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 78 毫秒
1.
We synthesized by solution-phase methods three analogues, [L-Leu(6)-OMe], [L-(alphaMe)Leu(3), L-Leu(6)-OMe], and [L-(alphaMe)Val(4), L-Leu(6)-OMe] of halovir A. The [L-Leu(6)-OMe] analogue is known to be biologically equipotent to its naturally occurring, antiviral, lipopentapeptide amide parent compound. The preferred conformations of the L-(alphaMe)Leu- and L-(alphaMe)Val-containing analogues, with a potentially reinforced helicity, were compared with those of [L-Leu(6)-OMe] halovir A and the natural peptide itself by use of a combination of FT-IR absorption and NMR techniques. Measurements of the antiviral activities against herpes simplex virus type-1 (HSV-1) of halovir A and its three analogues were also carried out. Interestingly, the [L-(alphaMe)Val(4), L-Leu(6)-OMe] analogue exhibits the most significant activity in reducing HSV-1 infectivity, notably higher than that of halovir A itself.  相似文献   

2.
Nandel FS  Khare B 《Biopolymers》2005,77(1):63-73
Conformational studies of the peptides constructed from achiral amino acid residues Aib and Delta(Z)Phe (I) Ac-Aib-Delta(Z)Phe-NHMe (II), and Ac-(Aib-Delta(Z)Phe)(3)-NHMe; peptides III-VI having L-Leu or D-Leu at either the N- or the C-terminal position and of peptides VII-X having Leu residues in different enantiomeric combinations at both the N- and the C-terminal positions in peptide II have been studied to design the peptide with the required helical sense. Peptide II, as expected, adopts degenerate left- and right-handed helical structures. It has been shown that the peptides IV and VI having D-Leu at either the N or the C terminus can be realized in the right-handed helical structure with the phi,psi values of -20 degrees and -60 degrees for the Aib/Delta(Z)Phe residues. L-Leu and D- Leu at both the terminals in peptides VII and VIII, respectively, have hardly any effect as both the left- and the right-handed structures are found to be degenerate. Peptides III and IX can be realized in right- and left-handed helical structures, respectively, in solvents of low polarity whereas peptides V and X are predicted to be in the right-handed helical structures stabilized by carbonyl-carbonyl interactions without the formation of hydrogen bonds. The conformational states with the phi,psi values of 0 degrees and -85 degrees in peptide V are characterized by rise per residue of 2.03 A, rotation per residue of 117.5 degrees , and 3.06 residues per turn. In all peptides having Leu residue at the N terminus, the methyl moiety of the acetyl group is involved in the CH/pi interactions with the Cepsilon--Cdelta edge of the aromatic ring of Delta(Z)Phe (3) and the amino group NH of Delta(Z)Phe is involved in the NH/pi interactions with its own aromatic ring. The CH(3) groups of the Aib residues are also involved in CH/pi interactions with the i + 1th and i + 3th Delta(Z)Phe's aromatic side chains.  相似文献   

3.
A series of oligo-peptide based catalysts were prepared using Fmoc solid-phase peptide synthesis. It was found that peptides with N-terminal proline residues catalyzed an aldol reaction yielding enantiomeric enriched product. Peptide H-Pro-Glu-Leu-Phe-OH catalyzed the reaction with good activity and moderate enantioselectivity (66% ee). Furthermore, it was shown that an acidic side chain and/or C-termini are essential to catalysis.  相似文献   

4.
A bacterial strain (ECU1001) capable of utilizing phenyl glycidyl ether as sole carbon source and energy source was isolated from soil samples through two steps of screening and was identified as a Bacillus megaterium. The epoxide hydrolase from Bacillus megaterium ECU1001 was biosynthesized in parallel with cell growth and a maximum activity of 31.0 U/l was reached after 30 h of culture when the biomass (DCW) was 9.1 g/l. A temperature of 35°C and pH 8.0 were optimal for the bioconversion. The lyophilized whole cells of Bacillus megaterium ECU1001 could preferentially hydrolyze the (R)-enantiomer of phenyl glycidyl ether, yeilding (S)-epoxide and (R)-diol with high enantioselectivity (E=47.8). The (S)-enantiomer of the epoxide remained in the reaction mixture with >99.5% ee (enantiomeric excess) at a conversion of 55.9%. The substrate concentration could be increased up to 60 mM without affecting the ee and (S)-phenyl glycidyl ether could be obtained with an optical purity of 100% ee and 25.6% yield. Therefore, the method is potentially useful for the preparative resolution of epoxides.  相似文献   

5.
As an additional step toward the dissection of the factors responsible for the onset of 3(10)-helix vs alpha-helix in peptides, in this paper we describe the results of a three-dimensional (3D) structural analysis by x-ray diffraction of the N(alpha)-acylated heptapeptide alkylamide mBrBz-L-Iva-L-(alphaMe)Val-L-Abu-L-(alphaMe)Val-L-(alphaMe)Phe-L-(alphaMe)Val-L-Iva-NHMe characterized by a single (L-Abu3) C(alpha)-trisubstituted and six C(alpha)-tetrasubstituted alpha-amino acids. We find that in the crystal state this peptide is folded in a mixed helical structure with short elements of 3(10)-helix at either terminus and a central region of alpha-helix. This finding, taken together with the published NMR and x-ray diffraction data on the all C(alpha)-methylated parent sequence and its L-Val2 analog (also the latter heptapeptide has a single C(alpha)-trisubstituted alpha-amino acid) strongly supports the view that one C(alpha)-trisubstituted alpha-amino acid inserted near the N-terminus of an N(alpha)-acylated heptapeptide alkylamide sequence may be enough to switch a regular 3(10)-helix into an essentially alpha-helical conformation. As a corollary of this work, the x-ray diffraction structure of the N(alpha)-protected, C-terminal tetrapeptide alkylamide Z-L-(alphaMe)Val-L-(alphaMe)Phe-L-(alphaMe)Val-L-Iva-NHMe, also reported here, is clearly indicative of the preference of this fully C(alpha)-methylated, short peptide for the 3(10)-helix. As the same terminally blocked sequence is mixed 3(10)/alpha-helical in the L-Abu3 heptapeptide amide but regular 3(10)-helical in the tetrapeptide amide and in the parent heptapeptide amide, these results point to an evident plasticity even of a fully C(alpha)-methylated short peptide.  相似文献   

6.
The conformational behaviour of deltaZPhe has been investigated in the model dipeptide Ac-deltaZPhe-NHMe and in the model tripeptides Ac-X-deltaZPhe-NHMe with X=Gly,Ala,Val,Leu,Abu,Aib and Phe and is found to be quite different. In the model tripeptides with X=Ala,Val,Leu,Abu,Phe the most stable structure corresponds to phi1=-30 degrees, psi1=120 degrees and phi2=psi2=30 degrees. This structure is stabilized by the hydrogen bond formation between C=O of acetyl group and the NH of the amide group, resulting in the formation of a 10-membered ring but not a 3(10) helical structure. In the peptides Ac-Aib-deltaZPhe-NHMe and Ac-(Aib-deltaZPhe)3-NHMe, the helical conformers with phi = +/-30 degrees, psi = +/-60 degrees for Aib residue and phi=psi= +/-30 degrees for deltaZPhe are predicted to be most stable. The computational studies for the positional preferences of deltaZPhe residue in the peptide containing one deltaZPhe and nine Ala residues reveal the formation of a 3(10) helical structure in all the cases with terminal preferences for deltaZPhe. The conformational behaviour of Ac-(deltaZPhe)n-NHMe with n< or =4 is predicted to be very labile. With n > 4, degenerate conformational states with phi,psi values of 0 degrees +/- 90 degrees adopt helical structures which are stabilized by carbonyl-carbonyl interactions and the N-H-pi interactions between the amino group of every deltaZPhe residue with one C-C edge of its own phenyl ring. The results are in agreement with the experimental finding that screw sense of helix for peptides containing deltaZPhe residues is ambiguous in solution. The helical structures stabilized by hydrogen bond formation are found to be at least 3kCalmol(-1) less stable. Conformational studies have also been carried out for the peptide Ac-(deltaEPhe)6-NHMe and the peptide Ac-deltaAla-(deltaZPhe)6-NHMe containing deltaAla residue at the N-terminal. The N-H-pi interactions are absent in peptide Ac-(deltaEPhe)6-NHMe.  相似文献   

7.
In continuation of our studies on the determination of the structural features of functionalized peptides in solution by combining time-resolved fluorescence data and molecular mechanics results, the conformational features of a series of linear, L-(alphaMe)Val-based peptides have been investigated in methanol. These foldamers have the general formula F[(alphaMe)Val](r)-T-[(alphaMe)Val](2)NHtBu, where (alphaMe)Val = C(alpha)-methylvaline and r = 0-3, while F [= fluoren-9-ylmethoxycarbonyl (Fmoc)] and T [= 2,2,6,6-tetramethylpiperidine-1-oxyl-4-amino-carboxylic (Toac)] are a fluorophoric N(alpha)-protecting group and a nitroxide-based alpha-amino acid quencher, respectively. According to ir and CD spectra, the longest term of the series (r = 3) attains a 3(10)-helical structure, while the other peptides populate an intramolecularly H-bonded, 3(10)-helix-like conformation affected by dynamic helical distortions, which are enhanced by the shortness of the backbone chain. Such distortions are reflected in both the energy of the stretching mode and the molar extinction coefficient of the H-bonded N-H groups, the former being higher and the latter smaller than those of a stable 3(10)-helix. Steady-state and time-resolved fluorescence measurements in methanol show a strong quenching of Fmoc by the Toac residue, located at different helix positions, depending on the r value. Comparison of quenching efficiencies and lifetime preexponents with those theoretically obtained from the deepest energy minimum conformers, assuming a F?rster mechanism, is satisfactory. The computed structures exhibit a rather compact arrangement, which accounts for the few sterically favored conformations for each peptide, in full agreement with the time-resolved fluorescence data. Orientational effects between the probes must be taken into account for a correct interpretation of the fluorescence decay results, implying that interconversion among conformational substates involving the probes is slower than the energy transfer rate.  相似文献   

8.
The crystal-state conformations of two octapeptides, pBrBz-(D-Iva)8-OtBu (8I) and Ac-[L-(alphaMe)Val]8-OH (8II), the heptapeptide Z-[L-(alphaMe)Val]7-OH (7), the hexapeptide Z-[L-(alphaMe)Leu]6-OtBu (6) and the tetrapeptide alkylamide Z-(Aib)2-L-Glu(OMe)-L-Ala-L-Lol (5) were assessed by x-ray diffraction analyses. Two independent molecules are observed in the asymmetric unit of each L-(alphaMe)Val homo-peptide. All four homo-peptides are folded in a regular 3(10)-helical structure (only the C-terminal H-bonded conformation of the D-Iva octapeptide is distorted to a type-I beta-turn). The hydroxyl groups of the C-terminal carboxyl moieties of the two L-(alphaMe)Val homo-peptides participate in an oxy-analogue of the type-III beta-turn conformation. While the two L-(alphaMe)Val 3(10)-helices are right-handed, the D-Iva and L-(alphaMe)Leu helices are left-handed. The tetrapeptide alkylamide is 3(10)-helical at the N-terminus, but it is mixed 3(10)/alpha-helical at the C-terminus.  相似文献   

9.
Microbial isolates from biofilters and petroleum-polluted bioremediation sites were screened for the presence of enantioselective epoxide hydrolases active towards tert-butyl glycidyl ether, benzyl glycidyl ether, and allyl glycidyl ether. Out of 270 isolated strains, which comprised bacteria, yeasts, and filamentous fungi, four were selected based on the enantioselectivities of their epoxide hydrolases determined in biotransformation reactions. The enzyme of Aspergillus niger M200 preferentially hydrolyses (S)-tert-butyl glycidyl ether to (S)-3-tert-butoxy-1,2-propanediol with a relatively high enantioselectivity (the enantiomeric ratio E is about 30 at a reaction temperature of 28 °C). Epoxide hydrolases of Rhodotorula mucilaginosa M002 and Rhodococcus fascians M022 hydrolyse benzyl glycidyl ether with relatively low enantioselectivities, the former reacting predominantly with the (S)-enantiomer, the latter preferring the (R)-enantiomer. Enzymatic hydrolysis of allyl glycidyl ether by Cryptococcus laurentii M001 proceeds with low enantioselectivity (E = 3). (R)-tert-Butyl glycidyl ether with an enantiomeric excess (ee) of over 99%, and (S)-3-tert-butoxy-1,2-propanediol with an ee-value of 86% have been prepared on a gram-scale using whole cells of A. niger M200. An enantiomeric ratio of approximately 100 has been determined under optimised biotransformation conditions with the partially purified epoxide hydrolase from A. niger M200. The regioselectivity of this enzyme was determined to be total for both (S)-tert-butyl glycidyl ether and (R)-tert-butyl glycidyl ether.  相似文献   

10.
In previous work, no chiral differences were found between D and L enantiomers of Leu in their ability to displace one another from the acid-extractable pool in mammalian cells. Recent evidence suggested otherwise. Our aim is to examine whether, in physiological range, D-amino acids have an equivalent ability to displace L-amino acids from the acid-extractable pool of HeLa cells, and vice versa. In the millimolar range, D-Leu and L-Leu have similar uptake and displacement properties with regard to the acid-extractable pool in HeLa cells, despite only the latter isomer being incorporated into protein. Below millimolar concentrations however, a distinct difference was found in the displacement of tritium-labelled L-Leu from the pool by unlabelled D-Leu compared with unlabelled L-Leu. Thus, unlabelled L-Leu in the external medium at 10−4 or 10−5 M displaced an equivalent amount of label from the pool as D-Leu introduced at a concentration approx. one order of magnitude higher, respectively. Reciprocal experiments, in which the acid-extractable pool was preloaded with 3H-D-Leu, confirmed this finding. The chirality difference was noted whether pool prelabelling was carried out at 37 or 0°C; but in order to avoid the complications of active transport mechanisms, the competition work reported here was done at 0°C. Similar chirality differences were observed with other hydrophobic amino acids, including His, Ile and Phe, such as, preferential displacement by the L-Leu racemer compared with the D-Leu racemer below mM levels. This was also true for the D and L forms of the non-utilisable isomer of Leu, norleucine (nLeu). We conclude that D-forms of hydrophobic amino acids have lower affinity for similar or the same intracellular binding sites involved in the acid-extractable pool than their L-forms. The significance of these chirality findings to amino acid pools in cells, and to the predominance of L-forms of amino acids in the biosphere is considered.  相似文献   

11.
Potent, broadly HIV-1 neutralizing antibodies (nAbs) may be invaluable for the design of an AIDS vaccine. 4E10 is the broadest HIV-1 nAb known to date and recognizes a contiguous and highly conserved helical epitope in the membrane-proximal region of gp41. The 4E10 epitope is thus an excellent target for vaccine design as it is also highly amenable to peptide engineering to enhance its helical character. To investigate the structural effect of both increasing the peptide length and of introducing helix-promoting constraints in the 4E10 epitope, we have determined crystal structures of Fab 4E10 bound to an optimized peptide epitope (NWFDITNWLWYIKKKK-NH(2)), an Aib-constrained peptide epitope (NWFDITNAibLWRR-NH(2)), and a thioether-linked peptide (NWFCITOWLWKKKK-NH(2)) to resolutions of 1.7 A, 2.1 A, and 2.2 A, respectively. The thioether-linked peptide is the first reported structure of a cyclic tethered helical peptide bound to an antibody. The introduced helix constraints limit the conformational flexibility of the peptides without affecting interactions with 4E10. The substantial increase in affinity (10 nM versus 10(4) nM of the IC(50) of the original KGND peptide template) is largely realized by 4E10 interaction with an additional helical turn at the peptide C terminus that includes Leu679 and Trp680. Thus, the core 4E10 epitope was extended and modified to a WFX(I/L)(T/S)XX(L/I)W motif, where X does not play a major role in 4E10 binding and can be used to introduce helical-promoting constraints in the peptide epitope.  相似文献   

12.
高对映选择性环氧化物水解酶产生菌的筛选及特性研究   总被引:6,自引:0,他引:6  
从土壤中分离的芽孢杆菌Bacillus megaterium ECU1001所产五氧化物水解酶能高对映选择性水解缩水甘油苯基醚(对映选择率E值可达47.8),当转化率为55.9%时,剩余的(S)-缩水甘油苯基醚的光学纯度(对映体过量值,ee)可达99.5%;当底物浓度提高到60mmol/L时,光学纯(S)-缩水基油苯基醚的收率达到25.6%。  相似文献   

13.
Epoxide hydrolase activity of Streptomyces strains   总被引:2,自引:0,他引:2  
The discovery of epoxide hydrolases within a Streptomyces sp. strain collection is described. Screening was performed in 96 well microtiter plates using a modified 4-(p-nitrobenzyl)pyridine assay with styrene oxide, 1,2-epoxy-hexane or 3-phenyl ethylglycidate (3-PEG) as substrates. Out of 120 strains investigated, S. antibioticus Tü4, S. arenae Tü495 and S. fradiae Tü27 exhibited epoxide hydrolase activity. These strains were further investigated by performing laboratory-scale biotransformations utilizing styrene oxide, 1,2-epoxy-hexane and 3-PEG followed by subsequent quantitative analysis employing chiral gas chromatography. The highest conversions were achieved with whole cells from S. antibioticus Tü4 in the presence of 10% (v/v) DMSO. However, enantioselectivity was only satisfying (E = 31) in the presence of 5% (v/v) acetone, which allowed isolation of optically pure non-hydrolyzed (R)-styrene oxide (99% enantiomeric excess (ee)) and (S)-phenyl-1,2-ethandiol (72% ee) at 55% conversion after 24 h. The resolution of 3-PEG proceeded with slightly lower enantioselectivity albeit higher reaction rates. With S. fradiae Tü27 and S. arenae Tü495 enantioselectivity towards styrene oxide was only E = 3-4.  相似文献   

14.
Using different stereoselective chemical and chemoenzymatic approaches we synthesized the chiral, Calpha-methylated alpha-amino acid L-(alphaMe)Nva with a short, linear side-chain. A set of terminally protected model peptides to the pentamer level containing either (alphaMe)Nva or Nva in combination with Ala and/or Aib was prepared using solution methods and characterized fully. Two (alphaMe)Nva peptides were also synthesized using side-chain hydrogenation of the corresponding Calpha-methyl, Calpha-allylglycine (Mag) peptides. A detailed solution and crystal-state conformational analysis based on FT-IR absorption, 1H NMR and X-ray diffraction techniques allowed us to define that: (i) (alphaMe)Nva is an effective beta-turn and 3(10)-helix former; and (ii) the relationship between (alphaMe)Nva chirality and the screw sense of the turn/helix formed is that typical of protein amino acids, i.e. L-(alphaMe)Nva induces the preferential formation of right-handed folded structures. In more general terms, this study reinforced previous conclusions that peptides based on alpha-amino acids with a Calpha-methyl substituent and a Calpha-linear alkyl substituent are characterized by a strong tendency to fold into turn and helical structures.  相似文献   

15.
Summary A series of amphiphilic, helical peptides was designed and synthesized to investigate the components necessary for formation of helical bundles with differing aggregation states. Minimalistic sequences were employed for the peptides which contained either four (Leu4), six (Leu6) or eight (Leu8) leucine residues within a sixteen amino acid sequence. All peptides were highly helical as evaluated by circular dichroism, and the helical content of each peptide exhibited a concentration dependence. Size exclusion chromatography confirmed aggregation states of dimer/trimer forLeu4, tetramer forLeu6, and hexamer octamer forLeu8. Disulfide crosslinking studies also confirmed that the dimer ofLeu4 favored a parallel orientation with respect to the helical dipole. This systematic study clearly defines the role of hydrophobicity in the self assembly of helical peptides; peptides with a small hydrophobic face favor small bundle sizes, whereas peptides containing larger hydrophobic faces form correspondingly larger helical bundles.  相似文献   

16.
Be X  Hong Y  Wei J  Androphy EJ  Chen JJ  Baleja JD 《Biochemistry》2001,40(5):1293-1299
E6AP is a cellular protein that binds cancer-related papillomaviral E6 proteins. The E6 binding domain, called E6ap, is located on an 18-amino acid segment of E6AP. The corresponding peptide was synthesized and its structure determined by nuclear magnetic resonance spectroscopy. The overall structure of the peptide is helical. A consensus E6-binding sequence among different E6 interacting proteins contains three conserved hydrophobic residues. In the structure of the E6AP peptide, the three conserved leucines (Leu 9, Leu 12, and Leu 13) form a hydrophobic patch on one face of the alpha-helix. Substitution of any of these leucines with alanine abolished binding to E6 protein, indicating that the entire hydrophobic patch is necessary. Mutation of a glutamate to proline, but not alanine, also disrupted the interaction between E6 and E6AP protein, suggesting that the E6-binding motif of the E6AP protein must be helical when bound to E6. Comparison of the E6ap structure and mutational results with those of another E6-binding protein (E6BP/ERC-55) indicates the existence of a general E6-binding motif.  相似文献   

17.
Four diastereomeric‐Leu‐Leu‐Aib‐Leu‐Leu‐Aib‐peptides, Boc‐D ‐Leu‐L ‐Leu‐Aib‐L ‐Leu‐L ‐Leu‐Aib‐OMe (1), Boc‐L ‐Leu‐D ‐Leu‐Aib‐L ‐Leu‐L ‐Leu‐Aib‐OMe (2), Boc‐L ‐Leu‐L ‐Leu‐Aib‐D ‐Leu‐L ‐Leu‐Aib‐OMe (3), and Boc‐L ‐Leu‐L ‐Leu‐Aib‐L ‐Leu‐D ‐Leu‐Aib‐OMe (4), were synthesized. The crystals of the four hexapeptides were characterized by X‐ray crystallographic analysis. Two diastereomeric hexapeptides 1 and 2 having D ‐Leu(1) or D ‐Leu(2) were folded into right‐handed (P) 3 10 ‐helical structures, while peptide 3 having D ‐Leu(4) was folded into a turn structure nucleated by type III′ and I$' \bf{\beta}$ ‐turns, and peptide 4 having D ‐Leu(5) was folded into a left‐handed (M) 3 10 ‐helical structure. Copyright © 2011 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

18.
The effects of the hydrophobicity and the distribution of hydrophobic residues on the surfaces of some designed alpha-helical transmembrane peptides (acetyl-K2-L(m)-A(n)-K2-amide, where m + n = 24) on their solution behavior and interactions with phospholipids were examined. We find that although these peptides exhibit strong alpha-helix forming propensities in water, membrane-mimetic media, and lipid model membranes, the stability of the helices decreases as the Leu content decreases. Also, their binding to reversed phase high-performance liquid chromatography columns is largely determined by their hydrophobicity and generally decreases with decreases in the Leu/Ala ratio. However, the retention of these peptides by such columns is also affected by the distribution of hydrophobic residues on their helical surfaces, being further enhanced when peptide helical hydrophobic moments are increased by clustering hydrophobic residues on one side of the helix. This clustering of hydrophobic residues also increases peptide propensity for self-aggregation in aqueous media and enhances partitioning of the peptide into lipid bilayer membranes. We also find that the peptides LA3LA2 [acetyl-K2-(LAAALAA)3LAA-K2-amide] and particularly LA6 [acetyl-K2-(LAAAAAA)3LAA-K2-amide] associate less strongly with and perturb the thermotropic phase behavior of phosphatidylcholine bilayers much less than peptides with higher L/A ratios. These results are consistent with free energies calculated for the partitioning of these peptides between water and phospholipid bilayers, which suggest that LA3LA2 has an equal tendency to partition into water and into the hydrophobic core of phospholipid model membranes, whereas LA6 should strongly prefer the aqueous phase. We conclude that for alpha-helical peptides of this type, Leu/Ala ratios of greater than 7/17 are required for stable transmembrane associations with phospholipid bilayers.  相似文献   

19.
To understand the terminal effect of chiral residue for determining a helical screw sense, we adopted five kinds of peptides IV containing N‐ and/or C‐terminal chiral Leu residue(s): Boc–L ‐Leu–(Aib–ΔPhe)2–Aib–OMe ( I ), Boc–(Aib–ΔPhe)2–L ‐Leu–OMe ( II ), Boc–L ‐Leu–(Aib–ΔPhe)2–L ‐Leu–OMe ( III ), Boc–D ‐Leu–(Aib–ΔPhe)2–L ‐Leu–OMe ( IV ), and Boc–D ‐Leu–(Aib–ΔPhe)2–Aib–OMe ( V ). The segment –(Aib–ΔPhe)2– was used for a backbone composed of two “enantiomeric” (left‐/right‐handed) helices. Actually, this could be confirmed by 1H‐nmr [nuclear Overhauser effect (NOE) and solvent accessibility of NH resonances] and CD spectroscopy on Boc–(Aib–ΔPhe)2–Aib–OMe, which took a left‐/right‐handed 310‐helix. Peptides IV were also found to take 310‐type helical conformations in CDCl3, from difference NOE measurement and solvent accessibility of NH resonances. Chloroform, acetonitrile, methanol, and tetrahydrofuran were used for CD measurement. The CD spectra of peptides IIII in all solvents showed marked exciton couplets with a positive peak at longer wavelengths, indicating that their main chains prefer a left‐handed screw sense over a right‐handed one. Peptide V in all solvents showed exciton couplets with a negative peak at longer wavelengths, indicating it prefers a right‐handed screw sense. Peptide IV in chloroform showed a nonsplit type CD pattern having only a small negative signal around 280 nm, meaning that left‐ and right‐handed helices should exist with almost the same content. In the other solvents, peptide IV showed exciton couplets with a negative peak at longer wavelengths, corresponding to a right‐handed screw sense. From conformational energy calculation and the above 1H‐nmr studies, an N‐ or C‐terminal L ‐Leu residue in the lowest energy left‐handed 310‐helical conformation was found to take an irregular conformation that deviates from a left‐handed helix. The positional effect of the L ‐residue on helical screw sense was discussed based on CD data of peptides IV and of Boc–(L ‐Leu–ΔPhe)n–L ‐Leu–OMe (n = 2 and 3). © 1999 John Wiley & Sons, Inc. Biopoly 49: 551–564, 1999  相似文献   

20.
Peptides representing both loop and the sixth transmembrane regions of the α-factor receptor of Saccharomyces cerevisiae were synthesized by solid-phase procedures and purified to near homogeneity. CD, nmr, and modeling analysis indicated that in aqueous media the first extracellular loop peptide E1(107–125), the third intracellular loop peptide I3(231–243), and the carboxyl terminus peptide I4(350–372) were mostly disordered. In contrast, the second extracellular loop peptide E2(191–206) assumed a well-defined structure in aqueous medium and the sixth transmembrane domain peptide receptor M6(252-269, C252A) was highly helical in trifluoroethanol/water (4:1), exhibiting a kink at Pro258. A synthetic peptide containing a sequence similar to that of the sixth transmembrane domain of a constitutively active α-factor receptor M6(252–269, C252A, P258L) in which Leu replaces Pro258 exhibited significantly different biophysical properties than the wild-type sequence. In particular, this peptide had very low solubility and gave CD resembling that of a β-sheet structure in hexafluoroacetone/water (1:1) whereas the wild-type peptide was partially helical under identical conditions. These results would be consistent with the hypothesis that the constitutive activity of the mutant receptor is linked to a conformational change in the sixth transmembrane domain. The study of the receptor segments also indicate that peptides corresponding to loops of the α-factor receptor do not appear to assume turn structures. © 1998 John Wiley & Sons, Inc. Biopoly 46: 343–357, 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