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 19 毫秒
1.
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   

2.
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   

3.
Starch from tubers potato (Solanum tuberosum), taro (Alocassia indica), new cocoyam (Xanthosoma sagitifolium), true yam (Dioscorea alata), and root cassava, (Manihot esculenta) crops was isolated and its morphology, composition and physicochemical properties were investigated before and after heat–moisture treatment (HMT) (100 °C, for 10 h at a moisture content of 30%). Native starch granules were round to oval to polygonal with smooth surfaces. The granule size (diameter) ranged from 3.0 to 110 μm.The total amylose content ranged from 22.4 to 29.3%, of which 10.1–15.5% was complexed by native lipid. The phosphorus content ranged from 0.01 to 0.1%. The X-ray pattern of potato and true yam was of the ‘B’-type. Whereas, that of new cocoyam and taro was of the ‘A’-type. Cassava exhibited a mixed ‘A+B’-type X-ray pattern. The relative crystallinity, swelling factor (SF), amylose leaching (AML), gelatinization temperature range and the enthalpy of gelatinization of the native starches ranged from 30 to 46, 22 to 54, 5 to 23%, 13 to 19 °C and 12 to 18 J/g, respectively. Susceptibility of native starches towards hydrolysis by 2.2N HCl and porcine pancreatic -amylase were 60–86% (after 12 days), and 4–62% (after 72 h), respectively. Retrogradation was most pronounced in the B-type starches. Granule morphology remained unchanged after HMT. The X-ray pattern of the B-type starches was altered (B→A+B) on HMT. However, that of the other starches remained unchanged. HMT decreased SF, AML, gelatinization enthalpy and susceptibility towards acid hydrolysis, but increased gelatinization temperatures and enzyme susceptibility. Extent of retrogradation and relative crystallinity decreased on HMT of true yam and potato starches, but remained unchanged in the other starches. The foregoing data showed that changes in physicochemical properties on HMT are influenced by the interplay of crystallite disruption, starch chain associations and disruption of double helices in the amorphous regions.  相似文献   

4.
Modification of maize starch by thermal processing in glacial acetic acid   总被引:3,自引:0,他引:3  
Differential scanning calorimetry (DSC) and X-ray diffraction (XRD) methods were used to determine if corn starch–glacial acetic acid mixtures can be melted and thermally processed at reasonable temperatures. DSC studies showed that the melting temperature of dry starch was reduced from about 280 to 180°C in the presence of >30% acetic acid. Glass transition temperatures varied from 110 to 40°C at 15 and 45% acetic acid, respectively. XRD showed the loss of native starch crystallinity and the formation of V-type complexes. Addition of 10% water decreased the melting temperatures to 140–150°C while addition of a base (sodium acetate) had little effect. Some possible applications of processing starch in glacial acetic acid will be discussed.  相似文献   

5.
Binary systems of polymers often display spherulitic morphologies after cooling from the melt, but these phenomena have rarely been reported among food polymers of native-size. Here we report the observation of spherulitic and other morphologies in gelatinized maize starch. The morphology could be manipulated by choosing polymer compositions and kinetic regimes. Spherulites (10 μm diameter) formed from gelatinized high-amylose maize starches and purified amylose at cooling rates of order of magnitude 100 °C/min. They were more numerous and exhibited a higher melting point the greater the ratio of amylose to amylopectin. Rapid cooling rates (150–500 °C/min) resulted in a more even distribution of smaller spherulites. Very rapid (liquid nitrogen quench) or slow (0.1–1 °C/min) cooling rates resulted in mixed morphology, as did addition of 15 or 60% (w/w) sucrose to a 10% (w/w) dispersion of high-amylose starch (HAS). Spherulites were observed in aqueous suspensions of high-amylose maize starch between 5 and 30% (w/w). Lower starch concentrations resulted in a broader size distribution and spherulites of more distinct shape. WAXS patterns of B-type were observed. Negatively birefringent spherulites predominated, but positive spherulites were found. The spherulite melting range overlapped with that for amylose–lipid complex. Evidence indicated that micro-phase separation takes place when a holding period at 95 °C follows gelatinization at 180 °C. Despite the high maximum temperature of treatment (180 °C) there was evidence for a memory effect in samples of 30% HAS. Spherulite morphology closely resembled that of native starch granules in very early stages of development.  相似文献   

6.
Raw cassava starch, having 74.94 and 0.44 g/100 g resistant starch type II and III (RS II and RS III), respectively, was autoclaved at 121 °C in water, 1, 10 or 100 mmol/L lactic acid. The formation of RS III was evaluated in relation to variable incubation temperature (−20 to 100 °C), incubation time (6–48 h) and autoclaving time (15–90 min). Negligible to low quantities of RS III (0.59–2.42 g/100 g) were formed from autoclaved starch suspended in 100 mmol/L lactic acid, whereas intermediate to high quantities (2.68–9.97 g/100 g) were formed from autoclaved starch suspended in water, 1 or 10 mmol/L lactic acid, except for treatments with water or 10 mmol/L lactic acid incubated at 100 °C for 6 h (1.74 g/100 g). Autoclaving times corresponding to maximum RS III contents were 15 and 45 min for water and 10 mmol/L lactic acid, respectively. Whereas, the RS III fractions from cassava starch suspended in water had melt transitions between 158 and 175 °C with low endothermic enthalpies (0.2–1.6 J/g), the thermal transitions of the acid treated samples were indistinct.  相似文献   

7.
粉末X射线衍射图谱计算植物淀粉结晶度方法的探讨   总被引:6,自引:0,他引:6  
徐斌  满建民  韦存虚 《植物学报》2012,47(3):278-285
植物淀粉有A-型、B-型和C-型3种晶体。以水稻(Oryza sativa)、马铃薯(Solanum tuberosum)、豌豆(Pisum sativum) 和莲藕(Nelumbo nucifera)淀粉为材料, 利用粉末X-射线衍射仪(XRD)调查了不同晶体类型淀粉的波谱特征, 探讨XRD波谱相对结晶度的计算方法。软件峰拟合法、软件曲线法、直线作图法和曲线作图法均可用于计算淀粉XRD波谱的相对结晶度, 以曲线作图法计算结果较为可靠。利用曲线作图法得出的结果表明, 稻米淀粉的结晶度与直链淀粉含量呈显著线性负相关, 酸解莲藕淀粉的结晶度与淀粉酸水解度呈显著线性正相关。酸水解使莲藕淀粉的C-型晶体转变为A-型晶体。上述研究结果为利用XRD分析植物淀粉晶体类型和计算相对结晶度提供了重要参考。  相似文献   

8.
Starch isolated from mature Ginkgo biloba seeds and commercial normal maize starches were subjected to α-amylolysis and acid hydrolysis. Ginkgo starch was more resistant to pancreatic α-amylase hydrolysis than the normal maize starch. The chain length distribution of debranched amylopectin of the starches was analyzed by using high performance anion-exchange chromatography equipped with an amyloglucosidase reactor and a pulsed amperometric detector. The chain length distribution of ginkgo amylopectin showed higher amounts of both short and long chains compared to maize starch. Naegeli dextrins of the starches prepared by extensive acid hydrolysis over 12 days demonstrated that ginkgo starch was more susceptible than normal maize to acid hydrolysis. Ginkgo dextrins also demonstrate a lower concentration of singly branched chains than maize dextrins, and unlike maize dextrin, debranched ginkgo shows no multiple branched chains. The ginkgo starch displayed a C-type X-ray diffraction pattern, compared to an A-type pattern for maize. Ginkgo starch and maize starch contained 24.0 and 17.6% absolute amylose contents, respectively.  相似文献   

9.
Activity of starch synthase and the amylose content in rice endosperm   总被引:30,自引:0,他引:30  
The content of amylose in endosperm of non-waxy japonica rice (Oryza sativa cv Akitakomachi) was increased by lowering the growth temperature from 25° to 15° during the ripening period. The activities of sucrose synthase, ADPglucose pyrophosphorylase, starch branching enzyme (Q-enzyme) and soluble starch synthase in endosperm developed at 15° were lower than or similar to those at 25°, when compared on a endosperm basis at the similar ripening stage. In contrast, the activity of starch granule-bound starch synthase, which is considered to be indispensable for amylose synthesis, was higher by 3–3.5-fold in the endosperm developed at the low temperature than that at the high ambient temperature. The results suggest that the low temperature specifically accelerates the expression of the bound starch synthase gene (waxy gene) in rice endosperm, which resulted in elevated amylose biosynthesis in the endosperm when developed at lower temperatures.  相似文献   

10.
Starch-poly (ethylene-co-acrylic acid) (EAA) complexes were prepared by jet-cooking mixtures of either cornstarch, waxy cornstarch or high amylose cornstarch with aqueous ammonia dispersions of EAA (4% EAA based on the weight of starch). Viscosities (η) were determined at temperatures ranging from 80°C to 22°C, and plots of log η versus 1/T (K−1) were prepared. When cooked with EAA, cornstarch and waxy cornstarch showed major changes in viscous behavior between 50°C and 60°C. Above 50–60°C, viscosity increased markedly with a reduction in temperature; however, viscosity increased slowly below 50–60°C with an apparent activation energy for the process approximating that of water itself. The temperature dependence of the measured viscosity from 80°C to 60°C could be attributed to the large increase in size and complexity of the flowing particles as individual amylopectin molecules were bound together by complexed EAA. Apparently, complexing is essentially complete at 50°C. When high amylose cornstarch was cooked in the absence of EAA, retrogradation produced a sharp increase in log η at temperatures below about 50°C. However, if EAA is present, association between amylose molecules apparently takes place via complex formation rather than retrogradation, since log η increases sharply at about 70–80°C. Also, in contrast to cornstarch and waxy cornstarch, log η versus 1/T plots for high amylose cornstarch did not level off at low temperatures. In general, viscosities increased with the pH of the system, particularly when η was measured at high temperatures. This could result from improved complexing ability of EAA under high pH conditions, possibly due to reduced micelle size and maximum extension of polymer chains from micelle surfaces.  相似文献   

11.
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   

12.
Water-soluble (WSP) and insoluble non-starch polysaccharides (WIP) were isolated from wheat flour to evaluate the effects of WSP and WIP on starch gel properties. Isolated WSP and WIP were added to two types of isolated wheat starch with different amylose content at a concentration of 3% based on the dry weight of starch. 30% starch gels were prepared and stored at 5 °C for 1, 8, or 24 h. The dynamic viscoelasticity of 30% starch gels mixed with WSP and WIP was measured using parallel plate geometry, showing that WSP and WIP affected the elastic component of starch gels in opposite ways. Adding WIP increased the storage shear modulus (G′) of starch gels, while adding WSP decreased G′ and dramatically increased the loss tangent (tan δ=G″/G′).  相似文献   

13.
The effects of starch granules on the rheological behaviour of gels of native potato and high amylopectin potato (HAPP) starches have been studied with small deformation oscillatory rheometry. The influence of granule remnants on the rheological properties of samples treated at 90 °C was evident when compared with samples treated at 140 °C, where no granule remnants were found. The presence of amylose in native potato starch gave to stronger network formation since potato starch gave higher moduli values than HAPP, after both 90 and 140 °C treatments. In addition, amylose may have strengthened the network of HAPP because higher moduli values were obtained when native potato starch was added to the system. The moduli values of the mixtures also increased with increasing polysaccharide concentration in the system, which is due to an increment in the polysaccharide chain contacts and entanglements. Finally, it was found that a mixture of commercial amylose from potato starch and HAPP resulted in lower values of G′ compared to native potato starch. This indicates that the source of amylose is important for the properties in a blend with native amylopectin.  相似文献   

14.
Physico-chemical characterisation of sago starch   总被引:3,自引:0,他引:3  
The physico-chemical characteristics of various sago starch samples from South East Asia were determined and compared to starches from other sources. X-ray diffraction studies showed that all the sago starches exhibited a C-type diffraction pattern. Scanning electron microscopy showed that they consist of oval granules with an average diameter around 30 μm. Proximate composition studies showed that the moisture content in the sago samples varied between 10.6% and 20.0%, ash between 0.06% and 0.43%, crude fat between 0.10% and 0.13%, fiber between 0.26% and 0.32% and crude protein between 0.19% and 0.25%. The amylose content varied between 24% and 31%. The percentage of amylose obtained by colourimetric determination agreed well with the values obtained by fractionation procedures and potentiometric titration. Intrinsic viscosities and weight average molecular weight were determined in 1M KOH. Intrinsic viscosity for amylose from sago starches varied between 310 and 460 ml/g while for amylopectin the values varied between 210 and 250 ml/g. The molecular weight for amylose was found to be in the range of 1.41×106 to 2.23×106 while for amylopectin it was in the range of 6.70×106 to 9.23×106. The gelatinisation temperature for the sago starches studied varied between 69.4°C and 70.1°C. The exponent ‘a’ in the Mark–Houwink equation and the exponent ‘’ in the equation Rg=kM was found to be 0.80 and 0.58, respectively for amylose separated from sago starch and these are indicative of a random coil conformation. Two types of pasting properties were observed. The first was characterised by a maximum consistency immediately followed by sharp decrease in consistency while the second type was characterised by a plateau when the maximum consistency was reached.  相似文献   

15.
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   

16.
The V-type X-ray diffraction pattern associated with amylose-lipid complexes disappeared when an amylose-lysolecithin complex was heated in the presence of cyclomaltoheptaose. When cyclomaltoheptaose was added to wheat starch or an amylose-lysolecithin complex, the melting enthalpy of the amylose-lysolecithin complex was lessened. Turbidity measurements showed that cyclomaltoheptaose disrupted the amylose-lysolecithin complexes at temperatures above 65°. Using differential scanning calorimetry and turbidity measurements, a molar binding ratio for the cyclomaltoheptaose-lysolecithin complex of 3:1 was obtained. This complex appeared to dissociate over a temperature range of 33–37° in an excess of water.  相似文献   

17.
Comparative studies of native maize starches with different amylose contents were carried out using X-ray powder diffraction. The results show a transition of crystalline type from A through C to B, accompanying a decrease in degree of crystallinity from 41.8% to 17.2% across a range of apparent amylose content from 0% to 84%. Hydration induces an increase in degree of granule crystallinity, but does not change the transition of crystal type. Progressively from A-type to C-type, crystallinity decreases rapidly with an increase in amylose content. From C-type to B-type, overall crystallinity decreases more slowly. The crystal type is strongly dependent on amylose content and on average chain length of the respective amylopectin. Waxy A-types have an average chain length of about 20, while in high amylose B-types this rises to ≈35. The proportion of short chains (10–13 glucose units) appears to affect crystal type significantly. Some V-type material was detected at high amylose levels. The proportion of this increased after prolonged exposure of the granules to iodine vapour. Implications for the arrangement of starch components in the granule are discussed.  相似文献   

18.
The formation of amylose–lipid complexes in a gelatinized potato starch matrix was investigated using potato starch and glycerol monopalmitin. These complexes exist in two forms, with the amounts of each of the forms being dependent on the temperatures and durations of the pre-treatments.

Differential scanning calorimetry (DSC) was used to analyze transition temperatures and melting enthalpies, and thereby determine the amount of the complexes in the samples. X-ray diffraction analysis was used to investigate their crystallinity.

In measurements with DSC, form I started to melt at 88.5°C, and form II at 112.9°C. When complex form II was preheated at 100 or 110°C, its melting point rose to 116.3 and 119.7°C, respectively, because of an annealing effect. The same phenomenon occurred with complex form I: when preheated at 90°C, its melting point rose to 96.8°C. The crystal formation of form II appeared to be slower when treated at 110°C than at 100°C. Their maximum melting enthalpies were reached after about 24 h and 4 h of preheating, respectively. In X-ray diffraction analyses, form II showed a V-pattern, but form I did not. This indicates that form II is more crystalline than form I. It was possible to transform form I into form II when it was heat treated, because form I was then partially or totally melted.

As a comparison, the charged substance cetyltrimethylammonium bromide created complex form I with amylose in the starch matrix, but not form II.  相似文献   


19.
The degradation of pea starch granules by acid hydrolysis has been investigated using a range of chemical and structural methods, namely through measuring changes in amylose content by both the iodine binding and concanavalin A precipitation methods, along with small angle X-ray scattering (SAXS), wide angle X-ray diffraction (XRD) and field emission scanning electron microscopy (FE-SEM). The relative crystallinity, intensity of the lamellar peak and the low-q scattering increased during the initial stages of acid hydrolysis, indicating early degradation of the amorphous regions (growth rings and lamellae). In the first 2 days of hydrolysis, there was a rapid decline in amylose content, a concomitant loss of precipitability of amylopectin by concanavalin A, and damage to the surface and internal granular structures was evident. These observations are consistent with both amylose and amylopectin being located on the surface of the granules and attacked simultaneously in the early stages of acid hydrolysis. The results are also consistent with amylose being more concentrated at the core of the granules. More extensive hydrolysis resulted in the simultaneous disruption of amorphous and crystalline regions, which was indicated by a decrease in lamellar peak intensity, decrease in interhelix peak intensity and no further increase in crystallinity. These results provide new insights into the organization of starch granules.  相似文献   

20.
Compatibility, flow and visco-elastic properties of a pregelatinised maize starch mixed with κ-carrageenan were investigated. After cooking of the pregelatinised starch, some undissolved granules remained in solution. Aqueous mixtures of κ-carrageenan and starch were studied at 60 °C and 20 °C by combining rheological measurements and microscopic observations under conditions allowing gelation of carrageenan and non-gelation of starch. The viscometric study of mixed dilute solutions of amylose from pregelatinised starch and carrageenan showed that the components are slightly incompatible. Mixture viscosity and elastic modulus were studied at 60 °C in details as a function of mixture composition for a total polymer concentration of 3%; both were found to be significantly higher than the corresponding theoretical additive values. This finding was interpreted by starch granules excluded volume effect. At 20 °C, no noticeable increase of mixture elastic modulus was found as compared with the additive value. The absence of the synergistic effect is supposed to be due to the formation of highly inhomogeneous gels with agglomerates of undissolved granules.  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