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 547 毫秒
1.
The depsipeptide Boc(1)-Leu(2)-Leu(3)-Ala(4)-Leu(5)-Leu(6)-Lac(7)-Leu(8)-Leu(9)-Lac(10)-Leu(11)-Leu(12)-Lac(13)-Leu(14)-Leu(15)-OEt(16) (1) (Boc = tert-butyloxycarbonyl, Lac = L-lactic acid residue) has been synthesized from the peptide Boc-Leu-Leu-Ala-OEt (2) and a depsipeptide, Boc-(Leu-Leu-Lac)(3)-Leu-Leu-OEt (3). Single crystals of 1 were successfully obtained and the structure has been solved by direct methods (such as Sir2002 and Shake-and-Bake). Interestingly, 1 adopts an alpha/3(10)-conjugated helix containing a kink at the junction of peptide and depsipeptide segments, Leu3-Lac7. This is significantly different from the conformation of 3, which has a straight alpha-helical structure with standard phi and psi angles. Microcrystalline CD spectra were also studied to compare structural properties of 1 and 3. The differences between alpha/3(10)- and alpha-helices appear in these CD spectra.  相似文献   

2.
Ohyama T  Oku H  Yoshida M  Katakai R 《Biopolymers》2001,58(7):636-642
A sequential polydepsipeptide, Boc-(Leu-Leu-Lac)3-Leu-Leu-OEt (1) (Lac = L-lactic acid residue) has been synthesized by the segment condensation method. The sequential unit of 1, -Leu-Leu-Lac-, is consisted of two amino acid residues and one hydroxy acid residue. X-ray diffraction measurement with an imaging plate detector and a direct-methods procedure of Shake-and-Bake successfully revealed the crystal structure of 1. In the solid state, the 11-mer depsipeptide, 1, have clear alpha-helical conformation even with the three ester linkages.  相似文献   

3.
The crystal-state conformations of the hexapeptide amide Pht-(Aib)(6)-NH-C(CH(3))(2)-O-OtBu (7), the hexapeptide Ac-L-aIle-(Aib)(5)-OtBu (6), the pentapeptide Z-(Aib)(3)-L-Glu(OtBu)-Aib-O-(CH(2))(2)-(1)Nap (5), the tetrapeptides Z-(Aib)(2)-L-His(N(tau)-Trt)-Aib-OMe (4 I) and Z-(Aib)(2)-L-Nva-Aib-OtBu (4 II), the tripeptide Pyr-(Aib)(3)-OtBu (3 I), the dipeptide amides Pyr-(Aib)(2)-(4)NH-TEMPO (3 II) and Piv-(Aib)(2)-NH-C(CH(3))(2)-O-OtBu (3 III), and the dipeptides Pht-Aib-betaAc(6)c-OtBu (2 I), Pht-Aib-NH-C(CH(3))(2)-O-OtBu (2 II) and Boc-gGly-mAib-OH (2 III) have been determined by X-ray diffraction analyses. All peptides investigated are characterized by one or more turn/helix forming Aib residues. Except the three short dipeptides, all are folded into C==O...H--N intramolecularly H-bonded 3(10)-helices, or into various types of beta-turns. In the structure of 6, two independent molecules of opposite screw sense were observed in the asymmetric unit, generating diastereomeric 3(10)-helices.  相似文献   

4.
The relative stability of alpha-helix and beta-sheet secondary structure in the solid state was investigated using poly(L-alanine) (PLA) as a model system. Protein folding and stability has been well studied in solution, but little is known about solid-state environments, such as the core of a folded protein, where peptide packing interactions are the dominant factor in determining structural stability. (13)C cross-polarization with magic angle spinning (CPMAS) NMR spectroscopy was used to determine the backbone conformation of solid powder samples of 15-kDa and 21.4-kDa PLA before and after various sample treatments. Reprecipitation from helix-inducing solvents traps the alpha-helical conformation of PLA, although the method of reprecipitation also affects the conformational distribution. Grinding converts the secondary structure of PLA to a final steady-state mixture of 55% beta-sheet and 45% alpha-helix at room temperature regardless of the initial secondary structure. Grinding PLA at liquid nitrogen temperatures leads to a similar steady-state mixture with 60% beta-sheet and 40% alpha-helix, indicating that mechanical shear force is sufficient to induce secondary structure interconversion. Cooling the sample in liquid nitrogen or subjecting it to high pressure has no effect on secondary structure. Heating the sample without grinding results in equilibration of secondary structure to 50% alpha-helix/50% beta-sheet at 100 degrees C when starting from a mostly alpha-helical state. No change was observed upon heating a beta-sheet sample, perhaps due to kinetic effects and the different heating rate used in the experiments. These results are consistent with beta-sheet approximately 260 J/mol more stable than alpha-helix in solid-state PLA.  相似文献   

5.
N-Acetyl-(E)-dehydrophenylalanine N'-methylamide [Ac-(E)-DeltaPhe-NHMe], one of a few representative (E)-alpha,beta-dehydroamino acids, was studied by FTIR in dichloromethane and acetonitrile. To support spectroscopic interpretations and to gain some deeper insight into the Ac-(E)-DeltaPhe-NHMe molecule, the Ramachandran potential energy surface was calculated by the B3LYP/6-31G*//HF/3-21G method and the conformers localized were fully optimized at the B3LYP/6-31 + G** level. The spectra and calculations were compared with those of the related molecules Ac-DeltaAla-NHMe and Ac-(Z)-DeltaPhe-NHMe. The title compound assumes two conformational states in equilibrium in dichloromethane solution with a predominance of the extended conformer E. The Ac-(E)-DeltaPhe-NHMe spectrum is like that of Ac-DeltaAla-NHMe, particularly in the region of bands AI and AII, and unlike that of Ac-(Z)-DeltaPhe-NHMe. The positions of bands AI and II together with the nu(s)(N1--H1) band proves that the conformers E of both DeltaAla and (E)-DeltaPhe compounds are stabilized by the quite strong C5 hydrogen bonds N1--H1...O2. The same conclusion is drawn from the Ramachandran diagrams. The conformers E of both compounds are placed in the global minima and the gaps in energy order between them and the second conformer are large. The conformers E of DeltaAla and (E)-DeltaPhe, apart from the N1--H1...O2 hydrogen bond, show the Cbeta--H...O1 interaction, and Ac-(E)-DeltaPhe-NHMe displays the NH/pi interaction with the N2--H2 projecting in the first carbon atom of the phenyl ring. The C5 hydrogen bond is stronger in (E)-DeltaPhe than that in the DeltaAla compound. This is in agreement with interactions found in the calculated structures and can be explained by the influence of the phenyl ring in position (E). In acetonitrile, the molecule of Ac-(E)-DeltaPhe-NHMe loses its C5 hydrogen bond and becomes unfolded, whereas that of Ac-DeltaAla-NHMe does not vary practically. Adopting conformation E in a non-polar solvent seems to be a general feature of the (E)-DeltaXaa residues.  相似文献   

6.
A structural transition from a 3(10)-helix to an alpha-helix has been characterized at high resolution for an octapeptide segment located in 3 different sequences. Three synthetic peptides, decapeptide (A) Boc-Aib-Trp-(Leu-Aib-Ala)2-Phe-Aib-OMe, nonapeptide (B) Boc-Trp-(Leu-Aib-Ala)2-Phe-Aib-OMe, and octapeptide (C) Boc-(Leu-Aib-Ala)2-Phe-Aib-OMe, are completely helical in their respective crystals. At 0.9 A resolution, R factors for A, B, and C are 8.3%, 5.4%, and 7.3%, respectively. The octapeptide and nonapeptide form ideal 3(10)-helices with average torsional angles phi(N-C alpha) and psi(C alpha-C') of -57 degrees, -26 degrees C and -60 degrees, -27 degrees for B. The 10-residue peptide (A) begins as a 3(10)-helix and abruptly changes to an alpha-helix at carbonyl O(3), which is the acceptor for both a 4-->1 hydrogen bond with N(6)H and a 5-->1 hydrogen with N(7)H, even though the last 8 residues have the same sequence in all 3 peptides. The average phi, psi angles in the decapeptide are -58 degrees, -28 degrees for residues 1-3 and -63 degrees, -41 degrees for residues 4-10. The packing of helices in the crystals does not provide any obvious reason for the transition in helix type. Fourier transform infrared studies in the solid state also provide evidence for a 3(10)- to alpha-helix transition with the amide I band appearing at 1,656-1,657 cm-1 in the 9- and 10-residue peptides, whereas in shorter sequences the band is observed at 1,667 cm-1.  相似文献   

7.
The effect of insertion of lactic acid (Lac) residues into peptide helices has been probed using specifically designed sequences. The crystal structures of 11-residue and 14-residue depsipeptides Boc-Val-Val-Ala-Leu-Val-Lac-Leu-Aib-Val-Ala-Leu-OMe (1) and Boc-Val-Ala-Leu-Aib-Val-Ala-Leu-Val-Lac-Leu-Aib-Val-Ala-Leu-OMe (3), containing centrally positioned Lac residues, have been determined. The structure of an 11-residue peptide Boc-Val-Ala-Leu-Aib-Val-Ala-Leu-Aib-Val-Ala-Leu-OMe (2), analog of a which is an amide previously determined Lac-containing depsipeptide, Boc-Val-Ala-Leu-Aib-Val-Lac-Leu-Aib-Val-Ala-Leu-OMe I. L. Karle, C. Das, and P. Balaram, Biopolymers, Vol. 59, (2001) pp. 276-289], is also reported. Peptide 1 adopts a helical fold, which is stabilized by mixture of 4-->1 and 5-->1 hydrogen bonds. Peptide 2 adopts a completely alpha-helical conformation stabilized by eight successive 5-->1 hydrogen bonds. Peptide 3 appears to be predominately alpha-helical, with seven 5-->1 hydrogen bonds and three 4-->1 interaction interspersed in the sequence. In the structure of peptide 3 in addition to water molecules in the head-to-tail region, hydration at an internal segment of the helix is also observed. A comparison of five related peptide helices, containing a single Lac residue, reveals that the hydroxy acid can be comfortably accommodated at interior positions in the helix, with the closest C=O...O distances lying between 2.8 and 3.3 A.  相似文献   

8.
The design, synthesis, conformational, crystallographic, and ion transport studies of 30-membered, proline containing depsipeptides that incorporate the rigid low molecular weight lipophilic adamantane (Adm) building blocks are reported. The adamantyl groups provide the desired membrane permeability and conformational constraint for efficient transport in lipid membranes. The novel cyclic depsipeptides are: c[--Adm--C(O)--Pro-- O--CH(2)-- CHR--NH--C(O)--Pro--C(O)-- Adm--C(O)--Pro--C(O)--NH--CHR--CH(2)-- O--Pro--C(O)--] where R==H for A and R==CONH--Adm for B. Crystal structure analysis of A established that the two peptide segments are identical in formula and in conformation and that the peptides are bonded to the interleaving Adm at the 1 and 3 positions. However, the complete ring is highly asymmetric in shape since bonds for both Peptide-Adm-Peptide segments have the syn-anti motif. Torsional angles for the connecting bonds to Adm are -162 degrees , +71 degrees and -169 degrees , -48 degrees . The irregular clamshell shape of the molecule has three internal C==O moieties directed in a manner that could provide three Na(+)--O ligands. While A exhibited negligible transport of Na(+) ions across membranes, peptide B endowed with two additional adamantanes in the periphery did transport Na(+) ions from outside to inside.  相似文献   

9.
Ganesh S  Jayakumar R 《Biopolymers》2003,70(3):336-345
Self-assembly of two tripeptide derivatives containing three nonpolar isoleucine moieties and polar oxyethylene groups are studied in methanol. Peptide A [CH3(OCH2CH2)3OCH2CO(Ile)3OCH3] and peptide B [CH3(OCH2CH2)3OCH2CO(Ile)3NH (CH2CH2O)3CH3] take a mixture of unordered and helical conformation at low concentration (8.5 x 10(-4) M). However, at high concentration (2 x 10(-3) M), both the peptide showed significant increase in the helical conformation. An interesting conformational transition of peptides A and B at various methanol contents was observed in the solvated films of these compounds by spectroscopic methods like the far-uv circular dichroism and Fourier transform infrared (FT-IR) techniques. Peptide B, which contains more polar oxyethylene groups than A, showed a highly cooperative conformational transition when the methanol content was decreased. This transition was characterized by a large increase of beta-sheet, retaining a alpha-helical contribution. Peptide A showed a conformational transition resulting in a beta-sheet in the aggregated state. From the CD spectra, the ratio in the ellipticity indicates that peptide B forms twisted antiparallel beta-sheet conformation, whereas peptide A takes a parallel beta-sheet conformation. The results obtained in this work indicates the role of polar derivatization on the conformational preference of peptides having similar sequence.  相似文献   

10.
An apolar helical decapeptide with different end groups, Boc- or Ac-, crystallizes in a completely parallel fashion for the Boc-analog and in an antiparallel fashion for the Ac-analog. In both crystals, the packing motif consists of rows of parallel molecules. In the Boc-crystals, adjacent rows assemble with the helix axes pointed in the same direction. In the Ac-crystals, adjacent rows assemble with the helix axes pointed in opposite directions. The conformations of the molecules in both crystals are quite similar, predominantly alpha-helical, except for the tryptophanyl side chain where chi 1 congruent to 60 degrees in the Boc- analog and congruent to 180 degrees in the Ac-analog. As a result, there is one lateral hydrogen bond between helices, N(1 epsilon)...O(7), in the Ac-analog. The structures do not provide a ready rationalization of packing preference in terms of side-chain interactions and do not support a major role for helix dipole interactions in determining helix orientation in crystals. The crystal parameters are as follow. Boc-analog: C60H97N11O13.C3H7OH, space group Pl with a = 10.250(3) A, b = 12.451(4) A, c = 15.077(6) A, alpha = 96.55(3) degrees, beta = 92.31(3) degrees, gamma = 106.37(3) degrees, Z = 1, R = 5.5% for 5581 data ([F] greater than 3.0 sigma(F)), resolution 0.89 A. Ac-analog: C57H91N11O12, space group P2(1) with a = 9.965(1) A, b = 19.707(3) A, c = 16.648(3) A, beta = 94.08(1), Z = 2, R = 7.2% for 2530 data ([F] greater than 3.0 sigma(F)), resolution 1.00 A.  相似文献   

11.
Karle IL  Urry DW 《Biopolymers》2005,77(4):198-204
Tropoelastin is a complex polymeric protein composed primarily of repeating segments of Val-Pro-Gly-Gly, Val-Pro-Gly-Val-Gly, and Ala-Pro-Gly-Val-Gly-Val that occurs in connective tissue and arteries. It has rubber-like extensible properties. A synthetic cyclic dodecapeptide, with a double repeat of the hexapeptide sequence, has been shown to undergo a reversible inverse temperature transition; that is, crystals grow at 60 degrees C and dissolve in the mother liquor upon cooling. An x-ray crystal structure analysis established that the cyclic backbone formed an elongated loop with a Pro-Gly, type II beta turn at both ends. Six internal cross strand NH...OC hydrogen bonds form between six NH donors and four O=C acceptors where two of the carbonyl O atoms are bifurcated acceptors. As a result, the molecule is pulled up into a corrugated profile. The corrugated loops form extended beta-sheets by additional intermolecular hydrogen bonds. An analysis of the dome region in a corrugated sheet suggests a reversible mechanism for extending and contracting the length of the whole molecule, akin to the motion of opening and closing an umbrella, caused by the motion of a water molecule with its associated hydrogen bonds acting as spokes. Crystal parameters: C44H72N12O12.3H2O, sp. gr. P2(1)2(1)2(1), a = 9.212 angstroms, b = 19.055 angstroms, c = 32.247 angstroms, d = 1.157 g/cm3.  相似文献   

12.
Fifteen years ago it was shown that an alpha-aminoisobutyric acid (Aib) residue is significantly more effective than an L-Pro or a D-amino acid residue in inducing beta-sheet disruption in short model peptides. As this secondary structure element is known to play a crucial role in the neuropathology of Alzheimer's disease, it was decided to check the effect of Aib (and other selected, helix inducer, C(alpha)-tetrasubstituted alpha-amino acids) on the beta-sheet conformation adopted by a protected pentapeptide related to the sequence 17-21 of the beta-amyloid peptide. By use of FT-IR absorption and 1H NMR techniques it was found that the strong self-association characterizing the pentapeptide molecules in weakly polar organic solvents is completely abolished by replacing a single residue with Aib or one of its congeners.  相似文献   

13.
In order to check our current knowledge on the principles involved in beta-hairpin formation, we have modified the sequence of a 3:5 beta-hairpin forming peptide with two different purposes, first to increase the stability of the formed 3:5 beta-hairpin, and second to convert the 3:5 beta-hairpin into a 2:2 beta-hairpin. The conformational behavior of the designed peptides was investigated in aqueous solution and in 30% trifluoroethanol (TFE) by analysis of the following nuclear magnetic resonance (NMR) parameters: nuclear Overhauser effect (NOE) data, and C(alpha)H, (13)C(alpha), and (13)C(beta) conformational shifts. From the differences in the ability to adopt beta-hairpin structures in these peptides, we have arrived to the following conclusions: (i) beta-Hairpin population increases with the statistical propensity of residues to occupy each turn position. (ii) The loop length, and in turn, the beta-hairpin type, can be modified as a function of the type of turn favored by the loop sequence. These two conclusions reinforce previous results about the importance of beta-turn sequence in beta-hairpin folding. (iii) Side-chain packing on each face of the beta-sheet may play a major role in beta-hairpin stability; hence simplified analysis in terms of isolated pair interactions and intrinsic beta-sheet propensities is insufficient. (iv) Contributions to beta-hairpin stability of turn and strand sequences are not completely independent. (v) The burial of hydrophobic surface upon beta-hairpin formation that, in turn, depends on side-chain packing also contributes to beta-hairpin stability. (vi) As previously observed, TFE stabilizes beta-hairpin structures, but the extent of the contribution of different factors to beta-hairpin formation is sometimes different in aqueous solution and in 30% TFE.  相似文献   

14.
Pathogenic prion proteins (PrP(Sc)) are thought to be produced by alpha-helical to beta-sheet conformational changes in the normal cellular prion proteins (PrP(C)) located solely in the caveolar compartments. In order to inquire into the possible conformational changes due to the influences of hydrophobic environments within caveolae, the secondary structures of prion protein peptides were studied in various kinds of detergents by CD spectra. The peptides studied were PrP(129-154) and PrP(192-213); the former is supposed to assume beta-sheets and the latter alpha-helices, in PrP(Sc). The secondary structure analyses for the CD spectra revealed that in buffer solutions, both PrP(129-154) and PrP(192-213) mainly adopted random-coils (approximately 60%), followed by beta-sheets (30%-40%). PrP(129-154) showed no changes in the secondary structures even in various kinds of detergents such as octyl-beta-D-glucopyranoside (OG), octy-beta-D-maltopyranoside (OM). sodium dodecyl sulfate (SDS), Zwittergent 3-14 (ZW) and dodecylphosphocholine (DPC). In contrast, PrP(192-213) changed its secondary structure depending on the concentration of the detergents. SDS, ZW, OG and OM increased the alpha-helical content, and decreased the beta-sheet and random-coil contents. DPC also increased the alpha-helical content, but to a lesser extent than did SDS, ZW, OG or OM. These results indicate that PrP(129-154) has a propensity to adopt predominantly beta-sheets. On the other hand, PrP(192-213) has a rather fickle propensity and varies its secondary structure depending on the environmental conditions. It is considered that the hydrophobic environments provided by these detergents may mimic those provided by gangliosides in caveolae, the head groups of which consist of oligosaccharide chains containing sialic acids. It is concluded that PrP(C) could be converted into a nascent PrP(Sc) having a transient PrP(Sc) like structureunder the hydrophobic environments produced by gangliosides.  相似文献   

15.
Using pulse radiolysis and laser flash photolysis, we have investigated the reactions of the deleterious species, e(-)(aq), HO&z.rad;, O(2)(*)(-) and O(2)((1)Delta(g)) with 10 water-soluble cyclopropyl-fused C(60) derivatives including a mono-adduct dendro[60]fullerene (d) and C(60) derivatives based on C(60)[C(COOH)(2)](n=2-6), some of which are known to be neuroprotective in vivo. The rate constants for reactions of e(-)(aq) and HO&z.rad; lie in the range 0.5-3.3 x 10(10) M(-1) s(-1). The d and bis-adduct monoanion radicals display sharp absorption peaks around 1000 nm (epsilon = 7 000-11 500 M(-1) cm(-1)); the anions of the tris-, tetra-, and penta-adduct derivatives have broader, weaker absorptions. The monohydroxylated radicals have their most intense absorption maxima around 390-440 nm (epsilon = 1000-3000 M(-1) cm(-1)). The anion and hydroxylated radical absorption spectra display a blue-shift as the number of addends increases. The radical anions react with oxygen (k approximately 10(7)-10(9) M(-1) s(-1)). The reaction of O(2)(*)(-) with the C(60) derivatives does not occur via an electron transfer. The rate constants for singlet oxygen reaction with the dendrofullerene and eee-derivative in D(2)O at pH 7.4 are k approximately 7 x 10(7) and approximately 2 x 10(7) M(-1) s(-1) respectively, in contrast to approximately 1.2 x 10(5) M(-1) s(-1) for the reaction with C(60) in C(6)D(6). The large acceleration of the rates for electron reduction and singlet oxygen reactions in water is due to a solvophobic process.  相似文献   

16.
Li HT  Du HN  Tang L  Hu J  Hu HY 《Biopolymers》2002,64(4):221-226
Amyloid-like aggregation of alpha-synuclein and deposit in Lewy bodies are thought to be the major cause of Parkinson's disease. Here we describe the secondary structural transformation and aggregation of human alpha-synuclein and its C-terminus truncated fragments in trifluoroethanol. Proteins containing the NAC (non-amyloid component) segment undergo a three-state transition: from native random coil to beta-sheet and to alpha-helical structure, while the NAC deficient fragment and gamma-synuclein undergo a typical two-state coil-to-alpha transition. The beta-sheet form is highly hydrophobic that strongly binds to 1-anilinonaphthalene-8-sulfonic acid (ANS) and is prone to self-aggregation. The results suggest that the NAC sequence is essential to beta-sheet formation and the aggregation originates from the beta-sheet intermediate, which may be implicated in the pathogenesis of Parkinson's disease.  相似文献   

17.
Model peptides based on -(Aib-Ala)(n)-, and (Aib)(n)-Leu-(Aib)(2) sequences, which have varying amounts of 3(10)-helical character, were studied by use of vibrational and electronic circular dichroism (VCD and ECD) and Fourier transform infrared (FTIR) absorption spectroscopies to test the correlation of spectral response and conformation. The data indicate that these peptides, starting from a length of about four to six residues, predominantly adopt a 3(10)-helical conformation at room temperature. The longest model peptides, depending on the series, may evidence some alpha-helical contribution to the spectra, while the shorter ones, with less than six residues, have much less order. The IR absorption spectra (as supported by theory) showed only small frequency changes between 3(10)- and alpha-helices. By contrast, solvent effects are a source of much bigger perturbations. The ECD results show that the intensity ratio for the approximately 222-nm to approximately 208-nm bands, while useful for distinguishing between these two helical types in some sequences, may have a narrower range of application than VCD. However, the VCD data presented here continue to support the proposed discrimination between alpha- and 3(10)-helices based on qualitative amide I and II bandshape differences. The present study shows the intensities of the 3(10)-helical amide I (peak-to-peak) to its amide II VCD to be of the same order and useful for discriminating them from alpha-helices, whose amide I dominates the amide II in intensity. This qualitative result is experimentally independent of the amount of alphaMe-substituted residues in the sequence. These experimental VCD results are consistent in detail with theoretical spectral simulations for Ac-(Ala)(8)-NH(2), Ac-(Aib-Ala)(4)-NH(2), and Ac-(Aib)(8)-NH(2) in 3(10)- and alpha-helical conformations.  相似文献   

18.
Oligopeptide-mediated helix stabilization of peptides in hydrophobic solutions was previously found by NMR and CD spectroscopic studies. The oligopeptide included the hydrophobic amino acids found in its parent peptide and were interposed by relevant basic oracidic amino acids. The strength of the interactions depended on the amino acid sequences. However, no helix-stabilizing effect was seen for the peptides in phosphate buffer solution, because the peptides assumed a random-coil structure. In order to ascertain whether the helix-stabilizing effect of an oligopeptide on its parent peptide could operate in aqueous solution, model peptides EK17 (Ac-AEAAAAEAAAKAAAAKA-NH2) and IFM17 (Ac-AEAAAAEIFMKAAAAKA-NH2) that may assume an alpha-helix in aqueous solutions were synthesized. Interactions were examined between various oligopeptides (EAAAK, KAAAE, EIFMK, KIFME, KIFMK and EYYEE) and EK17 or IFM17 in phosphate buffer and in 80% trifluoroethanol (TFE)-20% H2O solutions by CD spectra. EAAAK had little effect on the secondary structures of EK17 in both buffer and TFE solutions, while KAAAE, which has the reverse amino acid sequence of EAAAK, had a marked helix-destabilizing effect on EK17 in TFE. EIFMK and KIFME were found to stabilize the alpha-helical structure of EK17 in phosphate buffer solutions, whereas KIFMK and EYYEE destabilized the alpha-helical structure of EK17. EIFMK and KIFME had no effect on IFM17, because unexpectedly, IFM17 had appreciable amounts of beta-sheet structure in buffer solution. It was concluded that in order for the helix-stabilizing (1) the model peptide, the alpha-helical conformation of which is to be stabilized, should essentially assume an alpha-helical structure by nature, and (2) the hydrophobicity of the side-chains of the oligopeptide should be high enough for the oligopeptide to perform stable specific side chain-side chain intermolecular hydrophobic interactions with the model peptide.  相似文献   

19.
Li Q  Li F  Jia L  Li Y  Liu Y  Yu J  Fang Q  Cao A 《Biomacromolecules》2006,7(8):2377-2387
This study presents new investigations on chemical syntheses and characterization of new asymmetric AB(n)-shaped amphiphilic diblock methoxy poly(ethylene glycol)-b-[poly(l-lactide)](n), MPEG-b-(PLLA)(n) (n = 2, 4, and 8), bridged with dendritic ester linkages. First, a new series of A(OH)(n)-shaped hydroxy end-capped MPEG-(OH)(2), MPEG-(OH)(4), and MPEG-(OH)(8) bearing corresponding one- to three-generation dendritic ester moieties were efficiently derived from the starting MPEG (M(n) = 2 KDa) and 2,2'-bis(hydroxymethyl)propionic acid (Bis-HMPA) via ester coupling and a facile hydroxy protection-deprotection cycle, and then, chemical structures of these functional MPEG-(OH)(n) were characterized by nuclear magnetic resonance spectrometry (NMR) and MALDI-FTMS. Subsequently, by employing these MPEG-(OH)(n) as functional macroinitiators, new asymmetric AB(n)()-shaped amphiphilic MPEG-b-(PLLA)(2) S1, MPEG-b-(PLLA)(4) S2, and MPEG-b-(PLLA)(8) S3 bridged with dendritic Bis-HMPA ester linkages of L1-L3 as well as linear structural MPEG-b-PLLA references (R1-R3) were synthesized through the SnOct(2)-catalyzed ring-opening polymerization (ROP) of l-lactide at 130 degrees C in m-xylene solution, and their structures were further examined by NMR and gel permeation chromatography (GPC). It was demonstrated that the functional MPEG-(OH)(n) efficiently initiated the ROP of LLA, finally leading to successful formation of the AB(n)-shaped amphiphilic MPEG-b-(PLLA)(n) (n = 2, 4, and 8) with each PLLA arm weight close to 2 KDa and very narrow molecular weight distribution. Moreover, thermal history, crystallization, and spherulite morphologies were studied by means of differential scanning calorimeter (DSC), thermal gravimetric analyzer (TGA), and polarized microscope (POM) for these new structural amphiphilic S1-S3 as well as the linear R1-R3, intriguingly indicating a strong molecular architecture dependence of segmental crystallizability, spherulite morphology, and apparent crystal growth rate. Due to the favorable biodegradability and biocompatibility of the PLLA and MPEG, these results may therefore create new possibilities for these novel structural AB(n)-shaped amphiphilic MPEG-b-(PLLA)(n) as potential biomaterials.  相似文献   

20.
We synthesized three different lengths of poly(L-lysine) containing an -SH group at the terminal (PLL(n)-SH, n (polymerization degree) = 4, 10, 30) and adsorbed them on an Au(111) surface. To analyze the formation process and the structure of self-assembled monolayers (SAMs), we used atomic force microscopy (AFM) and Fourier transform infrared reflection absorption spectra (FT-IR RAS). At the initial stage of SAM growth, formation of nanosize domains was confirmed by AFM imaging. The alpha-helical PLL(30)-SH exhibited a well-defined SAM structure after adsorption reached equilibrium. The alpha-helical PLL(30)-SH was almost perpendicular to the gold surface and exhibited interesting molecular packing due to the secondary structure of PLL(30)-SH and the underlying Au(111) array. The tilt angle of the helix axis from the substrate normal was estimated to be about 50 degrees (AFM) and 44 degrees (FT-IR RAS) respectively. On the other hand, PLL(4)-SH and PLL(10)-SH formed beta-sheet-type SAMs on the Au(111) surface based on the structure determined by FT-IR RAS spectrum.  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