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.
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   

2.
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   

3.
Starch phosphate is important in starch metabolism and in order to deduce its location and structural effects in clusters and building blocks of amylopectin, these were isolated from a normal potato (WT) and two starches with antisense suppressed glucan water dikinase (asGWD) activity and starch branching enzyme (asSBE) activity possessing suppressed and increased phosphate contents, respectively. Neutral N-chains and phosphorylated P-chains of the amylopectin macromolecules were similar in WT and asGWD, whereas asSBE possessed considerably longer P-chains. Cluster β-limit dextrins were isolated by α-amylase treatment and successive β-amylolysis. Cluster sizes were generally smaller in asSBE. The building block composition of neutral N-clusters were very similar in WT and asGWD, while asSBE was different, containing less blocks with degree of polymerization (DP)>14. Phosphate content of the P-clusters of WT and asGWD was rather similar, while asSBE contained highly phosphorylated P-clusters with proportionally more P-chains and a low degree of branching. The average chain lengths of the P-clusters were, however, similar in all samples. Our data demonstrate only minor effect on the cluster structure in relation to phosphate deposition suggesting conserved reaction patterns of starch phosphorylation. Models are suggested to account for the principle structural and functional effects of starch phosphate esters.  相似文献   

4.
Human estrogenic 17β-hydroxysteroid dehydrogenase (17β-HSD1, EC1.1.1.62) is an important enzyme that catalyses the last step of active estrogen formation. 17β-HSD1 plays a key role in the proliferation of breast cancer cells. The three-dimensional structures of this enzyme and of the enzyme-estradiol complex have been solved (Zhu et al., 1993, J. Mol. Biol. 234:242; Ghosh et al., 1995, Structure 3:503; Azzi et al., 1996, Nature Struct. Biol. 3:665). The determination of the non-reactive ternary complex structure, which could mimic the transition state, constitutes a further critical step toward the rational design of inhibitors for this enzyme (Ghosh et al. 1995, Structure 3:503; Penning, 1996, Endocrine-Related Cancer, 3:41).

To further study the transition state, two non-reactive ternary complexes, 17β-HSD1–EM519-NADP+ and 17β-HSD1–EM553-NADP+ were crystallized using combined methods of soaking and co-crystallization. Although they belong to the same C2 space group, they have different unit cells, with a=155.59 Å, b=42.82 Å, c=121.15 Å, β=128.5° for 17β-HSD1–EM519-NADP+, and a=124.01 Å, b=45.16 Å, c=61.40 Å, β=99.2° for 17β-HSD1–EM553-NADP+, respectively. Our preliminary results revealed that the inhibitors interact differently with the enzyme than do the natural substrates.  相似文献   


5.
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   

6.
This study was conducted to determine the utility of deletion spectrum and mutant frequency (MF) of the hypoxanthine phosphoribosyl transferase gene (HPRT) as indicators of radiation exposure in Russian Liquidators who served in 1986 or 1987 in the clean up effort following the nuclear power plant accident at Chernobyl. HPRT MF was determined using the cloning assay for 117 Russian Controls and 122 Liquidators whose blood samples were obtained between 1991 and 1998. Only subjects from whom mutants were obtained for deletion analysis are included. Multiplex PCR analysis was performed on cell extracts of 1080 thioguanine resistant clones from Controls and 944 clones from Liquidators. Although the deletion spectra of Liquidators and Controls were similar overall, the Liquidator deletion spectrum was heterogeneous over time. Most notable, the proportion of total gene deletions was higher in 1991–1992 Liquidators than in Russian Controls (χ2=10.5, p=0.001) and in 1993–1994 Liquidators (χ2=8.3, p=0.004), and was marginally elevated relative to 1995–1996 Liquidators (χ2=3.3, p=0.07). This type of mutation has been highly associated with radiation exposure. Total gene deletions were not increased after 1992. Band shift mutations were also increased in the 1991–1992 Liquidators but were associated with increased MF of both Liquidators and Controls (p=0.009), not with increased MF in 1991–1992 Liquidators (p=0.7), and hence are not believed to be associated with radiation exposure. Regression analysis demonstrated that relative to Russian Controls HPRT MF was elevated in Liquidators overall when adjusted for age and smoking status (37%, p=0.0001), and also was elevated in Liquidators sampled in 1991–1992 (72%, p=0.0076), 1993–1994 (22%, p=0.037), and 1995–1996 (62%, p=0.0001). In summary, HPRT MF was found to be the more sensitive and persistent indicator of radiation exposure, but the specificity of total gene deletions led to detection of probable heterogeneity of radiation exposure within the exposed population.  相似文献   

7.
本研究对金针菇淀粉酶家族基因进行了信息分析,并选用金针菇双核菌株H1123作为实验材料,分析了菌丝生长过程中淀粉酶活性和淀粉酶基因表达特性之间的关系。结果表明,金针菇淀粉酶家族包含6个α淀粉酶和1个γ淀粉酶。7个淀粉酶基因的表达量均在菌丝接种后第10天出现峰值,并与胞外淀粉酶活性呈同步变化,说明基质中淀粉的分解和利用是淀粉酶家族各成员之间相互协调的结果。其中α-Amy-1α-Amy-4α-Amy-5的上调幅度最大,为淀粉降解和代谢过程的主效基因。值得注意的是胞内淀粉酶基因α-Amy-1在第10天时达到约90倍的上调表达水平。我们推测:金针菇胞外淀粉酶将淀粉分解为小分子单糖的同时,其胞内淀粉酶也参与了这些糖类的吸收和运输过程。  相似文献   

8.
9.
A new flexible sensor for in vitro experiments was developed to measure the surface potential, Φ, and its gradient, E (electric near field), at given sites of the heart. During depolarisation, E describes a vector loop from which direction and magnitude of local conduction velocity θ can be computed. Four recording silver electrodes (14 μm × 14 μm) separated by 50 μm, conducting leads, and solderable pads were patterned on a 50 μm thick polyimide film. The conductive structures, except the electrodes, were isolated with polyimide, and electrodes were chlorided. Spacer pillars mounted on the tip fulfil two functions: they keep the electrodes 70 μm from the tissue allowing non-contact recording of Φ and prevent lateral slipping. The low mass (9.1 mg) and flexibility (6.33 N/m) of the sensor let it easily follow the movement of the beating heart without notable displacement. We examined the electrodes on criteria like rms-noise of Φ, signal-to-noise ratio of Φ and E, maximum peak-slope recording dΦ/dt, and deviation of local activation time (LAT) from a common signal and obtained values of 24–28 μV, 46 and 41 dB, 497–561 V/s and no differences, respectively. With appropriate data acquisition (sampling rate 100 kHz, 24-bit), we were able to record Φ and to monitor E and θ on-line from beat-to-beat even at heart rates of 600 beats/min. Moreover, this technique can discriminate between uncoupled cardiac activations (as occur in fibrotic tissue) separated by less than 1 mm and 1 ms.  相似文献   

10.
The interaction of β1 integrin receptors and different extracellular matrix molecules during neuronal development was investigated by comparing both migration and morphological differentiation of D3 wild-type embryonic stem (ES) cell line-derived neural precursor cells with those of the β1 integrin knockout ES cell line G201. Analysing neurosphere explants on laminin and fibronectin as major β1 integrin ligands, the maximal spreading of outward migrating neuronal cells was determined. Compared with gelatine as a standard substrate, migration was found to be significantly increased for D3-derived neurospheres on fibronectin and laminin-1. These matrix effects were found to be even enhanced for G201 preparations. In addition, also the differentiation of wild-type and β1 integrin −/− neurones – as determined by MAP-2- and HNK-1-immunoreactive processes – was found to be increased on fibronectin and laminin when compared to gelatine standards. In the respective knockout preparations on these matrices, again perturbation effects were less pronounced than on gelatine. Our observations indicate that laminin and fibronectin are involved both in β1 integrin-dependent and -independent signalling mechanisms during neurogenesis. Upregulation of compensatory mechanisms such as β1 integrin-independent receptors for laminin and fibronectin might be responsible for the much less pronounced perturbations of G201 neural precursor migration and differentiation on these two substrates than on gelatine.  相似文献   

11.
Fifteen starches from different botanical sources were selected to study the influence of structural features on thermal properties and enzyme digestibility. Morphological appearance, X-ray diffraction pattern, apparent amylose content, unit-chain length distribution of amylopectin, thermal properties and enzyme digestibility of starch varied with botanical source. It was demonstrated that the distribution of unit-chains of amylopectin significantly correlated with functional properties of the starches. Gelatinization temperature of native and retrograded starches decreased and increased with a relative abundance of unit-chains with an approximate degree of polymerization (DP) of 8–12 and 16–26, respectively (P<0.01). Similar unit-chain lengths also affected the enzyme digestibility of starch granules (P<0.01).  相似文献   

12.
In this study, the structures of the hot-water (80°C) soluble starch fractions (HWS) of six new Japanese rice cultivars (Saikai 194, Saikai 198, Hokuriku 149, Suigen 258, Hoshiyutaka, and Saikai 184) were investigated following a previous study [Mizukami, H., Hizukuri, S. and Takeda, Y. Structures and pasting properties of starches from new characteristic rice cultivars, Oyo Toshitu Kagaku (J. Appl. Glycosci.) 43 (1996) 15–23]. The HWS were subfractionated into 1-butanol-precipitate (SAM) and supernatant (SAP) fractions. The yields of the SAM and SAP fractions were 0.3%–2.4% and 3.1%–4.1% by starch weight, respectively. The Hoshiyutaka and the Saikai 184 yielded both relatively large (2.4%) and small amounts (0.3%) of SAM. The SAM were small amylose molecules with a n between 320 and 420 and a w between 950 and 1850. The SAM from the Hoshiyutaka and the Saikai 184 were the larger molecules with n 390 and 420, respectively, and having slightly more branches (6.0 and 8.1) than those from the others (1.5–4.5). The SAP were smaller molecules having a n between 60 and 190 as compared to the SAM. The SAP was composed of small amylopectin molecules ( n 280–790, CL 17–32, β-amylolysis limit (β-AL), 54%–68%) including very small amylose molecules ( n 24–34) having an average number of branch linkage ( ) of between 0.4 and 0.5. Both the amount and the structures of hot-water-extractable rice starch fractions vary with cultivar, and may influence their cooking properties.  相似文献   

13.
The rheological behaviour of liquid two phase gelatin–locust bean gum (LBG) systems, comprising of (a) liquid LBG enriched continuous phase, and (b) flow-deformable gelatin-enriched dispersed particles seems to be determined, at the same phase composition, by phase viscosity ratio (μ). In the μ range from 0.03 to 0.21, viscosity dropped to values noticeably lower (13–40 times) than those of the corresponding LBG solution. Decrease in the viscosity of the mixtures was not observed at μ=0.5–0.6, corresponding that to the maximum energy scatter inside the droplets, in agreement with Mason’s conception of droplet deformation and disruption of liquid Newtonian emulsions.  相似文献   

14.
Accurate and precise determination of phospholipid composition by 31P NMR spectroscopy requires correct assignments and adequate spectral resolution. Because temperature and pH may affect chemical shifts (δ), our first aim was to establish the temperature coefficient (ΔδT) of common phospholipid classes when using sodium cholate as detergent. This parameter can then be used to aid in resonance assignments. The second goal was to investigate the pH dependence of δ so that, in addition to temperature, pH control can be used to minimize spectral overlap. For phosphatidylcholine, sphingomyelin, dihydrosphingomyelin and phosphatidylglycerol, δ values were invariant with pH and temperature. Whereas the ΔδT for phosphatidylinositol was 4 × 10−3 ppm/°C, regardless of pH, these coefficients were highly pH-dependent for phosphatidic acid, phosphatidylethanolamine and phosphatidylserine, exhibiting maximal variations with the deprotonation of the headgroup, particularly for phosphatidic acid. These trends indicate the importance of H-bonding on δ and ΔδT for phospholipid resonances.  相似文献   

15.
The aim of this study was to examine the protection of the yeast lacking the “antioxidant-like” prion precursor protein (Ure2p), by antioxidants and to elucidate how modification of redox homeostasis affects toxicity of agents inducing oxidative stress in the Δure2 cells. We found a diverse ability of a range of antioxidants to ameliorate the hypersensitivity of the Δure2 disruptant to oxidants and heavy metal ions. Glutathione and then ascorbate were the most effective antioxidants; Tempol, Trolox and melatonin were much less effective or even hampered the growth of the Δure2 cells exposed to tested agents. The intracellular level of ROS was augmented in the Δure2 mutant under normal growth conditions (1.7-fold), and after treatment with H2O2 (2.3-fold) and Cd(II) (2.8-fold), with respect to its wild-type counterpart. Glutathione was unable to prevent the increase in ROS production caused by CdCl2. The Δure2 disruptant was also hypersensitive to heat shock, like mutants lacking glutathione S-transferases.  相似文献   

16.
Effects of proportions of neutral detergent fibre (aNDFom) and starch, as well as their degradation rates, on rumen fermentation were tested using an in vitro rumen simulation system (SIMCO). The in vitro system was designed to simulate selective particle retention and had an average fluid volume of 1150 ml with a liquid dilution rate of approximately 0.07 h−1. Two types of hay (aNDFom sources) and two types of starch were each included at two different levels in the diet and were examined in an experiment following a 2×2×2 factorial arrangement of treatments (eight diet combinations). The hay was either late-cut timothy (Phleum pretense L.) or early cut meadow grass (Poa pratensis L.), with ruminal in situ aNDFom digestion rates of 0.03–0.04 and 0.07–0.08 h−1, respectively. The two starch types were raw (R) and cooked (C) potato starch with previously determined in vitro ruminal digestion rates of 0.04 and 0.20 h−1, respectively. The starch levels were 300 and 600 g/kg diet dry matter (DM) with the remaining being hay (282–682 g/kg DM) and peptone (14–111 g/kg DM). The aNDFom level varied among the diets with different starch levels and hay types. The peptone acted as a source of peptides and, together with ammonia salts from buffer, was used to balance the N contents of the diets. The feeding level for each of the eight vessels was 28 g DM/d. Two 10-day simulations were made with the system. The average pH was higher (P<0.05) for all treatments with raw potato starch (6.19) versus cooked starch (6.07). Protozoa scores, on a qualitative scale, declined faster at the higher starch level. The aNDFom digestibility was, as expected, higher (P<0.001) for meadow hay (0.57) than timothy (0.32), and was also higher (P<0.001) at the lower starch level (0.54) versus the higher (0.35). Microbial protein production efficiency (mg microbial N/g organic matter truly digested) was higher for the faster degrading aNDFom (P<0.01) and starch (P<0.05) sources, but was not affected by starch level. Cooked starch resulted in a lower acetate proportion (449 mmol/mol versus 591 mmol/mol VFA; P<0.001) but higher proportions of propionate (297 mmol/mol versus 236 mmol/mol VFA; P<0.001), and butyrate (169 mmol/mol versus 127 mmol/mol VFA; P<0.01). Butyrate increased with starch level (127 mmol/mol versus 169 mmol/mol VFA; P<0.01), and was also higher for meadow hay versus timothy (168 mmol/mol versus 128 mmol/mol VFA; P<0.01). Interactions between the treatments demonstrate that the response in VFA pattern to starch level is dependent on starch and aNDFom sources. Substrates such as starch and aNDFom are fermented differently depending on their rates of ruminal degradation.  相似文献   

17.
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   

18.
19.
Four ergosterol derivatives (1–4) have been isolated for the first time from the fruiting bodies of a basidiomycete fungus, Lactarius hatsudake, through activity-guided fractionation. Their structures were determined, using spectroscopic analysis, as: (22E,24R)-ergosta-5,7,22-dien-3β-ol (ergosterol, 1); 5,8-epidioxy-(22E,24R)-ergosta-6,22-dien-3β-ol (ergosterol peroxide, 2); 5,8-epidioxy-(24S)-ergosta-6-en-3β-ol (3); and (22E,24R)-ergosta-7,22-dien-3β,5,6β-triol (cerevisterol, 4). Compounds 2 and 3 showed selective inhibitory activity against Crotalus adamenteus venom phospholipase A2 (PLA2) enzyme, but not against Apis mellifcra bee venom PLA2. The antiphospholipase A2 activity of compounds 2 and 3 are reported here for the first time.  相似文献   

20.
Synthesis of chitosan sulfates with low molecular weight (Mv 9000–35,000 Da) was carried out by sulfation of low molecular weight chitosan (Mv 10,000–50,000 Da). The oleum was used as sulfating agent and dimethylfornamide as medium. The chitosans were prepared by enzymatic and acidic hydrolysis of initial high molecular weight chitosan as well as by extrusion solid-state deacetylation of chitin. As was shown by FT-IR and NMR-methods and elemental analysis, the sulfation occurred at C-6 and C-3 positions and substitution degree is 1.10–1.63. The molecular weight sulfated chitosan was determined by viscometric method and the Mark–Houwink equation [η]=10−5 4.97 M0.77. Study of anticoagulant activity showed that chitosan sulfates with lowered molecular weight demonstrated a regular increase of anti-Xa activity like heparins.  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