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.
This paper presents the results of a stereochemical analysis of local interactions in unfolded protein chains (sterical repulsions, hydrogen, and hydrophobic bonds, etc.) by means of space-filling modeles. On the basis of this analysis, an evaluation is made of thermodynamic parameters controlling the building-in of all the 20 natural amino acid residues in all the physically possible position of local secondary structures (α-helices, including α-helices with short fragments of helices 310 at the C-terminus; β-bends of different types, helices 310, and their combinations) as well as thermodynamic parameters of separate hydrogen bonds of polar side groups with the neighbor peptide groups (“local contacts”). The accuracy of the obtained results is discussed.  相似文献   

2.
Molecular processes depending on protein–protein interactions can use consensus recognition sequences that possess defined secondary structures. Left-handed polyproline II (PPII) helices are a class of secondary structure commonly involved with cellular signal transduction. However, unlike -helices, for which a substantial body of work exists regarding applications of ring-closing metathesis (RCM), there are few reports on the stabilization of PPII helices by RCM methodologies. The current study examined the effects of RCM macrocyclization on left-handed PPII helices involved with the SH3 domain-mediated binding of Sos1–Grb2. Starting with the Sos1-derived peptide “Ac-V1-P2-P3-P4-V5-P6-P7-R8-R9-R10-amide,” RCM macrocyclizations were conducted using alkenyl chains of varying lengths originating from the pyrrolidine rings of the Pro4 and Pro7 residues. The resulting macrocyclic peptides showed increased helicity as indicated by circular dichroism and enhanced abilities to block Grb2–Sos1 interactions in cell lysate pull-down assays. The synthetic approach may be useful in RCM macrocyclizations, where maintenance of proline integrity at both ring junctures is desired.  相似文献   

3.
A general theory of the molecular structure of receptors for transmitters based only on protein has been presented elsewhere (Smythies, 1974a,b). The acetylcholine receptor at the neuromuscular junction is postulated in particular to be based on a Kusnetsov-Ghokov grid with four sequencestwo “primary” chains A-x-B-cys-A-x-B where A = arg or lys and B = glu or phosphoser and two “secondary” chains of sequence -gly-x-gly-pro-x-ile-cys-asp-x- forming a symmetrical receptor cup of rectangular form. The present paper extends the model to include the gate over the adjacent ionophore (or “ion conductance modulator”: ICM) and the linking mechanism from receptor to gate. These are postulated to consist of a second Kusnetsov-Ghokov grid generated by a third “primary” chain along the side that covers the orifice to the ion conducting channel. The action of ACh is postulated to be to displace an hydrated Ca++ ion from the receptor cup and to disrupt the AB rungs in the receptor grid. The middle primary chain then slides 14 Å and the AB links reform. This replaces a bulky amino acid pair normally blocking the ion channel by a less bulky amino acid pair and so hydrated ions can be transmitted. It is further postulated that snake neurotoxins (ACh blockers) in a specified conformation bind mainly to the ionophore grid and prevent the sliding filament mechanism from opening; whereas the snake “cardiotoxins” (ACh agonists)—in a specified conformation—bind to the same sliding filament mechanism in its “open” ionophore gate and prevent it being closed: and histrionicotoxin binds to the same open “gate” but blocks it physically. The hypothesis may rigorously be tested by experiment as it makes detailed predictions on the X-ray structure of the snake neurotoxins and cardiotoxins.  相似文献   

4.
Abstract

The operator-binding domain of the λ-repressor contains five α-helices and an extended N-terminal arm in the crystal structure determined by Pabo and Lewis reported in Nature 298, 443,1982 (1). The four helices form a “box” enclosing a hydrophobic core, with the fifth helix interacting with the equivalent helix in a dimer. With a small number of well-defined secondary structure elements (microdomains), the repressor is well suited for an analysis of its folding pathways and kinetics by use of the diffusion-collision model. In this paper, the basic elements of the model appropriate to a several microdomain protein are formulated and applied to a set of folding pathways consistent with the crystal structure of the operator- binding domain. The overall kinetics, as well as the time-dependence of intermediate states are determined as a function of the microdomain stability parameter.  相似文献   

5.
Ancestral β-subunit (Anbu) is homologous to HslV and 20S proteasomes. Based on its phylogenetic distribution and sequence clustering, Anbu has been proposed as the “ancestral” form of proteasomes. Here, we report biochemical data, small-angle X-ray scattering results, negative-stain electron microscopy micrographs and a crystal structure of the Anbu particle from Yersinia bercovieri (YbAnbu). All data are consistent with YbAnbu forming defined 12–14 subunit multimers that differ in shape from both HslV and 20S proteasomes. The crystal structure reveals that YbAnbu subunits form tight dimers, held together in part by the Anbu specific C-terminal helices. These dimers (“protomers”) further assemble into a low-rise left-handed staircase. The lock-washer shape of YbAnbu is consistent with the presence of defined multimers, X-ray diffraction data in solution and negative-stain electron microscopy images. The presented structure suggests a possible evolutionary pathway from helical filaments to highly symmetric or pseudosymmetric multimer structures. YbAnbu subunits have the Ntn-hydrolase fold, a putative S1 pocket and conserved candidate catalytic residues Thr1, Asp17 and Lys32(33). Nevertheless, we did not detect any YbAnbu peptidase or amidase activity. However, we could document orthophosphate production from ATP catalyzed by the ATP-grasp protein encoded in the Y. bercovieri Anbu operon.  相似文献   

6.
The Alacoil is an antiparallel (rather than the usual parallel) coiled-coil of α-helices with Ala or another small residue in every seventh position, allowing a very close spacing of the helices (7.5–8.5 Å between local helix axes), often over four or five helical turns. It occurs in two distinct types that differ by which position of the heptad repeat is occupied by Ala and by whether the closest points on the backbone of the two helices are aligned or are offset by half a turn. The aligned, or ROP, type has Ala in position “d” of the heptad repeat, which occupies the “tip-to-tip” side of the helix contact where the Cα–Cβ bonds point toward each other. The more common offset, or ferritin, type of Alacoil has Ala in position “a” of the heptad repeat (where the Cα-Cβ bonds lie back-to-back, on the “knuckle-touch” side of the helix contact), and the backbones of the two helices are offset vertically by half a turn. In both forms, successive layers of contact have the Ala first on one and then on the other helix. The Alacoil structure has much in common with the coiled-coils of fibrous proteins or leucine zippers: both are α-helical coiled-coils, with a critical amino acid repeated every seven residues (the Leu or the Ala) and a secondary contact position in between. However, Leu zippers are between aligned, parallel helices (often identical, in dimers), whereas Alacoils are between antiparallel helices, usually offset, and much closer together. The Alacoil, then, could be considered as an “Ala anti-zipper.” Leu zippers have a classic “knobs-into-holes” packing of the Leu side chain into a diamond of four residues on the opposite helix; for Alacoils, the helices are so close together that the Ala methyl group must choose one side of the diamond and pack inside a triangle of residues on the other helix. We have used the ferritin-type Alacoil as the basis for the de novo design of a 66-residue, coiled helix hairpin called “Alacoilin.” Its sequence is: cmSP DQWDKE A AQYDAHA QE FEKKS HRNng TPEA DQYRHM A SQY QAMA QK LKAIA NQLKK Gseter (with “a” heptad positions underlined and nonhelical parts in lowercase), which we will produce and test for both stability and uniqueness of structure.  相似文献   

7.
We present an analysis of 10 blind predictions prepared for a recent conference, “Critical Assessment of Techniques for Protein Structure Prediction.”1 The sequences of these proteins are not detectably similar to those of any protein in the structure database then available, but we attempted, by a threading method, to recognize similarity to known domain folds. Four of the 10 proteins, as we subsequently learned, do indeed show significant similarity to then-known structures. For 2 of these proteins the predictions were accurate, in the sense that a similar structure was at or near the top of the list of threading scores, and the threading alignment agreed well with the corresponding structural alignment. For the best predicted model mean alignment error relative to the optimal structural alignment was 2.7 residues, arising entirely from small “register shifts” of strands or helices. In the analysis we attempt to identify factors responsible for these successes and failures. Since our threading method does not use gap penalties, we may readily distinguish between errors arising from our prior definition of the “cores” of known structures and errors arising from inherent limitations in the threading potential. It would appear from the results that successful substructure recognition depends most critically on accurate definition of the “fold” of a database protein. This definition must correctly delineate substructures that are, and are not, likely to be conserved during protein evolution. © 1995 Wiley-Liss, Inc.  相似文献   

8.
Examples of homomeric β‐helices and β‐barrels have recently emerged. Here we generalize the theory for the shear number in β‐barrels to encompass β‐helices and homomeric structures. We introduce the concept of the “β‐strip,” the set of parallel or antiparallel neighboring strands, from which the whole helix can be generated giving it n‐fold rotational symmetry. In this context, the shear number is interpreted as the sum around the helix of the fixed register shift between neighboring identical β‐strips. Using this approach, we have derived relationships between helical width, pitch, angle between strand direction and helical axis, mass per length, register shift, and number of strands. The validity and unifying power of the method is demonstrated with known structures including α‐hemolysin, T4 phage spike, cylindrin, and the HET‐s(218‐289) prion. From reported dimensions measured by X‐ray fiber diffraction on amyloid fibrils, the relationships can be used to predict the register shift and the number of strands within amyloid protofilaments. This was used to construct models of transthyretin and Alzheimer β(40) amyloid protofilaments that comprise a single strip of in‐register β‐strands folded into a “β‐strip helix.” Results suggest both stabilization of an individual β‐strip helix and growth by addition of further β‐strip helices can involve the same pair of sequence segments associating with β‐sheet hydrogen bonding at the same register shift. This process would be aided by a repeat sequence. Hence, understanding how the register shift (as the distance between repeat sequences) relates to helical dimensions will be useful for nanotube design.  相似文献   

9.
Amino-acid Sequence Invariance and Genetic Code   总被引:1,自引:0,他引:1  
FROM the complete nucleotide sequence of the gene coding for the coat protein of coliphage MS2, Min Jou et al.1 have suggested a scheme for base-pairing between certain stretches of the polynucleotide chain. Their hypothesis leads to the formation of double helices in their “flower” model and to the consequent implication that nucleic acids have a well-defined tertiary structure. Such a scheme remains to be proved by direct observation of the structure, but the indirect arguments of Min Jou et al. and others appear so convincing and the notion that the genetic material has a definite tertiary structure seems so probable, that there can be little doubt that the “flower” model is on the right lines in principle, even though it may be incorrect in detail.  相似文献   

10.
A protein is defined as an indexed string of elements at each level in the hierarchy of protein structure: sequence, secondary structure, super-secondary structure, etc. The elements, for example, residues or secondary structure segments such as helices or beta-strands, are associated with a series of properties and can be involved in a number of relationships with other elements. Element-by-element dissimilarity matrices are then computed and used in the alignment procedure based on the sequence alignment algorithm of Needleman & Wunsch, expanded by the simulated annealing technique to take into account relationships as well as properties. The utility of this method for exploring the variability of various aspects of protein structure and for comparing distantly related proteins is demonstrated by multiple alignment of serine proteinases, aspartic proteinase lobes and globins.  相似文献   

11.
In prion replication, the cellular form of prion protein (PrPC) must undergo a full conformational transition to its disease-associated fibrillar form. Transmembrane forms of PrP have been implicated in this structural conversion. The cooperative unfolding of a structural core in PrPC presents a substantial energy barrier to prion formation, with membrane insertion and detachment of parts of PrP presenting a plausible route to its reduction. Here, we examined the removal of residues 119–136 of PrP, a region which includes the first β-strand and a substantial portion of the conserved hydrophobic region of PrP, a region which associates with the ER membrane, on the structure, stability and self-association of the folded domain of PrPC. We see an “open” native-like conformer with increased solvent exposure which fibrilises more readily than the native state. These data suggest a stepwise folding transition, which is initiated by the conformational switch to this “open” form of PrPC.  相似文献   

12.
A general theory of the structural changes and fluctuations of proteins has been proposed based on statistical thermodynanic considerations at the chain level.The “structure” of protein was assumed to be characterized by the state of secondary bonds between unique pairs of specific sites on peptide chains. Every secondary bond changes between the bonded and unboned states by thermal agitation and the “structure” is continuously fluctuating. The free energy of the “structural state” that is defined by the fraction of secondary bonds in the bonded state has been expressed by the bond energy, the cooperative interaction between bonds, the mixing entropy of bonds, and the entropy of polypeptide chains. The most probable “structural state” can be simply determined by graphical analysis and the effect of temperature or solvent composition on it is discussed. The temperature dependence of the free energy, the probability distribution of structural states and the specific heat have been calculated for two examples of structural change.The theory predicts two different types of structural changes from the ordered to disordered state, a “structural transition” and a “gradual structural change” with rising temperature, In the “structural transition”, the probability distribution has two maxima in the temperature range of transition. In the “gradual structural change”, the probability distribution has only one maximum during the change.A considerable fraction of secondary bonds is in the unbonded state and is always fluctuating even in the ordered state at room temperature. Such structural fluctuations in a single protein molecule have been discussed quantitatively.The theory is extended to include small molecules which bind to the protein molecule and affect the structural state. The changes of structural state caused by specific and non-specific binding and allosteric effects are explained in a unified manner.  相似文献   

13.
Summary All the backbone 1H and 15N magnetic resonances (except for Pro residues) of the GDP-bound form of a truncated human c-Ha-ras proto-oncogene product (171 amino acid residues, the Ras protein) were assigned by 15N-edited two-dimensional NMR experiments on selectively 15N-labeled Ras proteins in combination with three-dimensional NMR experiments on the uniformly 15N-labeled protein. The sequence-specific assignments were made on the basis of the nuclear Overhauser effect (NOE) connectivities of amide protons with preceding amide and/or Cprotons. In addition to sequential NOEs, vicinal spin coupling constants for amide protons and C protons and deuterium exchange rates of amide protons were used to characterize the secondary structure of the GDP-bound Ras protein; six strands and five helices were identified and the topology of these elements was determined. The secondary structure of the Ras protein in solution was mainly consistent with that in crystal as determined by X-ray analyses. The deuterium exchange rates of amide protons were examined to elucidate the dynamic properties of the secondary structure elements of the Ras protein in solution. In solution, the -sheet structure in the Ras protein is rigid, while the second helix (A66-R73) is much more flexible, and the first and fifth helices (S17-124 and V152-L171) are more rigid than other helices. Secondary structure elements at or near the ends of the effector-region loop were found to be much more flexible in solution than in the crystalline state.  相似文献   

14.
Poly A was found to precipitate upon freezing acidic solutions at pH values where it is normally soluble; this precipitate tends to have the form of small thin plates of irregular outline (“plates”). The X-ray diffraction pattern and solubility properties of the “plates” were compared with those of poly A precipitated solely by exposure to lower pH values, and with fibers drawn from acidic solution. There is considerable molecular order in each of these three types of preparation. In all cases, the diffraction patterns are consistent with the presence of the double-stranded helical structure proposed by Rich, Davies, Crick, and Watson (J. Mol. Biol., 3 , 71 (1961)) based on fiber diffraction data. The diffraction pattern from the “plates” is compared in detail with that of the fibers, and is shown to be in accord with a packing scheme having the chain axis of the molecular structure confined to the plane of the “plate,” but oriented randomly in that plane.  相似文献   

15.
The solid state secondary structure of myoglobin, RNase A, concanavalin A (Con A), poly(L -lysine), and two linear heterooligomeric peptides were examined by both far-uv CD spectroscopy1 and by ir spectroscopy. The proteins associated from water solution on glass and mica surfaces into noncrystalline, amorphous films, as judged by transmission electron microscopy of carbon-platinum replicas of surface and cross-fractured layer. The association into the solid state induced insignificant changes in the amide CD spectra of all α-helical myoglobin, decreased the molar ellipticity of the α/β RNase A, and increased the molar ellipticity of all-β Con A with no change in the positions of the bands' maxima. High-temperature exposure of the films induced permanent changes in the conformation of all proteins, resulting in less α-helix and more β-sheet structure. The results suggest that the protein α-helices are less stable in films and that the secondary structure may rearrange into β-sheets at high temperature. Two heterooligomeric peptides and poly (L -lysine), all in solution at neutral pH with “random coil” conformation, formed films with variable degrees of their secondary structure in β-sheets or β-turns. The result corresponded to the protein-derived Chou-Fasman amino acid propensities, and depended on both temperature and solvent used. The ir and CD spectra correlations of the peptides in the solid state indicate that the CD spectrum of a “random” structure in films differs from random coil in solution. Formic acid treatment transformed the secondary structure of the protein and peptide films into a stable α-helix or β-sheet conformations. The results indicate that the proteins aggregate into a noncrystalline, glass-like state with preserved secondary structure. The solid state secondary structure may undergo further irreversible transformations induced by heat or solvent. © 1993 John Wiley & Sons, Inc.  相似文献   

16.
The crystal structure of the glutamine-binding protein (GlnBP) fromEscherichia coliin a ligand-free “open” conformational state has been determined by isomorphous replacement methods and refined to anR-value of 21.4% at 2.3 Å resolution. There are two molecules in the asymmetric unit, related by pseudo 4-fold screw symmetry. The refined model consists of 3587 non-hydrogen atoms from 440 residues (two monomers), and 159 water molecules. The structure has root-mean-square deviations of 0.013 Å from “deal” bond lengths and 1.5° from “ideal” bond angles.The GlnBP molecule has overall dimensions of approximately 60 Å × 40 Å × 35 Å and is made up of two domains (termed large and small), which exhibit a similar supersecondary structure, linked by two antiparallel β-strands. The small domain contains three α-helices and four parallel and one antiparallel β-strands. The large domain is similar to the small domain but contains two additional α-helices and three more short antiparallel β-strands. A comparison of the secondary structural motifs of GlnBP with those of other periplasmic binding proteins is discussed.A model of the “closed form” GlnBP-Gln complex has been proposed based on the crystal structures of the histidine-binding protein-His complex and “open form” GlnBP. This model has been successfully used as a search model in the crystal structure determination of the “closed form” GlnBP-Gln complex by molecular replacement methods. The model agrees remarkably well with the crystal structure of the Gln-GlnBP complex with root-mean-square deviation of 1.29 Å. Our study shows that, at least in our case, it is possible to predict one conformational state of a periplasmic binding protein from another conformational state of the protein. The glutamine-binding pockets of the model and the crystal structure are compared and the modeling technique is described.  相似文献   

17.
Helices are the most common elements of RNA secondary structure. Despite intensive investigations of various types of RNAs, the evolutionary history of the formation of new helices (novel helical structures) remains largely elusive. Here, by studying the nuclear ribosomal Internal Transcribed Spacer 2 (ITS2), a fast-evolving part of the eukaryotic nuclear ribosomal operon, we identify two possible types of helix formation: one type is “dichotomous helix formation”—transition from one large helix to two smaller helices by invagination of the apical part of a helix, which significantly changes the shape of the original secondary structure but does not increase its complexity (i.e., the total length of the RNA). An alternative type is “lateral helix formation”—origin of an extra helical region by the extension of a bulge loop or a spacer in a multi-helix loop of the original helix, which does not disrupt the pre-existing structure but increases RNA size. Moreover, we present examples from the RNA sequence literature indicating that both types of helix formation may have implications for RNA evolution beyond ITS2.  相似文献   

18.
The sensitivity of bacteriophage T4 lysozyme function to amino acid substitutions at defined positions in and around the longitudinal, hydrophobic strips of 9 alpha-helices was assessed after systematic replacement of each residue in the protein with a series of 13 amino acids. The hydrophobic strips were defined by identifying the longitudinal sectors in the helices with the highest mean residue hydrophobicities. Sensitivity to mutation (the percentage of replacements leading to loss of function) was calculated for each residue in the following positions: whole protein, helices, hydrophobic strips, other positions within the helices, and various positions within the hydrophobic strips as well as their extensions beyond the helices. Substitutions at positions in the hydrophobic strips led more frequently to loss of function than substitutions in the protein as a whole. One subset, the COOH-terminal hydrophobic strip residues, is apparently critical; substitutions of these residues (but not of their NH2-terminal counterparts) led at least as frequently to loss of function as substitutions of solvent-inaccessible residues, and nearly as frequently as substitutions of the most highly conserved residues.  相似文献   

19.
In striated muscle, the binding of calcium to troponin C (TnC) results in the removal of the C‐terminal region of the inhibitory protein troponin I (TnI) from actin. While structural studies of the muscle system have been successful in determining the overall organization of most of the components involved in force generation at the atomic level, the structure and dynamics of the C‐terminal region of TnI remains controversial. This domain of TnI is highly flexible, and it has been proposed that this intrinsically disordered region (IDR) regulates contraction via a “fly‐casting” mechanism. Different structures have been presented for this region using different methodologies: a single α‐helix, a “mobile domain” containing a small β‐sheet, an unstructured region, and a two helix segment. To investigate whether this IDR has in fact any nascent structure, we have constructed a skeletal TnC‐TnI chimera that contains the N‐domain of TnC (1–90), a short linker (GGAGG), and the C‐terminal region of TnI (97–182) and have acquired 15N NMR relaxation data for this chimera. We compare the experimental relaxation parameters with those calculated from molecular dynamic simulations using four models based upon the structural studies. Our experimental results suggest that the C‐terminal region of TnI does not contain any defined secondary structure, supporting the “fly‐casting” mechanism. We interpret the presence of a “plateau” in the 15N NMR relaxation data as being an intrinsic property of IDRs. We also identified a more rigid adjacent region of TnI that has implications for muscle performance under ischemic conditions. Proteins 2011. © 2011 Wiley‐Liss, Inc.  相似文献   

20.
Hicks JM  Hsu VL 《Proteins》2004,55(2):330-338
The poly-proline type II extended left-handed helical structure is well represented in proteins. In an effort to determine the helix's role in nucleic acid recognition and binding, a survey of 258 nucleic acid-binding protein structures from the Protein Data Bank was conducted. Results indicate that left-handed helices are commonly found at the nucleic acid interfacial regions. Three examples are used to illustrate the utility of this structural element as a recognition motif. The third K homology domain of NOVA-2, the Epstein-Barr nuclear antigen-1, and the Drosophila paired protein homeodomain all contain left-handed helices involved in nucleic acid interactions. In each structure, these helices were previously unidentified as left-handed helices by secondary structure algorithms but, rather, were identified as either having small amounts of hydrogen bond patterns to the rest of the protein or as being "unstructured." Proposed mechanisms for nucleic acid interactions by the extended left-handed helix include both nonspecific and specific recognition. The observed interactions indicate that this secondary structure utilizes an increase in protein backbone exposure for nucleic acid recognition. Both main-chain and side-chain atoms are involved in specific and nonspecific hydrogen bonding to nucleobases or sugar-phosphates, respectively. Our results emphasize the need to classify the left-handed helix as a viable nucleic acid recognition and binding motif, similar to previously identified motifs such as the helix-turn-helix, zinc fingers, leucine zippers, and others.  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