. (nuclear magnetic resonance) spectroscopy. A total of 1244 experimental constraints were used, including 1120 distance constraints, 103 dihedral angle constraints and 21 hydrogen bond constraints. Stereospecific assignments were made for 26 methylene groups and the methyls of 11 valines. Additional constraints on copper co-ordination were included in the restrained dynamics calculations. The structures are well defined with average atomic root-mean-square deviations from the mean of 0.45 A for all backbone heavy atoms and 1.08 A for side-chain heavy atoms. French bean plastocyanin adopts a beta-sandwich structure in solution that is similar to the X-ray structure of reduced poplar plastocyanin; the average atomic root-mean-square difference between 16 n.m.r. structures and the X-ray structure is 0.76 A for all backbone heavy atoms. The conformations of the side-chains that constitute the hydrophobic core of French bean plastocyanin are very well defined. Of 47 conserved residues that populate a single chi 1 angle in solution, 43 have the same rotamer in the X-ray structure. Many surface side-chains adopt highly preferred conformations in solution, although the 3J alpha beta coupling constants often indicate some degree of conformational averaging. Some surface side-chains are disordered in both the solution and crystal structures of plastocyanin. There is a striking correlation between measures of side-chain disorder in solution and side-chain temperature factors in the X-ray structure. Side-chains that form a distinctive acidic surface region, believed to be important in binding other electron transfer proteins, appear to be disordered. Fifty backbone amide protons form hydrogen bonds to carbonyls in more than 60% of the n.m.r. structures; 45 of these amide protons exchange slowly with solvent deuterons. Ten hydrogen bonds are formed between side-chain and backbone atoms, eight of which are correlated with decreased proton exchange. Of the 60 hydrogen bonds formed in French bean plastocyanin, 56 occur in the X-ray structure of the poplar protein; two of the missing hydrogen bonds are absent as a result of mutations. It appears that molecular dynamics refinement of highly constrained n.m.r. structures allows accurate prediction of the pattern of hydrogen bonding.  相似文献   

7.
The jigsaw puzzle model postulates that the predominant factor relating primary sequence to three-dimensional fold lies in the stereospecific packing of interdigitating side-chains within densely packed protein interiors. An attempt has been made to check the validity of the model by means of a surface complementarity function. Out of a database of 100 highly resolved protein structures the contacts between buried hydrophobic residues (Leu, Ile, Val, Phe) and their neighbours have been categorized in terms of the extent of side-chain surface area involved in a contact (overlap) and their steric fit (Sm). The results show that the majority of contacts between a buried residue and its immediate neighbours (side-chains) are of high steric fit and in the case of extended overlap at least one of the angular parameters characterizing interresidue geometry to have pronounced deviation from a random distribution, estimated by chi(2). The calculations thus tend to support the "jigsaw puzzle" model in that 75-85% of the contacts involving hydrophobic residues are of high surface complementarity, which, coupled to high overlap, exercise fairly stringent constraints over the possible geometrical orientations between interacting residues. These constraints manifest in simple patterns in the distributions of orientational angles. Approximately 60-80% of the buried side-chain surface packs against neighbouring side-chains, the rest interacting with main-chain atoms. The latter partition of the surface maintains an equally high steric fit (relative to side-chain contacts) emphasizing a non-trivial though secondary role played by main-chain atoms in interior packing. The majority of this class of contacts, though of high complementarity, is of reduced overlap. All residues whether hydrophobic or polar/charged show similar surface complementarity measures upon burial, indicating comparable competence of all amino acids in packing effectively with their atomic environments. The specificity thus appears to be distributed over the entire network of contacts within proteins. The study concludes with a proposal to classify contacts as specific and non-specific (based on overlap and fit), with the former perhaps contributing more to the specificity between sequence and fold than the latter.  相似文献   

8.
《Biophysical journal》2020,118(10):2502-2512
Proline-rich motifs (PRMs) are widely used for mediating protein-protein interactions with weak binding affinities. Because they are intrinsically disordered when unbound, conformational entropy plays a significant role for the binding. However, residue-level differences of the entropic contribution in the binding of different ligands remain not well understood. We use all-atom molecular dynamics simulation and the maximal information spanning tree formalism to analyze conformational entropy associated with the binding of two PRMs, one from the Abl kinase and the other from the nonstructural protein 1 of the 1918 Spanish influenza A virus, to the N-terminal SH3 (nSH3) domain of the CrkII protein. Side chains of the stably folded nSH3 experience more entropy change upon ligand binding than the backbone, whereas PRMs involve comparable but heterogeneous entropy changes among the backbone and side chains. In nSH3, two conserved nonpolar residues forming contacts with the PRM experience the largest side-chain entropy loss. In contrast, the C-terminal charged residues of PRMs that form polar contacts with nSH3 experience the greatest side-chain entropy loss, although their “fuzzy” nature is attributable to the backbone that remains relatively flexible. Thus, residues that form high-occupancy contacts between nSH3 and PRM do not reciprocally contribute to entropy loss. Furthermore, certain surface residues of nSH3 distal to the interface with PRMs gain entropy, indicating a nonlocal effect of ligand binding. Comparing between the PRMs from cAbl and nonstructural protein 1, the latter involves a larger side-chain entropy loss and forms more contacts with nSH3. Consistent with experiments, this indicates stronger binding of the viral ligand at the expense of losing the flexibility of side chains, whereas the backbone experiences less entropy loss. The entropy “hotspots” as identified in this study will be important for tuning the binding affinity of various ligands to a receptor.  相似文献   

9.
The structures of the native fructose-1,6-bisphosphatase (Fru-1,6-Pase), from pig kidney cortex, and its fructose 2,6-bisphosphate (Fru-2,6-P2) complexes have been refined to 2.8 A resolution to R-factors of 0.194 and 0.188, respectively. The root-mean-square deviations from the standard geometry are 0.021 A and 0.016 A for the bond length, and 4.4 degrees and 3.8 degrees for the bond angle. Four sites for Fru-2,6-P2 binding per tetramer have been identified by difference Fourier techniques. The Fru-2,6-P2 site has the shape of an oval cave about 10 A deep, and with other dimensions about 18 A by 12 A. The two Fru-2,6-P2 binding caves of the dimer in the crystallographically asymmetric unit sit next to one another and open in opposite directions. These two binding sites mutually exchange their Arg243 side-chains, indicating the potential for communication between the two sites. The beta, D-fructose 2,6-bisphosphate has been built into the density and refined well. The oxygen atoms of the 6-phosphate group of Fru-2,6-P2 interact with Arg243 from the adjacent monomer and the residues of Lys274, Asn212, Tyr264, Tyr215 and Tyr244 in the same monomer. The sugar ring primarily contacts with the backbone atoms from Gly246 to Met248, as well as the side-chain atoms, Asp121, Glu280 and Lys274. The 2-phosphate group interacts with the side-chain atoms of Ser124 and Lys274. A negatively charged pocket near the 2-phosphate group includes Asp118, Asp121 and Glu280, as well as Glu97 and Glu98. The 2-phosphate group showed a disordered binding perhaps because of the disturbance from the negatively charged pocket. In addition, Asn125 and Lys269 are located within a 5 A radius of Fru-2,6-P2. We argue that Fru-2,6-P2 binds to the active site of the enzyme on the basis of the following observations: (1) the structure similarity between Fru-2,6-P2 and the substrate; (2) sequence conservation of the residues directly interacting with Fru-2,6-P2 or located at the negatively charged pocket; (3) a divalent metal site next to the 2-phosphate group of Fru-2,6-P2; and (4) identification of some active site residues in our structure, e.g. tyrosine and Lys274, consistent with the results of the ultraviolet spectra and the chemical modification. The structures are described in detail including interactions of interchain surfaces, and the chemically modifiable residues are discussed on the basis of the refined structures.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

10.
Exposure of proteins to radicals in the presence of O(2) results in side-chain oxidation and backbone fragmentation; the interrelationship between these processes is not fully understood. Recently, initial attack on Ala side-chains was shown to give alpha-carbon radicals (and hence backbone cleavage) and formaldehyde, via the formation and subsequent beta-scission, of C-3 alkoxyl radicals. We now show that this side-chain to backbone damage transfer, is a general mechanism for aliphatic side-chains. Oxidation of Val, Leu, and Asp residues by HO(*)/O(2) results in the release of a family of carbonyls (including formaldehyde, acetone, isobutyraldehyde, and glyoxylic acid) via the formation, and subsequent beta-scission of alkoxyl radicals. The concentration of these products increases with the HO(*) flux. The release of multiple carbonyls confirms the occurrence of oxidation at C-3 and C-4 for Val, and these sites, plus C-5, for Leu. The detection of glyoxylic acid and CO(2)(-*) from Asp demonstrates the occurrence of competing beta-scission processes for the Asp C-3 alkoxyl radical. The yield of hydroperoxides and released carbonyls account for 10-145% of the initial HO(*). The greater than 100% yields confirm the occurrence of chain reactions in peptide/protein oxidation, with more than one residue being damaged per initiating radical.  相似文献   

11.
Packing interactions in bacteriophage T4 lysozyme were explored by determining the structural and thermodynamic effects of substitutions for Ala98 and neighboring residues. Ala98 is buried in the core of T4 lysozyme in the interface between two alpha-helices. The Ala98 to Val (A98V) replacement is a temperature-sensitive lesion that lowers the denaturation temperature of the protein by 15 degrees C (pH 3.0, delta delta G = -4.9 kcal/mol) and causes atoms within the two helices to move apart by up to 0.7 A. Additional structural shifts also occur throughout the C-terminal domain. In an attempt to compensate for the A98V replacement, substitutions were made for Val149 and Thr152, which make contact with residue 98. Site-directed mutagenesis was used to construct the multiple mutants A98V/T152S, A98V/V149C/T152S and the control mutants T152S, V149C and A98V/V149I/T152S. These proteins were crystallized, and their high-resolution X-ray crystal structures were determined. None of the second-site substitutions completely alleviates the destabilization or the structural changes caused by A98V. The changes in stability caused by the different mutations are not additive, reflecting both direct interactions between the sites and structural differences among the mutants. As an example, when Thr152 in wild-type lysozyme is replaced with serine, the protein is destabilized by 2.6 kcal/mol. Except for a small movement of Val94 toward the cavity created by removal of the methyl group, the structure of the T152S mutant is very similar to wild-type T4 lysozyme. In contrast, the same Thr152 to Ser replacement in the A98V background causes almost no change in stability. Although the structure of A98V/T152S remains similar to A98V, the combination of T152S with A98V allows relaxation of some of the strain introduced by the Ala98 to Val replacement. These studies show that removal of methyl groups by mutation can be stabilizing (Val98----Ala), neutral (Thr152----Ser in A98V) or destabilizing (Val149----Cys, Thr152----Ser). Such diverse thermodynamic effects are not accounted for by changes in buried surface area or free energies of transfer of wild-type and mutant side-chains. In general, the changes in protein stability caused by a mutation depend not only on changes in the free energy of transfer associated with the substitution, but also on the structural context within which the mutation occurs and on the ability of the surrounding structure to relax in response to the substitution.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

12.
It is widely believed that the dominant force opposing protein folding is the entropic cost of restricting internal rotations. The energetic changes from restricting side-chain torsional motion are more complex than simply a loss of conformational entropy, however. A second force opposing protein folding arises when a side-chain in the folded state is not in its lowest-energy rotamer, giving rotameric strain. chi strain energy results from a dihedral angle being shifted from the most stable conformation of a rotamer when a protein folds. We calculated the energy of a side-chain as a function of its dihedral angles in a poly(Ala) helix. Using these energy profiles, we quantify conformational entropy, rotameric strain energy and chi strain energy for all 17 amino acid residues with side-chains in alpha-helices. We can calculate these terms for any amino acid in a helix interior in a protein, as a function of its side-chain dihedral angles, and have implemented this algorithm on a web page. The mean change in rotameric strain energy on folding is 0.42 kcal mol-1 per residue and the mean chi strain energy is 0.64 kcal mol-1 per residue. Loss of conformational entropy opposes folding by a mean of 1.1 kcal mol-1 per residue, and the mean total force opposing restricting a side-chain into a helix is 2.2 kcal mol-1. Conformational entropy estimates alone therefore greatly underestimate the forces opposing protein folding. The introduction of strain when a protein folds should not be neglected when attempting to quantify the balance of forces affecting protein stability. Consideration of rotameric strain energy may help the use of rotamer libraries in protein design and rationalise the effects of mutations where side-chain conformations change.  相似文献   

13.
The importance of amino acid side-chains in helix stability has been investigated by making a series of mutations at the N-caps, C-caps and internal positions of the solvent-exposed faces of the two alpha-helices of barnase. There is a strong positional and context dependence of the effect of a particular amino acid on stability. Correlations have been found that provide insight into the physical basis of helix stabilization. The relative effects of Ala and Gly (or Ser) may be rationalized on the basis of solvent-accessible surface areas: burial of hydrophobic surface stabilizes the protein as does exposure to solvent of unpaired hydrogen bond donors or acceptors in the protein. There is a good correlation between the relative stabilizing effects of Ala and Gly at internal positions with the total change in solvent-accessible hydrophobic surface area of the folded protein on mutation of Ala----Gly. The relationship may be extended to the N and C-caps by including an extra term in hydrophilic surface area for the solvent exposure of the non-intramolecularly hydrogen-bonded main-chain CO, NH or protein side-chain hydrogen bonding groups. The requirement for solvent exposure of the C-cap main-chain CO groups may account for the strong preference for residues having positive phi and psi angles at this position, since this alpha L-conformation results in the largest solvent exposure of the C-terminal CO groups. Glycine in an alpha L-conformation results in the greatest exposure of these CO groups. Further, the side-chains of His, Asn, Arg and Lys may, with positive phi and psi-angles, form a hydrogen bond with the backbone CO of residue in position C -3 (residues are numbered relative to the C-cap). The preferences at the C-cap are Gly much greater than His greater than Asn greater than Arg greater than Lys greater than Ala approximately Ser approximately greater than Asp. The preferences at the N-cap are determined by hydrogen bonding of side-chains or solvent to the exposed backbone NH groups and are: Thr approximately Asp approximately Ser greater than Gly approximately Asn greater than Gln approximately Glu approximately His greater than Ala greater than Val much greater than Pro. These general trends may be obscured when mutation allows another side-chain to become a surrogate cap.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

14.
Automated protein redesign, as implemented in the program ORBIT, was used to redesign the core of phage T4 lysozyme. A total of 26 buried or partially buried sites in the C-terminal domain were allowed to vary both their sequence and side-chain conformation while the backbone and non-selected side-chains remained fixed. A variant with seven substitutions ("Core-7") was identified as having the most favorable energy. The redesign experiment was repeated with a penalty for the presence of methionine residues. In this case the redesigned protein ("Core-10") had ten amino acid changes. The two designed proteins, as well as the constituent single mutants, and several single-site revertants were over-expressed in Escherichia coli, purified, and subjected to crystallographic and thermal analyses. The thermodynamic and structural data show that some repacking was achieved although neither redesigned protein was more stable than the wild-type protein. The use of the methionine penalty was shown to be effective. Several of the side-chain rotamers in the predicted structure of Core-10 differ from those observed. Rather than changing to new rotamers predicted by the design process, side-chains tend to maintain conformations similar to those seen in the native molecule. In contrast, parts of the backbone change by up to 2.8A relative to both the designed structure and wild-type.Water molecules that are present within the lysozyme molecule were removed during the design process. In the redesigned protein the resultant cavities were, to some degree, re-occupied by side-chain atoms. In the observed structure, however, water molecules were still bound at or near their original sites. This suggests that it may be preferable to leave such water molecules in place during the design procedure. The results emphasize the specificity of the packing that occurs within the core of a typical protein. While point substitutions within the core are tolerated they almost always result in a loss of stability. Likewise, combinations of substitutions may also be tolerated but usually destabilize the protein. Experience with T4 lysozyme suggests that a general core repacking methodology with retention or enhancement of stability may be difficult to achieve without provision for shifts in the backbone.  相似文献   

15.
The mechanisms of the three-dimensional crambin structure alterations in the crystalline environments and in the trajectories of the molecular dynamics simulations in the vacuum and crystal surroundings have been analyzed. In the crystalline state and in the solution the partial regrouping of remote intramolecular packing contacts, involved in the formation and stabilization of the tertiary structure of the crambin molecule, occurs in NMR structures. In the crystalline state it is initiated by the formation of the intermolecular contacts, the conformational influence of its appearance is distributed over the structure. The changes of the conformations and positions of the residues of the loop segments, where the intermolecular contacts of the crystal surroundings are preferably concentrated, are most observable. Under the influence of these contacts the principal change of the regular secondary structure of crambin is taking place: extension of the two-strand β structure to the three-strand structure with the participation of the single last residue N46 of the C-terminal loop. In comparison with the C-terminal loop the more profound changes are observed in the conformation and the atomic positions of the backbone atoms and in the solvent accessibility of the residues of the interhelical loop. In the solution of the ensemble of the 8 NMR structures relative accessibility to the solvent differs more noticeably also in the region of the loop segments and rather markedly in the interhelical loop. In the crambin cryogenic crystal structures the positions of the atoms of the backbone and/or side chain of 14–18 of 46 residues are discretely disordered. The disorganizations of at least 8 of 14 residues occur directly in the regions of the intermolecular contacts and another 5 residues are disordered indirectly through the intramolecular contacts with the residues of the intermolecular contacts. Upon the molecular dynamics simulation in the vacuum surrounding as in the solution of the crystalline structure of crambin the essential changes of the backbone conformation are caused by the intermolecular contacts absence, but partly masked by the structure changes owing to the nonpolar H atoms absence on the simulated structure. The intermolecular contact absence is partly manifested upon the molecular dynamics simulation of the crambin crystal with one protein molecule. Compared to the crystal structure the lengths of the interpeptide hydrogen bonds and other interresidue contacts in an average solution NMR structure are somewhat shorter and accordingly the energy of the interpeptide hydrogen bonds is better. This length shortening can occur at the stage of the refinement of the NMR structures of the crambin and other proteins by its energy minimizations in the vacuum surroundings and not exist in the solution protein structures.  相似文献   

16.
The reactions of hydroxyl radicals with 30 dipeptides and several larger peptides were studied in aqueous solutions. The OH radicals were generated by U.V. photolysis of H2O2. The short-lived peptide radicals were spin-trapped using t-nitrosobutane and identified by e.s.r. For dipeptides containing the amino terminal residues glycine, alanine and phenylalanine, abstraction of the hydrogen from the carbon adjacent to the peptide nitrogen was the major process leading to the spin-adducts. Such radicals will be referred to as backbone radicals. Dipeptides with a carbonyl terminal serine residue and also glycylglutamic acid form both backbone and side-chain radicals, with the latter being formed in larger quantities. For dipeptides, side-chain radicals were detected on either the carboxyl or amino terminal residues of both. The effect of pD on the e.s.r. sectrum of the spin-adducts of glycylglycine was studied and the pK of the carboxyl group of this radical was determined to be 2.5. For (Ala)3 and (Ala)n, with an average value of n = 1800, backbone and minor side-chain radicals were observed. For ribonucleases-S-peptide, containing 20 amino acid residues, both backbone and side-chain radicals were detected.  相似文献   

17.
The structure of wild-type bacteriophage T4 glutaredoxin (earlier called thioredoxin) in its oxidized form has been refined in a monoclinic crystal form at 2.0 A resolution to a crystallographic R-factor of 0.209. A mutant T4 glutaredoxin gives orthorhombic crystals of better quality. The structure of this mutant has been solved by molecular replacement methods and refined at 1.45 A to an R-value of 0.175. In this mutant glutaredoxin, the active site residues Val15 and Tyr16 have been substituted by Gly and Pro, respectively, to mimic that of Escherichia coli thioredoxin. The main-chain conformation of the wild-type protein is similar in the two independently determined molecules in the asymmetric unit of the monoclinic crystals. On the other hand, side-chain conformations differ considerably between the two molecules due to heterologous packing interactions in the crystals. The structure of the mutant protein is very similar to the wild-type protein, except at mutated positions and at parts involved in crystal contacts. The active site disulfide bridge between Cys14 and Cys17 is located at the first turn of helix alpha 1. The torsion angles of these residues are similar to those of Escherichia coli thioredoxin. The torsion angle around the S-S bond is smaller than that normally observed for disulfides: 58 degrees, 67 degrees and 67 degrees for wild-type glutaredoxin molecule A and B and mutant glutaredoxin, respectively. Each sulfur atom of the disulfide cysteines in T4 glutaredoxin forms a hydrogen bond to one main-chain nitrogen atom. The active site is shielded from solvent on one side by the beta-carbon atoms of the cysteine residues plus side-chains of residues 7, 9, 21 and 33. From the opposite side, there is a cleft where the sulfur atom of Cys14 is accessible and can be attacked by a nucleophilic thiolate ion in the initial step of the reduction reaction.  相似文献   

18.
Previous examination of the accessibility of a panel of single-Cys mutants in transmembrane domain III (TMDIII) of the yeast mitochondrial citrate transport protein to the hydrophilic, cysteine-specific methanethiosulfonate reagent MTSES enabled identification of the water-accessible surface of this TMD. Further studies on the effect of citrate on MTS reagent accessibility, indicated eight sites within TMD III at which citrate conferred temperature-independent protection, thus providing strong evidence for participation of these residues in the formation of a portion of the substrate translocation pathway. Unexpectedly, citrate did not protect against inhibition of the Leu120Cys variant, despite its location on a water- and citrate-accessible surface of the TMDIII helix. This led to the hypothesis that in the 3-dimensional CTP structure, TMDIV packs against TMDIII in a manner such that the Leu120 side-chain folds behind the side-chain of Gln182. The present investigations addressed this hypothesis by examining the properties of the Gln182Cys single mutant and the Leu120Cys/Gln182Ala double mutant. We observed that in contrast to our findings with the Leu120Cys mutant, citrate did protect the Gln182Cys variant against MTSES-mediated inhibition. Importantly, truncation of the Gln182 side-chain to Ala enabled citrate to protect the Leu120Cys double mutant against inhibition. In combination these data support the idea that the Gln182 side-chain lines the transport path and sterically blocks access of citrate to the Leu120 side-chain. In a parallel series of investigations, we constructed 24 single-Cys substitution mutants that were chosen based on their hypothesized importance in substrate binding and/or translocation. We observed that substitution of Cys for residues E34, K37, K83, R87, Y148, D236, K239, T240, R276, and R279 resulted in > or =98% inactivation of CTP function, suggesting an essential structural and/or mechanistic role for these native residues. Superposition of this functional data onto a detailed 3-dimensional homology model of the CTP structure indicates that the side-chains of each of these residues project into the putative transport pathway. We hypothesize that a subset of these residues, in combination with four previously identified essential residues, define the citrate binding site(s) within the CTP.  相似文献   

19.
We have analyzed the structure of the interface between VL and VH domains in three immunoglobulin fragments: Fab KOL, Fab NEW and Fab MCPC 603. About 1800 A2 of protein surface is buried between the domains. Approximately three quarters of this interface is formed by the packing of the VL and VH beta-sheets in the conserved "framework" and one quarter from contacts between the hypervariable regions. The beta-sheets that form the interface have edge strands that are strongly twisted (coiled) by beta-bulges. As a result, the edge strands fold back over their own beta-sheet at two diagonally opposite corners. When the VL and VH domains pack together, residues from these edge strands form the central part of the interface and give what we call a three-layer packing; i.e. there is a third layer composed of side-chains inserted between the two backbone side-chain layers that are usually in contact. This three-layer packing is different from previously described beta-sheet packings. The 12 residues that form the central part of the three observed VL-VH packings are absolutely or very strongly conserved in all immunoglobulin sequences. This strongly suggests that the structure described here is a general model for the association of VL and VH domains and that the three-layer packing plays a central role in forming the antibody combining site.  相似文献   

20.
A high-quality three-dimensional structure of the bovine pancreatic trypsin inhibitor (BPTI) in aqueous solution was determined by 1H nuclear magnetic resonance (n.m.r.) spectroscopy and compared to the three available high-resolution X-ray crystal structures. A newly collected input of 642 distance constraints derived from nuclear Overhauser effects and 115 dihedral angle constraints was used for the structure calculations with the program DIANA, followed by restrained energy minimization with the program AMBER. The BPTI solution structure is represented by a group of 20 conformers with an average root-mean-square deviation (RMSD) relative to the mean solution structure of 0.43 A for backbone atoms and 0.92 A for all heavy atoms of residues 2 to 56. The pairwise RMSD values of the three crystal structures relative to the mean solution structure are 0.76 to 0.85 A for the backbone atoms and 1.24 to 1.33 A for all heavy atoms of residues 2 to 56. Small local differences in backbone atom positions between the solution structure and the X-ray structures near residues 9, 25 to 27, 46 to 48 and 52 to 58, and conformational differences for individual amino acid side-chains were analyzed for possible correlations with intermolecular protein-protein contacts in the crystal lattices, using the pairwise RMSD values among the three crystal structures as a reference.  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