首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
A ternary complex consisting of amylose, whey protein, and free fatty acids (FFA) has been identified in our previous investigations, and its iodine binding properties were investigated. After reaction with iodine solution, an absorption peak (λmax) at 620 nm was shown for pure amylose whereas the λmax decreased to 510 nm when amylose was first complexed with FFA. Interestingly, a λmax of 550 nm with an intermediate absorbance was observed for the ternary complex indicating its intermediate spectrophotometric property. Consistently, the amount of iodine bound by the ternary complex was between free amylose and typical amylose–FFA complex from potentiometric titration indicating the amylose–FFA complex within the ternary complex is less compact and more space is left for iodine binding. This in-between property of the ternary complex suggests it can be used as a molecular carrier to accommodate a forth component in addition to its functional lipids carrying capability in food product development.  相似文献   

2.
This protocol describes a high through put colorimetric method that relies on the formation of a complex between iodine and chains of glucose molecules in starch. Iodine forms complexes with both amylose and long chains within amylopectin. After the addition of iodine to a starch sample, the maximum absorption of amylose and amylopectin occurs at 620 and 550 nm, respectively. The amylose/amylopectin ratio can be estimated from the ratio of the 620 and 550 nm absorbance values and comparing them to a standard curve in which specific known concentrations are plotted against absorption values. This high throughput, inexpensive method is reliable and reproducible, allowing the evaluation of large populations of potato clones.   相似文献   

3.
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   

4.
C.d. measurements on amylose iodine solutions carried out at different degrees of iodine saturation show that the decrease of the Cotton effect, observed earlier below DP 5011,12, is mainly due to the decreasing complex formation constants. The continuous decrease of the Cotton effect above DP 50 is even more pronounced at higher iodine concentration. Slow addition of iodine leads to especially high Cotton effects with a marked maximum at DP 47–50 and only a small increase of the Cotton effect upon standing of the solutions. Rapid addition of iodine leads to considerably lower Cotton effects but a prolonged time-dependent increase. Further ordering, however, does not reach the same high degree as when ioidine is added slowly. The results are discussed in the light of a single and multichain initiation process. Studies with aggregating solutions on the one hand and with very dilute solutions on the other hand, strongly indicate that the c.d. intensity is related to a conformational state of helix ordering and not to an ordering by chain folding or regular intermolecular association. Thus, the slow, time-dependent increase of the Cotton effect reflects substantial conformational changes of amylose on binding iodine. The results indicate that in the range of DP 47–50 a particularly well ordered helix is formed. For production of the typical blue colour in aqueous soluton and stabilization of aligned polyiodide subunits there is no need to stipulate a special colloid or crystalline state of the complex.  相似文献   

5.
Most starch granules exhibit a natural crystallinity, with different diffraction patterns according to their botanical origin: A‐type from cereals and B‐type from tubers. The V polymorph results essentially from the complexing of amylose with compounds such as iodine, alcohols, or lipids. The intensity and nature of phase transitions (annealing, melting, polymorphic transitions, recrystallization, etc.) induced by hydrothermal treatments in crystalline structures are related to temperature and water content. Despite its small concentration, the lipid phase present mainly in cereal starches has a large influence on starch properties, particularly in complexing amylose. The formation of Vh crystalline structures was observed by synchrotron x‐ray diffraction in native maize starch heated at intermediate and high moisture contents (between 19 and 80%). For the first time, the crystallization of amylose–lipid complexes was evidenced in situ by x‐ray diffraction without any preliminary cooling, at heating rates corresponding to the usual conditions for differential scanning calorimetry experiments. For higher water contents, the crystallization of Vh complexes clearly occurred at 110–115°C. For intermediate water contents, mixed A + Vh (or B + Vh for high amylose starch) diffraction diagrams were recorded. Two mechanisms can be involved in amylose complexing: the first relating to crystallization of the amylose and lipid released during starch gelatinization, and the second to crystalline packing of separate complexed amylose chains (amorphous complexes) present in native cereal starches. © 1999 John Wiley & Sons, Inc. Biopoly 50: 99–110, 1999  相似文献   

6.
When 2-mm apical segments of the primary roots of Zea mays L.(cv. Golden Cross Bantam 70) were irradiated successively withred and far-red light, a photoirreversible absorbance decreasewas separated from the red far-red reversible absorbance changetypical of phytochrome. The difference spectrum of the reversiblechange showed maximum absorbance changes at 666 and 730 nm,while the photoirreversible change induced by red light showeda maximum decrease at 640 nm. The photoreversible absorbancechange was linearly proportional to the fluence of red lightbetween 1 and 6 J m–2, while the photoirreversible absorbancechange was proportional to its logarithm. Red light of approximately6 J m–2 induced 50% of the maximum photoirreversible absorbancechange at 640 nm but only about 25% of the maximum photoreversibleabsorbance change. Moreover, no effect of ascorbate on the twoabsorbance changes was observed. 1Faculty of Education, University of Yamagata, Yamagata 990,Japan. (Received November 2, 1980; )  相似文献   

7.
When 2-mm apical segments of the primary roots of Zea mays L.(cv. Golden Cross Bantam 70) were irradiated successively withred and far-red light, a photoirreversible absorbance decreasewas separated from the red far-red reversible absorbance changetypical of phytochrome. The difference spectrum of the reversiblechange showed maximum absorbance changes at 666 and 730 nm,while the photoirreversible change induced by red light showeda maximum decrease at 640 nm. The photoreversible absorbancechange was linearly proportional to the fluence of red lightbetween 1 and 6 J m–2, while the photoirreversible absorbancechange was proportional to its logarithm. Red light of approximately6 J m–2 induced 50% of the maximum photoirreversible absorbancechange at 640 nm but only about 25% of the maximum photoreversibleabsorbance change. Moreover, no effect of ascorbate on the twoabsorbance changes was observed. 1Faculty of Education, University of Yamagata, Yamagata 990,Japan. (Received November 2, 1980; )  相似文献   

8.
Lamellar single crystals of a high temperature polymorph of synthetic dextran were prepared at temperatures ranging from 120 to 200°C in a mixture of water and polyethylene glycol. Individual crystals with lath-like shapes gave well resolved electron diffraction diagrams from which the reciprocal unit cell parameters a*, b* and γ* could be measured. The direct cell parameters were then determined from a series of electron diffraction diagrams obtained by sequential tilting of the crystal about the b* axis. This gave a = 0·922 ± 0·001 nm, b = 0·922 ± 0·001 nm, c (chain axis) = 0·78 ± 0·01 nm, α = γ = 90° and β = 91·3° ± 0·5°. The crystal symmetry was P21 with b as the unique monoclinic axis. These data coupled with the observed density of the crystals, indicated that the unit cell contained two antiparallel dextran chains of two residues each. When the crystals were grown at temperatures between 90 and 120°C, a percentage of crystals containing both low and high temperature polymorphs were obtained. These mixed crystals had most likely grown in syntaxy.  相似文献   

9.
When amylose was synthesized using potato phosphorylase in the presence of amylose complexing lipids, monodisperse populations of amylose–lipid complexes were formed. Enzyme dosage and glucose-1-phosphate (glc-1-P)/primer ratio influenced the reaction rate of the enzymic synthesis, presumably by changing the balance between amylose synthesis and amylose–lipid complexation and precipitation, and impacted the molecular weight of the complexes. Lipid characteristics affected the dissociation properties and amylose chain lengths of the amylose–lipid complexes presumably by determining the minimal amylose chain length necessary for complexation and precipitation. Tailor made short chain amylose–lipid complexes can hence be produced by choosing the appropriate reaction conditions. We propose a synthesis mechanism in which the primer is elongated until an amylose chain is obtained which is of sufficient length to complex a first lipid. Further chain extension then occurs, together with subsequent complexation until the complex becomes insoluble and precipitates.  相似文献   

10.
The starch–iodine blue complex formation does not involve negatively charged iodine species like I, I, or I; rather, neutral iodine units are involved. The heat of reaction is determined to be about ?110 kJ for every mole of I-I unit in the amylose helix, which suggests that the dissociation of I2 (binding energy 149 kJ/mol) does not take place during the complex formation. Quantum mechanical (INDO CI) calculations indicate that the linear as well as nonlinear polyiodine units, I6, with interiodine distance of 3.0 Å are responsible for characteristic absorbance bands of the starch–iodine complex. Based on our previous article [(1989) J. Polym. Sci. A 27 , 4161] and the present studies we identify (C6H10O5)16.5I6 to be the polymeric unit responsible for the characteristic blue color of the complex.  相似文献   

11.
T. D. Simpson 《Biopolymers》1970,9(9):1039-1047
Solutions of amylose in ethylenediamine yield a crystalline film complex upon evaporation of solvent. The x-ray analysis indicates the presence of a tetragonal-shaped cell with a symmetry approximating that of space group P212121. The amylose sixfold helix has a diameter of 13.3 Å and a translation period of 8.0Å. Chemical and physical analyses support a complexing ratio of one ethylenediamine molecule to every two glucose units. The structure is nearly identical to any amylose–dimethyl sulfoxide complex previously examined. The square mode of packing arrangement appears to result from complexation between amylose chains. Such complexing indicates a much greater degree of amylose interaction than is observed in amylose complex structures having a hexagonal close-packing arrangement.  相似文献   

12.
The crystal structures of (1→3)-α-d-glucan triacetates were studied by X-ray diffraction measurements on fibre diagrams. The oriented films annealed in water at high temperature were of higher crystallinity and occurred as two crystalline polymorphs (GTA I and GTA II) depending on the samples and also the annealing temperature. All reflections in GTA I were indexed with a pseudo-orthorhombic unit cell with a = 1·753, b = 3·018 and c(fibre axis) = 1·205 nm. From the fibre repeat data coupled with the density data and the presence of only the (003) reflection on the meridian, an extended three-fold helical structure was proposed. Although some reflections in GTA II split from the layer lines, the basic unit cell was a monoclinic system with a = 1·685, b = 3·878, c (fibre axis) = 1·210 nm and γ = 112·2°. A similar three-fold structure to GTA I was proposed from the almost identical fibre repeat and the conformational analysis on (1→3)-α-d-glucan. It was concluded that, on acetylation, the d-glucan structure changed from the fully extended two-fold helix to the extended three-fold accompanied by some extent of chain shrinking.  相似文献   

13.
The extracellular starch released by autolysis of glycerol-grown Chroomonas salina (Wislouch) Butcher was purified and chemically characterized. It contained 99%D-glucose, formed a blue complex (λmax at 605 nm) with iodine, and showed an optical rotation value more positive than that of potato starch. Methylation analysis established the presence of 1,4- and 1,4,6-glucoside linkages in the proportion 15:1. These properties indicate the cryptomonad starch to be an iodophilic α-(1,4)-glucan, composed of ca. 30% amylose with amylopectin, and broadly resembling potato, Chilomonas and other unicellular algal starches.  相似文献   

14.
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   

15.
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   

16.
Oxygen evolving photosystem II (PSII-OEC) complexes and PSII core complexes were isolated from spinach and the thermophilic cyanobacteriumSynechococcussp. OD24 and characterized by gel electrophoresis, immunoblotting, and absorbance spectroscopy. The mass of the core complexes was determined by scanning transmission electron microscopy (STEM) and found to be 281 ± 65 kDa for spinach and 313 ± 52 kDa forSynechococcussp. OD24. The mass of the spinach PSII-OEC complex was 327 ± 64 kDa. Digital images of negatively stained PSII-OEC and PSII core complexes were recorded by STEM and analyzed by single particle averaging. All monomeric complexes showed similar morphologies and were of comparable length (14 nm) and width (10 nm). The averages revealed a pseudo-twofold symmetry axis, which is a prominent structural element of the monomeric form. Difference maps between the averaged projections of the oxygen evolving complexes and the core complexes from both species indicated where the 33-kDa extrinsic manganese stabilizing protein is bound. A symmetric organization of the PSII complex, with the PsbA and the PsbD proteins in the center and symmetrically arranged PsbB and PsbC proteins at the periphery of the monomeric complex, is proposed.  相似文献   

17.
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   

18.
Iodine has been used as an effective tool for studying both the structure and composition of dispersed starch and starch granules. In addition to being employed to assess relative amylose contents for starch samples, it has been used to look at the molecular mobility of the glucose polymers within intact starch granules based on exposure to iodine vapor equilibrated at different water activities. Starches of different botanical origin including corn, high amylose corn, waxy corn, potato, waxy potato, tapioca, wheat, rice, waxy rice, chick pea and mung bean were equilibrated to 0.33, 0.75, 0.97 water activities, exposed to iodine vapor and then absorbance spectra and LAB color were determined. In addition, a new iodine quantification method sensitive to <0.1% iodine (w/w) was employed to measure bound iodine within intact granular starch. Amylose content, particle size distribution of granules, and the density of the starch were also determined to explore whether high levels of long linear glucose chains and the surface area-to-volume ratio were important factors relating to the granular iodine binding. Results showed, in all cases, starches complexed more iodine as water content increased and waxy starches bound less iodine than their normal starch counterparts. However, much more bound iodine could be measured chemically with waxy starches than was expected based on colorimetric determination. Surface area appeared to be a factor as smaller rice and waxy rice starch granules complexed more iodine, while the larger potato and waxy potato granules complexed less than would be expected based on measured amylose contents. Corn, high amylose corn, and wheat, known to have starch granules with extensive surface pores, bound higher levels of iodine suggesting pores and channels may be an important factor giving iodine vapor greater access to bind within the granules. Exposing iodine vapor to moisture-equilibrated native starches is an effective tool to explore starch granule architecture.  相似文献   

19.
Phototransformation of the red-light-absorbing form (PR) tothe far-red-light-absorbing form (PFR) of phytochrome in 7-day-oldetiolated pea epicotyl hook segments was examined at 0.5C aftera red laser flash excitation using a multichannel transientspectra analyser with electrically gated photomultiplier. Effectsof a red laser pulse on the induction of phototransformationfrom PR to PFR were saturated at Ca. 15 mJ for flash wavelengthsof both 640 and 655 nm. The amount of PFR induced by a saturatinglaser pulse was ca. 50% of that obtained at the photostationaryequilibrium. A difference spectrum measured 15 µs afterthe flash showed an absorbance increase at 697 nm and a decreaseat 663 nm. A difference spectrum determined 200 ms after theflash showed no such major absorbance increase. Kinetic analysisof the rapid absorbance decrease at 700 and 710 nm gave onesimple first-order reaction component having a rate constantof 2,500 s–1. Kinetics of PFR appearance measured by absorbanceincrease at 750 nm was resolved into three first-order reactionshaving rate constants of 5, 1.8 and 0.4 s–1. The secondflash light of 710 nm given 2 µs and 2 ms after the firstred flash irradiation on PR resulted in the formation of PRrather than PFR. (Received February 8, 1985; Accepted April 11, 1985)  相似文献   

20.
A photoactive substrate-enzyme complex of the NADPH:protochlorophyllide oxidoreductase (POR; EC 1. 3. 1. 33) was purified from etiolated Triticum aestivum L. by gel chromatography after solubilization of prolamellar bodies by dodecyl-maltoside. Irradiation by a 1-ms flash induced the phototransformation of protocholorophyllide a (Pchlide) with −196 °C absorbance and emission maxima at 640 and 643 nm, respectively. The apparent molecular weight of this complex was 112 ± 24 kDa, which indicates aggregation of enzyme subunits. By lowering the detergent concentration in the elution buffer, a 1080 ± 250-kDa particle was obtained which displayed the spectral properties of the predominant form of photoactive Pchlide in vivo (−196 °C absorbance and fluorescence maxima at 650 and 653 nm). In this complex, POR was the dominant polypeptide. Gel chromatography in the same conditions of an irradiated sample of solubilized prolamellar bodies indicated rapid disaggregation of the complex after Pchlide phototransformation. High performance liquid chromatographic analysis of the POR complexes obtained using two detergent concentrations indicates a possible association of zeaxanthin and violaxanthin with the photoactive complex. Received: 25 February 1998 / Accepted: 8 June 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