共查询到20条相似文献,搜索用时 0 毫秒
1.
Bourbigot S Dodd E Horwood C Cumby N Fardy L Welch WH Ramjan Z Sharma S Waring AJ Yeaman MR Booth V 《Biopolymers》2009,91(1):1-13
Topologically, platelet factor-4 kinocidins consist of distinct N-terminal extended, C-terminal helical, and interposing gamma-core structural domains. The C-terminal alpha-helices autonomously confer direct microbicidal activity, and the synthetic antimicrobial peptide RP-1 is modeled upon these domains. In this study, the structure of RP-1 was assessed using several complementary techniques. The high-resolution structure of RP-1 was determined by NMR in anionic sodium dodecyl sulfate (SDS) and zwitterionic dodecylphosphocholine (DPC) micelles, which approximate prokaryotic and eukaryotic membranes, respectively. NMR data indicate the peptide assumes an amphipathic alpha-helical backbone conformation in both micelle environments. However, small differences were observed in the side-chain orientations of lysine, tyrosine, and phenylalanine residues in SDS versus DPC environments. NMR experiments with a paramagnetic probe indicated differences in positioning of the peptide within the two micelle types. Molecular dynamics (MD) simulations of the peptide in both micelle types were also performed to add insight into the peptide/micelle interactions and to assess the validity of this technique to predict the structure of peptides in complex with micelles. MD independently predicted RP-1 to interact only peripherally with the DPC micelle, leaving its spherical shape intact. In contrast, RP-1 entered deeply into and significantly distorted the SDS micelle. Overall, the experimental and MD results support a preferential specificity of RP-1 for anionic membranes over zwitterionic membranes. This specificity likely derives from differences in RP-1 interaction with distinct lipid systems, including subtle differences in side chain orientations, rather than gross changes in RP-1 structure in the two lipid environments. 相似文献
2.
Mechanical stretching of secondary structures is studied through molecular dynamics simulations of a Go-like model. Force versus displacement curves are studied as a function of the stiffness and velocity of the pulling device. The succession of stretching events, as measured by the order in which contacts are ruptured, is compared to the sequencing of events during thermal folding and unfolding. Opposite cross-correlations are found for an alpha-helix and a beta-hairpin structure. In a tandem of two alpha-helices, the two constituent helices unravel nearly simultaneously. A simple condition for simultaneous versus sequential unraveling of repeat units is presented. 相似文献
3.
Two synthetic peptides, SNasealpha1 and SNasealpha2, corresponding to residues G55-I72 and K97-A109, respectively, of staphylococcal nuclease (SNase), are adopted for detecting the role of helix alpha1 (E57-A69) and helix alpha2 (M98-Q106) in the initiation of folding of SNase. The helix-forming tendencies of the two SNase peptide fragments are investigated using circular dichroism (CD) and two-dimensional (2D) nuclear magnetic resonance (NMR) methods in water and 40% trifluoroethanol (TFE) solutions. The coil-helix conformational transitions of the two peptides in the TFE-H2O mixture are different from each other. SNasealpha1 adopts a low population of localized helical conformation in water, and shows a gradual transition to helical conformation with increasing concentrations of TFE. SNasealpha2 is essentially unstructured in water, but undergoes a cooperative transition to a predominantly helical conformation at high TFE concentrations. Using the NMR data obtained in the presence of 40% TFE, an ensemble of alpha-helical structures has been calculated for both peptides in the absence of tertiary interactions. Analysis of all the experimental data available indicates that formation of ordered alpha-helical structures in the segments E57-A69 and M98-Q106 of SNase may require nonlocal interactions through transient contact with hydrophobic residues in other parts of the protein to stabilize the helical conformations in the folding. The folding of helix alpha1 is supposed to be effective in initiating protein folding. The formation of helix alpha2 depends strongly on the hydrophobic environment created in the protein folding, and is more important in the stabilization of the tertiary conformation of SNase. 相似文献
4.
Jane R Allison Marlen M��ller Wilfred F van Gunsteren 《Protein science : a publication of the Protein Society》2010,19(11):2186-2195
The right‐handed α‐helix is the dominant helical fold of α‐peptides, whereas the left‐handed 314‐helix is the dominant helical fold of β‐peptides. Using molecular dynamics simulations, the properties of α‐helical α‐peptides and 314‐helical β‐peptides with different C‐terminal protonation states and in the solvents water and methanol are compared. The observed energetic and entropic differences can be traced to differences in the polarity of the solvent‐accessible surface area and, in particular, the solute dipole moments, suggesting different reasons for their stability. 相似文献
5.
Agata Bahyrycz Yoshikatsu Matsubayashi Mari Ogawa Youji Sakagami Danuta Konopińska 《Journal of peptide science》2005,11(9):589-592
Phytosulfokine-alpha (PSK-alpha), a sulfated growth factor of structure H-Tyr(SO3H)-Ile-Tyr(SO3H)-Thr-Gln-OH universally found in both monocotyledons and dicotyledons, strongly promotes proliferation of plant cells in culture. In studies on the structure/activity relationship of PSK-alpha the synthesis was performed of a series of a further 23 analogues modified in position 1, 3 or 4 as well as simultaneously in positions 1 and 3 of the peptide chain. Peptides were synthesized by the solid phase method according to the Fmoc procedure on a Wang-resin. Free peptides were released from the resin by 95% TFA in the presence of EDT. All peptides were tested by competitive binding assay to the carrot membrane using 3H-labelled PSK-alpha according to the test of Matsubayashi et al. Among these peptide analogues, [H-Phe(4-Cl)1]-PSK-alpha (IV), [H-Phe(4-I)1]-PSK-alpha (VII), and [Phe(4-Cl)3]-PSK-alpha (XI) retained 30% PSK-alpha activity. Analogue [Tyr(PO3H2)3]-PSK-alpha (IX) showed 10% of PSK-alpha activity. 相似文献
6.
A systematic analysis of the hydrogen-bonding geometry in helices and beta sheets has been performed. The distances and angles between the backbone carbonyl O and amide N atoms were correlated considering more than 1500 protein chains in crystal structures determined to a resolution better than 1.5 A. They reveal statistically significant trends in the H-bond geometry across the different secondary structural elements. The analysis has been performed using Secbase, a modular extension of Relibase (Receptor Ligand Database) which integrates information about secondary structural elements assigned to individual protein structures with the various search facilities implemented into Relibase. A comparison of the mean hydrogen-bond distances in alpha helices and 3(10) helices of increasing length shows opposing trends. Whereas in alpha helices the mean H-bond distance shrinks with increasing helix length and turn number, the corresponding mean dimension in 3(10) helices expands in a comparable series. Comparing similarly the hydrogen-bond lengths in beta sheets there is no difference to be found between the mean H-bond length in antiparallel and parallel beta sheets along the strand direction. In contrast, an interesting systematic trend appears to be given for the hydrogen bonds perpendicular to the strands bridging across an extended sheet. With increasing number of accumulated strands, which results in a growing number of back-to-back piling hydrogen bonds across the strands, a slight decrease of the mean H-bond distance is apparent in parallel beta sheets whereas such trends are obviously not given in antiparallel beta sheets. This observation suggests that cooperative effects mutually polarizing spatially well-aligned hydrogen bonds are present either in alpha helices and parallel beta sheets whereas such influences seem to be lacking in 3(10) helices and antiparallel beta sheets. 相似文献
7.
Molecular dynamics as a tool to detect protein foldability. A mutant of domain B1 of protein G with non-native secondary structure propensities 下载免费PDF全文
The usefulness of molecular dynamics to assess the structural integrity of mutants containing several mutations has been investigated. Our goal was to determine whether molecular dynamics would be able to discriminate mutants of a protein having a close-to-wild-type fold, from those that are not folded under the same conditions. We used as a model the B1 domain of protein G in which we replaced the unique central alpha-helix by the sequence of the second beta-hairpin, which has a strong intrinsic propensity to form this secondary structure in solution. In the resulting protein, one-third of the secondary structure has been replaced by a non-native one. Models of the mutants were built based on the three-dimensional structure of the wild-type GB1 domain. During 2 ns of molecular dynamics simulations on these models, mutants containing up to 10 mutations in the helix retained the native fold, while another mutant with an additional mutation unfolded. This result is in agreement with our circular dichroism and NMR experiments, which indicated that the former mutants fold into a structure similar to the wild-type, as opposed to the latter mutant which is partly unfolded. Additionally, a mutant containing six mutations scattered through the surface of the domain, and which is unfolded, was also detected by the simulation. This study suggests that molecular dynamics calculations could be performed on molecular models of mutants of a protein to evaluate their foldability, prior to a mutagenesis experiment. 相似文献
8.
Helix-capping motifs are believed to play an important role in stabilizing alpha-helices and defining helix start and stop signals. We performed microsecond scale Brownian dynamics simulations to study ten XAAD sequences, with X = (A,E,I,L,N,Q,S,T,V,Y), to examine their propensity to form helix capping motifs and correlate these results with those obtained from analyzing a structural database of proteins. For the widely studied capping box motif S**D, where the asterisk can be any amino acid residue, the simulations suggested that one of the two hydrogen bonds proposed earlier as a stabilizing factor might not be as important. On the other hand, side-chain interactions between the capping residue and the third residue downstream on the polypeptide chain might also play a role in stabilizing this motif. These results are consistent with explicit-solvent molecular dynamics simulations of two capping box motifs found in the proteins BPTI and alpha-dendrotoxin. Principal component analysis of the SAAD trajectory showed that the first three principal components, after those corresponding to translational-rotational motion were removed, accounted for more than half of the conformational fluctuations. The first component separated the conformational space into two parts with the all-helical conformation and the capping box motif lying largely in one part. The second component, on the other hand, could be used to describe conformational transitions between the all-helical form and the capping box motif. 相似文献
9.
Bicudo TC Bicudo RC Forato LA Beltramini LM Batista LA Filho RB Colnago LA 《Biopolymers》2008,89(3):175-178
The proline-rich N-terminal domain of gamma-zein has been reported in relevant processes, which include its ability to cross the cell membranes. Evidences indicate that synthetic hexapeptide (PPPVHL), naturally found in N-terminal portion of gamma-zein, can adopt the polyproline II (PPII) conformation in aqueous solution. The secondary structure of gamma-zein in maize protein bodies had been analyzed by solid state Fourier transform infrared and nuclear magnetic resonance spectroscopies. However, it was not possible to measure PPII content in physiological environment since the beta-sheet and PPII signals overlap in both solid state techniques. Here, the secondary structure of gamma-zein has been analyzed by circular dichroism in SDS aqueous solution with and without ditiothreitol (DTT), and in 60% of 2-propanol and water with DTT. The results show that gamma-zein has high helical content in all solutions. The PPII conformation was present at about 7% only in water/DTT solution. 相似文献
10.
Douglas Poland 《Biopolymers》2002,63(1):59-65
We have recently shown that one can construct the enthalpy distribution for protein molecules from experimental knowledge of the temperature dependence of the heat capacity. For many proteins the enthalpy distribution evaluated at the midpoint of the denaturation transition (corresponding to the maximum in the heat capacity vs temperature curve) is broad and biphasic, indicating two different populations of molecules (native and unfolded) with distinctly different enthalpies. At temperatures above the denaturation point, the heat capacity for the unfolded state in many proteins is quite large and using the analysis just mentioned, we obtain a gaussian-like enthalpy distribution that is very broad. A large value of the heat capacity indicates that there are structural changes going on in the unfolded state above the transition temperature. In the present paper we investigate the origin of this large heat capacity by considering the presence of changing amounts of secondary structure (specifically, alpha-helix) in the unfolded state. For this purpose we use the empirical estimates of the Zimm-Bragg sigma and s factors for all of the native amino acids in water as determined by Scheraga and co-workers. Using myoglobin as an example, we calculate probability profiles and distribution functions for the total number of helix states in the specific-sequence molecule. Given the partition function for the specific-sequence molecule, we can then calculate a set of enthalpy moments for the molecule from which we obtain a good estimate of the enthalpy distribution in the unfolded state. This distribution turns out to be quite narrow when compared with the distribution obtained from the raw heat capacity data. We conclude that there must be other major structural changes (backbone and solvent) that are not accounted for by the inclusion of alpha-helix in the unfolded state. 相似文献
11.
Fourier transform infrared (FTIR) experiments in dimethylsulfoxide, a solvent incapable of H donation, demonstrate that H --> D isotopic replacement on the amide side of peptide bonds involves modifications of both the position and intensity of the amide I band. The effect of the isotopic substitution is particularly significant in the 1710-1670 and 1670-1650 cm(-1) regions, which are generally associated with beta-turns and alpha-helices. This behavior, attributed to the existence of intramolecular H-bonds in the polypeptide chain, is directly correlated to the presence of different secondary structures. Utilizing the effects induced by isotopic substitution, a method for the quantitative determination of the percentage of intramolecular H-bonds and the correlated secondary structures is proposed. The method consists of three principal steps: resolution of the fine structure of the amide I band with the determination of the number and position of the different components; reconstruction of the experimentally measured amide I band as a combination of Gaussian and Lorentzian functions, centered on the wave numbers set by band-narrowing methods, through a curve-fitting program; and quantitative determination of the population of the H-bonded carbonyls and the correlated secondary structures by comparison of the integrated intensities pertaining to the components with homologous wave numbers before and after isotopic exchange. The method is tested on a synthetic fragment of proocytocin that was previously analyzed by NMR techniques using the same solvent systems. 相似文献
12.
Loss of conformational entropy is one of the primary factors opposing protein folding. Both the backbone and side-chain of each residue in a protein will have their freedom of motion restricted in the final folded structure. The type of secondary structure of which a residue is part will have a significant impact on how much side-chain entropy is lost. Side-chain conformational entropies have previously been determined for folded proteins, simple models of unfolded proteins, alpha-helices, and a dipeptide model for beta-strands, but not for polyproline II (PII) helices. In this work, we present side-chain conformational estimates for the three regular secondary structure types: alpha-helices, beta-strands, and PII helices. Entropies are estimated from Monte Carlo computer simulations. Beta-strands are modeled as two structures, parallel and antiparallel beta-strands. Our data indicate that restraining a residue to the PII helix or antiparallel beta-strand conformations results in side-chain entropies equal to or higher than those obtained by restraining residues to the parallel beta-strand conformation. Side-chains in the alpha-helix conformation have the lowest side-chain entropies. The observation that extended structures retain the most side-chain entropy suggests that such structures would be entropically favored in unfolded proteins under folding conditions. Our data indicate that the PII helix conformation would be somewhat favored over beta-strand conformations, with antiparallel beta-strand favored over parallel. Notably, our data imply that, under some circumstances, residues may gain side-chain entropy upon folding. Implications of our findings for protein folding and unfolded states are discussed. 相似文献
13.
Toshihiko Sakurai Sayaka Oka Atsushi Kubo Katsuhiko Nishiyama Isao Taniguchi 《Journal of peptide science》2006,12(6):396-402
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
14.
The importance of accurately modeling membrane proteins cannot be overstated, in lieu of the difficulties in solving their structures experimentally. Often, however, modeling procedures (e.g., global searching molecular dynamics) generate several possible candidates rather then pointing to a single model. Herein we present a new approach to select among candidate models based on the general hypothesis that silent amino acid substitutions, present in variants identified from evolutionary conservation data or mutagenesis analysis, do not affect the stability of a native structure but may destabilize the non-native structures also found. The proof of this hypothesis has been tested on the alpha-helical transmembrane domains of two homodimers, human glycophorin A and human CD3-zeta, a component of the T-cell receptor. For both proteins, only one structure was identified using all the variants. For glycophorin A, this structure is virtually identical to the structure determined experimentally by NMR. We present a model for the transmembrane domain of CD3-zeta that is consistent with predictions based on mutagenesis, homology modeling, and the presence of a disulfide bond. Our experiments suggest that this method allows the prediction of transmembrane domain structure based only on widely available evolutionary conservation data. 相似文献
15.
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
16.
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
17.
Yoshihiro Kuroda Yoshitaka Maeda Shinichi Sawa Kiyohiro Shibata Kazuhide Miyamoto Terumichi Nakagawa 《Journal of peptide science》2003,9(4):212-220
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
18.
A new method for the quantitative determination of the percentage of intramolecular H-bonds, based on Fourier transform infrared techniques, is applied to the conformational analysis of a series of synthetic peptides spanning the processing site of the ocytocin and neurophysin precursor. Even though the method uses traditional tools such as Fourier self-deconvolution, the Nth derivative, and curve-fitting procedures for the analysis of the spectra, the assignment of the absorptions due to peptide groups participating into secondary structures is based on the direct observation and quantification of the isotopic effect induced on the groups participating in intramolecular H-bonds in the presence of organic solvents. This permits the quantification of the different populations of molecules containing intramolecular H-bonds involved in beta-turns and alpha-helices. The results are consistent with those previously obtained by NMR techniques in the same solvent systems. 相似文献
19.
Fernando Formaggio Andrea Bettio Vittorio Moretto Marco Crisma Claudio Toniolo Quirinus B Broxterman 《Journal of peptide science》2003,9(7):461-466
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
20.
Hansen PI Larsen FH Motawia SM Blennow A Spraul M Dvortsak P Engelsen SB 《Biopolymers》2008,89(12):1179-1193
To gain insight into the molecular details and hydration of amylopectin, the five constituting trisaccharides have been chemically synthesized as their methyl alpha-glycosides. All five trisaccharides were subjected to 950 MHz NMR spectroscopy for complete assignment and nanosecond molecular dynamics trajectories were calculated to study the structure and dynamics of the trisaccharides in aqueous solution. Systematic analysis of the simulation data revealed several examples of bridging water molecules playing an important role in the stabilization of specific amylopectin conformations, which was also supported by the experimental NMR data such as interresidue NOE's and heteronuclear scalar couplings between nuclei from neighboring residues. Although alpha-maltotriose, alpha-iso-maltotriose, alpha-panose and alpha-isopanose are relatively well characterized structures, the study also includes one less characterized trisaccharide with the structure alphaGlcp(1-->4)alphaGlcp(1-->6)alphaGlcp. This trisaccharide, tentatively labelled alpha-forkose, is located at the branch point of amylopectin, forking the amylopectin into two strands that align into double-helical segments. The results show that the conformation of alpha-forkose takes a natural bend form which fits well into the structure of the double-helical segment of amylopectin. As the only trisaccharide in this study the structure of alpha-forkose is not significantly influenced by the hydration. In contrast, alpha-isopanose takes a restricted, but rather extended form due to an exceptionally strong localized water density. The two homo-linkage oligomers, alpha-maltotriose and alpha-iso-maltotriose, showed to be the most extended and the most flexible trimers, respectively, providing regular structure for crystalline domains and maximum linker flexibility for amorphous domains. 相似文献