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.
The gelatin-starch-water system has been studied at different temperatures, at a total biopolymer concentration of 5.0 wt%. The weight ratios (W) of gelatin/ starch used were 9:1, 8:2.4. 2:8, 1:9, with pH values between 5.82 (at W = 9:1) and 6.50 (at W = 1:9). The systems were characterized rheologically and by turbidity measurements to construct a phase diagram in the temperature (T) and composition (W) variables. The T-W quadrant consists of three regions: a singlephase solutions region (A) and regions of complete and incomplete phase separation (B and C, respectively). The system in region C is a gel. Region B, lying between A and C, corresponds to two co-existing liquid phases. The transition from A to C (obtained by cooling the system at constant W) involves crossing region B. The properties of the resulting gels depend on the rate of this intersection. Gels formed on rapid cooling have an even distribution of turbidity, whereas slow cooling gives two gel layers of different turbidity. The gelation temperature and gel strength of the mixed systems are dominated by the gelatin component, with no indication of network formation by starch.  相似文献   

2.
The thermophilic blue-green alga Synechococcus lividus was grown at 38 and 55°C. The reaction center chlorophyll-protein complexes (CP) of Photosystem (PS I) and PS II, CP aI and CP aII, were isolated by sodium dodecyl sulphate (SDS) polyacrylamide gel electrophoresis at 4°C. SDS solubilization of thylakoids was performed in the temperature range 0–65°C. The low-temperature absorption and fluorescence emission spectral properties of the isolated chlorophyll-protein complexes were analyzed. Only traces of CP aI were solubilized at temperatures below the lipid phase transition temperature. Instead, a minor PS I component, CP aI, was obtained that had absorption and fluorescence characteristics similar to those of CP aI. CP aI had a slightly lower mobility than CP aI in SDS-polyacrylamide gel electrophoresis. The amount of CP aI in the gel scan profile increased dramatically when solubilization was carried out above the phase transition temperatures, but started to decrease above 60°C. CP aII, on the other hand, could be efficiently extracted even at 0°C and was stable in the scan profile up to extraction temperatures of 30–40°C. Low-temperature absorption and fluorescence emission spectra were typical for CP aI and CP aII and no specific effects of the two growth temperatures on these properties were observed. The phase transition temperature was considered to be critical for the solubilization of CP aI, either because of the difficulties of SDS (especially as it forms micelles at low temperatures) in penetrating the solidified membrane lipids at temperatures below that of the phase transition or because the CP aI monomers of the PS I antennae are so strongly bound to each other that they cannot be dissociated by SDS before thermal agitation has reached a certain level that is achieved above the phase transition temperature. We consider both the difficulties in solubilizing CP aI at sub-transition temperatures and the heat stability of the two complexes as adaptations which enable Synechococcus to grow under extreme high-temperature regimes.  相似文献   

3.
A chlorophyll-protein was isolated from a Synechococcus P700-chlorophyll a-protein complex free from small subunits (CP1-e) by sodium dodecyl sulfate (SDS)-polyacrylamide gel electrophoresis after treatment with 2% 2-mercaptoethanol and 2% SDS. In contrast to CP1-e which, when electrophoresed under denaturating conditions, showed two polypeptide bands of 62 and 60 kDa, the chlorophyll-protein contained only the 60-kDa polypeptide and hence is called CP60. The yield of CP60 was maximal with 1-2% SDS and 2-4% sulfhydryl reagents because the chlorophyll-protein was denatured at higher concentrations of the reagents. The absorption spectrum of CP60, which retained more than half of the chlorophyll alpha molecules originally associated with the 60-kDa subunit of the photosystem I reaction center complex, showed a red band maximum at 672 nm and a small absorption band around 700 nm at liquid nitrogen temperature. CP60 emitted a fluorescence band at 717 to 725 nm at 77 degrees K. The temperature dependence of the far red band of CP60 was essentially the same as that of CP1-e between 77 and 273 degrees K. No photoresponse of P700 was detected in CP60. The results suggest that the two polypeptides resolved by SDS-gel electrophoresis from CP1-e are apoproteins of two distinct chlorophyll-proteins and that CP60 represents a chlorophyll-bearing 60-kDa subunit functioning as an intrinsic antenna protein of the photosystem I reaction center complex. It will also be shown that the temperature dependence of the far red fluorescence band is not related to the photosystem I photochemistry.  相似文献   

4.
Two lab-scale plug flow activated sludge reactors were run in parallel for 4 months at 30 and 55 degrees C. Research focussed on: (1) COD (chemical oxygen demand) removal, (2) effluent turbidity at both temperatures, (3) the origin of effluent colloidal material and (4) the possible role of protozoa on turbidity levels. Total COD removal percentages over the whole experimental period were 66+/-7% at 30 degrees C and 53+/-11% at 55 degrees C. Differences in total COD removal between both systems were due to less removal of soluble and colloidal COD at 55 degrees C compared to the reference system. Thermophilic effluent turbidity was caused by a combination of influent colloidal particles that were not effectively retained in the sludge flocs, and erosion of the thermophilic activated sludge itself, as shown by denaturing gradient gel electrophoresis (DGGE) profiles. DGGE analysis of PCR-amplified 16S rDNA fragments from mesophilic and thermophilic sludge differed, indicating that different microbial communities were present in the two reactor systems. The effects of protozoal grazing on the effluent turbidity of both reactors was negligible and thus could not account for the large turbidity differences observed.  相似文献   

5.
The effects of temperature acclimation at 10, 20 and 30 degrees C on the concentration and activity of the mixed function oxidase system in bluegill are as follows. Liver weight/body weight varied inversely with temperature. Significant (P less than 0.05) differences in concentration of cytochrome P-450 of hepatic microsomes were seen and varied inversely with temperature. Benzo(a)pyrene hydroxylase activity tested in vitro at incubation temperatures of 10, 20 and 30 degrees C showed significant differences in Km, but no differences in Vmax. SDS polyacrylamide gel electrophoresis revealed some quantitative differences in cytochrome P-450 isozymes between groups.  相似文献   

6.
研究了辣根过氧化物酶在三种表面活性剂(SDS,TritonX-100及CTAB)的水相胶束中催化联苯胺聚合反应的动力学。结果表明水相胶束体系有利于反应的进行。辣根过氧化物酶在水相胶束体系中遵循米氏反应,K_m在SDS、TritonX-100及CTAB三种体系中分别为3.014×10~(-4)mol/L、1.728×10~(-4)mol/L和5.664×10~(-5)mol/L。由于微环境的不同,HRP在三种体系中表现出不同的最适反应温度和最适pH。  相似文献   

7.
Sepia cartilage collagen (pepsin-extracted) in acetate buffer (pH = 2.98) forms micelles at a particular concentration below which they do not normally form. The critical micelle concentration (cmc) of the collagen was determined in buffer as well as in SDS, cetyltrimethylammonium bromide (CTAB) and Tween-80 micellar environments at different temperatures. Mutual interaction of collagen micelles with the ionic and nonionic micelles through the formation of the mixed micelle concept has also been found. The cmc of collagen decreased in the presence of SDS and Tween-80 micelles whereas it increased in the presence of CTAB micelles. This clearly suggests that the micelle formation of collagen is facilitated by the presence of SDS and Tween-80 and hindered by CTAB micelles. The various thermodynamic parameters were estimated from viscosity measurements and the transfer of collagen into the micelles of various surfactants and the reverse phenomenon was analyzed. This analysis has also been modelled conceptually as a different phase and the results have supported the above phenomenon. Our thermodynamic results are also able to predict the exact denaturation temperature as well as the structural order of water in the collagen in various environments. The hydrated volumes, Vh, of collagen in the above environments and intrinsic viscosity were also calculated. The low intrinsic viscosity, [eta], of collagen in an SDS environment compared to buffer and other surfactant environments suggested more workable systems in cosmetic and dermatological skin care preparations. The one and two-hydrogen-bonded models of this collagen in various environments have been analyzed. The calculated thermodynamic parameters varied with the concentration of collagen. The change of thermodynamic parameters from coil-coil to random-coil conformation upon denaturation of collagen were calculated from the amount of proline and hydroxyproline residues and compared with viscometric results. Thermodynamic results suggest that the stability of the collagen in the additive environments is in the following order: SDS greater than Tween-80 greater than buffer greater than CTAB.  相似文献   

8.
Turbidity, structure, and rheological features during gelation via the Ugi multicomponent condensation reaction of semidilute solutions of alginate have been investigated at different polymer and cross-linker concentrations and reaction temperatures. The gelation time of the system decreased with increasing polymer and cross-linker concentrations, and a temperature rise resulted in a faster gelation. At the gel point, a power law frequency dependence of the dynamic storage modulus (G' proportional, variant omega(n)(')) and loss modulus (G' ' proportional, variant omega(n)(' ')) was observed for all gelling systems with n' = n' ' = n. By varying the cross-linker density at a fixed polymer concentration (2.2 wt %), the power law exponent is consistent with that predicted (0.7) from the percolation model. The value of n decreases with increasing polymer concentration, whereas higher temperatures give rise to higher values of n. The elastic properties of the gels continue to grow over a long time in the postgel region, and at later stages in the gelation process, a solidlike response is observed. The turbidity of the gelling system increases as the gel evolves, and this effect is more pronounced at higher cross-linker concentration. The small-angle neutron scattering results reveal large-scale inhomogeneities of the gels, and this effect is enhanced as the cross-linker density increases. The structural, turbidity, and rheological features were found to change over an extended time after the formation of the incipient gel. It was demonstrated that temperature, polymer, and cross-linker concentrations could be utilized to tune the physical properties of the Ugi gels such as structure, transparency, and viscoelasticity.  相似文献   

9.
The effects of temperature on the assembly of collagen fibrils were examined in a system in which collagen monomers are generated de novo and in a physiological buffer by specific enzymic cleavage of type I pC-collagen, an intermediate in the normal processing of type I procollagen to type I collagen. Increasing the temperature of the reaction in the range of 29-35 degrees C decreased the turbidity lag and increased the rate of propagation as assayed by turbidity. The effect of temperature on the turbidity propagation rate gave a linear Arrhenius plot with a negative slope. The predicted value of the activation energy of propagation was 113 kJ/mol. However, the effects of temperature on the rate of assembly above 37 degrees C were opposite to the effects seen at temperatures below 37 degrees C. In the range of 37-41 degrees C, the turbidity propagation rate decreased markedly with temperature. Also, the turbidity lag increased. Therefore, much longer times were required for monomers to reach equilibrium with fibrils. A large fraction of the collagen monomers remaining in solution at temperatures above 37 degrees C was sensitive to rapid digestion by trypsin and alpha-chymotrypsin. Cooling the solutions to 25 degrees C made the monomers resistant to protease digestion. The results are consistent with the conclusion that, although formation of collagen fibrils is a classical example of an entropy-driven process of self-assembly, the rate of assembly between 37 and 41 degrees C is limited by reversible micro-unfolding of the monomer.  相似文献   

10.
Effects of low SDS concentrations on amorphous aggregation of tobacco mosaic virus (TMV) coat protein (CP) at 52 degrees C and on the protein structure were studied. It was found that SDS completely inhibits the TMV CP (11.5 microM) unordered aggregation at the detergent/CP molar ratio of 15 : 1 (0.005% SDS). As judged by fluorescence spectroscopy, these SDS concentrations did not prevent heating-induced disordering of the large-distance part of the TMV CP subunit, including the so-called "hydrophobic girdle". At somewhat higher SDS/protein ratio (40 : 1) the detergent completely disrupted the TMV CP hydrophobic girdle structure even at room temperature. At the same time, these low SDS concentrations (15 : 1, 40 : 1) strongly stabilized the structure of the small-distance part of the TMV CP molecule (the four alpha-helix bundle) against thermal disordering as judged by the far-UV (200-250 nm) CD spectra. Possible mechanisms of TMV CP heating-induced unordered aggregation initiation are discussed.  相似文献   

11.
Phospholipid exchange between bilayer membrane vesicles.   总被引:7,自引:0,他引:7  
The turbidity of lipid vesicles, freshly prepared by sonicating purified dimyristoyllecithin (DML) in dilute KCl solutions, was measured as a function of time at various temperatures. A sharp maximum in the rate of increase of turbidity is found just above the crystal:liquid-crystal phase transition temperature (Tm). The initial rate of turbidity increase is first order with respect to DML concentration. Electron and light microscopy reveal large vesicles which are not present before incubation or after incubation at temperatures far from the Tm. When temperature, rather than time, is the independent variable, a sharp drop in turbidity is seen at the Tm. The magnitude of this drop and the temperature at which it occurs were used to measure the rate of lipid transfer between vesicles composed of different lipids. A mixture of DML vesicles and dipalmitoyllecithin (DPL) vesicles exhibits sharp drops in turbidity at 24 and 41 degrees, the corresponding Tm's. With time, the magnitude of the transition at 24 degrees decreases while that which was originally at 41 degrees moves to lower temperatures and increases in magnitude. At equilibrium there is a single transition at 32.5 degrees characteristic of vesicles composed of equimolar DPL and DML. The rate at which equilibrium is approached increases at around 24 degrees and again around 41 degrees. These observations indicate that vesicles are in equilibrium with monomolecular lipid, the concentration of the latter being higher the shorter the lipid acyl group or the smaller the vesicle. DML molecules are therefore lost from small vesicles to large vesicles (DML system) or lost from DML vesicles to DML-DPL vesicles (mixed system). When DML vesicles containing a few percent brain gangliosides were studied, different behavior was observed; the initial rate of increase of turbidity becomes second order in lipid concentration, and the rate constant increases with increasing concentrations of KCl. The kinetic order, coupled with the fact that electrolyte reduces intervesicle electrostatic repulsion, argues that in this situation the mechanism of vesicle growth requires vesicle collision.  相似文献   

12.
Extent of binding (gamma 2(1)) of sodium dodecyl sulphate (SDS) to the binary complex formed between calfthymus DNA and cetyltrimethylammonium bromide (CTAB) has been measured in mole per mole of nucleotide in the complex as function of concentration of SDS by using equilibrium dialysis technique at different temperatures and pH. Binding of SDS to thermally denatured DNA-CTAB complex has also been studied. The most interesting aspect to be noted in this experiment is that the water insoluble DNA-CTAB binary complex gets solubilized in the ternary mixture in presence of SDS but when DNA is thermally denatured, the ternary system DNA-CTAB-SDS remains insoluble. Significant change in the extent of binding has been noted with the variation of the relative composition of DNA and CTAB in their binary mixture. The data of binding of SDS to DNA-CTAB complex are compared more precisely in terms of the standard Gibbs' free energy decrease (-delta G degree) for the saturation of the binding sites in the complex with the change of SDS activity from zero to unity in the rational mole fraction scale.  相似文献   

13.
The precipitating of effect of sodium dodecyl sulphate (SDS) on the egg white proteins ovalbumin, conalbumin and lysozyme was studied at 25 degrees C and at different pH values. The proteins precipitated below their respective isolectric points, provided the detergent to protein ratio was appropriate. The pH profile of precipitation was different for the three proteins reflecting net charge differences. The binding of SDS to the proteins was studied with [35S]-labelled SDS and for ovalbumin a ratio of 21--28 SDS molecules/protein molecule, dependent on the concentration of SDS initially used, seem to be required for precipitation at pH 4.5. This number, however, is dependent on pH and increases with an increased positive net charge of the protein. The precipitating effect of SDS was identical for ovalbumin, conalbumin and lysozyme when compared on a gram to gram basis (0.1--0.15 g SDS/g precipitated protein). The precipitated protein was denatured as measured by differential scanning calorimetry, but was also completely redissolved if pH was increased to above the isoelectric point. The precipitating effecto f SDS was also examined at elevated temperatures. The two-phase systems of the proteins induced by SDS at 25 degrees C were heated from 25 degrees C to 90 degrees C at a rate of 1.25 degrees C/min. The precipitation behaviour was similar for the three proteins upon heating. When the SDS concentration was increased the precipitation curves were transferred towards lower temperatures and the courses of precipitation became less sharp. The synergistic effect of SDS and heat on protein precipitation was differentiated by denaturation measurements and radioactive labelling. The ratio SDS to precipitated protein gradually diminished towards higher temperatures but no purely thermal precipitation was found.  相似文献   

14.
The temperature dependence of octyl glucoside micellization was determined and compared to the phase behavior of the octyl glucoside--egg phosphatidylcholine (PC) mixed system in excess water to help elucidate the process of vesicle formation from mixed surfactant-phospholipid micelles. The critical micelle concentration of octyl glucoside (OG) was determined from the sharp increase of ANS fluorescence at micellization in an NaCl buffer at temperatures ranging from 5 to 40 degrees C. The cmc decreased with increasing temperature from 31 mM at 5 degrees C to 16 mM at 40 degrees C. A similar but less steep temperature dependence is observed for the solubilization of egg PC vesicles by OG as monitored by the surfactant-dependent changes in (1) solution turbidity and (2) the resonance energy transfer between NBD-PE and Rho-PE incorporated in the vesicles. These assays identify two breakpoints, most likely the boundaries of the cylindrical micelle and spheroidal micelle coexistence region. The [OG]aq values at these two breakpoints have similar temperature dependencies. However, the cylindrical mixed micelles at the boundary have nearly identical OG:PC ratios over the temperature range studied, whereas the spheroidal mixed micelles have relatively more OG at the higher temperatures (OG:PC ratio increases from 2.92 to 3.72 between 5 and 35 degrees C). Estimation of the acyl volume to surface area ratio for the compositions observed suggests that this parameter remains constant over temperature. The spheroidal mixed micelles, but not the cylindrical PC-OG micelles, exhibit ideal mixing between the two components at all temperatures (5-35 degrees C). This temperature sensitivity may be utilized to improve the efficacy of membrane protein reconstitution.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

15.
Calorimetric titrations have been performed on the binding of ethidium and propidium to calf thymus DNA at temperatures in the 15-60 degrees C range. Enthalpy changes (delta HB) derived from these experiments performed with the new Omega reaction calorimeter have a precision of +/- 0.10 kcal/mol or less at all temperatures. For ethidium (a monocation), delta HB varies little with temperature, and the heat capacity change (delta CP) for the binding reaction derived from these parameters is 10 cal/deg/mol. In contrast, delta HB changes from -6.5 to -8.1 kcal/mol for DNA binding of propidium (a dication due to a charged amine group at the end of an alkyl chain attached to the phenanthridine ring nitrogen), and delta CP is -57 cal/deg/mol. At 21 degrees C a plot of delta HB vs mole ratio is curved downward for propidium in the 0.08-0.25 range, whereas the same plot at 45 degrees C is a straight line from 0.05 to 0.15 and sharply downward thereafter. Similar plots for ethidium follow the latter pattern between 25 and 50 degrees C. These observations and our analyses of delta HB and delta SB are consistent with the hypothesis that the location in the DNA complex and the rotational motion of the alkylamine chain change substantially over the temperature range in this study. Only near 50 degrees C is delta HB equal for the binding of these two cations to DNA, and caution must be used in analyses of enthalpic effects when the temperature dependence for delta HB is not available.  相似文献   

16.
Six male New Zealand white rabbits were individually exposed to 600 MHz radiofrequency (RF) radiation for 90 min in a waveguide exposure system at an ambient temperature (Ta) of 20 or 30 degrees C. Immediately after exposure, the rabbit was removed from the exposure chamber and its colonic and ear skin temperatures were quickly measured. The whole-body specific absorption rate (SAR) required to increase colonic and ear skin temperature was determined. At a Ta of 20 degrees C the threshold SAR for elevating colonic and ear skin temperature was 0.64 and 0.26 W/kg, respectively. At a Ta of 30 degrees C the threshold SARs were slightly less than at 20 degrees C, with values of 0.26 W/kg for elevating colonic temperature and 0.19 W/kg for elevating ear skin temperature. The relationship between heat load and elevation in deep body temperature shown in this study at 600 MHz is similar to past studies which employed much higher frequencies of RF radiation (2450-2884 MHz). On the other hand, comparison of these data with studies on exercise-induced heat production and thermoregulation in the rabbit suggest that the relationship between heat gain and elevation in body temperature in exercise and from exposure to RF radiation may differ considerably. When combined with other studies, it was shown that the logarithm of the SAR required for a 1.0 degree C elevation in deep body temperature of the rabbit, rat, hamster, and mouse was inversely related to the logarithm of body mass. The results of this study are consistent with the conclusion that body mass strongly influences thermoregulatory sensitivity of the aforementioned laboratory mammals during exposure to RF radiation.  相似文献   

17.
M D King  D Marsh 《Biochemistry》1989,28(13):5643-5647
The polymorphic phase behavior of 1-palmitoyl-2-lyso-sn-glycero-3-phosphocholine dispersions in excess water has been studied as a function of temperature and poly(ethylene glycol) (PEG) concentration, using proton dipolar-decoupled 31P NMR spectroscopy and turbidity measurements. The phase behavior was found to depend on both lipid concentration and PEG concentration, and most of the NMR experiments were conducted at a lipid concentration of 15 mg/mL. At low PEG concentrations (0-12 wt %), a thermotropic transition occurs at 3-5 degrees C with increasing temperature, from an interdigitated lamellar gel (L beta i) phase to a normal micellar phase. At intermediate PEG concentrations (12-20 wt %), thermotropic transitions take place with increasing temperature, first from the lamellar gel phase to a fluid cubic (Q alpha) phase and then at higher temperatures from the cubic phase to the micellar phase. At intermediate PEG concentrations above the former range (20-30 wt %), thermotropic transitions take place with increasing temperature, first from the lamellar gel phase to the cubic phase, then from the cubic phase to a normal hexagonal (HI) phase, and finally from the hexagonal phase to the micellar phase. At high PEG concentrations (greater than 30 wt %), a thermotropic transition takes place with increasing temperature from the lamellar gel phase directly to the fluid hexagonal phase. At these high PEG concentrations, the micellar phase is not attained within the accessible temperature range (less than or equal to 90 degrees C).(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

18.
Due to their broad applications, ionic surfactants have already been released into or utilized in soil and environmental systems. However, current understanding on the sorption behavior of surfactants onto soils is still limited. This work systematically investigated the sorption kinetics and isotherms of one cationic surfactant, cetyltrimethylammonium bromide (CTAB), and one anionic surfactant, sodium dodecyl sulfate (SDS), onto a silt loamy soil to determine the governing sorption mechanisms. The pseudo-second-order rate equation described the sorption kinetics data better than the pseudo-first-order rate equation. Experimental data showed that the sorption equilibrium for CTAB and SDS were reached at 24 and 240 h, respectively. Langmuir equation was better than Freundlich equation in simulating the sorption isotherms of CTAB and SDS on the soil. Soil Langmuir maximum sorption capacity of CTAB was much higher than that to SDS. When the experimental temperature increased, the sorption of CTAB and SDS on the soil decreased. In addition, the sorptive process of the surfactants on the soil was spontaneous and exothermal, as indicated by the absolute values of Gibbs free energy and enthalpy. The results also indicated that physical sorption was the dominant mechanism for the sorption of the two surfactants on the soil. Findings from this work are crucial to understand the environmental behaviors of ionic surfactants.  相似文献   

19.
药用植物艾纳香基因组DNA提取方法研究   总被引:3,自引:0,他引:3  
以药用植物艾纳香为研究对象,以-20℃保存、4℃保存、室温自然干燥和硅胶干燥四种样品保存方式,并采用SDS法、CTAB法、SDS-CTAB法和改良CTAB法4种不同的基因组DNA提取方法进行了对比试验,以期建立艾纳香的较好的样品保存方法和基因组DNA提取方法。结果表明,-20℃保存是艾纳香的较理想的样品保存方式;改良CTAB法是艾纳香基因组DNA提取较适宜的方法,该方法提取的DNA经紫外检测,其A_(260)/A_(280)为1.8左右,明显优于SDS法(1.1~1.5)、CTAB法(1.2~1.5)和SDS-CTAB法(1.4~1.6),琼脂糖凝胶电泳、酶切检测和PCR扩增也得出了同样的结论。  相似文献   

20.
Native collagen polypeptides exist in a unique triple helical conformation resistant to most proteinases. In this study, the stability of type I collagen triple helix, employing a mixture of trypsin and alpha-chymotrypsin as a proteolytic probe, was examined. The degradation of type I [3H]collagen was monitored as 3H-labeled peptides soluble in trichloroacetic acid (TCA) or by sodium dodecyl sulfate (SDS)-polyacrylamide slab gel electrophoresis. In one set of experiments, collagen substrates were preincubated at various temperatures for up to 8 h, followed by a 15-min proteolytic treatment at the same temperature. At 43 degrees C, most of the collagen was degraded, while the fraction of the substrate degraded at 40, 38, and 35 degrees C was 53, 41 and 19%, respectively. This fraction was independent of the preincubation time which varied from 10 to 480 min. Thus, at any given temperature, a constant fraction of the collagen substrate was susceptible to proteolysis. Measurement of the midpoint temperature (Tm) of the helix to coil transformation for type I collagen, at neutral pH employing an increasing temperature gradient and brief proteolysis at the individual temperatures, indicated a value of 38.8 degrees C. However, determination of the Tm by employing proteolytic digestions at a constant temperature (30 degrees C) using conditions under which the nonhelical peptides are readily digested to TCA-soluble peptides while native collagen resists such proteolysis, indicated a value of 42.7 degrees C. In further studies, collagen was subjected to continuous proteolysis for up to 24 h. A large fraction of collagen was digested at 30 or 34 degrees C, temperatures well below the Tm of the helix to coil transformation. SDS-polyacrylamide gel electrophoresis of the degradation products obtained at these temperatures revealed multiple cleavage fragments. Finally, temperature double-jump experiments indicated that the destabilization of the triple helix is reversible provided that the Tm of the substrate is not exceeded. The results provide evidence for reversible and local relaxation of the collagen triple helix.  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