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.
Aims: For this study, we performed a genetic screen of S. cerevisiae’s deletion library for mutants sensitive to dehydration stress, with which we aimed to discover cell dehydration–tolerant genes. Methods and Results: We used a yeast gene deletion set (YGDS) of 4850 viable mutant haploid strains to perform a genome‐wide screen for the identification of desiccation stress modifiers. SIP18 is among the genes identified as essential for cells surviving to drying/rehydration process. Deletion of SIP18 promotes the accumulation of reactive oxygen species and enhances apoptotic and necrotic cell death in response to dehydration/rehydration process. Conclusions: SIP18p acts as an inhibitor of apoptosis in yeast under dehydration stress, as suggested by its antioxidative capacity through the ROS accumulation reduction after an H2O2 attack. Significance and Impact of the Study: To our knowledge, this is the first systematic screen for the identification of putative genes essential to overcoming cell dehydration process. A broad range of identified genes could help to understand why some strains of high biotechnological interest cannot cope with the drying and rehydration treatments. Dehydration sensitivity makes these strains not suitable to be commercialized by yeast manufactures.  相似文献   

2.
The State of Water in Human and Dog Red Cell Membranes   总被引:8,自引:8,他引:0  
The apparent activation energy for the water diffusion permeability coefficient, Pd, across the red cell membrane has been found to be 4.9 ± 0.3 kcal/mole in the dog and 6.0 ± 0.2 kcal/mole in the human being over the temperature range, 7° to 37°C. The apparent activation energy for the hydraulic conductivity, Lp, in dog red cells has been found to be 3.7 ± 0.4 kcal/mole and in human red cells, 3.3 ± 0.4 kcal/mole over the same temperature range. The product of Lp and the bulk viscosity of water, η, was independent of temperature for both dog and man which indicates that the geometry of the red cell membrane is not temperature-sensitive over our experimental temperature range in either species. In the case of the dog, the apparent activation energy for diffusion is the same as that for self-diffusion of water, 4.6–4.8 kcal/mole, which indicates that the process of water diffusion across the dog red cell membrane is the same as that in free solution. The slightly, but significantly, higher activation energy for water diffusion in human red cells is consonant with water-membrane interaction in the narrower equivalent pores characteristic of these cells. The observation that the apparent activation energy for hydraulic conductivity is less than that for water diffusion across the red cell membrane is characteristic of viscous flow and suggests that the flow of water across the membranes of these red cells under an osmotic pressure gradient is a viscous process.  相似文献   

3.
Calorimetric studies of the reduction of free oxygen in solution by sodium dithionite are in agreement with a stoichiometry of 2 moles Na2S2O4 per mole of oxygen. The reaction is biphasic with ΔHt - 118±7 kcal mol?1 (?494 ± 29 kJ mol?1). The initial phase of the reaction proceeds with an enthalpy change of ca ?20 kcal (?84 kJ) and occurs when 0.5 moles of dithionite have been added per mole dioxygen present. This could be interpreted as the enthalpy change for the addition of a single electron to form the superoxide anion. Further reduction of the oxygen to water by one or more additional steps is accompanied by an enthalpy change of ca ?100 kcal (?418. 5 kJ). Neither of these reductive phases is consistent with the formation of hydrogen peroxide as an intermediate. The reduction of hydrogen peroxide by dithionite in 0.1 M phosphate buffer, pH 7.15, is a much slower process and with an enthalpy change of ca ? 74 kcal mol?1 (?314 kJ mol?1). Dissociation of oxyhemoglobin induced by the reduction of free oxygen tension with dithionite also shows a stoichiometry of 2 moles dithionite per mole oxygen present and an enthalpy change of ca. ?101 ±9 kcal mol?1 (?423± 38 kJ mol?1). The difference in the observed enthalpies (reduction of dioxygen vs. oxyhemoglobin) has been attributed to the dissociation of oxyhemoglobin, which is 17 kcal mol?1 (71 kJ mol?1).  相似文献   

4.
Yeast tRNA3Leu is one of several tRNA molecules which can adopt a stable, biologically inactive, denatured conformation. The circular dichroism of the native and denatured conformers differs, providing the basis for the present study of the mechanism for the renaturation process. Conversion of the denatured structure to the native takes place in two steps: a rapid change occurring immediately on addition of Mg++, followed by a slower, strongly temperature-dependent step which returns the molecule to its biologically active state. Optimal kinetic data for the second step could be obtained at 285 nm. Analysis of the time dependence of Δε285 by the Guggenheim method demonstrated that this step follows first-order kinetics. The temperature dependence of the rate constants over the range 32–41°C yielded the following parameters for the rate-limiting step: Ea = 69 kcal/mole, ΔH? = 69 kcal/mole, and ΔS? = 146 cal/mole deg. Values of this magnitude are typical of order—order transitions in nucleic acids.  相似文献   

5.
Absorbance-temperature profiles have been determined for the following self-complementary oligonucleotides or equimolar paris of complementary oligonucleotides containing GC base pairs: A2GCU2, A3GCU3, A4GCU4, A6CG + CGU6, A8CG + CGU8, A4G2 + C2U4, A5G2 + C2U5, A4G3 + C3U4, and A5G3 + C3U5. In all cases cooperative melting transitions indicate double-helix formation. As was found previously, the stability of GC containing oligomer helices is much higher than that of AU helices of corresponding length. Moreover, helices with the same length and base composition but different sequences also have quite different stabilites. The melting curves were andlyzed using a zipper model and the thermodynamic parameters for the AU pairs determined previously. The effect of single-strand stacking was considered separately. According to this model, the formation of a GC pair from unstacked single strands is associated with an ethalpy change of ?15 kcal/mole. Due to the high degree of single-strand stacking at room temperature the enthalpy change for the formation of GC pairs from unstacked single strands is only ?5 to ?6 kcal/mole. (The corresponding parameters for AU pairs are ?10.7 kcal/mole and ?5 to ?6 kcal/mole.) The sequence dependence of helix stability seems to be primarily entropic since no differences in ΔH were seen among the sequence isomers. The kinetics of helix formation was investigated for the same molecules using the temperature jump technique. Recombination of strands is second order with rate constants in the range of 105 to 107M?1 sec?1 depending on the chain length and the nucleotide sequence. Within a series of oligomers of a given type, the rates of recombination decrease with increasing chain length. Oligomers with the sequence AnGCUn recombine six to eight times slower than the other oligomers of corresponding chain length. The experimental enthalpies of activation of 6 to 9 kcal/mole suggest a nucleation length of one or two GC base pairs. The helix dissociation process has rate constants between 0.5 and 500 sec?1 and enthalpies of activation of 25 to 50 kcal/mole. An increase of chain length within a given nucleotide series leads to decreased rates of dissociation and increased enthalpies of activation. An investigation of the effect of ionic strength on AnGCUn helix formation showed that the rates of recombination increase considerably with increased ionic strength.  相似文献   

6.
Yeast cells are often used as a model system in various experiments. Moreover, due to their high metabolic activity, yeast cells have a potential to be applied as elements in the design of biofuel cells and biosensors. However a wider application of yeast cells in electrochemical systems is limited due to high electric resistance of their cell wall. In order to reduce this problem we have polymerized conducting polymer polypyrrole (Ppy) directly in the cell wall and/or within periplasmic membrane. In this research the formation of Ppy was induced by [Fe(CN)6]3−ions, which were generated from K4[Fe(CN)6], which was initially added to polymerization solution. The redox process was catalyzed by oxido-reductases, which are present in the plasma membrane of yeast cells. The formation of Ppy was confirmed by spectrophotometry and atomic force microscopy. It was confirmed that the conducting polymer polypyrrole was formed within periplasmic space and/or within the cell wall of yeast cells, which were incubated in solution containing pyrrole, glucose and [Fe(CN)6]4−. After 24 h drying at room temperature we have observed that Ppy-modified yeast cell walls retained their initial spherical form. In contrast to Ppy-modified cells, the walls of unmodified yeast have wrinkled after 24 h drying. The viability of yeast cells in the presence of different pyrrole concentrations has been evaluated.  相似文献   

7.
One important economical method for producing singlecell protein is to spray dry the cultured cells. This study presents some preliminary data on the effects of spray drying on cell viability. Under conditions similar to those for the production of spray-dried milk, 4–5 log cycles destruction occurred. The results indicate that, the activation energy for thermal destruction of yeast was reduced from the normal heat treatment value of 84 kcal/°K mole to about 38 kcal/°K mole.  相似文献   

8.
In this study Saccharomyces cerevisiae yeast cells was used as a novel vehicle for encapsulation of vitamin D3. The effects of initial cholecalciferol concentration (100,000 and 500,000 IU/g yeast), yeast cell pretreatment (plasmolysis with NaCl) and drying method (spray or freeze drying) on microcapsules properties were investigated. It was found that the vitamin concentration and drying method had significant influence on encapsulation efficiency (EE) and size of yeast microcapsules. Furthermore, EE values were more increased by the plasmolysis treatment. The highest EE was obtained for plasmolysed and spray dried yeast cells prepared using initial cholecalciferol concentration of 2.5 mg per gram of yeast cells (76.10?±?6.92%). The values of mean particle size were 3.43–7.91 μm. The presence of cholecalciferol in yeast microcapsules was confirmed by X-ray diffraction (XRD) and Fourier transform-infrared (FT-IR) analyses. The in vitro cholecalciferol release from yeast microcapsules in phosphate buffer saline solution (PBS) followed a controlled release manner consistent with a Fickian diffusion mechanism. In addition, the release studies in simulated gastrointestinal tract showed sustained release of cholecalciferol in the stomach condition and significant release in intestinal medium.  相似文献   

9.
Summary Water transport mechanisms in rabbit proximal convoluted cell membranes were examined by measurement of: (1) osmotic (P f ) and diffusional (P d ) water permeabilities, (2) inhibition ofP f by mercurials, and (3) activation energies (E a ) forP f .P f was measured in PCT brush border (BBMV) and basolateral membrane (BLMV) vesicles, and in viable PCT cells by stopped-flow light scattering;P d was measured in PCT cells by proton NMR Ti relaxation times using Mn as a paramagnetic quencher. In BLMV,P f (0.019 cm/sec, 23°C) was inhibited 65% by 5mm pCMBS and 75% by 300 m HgCl2 (K l =42 m);E a increased from 3.6 to 7.6 kcal/mole (15–40°C) with 300 m HgCl2. In BBMV,P f (0.073 cm/sec, 23°C,E a =2.8 kcal/mole, <33°C and 13.7 kcal/mole, >33°C) was inhibited 65% with HgCl2 withE a =9.4 kcal/mole (15–45°C). Mercurial inhibition in BLMV and BBMV was reversed with 10 m mercaptoethanol. Viable PCT cells were isolated from renal cortex by Dounce homogenization and differential seiving. Impedence sizing studies show that PCT cells are perfect osmometers (100–1000 mOsm). Assuming a cell surface-to-volume ratio of 25,000 cm–1,P f was 0.010±0.002 cm/sec (37°C) andP d was 0.0032 cm/sec.P f was independent of osmotic gradient size (25–1000 mOsm) withE a 2.5 kcal/mole (<27°C) and 12.7 kcal/mole (>27°C). CellP f was inhibited 53% by 300 m HgCl2 (23°C) withE a 6.2 kcal/mole. These findings indicate that cellP f is not restricted by extracellular or cytoplasmic unstirred layers and that cellP f is not flow-dependent. The high BLMV and BBMVP f , inhibition by HgCl2, lowE a which increases with inhibition, and the measuredP f /P d >1 in cells in the absence of unstirred layers provide strong evidence for the existence of water channels in proximal tubule brush border and basolateral membranes. These channels are similar to those found in erythrocytes and are likely required for rapid PCT transcellular water flow.  相似文献   

10.
H J Hinz  W Haar  T Ackermann 《Biopolymers》1970,9(8):923-936
The enthalpies of the helix-coil transitions of the ordered polynucleotide systems of poly(inosinic acid)–poly(cytidylic acid) [poly(I + C)], (helical duplex), and of poly (inosinic acid) [poly(I + I + I)], (proposed secondary structure: a triple-stranded helical complex), were determined by using an adiabatic twin-vessel differential calorimeter. Measuring the temperature course of the heat capacity of the aqueous polymer solutions, the enthalpy values for the dissociation of the helical duplex poly (I + C) and the three-stranded helical complex poly(I + 1 + 1), respectively, were obtained by evaluating the additional heat capacity involved in the conformational change of the polynucleotide system in the transition range. The ΔH values of the helix-coil transition of poly (I + C) resulting from the analysis of the calorimetric measurements vary between the limits 6.5 ± 0.4 kcal/mole (I + C) and 8.4 ± 0.4 kcal/mole (I + C). depending on the variation of the cation concentration ranging from 0.063 mole cations kg H2O to 1.003 mole cations/kg H2O. The calorimetric investigation of an aqueous poly I solution (cation concentration 1.0 mole/kg H2O) yielded the enthalpy value ΔH = 1.9 ± 0.4 kcal/mole (I), a result which has been interpreted qualitatively following current models of inter- and intramolecular forces of biologically significant macromolecules. Additional information on the transition behavior of poly(I+ C)Was obtained by ultraviolet and infrared absorption measurements.  相似文献   

11.
Aims: The aim of the present study was to evaluate and compare freezing and freeze‐drying treatments for conserving Rahnella aquatilis (BNM 0523) with the goal to achieve an adequate commercial formulation of this biocontrol agent. Methods and Results: The effect of several protective agents, rehydration media and freezing temperatures on the viability and functional activity of the R. aquatilis was investigated. The storage stability at 3 months and 4 years was determined by checking the viability of the cells and their biocontrol capability against Botrytis cinerea by measuring the percentage of reduction of disease severity on apple. The best results were obtained by the freeze‐drying of the cells using a mixture of skimmed nonfat milk 10%, yeast extract 0·5% and glucose 1% as the protecting and rehydrating medium, and a quickly freezing (?70°C) before the freeze‐drying. In this case, the viability of the cells after 4 years was 98%, and their antagonistic ability showed a little decrease with respect fresh cells. Conclusions: The studies showed that R. aquatilis was resistant to freezing and freeze‐drying when it was used a mixture of cryoprotectants and that it was possible to obtain inoculums with high viability and good effectiveness for reduction of decay caused by B. cinerea. Significance and Impact of the study: This study is probably the first report about the resistance of R. aquatilis to freezing and freeze‐drying treatments and shows that these operations could be useful for obtaining a commercial formulation of this biocontrol agent.  相似文献   

12.
Studies were made of the death kinetics of Escherichia coli cells heated at 46 to 56°C in 0.05M phosphate buffer (pH 7.0) containing either an amphoteric surfactant (Tego 15DL, 1–10 μg/ml) or sorbic acid (0.5 to 3%). A linear relationship was obtained on the Arrhenius plot for the death of cells heated with each antimicrobial agent. The kinetics of the action of the surfactant, however, differed from that of sorbic acid. With the amphoteric surfactant, the activation enthalpy of the death reaction decreased from 108 to 51 kcal/mol as the concentration of surfactant was increased in the range tested although the death rate remained high; whereas with sorbic acid the activation enthalpy remained fairly constant (104 ± 9 kcal/mol) independent of its concentration and the death rate was similarly high. Further, in the action of the amphoteric surfactant, a thermodynamic compensation effect was observed, the compensation temperature being 334°K (61°C), i.e., relatively close to the observed temperatures. For sorbic acid, however, this temperature seemed to be too high to observe when determined from the Arrhenius plot. The data of the dependency of the death-rate constant upon the concentration of antimicrobial agent indicated a similar difference in the action of the two agents. Based on our results and on data obtained by other workers, we propose that antimicrobial agents that enhance cellular death induced by heating can be characterized into two types.  相似文献   

13.
M W Springgate  D Poland 《Biopolymers》1973,12(10):2241-2260
The cooperative nature of the binding between polycytidylate and the oligoinosinates I(pI)5–10 has been determined. Using the data of Tazawa, Tazawa, and Ts'o, it is shown that knowledge of the slope of the adsorption isothern allows one to determine the oligomer-polymer binidng constant, the oligomer–oligomer interaction constant, and the average degree of association (cooperative clustering) of the oligomers on the polymer. Knowledge of the above equilibrium constants as a function of temperature yields the respective thermodynamic parameters; no assumptions need to be made about the nature of the equilibrium constants or the thermodynamic parameters. For very long chains of polycytidylate, simple, explicit relations are given for the determination of the equilibrium constants involved. For finite chains of polycytidylate, the calculation of a single graph for each oligomer and polymer size allows the equilibrium constants to be determined for all experimental conditions of temperature and concentration. We find that the enthalpy and entropy of binding an oligomer n, bases to be δHn = ±13.7 ? n(6.65) and δSn = +32.5 ? n(18.8) given, respectively, in kcal/mole and e.u.; these parameters predict a melting temperature of 81°C for the poly(I)·poly(C) complex compared with the experimental value of 75°C. If the enthalpy is interpreted as arising from a sum of hydrogen bonding and stacking interactions, then the enthalpy of stacking is ?13.7 kcal/mole while the enthalpy of hydrogen bonding is +7 ± 4 kcal/mole; the positive enthalpy of hydrogen bonding presumably is a result of the fact that in the inosine-cytosine base pair, only two of the three sites on cytosine can hydrogen bond, the third being blocked from hydrogen bonding with water. The enthalpy of interaction between neighboring bound oligomers is found to be ?10.4 kcal/mole while the corresponding entropy is ?26.1 e.u. The binding is bound to be cooperative, though the extent of clustering varies markedly with temperature; the average number of oligomers in a cluster on the polymer is found to about five at a melting temperature of 25°C. The approach and equations given have generally applicability to oligomer-polymer associations.  相似文献   

14.
A soluble NAD-dependent alcohol dehydrogenase (ADH) activity was detected in mycelium and yeast cells of wild-type Mucor rouxii. In the mycelium of cells grown in the absence of oxygen, the enzyme activity was high, whereas in yeast cells, ADH activity was high regardless of the presence or absence of oxygen. The enzyme from aerobically or anaerobically grown mycelium or yeast cells exhibited a similar optimum pH for the oxidation of ethanol to acetaldehyde (∼pH 8.5) and for the reduction of acetaldehyde to ethanol (∼pH 7.5). Zymogram analysis conducted with cell-free extracts of the wild-type and an alcohol-dehydrogenase-deficient mutant strain indicated the existence of a single ADH enzyme that was independent of the developmental stage of dimorphism, the growth atmosphere, or the carbon source in the growth medium. Purified ADH from aerobically grown mycelium was found to be a tetramer consisting of subunits of 43 kDa. The enzyme oxidized primary and secondary alcohols, although much higher activity was displayed with primary alcohols. K m values obtained for acetaldehyde, ethanol, NADH2, and NAD+ indicated that physiologically the enzyme works mainly in the reduction of acetaldehyde to ethanol. Received: 11 March 1999 / Accepted: 14 July 1999  相似文献   

15.
Isotopic exchange kinetics at equilibrium for E. coli native aspartate transcarbamylase at pH 7.8, 30 °C, are consistent with an ordered BiBi substrate binding mechanism. Carbamyl phosphate binds before l-Asp, and carbamyl-aspartate is released before inorganic phosphate. The rate of [14C]Asp C-Asp exchange is much faster than [32P]carbamyl phosphate Pi exchange. Phosphate, and perhaps carbamyl phosphate, appears to bind at a separate modifier site and prevent dissociation of active-site bound Pi or carbamyl phosphate. Initial velocity studies in the range of 0–40 °C reveal a biphasic Arrhenius plot for native enzyme: Ea (>15 °C) = 6.3 kcal/ mole and Ea (<15 °C) = 22.1 kcal/mole. Catalytic subunits show a monophasic plot with Ea ? 20.2 kcal/mole. This, with other data, suggests that with native enzyme a conformational change accompanying aspartate association contributes significantly to rate limitation at t > 15 °C, but that catalytic steps become definitively slower below 15 °C. Model kinetics are derived to show that this change in mechanism at low temperature can force an ordered substrate binding system to produce exchange-rate patterns consistent with a random binding system with all exchange rates equal. The nonlinear Arrhenius plot also has important consequences for current theories of catalytic and regulatory mechanisms for this enzyme.  相似文献   

16.
Unlike technical grade yeast RNA, which was confirmed to contain several per cent of 2′–5′ phosphodiester linkages, RNA prepared from different kinds of commercial yeast in a cold room consisted exclusively of 3′–5′ phosphodiester linkages. Heat treatment of the 3′–5′ linked RNA solution resulted in partial isomerization of the internucleotide linkage of the polynucleotide chain (C3′-C5′->C2′-C5′). The isomerization of RNA occurred in the presence of water, at high temperature, and under acidic conditions. Treatment of dry RNA at 100°C for 2hr did not result in any detectable isomerization. The isomerization was actually observed in yeast RNA when yeast cells suspended in sodium chloride solution were heated. It is concluded therefore that 2′-5′ phosphodiester linkages found in technical grade RNA had been formed neither at a step of precipitating RNA with acid nor at a step of drying RNA, but had been formed at a step of heat extraction of RNA from yeast. When 0.1 % poly (A) solution, pH 4.8, was heated for 20 hr in a boiling water bath, the isomerization proceeded during the first 6hr, and finally reached about 37%, irrespective of chain length.  相似文献   

17.
pKa1 values of 3-methoxy-N-desmethyldiazepam in acetonitrile and methanol containing various acid concentrations were determined by spectrophotometry to be 3.5 and 1.3, respectively. Temperature-dependent racemization of enantiomeric 3-methoxy-N-desmethyldiazepam in methanol containing 0.5 M H2SO4 was studied by circular dichroism spectropolorimetry and the racemization reactions were found to follow apparent first-order kinetics. Thermodynamic parameters of the racemization reaction were found to be: Eact = 18.8 kcal/mol, and at 25°C: ΔH? = 18.3 kcal/mol, ΔS? = ?14.8 entropy unit, and ΔG? = 22.7 kcal/mol, respectively. The racemization had an isotope effect (kH/kD) of 1.6 at 42°C. Based on the results of this report and those of earlier reports by other investigators, a nucleophilically solvated C3 carbocation intermediate resulting from either a P (plus) or an M (minus) conformation is proposed to be an intermediate and responsible for the stereoselective nucleophilic substitution and the subsequent racemization of 3-methoxy-N-desmethyldiazepam enantiomers. © 1993 Wiley-Liss, Inc.  相似文献   

18.
Dietmar Prschke 《Biopolymers》1971,10(10):1989-2013
The properties of oligonucleotide helices of adeuylic- and uridylic acid oligomers have been investigated by measurements of hypo-and hyperchromieity. High ionic strengths favor the formation of triple helices. Thus, the double helix-coil transition can be studied (without interference by triple helices) only at low ionic-strength. A “phase diagram” is given representing the Tm-values of the various transitions at different ionic strengths for the system A(pA)17 + U(pU)17. Oligonucleolides of chain lengths <8 always form both double and triple helices at the nucleotide concentrations required for base pairing. For this reason the double helix-coil transition without coupling of the triple helix equilibrium can only be measured for chain lengths higher than 7. Melting curves corresponding to this transition have been determined for chain lengths 8, 9, 10, 11, 14 and 18 at different concentrations. An increase in nucleotide concentration leads to an increase in melting temperature. The shorter the chain length the lower the Tm-value and the broader the helix-coil transition. The experimental transition curves have been analysed according to a staggering zipper model with consideration of the stacking of the adeuylic acid single strands and the electrostatic repulsion of tlip phosphate charges on opposite strands. The temperature dependence of the nucleation parameter has been accounted for by a slacking factor x. The stacking factor expresses the magnitude of the stacking enthalpy. By curve fitting xwas computed to be 0.7, corresponding to a stacking enthalpy of about S kcal/mole. The model described allows the reproduction of the experimental transition curves with relatively high accuracy. In an appendix the thermodynamic parameters of the stacking equilibrium of poly A and of the helix-coil equilibria of poly A + poly U at neutral pH are calculated (ΔHA = ?7.9 kcal/mole for the poly A stacking and ΔH12 = ?10.9 kcal/mole for the formation of the double helix from the randomly coiled single strands). A formula for the configurational entropy of polymers derived by Flory on the basis of a liquid lattice model is adapted to calculate the stacking entropies of adenylic oligomers.  相似文献   

19.
The biocontrol yeast Pichia anomala J121 can effectively reduce mould growth on moist cereal grains during airtight storage. Practical use of microorganisms requires formulated products that meet a number of criteria. In this study we compared different formulations of P. anomala. The best way to formulate P. anomala was freeze-drying. The initial viability was as high as 80%, with trehalose previously added to the yeast. Freeze-dried products could be stored at temperatures as high as 30 °C for a year, with only a minor decrease in viability. Vacuum-drying also resulted in products with high storage potential, but the products were not as easily rehydrated as freeze-dried samples. Upon desiccating the cells using fluidised-bed drying or as liquid formulations, a storage temperature of 10 °C was required to maintain viability. Dependent on the type of formulation, harvesting of cells at different nutritional stresses affected the initial viabilities, e.g. the initial viability for fluidised-bed-dried cells was higher when the culture was fed with excess glucose, but for freeze-drying it was superior when cells were harvested after depletion of carbon. Using micro-silos we found that the biocontrol activity remained intact after drying, storage and rehydration for all formulations.  相似文献   

20.
Synopsis Glutaraldehyde reacts readily with various proteins in solution. With high concentrations of both, the solutions become yellow and many proteins form a gel. At low concentrations the reactions may be followed by the changes in the u.v. spectrum between 250 and 300 nm. The reverse reaction does not proceed to any detectable extent. The kinetics are pseudo-first-order. The activation energies for the reactions between proteins and glutaraldehyde were found to be about II kcal/mole. This suggests that the proteins have not been denatured to any marked extent by the glutaraldehyde fixation. The rates of reactions increase with pH. The rate of formation of glutaraldehyde-protein links per protein molecule glutarated is approximately I sec–1 mol–1 1.  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