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 281 毫秒
1.
In laboratory cultures of ice nucleation-active (Ice+) Erwinia herbicola isolates, it has been difficult to achieve high-level expression of ice nuclei, especially nuclei active at temperatures warmer than −5°C (i.e., type 1 ice nuclei). Here we demonstrate that starvation for phosphate and exposure to low temperature triggers expression of ice nuclei in E. herbicola cultures. Starvation for nitrogen, sulfur, or iron was less effective. Under optimal conditions with two different strains, essentially all cells produced ice nuclei active at −10°C or warmer, with an average of 22% containing type 1 ice nuclei within 1 h of a low-temperature shift. These conditions did not greatly enhance the shedding of ice nucleation-active membrane vesicles that are known to be produced by Ice+ E. herbicola isolates. These results support the theory that the Ice+ phenotype may allow nutrient-limited epiphytes to trigger freezing damage, releasing nutrients from host plants. Received: 2 November 1997 / Accepted: 5 January 1998  相似文献   

2.
Release of ice nuclei in the growth medium of recombinant Halomonas elongata cells expressing the inaZ gene of Pseudomonas syringae was studied in an attempt to produce cell-free active ice nuclei for biotechnological applications. Cell-free ice nuclei were not retained by cellulose acetate filters of 0.2 microm pore size. Highest activity of cell-free ice nuclei was obtained when cells were grown in low salinity (0.5-5% NaCl, w/v). Freezing temperature threshold, estimated to be below -7 degrees C indicating class C nuclei, was not affected by medium salinity. Their density, as estimated by Percoll density centrifugation, was 1.018 +/- 0.002 gml(-1) and they were found to be free of lipids. Ice nuclei are released in the growth medium of recombinant H. elongata cells probably because of inefficient anchoring of the ice-nucleation protein aggregates in the outer membrane. The ice+ recombinant H. elongata cells could be useful for future use as a source of active cell-free ice nucleation protein.  相似文献   

3.
The growth of ice nucleation-active and near-isogenic ice nucleation-deficient (Ice) Pseudomonas syringae strains coexisting on leaf surfaces was examined to determine whether competition was sufficient to account for antagonism of phylloplane bacteria. The ice nucleation frequency spectra of 46 IceP. syringae mutants, obtained after mutagenesis with ethyl methanesulfonate, differed both quantitatively and qualitatively, but the mutants could be grouped into four distinct phenotypic classes. The numbers of ice nucleation-active bacteria and ice nuclei active at -5 degrees C were reduced on plants colonized with IceP. syringae mutant strains before challenge inoculations with an IceP. syringae wild-type strain. Frost injury to plants pretreated with IceP. syringae strains was also reduced significantly compared with that to control plants and was correlated with the population size of the IceP. syringae strain and with the numbers of ice nuclei active at -5 degrees C. An IceP. syringae strain colonized leaves, flowers, and young fruit of pears in field experiments and significantly reduced the colonization of these tissues by IceP. syringae strains and Erwinia amylovora as compared with untreated trees.  相似文献   

4.
5.
6.
Release of cell-free ice nuclei by Erwinia herbicola.   总被引:14,自引:3,他引:11       下载免费PDF全文
Several ice-nucleating bacterial strains, including Erwinia herbicola, Pseudomonas fluorescens, and Pseudomonas syringae isolates, were examined for their ability to shed ice nuclei into the growth medium. Only E. herbicola isolates shed cell-free ice nuclei active at -2 to -10 degrees C. These cell-free nuclei exhibited a freezing spectrum similar to that of ice nuclei found on whole cells, both above and below -5 degrees C. Partially purified cell-free nuclei were examined by density gradient centrifugation, chemical and enzymatic probes, and electron microscopy. Ice-nucleating activity in these cell-free preparations was associated with outer membrane vesicles shed by cells and was sensitive to protein-modifying reagents.  相似文献   

7.
A novel ice-nucleating bacterium (INB) was isolated from Ross Island, Antarctica. INBs could be isolated more frequently than was generally thought. INB strain IN-74 was found in the white colony group. Strain IN-74 was identified from its taxonomic characteristics as a novel INB, Pseudomonas antarctica IN-74. When strain IN-74 was cultured aerobically in a medium consisting of the ice-nucleating broth (pH 7.0) for 6 days at 4 degrees C, the ice-nucleating activity of strain IN-74 cells was obtained. Strain IN-74 cells produced ice nuclei only at extremely low growth temperatures. The nuclei appeared to be less thermolabile than those of INB Pseudomonas fluorescens KUIN-1. The freezing difference spectra in D2O and H2O at ice-nucleating temperature for strain IN-74 cells and conventional INBs (Pseudomonas fluorescens KUIN-1, Pseudomonas viridiflava KUIN-2, and Pseudomonas syringae C-9) exhibited different curves.  相似文献   

8.
We have analyzed the relationship between expression of the transformed phenotype and thyroid hormone (triiodothyronine, T3) inducibility of Na,K-ATPase and binding of 125I-epidermal growth factor (EGF) to cell membrane receptors in wild-type (wt) and mutant type 5 adenovirus (Ad5)-transformed CREF cells displaying a cold-sensitive (cs) expression of the transformed phenotype. CREF cells respond to thyroid hormone treatment with increased Na,K-ATPase activity and bind similar levels of 125I-EGF at 32 degrees C, 37 degrees C and 39.5 degrees C. In contrast, CREF cells transformed by wt Ad5 or the E1a plus E1b-transforming genes of wt Ad5 are refractile to T3 treatment and bind lower levels of 125I-EGF than CREF cells at all three temperatures. By employing a series of cloned CREF cell lines transformed by a host-range cold-sensitive mutant virus, H5hr1 or H5dl101, or the E1a or E1a plus E1b genes from these viruses, we have investigated expression of the transformed state and its relationship with hormone inducibility and EGF binding. When cs virus, cs E1a- or cs E1a plus E1b-transformed CREF clones were grown at 32 degrees C, a nonpermissive transforming temperature in which cs-transformed cells exhibit properties similar to untransformed CREF cells, T3 induced Na,K-ATPase activity and these cells bound similar levels of 125I-EGF as CREF cells. However, when cs virus- and cs Ela plus E1b-transformed CREF clones were incubated at 37 degrees C or 39.5 degrees C, temperatures at which cs-transformed cells exhibit properties similar to wt Ad5-transformed CREF cells, they did not respond to T3 and bound lower levels of 125I-EGF than CREF cells. In the case of cs E1a-transformed CREF clones, thyroid hormone responsiveness was observed at both 32 degrees C and 37 degrees C, but not at 39.5 degrees C. By performing temperature shift experiments--i.e. 32 degrees C to 37 degrees C, 32 degrees C to 39.5 degrees C, 37 degrees C to 32 degrees C, and 39.5 degrees C to 32 degrees C, it was demonstrated that after a shift from lower to higher temperature a 24-hr lag period was required for cs-transformed CREF cells to lose T3 inducibility and exhibit reduced EGF binding, whereas 96 hr after a shift from higher to lower temperature a 96-hr lag period was required for cs-transformed cells to regain T3 inducibility and increased 125I-EGF binding.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

9.
10.
Proliferation of mammalian cells can be controlled by low cultivation temperature. However, depending on cell type and expression system, varying effects of a temperature shift on heterologous protein production have been reported. Here, we characterize growth behavior and productivity of the Chinese hamster ovary (CHO) cell line XM111-10 engineered to synthesize the model-product-secreted alkaline phosphatase (SEAP). Shift of cultivation temperature from 37 degrees C to 30 degrees C caused a growth arrest mainly in the G1 phase of the cell cycle concomitant with an up to 1.7-fold increase of specific productivity. A low temperature cultivation provided 3.4 times higher overall product yield compared to a standard cultivation at 37 degrees C. The cellular and molecular mechanisms underlying the effects of low temperature on growth and productivity of mammalian cells are poorly understood. Separation of total protein extracts by two-dimensional gel electrophoresis showed altered expression levels of CHO-K1 proteins after decrease in cultivation temperature to 30 degrees C. These changes in the proteome suggest that mammalian cells respond actively to low temperature by synthesizing specific cold-inducible proteins. In addition, we provide the first evidence that the cold response of mammalian cells includes changes in postranslational protein modifications. Two CHO proteins were found to be phosphorylated at tyrosine residues following downshift of cultivation temperature to 30 degrees C. Elucidating cellular events during cold exposure is necessary for further optimization of host-cell lines and expression systems and can provide new strategies for metabolic engineering.  相似文献   

11.
A temperature increase from 35° to 40–42°C enhances the rise of cytoplasmic serine proteinase (ISP1) activity in Bacillus megaterium incubated in a sporulation medium. A temperature shift from 27°C in the growth medium to 35°C in the sporulation medium has the same effect. Elevated temperature stimulates the increase of ISP1 level when applied immediately after the transfer of cells from the growth to the sporulation medium (at T0) or at T3, when sporulation becomes irreversible. The cytoplasmic PMSF-resistant activity or the proteolytic activity associated with the membrane fraction is stimulated only slightly or not at all. A temperature increase to 45–47°C suppresses the rise of proteolytic activities in all cell fractions. In addition to the elevation of the ISP1 activity by an upward temperature shift, the rise of this enzyme in nongrowing cells is also stimulated by osmotic stress. In growing populations, in contrast to the rise of the ISP1 activity caused by elevated temperature in nongrowing cells, this proteinase is induced by low temperatures (24–27°C). The ISP1 activity roughly correlates with the enzyme protein concentration determined by immunoblotting.  相似文献   

12.
The heat shock protein ClpB is essential for acquired thermotolerance in cyanobacteria and eukaryotes and belongs to a diverse group of polypeptides which function as molecular chaperones. In this study we show that ClpB is also strongly induced during moderate cold stress in the unicellular cyanobacterium Synechococcus sp. strain PCC 7942. A fivefold increase in ClpB (92 kDa) content occurred when cells were acclimated to 25 degrees C over 24 h after being shifted from the optimal growth temperature of 37 degrees C. A corresponding increase occurred for the smaller ClpB' (78 kDa), which arises from a second translational start within the clpB gene of prokaryotes. Shifts to more extreme cold (i.e., 20 and 15 degrees C) progressively decreased the level of ClpB induction, presumably due to retardation of protein synthesis within this relatively cold-sensitive strain. Inactivation of clpB in Synechococcus sp. increased the extent of inhibition of photosynthesis upon the shift to 25 degrees C and markedly reduced the mutant's ability to acclimate to the new temperature regime, with a threefold drop in growth rate. Furthermore, around 30% fewer delta clpB cells survived the shift to 25 degrees C after 24 h compared to the wild type, and more of the mutant cells were also arrested during cell division at 25 degrees C, remaining attached after septum formation. Development of a cold thermotolerance assay based on cell survival clearly demonstrated that wild-type cells could acquire substantial resistance to the nonpermissive temperature of 15 degrees C by being pre-exposed to 25 degrees C. The same level of cold thermotolerance, however, occurred in the delta clpB strain, indicating ClpB induction is not necessary for this form of thermal resistance in Synechococcus spp. Overall, our results demonstrate that the induction of ClpB contributes significantly to the acclimation process of cyanobacteria to permissive low temperatures.  相似文献   

13.
Rates of assembly and degradation of bacterial ice nuclei   总被引:6,自引:0,他引:6  
The kinetics of ice-nucleus assembly from newly synthesized nucleation protein were observed following induction of nucleation gene expression in the heterologous host Escherichia coli. Assembly was significantly slower for the small proportion of ice nuclei active above -4.4 degrees C; this was consistent with the belief that these nuclei comprise the largest aggregates of nucleation protein. The kinetics of nucleus degradation were followed after inhibiting protein synthesis. Nucleation activity and protein showed a concerted decay, indicating that most of the functional ice nuclei are in equilibrium with a single cellular pool of nucleation protein. A minority of the ice nuclei decayed much more slowly than the majority; presumably their nucleation protein was distinct either by virtue of different structure or different subcellular compartmentalization, or because of its presence in a metabolically distinct subpopulation of cells.  相似文献   

14.
Membrane fluidity adaptation to the low growth temperature in Bacillus subtilis involves two distinct mechanisms: (1) long-term adaptation accomplished by increasing the ratio of anteiso- to iso-branched fatty acids and (2) rapid desaturation of fatty acid chains in existing phospholipids by induction of fatty acid desaturase after cold shock. In this work we studied the effect of medium composition on cold adaptation of membrane fluidity. Bacillus subtilis was cultivated at optimum (40 degrees C) and low (20 degrees C) temperatures in complex medium with glucose or in mineral medium with either glucose or glycerol. Cold adaptation was characterized by fatty acid analysis and by measuring the midpoint of phospholipid phase transition T(m) (differential scanning calorimetry) and membrane fluidity (DPH fluorescence polarization). Cells cultured and measured at 40 degrees C displayed the same membrane fluidity in all three media despite a markedly different fatty acid composition. The T(m) was surprisingly the highest in the case of a culture grown in complex medium. On the contrary, cultivation at 20 degrees C in the complex medium gave rise to the highest membrane fluidity with concomitant decrease of T(m) by 10.5 degrees C. In mineral media at 20 degrees C the corresponding changes of T(m) were almost negligible. After a temperature shift from 40 to 20 degrees C, the cultures from all three media displayed the same adaptive induction of fatty acid desaturase despite their different membrane fluidity values immediately after cold shock.  相似文献   

15.
16.
17.
Wowk B  Fahy GM 《Cryobiology》2002,44(1):14-23
The simple linear polymer polyglycerol (PGL) was found to apparently bind and inhibit the ice nucleating activity of proteins from the ice nucleating bacterium Pseudomonas syringae. PGL of molecular mass 750 Da was added to a solution consisting of 1 ppm freeze-dried P. syringae 31A in water. Differential ice nucleator spectra were determined by measuring the distribution of freezing temperatures in a population of 98 drops of 1 microL volume. The mean freezing temperature was lowered from -6.8 degrees C (control) to -8.0,-9.4,-12.5, and -13.4 degrees C for 0.001, 0.01, 0.1, and 1% w/w PGL concentrations, respectively (SE < 0.2 degrees C). PGL was found to be an ineffective inhibitor of seven defined organic ice nucleating agents, whereas the general ice nucleation inhibitor polyvinyl alcohol (PVA) was found to be effective against five of the seven. The activity of PGL therefore seems to be specific against bacterial ice nucleating protein. PGL alone was an ineffective inhibitor of ice nucleation in small volumes of environmental or laboratory water samples, suggesting that the numerical majority of ice nucleating contaminants in nature may be of nonbacterial origin. However, PGL was more effective than PVA at suppressing initial ice nucleation events in large volumes, suggesting a ubiquitous sparse background of bacterial ice nucleating proteins with high nucleation efficiency. The combination of PGL and PVA was particularly effective for reducing ice formation in solutions used for cryopreservation by vitrification.  相似文献   

18.
19.
Many plants acquire freezing tolerance through cold acclimatization (CA), a prolonged exposure to low but non-freezing temperatures at the onset of winter. CA is associated with gene expression that requires transient calcium influx into the cytosol. Alfalfa (Medicago sativa) cells treated with agents blocking this influx are unable to cold-acclimatize. Conversely, chemical agents causing increased calcium influx induce cold acclimatization-specific (cas) gene expression in alfalfa at 25 degrees C. How low temperature triggers calcium influx is, however, unknown. We report here that induction of a CA-specific gene (cas30), calcium influx and freezing tolerance at 4 degrees C are all prevented by cell membrane fluidization, but, conversely, are induced at 25 degrees C by membrane rigidification. cas30 expression and calcium influx at 4 degrees C are also prevented by jasplakinolide (JK), an actin microfilament stabilizer, but induced at 25 degrees C by the actin microfilament destabilizer cytochalasin D (CD). JK blocked the membrane rigidifier-induced, but not the calcium channel agonist-induced cas30 expression at 25 degrees C. These findings indicate that cytoskeleton re-organization is an integral component in low-temperature signal transduction in alfalfa cell suspension cultures, serving as a link between membrane rigidification and calcium influx in CA.  相似文献   

20.
When Tetrahymena thermophila cells growing at 30 degrees C are shifted to either 40 or 43 degrees C, the kinetics and extent of induction of heat shock mRNAs in both cases are virtually indistinguishable. However, the cells shifted to 40 degrees C show a typical induction of heat shock protein (HSP) synthesis and survive indefinitely (100% after 24 h), whereas those at 43 degrees C show an abortive synthesis of HSPs and die (less than 0.01% survivors) within 1 h. Cells treated at 30 degrees C with the drugs cycloheximide or emetine, at concentrations which are initially inhibitory to protein synthesis and cell growth but from which cells can eventually recover and resume growth, are after this recovery able to survive a direct shift from 30 to 43 degrees C (ca. 70% survival after 1 h). This induction of thermotolerance by these drugs is as efficient in providing thermoprotection to cells as is a prior sublethal heat treatment which elicits the synthesis of HSPs. However, during the period when drug-treated cells recover their protein synthesis ability and simultaneously acquire the ability to subsequently survive a shift to 43 degrees C, none of the major HSPs are synthesized. The ability to survive a 1-h, 43 degrees C heat treatment, therefore, does not absolutely require the prior synthesis of HSPs. But, as extended survival at 43 degrees Celsius depends absolutely on the ability of cells to continually synthesize HSPs, it appears that a prior heat shock as well as the recovery from protein synthesis inhibition elicits a change in the protein synthetic machinery which allows the translation of HSP mRNAs at what would otherwise be a nonpermissive temperature for protein synthesis.  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