首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 46 毫秒
1.
Inoue C  Suzuki T 《Cryobiology》2006,52(1):83-89
The enthalpy relaxation of freeze concentrated sucrose-water glass was investigated using 40% sucrose, differential scanning calorimetry (DSC) with isothermal ageing for 1-6 days at various temperatures (-70, -65, -60, and -55 degrees C). The enthalpy relaxation was observed as an endothermic peak superimposed on the endothermic step-wise change due to the glass transition around -47 degrees C. The enthalpy relaxation was found to increase with ageing time and temperature. An 80% sucrose glass was also investigated at ageing temperatures of -60 and -65 degrees C, and this material exhibited a similar glass transition and enthalpy relaxation to that observed with the frozen 40% sucrose solution. The calculated activation energy of the enthalpy relaxation of the sucrose-water glass was smaller than that reported for pure sucrose. These results suggest that the freeze concentrated sucrose-water glass could have a higher molecular mobility and less stability than pure sucrose glass.  相似文献   

2.
Dissolution of sucrose crystals in the anhydrous sorbitol melt   总被引:1,自引:0,他引:1  
The dissolution of a sugar (sucrose as a model) with higher melting point was studied in a molten food polyol (sorbitol as a model) with lower melting point, both in anhydrous state. A DSC and optical examination revealed the dissolution of anhydrous sucrose crystals (mp 192 degrees C) in anhydrous sorbitol (mp 99 degrees C) liquid melt. The sucrose-sorbitol crystal mixtures at the proportions of 10, 30, 60, 100 and 150 g of sucrose per 100 g of sorbitol were heat scanned in a DSC to above melting endotherm of sorbitol but well below the onset temperature of melting of sucrose at three different temperatures 110, 130 and 150 degrees C. The heat scanning modes used were with or without isothermal holding. The dissolution of sucrose in the sorbitol liquid melt was manifested by an increase in the glass transition temperature of the melt and corresponding decrease in endothermic melting enthalpy of sucrose. At given experimental conditions, as high as 25 and 85% of sucrose dissolved in the sorbitol melt during 1 h of isothermal holding at 110 and 150 degrees C, respectively. Optical microscopic observation also clearly showed the reduction in the size of sucrose crystals in sorbitol melt during the isothermal holding at those temperatures.  相似文献   

3.
Molecular mobility is known to be a key parameter in controlling the physical properties of materials and thus their quality and performance. Beyond glass transition related changes, attention should be called to the impact of local motions remaining in the glassy state. Gelatinized waxy maize starch at different sucrose contents (0-20% solids) was equilibrated between 0 and 14% water and sorption isotherms determined at 25 degrees C. The effect of water and sucrose content on the molecular mobility of glassy starch was investigated by differential scanning calorimetry through enthalpy relaxation studies and dynamical mechanical thermal analysis. The existence of sucrose-starch interactions was suggested by the sorption isotherms not following the expected additivity of the single component sorption curves. Contrary to the glass transition or associated alpha relaxation, water and sucrose affected differently the secondary relaxations. Indeed, the beta relaxation observed around -15 degrees C was shifted to lower temperature upon increasing hydration, and to higher temperature when sucrose content increased, suggesting a hindering of these local motions. Enthalpy relaxation of the ternary mixtures was studied following aging up to 668 h at Tg -15 degrees C. Ternary mixtures exhibited an enthalpy relaxation upon aging lower than starch alone as a sign of lower polymer mobility in the presence of small molecules, contrary to the free volume theory. Relaxation kinetics were characterized with the Cowie-Ferguson model and compared to literature data. The extent of the enthalpy relaxation appeared to be controlled by the distance between the aging temperature and the beta relaxation temperature.  相似文献   

4.
Two endothermic peaks could be observed for five commercial samples of bovine serum albumin (BSA). The smaller peak observed by differential scanning calorimetry (DSC) corresponded to enthalpy relaxation. This peak was followed on storage of BSA, in its glassy state, after it had been heated above its denaturation temperature. Enthalpy and peak temperature increased with duration of storage. On storage for one week at 60 degrees C, a sample at 8.3% moisture showed a peak at 100 degrees C with an energy value of approximately 2 J per g protein. BSA samples were heated within the DSC sufficiently to eliminate the lower enthalpy peak but without altering the denaturation enthotherm. The amount of physical aging shown by these BSA samples was similar to that of the heat-denatured samples. It was concluded that the heating endotherms of dry BSA reflect both the storage and thermal history of the sample. Possible implications of the enthalpy relaxation of BSA on the behavior of this important protein are considered.  相似文献   

5.
1. Differential scanning calorimetry has been used to study the thermal denaturation of lactate dehydrogenase. At pH 7.0 in 0.1 M potassium phosphate buffer, only one transition was observed. Both the enthalpy of denaturation and the melting temperature are linear function of heating rate. The enthalpy is 430 kcal/mol and the melting temperature 61 degrees C at 0 degrees C/min heating rate. The ratio of the calorimetric heat to the effective enthalpy indicated that the denaturation is highly cooperative. Subunit association does not appear to significantly contribute to the enthalpy of denaturation. 2. Both cofactor and sucrose addition stabilized the protein against thermal denaturation. Pyruvate addition produced no changes. Only a small time-dependent destabilization was observed at low concentrations of urea. Large effects were observed in concentrated NaCl solutions and with sulfhydryl-modified lactate dehydrogenase.  相似文献   

6.
Prothrombin denaturation was examined in the presence of Na2EDTA, 5mM CaCl2, and CaCl2 plus membranes containing 1-palmitoyl-2-oleoyl-3-sn-phosphatidylcholine (POPC) in combination with either bovine brain phosphatidylserine (PS) or 1,2-dioleoyl-phosphatidylglycerol (DOPG). Heating denaturation of prothrombin produced thermograms showing two peaks, a minor one at approximately 59 degrees C previously reported to correspond to denaturation of the fragment 1 region (Ploplis, V. A., D. K. Strickland, and F. J. Castellino 1981. Biochemistry. 20:15-21), and a main one at approximately 57-58 degrees C, reportedly due to denaturation of the rest of the molecule (prethrombin 1). The main peak was insensitive to the presence of 5mM Ca2+ whereas the minor peak was shifted to higher temperature (Tm approximately 65 degrees C) by Ca2+. Sufficient concentrations of POPC/bovPS (75/25) large unilamellar vesicles to guarantee binding of 95% of prothrombin resulted in an enthalpy loss in the main endotherm and a comparable enthalpy gain in the minor endotherm accompanying an upward shift in peak temperature (Tm approximately 73 degrees C). Peak deconvolution analysis on the prothrombin denaturation profile and comparison with isolated prothrombin fragment 1 denaturation endotherms suggested that the change caused by POPC/PS vesicles reflected a shift of a portion of the enthalpy of the prethrombin 1 domain to higher temperature (Tm approximately 77 degrees C). The enthalpy associated with this high-temperature endotherm increased in proportion to the surface concentration of PS. By contrast, POPC/DOPG (50/50) membranes shifted the prethrombin 1 peak by 4 degrees C to a lower temperature and the fragment 1 peak by 5 degrees C to a higher temperature. The data lead to a hypothesis that the fragment 1 and prethrombin 1 domains of prothrombin do not denature quite independently and that binding of prothrombin to acidic-lipid membranes disrupts the interaction between these domains. It is further hypothesized that PS containing membranes exert the additional specific effect of decoupling the denaturation of two subdomains of the prethrombin 1 domain of prothrombin.  相似文献   

7.
The glass transition temperature, T(g), and enthalpy relaxation of amorphous lactose glass were investigated by differential scanning calorimetry (DSC) for isothermal aging periods at various temperatures (25, 60, 75, and 90 degrees C) below T(g). Both T(g) and enthalpy relaxation were found to increase with increasing aging time and temperature. The enthalpy relaxation increased approximately exponentially with aging time at a temperature (90 degrees C) close to T(g) (102 degrees C). There was no significant change observed in the enthalpy relaxation around room temperature (25 degrees C) over an aging period of 1month. The Kohlrausch-Williams-Watts (KWW) model was able to fit the experimental enthalpy relaxation data well. The relaxation distribution parameter (beta) was determined to be in the range 0.81-0.89. The enthalpy relaxation time constant (tau) increased with decreasing aging temperature. The observed enthalpy relaxation data showed that molecular mobility in amorphous lactose glass was higher at temperatures closer to T(g). Lactose glass was stable for a long time at 25 degrees C. These findings should be helpful for improving the processing and storage stability of amorphous lactose and lactose containing food and pharmaceutical products.  相似文献   

8.
Solubility of fish (Labio rohita) myosin has been studied at varying temperatures in presence of various inorganic salts like NaCl, KCl, NaBr, Na2SO4, KI, and organic solutes like sucrose and urea. The effect of pH on the solubility has also been studied both in absence and presence of NaCl. Thermal denaturation temperatures of myosin in presence of NaCl, KCl, NaBr and Na2SO4 were found to be 40 degrees, 40 degrees, 45 degrees and 50 degrees C respectively. Thermodynamic parameters like changes in standard free energy (delta G degrees), enthalpy (delta H degrees) and entropy (delta S degrees) for precipitation of myosin from solution phase to gel phase have been evaluated and the physico-chemical aspects have been critically discussed. The average delta G degrees for gel formation varied only between -30 and -40 kJ/mole of myosin, although the nature of solutes, temperature and folding state of protein have been grossly altered. A compensation effect has also been exhibited from the linear plot of average values of delta H degrees against T delta S degrees for various solutes.  相似文献   

9.
Differential scanning calorimetry and cryomicroscopy were used to investigate the effects of type I antifreeze protein (AFP) from winter flounder on 58% solutions of hydroxyethyl starch. The glass, devitrification, and melt transitions noted during rewarming were unaffected by 100 micrograms/ml AFP. Isothermal annealing experiments were undertaken to detect the effects of AFP-induced inhibition of ice crystal growth using calorimetry. A premelt endothermic peak was detected during warming after the annealing procedure. Increasing the duration or the temperature of the annealing for the temperature range from -28 and -18 degrees C resulted in a gradual increase in the enthalpy of the premelt endotherm. This transition was unaffected by 100 micrograms/ml AFP. Annealing between -18 and -10 degrees C resulted in a gradual decrease in the premelt peak enthalpy. This process was inhibited by 100 micrograms/ml AFP. Cryomicroscopic examination of the samples revealed that AFP inhibited ice recrystallization during isothermal annealing at -10 degrees C. Annealing at lower temperatures resulted in minimal ice recrystallization and no visible effect of AFP. Thus, the 100 micrograms/ml AFP to have a detectable influence on thermal events in the calorimeter, conditions must be used that result in significant ice growth without AFP and visible inhibition of this process by AFP.  相似文献   

10.
串珠镰刀菌可利用羟基脯氨酸、蔗糖、甘油和珍珠岩(P)等组成的P液体培养基合成镰刀菌素c(Fc),其最高量为93 6mg/kg有机物,比在玉米渣培养基中形成的Fc量较高。用P液体培养基制备Fc,受蔗糖浓度、胺类和培养温度及培养时间等的影响。实验证明,由1g百姓遭基因氨酸、40g蔗糖和珍珠岩组或的P液体培养基.在28℃培养二周是形成Fc的理想条件。液体培养基中加八珍珠岩,Fc的形成量增加500多倍。  相似文献   

11.
The vesicular stomatitis virus glycoprotein reconstituted into dipalmitoylphosphatidylcholine (DPPC) vesicles exerts a profound effect upon the DPPC gel to liquid-crystalline phase transition. The glycoprotein was reconstituted into DPPC vesicles by octyl glucoside dialysis. The gel to liquid-crystalline phase transition of these vesicles was monitored by differential scanning calorimetry. Vesicles formed in the absence of glycoprotein (600--2100-A diameter) underwent the phase transition at 41.0 degrees C and had an associated enthalpy change of 8.0 +/- 1.6 kcal/mol. Increasing the mole ratio of glycoprotein to DPPC in the vesicles to 0.15 mol % reduced both the transition temperature and the transition enthalpy change. The enthalpy change as a function of the mole percent glycoprotein could be fit to a straight line by a least-squares procedure. Extrapolation of the results to the glycoprotein concentration where the enthalpy change was zero indicated one glycoprotein molecule bound 270 +/- 150 molecules of DPPC.  相似文献   

12.
The effect of pH and temperature on the thermal denaturation of micrococcal nuclease wer4e investigated. The ranges employed were between pH3.30 and pH9.70 and between 10 degrees C and 85 degrees C, respectively. The reversible denaturation involved in the whole process was clearly discriminated from the irreversible one. The former took place with a large enthalpy change of 384 kJ mol(-1) at pH 9.70, where the enzyme exhibited it s maximum activity. The latter probably led to aggregation because the successive long incubation after complete deactivation caused precipitation. A reasonable scheme explaining the process involving both denaturations was proposed and the kinetic on the irreversible deactivation was performed. It was revealed that the irreversible deactivation involved two types of reactions whose activation energies were relatively small: 22.2 kJ mol(-1) and 18.8kJ mol(-1). The presence of sucrose suppressed the reversible denaturation without significant influence on enthalpy change, whereas it affected little the irreversible deactivation kinetically. The effects of pH change and addition of sucrose on the denaturation were discussed thermodynamically, especially in terms of the entropy change. (c) 1994 John Wiley & Sons, Inc.  相似文献   

13.
Salts are present in most amorphous biomaterials such as dried or frozen solid foods, plant seeds, and bacterial spores, and in some pharmaceutical formulations. However, knowledge of how salts modulate the physical properties of amorphous solid sugars, a major component in these systems, is lacking. We have used phosphorescence of the triplet probe erythrosin B (Ery B) to monitor molecular mobility in amorphous sucrose films (dried against P(2)O(5)) containing the salts NaCl, MgCl(2), CaCl(2), NaAcetate, Na(3)Citrate, NaH(2)PO(4), or Na(2)HPO(4) at a mole ratio of 0.2:1 (salt/sucrose). All the salts examined, except NaH(2)PO(4), significantly increased the phosphorescence lifetime of Ery B over the temperature range from 5 to 100 degrees C. This increase is due to a reduction in the rate of collisional quenching of the triplet state due to interactions with the matrix, indicating that these salts decreased the matrix molecular mobility. NaAcetate, Na(3)Citrate, and Na(2)HPO(4) decreased mobility more than NaCl, CaCl(2), or MgCl(2), perhaps due to specific hydrogen bonding interactions between the anion and sucrose. Systematic variations in the probe emission lifetime across the excitation and emission bands at 25 degrees C indicate that there are sites of different mobilities within amorphous solid sucrose; this dynamic site heterogeneity was enhanced in the presence of the divalent cationic salts MgCl(2) and CaCl(2). These results suggest that salts may play a significant role in modulating the mobility, and thus the long-term stability, of amorphous biological matrixes.  相似文献   

14.
A new method was used for reconstituting active sodium deoxycholate solubilized Ca2+-ATPase of rabbit skeletal muscle sarcoplasmic reticulum. Removal of the detergent by dialysis at the pretransition temperature of the pure lipid (22 degrees C) favored the formation of sheet-like structures with a lipid and protein content close to that of the detergent-solubilized sample. Freeze-fracture electron micrographs revealed the Ca2+-ATPase to be organized in rows corresponding to the typical banded pattern seen in low-temperature freeze-fracture micrographs of pure lipid bilayers. Incubation of the sheetlike structures at a temperature (38 degrees C) above the pure lipid main phase transition (33.5 degrees C) caused closure of the sheets into vesicles displaying homogeneous intramembranous particle distributions, at least for membranes containing less than 150 lipids per Ca2+-ATPase. However, in membranes of higher lipid content, free lipid patches were seen both above and below the lipid phase transition. By use of high-sensitivity differential scanning calorimetry, three classes of excess heat capacity peaks were observed in the vesiculated samples. A broadened "free lipid" peak occurred for samples containing between 550 and 200 lipids per protein (Tm = 33.5 degrees C, as for the order-disorder transition in pure lipid vesicles). Between 200 and 150 lipids per Ca2+-ATPase, a broad shoulder became apparent in the range of 29-32 degrees C. Below 150 lipids per Ca2+-ATPase, a peak at 26-28 degrees C became increasingly prominent with lower lipid content. At a lipid to protein ratio of about 30, no peaks in heat capacity were observed. The temperature dependence of diphenylhexatriene fluorescence anisotropy revealed a similar pattern of membrane phase behavior, except that a phase transition was detected at 33.5 degrees C in all membranes studied. On the basis of these observations, we propose that the Ca2+-ATPase is surrounded by a "lipid annulus" of motionally inhibited lipid molecules that do not contribute to a calorimetrically detectable phase transition. Beyond the annulus, "secondary domains" of disrupted lipid packing account for the peak at 26-28 degrees C and the 29-32 degrees C shoulders. At high lipid to protein ratios, the secondary domains coexist with protein-free, lipid-bilayer patches, which account for the peak at 33.5 degrees C.  相似文献   

15.
Kernchen U  Lipps G 《Biochemistry》2006,45(2):594-603
The single-stranded DNA binding protein from Sulfolobus solfataricus (Sso-RPA) binds single-stranded DNA with dissociation constants in the range of 10-30 nM at room temperature. The affinity for DNA decreases at higher temperatures. At 85 degrees C, the optimal growth temperature of the crenarchaeot S. solfataricus, the dissociation constant is only about 1 microM. We analyzed the equilibrium between Sso-RPA and a fluorescently labeled 13 nucleotide oligonucleotide by fluorescence anisotropy measurements in the presence of four different salts and in the temperature range between 10 and 60 degrees C. In the presence of potassium chloride and choline chloride, three to four ions are released upon complexation, independent of the temperature. In contrast, in the presence of potassium fluoride and potassium glutamate, we observed a significant change of the number of ions released when the temperature was varied. The binding reaction is strongly exothermic with enthalpies of about -55 to -70 kJ/mol, depending upon the salt. Van't Hoff analysis suggests that the binding enthalpy is temperature independent.  相似文献   

16.
A Blume  J Tuchtenhagen 《Biochemistry》1992,31(19):4636-4642
The heat of dissociation of the second proton of 1,2-dimyristoylphosphatidic acid (DMPA) was studied as a function of temperature using titration calorimetry. The dissociation of the second proton of DMPA was induced by addition of NaOH. From the calorimetric titration experiment, the intrinsic pK0 for the dissociation reaction could be determined by applying the Gouy-Chapman theory. pK0 decreases with temperature from ca. 6.2 at 11 degrees C to 5.4 at 54 degrees C. From the total heat of reaction, the dissociation enthalpy, delta Hdiss, was determined by subtracting the heat of neutralization of water and the heat of dilution of NaOH. In the temperature range between 2 and 23 degrees C, delta Hdiss is endothermic with an average value of ca. 2.5 kcal.mol-1 and shows no clear-cut temperature dependence. In the temperature range between 23 and 52 degrees C, delta Hdiss calculated after subtraction of the heat of neutralization and dilution is not the true dissociation enthalpy but includes contributions from the phase transition enthalpy, delta Htrans, as the pH jump induces a transition from the gel to the liquid-crystalline phase. The delta Cp for the reaction enthalpy observed in this temperature range is positive. Above 53 degrees C, the pH jump induces again only the dissociation of the second proton, and the bilayers stay in the liquid-crystalline phase. In this temperature range, delta Hdiss seems to decrease with temperature. The thermodynamic data from titration calorimetry and differential scanning calorimetry as a function of pH can be combined to construct a complete enthalpy-temperature diagram of DMPA in its two ionization states.  相似文献   

17.
In this paper, we report on the suitability of solutions containing propane-1,2-diol (propylene glycol, PD), sugars, and salts for the vitrification of the human cell line, ECV304. Cooling (at 10 degrees C/min) and rewarming (at 80 degrees C/min) were at rates that are practicable for the tissues to be studied later. Under these conditions, 45% PD in phosphate-buffered saline (PBS) sometimes froze during cooling and always devitrified during rewarming but both events were avoided if the PBS salts were replaced by an osmotically equivalent concentration of sucrose or trehalose. The effect of such solutions on cells was evaluated using a cell culture assay in which the number of cells recovered after 3 days of culture was divided by the number cells plated, giving a cell multiplication factor or CMF. In the absence of PD the cells tolerated a low-salt concentration in solutions that were made isotonic with sugars, but they recovered poorly when 45% PD was also present. Trehalose gave significantly better recovery than sucrose. When 39% PD and 15% trehalose were included in a low-salt vehicle solution (LSV) that contained approximately 5% of the total salt concentration of PBS (this solution was designated LSV/39/15), the cells exhibited approximately 40% of untreated control CMF following exposure for 9min. LSV/39/15 vitrifies with a glass transition temperature of -102 degrees C, does not devitrify when warmed at 80 degrees C/min, and has suitable dielectric properties for uniform and rapid dielectric heating. An improved method for adding and removing LSV/39/15 gave a CMF of approximately 55% of untreated controls. Using this method, 1.0ml suspensions of ECV304 cells was cooled to, and stored briefly at, -120 degrees C and then rewarmed by immersion in a 37 degrees C water bath ( approximately 75 degrees C/min). The CMF of the cooled samples was similar to that of the exposure-only controls, approximately 50% of the untreated control CMF in both cases.  相似文献   

18.
The thermal denaturation, aggregation, and degradation of hen egg white ovalbumin dissolved in distilled and deionized water (60 mg/ml, pH 7.5) was investigated by differential scanning calorimetry (DSC), polyacrylamide gel electrophoresis (PAGE), and viscosity measurement. Two independent endothermic peaks were observed up to 180 degrees C by the DSC analysis. The first peak appeared at around 80 degrees C, corresponding to the denaturation temperature of ovalbumin. The second peak occurred around 140 degrees C due to the degradation of protein molecules as judged from the analysis by SDS-PAGE. The viscosity of the ovalbumin solution increased dramatically above 88 degrees C and maintained almost the same value up until heating to 140 degrees C. The increase in viscosity after heating to 88 degrees C was due to the denaturation and subsequent aggregation of ovalbumin molecules as observed by SDS-PAGE. The decrease in viscosity of the samples heated above 150 degrees C appears to have been the result of degradation of the ovalbumin molecules.  相似文献   

19.
Body surface temperature of individually marked wasps (Paravespula vulgaris, Vespidae, Hymenoptera) was measured by infrared thermography during repeated visits to a feeding bowl without injuring them or disturbing their behavior. Wasps were fed 0.5, 1 and 2 mol/l sucrose solution at two ambient temperatures.Thoracic temperature varied significantly in dependence on food quality (sucrose concentration of solution). At the higher ambient temperatures of 26.1-30.2 degrees C mean thoracic surface temperatures from different experiments were 35.3 and 38.0 degrees C when the wasps took a 0.5 mol/l sucrose solution, 37.0, 38.7 and 38.7 degrees C when they took a 1 mol/l solution, and 39.1 degrees C when they took a 2 mol/l sucrose solution. At the lower ambient temperatures of 17.6-21.0 degrees C thoracic temperatures were lower but the effect of different sucrose concentrations was similar: 34.7 degrees C with a 0.5 mol/l and 36.1 degrees C with a 1 mol/l sucrose solution. The concentration effect amounted to about 10-25% of the whole variability of thorax temperature. By contrast, the temperatures of the head and abdomen did not follow the changes in thorax temperature according to changes in sucrose concentration closely, which suggests that the pattern of haemolymph circulation may have changed after landing, during the wasps' stay at the feeder. At initial landing at the feeders thoracic temperatures where equal to (three of eight tests) or lower (five of eight tests) than at final departure.The correlation of thorax temperature with food quality probably reflects the wasps' level of excitement and motivation to collect the food, which allows them to balance energetic investment with profitability of foraging and the needs of flight muscle performance and motility.  相似文献   

20.
The thermal denaturation of clathrin-coated vesicles isolated from bovine brain tissue has been studied by differential scanning calorimetry and has been compared to basket structures reformed from isolated triskelion trimers of clathrin and to isolated triskelions. The coated vesicles and reformed baskets displayed similar, yet distinct, thermal behavior. Calorimetric data of the coated vesicles exhibited a single denaturation transition peak at 55.9 +/- 0.1 degrees C, skewed to low temperatures whereas the thermograms for the reformed baskets exhibited a broad transition peak at 53.1 +/- 0.1 degrees C and a peak at 56.3 +/- 0.1 degrees C. Neither transition was reversible. The specific transition enthalpy was 11.5 +/- 1.0 J g-1 for the coated vesicles and the total transition enthalpy was 9.1 +/- 0.3 J g-1 for the reformed baskets. In contrast, isolated triskelions showed no thermal transition between 15 and 90 degrees C. Although the coated vesicles and the reformed baskets have similar stability reflecting their similar structures, the coated vesicles appear to be marginally more stable than the reformed baskets. The complexity of the transition profiles and their lack of symmetry suggest the existence of several, somewhat independent, domains unique to the cage-like structure of the coated vesicles and reformed baskets.  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