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.
Abstract: This review briefly surveys the conformational properties of guest ω‐amino acid residues when incorporated into host α‐peptide sequences. The results presented focus primarily on the use of β‐ and γ‐residues in αω sequences. The insertion of additional methylene groups into peptide backbones enhances the range of accessible conformations, introducing additional torsional variables. A nomenclature system, which permits ready comparisons between α‐peptides and hybrid sequences, is defined. Crystal structure determination of hybrid peptides, which adopt helical and β‐hairpin conformations permits the characterization of backbone conformational parameters for β‐ and γ‐residues inserted into regular α‐polypeptide structures. Substituted β‐ and γ‐residues are more limited in the range of accessible conformation than their unsubstituted counterparts. The achiral β,β‐disubstituted γ‐amino acid, gabapentin, is an example of a stereochemically constrained residue in which the torsion angles about the Cβ–Cγ (θ1) and Cα–Cβ (θ2) bonds are restricted to the gauche conformation. Hybrid sequences permit the design of novel hydrogen bonded rings in peptide structures.  相似文献   

2.
Abstract: Replacement of Phe3 in the endogenous δ‐opioid selective peptide deltorphin I with four optically pure stereoisomers of the topographically constrained, highly hydrophobic novel amino acid β‐isopropylphenylalanine (β‐iPrPhe) produced four pharmacologically different deltorphin I peptidomimetics. Radiolabeled ligand‐binding assays and in vitro biological evaluation indicate that the stereoconfiguration of the iPrPhe residue plays a crucial role in determining the binding affinity, bioactivity and selectivity of [β‐iPrPhe3]deltorphin I analogs: a (2S,3R) configuration of the iPrPhe3 residue in [β‐iPrPhe3]deltorphin I provided the most desirable biological properties with binding affinity (IC50 = 2 n m ), bioassay potency (IC50 = 1.23 n m in MVD assay) and exceptional selectivity for the δ‐opioid receptor over the µ‐opioid receptor (30 000). Further conformational studies based on two‐dimensional NMR and computer‐assisted molecular modeling suggested a model for the possible bioactive conformation in which the Tyr1 and (2S,3R)‐β‐iPrPhe3 residues adopt trans side‐chain conformations, and the linear peptide backbone favors a distorted β‐turn conformation.  相似文献   

3.
Abstract: Alzheimer's disease, Huntington's disease and prion diseases are part of a growing list of diseases associated with formation of β‐sheet containing fibrils. In a previous publication, we demonstrated that the self‐association of the Alzheimer's β‐amyloid (Aβ) peptide is inhibited by peptides homologous to the central core domain of Aβ, but containing N‐methyl amino acids at alternate positions. When these inhibitor peptides are arrayed in an extended, β‐strand conformation, the alternating position of N‐methyl amino acids gives the peptide two distinct faces, one exhibiting a normal pattern of peptide backbone hydrogen bonds, but the other face having limited hydrogen‐bonding capabilities due to the replacement of the amide protons by N‐methyl groups. Here, we demonstrate, through two‐dimensional NMR and circular dichroic spectroscopy, that a pentapeptide with two N‐methyl amino acids, Aβ16–20m or Ac‐K(Me)LV(Me)FF‐NH2, does indeed have the intended structure of an extended β‐strand. This structure is remarkably stable to changes in solvent conditions and resists denaturation by heating, changes in pH (from 2.5 to 10.5), and addition of denaturants such as urea and guanindine‐HCl. We also show that this peptide, despite its hydrophobic composition, is highly water soluble, to concentrations > 30 mm , in contrast to the nonmethylated congener, Aβ16–20 (Ac‐KLVFF‐NH2). The striking water solubility, in combination with the hydrophobic composition of the peptide, suggested that the peptide might be able to pass spontaneously through cell membranes and model phospholipid bilayers such as unilamellar vesicles. Thus, we also demonstrate that this peptide is indeed able to pass spontaneously through both synthetic phospholipid bilayer vesicles and cell membranes. Characterization of the biophysical properties of the Aβ16–20m peptide may facilitate the application of this strategy to other systems as diverse as the HIV protease and chemokines, in which there is dimerization through β‐strand domains.  相似文献   

4.
Abstract: The chemical synthesis and X‐ray crystal structure analysis of a model peptide incorporating a conformationally flexible β‐Ala residue: Boc‐β‐Ala‐Pda, 1 (C23H46N2O3: molecular weight = 398.62) have been described. The peptide crystallized in the crystal system triclinic with space group P21: a = 5.116(3) Å, b = 5.6770(10) Å, c = 21.744(5) Å; α = 87.45°, β = 86.87°, γ = 90.0°; Z = 1. An attractive feature of the crystal molecular structure of 1 is the induction of a reasonably extended backbone conformation of the β‐Ala moiety, i.e. the torsion angles φ ≈ ?115°, µ ≈ 173° and ψ ≈ 122°, correspond to skew?, trans and skew+ conformation, respectively, by an unbranched hydrophobic alkyl chain, Pda, which prefers an all‐anti orientation (θ1 ≈ ?153°, θ2 ≈ … θ14 ≈ ±178°). The observation is remarkable because, systematic conformational investigations of short linear β‐Ala peptides of the type Boc‐β‐Ala‐Xaa‐OCH3 (Xaa = Aib or Acc6) have shown that the chemical and stereochemical characters of the neighboring moieties may be critical in dictating the overall folded and/or unfolded conformational features of the β‐Ala residue. The overall conformation of 1 is typical of a ‘bar’. It appears convincing that, in addition to a number of hydrophobic contacts between the parallel arranged molecules, an array of conventional N‐H…O=C intermolecular H‐bonding interactions stabilize the crystal molecular structure. Moreover, the resulting 14‐membered pseudo‐ring motif, generated by the amide–amide interactions between the adjacent molecules, is completely devoid of nonconventional C?H…O interaction. The potentials of the conformational adaptation of the β‐Ala residue, to influence and stabilize different structural characteristics have been highlighted.  相似文献   

5.
The use of peptides as drugs in pharmaceutical applications is hindered by their susceptibility to proteolysis and therefore low bioavailability. β‐Peptides that contain an additional methylene group in the backbone, are gaining recognition from a pharmaceutical stand point as they are considerably more resilient to proteolysis and metabolism. Recently, we reported two new classes of β ‐peptides, β 3‐ and β2‐peptides derived from l ‐aspartic acid and l ‐diaminopropionic acid, respectively. Here, we report the proteolytic stability of these β‐peptidic compounds and a mixed α /β‐peptide against three enzymes (pronase, trypsin and elastase), as well as, human serum. The stability of these peptides was compared to an α‐peptide. Peptides containing β‐linkages were resistant to all conditions. The mixed α /β‐peptide, however, exhibited proteolysis in the presence of trypsin and pronase but not elastase. The rate of degradation of the mixed α /β‐peptide was slower than that would be expected for an α‐peptide. In addition, these β‐peptides were not toxic to HeLa and COS‐1 cell lines as observed by MTT cytotoxicity assay. These results expand the scope of mixed α /β‐peptides containing β‐amino acids or small β‐peptide fragments as therapeutic peptides.  相似文献   

6.
Mimetics of β‐turn structures in proteins have been used to calibrate the relative reactivities toward deamidation of asparagine residues in the two central positions of a β‐turn and in a random coil. N‐Acetyl‐Asn‐Gly‐6‐aminocaproic acid, an acyclic analog of a β‐turn mimic undergoes deamidation of the asparaginyl residue through a succinimide intermediate to generate N‐acetyl‐Asp‐N‐Gly‐6‐aminocaproic acid (6‐aminocaproic acid, hereafter Aca) and N‐acetyl‐l ‐iso‐aspartyl (isoAsp)‐Gly‐Aca (pH 8.8, 37 °C) ≈ 3‐fold faster than does the cyclic β‐turn mimic cyclo‐[L‐Asn‐Gly‐Aca] with asparagine at position 2 of the β‐turn. The latter compound, in turn, undergoes deamidation ≈ 30‐fold faster than its positional isomer cyclo‐[Gly‐Asn‐Aca] with asparagine at position 3 of the β‐turn. Both cyclic peptides assume predominantly β‐turn structures in solution, as demonstrated by NMR and circular dichroism characterization. The open‐chain compound and its isomer N‐acetyl‐Gly‐Asn‐Aca assume predominantly random coil structures. The latter isomer undergoes deamidation 2‐fold slower than the former. Thus the order of reactivity toward deamidation is: asparagine in a random coil ≈ 3× asparagine in position 2 of a β‐turn ≈ 30× asparagine in position 3 of a β‐turn.  相似文献   

7.
Abstract: The chaperone activity and biophysical properties of the 19 amino acid peptide DFVIFLDVKHFSPEDLTVK, identified as the functional element in αA‐crystallin and here referred to as mini‐αA‐crystallin, were studied using light scattering and spectroscopic methods after altering its sequence and enantiomerism. The all‐d and all‐l conformers of the peptide do not show marked differences in their chaperone‐like activity against heat‐induced aggregation of alcohol dehydrogenase at 48°C and dithiothreitol‐induced aggregation of insulin. The retro peptide does not show any secondary structure and is also unable to act like a chaperone. Both all‐l and all‐d peptides lose their β‐sheet conformations, hydrophobicity and chaperone‐like activity at temperatures > 50°C. However, upon cooling, a significant portion of those properties was regained, suggesting temperature‐dependent, reversible structural alterations in the peptides under investigation. We propose that both the hydrophobicity and β‐sheet conformation of the functional element of αA‐crystallin are essential for chaperone‐like activity.  相似文献   

8.
Abstract: Crystal structure analysis of a model peptide: Boc‐β‐Ala‐Aib‐β‐Ala‐NHCH3 (β‐Ala: 3‐amino propionic acid; Aib: α‐aminoisobutyric acid) revealed distinct conformational preferences for folded [φ≈136°, µ ≈ ?62°, ψ ≈100°] and semifolded [φ ≈ 83°, µ ≈ ?177°, ψ ≈ ?117°] structures of the N‐ and C‐terminus β‐Ala residues, respectively. The overall folded conformation is stabilized by unusual Ni···H‐Ni+1 and nonconventional C–H···O intramolecular hydrogen bonding interactions.  相似文献   

9.
Abstract: The characterization of a four‐stranded β‐sheet structure in a designed 26‐residue peptide Beta‐4 is described. The sequence of Beta‐4 (Arg‐Gly‐Thr‐Ile‐Lys‐Dpro‐Gly‐Ile‐Thr‐Phe‐Ala‐DPro‐Ala‐Thr‐Val‐Leu‐Phe‐Ala‐Val‐DPro‐Gly‐Lys‐Thr‐Leu‐Tyr‐Arg) was chosen such that three strategically positioned DPro‐Xxx segments nucleate type II′β‐turns, which facilitate hairpin extension. A four‐stranded β‐sheet structure is determined in methanol from 500 MHz 1H NMR data using a total of 100 observed NOEs, 11 dihedral restraints obtained from vicinal JCαH‐NH values and 10 hydrogen bonding constraints obtained from H/D exchange data. The observed NOEs provide strong evidence for a stable four‐stranded sheet and a nonpolar cluster involving Ile8, Phe10, Val15 and Phe17. Circular dichroism studies in water–methanol mixtures provide evidence for melting of the β‐sheet structure at high water concentrations. NMR analysis establishes that the four‐stranded sheet in Beta‐4 is appreciably populated in 50% (v/v) aqueous methanol. In water, the peptide structure is disorganized, although the three β‐turn nuclei appear to be maintained.  相似文献   

10.
Alzheimer's disease is most common neurodegenerative disorder and is characterized by increased production of soluble amyloid‐β oligomers, the main toxic species predominantly formed from aggregation of monomeric amyloid‐β (Aβ). Increased production of Aβ invokes a cascade of oxidative damages to neurons and eventually leads to neuronal death. This study was aimed to investigate the neuroprotective effects of a β‐sheet breaker α/β‐hybrid peptide (BSBHp) and the underlying mechanisms against Aβ40‐induced neurotoxicity in human neuroblastoma SH‐SY5Y cells. Cells were pretreated with the peptide Aβ40 to induce neurotoxicity. Assays for cell viability, cell membrane damage, cellular apoptosis, generation of reactive oxygen species (ROS), intracellular free Ca2+, and key apoptotic protein levels were performed in vitro. Our results showed that pretreatment with BSBHp significantly attenuates Aβ40‐induced toxicity by retaining cell viability, suppressing generation of ROS, Ca2+ levels, and effectively protects neuronal apoptosis by suppressing pro‐apoptotic protein Bax and up‐regulating antiapoptotic protein Bcl‐2. These results suggest that α/β‐hybrid peptide has neuroprotective effects against Aβ40‐induced oxidative stress, which might be a potential therapeutic agent for treating or preventing neurodegenerative diseases.  相似文献   

11.
Abstract: The structural perturbation induced by CαH→Nα exchange in azaamino acid‐containing peptides was predicted by ab initio calculation of the 6‐31G* and 3‐21G* levels. The global energy‐minimum conformations for model compounds, For‐azaXaa‐NH2 (Xaa = Gly, Ala, Leu) appeared to be the β‐turn motif with a dihedral angle of φ = ± 90°, ψ = 0°. This suggests that incorporation of the azaXaa residue into the i + 2 position of designed peptides could stabilize the β‐turn structure. The model azaLeu‐containing peptide, Boc‐Phe‐azaLeu‐Ala‐OMe, which is predicted to adopt a β‐turn conformation was designed and synthesized in order to experimentally elucidate the role of the azaamino acid residue. Its structural preference in organic solvents was investigated using 1H NMR, molecular modelling and IR spectroscopy. The temperature coefficients of amide protons, the characteristic NOE patterns, the restrained molecular dynamics simulation and IR spectroscopy defined the dihedral angles [ (φi+1, ψi+1) (φi+2, ψi+2)] of the Phe‐azaLeu fragment in the model peptide, Boc‐Phe‐azaLeu‐Ala‐OMe, as [(?59°, 127°) (107°, ?4°)]. This solution conformation supports a βII‐turn structural preference in azaLeu‐containing peptides as predicted by the quantum chemical calculation. Therefore, intercalation of the azaamino acid residue into the i + 2 position in synthetic peptides is expected to provide a stable β‐turn formation, and this could be utilized in the design of new peptidomimetics adopting a β‐turn scaffold.  相似文献   

12.
Abstract: Rs‐AFPs are antifungal proteins, isolated from radish (Raphanus sativus) seed or leaves, which consist of 50 or 51 amino acids and belong to the plant defensin family of proteins. Four highly homologous Rs‐AFPs have been isolated (Rs‐AFP1–4). The structure of Rs‐AFP1 consists of three β‐strands and an α‐helix, and is stabilized by four cystine bridges. Small peptides deduced from the native sequence, still having biological activity, are not only important tools to study structure?function relationships, but may also constitute a commercially interesting target. In an earlier study, we showed that the antifungal activity of Rs‐AFP2 is concentrated mainly in the β2?β3 loop. In this study, we synthesized linear 19‐mer peptides, spanning the entire β2?β3 loop, that were found to be almost as potent as Rs‐AFP2. Cysteines, highly conserved in the native protein, are essential for maintaining the secondary structure of the protein. Surprisingly, in the 19‐mer loop peptides, cysteines can be replaced by α‐aminobutyric acid, which even improves the antifungal potency of the peptides. Analogous cyclic 19‐mer peptides, forced to adopt a hairpin structure by the introduction of one or two non‐native disulfide bridges, were also found to possess high antifungal activity. The synthetic 19‐mer peptides, like Rs‐AFP2 itself, cause increased Ca2+ influx in pregerminated fungal hyphae.  相似文献   

13.
Abstract: A brief survey of the history of peptide chemistry from Theodore Curtius to Emil Fischer to Bruce Merrifield is first presented. The discovery and development of peptide ligation, i.e. of actual chemical synthesis of proteins are described. In the main chapter, ‘ Synthesis of Proteins by Chemical Ligation ’ a detailed discussion of the principles, reactivities and mechanisms involved in the various coupling strategies now applied (ligation, chemical ligation, native chemical ligation) is given. These include coupling sites with cysteine and methionine (as well as the seleno analogs), histidine, glycine and pseudo‐prolines, ‘unrestricted’ amino‐acid residues (using the Staudinger reaction), as well as solid‐phase segment coupling by thioligation of unprotected peptides. In another section, ‘ Synthesis of β‐peptides by Thioligation ’, couplings involving β2‐ and β3‐peptides are described (with experimental details).  相似文献   

14.
Abstract: To assess the relative importance of backbone hydrogen bonding (H‐bonding) vs. side chain hydrophobicity in protein structural formation, a method called side chain‐backbone swap is proposed. Such a method swaps the side chain and backbone portions of certain amino acid residues, such as Asp, Glu, Asn, Gln, Lys, and Arg. Such a swap retains the sequence of a polypeptide and preserves the identity of the backbone linkage. On the other hand, the swap disrupts backbone H‐bonding geometry because of the introduction of extra methylene groups into the peptide backbone. In this project, we chose the two‐stranded α‐helical coiled‐coil to implement side chain‐backbone swap. A pair of 36‐residue peptides was designed. The two peptides have identical sequence with four residues in each heptad repeat occupied by glutamyl residues. Each glutamic acid was incorporated either as α‐glutamyl residue (the peptide is denoted as α‐Glu‐36) or as γ‐glutamyl residue (the peptide is denoted as γ‐Glu‐36). The inter‐conversion between the two peptides constitutes a side chain‐backbone swap. Residues constituting the hydrophobic core of the coiled‐coil, however, are left unchanged. The peptide pair was characterized by circular dichroism spectroscopy, reversed‐phase liquid chromatography (RPLC), and two‐dimensional nuclear magnetic resonance (NMR). The results indicate that α‐Glu‐36 is a two‐stranded α‐helical coiled‐coil while γ‐Glu‐36 lacks stable structural elements. It is concluded that, at least for coiled‐coils where hydrophobic interactions are predominantly long‐range, local backbone H‐bonding is a required for structural formation, consistent with a hierarchic folding mechanism. The methodological implication of side chain‐backbone swap is also discussed.  相似文献   

15.
Alzheimer's disease is a neurodegenerative disorder associated with amyloid‐β (Aβ) fibrillation. N‐Methylated amyloid‐β peptides are potent inhibitors of amyloid‐β fibrillation. We investigated the inhibitory effect of N‐Methylated Aβ30–40 peptides on Aβ1–40 fibrillation. N‐Methylated Aβ30–40 peptides affected the fibrillation, and this effect was dependent on the concentration of N‐Methylated peptide and the number and position of N‐Methylated groups. N‐Methylated Aβ30–40 peptides were co‐aggregated with Aβ1–40. Spectroscopic technique was adopted to investigate an origin of the observed dependence. Suppression of thioflavin T (ThT) fluorescence count was correlated with the dissociation constant Kd of monomer–dimer equilibrium of each N‐Methylated Aβ30–40 peptide. Monomeric N‐Methylated peptides decreased ThT fluorescence count during Aβ1–40 fibrillation. Secondary structure content was not largely different between Aβ1–40 fibrils and co‐aggregates. These results suggested that N‐Methylated Aβ30–40 peptides disrupted the regular β‐sheet structure of Aβ1–40 fibrils and affected the ThT fluorescence count. The monomer–dimer equilibrium of N‐Methylated peptides was (partly) responsible for the observed dependence of their inhibitory effect on the concentration of N‐Methylated peptide and the number and position of N‐Methylated groups. Our study provides a hint to design new N‐Methylated inhibitor peptides of fibrillation.  相似文献   

16.
Abstract: A protected tridehydropeptide containing (Z)‐β‐(3‐pyridyl)‐α,β‐dehydroalanine (ΔZ3Pal) residue, Boc‐Leu‐ΔZ3Pal‐Leu‐OMe ( 1 ), was synthesized via Erlenmeyer azlactone method. X‐ray crystallographic analysis revealed that the peptide 1 adopts an extended conformation, which is similar to that of a ΔZPhe analog, Boc‐Leu‐ΔZPhe‐Leu‐OMe ( 2 ).  相似文献   

17.
Abstract: A model peptide AAGDYY‐NH2 (B1), which is found to adopt a β‐turn conformation in the TEM‐1 β‐lactamase inhibitor protein (BLIP) in the TEM‐1/BLIP co‐crystal, was synthesized to elucidate the mechanism of its β‐turn formation and stability. Its structural preferences in solution were comprehensively characterized using CD, FT‐IR and 1H NMR spectroscopy, respectively. The set of observed diagnostic NOEs, the restrained molecular dynamics simulation, CD and FT‐IR spectroscopy confirmed the formation of a β‐turn in solution by the model peptide. The dihedral angles [(φ3, ?3) (φ4, ?4)] of [(?52°, ?32°) (?38°, ?44°)] of Gly‐Asp fragment in the model peptide are consistent with those of a type III β‐turn. In a conclusion, the conformational preference of the linear hexapeptide B1 in solution was determined, and it would provide a simple template to study the mechanism of β‐turn formation and stability.  相似文献   

18.
Aβ42 peptide, with two additional residues at C‐terminus, aggregates much faster than Aβ40. We performed equilibrium replica‐exchange molecular dynamics simulations of their monomers using our residue‐specific force field. Simulated 3JHNHα‐coupling constants agree excellently with experimental data. Aβ40 and Aβ42 have very similar local conformational features, with considerable β‐strand structures in the segments: A2‐H6 ( A ), L17‐A21 ( B ), A30‐V36 ( C ) of both peptides and V39‐I41 ( D ) of Aβ42. Both peptides have abundant A‐B and B‐C contacts, but Aβ40 has much more contacts between A and C than Aβ42, which may retard its aggregation. Only Aβ42 has considerable A ‐ B ‐ C ‐ D topology. Decreased probability of A ‐ C contact in Aβ42 relates to the competition from C ‐ D contact. Increased A ‐ C contact probability may also explain the slower aggregation of A2T and A2V mutants of Aβ42.  相似文献   

19.
Solution conformations of three series of model peptides, homochiral Ac-Pro-L-Xaa-NHCH3 and heterochiral Ac-Pro-D-Xaa-NHcH3 (Xaa = Val, Phe, Leu, Abu. Ah) as well as αβ-unsaturated Ac-Pro-ΔXaa-NHCH3 [Δ Xaa =ΔVal, (Z)-ΔPhe, (Z)-ΔLeu, (Z)-ΔAbu] were investigated in CDCl3 and CH2Cl2 by 1H-, 13C-NMR, and FTIR spectroscopy. NH stretching absorption spectra, solvent shifts Δδ for NH (Xaa) and NHCH3 on going from CDCl3 to (CD3)2SO, diagnostic interresidue proton NOEs, and trans-cis isomer ratios were examined. These studies performed showed the essential difference in conformational propensities between homochiral peptides (L-Xaa) on the one hand and heterochiral (D-Xaa) and αβ-dehydropeptides (ΔXaa) on the other. Former compounds are conformationally flexible with an inverse γ-bend, a β-turn, and open forms in an equilibrium depending on the nature of the Xaa side chain. Conformational preferences of heterochiral and αβ-dehydropeptides are very similar, with the type-II β-turn as the dominating structure. There is no apparent correlation between conformational properties and the nature of the Xaa side chain within the two groups. The β-turn formation propensity seems to be somewhat greater in αβ-unsaturated than in heterochiral peptides, but an estimation of β-folded conformers is risky.  相似文献   

20.
Abstract: Gramicidin S (GS) is a 10‐residue cyclic β‐sheet peptide with lytic activity against the membranes of both microbial and human cells, i.e. it possesses little to no biologic specificity for either cell type. Structure–activity studies of de novo‐designed 14‐residue cyclic peptides based on GS have previously shown that higher specificity against microbial membranes, i.e. a high therapeutic index (TI), can be achieved by the replacement of a single l ‐amino acid with its corresponding d ‐enantiomer [Kondejewski, L.H. et al. (1999) J. Biol. Chem. 274 , 13181]. The diastereomer with a d ‐Lys substituted at position 4 caused the greatest improvement in specificity vs. other l to d substitutions within the cyclic 14‐residue peptide GS14, through a combination of decreased peptide amphipathicity and disrupted β‐sheet structure in aqueous conditions [McInnes, C. et al. (2000) J. Biol. Chem. 275 , 14287]. Based on this information, we have created a series of peptide diastereomers substituted only at position 4 by a d ‐ or l ‐amino acid (Leu, Phe, Tyr, Asn, Lys, and achiral Gly). The amino acids chosen in this study represent a range of hydrophobicities/hydrophilicities as a subset of the 20 naturally occurring amino acids. While the d ‐ and l ‐substitutions of Leu, Phe, and Tyr all resulted in strong hemolytic activity, the substitutions of hydrophilic d ‐amino acids d ‐Lys and d ‐Asn in GS14 at position 4 resulted in weaker hemolytic activity than in the l ‐diastereomers, which demonstrated strong hemolysis. All of the l ‐substitutions also resulted in poor antimicrobial activity and an extremely low TI, while the antimicrobial activity of the d ‐substituted peptides tended to improve based on the hydrophilicity of the residue. d ‐Lys was the most polar and most efficacious substitution, resulting in the highest TI. Interestingly, the hydrophobic d ‐amino acid substitutions had superior antimicrobial activity vs. the l ‐enantiomers although substitution of a hydrophobic d ‐amino acid increases the nonpolar face hydrophobicity. These results further support the role of hydrophobicity of the nonpolar face as a major influence on microbial specificity, but also highlights the importance of a disrupted β‐sheet structure on antimicrobial activity.  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