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.
An analysis of possible tight packing of hydrophobic groups simultaneously at the both surfaces of β-hyperboloid-8 was conducted. This analysis shows that the disposition of amino acid side chains at the real β-structure's surface is unique. If we sign the mean distance between adjacent β-strands as “a,” and the mean distance along β-strand between Cα atoms, whose side chains are directed to one side of the β-sheet, as “b,” the ratio b/a = √2 very precisely. This ratio ensures the most efficient packing of side hydrophobic groups at the outer surface of β-hyperboloid-8, forming, at the same time, the second by efficiency packing at its inner surface. © 1995 Wiley-Liss, Inc.  相似文献   

2.
Both Type I' and Type II' β‐turns have the same sense of the β‐turn twist that is compatible with the β‐sheet twist. They occur predominantly in two residue β‐hairpins, but the occurrence of Type I' β‐turns is two times higher than Type II' β‐turns. This suggests that Type I' β‐turns may be more stable than Type II' β‐turns, and Type I' β‐turn sequence and structure can be more favorable for protein folding than Type II' β‐turns. Here, we redesigned the native Type II' β‐turn in GFP to Type I' β‐turn, and investigated its effect on protein folding and stability. The Type I' β‐turns were designed based on the statistical analysis of residues in natural Type I' β‐turns. The substitution of the native “GD” sequence of i+1 and i+2 residues with Type I' preferred “(N/D)G” sequence motif increased the folding rate by 50% and slightly improved the thermodynamic stability. Despite the enhancement of in vitro refolding kinetics and stability of the redesigned mutants, they showed poor soluble expression level compared to wild type. To overcome this problem, i and i + 3 residues of the designed Type I' β‐turn were further engineered. The mutation of Thr to Lys at i + 3 could restore the in vivo soluble expression of the Type I' mutant. This study indicates that Type II' β‐turns in natural β‐hairpins can be further optimized by converting the sequence to Type I'. Proteins 2014; 82:2812–2822. © 2014 Wiley Periodicals, Inc.  相似文献   

3.
Tannin acylhydrolase (EC 3.1.1.20) referred commonly as tannase catalyzes the hydrolysis of the galloyl ester bond of tannins to release gallic acid. Although the enzyme is useful for various industries, the tertiary structure is not yet determined. In this study, we determined the crystal structure of tannase produced by Lactobacillus plantarum. The tannase structure belongs to a member of α/β‐hydrolase superfamily with an additional “lid” domain. A glycerol molecule derived from cryoprotectant solution was accommodated into the tannase active site. The binding manner of glycerol to tannase seems to be similar to that of the galloyl moiety in the substrate. Proteins 2013; 81:2052–2058. © 2013 Wiley Periodicals, Inc.  相似文献   

4.
α-Prolamins are the major seed storage proteins of species of the grass tribe Andropogonea. They are unusually rich in glutamine, proline, alanine, and leucine residues and their sequences show a series of tandem repeats presumed to be the result of multiple intragenic duplication. Two new sequences of α-prolamin clones from Coix (pBCX25.12 and pBCX25.10) are compared with similar clones from maize and Sorghum in order to investigate evolutionary relationships between the repeat motifs and to propose a schematic model for their three-dimensional structure based on hydrophobic membrane-helix propensities and helical “wheels.” A scheme is proposed for the most recent events in the evolution of the central part of the molecule (repeats 3 to 8) which involves two partial intragenic duplications and in which contemporary odd-numbered and even-numbered repeats arise from common ancestors, respectively. Each pair of repeats is proposed to form an antiparallel α-helical hairpin and that the helices of the molecule as a whole are arranged on a hexagonal net. The majority of helices show six faces of alternating hydrophobic and polar residues, which give rise to intersticial holes around each helix which alternate in chemical character. The model is consistent with proteins which contain different numbers of repeats, with oligomerization and with the dense packaging of α-prolamins within the protein body of the seed endosperm. © 1993 Wiley-Liss, Inc.  相似文献   

5.
The structure of AF2331, a 11‐kDa orphan protein of unknown function from Archaeoglobus fulgidus, was solved by Se‐Met MAD to 2.4 Å resolution. The structure consists of an α + β fold formed by an unusual homodimer, where the two core β‐sheets are interdigitated, containing strands alternating from both subunits. The decrease in solvent‐accessible surface area upon dimerization is unusually large (3960 Å2) for a protein of its size. The percentage of the total surface area buried in the interface (41.1%) is one of the largest observed in a nonredundant set of homodimers in the PDB and is above the mean for nearly all other types of homo‐oligomers. AF2331 has no sequence homologs, and no structure similar to AF2331 could be found in the PDB using the CE, TM‐align, DALI, or SSM packages. The protein has been identified in Pfam 23.0 as the archetype of a new superfamily and is topologically dissimilar to all other proteins with the “3‐Layer (BBA) Sandwich” fold in CATH. Therefore, we propose that AF2331 forms a novel α + β fold. AF2331 contains multiple negatively charged surface clusters and is located on the same operon as the basic protein AF2330. We hypothesize that AF2331 and AF2330 may form a charge‐stabilized complex in vivo, though the role of the negatively charged surface clusters is not clear.  相似文献   

6.
The kinetics of α-helix formation in polyalanine and polyglycine eicosamers (20-mers) were examined using torsional-coordinate molecular dynamics (MD). Of one hundred fifty-five MD experiments on extended (Ala)20 carried out for 0.5 ns each, 129 (83%) formed a persistent α-helix. In contrast, the extended state of (Gly)20 only formed a right-handed α-helix in two of the 20 MD experiments (10%), and these helices were not as long or as persistent as those of polyalanine. These simulations show helix formation to be a competition between the rates of (a) forming local hydrogen bonds (i.e. hydrogen bonds between any residue i and its i + 2, i + 3, i + 4, or i + 5th neighbor) and (b) forming nonlocal hydrogen bonds (HBs) between residues widely separated in sequence. Local HBs grow rapidly into an α-helix; but nonlocal HBs usually retard helix formation by “trapping” the polymer in irregular, “balled-up” structures. Most trajectories formed some nonlocal HBs, sometimes as many as eight. But, for (Ala)20, most of these eventually rearranged to form local HBs that lead to α-helices. A simple kinetic model describes the rate of converting nonlocal HBs into α-helices. Torsional-coordinate MD speeds folding by eliminating bond and angle degrees of freedom and reducing dynamical friction. Thus, the observed 210 ps half-life for helix formation is likely to be a lower bound on the real rate. However, we believe the sequential steps observed here mirror those of real systems. Proteins 33:343–357, 1998. © 1998 Wiley-Liss, Inc.  相似文献   

7.
17-O-Acetyl testosterone, which has no susceptible hydroxyl or carboxyl group for glycosylation, was glycosylated with 2,3,4,6-tetra-O-acetyl-α- -glucopyranosyl bromide in the presence of a mixed catalyst, Hg(CN)2 and HgBr2, in benzene–nitromethane. Reaction occurred on the α,β-unsaturated ketone on the six–membered A-ring to give six 3-O-glycosides, each bearing a cyano group at the 3- or 5-position of the aglycon, and a 3-O-glycoside bearing a CONH2 group at the 3-position. Structural analyses of these products were carried out by various NMR (1H, 13C NMR, 1H–1H and 1H–13C COSY, HMBC, and DEPT), FABMS and X-ray analyses. The mechanisms of the formations of the products are discussed. It was determined that mercuric cyanide was essential as a catalyst for the progress of the cyanoglycosylation.  相似文献   

8.
Germline chimeric chickens can be constructed by injecting donor chicken blastodermal cells (CBCs) into recipient embryos and incubating to hatch. Transgenic chickens can be produced through chimeric intermediates if the donor cells are genetically manipulated; the chance of producing a transgenic chimera would be increased by enriching the donor population in transfected cells. To demonstrate that donor CBCs can be sorted according to the expression of a foreign gene, CBCs in suspension were subjected to transfection with plasmid DNA encoding bacterial β‐galactosidase (β‐gal). Following an overnight incubation, the CBCs were loaded with 5‐dodecanoylaminofluorescein di‐β‐D‐galactopyranoside (C12FDG), which is fluorescent after cleavage by β‐gal. The treated cells were subjected to fluorescence activated cell sorting (FACS) to give “positive” (fluorescent) and “negative” (non‐fluorescent) populations. Almost 100% of the “positive” population showed β‐gal activity. “Positive” cells were cultured on mouse SNL 76/7 fibroblast feeder cells and formed colonies, most of which still stained positively for β‐gal activity after three days. FACS‐sorted cells of Barred Plymouth Rock origin were injected into recipient White Leghorn embryos, resulting in chimeric embryos. Of the 298 embryos injected with sorted cells, 23 (8%; 18 injected with “positive cells, five with “negative”) survived to rearing. Somatic chimerism was seen in 12 of 18 (67%) “positive” and three of five (60%) “negative” birds with the proportion of black pigmentation averaging 19% overall. Twenty birds reached sexual maturity, of which 12 (60%) were somatically chimeric; seven (35%) of these produced donor‐derived chicks. Donor CBCs can, therefore, be sorted by FACS according to the expression of a selectable marker gene without impairing their ability to contribute to germline chimeras; this procedure could be incorporated into a practicable method by which to increase the chances of producing a transgenic chicken. Mol. Reprod. Dev. 52:33–42, 1999. © 1999 Wiley‐Liss, Inc.  相似文献   

9.
The properties of natural human interferon γ (IFN-γ) molecules dissolved in protein-denaturing and non-denaturing solvents were examined by high-performance size-exclusion chromatography on a gel permeation column. IFN-γ and tritium-labeled IFN-γ molecules formed either dimers (>90.5%) with the molecular mass of 60 kDa or probably tetramers (<9.5%) with the molecular mass of approximately 100 kDa in non-denaturing solvents, and no monomer was detected. These oligomers were dissociated in protein-denaturing solvents such as 6 M guanidine hydrochloride, and IFN-γ existed as monomers. There is no effect on formation of the monomer based on the dissociation of oligomers by acid treatment at pH 4.0. The monomers in protein-denaturing solvents formed dimers by association when applied to a column equilibrated with a non-denaturing solvent of phosphate buffer, pH 7.0. In conclusion, natural human IFN-γ forms oligomers, particularly dimers, in non-denaturing solution, and this oligomer formation is a reversible reaction.  相似文献   

10.
NagZ is an exo‐N‐acetyl‐β‐glucosaminidase, found within Gram‐negative bacteria, that acts in the peptidoglycan recycling pathway to cleave N‐acetylglucosamine residues off peptidoglycan fragments. This activity is required for resistance to cephalosporins mediated by inducible AmpC β‐lactamase. NagZ uses a catalytic mechanism involving a covalent glycosyl enzyme intermediate, unlike that of the human exo‐N‐acetyl‐β‐glucosaminidases: O‐GlcNAcase and the β‐hexosaminidase isoenzymes. These latter enzymes, which remove GlcNAc from glycoconjugates, use a neighboring‐group catalytic mechanism that proceeds through an oxazoline intermediate. Exploiting these mechanistic differences we previously developed 2‐N‐acyl derivatives of O‐(2‐acetamido‐2‐deoxy‐D ‐glucopyranosylidene)amino‐N‐phenylcarbamate (PUGNAc), which selectively inhibits NagZ over the functionally related human enzymes and attenuate antibiotic resistance in Gram‐negatives that harbor inducible AmpC. To understand the structural basis for the selectivity of these inhibitors for NagZ, we have determined its crystallographic structure in complex with N‐valeryl‐PUGNAc, the most selective known inhibitor of NagZ over both the human β‐hexosaminidases and O‐GlcNAcase. The selectivity stems from the five‐carbon acyl chain of N‐valeryl‐PUGNAc, which we found ordered within the enzyme active site. In contrast, a structure determination of a human O‐GlcNAcase homologue bound to a related inhibitor N‐butyryl‐PUGNAc, which bears a four‐carbon chain and is selective for both NagZ and O‐GlcNAcase over the human β‐hexosamnidases, reveals that this inhibitor induces several conformational changes in the active site of this O‐GlcNAcase homologue. A comparison of these complexes, and with the human β‐hexosaminidases, reveals how selectivity for NagZ can be engineered by altering the 2‐N‐acyl substituent of PUGNAc to develop inhibitors that repress AmpC mediated β‐lactam resistance.  相似文献   

11.
The crystal structure of the model tripeptide Boc-Aib-Gly-Leu-OMe ( 1 ) reveals two independent molecules in the asymmetric unit that adopt “enantiomeric” type I and type I′ β-turn conformations with the Aib and Gly residues occupying the corner (i + 1 and i + 2) positions. 13C cross polarization and magic angle sample spinning spectra in the solid state also support the coexistence of two conformational species. 13C-nmr in CDCl3 establishes the presence of a single species or rapid exchange between conformations. 400 MHz 1H-nmr provides evidence for conformational exchange involving a major and minor species, with β-turn conformations supported by the low solvent exposure of Leu(3) NH and the observation of NiH ↔ Ni+1H nuclear Overhauser effects. CD bands in the region 190–230 nm are positive, supporting a major population of type I′ β-turns. The isomeric peptide, Boc-Gly-Leu-Aib-OMe ( 2 ), adopts an “open” type II′ β-turn conformation in crystals. Solid state and solution nmr support population of a single conformational species. Chiral perturbation introduced outside the folded region of peptides may provide a means of modulating screw sense in achiral sequences. © 1998 John Wiley & Sons, Inc. Biopoly 45: 191–202, 1998  相似文献   

12.
The full‐length complementary DNA (cDNA) sequences encoding cd8α and cd8β molecules were sequenced and characterized from mandarin fish Siniperca chuatsi. Conserved motifs and residues were found to be present in derived peptides of the Cd8 molecules. For example, WXR motif, DXGXYXC motif, and four cysteine residues were present in the extracellular region of the Cd8 protein. Threonine, serine and proline residues involved in multiple O‐linked glycosylation events were located in the membrane proximal hinge region. The common CPH motif in the cytoplasmic tail was detected similar to other teleost Cd8 molecules. Different from those in mammals, S. chuatsi Cd8 sequences have many extra cysteine residues (C149 in Cd8α sequence and C46, C51 and C158 in Cd8β sequence), which also exist in other teleost Cd8 molecules. Real‐time polymerase chain reaction (RT‐PCR) and Western blot analyses revealed that the thymus had the highest expression of cd8 messenger (m)RNA and protein. After stimulated with phytohaemagglutinin, polyriboinsine‐polyribocyaidylic acid and concanavalin A (ConA), the expression level of cd8 mRNA increased significantly in head‐kidney lymphocytes at 4 and 8 h, but decreased to normal level at 12 h. Similarly, stimulation with ConA in vivo also led to an increase in the cd8 mRNA level in the spleen. Immunohistochemistry analysis demonstrated that Cd8α‐positive cells can be detected in the thymus, spleen and intestine by using polyclonal anti‐Cd8α antibody.  相似文献   

13.
This work describes an investigation of pathways and binging pockets (BPs) for dioxygen (O2) through the cofactorless oxygenase 3‐hydroxy‐2‐methylquinolin‐4‐one 2,4‐dioxygenase in complex with its natural substrate, 3‐hydroxy‐2‐methylquinolin‐4(1H)‐one, in aqueous solution. The investigation tool was random‐acceleration molecular dynamics (RAMD), whereby a tiny, randomly oriented external force is applied to O2 in order to accelerate its movements. In doing that, care was taken that the external force only continues, if O2 moves along a direction for a given period of time, otherwise the force changed direction randomly. Gates for expulsion of O2 from the protein, which can also be taken as gates for O2 uptake, were found throughout almost the whole external surface of the protein, alongside a variety of BPs for O2. The most exploited gates and BPs were not found to correspond to the single gate and BP proposed previously from the examination of the static model from X‐ray diffraction analysis of this system. Therefore, experimental investigations of this system that go beyond the static model are urgently needed.  相似文献   

14.
The high cost of enzymes is a major bottleneck preventing the development of an economically viable lignocellulosic ethanol industry. Commercial enzyme cocktails for the conversion of plant biomass to fermentable sugars are complex mixtures containing more than 80 proteins of suboptimal activities and relative proportions. As a step toward the development of a more efficient enzyme cocktail for biomass conversion, we have developed a platform, called GENPLAT, that uses robotic liquid handling and statistically valid experimental design to analyze synthetic enzyme mixtures. Commercial enzymes (Accellerase 1000 +/? Multifect Xylanase, and Spezyme CP +/? Novozyme 188) were used to test the system and serve as comparative benchmarks. Using ammonia‐fiber expansion (AFEX) pretreated corn stover ground to 0.5 mm and a glucan loading of 0.2%, an enzyme loading of 15 mg protein/g glucan, and 48 h digestion at 50°C, commercial enzymes released 53% and 41% of the available glucose and xylose, respectively. Mixtures of three, five, and six pure enzymes of Trichoderma species, expressed in Pichia pastoris, were systematically optimized. Statistical models were developed for the optimization of glucose alone, xylose alone, and the average of glucose + xylose for two digestion durations, 24 and 48 h. The resulting models were statistically significant (P < 0.0001) and indicated an optimum composition for glucose release (values for optimized xylose release are in parentheses) of 29% (5%) cellobiohydrolase 1, 5% (14%) cellobiohydrolase 2, 25% (25%) endo‐β1,4‐glucanase 1, 14% (5%) β‐glucosidase, 22% (34%) endo‐β1,4‐xylanase 3, and 5% (17%) β‐xylosidase in 48 h at a protein loading of 15 mg/g glucan. Comparison of two AFEX‐treated corn stover preparations ground to different particle sizes indicated that particle size (100 vs. 500 µm) makes a large difference in total digestibility. The assay platform and the optimized “core” set together provide a starting point for the rapid testing and optimization of alternate core enzymes from other microbial and recombinant sources as well as for the testing of “accessory” proteins for development of superior enzyme mixtures for biomass conversion. Biotechnol. Bioeng. 2010;106: 707–720. © 2010 Wiley Periodicals, Inc.  相似文献   

15.
The Tyr corner is a conformation in which a tyrosine (residue “Y”) near the beginning or end of an antiparallel β-strand makes an H bond from its side-chain OH group to the backbone NH and/or CO of residue Y – 3, Y – 4, or Y – 5 in the nearby connection. The most common “classic” case is a Δ4 Tyr corner (more than 40 examples listed), in which the H bond is to residue Y – 4 and the Tyr x1 is near ?60°. Y – 2 is almost always a glycine, whose left-handed β or very extended β conformation helps the backbone curve around the Tyr ring. Residue Y – 3 is in polyproline II conformation (often Pro), and residue Y – 5 is usually a hydrophobic (often Leu) that packs next to the Tyr ring. The consensus sequence, then, is LxPGxY, where the first x (the H-bonding position) is hydrophilic. Residues Y and Y – 2 both form narrow pairs of β-sheet H-bonds with the neighboring strand, Δ5 Tyr corners have a 1-residue insertion between the Gly and the Tyr, forming a β-bulge. One protein family has a Δ4 corner formed by a His rather than a Tyr, and several examples use Trp in place of Tyr. For almost all these cases, the protein or domain is a Greek key β-barrel structure, the Tyr corner ends a Greek key connection, and it is well-conserved in related proteins. Most low-twist Greek key β-barrels have 1 Tyr corner. “Reverse” Δ4 Tyr corners (H bonded to Y + 4) and other variants are described, all less common and less conserved. It seems likely that the more classic Tyr corners (Δ4, Δ5, and Δ3 Tyr, Trp, or His) contribute to the stability of a Greek key connection over a hairpin connection, and also that they may aid in the process of folding up Greek key structures.  相似文献   

16.
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   

17.
Biological diversity analysis is among the most informative approaches to describe communities and regional species compositions. Soil ecosystems include large numbers of invertebrates, among which soil bugs (Crustacea, Isopoda, Oniscidea) play significant ecological roles. The aim of this study was to provide advices to optimize the sampling effort, to efficiently monitor the diversity of this taxon, to analyze its seasonal patterns of species composition, and ultimately to understand better the coexistence of so many species over a relatively small area. Terrestrial isopods were collected at the Natural Reserve “Saline di Trapani e Paceco” (Italy), using pitfall traps monthly monitored over 2 years. We analyzed parameters of α‐ and β‐diversity and calculated a number of indexes and measures to disentangle diversity patterns. We also used various approaches to analyze changes in biodiversity over time, such as distributions of species abundances and accumulation and rarefaction curves. As concerns species richness and total abundance of individuals, spring resulted the best season to monitor Isopoda, to reduce sampling efforts, and to save resources without losing information, while in both years abundances were maximum between summer and autumn. This suggests that evaluations of β‐diversity are maximized if samples are first collected during the spring and then between summer and autumn. Sampling during these coupled seasons allows to collect a number of species close to the γ‐diversity (24 species) of the area. Finally, our results show that seasonal shifts in community composition (i.e., dynamic fluctuations in species abundances during the four seasons) may minimize competitive interactions, contribute to stabilize total abundances, and allow the coexistence of phylogenetically close species within the ecosystem.  相似文献   

18.
The synthesis of new dermorphin analogues is described. The (R)‐alanine or phenylalanine residues of natural dermorphin were substituted by the corresponding α‐methyl‐β‐azidoalanine or α‐benzyl‐β‐azido(1‐piperidinyl)alanine residues. The potency and selectivity of the new analogues were evaluated by a competitive receptor binding assay in rat brain using [3H]DAMGO (a μ ligand) and [3H]DELT (a δ ligand). The most active analogue in this series, Tyr‐(R)‐Ala‐(R)‐α‐benzyl‐β‐azidoAla‐Gly‐Tyr‐Pro‐Ser‐NH2 and its epimer were analysed by 1H and 13C NMR spectroscopy and restrained molecular dynamics simulations. The dominant conformation of the investigated peptides depended on the absolute configuration around Cα in the α‐benzyl‐β‐azidoAla residue in position 3. The (R) configuration led to the formation of a type I β‐turn, whilst switching to the (S) configuration gave rise to an inverse β‐turn of type I′, followed by the formation of a very short β‐sheet. The selectivity of Tyr‐(R)‐Ala‐(R) and (S)‐α‐benzyl‐β‐azidoAla‐Gly‐Tyr‐Pro‐Ser‐NH2 was shown to be very similar; nevertheless, the two analogues exhibited different conformational preferences. Copyright © 2016 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

19.
20.
Results of energy calculations for α-MSH (α-melanocyte stimulating hormone, Ac-Ser1-Tyr2-Ser3-Met4-Glu5-His6-Phe7-Arg8-Trp9-Gly10-Lys11-Pro12-Val13-NH2) and [D -Phe7]α-MSH were used for design of cyclic peptides with the general aim to stabilize different conformational isomers of the parent compound. The minimal structural modifications of the conformationally flexible Gly10 residue, as substitutions for L -Ala, D -Ala, or Aib (replacing of hydrogen atoms by methyl groups), were applied to obtain octa- and heptapeptide analogues of α-MSH(4–11) and α-MSH(5–11), which were cyclized by lactam bridges between the side chains in positions 5 and 11. Some of these analogues, namely those with substitutions of the Gly10 residue with L -Ala or Aib, showed biological activity potencies on frog skin comparable to the potency of the parent tridecapeptide hormone. Additional energy calculations for designed cyclic analogues were used for further refinement of the model for the biologically active conformations of the His-Phe-Arg-Trp “message” sequence within the sequences of α-MSH and [D -Phe7]α-MSH. In such conformations the aromatic moieties of the side chains of the His6, L/D -Phe7, and Trp9 residues form a continuous hydrophobic “surface,” presumably interacting with a complementary receptor site. This feature is characteristic for low-energy conformers of active cyclic analogues, but it is absent in the case of inactive analogues. This particular spatial arrangement of functional groups involved in the message sequence is very close for α-MSH and [D -Phe7]α-MSH, as well as for biologically active cyclic analogues despite differences of dihedral angle values for corresponding low-energy conformations. © 1998 John Wiley & Sons, Inc. Biopoly 46: 155–167, 1998  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