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 31 毫秒
1.
A variety of novel opaque, white polymers ranging from rubbery materials to tough and rigid plastics have been prepared by the thermal polymerization at 85-160 degrees C of varying amounts of 87% conjugated linseed oil, styrene, and divinylbenzene. Gelation of the reactants typically occurs at temperatures higher than 120 degrees C, and fully cured thermosets are obtained after postcuring at 160 degrees C. The fully cured thermosets have been determined by Soxhlet extraction to contain approximately 35-85% cross-linked materials. The microcomposition of these polymers, as determined by 1H NMR spectroscopy, indicates that the cross-linked materials are composed of both soft oily and hard aromatic phases. After solvent extraction, the insoluble materials exhibit nanopores well distributed throughout the polymer matrixes. Dynamic mechanical analysis of these polymers indicates that they are phase separated with a soft rubbery phase having a sharp glass transition temperature of -50 degrees C and a hard brittle plastic phase with a broadened glass transition temperature of 70-120 degrees C. These polymers possess cross-link densities of 0.15-2.41 x 10(4) mol/m3, compressive Young's moduli of 12-438 MPa, and compressive strengths of 2-27 MPa. These materials are thermally stable below 350 degrees C and exhibit a major thermal degradation of 72-90% at 493-500 degrees C.  相似文献   

2.
Novel thermosetting copolymers, ranging from tough and ductile to very soft rubbers, have been prepared by the cationic copolymerization of regular (SOY) and 100% conjugated soybean oils (C(100)SOY) with dicyclopentadiene (DCP) catalyzed by Norway fish oil (NFO)-modified and SOY- and C(100)SOY-diluted boron trifluoride diethyl etherate (BFE). The gelation time of the reactions varies from 4 to 991 min at 110 degrees C. The yields of the bulk copolymers are essentially quantitative, while the yields of the cross-linked copolymers remaining after Soxhlet extraction with methylene chloride range from 69% to 88%, depending on the monomer stoichiometry and the catalyst used. (1)H NMR spectroscopy and Soxhlet extraction data indicate that these copolymers consist of a cross-linked soybean oil-DCP network plasticized by certain amounts of methylene chloride-soluble linear or less cross-linked soybean oil-DCP copolymers, unreacted oil, and some low molecular weight hydrolyzed oil. The molecular weights of these soluble fractions are in the range from 400 to 10,000 g/mol based on polystyrene standards. The bulk copolymers have glass transition temperatures ranging from -22.6 to 56.6 degrees C, while their tan delta peak values range from 0.7 to 1.2. Thermogravimetric analysis (TGA) indicates that these soybean oil-DCP copolymers are thermally stable below 200 degrees C, with 10% and 50% weight loss temperatures ranging from 280 to 372 degrees C and 470-554 degrees C, respectively. These properties suggest that these biobased thermosets may prove useful alternatives to current petroleum-based plastics and find widespread utility.  相似文献   

3.
Ceric ammonium nitrate (CAN) in the presence of nitric acid has been used as efficient initiator for graft copolymerization of the ethyl acrylate onto cellulose at 35.0 +/- 0.1 degrees C. Graft copolymerization of ethyl acrylate onto cellulose has taken place through the radical initiation process. The graft yield and other grafting parameters have been evaluated by varying concentration of ethyl acrylate from 2.5 x 10(-1) to 15.0 x 10(-1) mol dm(-3) and ceric ammonium nitrate from 5.0 x 10(-3) to 25.0 x 10(-3) mol dm(-3) at constant concentration of the nitric acid (8.0 x 10(-2) mol dm(-3)). The rate of graft copolymerization has shown 1.5 order with respect to the concentration of the ceric ammonium nitrate. The graft copolymerization data obtained at different temperatures were used to calculate the energy of activation, which has been found to be 28.9 kJ mol(-1) within the temperature range from 20 to 50 degrees C. The effect of addition of cationic and anionic surfactants on graft copolymerization has also been studied. On the basis of the experimental observations, reaction steps have been proposed and a suitable rate expression for graft copolymerization has been derived.  相似文献   

4.
Temperature-responsive cellulose has been obtained by graft copolymerization of N-isopropylacrylamide (NIPAAm) monomer using ceric ammonium nitrate (CAN) as initiator at 25.0 +/- 0.1 degrees C in acidic medium. Kinetic and grafting parameters were evaluated at different concentrations of NIPAAm ranging from 1.25 x 10(-3) to 12.5 x 10(-3) mol dm(-3) and varying concentrations of CAN from 1.5 x 10(-3) to 9.0 x 10(-3) mol dm(-3) at constant concentration of nitric acid (2.5 x 10(-2) mol dm(-3)). The graft copolymerization of NIPAAm onto cellulose has shown a significant increasing trend below lower critical solution temperature (LCST) of poly(N-isopropylacrylamide) (PNIPAAm) and shown low energy of activation (18.0 kJ mol(-1)) for graft copolymerization within the temperature range of 10-35 degrees C as determined with Arrhenius plot. The PNIPAAm-grafted cellulose has shown improved thermal stability and shown temperature-dependent degree of swelling. Variation in degree of swelling of PNIPAAm-grafted cellulose as a function of temperature has been used to determine LCST of PNIPAAm-grafted cellulose. The contact angle (theta) has shown variation on increasing the graft yield and temperature. On the basis of experimental observations, the reaction steps for graft copolymerization have been proposed and a rate expression has been derived.  相似文献   

5.
Noronha AM  Wilds CJ  Miller PS 《Biochemistry》2002,41(27):8605-8612
Short DNA duplexes containing a 1,3-N(4)C-alkyl-N(4)C interstrand cross-link that joins the two C residues of a -CNG- sequence were prepared using either a phosphoramidite or convertible nucleoside approach. The alkyl cross-link consists of 2, 4, or 7 methylene groups. The duplexes, which contain a seven-base-pair core and A(3)/T(3) complementary 3'-overhanging ends, were characterized by enzymatic digestion and MALDI-TOF mass spectrometry. Ultraviolet thermal denaturation studies showed that the duplexes denature in a cooperative manner and that the length of the cross-link affects the thermal stability. Thus, the transition temperature of the ethyl cross-linked duplex, 42 degrees C, is 16 degrees C higher than the melting temperature of the corresponding non-cross-linked control, whereas the transition temperatures of the butyl and heptyl cross-linked duplexes, 73 and 72 degrees C, respectively, are 46-47 degrees C higher. The reduced molecularity of denaturation of the cross-linked duplexes versus melting of the non-cross-linked duplex most likely accounts for these differences. Examination of molecular models suggests that the ethyl cross-link is too short to span the distance between the two C residues at the site of the cross-link in B-form DNA without causing distortion of the helix, whereas less and no distortion would be expected for the butyl and heptyl cross-links, respectively. The circular dichroism spectra, which show greatest deviation in the ethyl cross-linked duplex from B-form DNA, are consistent with this expectation. Anomalous mobilities on native polyacrylamide gels of multimers produced by self-ligation of each of the cross-linked duplexes suggest that the ethyl and butyl cross-linked duplexes undergo bending deformations, whereas multimers derived from the heptyl cross-linked duplex migrated normally. The bending angle was estimated to be 20 degrees, 13 degrees, and 0 degrees for the ethyl, butyl, and heptyl cross-linked duplexes, respectively. Thus, it appears that the degree of bending in these N(4)C-alkyl-N(4)C cross-linked duplexes is controlled by the length of the cross-link.  相似文献   

6.
The characteristics of horseradish peroxidase (HRP) immobilized onto Sepharose by a photochemical-initiated graft copolymerization are presented. Active copolymers were synthesized using different amounts of glycidylmethacrylate (GMA), bisacryloylpiperazine (BAP), or 1,3,5-hexhydrotriacryloyl-(s)-triazine (HTsT) as functional monomer. The activities, the K'(m) values (pGMA) copolymers: 0.53-0.76 x 10(-4)M; pBAP copolymers: 0.90-1.4 x 10(-4); pHTsT copolymers: 1.8-2.6 x 10(-4)M and the thermal stabilities of the enzyme copolymers were strictly connected to the type of polymer. By varying the polymer amount present in a given copolymer, significant differences were found in the thermostability properties of pBAP and pHTsT copolymers both when checked in water or in phosphate buffer. No differences were found for pGMA copolymers. The samples in which there are the lowest pBAP or pHTsT content resulted the most stable. The activity retained after 240 min at 60 degrees C by free HRP and pGMA-HRP was 30% whereas by pBAP-HRP and pHTsT-HRP it was 50 and 75% of the original. Operational stability of the materials was in agreement with thermostability data. These results are discussed in terms of enzyme microenvironment which is strongly affected by the different network of the three polymers.  相似文献   

7.
Characterization of poly-3-hydroxybutyric acid (PHB) and poly-3-hydroxybutyric-co-valeric acid (PHBV, 13% valerate) in chloroform was performed using size exclusion chromatography coupled to a multi-angle light scattering detector (SEC-MALS). Absolute molar mass averages, molar mass distribution, and the radius of gyration were determined. Three sample preparation methods were examined: dissolution in chloroform (1) at room temperature, (2) at 60 degrees C, and (3) after thermal pretreatment of samples (annealing at 180 degrees C with subsequent quenching in liquid nitrogen). Dissolution at 60 degrees C and dissolution of thermally pretreated samples gave molecularly dissolved PHB and PHBV. At 60 degrees C using acid free chloroform, there was no indication of degradation for up to 120 min dissolution time, whereas thermal degradation of polymers did take place during annealing at 180 degrees C. The degradation rate constants for number and weight average degree of polymerization at 180 degrees C were slightly higher for PHB (5.19 x 10(-5) min(-1), 4.95 x 10(-5) min(-1)) than for PHBV (4.99 x 10(-5) min(-1), 4.54 x 10(-5) min(-1)). The dependence of the radii of gyration on molar mass showed that both polymers form random coils in chloroform. The relationship between the absolute molar masses and relative SEC results was determined. DSC and NMR characterization also gave evidence of the progress of degradation.  相似文献   

8.
To study the thermal response of interscapular brown fat (IBF) to norepinephrine (NE), urethan-anesthetized rats (1.2 g/kg ip) maintained at 28-30 degrees C received a constant venous infusion of NE (0-2 x 10(4) pmol/min) over a period of 60 min. IBF temperatures (T(IBF)) were recorded with a small thermistor fixed under the IBF pad. Data were plotted against time and expressed as maximal variation (Deltat degrees C). Saline-injected rats showed a decrease in T(IBF) of approximately 0.6 degrees C. NE infusion increased T(IBF) by a maximum of approximately 3.0 degrees C at a dose of 10(4) pmol x min(-1) x 100 g body wt(-1). Surgically thyroidectomized (Tx) rats kept on 0.05% methimazole showed a flat response to NE. Treatment with thyroxine (T(4), 0.8 microg x 100 g(-1) x day(-1)) for 2-15 days normalized mitochondrial UCP1 (Western blotting) and IBF thermal response to NE, whereas iopanoic acid (5 mg x 100 g body wt(-1) x day(-1)) blocked the effects of T(4). Treatment with 3,5, 3'-triiodothyronine (T(3), 0.6 microg x 100 g body wt(-1) x day(-1)) for up to 15 days did not normalize UCP1 levels. However, these animals showed a normal IBF thermal response to NE. Cold exposure for 5 days or feeding a cafeteria diet for 20 days increased UCP1 levels by approximately 3.5-fold. Nevertheless, the IBF thermal response was only greater than that of controls when maximal doses of NE (2 x 10(4) pmol/min and higher) were used. Conclusions: 1) hypothyroidism is associated with a blunted IBF thermal response to NE; 2) two- to fourfold changes in mitochondrial UCP1 concentration are not necessarily translated into heat production during NE infusion.  相似文献   

9.
1. The camel has insulin receptors that by multiple function criteria are very similar to those of the other mammals (rabbit and rat) and non-mammals (chicken and pigeon), with sharp pH dependence to insulin binding at pH 7.2-7.6. 2. Equilibrium binding was faster at higher temperatures (24-37 degrees C) than at lower (4 degrees C). 3. Binding data yielded curvilinear Scatchard plots with half maximal displacement of 125I-insulin at 9 x 10(-9) M, 2.5 x 10(-9) M, 6.3 x 10(-10) M for camel, rabbit, pigeon and chicken respectively, suggesting differences in mammalian and non-mammalian liver membranes. 4. Autoradiogram patterns showed the presence of an identical subunit structure with Mr 74,000 for all membranes studied. Pigeon membrane showed a band with Mr 110,000, the absence of which in other membranes could be due to the degradation factor or the concentration of disuccinimidyl suberate (DSS).  相似文献   

10.
H Yu  N Soong    W F Anderson 《Journal of virology》1995,69(10):6557-6562
A quantitative analysis of the binding kinetics of intact Moloney murine leukemia retrovirus (MoMuLV) particles with NIH 3T3 cells was performed with an immunofluorescence flow cytometry assay. The virus-cell binding equilibrium dissociation constant (KD), expressed in terms of virus particle concentration, was measured to be 8.5 (+/- 6.4) x 10(-12) M at 4 degrees C and was three- to sixfold lower at temperatures above 15 degrees C. The KD of virus binding is about 1,000-fold lower than the KD of purified MoMuLV envelope. The association rate constant was determined to be 2.5 (+/- 0.9) x 10(9) M-1 min-1 at 4 degrees C and was 5- to 10-fold higher at temperatures above 15 degrees C. The apparent dissociation rate constant at 4 degrees C was 1.1 (+/- 0.4) x 10(-3) min-1 and was doubled for every 10 degrees C increase in temperature over the range tested (15 to 37 degrees C).  相似文献   

11.
Thermal degradation, reactivity and kinetics for biomass materials cotton stalk (CS), sugarcane bagasse 1 (SB1), sugarcane bagasse 2 (SB2) and shea meal (SM) have been evaluated under pyrolysis (N(2)) and oxidising (dry air) conditions, using a non-isothermal thermogravimetric method (TGA). In the cases of CS and SB1 the peak temperatures were 51 degrees C higher for pyrolysis compared with oxidative degradation, whereas for SB2 and SM the difference was approximately 38 degrees C. However, the differences in the rates of weight loss were significantly higher under oxidising conditions for all the materials studied. Maximum rate of weight loss (%s(-1)) under pyrolysis conditions ranged from 0.10 to 0.18 whereas these values accelerated to the range of 0.19-0.28 under oxidising conditions, corresponding to respective peak temperatures. Samples ranked in order of reactivity (R(M)x10(3)) (%s(-1) degrees C(-1)) are CS=1.31 approximately SM=1.30>SB2=1.14>SB1=0.94 for air and CS=0.54>SB2=0.49>SB1=0.45>SM=0.31 for nitrogen. Shea meal exhibited a complex char combustion behaviour indicating that there may be two distinct types of char derived from fibrous and woody components in the original material. Activation energy calculations were based on the Arrhenius correlation.  相似文献   

12.
Sugar-substituted epoxides 5-8 were synthesized by asymmetric epoxidation (in CH(2)Cl(2)/water) of alpha-olefins having neighboring sugars (1-4) by use of an achiral oxidant (MCPBA), in which the sugar moiety acted as a chiral template. The diastereoselectivities depend on the methylene spacer between vinyl group and carbohydrate derivatives. The methylene spacer between sugar and vinyl groups influenced the diastereoselectvity. In the case of epoxidation of 4 at 27 degrees C for 24 h, the diastereoselectivity was the highest (99/1). Copolymerizations of 5-8 with succinic anhydride were attained at 100 degrees C for 72 h to give poly(ethylene succinate) having pendant carbohydrate [poly(SAn-alt-5), M(n) = 1.4 x 10(3); poly(SAn-alt-6), M(n) = 2.2 x10(3); poly(SAn-alt-7), M(n) = 2.9 x 10(3); poly(SAn-alt-8), M(n) = 1.8 x 10(3)]. The methylene spacer between sugar and epoxide has an effect on the reactivity of epoxide in copolymerization as well as the diastereoselectivity. Alternating copolymerization of 7 and glutaric anhydride gave a polyester of M(n) 4.2 x10(3).  相似文献   

13.
The effect of chlorpromazine on the development of cold shock in erythrocytes exposed to sodium chloride was shown to depend on the tonicity of the medium in which the cells were cooled from 37 degrees C down to 0 degrees C as well as on the amphipate concentration. After cooling of erythrocytes in a NaCl (0.75-1.5 M)-containing medium with chlorpromazine (7 x 10(-5) M, 2.1 x 10(-4) M and 3.5 x 10(-4) M) the hypertonic cold shock was inhibited, the protective effect of the amphipate being less pronounced at its increasing concentrations. After cooling of cells under conditions of moderate hypertonicity (0.3-0.6 M NaCl) no modifying effect of chlorpromazine on the sensitivity of erythrocytes to the temperature decrease from 37 degrees C down to 0 degrees C was manifested. However, under iso- and hypertonic conditions chlorpromazine used at 2.1 x 10(-4) M and 3.5 x 10(-4) M stimulated the cold shock development in erythrocytes. A sharp increase in the medium tonicity (up to 1.8-3.0 M and higher) the cells underwent isothermal hemolysis which was more expressed at 0 degrees C than at 37 degrees C. These data suggest that chlorpromazine significantly activates the hemolytic process at low temperatures.  相似文献   

14.
1. The thermal dependence of contractile behavior at different stimulation frequencies was investigated in isolated frog sartorius muscles. 2. Increasing incubation temperature (10-30 degrees C) produced decreases in Pt (43.7%) and P15 (70.3%), and an increase in Po (26.0%). 3. Thermal ratios (R10) calculated for Pt, P15 and Po indicated high thermal dependence at lower temperatures (10-20 degrees C; 0.60, 0.44 and 1.38, respectively) but relative thermal independence at higher temperatures (20-30 degrees C; 0.95, 0.75 and 0.95, respectively). 4. Contractile ratios (Pt/Po and P15/Po) decreased with increased temperature (10-30 degrees C; 56.3% and 76.0%, respectively). 5. Thermal ratios (R20) calculated for peak tension at different stimulation frequencies demonstrated high thermal dependence at the lower frequencies (10-30 pps, 0.46-0.48) and decreasing dependence at higher frequencies (40-50 pps, 0.69-0.82).  相似文献   

15.
A thrombin receptor in resident rat peritoneal macrophages.   总被引:2,自引:0,他引:2  
Resident rat peritoneal macrophages possess 6 x 10(2) high-affinity binding sites per cell for bovine thrombin with a Kd of 11 pM, and 7.5 x 10(4) low-affinity sites with a Kd of 5.8 nM. These binding sites are highly specific for thrombin. Half-maximal binding of 125I-labeled bovine thrombin is achieved after 1 min at 37 degrees C, and after 12 min at 4 degrees C. The reversibly bound fraction of the ligand dissociates according to a biexponential time course with the rate constants 0.27 and 0.06 min-1 at 4 degrees C. Part of the tracer remains cell-associated even after prolonged incubation, but all cell-associated radio-activity migrates as intact thrombin upon sodium dodecyl sulfate-polyacrylamide gel electrophoresis. The bound thrombin is minimally endocytosed as judged by the resistance to pH 3 treatment, and the receptor does not mediate a quantitatively important degradation of the ligand. The binding is not dependent on the catalytic site of thrombin, since irreversibly inactivated thrombin also binds to the receptor. 125I-labeled thrombin covalently cross-linked to its receptor migrates in sodium dodecyl sulfate-polyacrylamide gel electrophoresis with a Mr 160,000, corresponding to an approximate receptor size of Mr 120,000.  相似文献   

16.
Uniformly degrading biomaterials with adjustable degradation product migration rates were customized by combining the advantages of cross-linked poly(epsilon-caprolactone) with the hydrophilic character of poly(1,5-dioxepan-2-one). Hydrolytic degradation of these random cross-linked networks using 2,2'-bis-(epsilon-caprolactone-4-yl) propane (BCP) as the cross-linking agent was studied for up to 546 days in phosphate buffer solution at pH 7.4 and 37 degrees C. The hydrophilicity of the materials was altered by varying the copolymer compositions. After different hydrolysis times the materials were characterized, and the degradation products were extracted from the buffer solution and analyzed. Fourier transform infrared spectroscopy, differential scanning calorimetry, atomic force microscopy, scanning electron microscopy, and gas chromatography-mass spectrometry were used to observe the changes taking place during the hydrolysis. From the results it was concluded that degradation profiles and migration of degradation products are controllable by tailoring the hydrophilicity of cross-linked polyester-ether networks.  相似文献   

17.
Temperature dependence of the isometric tension was examined in chemically skinned, glycerinated, rabbit Psoas, muscle fibers immersed in relaxing solution (pH approximately 7.1 at 20 degrees C, pCa approximately 8, ionic strength 200 mM); the average rate of heating/cooling was 0.5-1 degree C/s. The resting tension increased reversibly with temperature (5-42 degrees C); the tension increase was slight in warming to approximately 25 degrees C (a linear thermal contraction, -alpha, of approximately 0.1%/degree C) but became more pronounced above approximately 30 degrees C (similar behavior was seen in intact rat muscle fibers). The extra tension rise at the high temperatures was depressed in acidic pH and in the presence of 10 mM inorganic phosphate; it was absent in rigor fibers in which the tension decreased with heating (a linear thermal expansion, alpha, of approximately 4 x 10(-5)/degree C). Below approximately 20 degrees C, the tension response after a approximately 1% length increase (complete < 0.5 ms) consisted of a fast decay (approximately 150.s-1 at 20 degrees C) and a slow decay (approximately 10.s-1) of tension. The rate of fast decay increased with temperature (Q10 approximately 2.4); at 35-40 degrees C, it was approximately 800.s-1, and it was followed by a delayed tension rise (stretch-activation) at 30-40.s-1. The linear rise of passive tension in warming to approximately 25 degrees C may be due to increase of thermal stress in titin (connectin)-myosin composite filament, whereas the extra tension above approximately 30 degrees C may arise from cycling cross-bridges; based on previous findings from regulated actomyosin in solution (Fuchs, 1975), it is suggested that heating reversibly inactivates the troponin-tropomyosin control mechanism and leads to Ca-independent thin filament activation at high temperatures. Additionally, we propose that the heating-induced increase of endo-sarcomeric stress within titin-myosin composite filament makes the cross-bridge mechanism stretch-sensitive at high temperatures.  相似文献   

18.
C(4) plants are rare in the cool climates characteristic of high latitudes and elevations, but the reasons for this are unclear. We tested the hypothesis that CO(2) fixation by Rubisco is the rate-limiting step during C(4) photosynthesis at cool temperatures. We measured photosynthesis and chlorophyll fluorescence from 6 degrees C to 40 degrees C, and in vitro Rubisco and phosphoenolpyruvate carboxylase activity from 0 degrees C to 42 degrees C, in Flaveria bidentis modified by an antisense construct (targeted to the nuclear-encoded small subunit of Rubisco, anti-RbcS) to have 49% and 32% of the wild-type Rubisco content. Photosynthesis was reduced at all temperatures in the anti-Rbcs plants, but the thermal optimum for photosynthesis (35 degrees C) did not differ. The in vitro turnover rate (kcat) of fully carbamylated Rubisco was 3.8 mol mol(-)(1) s(-)(1) at 24 degrees C, regardless of genotype. The in vitro kcat (Rubisco Vcmax per catalytic site) and in vivo kcat (gross photosynthesis per Rubisco catalytic site) were the same below 20 degrees C, but at warmer temperatures, the in vitro capacity of the enzyme exceeded the realized rate of photosynthesis. The quantum requirement of CO(2) assimilation increased below 25 degrees C in all genotypes, suggesting greater leakage of CO(2) from the bundle sheath. The Rubisco flux control coefficient was 0.68 at the thermal optimum and increased to 0.99 at 6 degrees C. Our results thus demonstrate that Rubisco capacity is a principle control over the rate of C(4) photosynthesis at low temperatures. On the basis of these results, we propose that the lack of C(4) success in cool climates reflects a constraint imposed by having less Rubisco than their C(3) competitors.  相似文献   

19.
The kinetics of thermal denaturation of a biliprotein, C-phycocyanin (C-PC) isolated from Spirulina platensis were studied at different pH values, ranging from 4.0 to 8.0. The denaturation of C-PC follows the first order kinetics and rate constant at pH 5.0 and temperature 55 degrees C is found to be 4.37 x 10(-5) s(-1), which increases to 5.46 x 10(-1) s(-1) at pH 7.0. The denaturation rate is much higher at 65 degrees C and pH 7.0 (7.96 x 10(-4)), as compared to at pH 5.0 (1.46 x 10(-4)). The thermal stability of C-PC is more at pH 5.0, as compared to other pH values. The observed differences in entropy values at pH 5.0, as compared to other pH values indicate a considerably close fit structure of the protein at pH 5.0, which increases the stability of native structure, even at higher temperature (65 degrees C).  相似文献   

20.
The effect of methanol on the folding of staphylococcal nuclease has been investigated. Equilibrium thermal unfolding transitions were monitored by fluorescence emission. The transition was very sensitive to the presence of methanol (at pH 7.0), the Tm decreased from above 50 degrees C for aqueous solution to below 0 degree C for 70% methanol. The transitions were fully reversible and conformed to two-state behavior. A linear relationship was observed between the hydrophobicity of the solvent and both the Tm and the change in delta G for unfolding. The effect of pH on the transition in 50% methanol at 0 degree C was essentially the same as for aqueous solution, with a cooperative transition in the vicinity of apparent pH (pH*) 4. The unfolding transition was determined as a function of guanidine thiocyanate in aqueous and 50% methanol solvents. The midpoints of the transitions were 0.30 and 0.20 M, respectively, at 2.1 degrees C. The kinetics of folding at 0 degree C were compared in aqueous, 50% methanol and 0.30 M guanidine thiocyanate solvents, by monitoring changes in the tryptophan fluorescence intensity. Triphasic kinetics for refolding in both aqueous and 50% methanol solutions were observed in stopped-flow experiments. In both solvent systems the slowest phase is ascribed to proline isomerization. The kinetics of refolding were monitored at subzero temperatures in 50% methanol at pH* 7.0 in manual mixing experiments. Biphasic kinetics were observed at temperatures between 0 and -35 degrees C. A third, faster phase, was inferred from the missing amplitude. The energies of activation were 20.0 and 17.2 kcal mol-1, respectively, for the two slower phases. At -33.8 degrees C, the observed pseudo first-order rate constants were 1.2 x 10(-3) and 2.1 x 10(-5) s-1. At temperatures above -35 degrees C, the sum of the observed amplitudes was essentially constant at 70-75% of the expected total amplitude. At lower temperatures the amplitude of the refolding reaction decreased, and the native state was not formed (unless the temperature was increased), due to the formation of a trapped intermediate state. This intermediate has circular dichroism and fluorescence properties consistent with a compact state with some molten globule characteristics.  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