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 15 毫秒
1.
Structure of rice starch and its relation to cooked-rice texture   总被引:18,自引:0,他引:18  
Starch from seven varieties of rice, known to cook from very soft to very hard texture, was fractionated by gel-permeation chromatography on Sepharose CL-2B column. The high-molecular-weight fraction (Sepharose FRI) and the low-molecular-weight fraction (Sepharose FRII, further sub-divided into FRIIa, IIb and IIc) were debranched using isoamylase and fractionated on Biogel P-10. All the four fractions in all the different varieties of rice gave a similar trimodal chain profile, indicating the presence of branched molecules in all of them. Clearly, the branched component of starch (‘amylopectin’) is not necessarily big in size, but includes very small to very big molecules. The presence or absence of the largely linear, and relatively small molecule, ‘amylose’, could not be settled either way with the technique employed. However, based on certain assumptions, amylose content was calculated to be in the range of 7%–11% in the samples, much less than generally thought. The content of long-B chains of the branched molecule in the four Sepharose fractions individually and in aggregate, as well as the calculated amylose content, correlated well with the sensory tenderness of cooked rice. It was observed that the content of all long linear chains, including amylose if any, govern the texture of cooked rice.  相似文献   

2.
Light scattering techniques were used for structural characterization of starches with diverse amylose and amylopectin level, dissolved in water by microwave heating in a high pressure vessel and stored during different times. In general, apparent molar mass ( ), gyration radius ( ) and hydrodynamic radius ( ) values decreased when storage time increased. This could be due to depolymerization of the samples during the storage time. The fractal dimension obtained from the – relationship showed that the samples presented, in general, a globular structure, with a higher level of branching when amylopectin level in the sample increased. The particle scattering factors and Kratky plots, well suited for studying the internal structure of a macromolecule, showed a depolymerization when storage time increased. The νRH values for Eurylon 5 (0.56) and Eurylon 7 (0.58) starches were close to the values reported for linear chains. For amylopectin (0.09) and normal corn starch (0.10) the νRH values were lower; these values would define a highly branched structure. The relaxation rate distribution of the samples showed that there are changes in the internal structure when storage time increases, and that these changes depend on amylose and amylopectin level present in the sample. The ρ values for the samples analyzed were between 0.88 and 1.3; these values are characteristic of a sphere or globular structure.  相似文献   

3.
Alpha-amylase of B. amyloliquefaciens was used for the structural characterization of the amylopectin from waxy-rice starch. Fractions of -dextrins with a degree of polymerization (d.p.) <5000 were isolated from amylopectin hydrolysates after 1 and 3 h. φ,β-Limit dextrins were prepared by successive phosphorolysis and beta-amylolysis of the fractions and these were analysed by a second alpha-amylolysis. Based on the hydrolysis pattern, the limit dextrins were divided into two major groups, A and B, which possessed units of clusters of d.p. 100–200 and 90–130, respectively. An extensive alpha-amylolysis resulted in characteristic distributions of dextrins with d.p. <80 which represented branched building blocks. Type A dextrins possessed more larger building blocks with d.p. 40, but less intermediate and small blocks, than type B. The φ,β-limit dextrin of the original amylopectin had a distinct distribution enriched in small building blocks. A model is proposed in which the two types of dextrins originate from regular and less regular structural domains of the amylopectin fraction within the starch granules.  相似文献   

4.
Conditions for the preparation of carboxymethyl derivatives of corn and amaranth starch were compared. The two starches differed considerably with respect to the optimum conditions such as temperature, pH, time, concentration of sodium monochloroacetate, and starch:liquor ratio. In both cases, isopropyl alcohol was the solvent of choice. Multistage carboxylation was also carried out. Amaranth starch differs from corn starch in two respects. It is waxy in nature and also has a small granule size of 1–2 μm. However, comparison with rice starch, having a granule size of 1–2 μm and potato starch, having a similar amylose content as corn starch showed no correlation between any of these parameters.  相似文献   

5.
Influence of amylose content on starch films and foams   总被引:1,自引:0,他引:1  
After extraction of smooth pea starch and waxy maize starch from pure amylose and amylopectin fractions, films with various amylose contents were prepared by casting in the presence of water or water with glycerol. For unplasticized films, a continuous increase in tensile strength (40–70 MPa) and elongation (4–6%) was observed as amylose increased from 0 to 100%. Discrepancies with values obtained for native starches with variable amylose content and different botanical origins were attributable to variations in the molecular weights of components. Taking cell wall properties into account, the values obtained in the laboratory were used to improve the relation between the flexural behavior of extruded foams and the model of cellular solids with open cavities.

The properties of plasticized films were not improved by the presence of glycerol and remained constant when amylose content was higher than 40%. Results are interpreted on the basis of topological differences between amylose and amylopectin.  相似文献   


6.
The molecular structure and physicochemical properties of acid–alcohol treated maize and potato starches (0.36% HCl in methanol at 25 °C for 1–15 days) were investigated. The yields of the modified starches were ranging from 91 to 100%. The average granule size of modified starches decreased slightly. The solubility of starches increased with the increase of treatment time, and the pasting properties confirmed the high solubility of modified starches. The gelatinization temperatures and range of gelatinization increased with the increase of treatment time except To (onset temperature) of maize starch. Molecular structures of modified starches suggested the degradation of starches occurred mostly within the first 5 days of treatment, and degradation rate of potato starch was higher than maize starch both in amylopectin and in amylose. Maize starch was found less susceptible to acid–alcohol degradation than potato starch.  相似文献   

7.
Caldesmon is a component of the thin filaments of smooth muscles where it is believed to play an essential role in regulating the thin filaments’ interaction with myosin and hence contractility. We studied the effects of caldesmon and two recombinant fragments CaDH1 (residues 506–793) and CaDH2 (residues 683–767) on the structure of actin–tropomyosin by making measurements of the fluorescence polarisation of probes specifically attached to actin. CaDH1, like the parent molecule caldesmon, is an inhibitor of actin–tropomyosin interaction with myosin whilst CaDH2 is an activator. The F-actin in permeabilised and myosin free rabbit skeletal muscle ‘ghost’ fibres was labelled by tetramethyl rhodamine-isothiocyanate (TRITC)–phalloidin or fluorescein-5′-isothiocyanate (FITC) at lysine 61. Fluorescence polarisation measurements were made and the parameters ΦA, ΦE, Θ1/2 and N were calculated. ΦA and ΦE are angles between the fiber axis and the absorption and emission dipoles, respectively; Θ1/2 is the angle between the F-actin filament axis and the fiber axis; N is the relative number of randomly oriented fluorophores. Actin–tropomyosin interaction with myosin subfragment-1 induced changes in the parameters of the polarised fluorescence that are typical of strong binding of myosin to actin and of the ‘on’ conformational state of actin. Caldesmon and CaDH1 (as well as troponin in the absence of Ca2+) diminished the effect of S-1, whereas CaDH2 (as well as troponin in the presence of Ca2+) enhanced the effect of S1. Thus the structural evidence correlates with biochemical evidence that C-terminal actin-binding sites of caldesmon can modulate the structural transition of actin monomers between ‘off’ (caldesmon and CaDH1) and ‘on’ (S-1 and CaDH2) states in a manner analogous to troponin.  相似文献   

8.
Yield stress of 6% (w/w) waxy maize (WXM), cross-linked waxy maize (CLWM), and cold water swelling (CWS) starches in xanthan gum dispersions: 0%, 0.35%, 0.50%, 0.70%, and 1.0% was measured with the vane method at an apparent shear rate of 0.05 s−1. The intrinsic viscosity of the xanthan gum was determined to be: 112.3 dL/g in distilled water at 25 °C. Values of the static (σ0s) and dynamic (σ0d) yield stress of each dispersion were measured before and after breaking down its structure under continuous shear, respectively. The WXM and CWS starches exhibited synergistic behavior, whereas the CLWM starch showed antagonistic effect with xanthan gum. The difference (σ0s − σ0d) was the stress required to break the inter-particle bonding (σb). The contributions of the viscous (σv) and network (σn) components were estimated from an energy balance model. In general, values of σb of the starch–xanthan gum dispersions decreased and those of σn increased with increase in xanthan gum concentration.  相似文献   

9.
European agricultural landscapes are mosaics of intensively cultivated areas and semi-natural elements. Although comprising only a small fraction of the total area, semi-natural elements provide habitat for most of the landscape biodiversity. Agricultural intensification has increasingly fragmented semi-natural elements and species numbers are in decline. Insights into the effects of landscape structure on species’ distributions within and among semi-natural habitats are needed to conserve biodiversity in agricultural landscapes more effectively. We investigated the landscape- and habitat-specific diversity partitions of wild bees, true bugs, and carabid beetles in two differently structured agricultural landscapes in Switzerland. In each landscape, we partitioned the total species diversity (γ) into its additive components within (P) and among patches (βP) and among habitats (βH). In the landscape characterized by a patchy, isolated distribution of habitat elements, among-patch diversity (βP) explained 44% of the total species richness (γ) and was significantly higher than expected under a random distribution of samples among habitat patches; in the landscape with higher habitat connectivity, among-patch diversity (βP) comprised 32% of the total species richness (γ) and did not differ from the random expectation. Habitat-specific within-patch contributions to species richness were similarly low across habitat types (P=23–24%) in the patchy landscape, whereas in the more connected landscape within-patch partitions tended to be higher and differed among habitat types (P=22–38%). Functionally different groups of bees, true bugs, and carabids also responded differently to landscape structure in a manner that was consistent with known differences in resource specialization and dispersal ability. Differences in diversity partitions among landscapes and taxa indicate the need for flexible conservation strategies. Conservation of habitat-specific diversity may require more habitat patches in landscapes that have lower habitat connectivity and low within-patch diversity (P) than in landscapes with higher within-patch diversity (P).  相似文献   

10.
Molecular fractionation of starch by density-gradient ultracentrifugation   总被引:2,自引:0,他引:2  
Amylose and amylopectin in corn and potato starches were fractionated by centrifugation at 124,000g for 3-72 h at 40 degrees C in a gradient media, Nycodenz, based on their sedimentation rate differences. The fractions were collected from a centrifuge tube, and then analyzed by the phenol-sulfuric acid method and iodine-binding test. Amylopectin, a large and highly branched starch molecule, migrated faster than amylose and quickly reached its isopycnic point with a buoyant density of about 1.25 g/mL, exhibiting a sharp and stable carbohydrate peak. Amylose, which is a relatively small and linear molecule, however, migrated slowly in a broad density range and continued moving to higher density regions, eventually overlapping with amylopectin peak as the centrifugation continued. This could indicate that the buoyant density of amylose is similar to that of amylopectin. Under centrifugal conditions of 3 h and 124,000g, amylose and amylopectin molecules were clearly separated, and the presence of intermediate starch molecules (11.5 and 7.7% for corn and potato starch, respectively) was also observed between amylose and amylopectin fractions. The amylose content of corn and potato starches was 22.6 and 21.1%, respectively, based on the total carbohydrate analysis after the ultracentrifugation for 3 h. In alkaline gradients (pH 11 or 12.5), the sedimentation rate of starch molecules and the buoyant density of amylopectin were reduced, possibly due to the structural changes induced by alkali.  相似文献   

11.
Cross-linked waxy maize (CWM) starch dispersions (STDs) of concentration 50 g kg−1 were heated in sucrose solutions containing 0–600 g kg−1 (g sucrose/kg dispersion) at 85 °C at low shear and in intermittently agitated cans at 110 °C. The STDs heated in 0–300 g kg−1 sucrose exhibited antithixotropic behavior, while those heated in 400–600 g kg−1 sucrose exhibited thixotropic behavior. The mean starch granule diameter of the starch dispersions did not show strong dependence on sucrose concentration. The dispersions, especially those with high sucrose concentrations and heated at 110 °C, exhibited G′ versus frequency (ω) profiles of gels. The STDs exhibited first normal stress differences that increased in magnitude with the concentration of sucrose. Values of the first normal stress coefficient of canned dispersions calculated from dynamic rheological data plotted against ω and experimental values plotted against shear rate of some of the STDs overlapped.  相似文献   

12.
The reaction of thiamine with K2PtIICl4 and with PtIVCl4 in the presence of excess NaSCN in aqueous solution gave thiamine salts, (H-thiamine)[Pt(SCN)4] · 3H2O (1) and (H-thiamine)[Pt(SCN)6] · H2O (2), respectively, structures of which have been determined by X-ray diffraction. The thiamine molecule adopts the usual F conformation in each salt. In 1, [Pt(SCN)4]2− ions act as large planar spacers in the crystal lattice and interact scarcely with thiamine, except for a hydrogen bonding with the terminal hydroxy O(5γ). Instead, water molecules form two types of host–guest-like interactions with the pyrimidine and the thiazolium moieties of a thiamine molecule, one being a C(2)–Hwaterpyrimidine bridge and the other being an N(4′)–Hwaterthiazolium bridge. In 2, despite the much larger ion size, octahedral [Pt(SCN)6]2− ions form a C(2)–Hanionpyrimidine bridge and an N(4′)–Hanionthiazolium bridge. An additional hydrogen bonding between the anion and the terminal O(5γ) of thiamine creates a hydrogen-bonded macrocyclic ring {thiaminium–[Pt(SCN)6]2−}2, a supramolecule.  相似文献   

13.
Effects of different cellulose derivatives and carrageenans on the pasting, rheological, and textural properties of normal (NRS) and waxy (WRS) rice starches were investigated. When suspensions of both NRS and WRS were heated in a Rapid Visco Analyser (RVA) in the presence of the hydrocolloids, increases in apparent pasting temperatures and peak and final viscosities in the following decreasing order were observed: methylcellulose (MC) > carboxymethylcellulose (CMC) for cellulose derivatives and λ- > í- > κ-carregeenan for carrageenans. Slight decreases in peak and final viscosities were observed when hydroxypropylmethylcellulose (HPMC) was the hydrocolloid. Dynamic viscoelasticity measurements indicated that NRS–hydrocolloid pastes were less solid-like than the control, as evidenced by their higher tan δ values, whereas tan δ values of WRS–hydrocolloid pastes were the same as that of the control. Steady shear rheological measurements showed that addition of the different hydrocolloids used increased the apparent viscosity (ηa,100) and consistency coefficient (K) values of both starches with the same trend as that observed during pasting, whereas the opposite trend was observed for the flow behavior index (n) values. The hardness and adhesiveness of NRS pastes were significantly increased by addition of κ- and í-carrageenans, but were unaffected by the other hydrocolloids. A similar effect was observed for WRS, with the exception of κ-carregeenan, in which the hardness of the mixed paste was decreased. The starch–hydrocolloid pastes exhibited a phase-separated microstructure in which amylose- and amylopectin-rich domains were dispersed in a hydrocolloid-rich continuous phase.  相似文献   

14.
In order to elucidate early Aptian marine paleotemperature evolution across the period of enhanced organic carbon (Corg)-burial [Oceanic Anoxic Event (OAE) 1a], stable isotope analyses were performed on pelagic limestones at Deep Sea Drilling Project Site 463, central Pacific Ocean. The δ18O data exhibit a distinct anomaly by ~ − 2‰ spanning the OAE 1a interval (i.e., a ~ 6 m-thick, phytoplanktonic Corg-rich unit constrained by magneto-, bio- and δ13C stratigraphy). Elucidation of paleotemperature significance of the δ18O shift is made by taking account of recent Sr/Ca evidence at the same section, which revealed that geochemical signals in carbonate-poor lithologies are relatively unaltered against burial diagenesis. By discriminating δ18O values from carbonate-poor samples (CaCO3 contents = 5–30 wt.%), it appears that an abrupt rise in sea-surface temperatures (SSTs) by 8 °C (= − 1.7‰ shift in δ18O) occurred immediately before OAE 1a, whereas a cooling mode likely prevailed during the peak Corg-burial. In terms of its stratigraphic relationship as to the Corg-rich interval and to a pronounced negative δ13C excursion, as well as its timescale, the observed SST rise resembles those associated with the Paleocene–Eocene thermal maximum and, more strikingly, Jurassic Toarcian OAE. This observation is consistent with the hypothesis that these paleoenvironmental events were driven by a common causal mechanism, which was likely initiated by the greenhouse effect via massive release of CH4 or CO2 from the isotopically-light carbon reservoir and terminated by a negative productivity feedback.  相似文献   

15.
Redox enzyme mediated biocatalysis has the potential to regio- and stereo-specifically oxidize hydrocarbons producing valuable products with minimal by-product formation. In vitro reactions of the camphor (cytochrome P-450) 5-monooxygenase enzyme system with naphthalene-like substrates yield stereospecifically hydroxylated products from nonactivated hydrocarbons. Specifically, the enzyme system catalyzes the essentially stereospecific conversion of the cycloarene, tetralin (1,2,3,4-tetrahydronaphthalene) to (R)-1-tetralol ((R)-(−)-1,2,3,4-tetrahydro-1-naphthol). It is shown that this reaction obeys Michaelis–Menten kinetics and that interactions between the enzyme subunits are not affected by the identity of the substrate. This subunit independence extends to the efficiency of NADH usage by the enzyme system—subunit ratios do not effect efficiency, but substrate identity does. Tetralin is converted at an efficiency of 13±3%, whereas (R)-1-tetralol is converted at 7.8±0.7%. A model of this system based on Michaelis–Menten parameters for one subunit (Pdx: KM=10.2±2 μM) and both substrates (tetralin: KM=66±26 μM, νmax=0.11±0.04 s−1, and (R)-1-tetralol: KM=2800±1300 μM, νmax=0.83±0.22 s−1) is presented and used to predict the consumption and production of all substrates, products and cofactors.  相似文献   

16.
Endosperm starch and pericarp starch were isolated from maize (B73) kernels at different developmental stages. Starch granules, with small size (2–4 μm diameter), were first observed in the endosperm on 5 days after pollination (DAP). The size of endosperm-starch granules remained similar until 12DAP, but the number increased extensively. A substantial increase in granule size was observed from 14DAP (diameter 4–7 μm) to 30DAP (diameter10–23 μm). The size of starch granules on 30DAP is similar to that of the mature and dried endosperm-starch granules harvested on 45DAP. The starch content of the endosperm was little before 12DAP (less than 2%) and increased rapidly from 10.7% on 14DAP to 88.9% on 30DAP. The amylose content of the endosperm starch increased from 9.2% on 14DAP to 24.2% on 30DAP and 24.4% on 45DAP (mature and dried). The average amylopectin branch chain-length of the endosperm amylopectin increased from DP23.6 on 10DAP to DP26.9 on14DAP and then decreased to DP25.4 on 30DAP and DP24.9 on 45DAP. The onset gelatinization temperature of the endosperm starch increased from 61.3 °C on 8DAP to 69.0 °C on 14DAP and then decreased to 62.8 °C on 45DAP. The results indicated that the structure of endosperm starch was not synthesized consistently through the maturation of kernel. The pericarp starch, however, showed similar granule size, starch content, amylose content, amylopectin structure and thermal properties at different developmental stages of the kernel.  相似文献   

17.
The objective of this study was to develop a simple mechanistic model to predict the magnitude of ABA signalling ([X-ABA]) of potatoes (Solanum tuberosum L.) exposed to partial root-zone drying (PRD). Potatoes were grown in pots in a glasshouse with the roots split equally between two soil columns. At tuber initiation stage, plants were subjected to three irrigation treatments: (1) both soil columns were fully irrigated (FI) daily to a volumetric soil water content (θ) of 18.0%; (2) PRD, in which one soil column was irrigated daily to 18.0% while the other was allowed to dry, and the irrigation was shifted between columns when the θ of the drying soil column had decreased to 7–8%; (3) non-irrigation (NI), where irrigation was withheld after onset of treatments and lasted for 5 days until θ had decreased to 7%. In the PRD plants, the fraction of soil water extraction (FSWE) by the dry roots declined exponentially with declining soil water potential (Ψsoil-dry); however, after shifting of irrigation, the previously dried roots immediately recovered the full capacity of water uptake. During the first PRD drying cycle, FI plants had the highest stomatal conductance (gs), and followed by PRD plants and NI plants had the lowest gs. Photosynthesis (A) was similar for FI and PRD plants, and was significantly lower for the NI plants only on 3–4 days after treatment. In the NI plants, a linear relationship between Ψsoil and [X-ABA] was obtained. Based on these relationships, a simple model predicting [X-ABA] in the PRD plants ([X-ABA]PRD) was developed. Assuming that a constant [X-ABA] of 115 nM (similar to that found in the FI plants) originated from the wet roots; the simulation results indicated that irrigation should be shifted between the two sides when Ψsoil-dry had decreased to −80 kPa, and [X-ABA]PRD had reached a peak of ca. 150 nM. However, the [X-ABA]PRD predicted by the model was significantly lower than the measured value; whilst a simple average of [X-ABA] from the wet and the dry soil columns based on the [X-ABA]–Ψsoil relationship better predicted [X-ABA]PRD.  相似文献   

18.
B.Sridhara Rao   《Gene》1994,140(2):233-237
We have earlier demonstrated that a sequence bordering an amplified DNA segment and containing the unusual sequence (dG-dA)n·(dT-dC)n could slow replication fork movement [Rao et al., Nucleic Acids Res. 16 (1988) 8077–8094]. This was done by cloning the unusual sequence in simian virus 40 (SV40) and following the rate of incorporation of radioactively labeled nucleotides into various regions of the SV40 genome. In the present study, we have analyzed the in vivo replicative intermediates of the SV40 variants containing the unusual sequences by a two-dimensional gel electrophoretic technique. We found that the technique can be used to detect minor pauses in DNA replication and demonstrated that the cloned (dG-dA)n·(dT-dC)n tracts, that can potentially adopt triplex structures, could slow DNA replication fork movement. A sequence from the plasmid pUC18 did not slow fork movement when cloned in the same locus of SV40. The pause caused by the alternating guanosine-adenosine repeats might play a role in the regulation of DNA replication and gene amplification in vivo.  相似文献   

19.
R M Santos  E Rojas 《FEBS letters》1987,220(2):342-346
The effects of forskolin on electrical coupling among pancreatic β-cells were studied. Two microelectrodes were used to measure membrane potentials simultaneously in pairs of islet β-cells. Intracellular injection of a current pulse (ΔI) elicited a membrane response ΔV1 in the injected cell and also a response ΔV2 in a nearby β-cell confirming the existence of cell-to-cell electrical coupling among islet β-cells. In the presence of glucose (7 mM), application of forskolin evoked a transient depolarization of the membrane and electrical activity suggesting that the drug induced a partial inhibition of the β-cell membrane K+ conductance. Concomitant with this depolarization of the membrane there was a marked decrease in β-cell input resistance (ΔV2/ΔI) suggesting that exposure to forskolin enhanced intercellular coupling. Direct measurements of the coupling ratio ΔV2/ΔV1 provided further support to the idea that forskolin enhances electrical coupling among islet cells. Indeed, application of forskolin reversibly increased the coupling ratio. These results suggest that cAMP might be involved in the modulation of electrical coupling among islet β-cells.  相似文献   

20.
Paste of defatted and/or mildly acid-hydrolyzed high amylose corn starch was freeze-thawed, and then the starch was isolated by vacuum drying for the analysis in crystallization and pasting properties. X-ray diffraction pattern and differential scanning calorimetric analysis showed that the crystallinity of the freeze-thawed starch was increased as the degree of hydrolysis increased. The diffraction pattern revealed B- and V-crystals with patterns with diffraction peaks at 17, 20, and 23–25° (2θ), which were developed by amylose recrystallization during the freeze-thawing. The crystal melting enthalpies, for dual endothermic transitions above 100 °C, which resulted from the melting of amylose–lipids complex and amylose double helices were raised by the treatment. The isolated and dried starch formed a paste by aqueous heating under the ambient pressure, and its paste viscogram exhibited substantially higher resistance to shear-thinning, and rapid setback upon cooling. Acid hydrolysis, however, reduced overall paste viscosity, possibly due to the increased crystallinity. Enzyme-resistant starch content in the acid hydrolyzed starch was increased by the freeze-thawing, but not by acid hydrolysis. It was slightly increased by defatting.  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