首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Nucleic acid recognition is often mediated by α‐helices or disordered regions that fold into α‐helix on binding. A peptide bearing the DNA recognition helix of HPV16 E2 displays type II polyproline (PII) structure as judged by pH, temperature, and solvent effects on the CD spectra. NMR experiments indicate that the canonical α‐helix is stabilized at the N‐terminus, while the PII forms at the C‐terminus half of the peptide. Re‐examination of the dihedral angles of the DNA binding helix in the crystal structure and analysis of the NMR chemical shift indexes confirm that the N‐terminus half is a canonical α‐helix, while the C‐terminal half adopts a 310 helix structure. These regions precisely match two locally driven folding nucleii, which partake in the native hydrophobic core and modulate a conformational switch in the DNA binding helix. The peptide shows only weak and unspecific residual DNA binding, 104‐fold lower affinity, and 500‐fold lower discrimination capacity compared with the domain. Thus, the precise side chain conformation required for modulated and tight physiological binding by HPV E2 is largely determined by the noncanonical strained α‐helix conformation, “presented” by this unique architecture. © 2009 Wiley Periodicals, Inc. Biopolymers 91: 432–443, 2009. This article was originally published online as an accepted preprint. The “Published Online” date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com  相似文献   

2.
Salmon calcitonin (sCT) was selected as a model protein drug for investigating its intrinsic thermal stability and conformational structure in the solid and liquid states by using a Fourier transform infrared (FT‐IR) microspectroscopy with or without utilizing thermal analyzer. The spectral correlation coefficient (r) analysis between two second‐derivative IR spectra was applied to quantitatively estimate the structural similarity of sCT in the solid state before and after different treatments. The thermal FT‐IR microspectroscopic data clearly evidenced that sCT in the solid state was not effected by temperature and had a thermal reversible property during heating–cooling process. Moreover, the high r value of 0.973 or 0.988 also evidenced the structural similarity of solid‐state sCT samples before and after treatments. However, sCT in H2O exhibited protein instability and thermal irreversibility after incubation at 40°C. The temperature‐induced conformational changes of sCT in H2O was occurred to transform the α‐helix/random coil structures to β‐sheet structure and also resulted in the formation of intramolecular and intermolecular β‐sheet structures. © 2009 Wiley Periodicals, Inc. Biopolymers 93: 200–207, 2010. This article was originally published online as an accepted preprint. The “Published Online” date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com  相似文献   

3.
Common structural elements in proteins such as α‐helices or β‐sheets are characterized by uniformly repeating, energetically favorable main chain conformations which additionally exhibit a completely saturated hydrogen‐bonding network of the main chain NH and CO groups. Although polyproline or polyglycine type II helices (PPII or PGII) are frequently found in proteins, they are not considered as equivalent secondary structure elements because they do not form a similar self‐contained hydrogen‐bonding network of the main chain atoms. In this context our finding of an unusual motif of glycine‐rich PGII‐like helices in the structure of the acetophenone carboxylase core complex is of relevance. These PGII‐like helices form hexagonal bundles which appear to fulfill the criterion of a (largely) saturated hydrogen‐bonding network of the main‐chain groups and therefore may be regarded in this sense as a new secondary structure element. It consists of a central PGII‐like helix surrounded by six nearly parallel PGII‐like helices in a hexagonal array, plus an additional PGII‐like helix extending the array outwards. Very related structural elements have previously been found in synthetic polyglycine fibers. In both cases, all main chain NH and CO groups of the central PGII‐helix are saturated by either intra‐ or intermolecular hydrogen‐bonds, resulting in a self‐contained hydrogen‐bonding network. Similar, but incomplete PGII‐helix patterns were also previously identified in a GTP‐binding protein and an antifreeze protein.  相似文献   

4.
Recent research has implicated the C‐terminus of G‐protein coupled receptors in key events such as receptor activation and subsequent intracellular sorting, yet obtaining structural information of the entire C‐tail has proven a formidable task. Here, a peptide corresponding to the full‐length C‐tail of the human CB1 receptor (residues 400–472) was expressed in E.coli and purified in a soluble form. Circular dichroism (CD) spectroscopy revealed that the peptide adopts an α‐helical conformation in negatively charged and zwitterionic detergents (48–51% and 36–38%, respectively), whereas it exhibited the CD signature of unordered structure at low concentration in aqueous solution. Interestingly, 27% helicity was displayed at high peptide concentration suggesting that self‐association induces helix formation in the absence of a membrane mimetic. NMR spectroscopy of the doubly labeled (15N‐ and 13C‐) C‐terminus in dodecylphosphocholine (DPC) identified two amphipathic α‐helical domains. The first domain, S401‐F412, corresponds to the helix 8 common to G protein‐coupled receptors while the second domain, A440‐M461, is a newly identified structural motif in the distal region of the carboxyl‐terminus of the receptor. Molecular modeling of the C‐tail in DPC indicates that both helices lie parallel to the plane of the membrane with their hydrophobic and hydrophilic faces poised for critical interactions. © 2009 Wiley Periodicals, Inc. Biopolymers 91: 565–573, 2009. This article was originally published online as an accepted preprint. The “Published Online” date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com  相似文献   

5.
While end capping in α‐helices is well understood, the concept of capping a β‐hairpin is a relatively recent development; to date, favorable Coulombic interactions are the only example of sidechains at the termini influencing the overall stability of a β‐hairpin. While cross‐strand hydrophobic residues generally provide hairpin stabilization, particular when flanking the turn region, those remote from this location appear to provide little stabilization. While probing for an optimal residue at a hydrogen bond position near the terminus of a designed β‐hairpin a conservative, hydrophobic, V → I mutation was observed to not only result in a significant change in fold population but also effected major changes in the structuring shifts at numerous sites in the peptide. Mutational studies reveal that there is an interaction between the sidechain at this H‐bonded site and the sidechain at the C‐terminal non‐H‐bonded site of the hairpin. This interaction, which appears to be hydrophobic in character, requires a highly twisted hairpin structure. Modifications at the C‐terminal site, for example an E → A mutation (ΔΔGU = 6 kJ/mol), have profound affects on fold structure and stability. The data suggests that this may be a case of hairpin end capping by the formation of a hydrophobic cluster. © 2009 Wiley Periodicals, Inc. Biopolymers 91: 557–564, 2009. This article was originally published online as an accepted preprint. The “Published Online”date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com  相似文献   

6.
A 34‐residue α/β peptide [IG(28–61)], derived from the C‐terminal part of the B3 domain of the immunoglobulin binding protein G from Streptoccocus, was studied using CD and NMR spectroscopy at various temperatures and by differential scanning calorimetry. It was found that the C‐terminal part (a 16‐residue‐long fragment) of this peptide, which corresponds to the sequence of the β‐hairpin in the native structure, forms structure similar to the β‐hairpin only at T = 313 K, and the structure is stabilized by non‐native long‐range hydrophobic interactions (Val47–Val59). On the other hand, the N‐terminal part of IG(28–61), which corresponds to the middle α‐helix in the native structure, is unstructured at low temperature (283 K) and forms an α‐helix‐like structure at 305 K, and only one helical turn is observed at 313 K. At all temperatures at which NMR experiments were performed (283, 305, and 313 K), we do not observe any long‐range connectivities which would have supported packing between the C‐terminal (β‐hairpin) and the N‐terminal (α‐helix) parts of the sequence. Such interactions are absent, in contrast to the folding pathway of the B domain of protein G, proposed recently by Kmiecik and Kolinski (Biophys J 2008, 94, 726–736), based on Monte‐Carlo dynamics studies. Alternative folding mechanisms are proposed and discussed. © 2010 Wiley Periodicals, Inc. Biopolymers 93: 469–480, 2010. This article was originally published online as an accepted preprint. The “Published Online” date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com  相似文献   

7.
Chengcheng Hu  Patrice Koehl 《Proteins》2010,78(7):1736-1747
The three‐dimensional structure of a protein is organized around the packing of its secondary structure elements. Although much is known about the packing geometry observed between α‐helices and between β‐sheets, there has been little progress on characterizing helix–sheet interactions. We present an analysis of the conformation of αβ2 motifs in proteins, corresponding to all occurrences of helices in contact with two strands that are hydrogen bonded. The geometry of the αβ2 motif is characterized by the azimuthal angle θ between the helix axis and an average vector representing the two strands, the elevation angle ψ between the helix axis and the plane containing the two strands, and the distance D between the helix and the strands. We observe that the helix tends to align to the two strands, with a preference for an antiparallel orientation if the two strands are parallel; this preference is diminished for other topologies of the β‐sheet. Side‐chain packing at the interface between the helix and the strands is mostly hydrophobic, with a preference for aliphatic amino acids in the strand and aromatic amino acids in the helix. From the knowledge of the geometry and amino acid propensities of αβ2 motifs in proteins, we have derived different statistical potentials that are shown to be efficient in picking native‐like conformations among a set of non‐native conformations in well‐known decoy datasets. The information on the geometry of αβ2 motifs as well as the related statistical potentials have applications in the field of protein structure prediction. Proteins 2010. © 2010 Wiley‐Liss, Inc.  相似文献   

8.
A fluorescently labeled 20‐residue polyglutamic acid (polyE) peptide 20 amino acid length polyglutamic acid (E20) was used to study structural changes which occur in E20 as it co‐aggregates with other unlabeled polyE peptides. Resonance energy transfer (RET) was performed using an o‐aminobenzamide donor at the N‐terminus and 3‐nitrotyrosine acceptor at the C‐terminus of E20. PolyE aggregates were not defined as amyloid, as they were nonfibrillar and did not bind congo red. Circular dichroism measurements indicate that polyE aggregation involves a transition from α‐helical monomers to aggregated β‐sheets. Soluble oligomers are also produced along with aggregates in the reaction, as determined through size exclusion chromatography. Time‐resolved and steady‐state RET measurements reveal four dominant E20 conformations: (1) a partially collapsed conformation (24 Å donor–acceptor distance) in monomers, (2) an extended conformation in soluble oligomers (>29 Å donor–acceptor distance), (3) a minor partially collapsed conformation (22 Å donor‐acceptor distance) in aggregates, and (4) a major highly collapsed conformation (13 Å donor–acceptor distance) in aggregates. These findings demonstrate the use of RET as a means of determining angstrom‐level structural details of soluble oligomer and aggregated states of proteins. © 2009 Wiley Periodicals, Inc. Biopolymers 93: 299–317, 2010. This article was originally published online as an accepted preprint. The “Published Online” date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com  相似文献   

9.
Vasoactive intestinal peptide (VIP) and pituitary adenylate cyclase‐activating polypeptide‐27 (PACAP27) are members of the secretin‐glucagon family containing 28 and 27 residues, respectively. NMR spectroscopy studies suggest that the N‐terminus exhibit consecutive β‐turns whereas the central and C‐terminal parts of the VIP molecule have been characterized as being two α‐helices. In contrast, similar studies carried out on PACAP suggest that the shortest active peptide segment PACAP27 in the presence of trifluoroethanol (TFE) exhibits a disordered N‐terminal domain followed by a α‐helix expanding residues 9–26 with a discontinuity between residues 20 and 21. In the present study, a series of MD trajectories of VIP and PACAP27 were carried out using two different implicit models of the solvent: the Generalized Born that use an effective Born radii described by Onufriev, Bashford, and Case (GBOBC) and the Hawkins, Cramer, and Truhlar approximation (GBHCT) and two different force fields: AMBER ff99 and a modified version of the latter described by Sorin and Pande (Biophys J 2005, 88, 2472‐2493), ff99SP. Comparison of the structures obtained from the MD trajectories and those derived from the NMR studies in the literature indicates that the GBOBC method is more efficient in the exploration of the conformational space and presents a higher agreement with the experimental structure of VIP and PACAP27 in TFE. © 2009 Wiley Periodicals, Inc. Biopolymers 91: 391–400, 2009. This article was originally published online as an accepted preprint. The “Published Online” date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com  相似文献   

10.
A de novo polypeptide GH6[(GA)3GY(GA)3GE]8GAH6 (YE8) has a significant number of identical weakly interacting β‐strands with the turns and termini functionalized by charged amino acids to control polypeptide folding and aggregation. YE8 exists in a soluble, disordered form at neutral pH but is responsive to changes in pH and ionic strength. The evolution of YE8 secondary structure has been successfully quantified during all stages of polypeptide fibrillation by deep UV resonance Raman (DUVRR) spectroscopy combined with other morphological, structural, spectral, and tinctorial characterization. The YE8 folding kinetics at pH 3.5 are strongly dependent on polypeptide concentration with a lag phase that can be eliminated by seeding with a solution of folded fibrillar YE8. The lag phase of polypeptide folding is concentration dependent leading to the conclusion that β‐sheet folding of the 11‐kDa amyloidogenic polypeptide is completely aggregation driven. © 2010 Wiley Periodicals, Inc. Biopolymers 93: 607–618, 2010. This article was originally published online as an accepted preprint. The “Published Online” date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com  相似文献   

11.
Fourier transform infrared (FT‐IR) spectroscopy combined with 2D correlation spectroscopy has been used to offer some information about stability and structure of some soluble elastins. Temperature has been chosen as the perturbation to monitor the infrared behavior of various soluble elastins, namely, α‐elastin p, α‐elastin, and k‐elastin. In the 3800–2700 cm?1 region, the H‐containing groups were analyzed. The bonded hydroxyls are found to decrease prior to the NH‐related hydrogen bonds and also to the conformational reorganization of hydrocarbon chains. The transition temperatures were evaluated and they were found to agree with those obtained from DSC data. The FTIR spectra and their 2nd derivatives denote that α‐ elastins exhibited amide‐I, ‐II and ‐III bands at 1656, 1539 and 1236 cm?1, respectively, while in k‐elastin these bands were found at 1652 cm?1 for amide I, 1540 cm?1 for amide II and 1248 cm?1 for amide III. The macroscopic IR finger‐print method, which combines: general IR spectra, secondary derivative spectra, and 2D‐IR correlation spectra, is useful to discriminate different elastins. Thus using the differences of the position and intensity of the bands from “fingerprint region” of studied elastins, which include the peaks assigned to C?O, C? C groups from α‐helix, β‐turn, and the peaks assigned to the amide groups, it is possible to identify and discriminate elastins from each others. Furthermore, the pattern of 2D‐IR correlation spectra under thermal perturbation, allow their direct identification and discrimination. © 2010 Wiley Periodicals, Inc. Biopolymers 93: 1072–1084, 2010. This article was originally published online as an accepted preprint. The “Published Online” date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com  相似文献   

12.
The anti‐lipopolysaccharide factor ALF‐Pm3 is a 98‐residue protein identified in hemocytes from the black tiger shrimp Penaeus monodon. It was expressed in Pichia pastoris from the constitutive glyceraldehyde‐3‐phosphate dehydrogenase promoter as a folded and 15N uniformly labeled rALF‐Pm3 protein. Its 3D structure was established by NMR and consists of three α‐helices packed against a four‐stranded β‐sheet. The C34? C55 disulfide bond was shown to be essential for the structure stability. By using surface plasmon resonance, we demonstrated that rALF‐Pm3 binds to LPS, lipid A and to OM®‐174, a soluble analogue of lipid A. Biophysical studies of rALF‐Pm3/LPS and rALF‐Pm3/OM®‐174 complexes indicated rather high molecular sized aggregates, which prevented us to experimentally determine by NMR the binding mode of these lipids to rALF‐Pm3. However, on the basis of striking structural similarities to the FhuA/LPS complex, we designed an original model of the possible lipid A‐binding site of ALF‐Pm3. Such a binding site, located on the ALF‐Pm3 β‐sheet and involving seven charged residues, is well conserved in ALF‐L from Limulus polyphemus and in ALF‐T from Tachypleus tridentatus. In addition, our model is in agreement with experiments showing that β‐hairpin synthetic peptides corresponding to ALF‐L β‐sheet bind to LPS. Delineating lipid A‐binding site of ALFs will help go further in the de novo design of new antibacterial or LPS‐neutralizing drugs. © 2008 Wiley Periodicals, Inc. Biopolymers 91: 207–220, 2009. This article was originally published online as an accepted preprint. The “Published Online” date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com  相似文献   

13.
Hu C  Koehl P  Max N 《Proteins》2011,79(10):2828-2843
The three‐dimensional structure of a protein is organized around the packing of its secondary structure elements. Predicting the topology and constructing the geometry of structural motifs involving α‐helices and/or β‐strands are therefore key steps for accurate prediction of protein structure. While many efforts have focused on how to pack helices and on how to sample exhaustively the topologies and geometries of multiple strands forming a β‐sheet in a protein, there has been little progress on generating native‐like packings of helices on sheets. We describe a method that can generate the packing of multiple helices on a given β‐sheet for αβα sandwich type protein folds. This method mines the results of a statistical analysis of the conformations of αβ2 motifs in protein structures to provide input values for the geometric attributes of the packing of a helix on a sheet. It then proceeds with a geometric builder that generates multiple arrangements of the helices on the sheet of interest by sampling through these values and performing consistency checks that guarantee proper loop geometry between the helices and the strands, minimal number of collisions between the helices, and proper formation of a hydrophobic core. The method is implemented as a module of ProteinShop. Our results show that it produces structures that are within 4–6 Å RMSD of the native one, regardless of the number of helices that need to be packed, though this number may increase if the protein has several helices between two consecutive strands in the sequence that pack on the sheet formed by these two strands. Proteins 2011; Published 2011 Wiley‐Liss, Inc.  相似文献   

14.
The natural amino acids are primarily helix breakers at the low assignment temperatures characteristic of many studies, but recent genomic analyses of thermophilic proteins suggest that at high temperatures, some breakers may become strong helix formers. Moreover, the breaker/former inventory has not been previously characterized at the physiologically relevant temperature of 37°C. The versatility of 13C?O NMR chemical shifts as helicity reporters allows construction of two mutant peptide series, tailored to expand the range of temperature assignments for helical propensities and derived from the core hosts tL‐Ala9XxxAla9tL and tL‐AlaNva4XxxNva4Ala9tL, Nva = norvaline. For three limiting guests Xxx, the helix former Nva and the breakers Gly and Pro, we report wXxx[T] assignments at seven temperatures from 2 to 80°C, validating our reasoning and paving the way for assignment of a definitive wXxx[T] data‐base. © 2008 Wiley Periodicals, Inc. Biopolymers 91: 311–320, 2009. This article was originally published online as an accepted preprint. The “Published Online” date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com  相似文献   

15.
Astringency is one of the major organoleptic properties of food and beverages that are made from plants, such as tea, chocolate, beer, or red wine. This sensation is thought to be due to interactions between tannins and salivary proline‐rich proteins, which are natively unfolded proteins. A human salivary proline‐rich protein, namely IB‐5, was produced by the recombinant method. Its interactions with a model tannin, epigallocatechin gallate (EGCG), the major flavan‐3‐ol in green tea, were studied here. Circular dichroism experiments showed that IB‐5 presents residual structures (PPII helices) when the ionic strength is close to that in saliva. In the presence of these residual structures, IB‐5 undergoes an increase in structural content upon binding to EGCG. NMR data corroborated the presence of preformed structural elements within the protein prior to binding and a partial assignment was proposed, showing partial structuration. TOCSY experiments showed that amino acids that are involved in PPII helices are more likely to interact with EGCG than those in random coil regions, as if they were anchorage points for the ligand. The signal from IB‐5 in the DOSY NMR spectrum revealed an increase in polydispersity upon addition of EGCG while the mean hydrodynamic radius remained unchanged. This strongly suggests the formation of IB‐5/EGCG aggregates. © 2009 Wiley Periodicals, Inc. Biopolymers 91: 745–756, 2009. This article was originally published online as an accepted preprint. The “Published Online” date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com  相似文献   

16.
Pectin is a group of carbohydrate polymers constructing the primary cell walls and the middle lamella of terrestrial plants. Herein, we demonstrated the structure and immunomodulatory activity of the major pectic polysaccharide DL‐3B2 isolated from the leaves of Diospyros kaki. Based on composition analysis, methylation analysis, two‐step acid hydrolysis, lithium‐mediated selective degradation, 13C NMR spectroscopy, and electrospray ionization mass spectrometry, DL‐3B2 was found to contain an α‐1, 4‐linked galacturonic acid (GalA) backbone with some insertions of α‐1, 2‐linked rhamnose residues. The arabinan‐ and arabinogalactan‐side chains were attached to O‐4 of the rhamnose residues, whereas the linear arabinoxylan was probably linked to O‐3 of the GalA residues. Immunological tests in vitro showed that DL‐3B2 could help stimulate lipopolysaccharide‐induced B lymphocyte proliferation, but not ConA‐induced T lymphocyte proliferation, and that the arabinose residues play a role in maintaining this immunological activity. © 2010 Wiley Periodicals, Inc. Biopolymers 93: 649–656, 2010. This article was originally published online as an accepted preprint. The “Published Online” date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com  相似文献   

17.
The secondary structure transitions of regenerated silk fibroin (RSF) under different external perturbations have been studied extensively, except for pressure. In this work, time‐resolved infrared spectroscopy with the attenuated total reflectance (ATR) accessory was employed to follow the secondary structure transitions of RSF in its wet film under low pressure. It has been found that pressure alone is favorable only to the formation of β‐sheet structure. Under constant pressure there is an optimum amount of D2O in the wet film (D2O : film = 2:1) so as to provide the optimal condition for the reorganization of the secondary structure and to have the largest formation of β‐sheet structure. Under constant amount of D2O and constant pressure, the secondary structure transitions of RSF in its wet film can be divided into three stages along with time. In the first stage, random coil, α‐helix, and β‐turn were quickly transformed into β‐sheet. In the second stage, random coil and β‐turn were relatively slowly transformed into β‐sheet and α‐helix, and the content of α‐helix was recovered to the value prior to the application of pressure. In the third and final stage, no measurable changes can be found for each secondary structure. This study may be helpful to understand the secondary structure changes of silk fibroin in silkworm's glands under hydrostatic pressure.  相似文献   

18.
In this report, we describe the X‐ray crystal structures of two single domain camelid antibodies (VHH), F5 and F8, each in complex with ricin toxin's enzymatic subunit (RTA). F5 has potent toxin‐neutralizing activity, while F8 has weak neutralizing activity. F5 buried a total of 1760 Å2 in complex with RTA and made contact with three prominent secondary structural elements: α‐helix B (Residues 98–106), β‐strand h (Residues 113–117), and the C‐terminus of α‐helix D (Residues 154–156). F8 buried 1103 Å2 in complex with RTA that was centered primarily on β‐strand h. As such, the structural epitope of F8 is essentially nested within that of F5. All three of the F5 complementarity determining regions CDRs were involved in RTA contact, whereas F8 interactions were almost entirely mediated by CDR3, which essentially formed a seventh β‐strand within RTA's centrally located β‐sheet. A comparison of the two structures reported here to several previously reported (RTA‐VHH) structures identifies putative contact sites on RTA, particularly α‐helix B, associated with potent toxin‐neutralizing activity. This information has implications for rational design of RTA‐based subunit vaccines for biodefense. Proteins 2016; 84:1162–1172. © 2016 Wiley Periodicals, Inc.  相似文献   

19.
HP1454 is a protein of 303 amino acids found in the extracellular milieu of Helicobacter pylori. The protein structure, crystallized in the orthorhombic C2221 space group with one molecule per asymmetric unit, has been determined using the single‐wavelength anomalous dispersion method. HP1454 exhibits an elongated bent shape, composed of three distinct domains. Each domain possesses a fold already present in other structures: Domain I contains a three‐strand antiparallel β‐barrel flanked by a long α‐helix, Domain II is an anti‐parallel three‐helix bundle, and Domain III a β‐sheet flanked by two α‐helices. The overall assembly of the protein does not bear any similarity with known structures. Proteins 2014; 82:2868–2873. © 2014 Wiley Periodicals, Inc.  相似文献   

20.
Many of the bilayer spanning segments of membrane transport proteins contain proline residues, and most of them are believed to occur in α-helical form. A proline residue in the middle of an α-helix is known to produce a bend in the helix, and recent studies have focused on characterizing such a bend at atomic level. In the present case, molecular dynamics (MD) studies are carried out on helix F model of bacteriorhodopsin (BR) Ace-(Ala)7-Trp-(Ala)2-Tyr-Pro-(Ala)2-Trp-(Ala)8-NHMe and compared with Ace-(Ala)7-Trp-(Ala)2-Tyr-(Ala)3-Trp-(Ala)8-NHMe in which the proline is replaced by alanine. The bend in the helix is characterized by structural parameters such as kink angle (α), wobble angle (θ), virtual torsion angle (ρ), and the hydrogen bond distance d (Op?3 … Np+1). The average values and the flexibility involved in these parameters are evaluated. The correlation among the bend related parameters are estimated. The equilibrium side chain orientations of tryptophan and tyrosine residues are discussed and compared with those found in the recently proposed model of bacteriorhodopsin. Finally, a detailed characterization of the bend in terms of secondary structures such as αI, αII and goniometric helices are discussed, which can be useful in the interpretation of the experimental results on the secondary structures of membrane proteins involving the proline residue. © 1993 Wiley-Liss, Inc.  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