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 46 毫秒
1.
Some properties of cassava and cocoyam starches were studied and compared with a view to determining the functional applications in food systems for which they are suitable. The starches were compared in terms of their microscopic, thermal, physicochemical and rheological properties. Microscopy revealed smaller granule sizes of cocoyam starch compared with cassava. The amylose content was higher in cocoyam starch (33.3%) than in cassava starch (29.3%). Gelatinization in cassava starch occurred at a lower temperature range (60.11–72.67 °C) compared with cocoyam (72.96–80.25 °C) with the endothermic gelatinization enthalpy higher in cocoyam. The swelling power and solubility patterns indicated lower relaxation temperature, higher swelling and solubilization rates in cassava starch compared with cocoyam starch. The pasting characteristics of 8% (db, dry basis) starch slurry showed that cassava had higher peak viscosity but lower stability and setback ratios compared with cocoyam. This indicates that cocoyam starch paste is better in withstanding processing conditions and would present a superior thickening characteristic than cassava starch paste. The flow properties of both starch pastes showed non-Newtonian behaviour and could be best described by the Herschel–Bulkley model. The rate index and yield stress of cocoyam starch paste was higher than that of cassava. The storage modulus of cocoyam starch paste was higher than that of cassava indicating that cocoyam starch paste was more rigid than cassava starch paste. Cassava starch paste exhibited higher paste clarity and freeze-thaw stability than cocoyam starch paste. The properties of cassava and cocoyam starches dictate their food applications.  相似文献   

2.
Starch obtained from yellow and white plantain varieties were subjected to proximate analysis, physicochemical and rheological characterization in order to evaluate their properties. Yellow plantain variety gave higher yield of starch than the white variety. The two varieties differed in the purity of starch extract; white plantain starch contained: ash (1.09%), protein (0.640%) and fat (0.276%) while yellow plantain starch contained: ash (0.95%), protein (0.325%) and fat (0.403%). The amylose content of yellow plantain starch (24.36% (apparent), 26.13% (total)) was similar to that of white plantain starch (24.24% (apparent), 26.01% (total)). Scanning electron microscopy revealed bimodal irregular shaped granules (3.74–7.00 and 10.00–33.00 μm) in white plantain starch and elliptical granules (11.22–41.00 μm) in yellow plantain starch. Both starches differed markedly in their physicochemical properties. Their differences in gelatinization temperature (yellow plantain, 64.99–73.90 °C; white plantain, 68.08–77.15 °C), swelling and solubility patterns, and pasting characteristics indicated that yellow plantain starch had weaker granule architecture compared with white plantain starch. Further evidence of differences in properties was obtained from flow and viscoelastic properties of the starch gels, paste clarity and freeze–thaw stability.  相似文献   

3.
Starch from AC Hill oat grains (Avena nuda) was isolated and some of the characteristics determined. The yield of starch was 23·4% on a whole grain basis. The shape of the granule was polyhedral to irregular, with granules 6–10 μm in diameter. Lipids were extracted by acid hydrolysis and by selective solvent extraction with chloroform-methanol 2:1 v/v (CM) at ambient temperature, followed by n-propanol-water 3:1 v/v (PW) at 90–100°C. The acid hydrolyzed extracts which represented the total starch lipids (TSL) was 1·13%. The free lipids in the CM extract (1% TSL) was 6·2%, whereas the free and bound lipids in the PW extracts was 93.0%. Neutral lipids formed the major lipid class in the CM and PW extracts. The monoacyl lipid content in both CM and PW extracts was 61·0%. The total amylose content was 19·4%, of which 13·9% was complexed by native lipids. X-ray diffraction was of the ‘A’ type. Oat starch differed from wheat starch in showing a higher swelling factor, decreased amylose leaching, coleaching of a branched starch component and amylose during the pasting process, higher peak viscosity and set-back, low gel rigidity, greater susceptibility towards acid hydrolysis, greater resistance to -amylase action and a higher freeze-thaw stability. Furthermore, in comparison to wheat starch, the amylose chains of oat starch appear to be more loosely arranged in the amorphous regions, whereas in crystalline regions, oat starch chains are more compactly packed. Lipid removal from oat and wheat starches decreased their swelling factor, peak viscosity, set-back, gelatinization temperatures, freeze-thaw stability and paste clarity (at pH > 4·0), and increased their thermal stability, amylose leaching, enthalpy of gelatinization, susceptibility towards -amylase and paste clarity (at pH < 4·0). The results also showed that the properties of AC Hill oat starch is not representative of oat starch in general.  相似文献   

4.
This study was carried out in order to compare and establish the changes in physicochemical properties of starch from four different cultivars of yam at various stages of maturity during growth. The results showed that the starch content of the four yam tubers increased as growth progressed and were in the range of 70.5–85.3% on a dry basis. The shapes of the starch granules were round to oval or angular in the four yams and the size of starch granule increased with growth time ranging from 10 to 40 μm. The X-ray diffraction patterns could be classified as typical of B-type starch for the four cultivars of yam starch. The transition temperature of gelatinization of the four yam starches decreased during maturity. The RVA parameters suggested that yam starch paste showed a lower breakdown at an early harvest time. It appeared to be thermo-stable during heating but had a high setback after cooling, which might result in a tendency towards high retrogradation. The results for pasting behaviors showed that higher amylose content was associated with a lower pasting temperature and a higher peak viscosity in these starches.  相似文献   

5.
Pure starches were isolated from white and red sorghum cultivated in Tidikelt, a hyper arid region situated in south Algeria. Amylose content, X-ray pattern and rheological properties of starches were examined. The amylose content in white sorghum starch (27.1%) was slightly higher than that in red sorghum (24.8%). The swelling power and the solubility behavior of both starches were nearly similar below 65 °C. At higher temperatures, starch isolated from the white sorghum cultivar showed higher swelling power and lower solubility index than pigmented sorghum starch. The pasting properties of starches determined by RVA, Rapid Visco Analyser showed different viscosity peaks. Red sorghum starch had a higher value (4731 cP) than white sorghum starch (4093 cP). For both sorghum, X-ray diffractograms exhibit an A-type diffraction pattern, typical of cereal starches and the relative degrees of crystallinity were estimated at 22.72% and 28.91%, respectively, for local white and red sorghum starch. DSC analysis revealed that sorghum starches present higher temperatures at the peak (70.60 and 72.28 °C for white and red sorghum starches, respectively) and lower gelatinization enthalpies (9.087 and 8.270 J/g for white and red sorghum starches, respectively) than other cereal starches.The results showed that physicochemical and functional properties of sorghum cultivar starches were influenced by the genotype and the environment.  相似文献   

6.
The physico-chemical, morphological, thermal, pasting, textural, and retrogradation properties of the starches isolated from four traditional Taewa (Maori potato) cultivars (Karuparera, Tutaekuri, Huakaroro, Moemoe) of New Zealand were studied and compared with starch properties of a modern potato cultivar (Nadine). The relationships between the different starch characteristics were quantified using Pearson correlation and principal component analysis. Significant differences were observed among physico-chemical properties such as phosphorus content, amylose content, swelling power, solubility and light transmittance of starches from the different potato cultivars. The starch granule morphology (size and shape) for all the potato cultivars showed considerable variation when studied by scanning electron microscopy and particle size analysis. Starch granules from Nadine and Moemoe cultivars showed the presence of large and irregular or cuboid granules in fairly high number compared with the starches from the other cultivars. The transition temperatures (To; Tp; Tc) and the enthalpies (ΔHgel) associated with gelatinization suggested differences in the stability of the crystalline structures among these potato starches. The Moemoe starch showed the lowest To, while it was higher for Tutaekuri and Karuparera starches. Pasting properties such as peak, final and breakdown viscosity and texture profile analysis (TPA) parameters of starch gels such as hardness and fracturability were found to be higher for Nadine and Huakaroro starches. The Nadine and Huakaroro starch gels also had lower tendency towards retrogradation as evidenced by their lower syneresis (%) during storage at 4 °C. Principal component analysis showed that the Tutaekuri and Nadine cultivars differed to the greatest degree in terms of the properties of their starches.  相似文献   

7.
Physicochemical properties of etherified maize starches   总被引:2,自引:0,他引:2  
The changes in starch properties due to etherification with allyl glycidyl ether (AGE) have been investigated. After etherification of three different starches (containing 0.9%, 27% and 70% amylose), no appreciable differences in granular appearance were observed, but the granule crystallinity of these starches was changed. Furthermore, the incorporation of AGE in the starch significantly affects its physicochemical properties: the gelatinization temperatures were decreased and the pasting properties were altered. Both the swelling power and the solubility index increased as the degree of substitution (DS) increased. The rheology behaviour of the droplets of swollen granules suspension was studied under shear flow conditions.  相似文献   

8.
Potential roles of ultra high pressure (UHP) in starch granule reactivity and properties of acetylated starch were investigated. Corn starch was substituted with acetic anhydride at pressure range of 0.1–400 MPa for 15 min; also, conventional reaction (30 °C, 60 min) was conducted as reaction control. Native and acetylated corn starches were assessed with respect to degree of substitution (DS), X-ray diffraction pattern/relative crystallinity, starch solubility/swelling power, gelatinization, and pasting behavior. For the UHP-assisted acetylated starches, DS values increased along with increasing pressure levels from 200 to 400 MPa, and reaction at 400 MPa exhibited maximum reactivity (though lower than the DS value of the reaction control). Both UHP-assisted and conventional acetylation of starch likely occurred predominantly at amorphous regions within granules. Gelatinization and pasting properties of the UHP-assisted acetylated starches may be less influenced by UHP treatment in acetylation reaction, though restricted starch solubility/swelling were observed.  相似文献   

9.
In this work, the chayote tuber starch (CHS) was isolated and its chemical composition and its physical and microscopic characteristics were determined, and compared with potato (PS) and maize (MS) starches. The starch content in chayote tubers (728 g kg−1 dry weight) was similar to potato tubers (700 g kg−1 dry weight), with a high level of purity (>98%), while its phosphorous content was higher (0.15%) than PS (0.08%) and MS (0.01%). Starch granules were oval, irregular, truncated and rounded with sizes between 7 and 50 μm with smooth surfaces. CHS dispersions (1% and 4%, w/w) showed higher viscosity (75 and 1715 mPa s), than PS (350% and 50% lower) and MS (715% and 600% lower). The gelatinization temperature (65 to 74 °C) was similar in CHS and PS. The pasting properties (RVA) of the starches suggest that CHS showed better characteristics than the commercial potato and maize starches. Therefore, CHS could be used as a thickening agent and a substitute to PS in food dispersions where a high viscosity is needed.  相似文献   

10.
Native new cocoyam starch (nNCS) was subjected to annealing (aNCS) and heat moisture treatment at 18% moisture level (h18NCS), 21% moisture level (h21NCS), 24% moisture level (h24NCS) and 27% moisture level (h27NCS) as hydrothermal treatments. Scanning electron and light microscopy revealed round and polygonal shapes with sizes ranging from 15 to 40 μm for native and modified starches. nNCS showed “A” pattern X-ray diffraction and no significant differences were observed in the X-ray pattern of the modified starches. Swelling power and solubility reduced following heat moisture treatment. At all pH studied (2–12), unmodified new cocoyam starch exhibited higher swelling capacity and solubility than the modified derivatives. Hydrothermal modifications improved water absorption capacity but reduced oil absorption capacity. Pasting temperature of native starch shifted to higher values following annealing and heat moisture treatment. Hot paste viscosity (Hv), viscosity after 30 min holding at 95 °C (Hv30) and cold paste viscosity (Cv) reduced after annealing and heat moisture treatment. The result also indicates that hydrothermal treatments reduced the tendency for setback. As the number of days of storage of starch paste increased from 1 to 10, light transmittance of all the starches reduced but marked reduction of light transmittance was observed in native starch. DSC studies revealed increase in gelatinization temperature following annealing and heat moisture treatment. Starch hydrothermal modifications reduced retrogradation as enthalpies of regelatinization reduced following modifications. The regelatinization peak in the second day scanning shifted to lower temperature than the gelatinization peak in first run heating DSC curve for all samples. The regelatinization peak also became larger and shifted to higher temperature range when the storage days increased from 2 to 7.  相似文献   

11.
Recent novel mixed blooms of several species of toxic raphidophytes have caused fish kills and raised health concerns in the highly eutrophic Inland Bays of Delaware, USA. The factors that control their growth and dominance are not clear, including how these multi-species HAB events can persist without competitive exclusion occurring. We compared and contrasted the relative environmental niches of sympatric Chattonella subsalsa and Heterosigma akashiwo isolates from the bays using classic Monod-type experiments. C. subsalsa grew over a temperature range from 10 to 30 °C and a salinity range of 5–30 psu, with optimal growth occurring from 20 to 30 °C and 15 to 25 psu. H. akashiwo had similar upper temperature and salinity tolerances but also lower limits, with growth occurring from 4 to 30 °C and 5 to 30 psu and optimal growth between 16 and 30 °C and 10 and 30 psu. These culture results were confirmed by field observations of bloom occurrences in the Inland Bays. Maximum nutrient-saturated growth rates (μmax) for C. subsalsa were 0.6 d−1 and half-saturation concentrations for growth (Ks) were 9 μM for nitrate, 1.5 μM for ammonium, and 0.8 μM for phosphate. μmax of H. akashiwo (0.7 d−1) was slightly higher than C. subsalsa, but Ks values were nearly an order of magnitude lower at 0.3 μM for nitrate, 0.3 μM for ammonium, and 0.2 μM for phosphate. H. akashiwo is able to grow on urea but C. subsalsa cannot, while both can use glutamic acid. Cell yield experiments at environmentally relevant levels suggested an apparent preference by C. subsalsa for ammonium as a nitrogen source, while H. akashiwo produced more biomass on nitrate. Light intensity affected both species similarly, with the same growth responses for each over a range from 100 to 600 μmol photons m−2 s−1. Factors not examined here may allow C. subsalsa to persist during multi-species blooms in the bays, despite being competitively inferior to H. akashiwo under most conditions of nutrient availability, temperature, and salinity.  相似文献   

12.
Relationships between swelling capacities, pasting properties, rotational flow behaviour and textural properties of hydro-thermally heated wet-milled starch granules from corn dried between 60 and 130 °C were investigated. High-drying temperatures applied during the corn drying process conferred to the wet-milled starch granules (WSG) such a rigidity which reduced their swelling capacities, their water binding capacities and their water solubility index after gelatinization. These granules changes affected their pasting characteristics, their flow behaviour and several textural parameters of gel formed from the wet-milled starch granule after gelatinization. The rigidity of granules was a major factor determining the formation of either starch pastes or gels.  相似文献   

13.
Cryosurgery offers a promising therapeutic alternative for the treatment of prostate cancer. While often successful, complete cryoablation of cancerous tissues sometimes fails due to technical challenges. Factors such as the end temperature, cooling rate, duration of the freezing episode, and repetition of the freezing cycle have been reported to influence cryosurgical outcome. Accordingly, we investigated the effects of these variables in an in vitro prostate cancer model. Human prostate cancer PC-3 and LNCaP cultures were exposed to a range of sub-zero temperatures (−5 to −40 °C), and cells were thawed followed by return to 37 °C. Post-thaw viability was assessed using a variety of fluorescent probes including alamarBlue™ (metabolic activity), calceinAM (membrane integrity), and propidium iodide (necrosis). Freeze duration following ice nucleation was investigated using single and double freezing cycles (5, 10, and 20 min). The results demonstrated that lower freezing temperatures yielded greater cell death, and that LNCaP cells were more susceptible to freezing than PC-3 cells. At −15 °C, PC-3 yielded 55% viability versus 20% viability for LNCaP. Double freezing cycles were found to be more than twice as destructive versus a single freeze–thaw cycle. Both cell types experienced increased cell death when exposed to freezing temperatures for longer durations. When thawing rates were considered, passive (slower) thawing following freezing yielded greater cell death than active (faster) thawing. A 20% difference in viability between passive and active thawing was observed for PC-3 for a 10 min freeze. Finally, the results demonstrate that just reaching −40 °C in vitro may not be sufficient to obtain complete cell death. The data support the use of extended freeze times, multiple freeze–thaw cycles, and passive thawing to provide maximum cell destruction.  相似文献   

14.
Some hexavalent chromium [Cr(VI)]-containing compounds are lung carcinogens. Once within cells, Cr(VI) is reduced to trivalent chromium [Cr(III)] which displays an affinity for both DNA bases and the phosphate backbone. A diverse array of genetic lesions is produced by Cr including Cr–DNA monoadducts, DNA interstrand crosslinks (ICLs), DNA–Cr–protein crosslinks (DPCs), abasic sites, DNA strand breaks and oxidized bases. Despite the large amount of information available on the genotoxicity of Cr, little is known regarding the molecular mechanisms involved in the removal of these lesions from damaged DNA. Recent work indicates that nucleotide excision repair (NER) is involved in the processing of Cr–DNA adducts in human and rodent cells. In order to better understand this process at the molecular level and begin to identify the Cr–DNA adducts processed by NER, the incision of CrCl3 [Cr(III)]-damaged plasmid DNA was studied using a thermal-resistant UvrABC NER endonuclease from Bacillus caldotenax (Bca). Treatment of plasmid DNA with Cr(III) (as CrCl3) increased DNA binding as a function of dose. For example, at a Cr(III) concentration of 1 μM we observed 2 Cr(III)–DNA adducts per plasmid. At this same concentration of Cr(III) we found that 17% of the plasmid DNA contained ICLs (0.2 ICLs/plasmid). When plasmid DNA treated with Cr(III) (1 μM) was incubated with Bca UvrABC we observed 0.8 incisions/plasmid. The formation of endonuclease IV-sensitive abasic lesions or Fpg-sensitive oxidized DNA bases was not detected suggesting that the incision of Cr(III)-damaged plasmid DNA by UvrABC was not related to the generation of oxidized DNA damage. Taken together, our data suggest that a sub-fraction of Cr(III)–DNA adducts is recognized and processed by the prokaryotic NER machinery and that ICLs are not necessarily the sole lesions generated by Cr(III) that are substrates for NER.  相似文献   

15.
β-Turmerin from turmeric (Curcuma longa) waste grits obtained after extraction of curcumin was purified by successive gel permeation chromatography. Homogeneity of β-turmerin was confirmed by its movement as single band both in SDS-PAGE and as well as in native (basic) PAGE. The apparent molecular mass is 34 kDa by SDS-PAGE. It is more hydrophobic protein and showed sharp single peak in RP-HPLC with retention time of 62.17 min. It is a glycoprotein as it shows the presence of amino sugars up to 0.021 gm%. In three different model systems i.e., linolenic acid micelles, erythrocyte membrane systems and liposomes, β-turmerin at 0.125 μM offered 70%, 64%, and 60% inhibition of lipid peroxidation, which is 3200 times more efficient than the standard antioxidants BHA (400 μM) and α-tocopherol (400 μM). β-turmerin inhibited diene–triene and tetraene conjugation up to 54%, 72% and 47%, respectively. β-turmerin also effectively scavenges hydroxyl radicals when compared to BHA and α-tocopherol. β-turmerin (2.5 μM) further inhibited the activation of PMNL mediated by fMLP up to the extent of 75%, where as standards BHA (400 μM) and mannitol (10 μM) inhibited the same to 65% and 55%, respectively. At 0.125 μM dose β-turmerin prevented t-BOOH induced cell death at all time intervals. In addition to the above properties, it is non-toxic to lymphocytes as it did not affect the viability of cells. The mechanism of antioxidant action of β-turmerin could probably be by counteracting/quenching of reactive oxygen species (ROS). We report the purification and characterization of β-turmerin (34 kDa), a potent antioxidant protein from turmeric waste grits.  相似文献   

16.
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   

17.
Guan  Haining  Diao  Xiaoqin  Han  Jianchun  Kong  Baohua  Liu  Dengyong 《Food biophysics》2021,16(3):395-405

The influences of soy protein isolate hydrolysate (SPIH) obtained during different pressure treatments for 4 h on pasting and short-term retrogradation behaviors of maize starch (MS) were investigated. The results showed solubility of MS markedly increased, whereas swelling power decreased with increased SPIH concentration and pressure. Compared with native MS, the addition of SPIH led to decrease of peak viscosity, final viscosity, setback, and breakdown, whereas pasting temperature was increased. Meanwhile, differential scanning calorimetry (DSC) analysis also showed an increase in gelatinization temperature. In addition, low-field nuclear magnetic resonance (LF-NMR) analysis indicated that the tight association of water and starch molecules was observed with increasing pressures and additions of SPIH. Confocal laser scanning microscopy (CLSM) and atomic force microscope (AFM) images indicated that SPIH obtained at 200 MPa dispersed in the MS gel system to block the formation of hydrogen bonds and inhibit the recrystallization of MS. Fourier transform infrared (FTIR) spectroscopy analysis demonstrated that the addition of SPIH resulted in a decrease in hydrogen bonds within the starch molecules and the result supported above CLSM and AFM measurement results. The results proved that the addition of SPIH could effectively influence pasting characteristics and inhibit the short-term retrogradation of MS, which can be helpful to the application of SPIH in starch-based functional foods.

  相似文献   

18.
One and a half year-old Ginkgo saplings were grown for 2 years in 7 litre pots with medium fertile soil at ambient air CO2 concentration and at 700 μmol mol−1 CO2 in temperature and humidity-controlled cabinets standing in the field. In the middle of the 2nd season of CO2 enrichment, CO2 exchange and transpiration in response to CO2 concentration was measured with a mini-cuvette system. In addition, the same measurements were conducted in the crown of one 60-year-old tree in the field. Number of leaves/tree was enhanced by elevated CO2 and specific leaf area decreased significantly.CO2 compensation points were reached at 75–84 μmol mol−1 CO2. Gas exchange of Ginkgo saplings reacted more intensively upon CO2 than those of the adult Ginkgo. On an average, stomatal conductance decreased by 30% as CO2 concentration increased from 30 to 1000 μmol mol−1 CO2. Water use efficiency of net photosynthesis was positively correlated with CO2 concentration levels. Saturation of net photosynthesis and lowest level of stomatal conductance was reached by the leaves of Ginkgo saplings at >1000 μmol mol−1 CO2. Acclimation of leaf net CO2 assimilation to the elevated CO2 concentration at growth occurred after 2 years of exposure. Maximum of net CO2 assimilation was 56% higher at ambient air CO2 concentration than at 700 μmol mol−1 CO2.  相似文献   

19.
Biofilms were collected on discoloured limestone samples and on adhesive tape from historic buildings at the Mayan site of Edzna, in Campeche, Mexico. Grey, brown, and black areas were colonised predominantly by coccoid and colonial cyanobacteria, also detected as endoliths. The major biomass on the pink stone surface was Trentepohlia. At low power, surface micropits containing crustose-lichen-like black colonies were seen within these pink areas. These were apparently pure colonies of T. umbrina; no filamentous fungi were detected here. The black colour was produced by dense, amorphous pigment deposits in the colony centres. Insolation levels were high (400 μmphotons m−2 s−1) and the moisture content of the stone was 0.7%. The high carotenoid:chlorophyll-α ratio of Trentepohlia biofilms (3.9:1) is an adaptation protecting against environmental stress. Trentepohlia caused severe local erosion (pitting) and, when present as a more uniform biofilm, the well-known pink surface discoloration.  相似文献   

20.
Studies have been undertaken to investigate the effect of sugars on the thermal and rheological properties of sago starch. Sugars were found to increase the gelatinization temperature Tgel, and gelatinization enthalpy ΔH. Tgel and ΔH increased in the following order: control (water alone) < ribose < fructose < glucose < maltose < sucrose. The increase in ΔH was greater for 50% starch compared to 10% starch samples. The swelling factors in the presence of sugar were higher compared to the control for sugar concentrations below 25% but were lower at sugar concentration greater than 25%. These effects are discussed in terms of the antiplaticizing effect of the sugars compared to water, the influence of sugar–starch interactions and also the effect of the sugars on water structure. The storage modulus G′, the rate constant of gelation k, and the gel strength were significantly reduced in the presence of sugars. Generally G′ and k decreased in the following order: control (water alone) > hexose > disaccharide > pentoses. This has been attributed to the reduced proportion of amylose leached following gelatinizatison. In the presence of hexoses the freeze–thaw stability of starch gels decreased while in the presence of disaccharides and pentoses the freeze–thaw stability was slightly improved. © 1999 John Wiley & Sons, Inc. Biopoly 50: 401–412, 1999  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