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 25 毫秒
1.
The conformation of amylose in various solvents is discussed. It is shown that the changes in molecular volume of the polysaccharide (measured by viscosity) as potassium chloride is added to a solution of amylose at pH 12 are similar to those obtained on adding butan-1-ol to the solution. The viscosity number in both cases decreases to values less than that observed for amylose in water, in which Flory theta-conditions are approximated. The minimum value of the viscosity number, in fact, is identical to that observed on the addition of butan-1-ol and iodine to neutral aqueous solution of amylose—conditions known to result in a helical complex. It is concluded that amylose undergoes a coil-to-helix transition as potassium chloride is dadde to solutions of the polysaccharide at pH 12.  相似文献   

2.
The dependence on chain length of two characteristic properties of amylose, i.e., retrogradation and complex formation with iodine, have been studied by using enzymatically synthesized, homodisperse amyloses. The association rates of amyloses in water containing 5% dimethyl sulfoxide have a sharp maximum at a degree of polymerization P?n of 80; shorter and longer molecules are much more soluble. The iodine complexes of amylose exhibit a strong Cotton effect in the range of the long-wave absorption maximum (position depending on chain length) and two weaker Cotton effects at 480 and 350 nm. The long-wave Cotton effect is most intense at about P?n 50 and decreases rapidly for shorter and longer chains. This behavior is unexpected and is not in accordance with the further increase of λmax and λmax. The experiments can best be interpreted by assuming well ordered, stiff chains in the low molecular weight range (P?n 50–80). For longer chains, the findings are discussed in the light of current concepts of amylose conformation in aqueous solution, namely the model of the broken helical chain (alternating stiff helical segments and unordered regions) and the model of a flexible coil without a significant helical content. However, according to the results given in this paper, a wormlike helical chain seems to be the most adequate model for amylose conformation in neutral solution.  相似文献   

3.
The CD spectra of amylose, the maltose oligomers, cellobiose, and the cyclodextrins are measured in aqueous solution to 1640 Å. Two bands are found in this region. An examination of chromophorically equivalent but conformationally different glucans demonstrates that the CD of glucans is sensitive to conformation. However, the conformation of the subunits in the maltose oligomer series appears independent of chainlength. The conformations of the interior maltosyl groups of this series are all approximately equivalent. A comparison of the amylose and cycloamylose spectra indicates that amylose is substantially biased in its chirality. Since there is little chainlength dependence, the oligomers should have a similar chirality bias. The butanol complex of amylose, which is believed to have a V-form helical conformation, has a CD similar to free aqueous amylose. This is consistent with conformational calculations on amylose which predict that the maltosyl subunits of amylose are restricted to a small region on its conformational map which includes the V-form helix. It is also consistent with the idea that amylose a loosely wound and extended helix in aqueous solution.  相似文献   

4.
Amylose, a putative linear α-(1→4)-glucan and a component of most starches, was isolated from potato, rice, and wheat starches by forming the 1-butanol complex in a solution of the starches. It previously had been found that these amyloses were incompletely hydrolyzed by β-amylase, indicating that it was partially branched. Solubilization of the butanol complex in water and steam distillation of the 1-butanol, followed by cooling to 4 °C gave precipitation of the double helical, linear, retrograded amylose over a 15 h period, leaving the soluble amylose in solution. The soluble amyloses were precipitated with two volumes of ethanol, and the precipitate was solubilized and reprecipitated to remove traces of linear amylose. The precipitated, soluble amyloses, were partially branched and had properties intermediate between linear amylose and amylopectin. The water solubility of the potato amylose was 10.52 mg/mL, with a number-average degree of polymerization (DPn) of 8440 and 2.1% branch linkages that had a DPn of 48; the water solubility of the rice amylose was 8.83 mg/mL, with a DPn of 2911 and 1.4% branch linkages that had a DPn of 72; and the water solubility of wheat amylose was 6.33 mg/mL, with a DPn of 1160 and 1.6% branch linkages that had a DPn of 64. The three soluble amyloses have structures and properties intermediate between the nearly water insoluble (?1 mg/mL), linear amylose, and the highly water-soluble, 4-5% branched, amylopectin.  相似文献   

5.
Using a potato amylose fraction of 8 × 105, molecular-weight viscosity studies were carried out at 25°C on solutions containing 0.176–0.042% polymer, 8.67 mM KI, 1% ethanol, and different concentrations of iodine. By a novel extrapolation method, the intrinsic viscosities of the amylose/iodine complex were determined under various conditions of iodine binding (0–0.133 g I2/g amylose). Contrary to the view long held in this research area, it was found that the intrinsic viscosity of amylose solutions decreases significantly upon complex formation with iodine. Taking into account the results of our previous kinetic studies, the present findings are interpreted in terms of an amylose model characterized by loose, extended helical regions which are interrupted by short disordered regions. It is proposed that the intrinsic viscosity decrease observed is due to a shortening of the linear dimension of the polymer chain. This conformation change is apparently caused by the contraction of loose helical regions of the amylose macromolecule due to the entrapment of iodine (and perhaps other) atoms inside the helical cavities.  相似文献   

6.
Complex formation in the ternary system of amylose (degree of polymerization, DP, 1100), SDS, and iodine was studied statically by spectrophotometry and amperometric titration and kinetically by the pressure-jump method. It was clarified that (1) iodine (I3?) to some extent binds to amylose saturated with SDS to form an inclusion complex (ASI system); (2) the binding of SDS apparently transforms amylose of DP 1100 to that of much lower DP (less than 60) from the viewpoint of iodine binding; and (3) iodine binds to sites unoccupied by SDS in the center of the helical segment of amylose. Pressure-jump relaxation phenomenon was not observed in solutions in which iodine was dissolved prior to SDS (AIS system), but it was observed in the ASI system; it is ascribed to the association and dissociation of three molecules of iodine in the center of the amylose helix. Comparison of the rate constants in the ASI system with those in the amylose (DP 32) and iodine system indicates that iodine runs to and from the helical segment of amylose perpendicularly to the axial plane in the former, while it runs horizontally in the latter. We discuss the order of ligand mixing on the resulting structure of the ternary complexes of amylose, SDS, and iodine.  相似文献   

7.
《Carbohydrate polymers》1997,33(1):19-26
Commercial potato amylose was used to study temperature effects on the retrogradation of amylose solutions (3.5mg/ml). The retrogradation rate decreased as incubation temperature increased (5 to 45 °C). The degree of retrogradation within 24 h decreased from 58.8 to 7.1% as incubation temperature increased from 5 to 45 °C. In the amylose solution, different-sized molecular subfractions retrograded at different rates. After incubating at 5 °C for 100 days, the majority of the amylose molecules retrograded and precipitated from the solution; at 45 °C, only amylose of the small-molecular subfraction (number average, DPn = 110; weight average, DPw = 150) retrograded and precipitated. Entanglement of molecules was observed in size exclusion chromatograms. The morphology of retrograded amylose observed by using a scanning electron microscope differed with the retrogradation temperature. The chain length of amylose crystalline segments, prepared by hydrolysis of retrograded amylose, showed a narrow distribution (polydispersity from 1.21 to 1.67). The chain lengths of resistant segments increased DPn from 39 to 52 and DPw from 47 to 72 for α-amylolysis and DPn from 34 to 40 and DPw from 48 to 67 for 16% sulfuric acid hydrolysis, when incubation temperature increased from 5 to 45 °C.  相似文献   

8.
The effects on amylose conformation of percentage water in dimethyl sulfoxide (DMSO)/water mixtures were measured by following changes in specific optical rotation, limiting viscosity number, and 13C-NMR chemical shifts. The temperature dependence of specific optical rotation showed differences in amylose conformation at four chosen ratios of dimethyl sulfoxide/water. An amylose conformational change was also deduced from 13C-NMR chemical shift data. Changes in limiting viscosity of amylose in different proportions of DMSO/water, and the effect of tetramethylurea on the specific rotation of amylose, indicate that intramolecular hydrogen bonding decreases with increased water content. 66.6% DMSO appears to be a crossover concentration, below which the helical conformation is progressively lost as water is added. When water content is over 60%, transition to a conformation which allows iodine complexation to take place is complete. A transition of amylose conformation from helix to loose helix to random coil with increasing water content was deduced from the experimental results.  相似文献   

9.
The nonbonded interaction energy of disaccharides, mannobiose and galactobiose and polysaccharides mannan and galactan have been computed as a function of dihedral angles (?,ψ). The conformation (40°, ?20°) has been preferred for the mannan chain from nonbonded interaction energy considerations. The O5…O3′ type of intramolecular hydrogen bond has been found to be possible in the above conformation. Comparison of the allowed region of mannan with those of cellulose and xylan indicates that the monomer unit, in mannan chain has slightly higher freedom of rotation than that of cellulose and less than that of xylan. As in cellulose and mannan, the freedom of rotation of the monomer units in β-1,4′ galactan is highly restricted. Unlike mannan (which prefers an extended conformation) the β-1,4′ galactan prefers a helical conformation similar to amylose. Just as in amylose the O2…O3′ type hydrogen bond between contiguous residues is also possible in β-1,4′ galactan.  相似文献   

10.
Bernd Ebert  Hans Elmgren 《Biopolymers》1984,23(11):2543-2557
In the previous study [Elmgren, H. (1984) Biopolymers 23, 2525–2541] concerning the conformation of amylose in aqueous solution, it was stated that amylose in a neutral aqueous solution is a random coil consisting of helical segments. In terms of Kuhn statistics, each segment contains more than 100 monomers. The number of monomers per segment decreases by alkali addition. In an attempt to verify these statements, a combined electron spin resonance (esr) and ultracentrifugation (uc) study of a weakly hydroxyethylated amylose sample in water and alkaline solvents was performed. This combination of measuring techniques makes it possible to estimate the Brownian motion, and thus the mass of the polymer segments. As a control for the obtained esr data, fluorescence depolarization (fdp) measurements were performed on the polymer sample in a bicarbonate buffer at pH 10. The result of the study confirms that the amylose segments are very heavy in water. In strong alkaline solvents, the segment mass corresponds to that of a few monomers. Our findings thus support the statements made in the preceding article, and the data obtained by others. [Kitamura S., Yunokawa H., & Kuge T., (1982) Polym. J. 14, 85–91; Kitamura S., Yunokawa H., Mitsu'ie S., & Kuge T., (1982) Polym J. 14, 93–99].  相似文献   

11.
M. Kodama  H. Noda  T. Kamata 《Biopolymers》1978,17(4):985-1002
The conformation of amylose in aqueous solution has been found to be dependent on its molecular weight. When the molecular weight of amylose is outside of the so-called “dissolving gap” described by Burchard (6500<Mr<160,000) it behaves as a random coil, whereas when its molecular weight is within the “dissolving gap,” it easily aggregates forming a rigid coil which is the B-type (retrograded) amylose. The conformation of this rigid coil is suggested to be a double helix.  相似文献   

12.
Fluorescence of 2-p-toluidinylnaphthalene-6-sulfonate (TNS) was enhanced in the presence of maltooligosaccharides, amylose, and other α-glucans. The dependence of relative TNS fluorescence intensity per glucose unit on chain length of oligosaccharides was examined. The values of binding constant and thermodynamic parameters, assuming the 1:1 complex for TNS-amylose (number-average degree of polymerization, DPN = 17), were determined by the fluorescence titration. The values of thermodynamic parameters for 1:1 complex formation of TNS-α- and β-cyclodextrins were also determined and compared with those of TNS-amylose (DPN = 17). The fluorescence intensity of TNS in the presence of amylose (DPN = 600) decreased by the action of glucoamylase and taka-amylase A. The fluorescence of TNS-amylose (DPN = 17) system increased with the increased ionic strength. In the presence of pullulan, TNS fluorescence was also enhanced and decreased by the action of pullulanase. Amylopectin enhanced TNS fluorescence rather more strongly than amylose (DPN = 17) at the same concentration. In the presence of dextran, the fluorescence of TNS was scarcely enhanced. The degree of fluorescence enhancement of TNS in the presence of α-glucans seems to reflect the structures of α-glucans in solution, since TNS fluorescence is enhanced in the hydrophobic environment or by the disturbance of free intramolecular rotation.  相似文献   

13.
The electric birefringence of poly(L -glutamic acid) (PLGA) in dimethylsulfoxide (DMSO)–methanol mixtures has been measured by use of the rectangular pulse technique. The length distribution curve, the mean molecular length, and the mean apparent permanent dipole moment of PLGA in solution have been obtained from the decaycurve and field strength dependence of the steady-state birefringence according to the method developed for analyzing the electric birefringence of a polydisperse system. The length distribution curve exhibits one or two peaks. The length corresponding to a high peak and the mean length of PLGA undergo an abrupt change in the vicinity of 50 to 60 vol % DMSO at 30°C. Moreover, a sharp change of the Moffitt b0 parameter with the solvent composition is observed. These results provide evidence for the existence of a solvent-induced transition from a helical conformation (presumably α-helix) to another helical conformation with shorter length per amino acid residue. Further, the temperature dependence of the length distribution of PLGA in 50 vol % DMSO suggests the existence of a temperature-induced helix ? helix transition.  相似文献   

14.
We previously proposed a mechanism for the structural formation of cellulose from its solution using a molecular dynamics (MD) simulation and suggested that the initial structure from its solution plays a critical role in determining its final structure. Structural changes in the van der Waals-associated cellulose molecular sheet as the initial structure were examined by MD simulation; the molecular sheet was found to be disordered due to maltohexaoses as an amylose model in terms of the hydrogen bonding system of cellulose. The structure and properties of cellulose/amylose blends prepared from an aqueous NaOH solution were examined experimentally by wide-angle X-ray diffraction and dynamic viscoelasticity measurements. The crystallinity of cellulose in the cellulose/amylose blend films was lower than that of cellulose film. The diffraction peaks of the cellulose/amylose blends were slightly shifted; specifically, () was shifted to a higher angle, and (1 1 0) and (0 2 0) were shifted to lower angles. These experimental results probably resulted from the disordered molecular sheet, as revealed by MD simulations.  相似文献   

15.
Starch dextrins of different molecular sizes (DPn 311, 142 and 39) were prepared by hydrolyzing a high amylose maize starch in acidic alcohol solutions. The dextrins were dissolved in an aqueous dimethyl sulfoxide solution (90% DMSO), and then the solution was allowed to migrate down into n-butanol separated by a membrane filter. The complex was gradually formed between the dextrin and butanol, and precipitated in the butanol layer. The dextrin–butanol complex yielded V6-I type crystals with broad reflections (d-spacings 1.123, 0.657 and 0.429 nm) under X-ray diffractograms. Platelets of average length less than 100 nm, interspersed in amorphous matrices, were observed in complexes of DPn 311 and 142, but that of DPn 39 showed different morphology, and the formation of complexes was limited. By hydrolyzing the complex of DPn 311 with α-amylase, amorphous matrices were selectively removed, and crystallites of 23–72 nm showing a V6-I X-ray diffraction pattern were obtained. However, crystallites in complexes of DPn 142 and 39 were eroded by amylolysis, forming large aggregates.  相似文献   

16.
The 21-residue fragment Tyr-Gly-Ser-Thr-Ser-Gln-Glu-Val-Ala-Ser-Val-Lys-Gln-Ala-Phe-Asp-Ala-Val- Gly-Val-Lys, corresponding to sequence 296-316 of thermolysin and thus encompassing the COOH-terminal helical segment 301-312 of the native protein, was synthesized by solid-phase methods and purified to homogeneity by reverse-phase high performance liquid chromatography. The peptide 296-316 was then cleaved with trypsin at Lys307 and Staphylococcus aureus V8 protease at Glu302, producing the additional fragments 296-307, 308-316, 296-302, and 303-316. All these peptides, when dissolved in aqueous solution at neutral pH, are essentially structureless, as determined by circular dichroism (CD) measurements in the far-ultraviolet region. On the other hand, fragment 296-316, as well as some of its proteolytic fragments, acquires significant helical conformation when dissolved in aqueous trifluoroethanol or ethanol. In general, the peptides mostly encompassing the helical segment 301-312 in the native thermolysin show helical conformation in aqueous alcohol. In particular, quantitative analysis of CD data indicated that fragment 296-316 attains in 90% aqueous trifluoroethanol the same percentage (approximately 58%) of helical secondary structure of the corresponding chain segment in native thermolysin. These results indicate that peptide 296-316 and its subfragments are unable to fold into a stable native-like structure in aqueous solution, in agreement with predicted location and stabilities of isolated subdomains of the COOH-terminal domain of thermolysin based on buried surface area calculations of the molecule.  相似文献   

17.
Abstract

NMR and CD spectroscopy have been used to examine the conformation of the peptide, β(12–28), (VHHQKLVFFAEDVGSNK) in aqueous and 60% TFE/40% H20 solution at pH 2.4. In 60% TFE solution, the peptide is helical as confirmed by the CD spectrum and by the pattern of the NOE cross peaks detected in the NOESY spectrum of the peptide. In aqueous solution, the peptide adopts a more extended and flexible conformation. Broadening of resonances at low temperature, temperature-dependent changes in the chemical shifts of several of the CHα resonances and the observation of a number of NOE contacts between the hydrophobic side-chain protons of the peptide are indicative of aggregation in aqueous solution. The behavior of β(12–28) in 60% TFE and in aqueous solution are consistent with the overall conformation and aggregation behavior reported for the larger peptide fragment, β(1–28) and the parent β-amyloid peptide.  相似文献   

18.
Hylaseptin-4 (HSP-4, GIGDILKNLAKAAGKAALHAVGESL-NH2) is an antimicrobial peptide originally isolated from Hypsiboas punctatus tree frog. The peptide has been chemically synthetized for structural investigations by CD and NMR spectroscopies. CD experiments reveal the high helical content of HSP-4 in biomimetic media. Interestingly, the aggregation process seems to occur at high peptide concentrations either in aqueous solution or in presence of biomimetic membranes, indicating an increase in the propensity of the peptide for adopting a helical conformation. High-resolution NMR structures determined in presence of DPC-d38 micelles show a highly ordered α-helix from amino acid residues I2 to S24 and a smooth bend near G14. A large separation between hydrophobic and hydrophilic residues occurs up to the A16 residue, from which a shift in the amphipathicity is noticed. Oriented solid-state NMR spectroscopy show a roughly parallel orientation of the helical structure along the POPC lipid bilayer surface, with an insertion of the hydrophobic N-terminus into the bilayer core. Moreover, a noticeable pH dependence of the aggregation process in both aqueous and in biomimetic membrane environments is attributed to a single histidine residue (H19). The protonation degree of the imidazole side-chain might help in modulating the peptide-peptide or peptide-lipid interactions. Finally, molecular dynamics simulations confirm the orientation and preferential helical conformation and in addition, show that HSP-4 tends to self-aggregate in order to stabilize its active conformation in aqueous or phospholipid bilayer environments.  相似文献   

19.
The iodine which is added to an aqueous amylose solution is bound only partly by the amylose while forming the blue complex and partly remains free. The equilibrium normality of the free and the bound iodine at half-saturation of amylose by iodine is designated as [If]v and [Ib]w, respectively. The stability of the poly iodine chain formed within the axis of amylose helices depends on its length, i.e., indirectly on the DP of the amylose helices: the greater this stability, the lower the [If]v value. The amylose molecule consists of helical segments. Such a molecule may behave as a random coil. The average length of the helical segments in freshly prepared amylose-iodine complexes depends on temperature, pH, iodide concentration, the presence of other complex-forming agents, and the DP of the amylose. This latter factor is investigated in the present paper. By the aid of an automatically recording photometrictitrating device the coherent values of [Ib] and [If] were determined. Plotting these values against DP n for mechanochemically degraded as well as for periodateo-xidized amyloses resulted in curves consisting of two linear sections. The break of the curves occurred between DP n 110 and 130. It was concluded that below DP n = 100 the DP of helical segments (= sDP n) is identical to the DP n of the total molecule, i.e., the molecule consists of only a single, relatively stiff helix. Above this limit the molecule contains several helical segments. The DP of these helical segments can be calculated as follows: sDP n = 141.1 ? 10.2 × 105[If]v. This equation is considered to be valid for 0.5–0.6 mg. amylose in 100 ml. 0.1N HCl at 20°C., λ = 650 mμ, euuvet diameter 3.4 cm., the feed rate of the iodate-iodide titrating solution (in acid medium resulting in a 5 × 10?3N I2 solution with a molar iodide to iodine ratio of 1.5) is 0.4ml./min. Amylose molecules of, e.g., DP n = 1380 consist of an average of 11.4 segments having a DP of about 120 and consisting of an average of 15–18 helical turns.  相似文献   

20.
The Moffitt b0 parameter of poly-L-glutamic acid in the presumed helical state varied with solvent composition, ranging in magnitude from less than 600° in aqueous solution to 800° in methanol. b0 was also dependent on temperature throughout the excessable temperature range. The value in aqueous solution is at least 100° smaller than the values for a number of polypeptides in organic solvents, when compared at the same refractive index. Therefore the optical rotatory dispersion data do not provide evidence that the molecule is completely helical in aqueous solution. Since other types of evidence for helical content are not sufficient to establish that PLGA is a complete helix, the helical content of proteins and polypeptides determined by rotatory dispersion measurements should be regarded as uncertain by about 20 per cent.  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