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 46 毫秒
1.
Our goal was to gain a better understanding of how protein stability can be increased by improving β‐turns. We studied 22 β‐turns in nine proteins with 66–370 residues by replacing other residues with proline and glycine and measuring the stability. These two residues are statistically preferred in some β‐turn positions. We studied: Cold shock protein B (CspB), Histidine‐containing phosphocarrier protein, Ubiquitin, Ribonucleases Sa2, Sa3, T1, and HI, Tryptophan synthetase α‐subunit, and Maltose binding protein. Of the 15 single proline mutations, 11 increased stability (Average = 0.8 ± 0.3; Range = 0.3–1.5 kcal/mol), and the stabilizing effect of double proline mutants was additive. On the basis of this and our previous work, we conclude that proteins can generally be stabilized by replacing nonproline residues with proline residues at the i + 1 position of Type I and II β‐turns and at the i position in Type II β‐turns. Other turn positions can sometimes be used if the φ angle is near ?60° for the residue replaced. It is important that the side chain of the residue replaced is less than 50% buried. Identical substitutions in β‐turns in related proteins give similar results. Proline substitutions increase stability mainly by decreasing the entropy of the denatured state. In contrast, the large, diverse group of proteins considered here had almost no residues in β‐turns that could be replaced by Gly to increase protein stability. Improving β‐turns by substituting Pro residues is a generally useful way of increasing protein stability. Proteins 2009. © 2009 Wiley‐Liss, Inc.  相似文献   

2.
The Ramachandran map clearly delineates the regions of accessible conformational (φ–ψ) space for amino acid residues in proteins. Experimental distributions of φ, ψ values in high‐resolution protein structures, reveal sparsely populated zones within fully allowed regions and distinct clusters in apparently disallowed regions. Conformational space has been divided into 14 distinct bins. Residues adopting these relatively rare conformations are presented and amino acid propensities for these regions are estimated. Inspection of specific examples in a completely “arid”, fully allowed region in the top left quadrant establishes that side‐chain and backbone interactions may provide the energetic compensation necessary for populating this region of φ–ψ space. Asn, Asp, and His residues showed the highest propensities in this region. The two distinct clusters in the bottom right quadrant which are formally disallowed on strict steric considerations correspond to the gamma turn (C7 axial) conformation (Bin 12 ) and the i + 1 position of Type II′ β turns (Bin 13) . Of the 516 non‐Gly residues in Bin 13 , 384 occupied the i + 1 position of Type II′ β turns. Further examination of these turn segments revealed a high propensity to occur at the N‐terminus of helices and as a tight turn in β hairpins. The β strand–helix motif with the Type II′ β turn as a connecting element was also found in as many as 57 examples. Proteins 2014; 82:1101–1112. © 2013 Wiley Periodicals, Inc.  相似文献   

3.
Mimicry of structural motifs is a common feature in proteins. The 10‐membered hydrogen‐bonded ring involving the main‐chain C?O in a β‐turn can be formed using a side‐chain carbonyl group leading to Asx‐turn. We show that the N? H component of hydrogen bond can be replaced by a Cγ‐H group in the side chain, culminating in a nonconventional C? H···O interaction. Because of its shape this β‐turn mimic is designated as ω‐turn, which is found to occur ~three times per 100 residues. Three residues (i to i + 2) constitute the turn with the C? H···O interaction occurring between the terminal residues, constraining the torsion angles ?i + 1, ψi + 1, ?i + 2 and χ1(i + 2) (using the interacting Cγ atom). Based on these angles there are two types of ω‐turns, each of which can be further divided into two groups. Cβ‐branched side‐chains, and Met and Gln have high propensities to occur at i + 2; for the last two residues the carbonyl oxygen may participate in an additional interaction involving the S and amino group, respectively. With Cys occupying the i + 1 position, such turns are found in the metal‐binding sites. N‐linked glycosylation occurs at the consensus pattern Asn‐Xaa‐Ser/Thr; with Thr at i + 2, the sequence can adopt the secondary structure of a ω‐turn, which may be the recognition site for protein modification. Location between two β‐strands is the most common occurrence in protein tertiary structure, and being generally exposed ω‐turn may constitute the antigenic determinant site. It is a stable scaffold and may be used in protein engineering and peptide design. Proteins 2015; 83:203–214. © 2014 Wiley Periodicals, Inc.  相似文献   

4.
Elucidating protein function from its structure is central to the understanding of cellular mechanisms. This involves deciphering the dependence of local structural motifs on sequence. These structural motifs may be stabilized by direct or water‐mediated hydrogen bonding among the constituent residues. π‐Turns, defined by interactions between (i) and (i + 5) positions, are large enough to contain a central space that can embed a water molecule (or a protein moiety) to form a stable structure. This work is an analysis of such embedded π‐turns using a nonredundant dataset of protein structures. A total of 2965 embedded π‐turns have been identified, as also 281 embedded Schellman motif, a type of π‐turn which occurs at the C‐termini of α‐helices. Embedded π‐turns and Schellman motifs have been classified on the basis of the protein atoms of the terminal turn residues that are linked by the embedded moiety, conformation, residue composition, and compared with the turns that have terminal residues connected by direct hydrogen bonds. Geometrically, the turns have been fitted to a circle and the position of the linker relative to its center analyzed. The hydroxyl group of Ser and Thr, located at (i + 3) position, is the most prominent linker for the side‐chain mediated π‐turns. Consideration of residue conservation among homologous sequences indicates the terminal and the linker positions to be the most conserved. The embedded π‐turn as a binding site (for the linker) is discussed in the context of “nest,” a concave depression that is formed in protein structures with adjacent residues having enantiomeric main‐chain conformations. © 2013 Wiley Periodicals, Inc. Biopolymers 101: 441–453, 2014.  相似文献   

5.
An important nucleation event during the folding of staphylococcal nuclease involves the formation of a β‐hairpin by the sequence 21DTVKLMYKGQPMTFR35. Earlier studies show that the turn sequence ‘YKGQP’ has an important role in the folding of this β‐hairpin. To understand the active or passive nature of the turn sequence ‘YKGQP’ in the folding of the aforementioned β‐hairpin sequence, we studied glycine mutant peptides Ac‐2DTVKLMYGGQPMTFR16‐NMe (K9G:15), Ac‐2DTVKLMYKGGPMTFR16‐NMe (Q11G:15), Ac‐2DTVKLMYGGGPMTFR16‐NMe (K9G/Q11G:15), and Ac‐2DTVKLMGGGGGMTFR16‐NMe (penta‐G:15) by using molecular dynamics simulations, starting with two different unfolded states, polyproline II and extended conformational forms. Further, 5mer mutant turn peptides Ac‐2YGGQP6‐NMe (K3G:5), Ac‐2YKGGP6‐NMe (Q5G:5), Ac‐2YGGGP6‐NMe (K3G/Q5G:5), and Ac‐2GGGGG6‐NMe (penta‐G:5) were also studied individually. Our results show that an initial hydrophobic collapse and loop closure occurs in all 15mer mutants, but only K9G:15 mutant forms a stable native‐like β‐hairpin. In the other 15mer mutants, the hydrophobic collapsed state would not proceed to β‐hairpin formation. Of the different simulations performed for the penta‐G:15 mutant, in only one simulation a nonnative β‐hairpin conformation is sampled with highly flexible loop region (8GGGGG12), which has no specific conformational preference as a 5mer. While the sequence ‘YGGQP’ in the K3G:5 simulation shows relatively higher β‐turn propensity, the presence of this sequence in K9G:15 peptide seems to be driving the β‐hairpin formation. Thus, these results seem to suggest that for the formation of a stable β‐hairpin, the initial hydrophobic collapse is to be assisted by a turn propensity. Initial hydrophobic collapse alone is not sufficient to guide β‐hairpin formation. Copyright © 2013 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

6.
The introduction of non‐natural modules could provide unprecedented control over folding/unfolding behavior, conformational stability, and biological function of proteins. Success requires the interrogation of candidate modules in natural contexts. Here, expressed protein ligation is used to replace a reverse turn in bovine pancreatic ribonuclease (RNase A) with a synthetic β‐dipeptide: β2‐homoalanine–β3‐homoalanine. This segment is known to adopt an unnatural reverse‐turn conformation that contains a 10‐membered ring hydrogen bond, but one with a donor–acceptor pattern opposite to that in the 10‐membered rings of natural reverse turns. The RNase A variant has intact enzymatic activity, but unfolds more quickly and has diminished conformational stability relative to native RNase A. These data indicate that hydrogen‐bonding pattern merits careful consideration in the selection of beneficial reverse‐turn surrogates.  相似文献   

7.
Young Kee Kang    Byung Jin Byun 《Biopolymers》2012,97(12):1018-1025
The γ‐peptide β‐turn structures have been designed computationally by the combination of chirospecific γ 2 , 3 ‐residues of 2‐(aminomethyl)cyclohexanecarboxylic acid (γAmc6) with a cyclohexyl constraint on the Cα?Cβ bond using density functional methods in water. The chirospecific γAmc6 dipeptide with the (2S,3S)‐(2R,3R) configurations forms a stable turn structure in water, resembling a type II′ turn of α‐peptides, which can be used as a β‐turn motif in β‐hairpins of Ala‐based α‐peptides. The γAmc6 dipeptide with homochiral (2S,3S)‐(2S,3S) configurations but different cyclohexyl puckerings shows the capability to be incorporated into one of two β‐turn motifs of gramicidin S. The overall structure of this gramicidin S analogue is quite similar to the native gramicidin S with the same patterns and geometries of hydrogen bonds. Our calculated results and the recently observed results may imply the wider applicability of chirospecific γ‐peptides with a cyclohexyl constraint on the backbone to form various peptide foldamers. © 2012 Wiley Periodicals, Inc. Biopolymers 97:1018–1025, 2012.  相似文献   

8.
We study the differences in folding stability of β‐hairpin peptides, including GB1 hairpin and a point mutant GB1 K10G, as well as tryptophan zippers (TrpZips): TrpZip1, TrpZip2, TrpZip3‐1, and TrpZip4. By performing replica‐exchange molecular dynamics simulations with Amber03* force field (a modified version of Amber ff03) in explicit solvent, we observe ab initio folding of all the peptides except TrpZip3‐1, which is experimentally known to be the least stable among the peptides studied here. By calculating the free energies of unfolding of the peptides at room temperature and folding midpoint temperatures for thermal unfolding of peptides, we find that TrpZip4 and GB1 K10G peptides are the most stable β‐hairpins followed by TrpZip1, GB1, and TrpZip2 in the given order. Hence, the proposed K10G mutation of GB1 peptide results in enhanced stability compared to wild‐type GB1. An important goal of our study is to test whether simulations with Amber 03* model can reproduce experimentally predicted folding stability differences between these peptides. While the stabilities of GB1 and TrpZip1 yield close agreement with experiment, TrpZip2 is found to be less stable than predicted by experiment. However, as heterogenous folding of TrpZip2 may yield divergent thermodynamic parameters by different spectroscopic methods, mismatching of results with previous experimental values are not conclusive of model shortcomings. For most of the cases, molecular simulations with Amber03* can successfully reproduce experimentally known differences between the mutated peptides, further highlighting the predictive capabilities of current state‐of‐the‐art all‐atom protein force fields. Proteins 2015; 83:1307–1315. © 2015 Wiley Periodicals, Inc.  相似文献   

9.
The propensities of peptides that contain the Asn‐Gly segment to form β‐turn and β‐hairpin structures were explored using the density functional methods and the implicit solvation model in CH2Cl2 and water. The populations of preferred β‐turn structures varied depending on the sequence and solvent polarity. In solution, β‐hairpin structures with βI′ turn motifs were most preferred for the heptapeptides containing the Asn‐Gly segment regardless of the sequence of the strands. These preferences in solution are consistent with the corresponding X‐ray structures. The sequence, H‐bond strengths, solvent polarity, and conformational flexibility appeared to interact to determine the preferred β‐hairpin structure of each heptapeptide, although the β‐turn segments played a role in promoting the formation of β‐hairpin structures and the β‐hairpin propensity varied. In the heptapeptides containing the Asn‐Gly segment, the β‐hairpin formation was enthalpically favored and entropically disfavored at 25°C in water. The calculated results for β‐turns and β‐hairpins containing the Asn‐Gly segment imply that these structural preferences may be useful for the design of bioactive macrocyclic peptides containing β‐hairpin mimics and the design of binding epitopes for protein–protein and protein–nucleic acid recognitions. © 2016 Wiley Periodicals, Inc. Biopolymers 105: 653–664, 2016.  相似文献   

10.
The increasing interest in click chemistry and its use to stabilize turn structures led us to compare the propensity for β‐turn stabilization of different analogs designed as mimics of the β‐turn structure found in tendamistat. The β‐turn conformation of linear β‐amino acid‐containing peptides and triazole‐cyclized analogs were compared to ‘conventional’ lactam‐ and disulfide‐bridged hexapeptide analogs. Their 3D structures and their propensity to fold in β‐turns in solution, and for those not structured in solution in the presence of α‐amylase, were analyzed by NMR spectroscopy and by restrained molecular dynamics with energy minimization. The linear tetrapeptide Ac‐Ser‐Trp‐Arg‐Tyr‐NH2 and both the amide bond‐cyclized, c[Pro‐Ser‐Trp‐Arg‐Tyr‐D ‐Ala] and the disulfide‐bridged, Ac‐c[Cys‐Ser‐Trp‐Arg‐Tyr‐Cys]‐NH2 hexapeptides adopt dominantly in solution a β‐turn conformation closely related to the one observed in tendamistat. On the contrary, the β‐amino acid‐containing peptides such as Ac‐(R)‐β3‐hSer‐(S)‐Trp‐(S)‐β3‐hArg‐(S)‐β3‐hTyr‐NH2, and the triazole cyclic peptide, c[Lys‐Ser‐Trp‐Arg‐Tyr‐βtA]‐NH2, both specifically designed to mimic this β‐turn, do not adopt stable structures in solution and do not show any characteristics of β‐turn conformation. However, these unstructured peptides specifically interact in the active site of α‐amylase, as shown by TrNOESY and saturation transfer difference NMR experiments performed in the presence of the enzyme, and are displaced by acarbose, a specific α‐amylase inhibitor. Thus, in contrast to amide‐cyclized or disulfide‐bridged hexapeptides, β‐amino acid‐containing peptides and click‐cyclized peptides may not be regarded as β‐turn stabilizers, but can be considered as potential β‐turn inducers. Copyright © 2011 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

11.
To investigate the structural role played by isostructural unbranched alkyl‐chains on the conformational ensemble and stability of β‐turn structures, the conformational properties of a designed model peptide: Plm‐Pro‐Gly‐Pda ( 1 , Plm: H3C—(CH2)14—CONH—; Pda: —CONH— (CH2)14—CH3) have been examined and compared with the parent peptide: Boc‐Pro‐Gly‐NHMe ( 2 , Boc: tert‐butoxycarbonyl; NHMe: N‐methylamide). The characteristic 13C NMR chemical‐shifts of the Pro Cβ and Cγ resonances ascertained the incidence of an all‐trans peptide‐bond in low polarity deuterochloroform solution. Using FTIR and 1H NMR spectroscopy, we establish that apolar alkyl‐chains flanking a β‐turn promoting Pro‐Gly sequence impart definite incremental stability to the well‐defined hydrogen‐bonded structure. The assessment of 1H NMR derived thermodynamic parameters of the hydrogen‐bonded amide‐NHs via variable temperature indicate that much weaker hydrophobic interactions do contribute to the stability of folded reverse turn structures. The far‐UV CD spectral patterns of 1 and 2 in 2,2,2‐trifluoroethanol are consistent with Pro‐Gly specific type II β‐turn structure, concomitantly substantiate that the flanking alkyl‐chains induce substantial bias in enhanced β‐turn populations. In view of structural as well as functional importance of the Pro‐Gly mediated secondary structures, besides biochemical and biological significance of proteins lipidation via myristoylation or palmytoilation, we highlight potential convenience of the unbranched Plm and Pda moieities not only as main‐chain N‐ and C‐terminal protecting groups but also to mimic and stabilize specific isolated secondary and supersecondary structural components frequently observed in proteins and polypeptides. © 2013 Wiley Periodicals, Inc. Biopolymers 99: 419–426, 2013.  相似文献   

12.
Alzheimer's disease is the most common form of dementia in humans and is related to the accumulation of the amyloid‐β (Aβ) peptide and its interaction with metals (Cu, Fe, and Zn) in the brain. Crystallographic structural information about Aβ peptide deposits and the details of the metal‐binding site is limited owing to the heterogeneous nature of aggregation states formed by the peptide. Here, we present a crystal structure of Aβ residues 1–16 fused to the N‐terminus of the Escherichia coli immunity protein Im7, and stabilized with the fragment antigen binding fragment of the anti‐Aβ N‐terminal antibody WO2. The structure demonstrates that Aβ residues 10–16, which are not in complex with the antibody, adopt a mixture of local polyproline II‐helix and turn type conformations, enhancing cooperativity between the two adjacent histidine residues His13 and His14. Furthermore, this relatively rigid region of Aβ (residues, 10–16) appear as an almost independent unit available for trapping metal ions and provides a rationale for the His13‐metal‐His14 coordination in the Aβ1–16 fragment implicated in Aβ metal binding. This novel structure, therefore, has the potential to provide a foundation for investigating the effect of metal ion binding to Aβ and illustrates a potential target for the development of future Alzheimer's disease therapeutics aimed at stabilizing the N‐terminal monomer structure, in particular residues His13 and His14, and preventing Aβ metal‐binding‐induced neurotoxicity.Proteins 2013; 81:1748–1758. © 2013 Wiley Periodicals, Inc.  相似文献   

13.
The design of biomimetic materials through molecular self‐assembly is a growing area of modern nanotechnology. With problems of protein folding, self‐assembly, and sequence–structure relationships as essential in nanotechnology as in biology, the effect of the nucleation of β‐hairpin formation by proline on the folding process has been investigated in model studies. Previously such studies were limited to investigations of the influence of proline on the formation of turns in short peptide sequences. The effect of proline‐based triads on the folding of an 11‐kDa amyloidogenic peptide GH6[(GA)3GY(GA)3GE]8GAH6 ( YE8 ) was investigated by selective substitution of the proline‐substituted triads at the γ‐turn sites. The folding and fibrillation of the singly proline‐substituted polypeptides, e.g., GH6? [(GA)3GY(GA)3GE]7(GA)3GY(GA)3PD? GAH6 ( 8PD ), and doubly proline‐substituted polypeptides, e.g., GH6? [(GA)3GY(GA)3GE]3(GA)3GY(GA)3PD[(GA)3GY(GA)3GE]3(GA)3GY(GA)3PD? GAH6 ( 4,8PD ), were directly monitored by circular dichroism and deep UV resonance Raman and fluorescence spectroscopies. These findings were used to identify the essential folding domains, i.e., the minimum number of β‐strands necessary for stable folding. These experimental findings may be especially useful in the design and construction of peptidic materials for a wide range of applications as well as in understanding the mechanisms of folding critical to fibril formation. © 2015 Wiley Periodicals, Inc. Biopolymers 103: 339–350, 2015.  相似文献   

14.
Fibrillation of β‐amyloid is recognized as a key process leading to the development of Alzheimer's disease. Small peptides called β‐sheet breakers were found to inhibit the process of β‐amyloid fibrillation and to dissolve amyloid fibrils in vitro, in vivo, and in cell culture studies [1,2]. The mechanism by which peptide inhibition takes place remains elusive and a detailed model needs to be established. Here, we present new insights into the possible role of consecutive Phe residues, present in the structure of β‐sheet breakers, supported by the results obtained by means of MD simulations. We performed a 30‐ns MD of two β‐sheet breakers: iAβ5 (LPFFD) and iAβ6 (LPFFFD) which have two and three consecutive Phe residues, respectively. We have found that Phe rings in these peptides tend to form stacked conformations. For one of the peptides – iAβ6 – the calculated electrostatic contribution to free energy of one of the conformers with three rings stacked (c2) is significantly lower than that corresponding to the unstacked one (c1), two rings stacked (c0) and second conformer with three rings stacked (c3). This may favor the interaction of the c2 conformer with the target on amyloid fibril. We hypothesize that the mechanism of inhibition of amyloidogenesis by β‐sheet breaker involves competition among π‐stacked Phe residues of the inhibitor and π‐stacking within the β‐amyloid fibril. iAβ6 may be a promising candidate for a lead compound of amyloidogenesis inhibitors. Copyright © 2013 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

15.
In this study, proteinogenic amino acids residues of dimeric dermorphin pentapeptides were replaced by the corresponding β3homo‐amino acids. The potency and selectivity of hybrid α/β dimeric dermorphin pentapeptides were evaluated by competetive receptor binding assay in the rat brain using [3H]DAMGO (a μ ligand) and [3H]DELT (a δ ligand). Tha analog containing β3homo‐Tyr in place of Tyr (Tyr‐d ‐Ala‐Phe‐Gly‐β3homo‐Tyr‐NH‐)2 showed good μ receptor affinity and selectivity (IC50 = 0.302, IC50 ratio μ/δ = 68) and enzymatic stability in human plasma. Copyright © 2016 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

16.
Alzheimer's disease is a progressive neurodegenerative disease characterized by extracellular deposits of β‐amyloid (Aβ) plaques. Aggregation of the Aβ42 peptide leading to plaque formation is believed to play a central role in Alzheimer's disease pathogenesis. Anti‐Aβ monoclonal antibodies can reduce amyloid plaques and could possibly be used for immunotherapy. We have developed a monoclonal antibody C706, which recognizes the human Aβ peptide. Here we report the crystal structure of the antibody Fab fragment at 1.7 Å resolution. The structure was determined in two crystal forms, P21 and C2. Although the Fab was crystallized in the presence of Aβ16, no peptide was observed in the crystals. The antigen‐binding site is blocked by the hexahistidine tag of another Fab molecule in both crystal forms. The poly‐His peptide in an extended conformation occupies a crevice between the light and heavy chains of the variable domain. Two consecutive histidines (His4–His5) stack against tryptophan residues in the central pocket of the antigen‐binding surface. In addition, they form hydrogen bonds to the acidic residues at the bottom of the pocket. The mode of his‐tag binding by C706 resembles the Aβ recognition by antibodies PFA1 and WO2. All three antibodies recognize the same immunodominant B‐cell epitope of Aβ. By similarity, residues Phe–Arg–His of Aβ would be a major portion of the C706 epitope. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

17.
Lanreotide, a synthetic cyclic octapeptide, analogue of the peptide hormone somatostatin‐14 (SST‐14), is routinely used as a long‐acting medication in the management of neuroendocrine tumors. Despite its therapeutic importance, low concentration structural data is still lacking for lanreotide. In fact, the major part of the previous structural investigations were focused on the remarkable aggregation properties of this peptide, appearing at high concentrations (>5 mM). Here, we have applied three optical spectroscopic techniques, i.e. fluorescence, circular dichroism and Raman scattering, for analyzing the structural dynamics at the concentrations below 5 mM, where lanreotide exists either in a monomer state or at the first stages of aggregation. The obtained data from lanreotide were discussed through their comparison with those collected from SST‐14, leading us to the following conclusions: (i) The central D‐Trp residue, forming with its adjacent Lys the main receptor interacting part of lanreotide, keeps a constant high rotational freedom whatever the environment (water, water/methanol, methanol). (ii) A solvent‐dependent tight β‐turn, belonging to the type‐II' family, is revealed in lanreotide. (iii) Raman data analyzed by band decomposition in the amide (I and III) regions allowed estimation of different secondary structural elements within the millimolar range. Interestingly, the applied protocol shows a perfect agreement between the structural features provided by the amide I and amide III Raman markers. © 2014 Wiley Periodicals, Inc. Biopolymers 101: 1019–1028, 2014.  相似文献   

18.
The relationship between the conformation and biological activity of the peptide allosteric modulator of the interleukin‐1 receptor 101.10 (D ‐Arg‐D ‐Tyr‐D ‐Thr‐D ‐Val‐D ‐Glu‐D ‐Leu‐D ‐Ala‐NH2) has been studied using (R)‐ and (S)‐Bgl residues. Twelve Bgl peptides were synthesized using (R)‐ and (S)‐cyclic sulfamidate reagents derived from L ‐ and D ‐aspartic acid in an optimized Fmoc‐compatible protocol for efficient lactam installment onto the supported peptide resin. Examination of these (R)‐ and (S)‐Bgl 101.10 analogs for their potential to inhibit IL‐1β‐induced thymocyte cell proliferation using a novel fluorescence assay revealed that certain analogs exhibited retained and improved potency relative to the parent peptide 101.10. In light of previous reports that Bgl residues may stabilize type II′β‐turn‐like conformations in peptides, CD spectroscopy was performed on selected compounds to identify secondary structure necessary for peptide biological activity. Results indicate that the presence of a fold about the central residues of the parent peptide may be important for activity. Copyright © 2011 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

19.
Many years ago, β2/β3‐peptides, consisting of alternatively arranged β2‐ and β3h‐amino‐acid residues, have been found to undergo folding to a unique type of helix, the 10/12‐helix, and to exhibit non‐polar, lipophilic properties (Helv. Chim. Acta 1997 , 80, 2033). We have now synthesized such ‘mixed’ hexa‐, nona‐, dodeca‐, and octadecapeptides, consisting of Val‐Ala‐Leu triads, with N‐terminal fluorescein (FAM) labels, i.e., 1 – 4 , and studied their interactions with POPC (=1‐palmitoyl‐2‐oleoyl‐sn‐glycero‐3‐phosphocholine) giant unilamellar vesicles (GUVs) and with human white blood cancer cells U937. The methods used were microfluidic technology, fluorescence correlation spectroscopy (FCS), a flow‐cytometry assay, a membrane‐toxicity assay with the dehydrogenase G6PDH as enzymatic reporter, and visual microscopy observations. All β3/β2‐peptide derivatives penetrate the GUVs and/or the cells. As shown with the isomeric β3/β2‐, β3‐, and β2‐nonamers, 2, 5 , and 6 , respectively, the derivatives 5 and 6 consisting exclusively of β3‐ or β2‐amino‐acid residues, respectively, interact neither with the vesicles nor with the cells. Depending on the method of investigation and on the pretreatment of the cells, the β3/β2‐nonamer and/or the β3/β2‐dodecamer derivative, 2 and/or 3 , respectively, cause a surprising disintegration or lysis of the GUVs and cells, comparable with the action of tensides, viral fusion peptides, and host‐defense antimicrobial peptides. Possible sources of the chain‐length‐dependent destructive potential of the β3/β2‐nona‐ and β3/β2‐dodecapeptide derivatives, and a possible relationship with the phosphate‐to‐phosphate and hydrocarbon thicknesses of GUVs, and eukaryotic cells are discussed. Further investigations with other types of GUVs and of eukaryotic or prokaryotic cells will be necessary to elucidate the mechanism(s) of interaction of ‘mixed’ β3/β2‐peptides with membranes and to evaluate possible biomedical applications.  相似文献   

20.
5α‐Androst‐16‐en‐3α‐ol (α‐androstenol) is an important contributor to human axilla sweat odor. It is assumed that α‐andostenol is excreted from the apocrine glands via a H2O‐soluble conjugate, and this precursor was formally characterized in this study for the first time in human sweat. The possible H2O‐soluble precursors, sulfate and glucuronide derivatives, were synthesized as analytical standards, i.e., α‐androstenol, β‐androstenol sulfates, 5α‐androsta‐5,16‐dien‐3β‐ol (β‐androstadienol) sulfate, α‐androstenol β‐glucuronide, α‐androstenol α‐glucuronide, β‐androstadienol β‐glucuronide, and α‐androstenol β‐glucuronide furanose. The occurrence of α‐androstenol β‐glucuronide was established by ultra performance liquid chromatography (UPLC)/MS (heated electrospray ionization (HESI)) in negative‐ion mode in pooled human sweat, containing eccrine and apocrine secretions and collected from 25 female and 24 male underarms. Its concentration was of 79 ng/ml in female secretions and 241 ng/ml in male secretions. The release of α‐androstenol was observed after incubation of the sterile human sweat or α‐androstenol β‐glucuronide with a commercial glucuronidase enzyme, the urine‐isolated bacteria Streptococcus agalactiae, and the skin bacteria Staphylococcus warneri DSM 20316, Staphylococcus haemolyticus DSM 20263, and Propionibacterium acnes ATCC 6919, reported to have β‐glucuronidase activities. We demonstrated that if α‐ and β‐androstenols and androstadienol sulfates were present in human sweat, their concentrations would be too low to be considered as potential precursors of malodors; therefore, the H2O‐soluble precursor of α‐androstenol in apocrine secretion should be a β‐glucuronide.  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