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 484 毫秒
1.
The response to heat stress in six yeast species isolated from Antarctica was examined. The yeast were classified into two groups: one psychrophilic, with a maximum growth temperature of 20°C, and the other psychrotrophic, capable of growth at temperatures above 20°C. In addition to species-specific heat shock protein (hsp) profiles, a heat shock (15°C–25°C for 3 h) induced the synthesis of a 110-kDa protein common to the psychrophiles, Mrakia stokesii, M. frigida, and M. gelida, but not evident in Leucosporidium antarcticum. Immunoblot analyses revealed heat shock inducible proteins (hsps) corresponding to hsps 70 and 90. Interestingly, no proteins corresponding to hsps 60 and 104 were observed in any of the psychrophilic species examined. In the psychrotrophic yeast, Leucosporidium fellii and L. scottii, in addition to the presence of hsps 70 and 90, a protein corresponding to hsp 104 was observed. In psychrotrophic yeast, as observed in psychrophilic yeast, the absence of a protein corresponding to hsp 60 was noted. Relatively high endogenous levels of trehalose which were elevated upon a heat shock were exhibited by all species. A 10 Celsius degree increase in temperature above the growth temperature (15°C) of psychrophiles and psychrotrophs was optimal for heat shock induced thermotolerance. On the other hand, in psychrotrophic yeast grown at 25°C, only a 5 Celsius degree increase in temperature was necessary for heat shock induced thermotolerance. Induced thermotolerance in all yeast species was coincident with hsp synthesis and trehalose accumulation. It was concluded that psychrophilic and psychrotrophic yeast, although exhibiting a stress response similar to mesophilic Saccharomyces cerevisiae, nevertheless had distinctive stress protein profiles. Received: August 7, 1997 / Accepted: October 22, 1997  相似文献   

2.
To examine whether recent evolutionary history affects the expression of Hsp70, the major heat-induced-heat shock protein in Drosophila melanogaster, we measured Hsp70 expression, thermotolerance, and hsp70 gene number in replicate populations undergoing laboratory evolution at different temperatures. Despite Hsp70's ancient and highly conserved nature, experimental evolution effectively and replicably modified its expression and phenotype (thermotolerance). Among five D. melanogaster populations founded from a common ancestral population and raised at three different temperatures (one at 18°C, two each at 25°C and 28°C) for twenty years, Hsp70 expression varies in a consistent pattern: the replicate 28°C lines expressed 30–50% less Hsp70 than the other lines at a range of inducing temperatures. This modification was refractory to acclimation, and correlated with thermotolerance: the 28°C lines had significantly lower inducible tolerance of 38.5°C and 39°C. We verified the presence of five hsp70 genes in the genome of each line, excluding copy number variation as a candidate molecular basis of the evolved difference in expression. These findings support the ability of Hsp70 levels in D. melanogaster populations to change over microevolutionary time scales and implicate constancy of environmental temperature as a potentially important selective agent.  相似文献   

3.
Growth temperature has a marked influence on the thermotolerance of photosystem II (PSII), which is the most heat‐sensitive component of photosynthesis. Using Synechocystis sp. PCC 6803 we have established that thylakoids isolated from cells grown at 38°C have a greater degree of thermotolerance than those isolated from cells grown at 25°C. Reconstitution experiments using Triton X‐100 protein extracts of these thylakoids added to Triton‐treated thylakoid membranes further indicated that the 38°C Triton extract contains proteins that are directly capable of enhancing PSII thermotolerance. We have used 4‐plex iTRAQ, extensive off‐line fractionation and sample re‐injection to comprehensively identify the differences between these two preparations that may be responsible for the observed effects on PSII thermotolerance. This has resulted in the reproducible identification of 168 proteins out of a total of 385 distinct proteins. Our results have identified 15 proteins whose levels are increased in extracts that result in increased thermotolerance of PSII and 33 proteins whose levels decrease. Notably, components of the cytochrome b6/f and NADH dehydrogenase complexes, crucial components in electron transport, are approximately twofold more abundant in 38°C thylakoid extracts. The possible biological importance of these changes is discussed.  相似文献   

4.
Non‐lethal heat‐shock (HS) treatment has previously been shown to induce thermotolerance in soybean (Glycine max cv. Kaohsiung No.8) seedlings. This acquired thermotolerance correlates with the de novo synthesis of heat‐shock proteins (HSPs). Interestingly, we found that ethanol treatments also elicited HS‐like responses in aetiolated soybean seedlings at their normal growth temperature of 28 °C. Northern blot analyses revealed that the expression of HS genes hsp17.5, hsp70 and hsc 70 was induced by ethanol. Radioactive amino acids were preferentially incorporated into high molecular weight (HMW) HSPs rather than class I low molecular weight (LMW) HSPs during non‐lethal ethanol treatments. Immunoblot analysis confirmed that no accumulation of class I LMW HSPs occurred after non‐lethal ethanol treatment. Pre‐treatment with a non‐lethal dose of ethanol did not provide thermotolerance, as the aetiolated soybean seedlings could not survive a subsequent heat shock of 45 °C for 2 h. In contrast, non‐lethal HS pre‐treatment, 40 °C for 2 h, conferred tolerance on aetiolated soybean seedlings to otherwise lethal treatments of 7·5% ethanol for 8 h or 10% ethanol for 4 h. These results suggest that plant class I LMW HSPs may play important roles in providing both thermotolerance and ethanol tolerance.  相似文献   

5.
Small heat shock proteins are ubiquitous in all three domains (Archaea, Bacteria and Eukarya) and possess molecular chaperone activity by binding to unfolded polypeptides and preventing aggregation of proteins in vitro. The functions of a small heat shock protein (S.so-HSP20) from the hyperthermophilic archaeon, Sulfolobus solfataricus P2 have not been described. In the present study, we used real-time polymerase chain reaction analysis to measure mRNA expression of S.so-HSP20 in S. solfataricus P2 and found that it was induced by temperatures that were substantially lower (60°C) or higher (80°C) than the optimal temperature for S. solfataricus P2 (75°C). The expression of S.so-HSP20 mRNA was also up-regulated by cold shock (4°C). Escherichia coli cells expressing S.so-HSP20 showed greater thermotolerance in response to temperature shock (50°C, 4°C). By assaying enzyme activities, S.so-HSP20 was found to promote the proper folding of thermo-denatured citrate synthase and insulin B chain. These results suggest that S.so-HSP20 promotes thermotolerance and engages in chaperone-like activity during the stress response.  相似文献   

6.
Similarly to Helicobacter pylori but unlike Vibrio cholerae O1/O139, Campylobacter jejuni is non‐motile at 20°C but highly motile at ≥37°C. The bacterium C. jejuni has one of the highest swimming speeds reported (>100 μm/s), especially at 42°C. Straight and spiral bacterial shapes share the same motility. C. jejuni has a unique structure in the flagellate polar region, which is characterized by a cup‐like structure (beneath the inner membrane), a funnel shape (opening onto the polar surface) and less dense space (cytoplasm). Other Campylobacter species (coli, fetus, and lari) have similar motility and flagellate polar structures, albeit with slight differences. This is especially true for Campylobacter fetus, which has a flagellum only at one pole and a cup‐like structure composed of two membranes.  相似文献   

7.
Tic110 is a major component of the chloroplast protein import translocon. Two functions with mutually exclusive structures have been proposed for Tic110: a protein‐conducting channel with six transmembrane domains and a scaffold with two N‐terminal transmembrane domains followed by a large soluble domain for binding transit peptides and other stromal translocon components. To investigate the structure of Tic110, Tic110 from Cyanidioschyzon merolae (CmTic110) was characterized. We constructed three fragments, CmTic110A, CmTic110B and CmTic110C, with increasing N‐terminal truncations, to perform small‐angle X‐ray scattering (SAXS) and X‐ray crystallography analyses and Dali structural comparison. Here we report the molecular envelope of CmTic110B and CmTic110C determined by SAXS, and the crystal structure of CmTic110C at 4.2 Å. Our data indicate that the C‐terminal half of CmTic110 possesses a rod‐shaped helix‐repeat structure that is too flattened and elongated to be a channel. The structure is most similar to the HEAT‐repeat motif that functions as scaffolds for protein–protein interactions.  相似文献   

8.
A series of experiments were designed to study the effect of elevated temperatures on developmental competence of bovine oocytes and embryos produced in vitro. In experiment 1, the effect of heat shock (HS) by a mild elevated temperature (40.5°C) for 0, 30, or 60 min on the viability of in vitro matured (IVM) oocytes was tested following in vitro fertilization (IVF) and culture. No significant difference was observed between the control (39°C) and the heat‐treated groups in cleavage, blastocyst formation, or hatching (P > 0.05). In experiment 2, when the HS temperature was increased to 41.5°C, neither the cleavage rate nor blastocyst development was affected by treatment. However, the rate of blastocyst hatching appeared lower in the HS groups (13% in control group vs. 3.9% and 5.6% in 30 min and 60 min, respectively; P < 0.05). When IVM oocytes were treated at 43°C prior to IVF (experiment 3), no difference was detected in blastocyst and expanded blastocyst development following heat treatment for 0, 15, or 30 min, but heat treatment of oocytes for 45 or 60 min significantly reduced blastocyst and expanded blastocyst formation (P < 0.05). In experiment 4, the thermotolerance of day 3 and day 4 bovine IVF embryos were compared. When embryos were pre‐treated with a mild elevated temperature (40.5°C) for 1 hr, and then with a higher temperature (43°C) for 1 hr, no improvement in thermotolerance of the embryos was observed as compared to those treated at 43°C alone. However, a higher thermotolerance was observed in day 4 than day 3 embryos. In conclusion, treatment at 43°C, but not 40.5°C or 41.5°C significantly reduced oocyte developmental competence. An increase in thermotolerance was observed from day 3 to day 4 of in vitro embryonic development, which corresponds to the maternal to zygotic transition of gene expression in bovine embryos. Mol. Reprod. Dev. 53:336–340, 1999. © 1999 Wiley‐Liss, Inc.  相似文献   

9.
Abstract The whitefly Bemisia tabaci harbors Portiera aleyrodidarum, an obligatory symbiotic bacterium, as well as several secondary symbionts, including Rickettsia, Hamiltonella, Wolbachia, Arsenophonus, Cardinium and Fritschea, the function of which is unknown. In Israel, Rickettsia is found in both the B and Q of B. tabaci biotypes, and while all other secondary symbionts are located in the bacteriomes, Rickettsia can occupy most of the body cavity of the insect. We tested whether Rickettsia influences the biology of B. tabaci and found that exposing a Rickettsia‐containing population to increasing temperatures significantly increases its tolerance to heat shock that reached 40°C, compared to a Rickettsia‐free population. This increase in tolerance to heat shock was not associated with specific induction of heat‐shock protein gene expression; however, it was associated with reduction in Rickettsia numbers as was assessed by quantitative real‐time polymerase chain reaction and fluorescence in situ hybridization analyses. To assess the causes for thermotolerance when Rickettsia is reduced, we tested whether its presence is associated with the induction of genes required for thermotolerance. We found that under normal 25°C rearing temperature, genes associated with response to stress such as cytoskeleton genes are induced in the Rickettsia‐containing population. Thus, the presence of Rickettsia in B. tabaci under normal conditions induces the expression of genes required for thermotolerance that under high temperatures indirectly lead to this tolerance.  相似文献   

10.
Pre‐exposure to mild heat stress enhances the thermotolerance of insects. Stress hardening is a beneficial physiological plasticity, but the mechanism underlying it remains elusive. Here we report that reactive oxygen species (ROS) concentrations were quickly and transiently elevated in the armyworms, Mythimna separata, by exposing them to 40°C, but not other tested temperatures. Larvae exposed to 40°C had subsequently elevated antioxidant activity and the highest survival of all tested heating conditions. The elevation of ROS after lethal heating at 44°C for 1 h was approximately twofold compared to heating at 40°C. Injection of an optimal amount of hydrogen peroxide (H2O2) similarly caused sequential elevation of ROS and antioxidant activity in the test larval hemolymph, which led to significantly enhanced survival after lethal heat stress. The H2O2‐induced thermotolerance was abolished by coinjection of potent antioxidants such as ascorbic acid or N‐acetylcysteine. Both preheating at 40°C and H2O2 injection enhanced expression of genes encoding superoxide dismutase 1, catalase, and heat shock protein 70 in the fat body of test larvae, indicating the adequate heat stress induced a transient elevation of ROS, followed by upregulation of antioxidant activity. We infer that thermal stress hardening is induced by a small timely ROS elevation that triggers a reduction–oxidation signaling mechanism.  相似文献   

11.
The sycamore lace bug, Corythucha ciliata (Say) (Hemiptera: Tingidae), is an invasive exotic pest on Platanus trees in China. This study assessed the thermotolerance of C. ciliata in the laboratory. Detailed experiments were conducted on the effects of high temperature (35, 37, 39, 41, 43, and 45 °C), duration of exposure (0.5, 1, 2, 4, 6, and 8 h), and developmental stage (egg, nymph, and adult) on survival of the bug. Meanwhile, the effects of heat hardening on survival at lethal temperature (exposure to 33, 35, 37, 39, and 41 °C for 1 h prior to transfer to 43 °C for 2 h) were also assessed for nymphs and adults. Survival of eggs, nymphs, and adults was not affected by temperatures between 35 and 39 °C, but declined rapidly with increasing duration of exposure (from 0.5 to 8 h) at temperatures ≥41 °C. The lethal temperature that caused mortality of 50% (Ltemp50) of all developmental stages decreased with increasing duration of exposure from 0.5 to 8 h. The Ltemp50 for nymphs was 44.3, 42.0, and 39.0 °C after 0.5, 2, and 8 h exposure, respectively. Thermotolerance was the highest in eggs, followed by adults and then nymphs. Thermotolerance was slightly greater for adult males than for adult females. The ability of nymphs, females, and males to survive exposure to 43 °C for 2 h significantly increased by heat hardening, i.e., by exposure to a non‐lethal high temperature for 1 h; the optimal heat‐hardening temperature was 37 °C. The results indicate that survival of C. ciliata at heat‐shock temperatures depended on both the temperature and the duration of exposure, and the tolerance to heat shock was enhanced by heat hardening. The thermotolerance of C. ciliata may partially explain why C. ciliata has been rapidly spreading on Platanus trees in southern provinces of China.  相似文献   

12.
Thermotolerance of photosynthesis in salt‐adapted Atriplex centralasiatica plants (100–400 mm NaCl) was evaluated in this study after detached leaves and whole plants were exposed to high temperature stress (30–48 °C) either in the dark or under high light (1200 mol m?2 s?1). In parallel with the decrease in stomatal conductance, intercellular CO2 concentration and CO2 assimilation rate decreased significantly with increasing salt concentration. There was no change in the maximal efficiency of PSII photochemistry (Fv/Fm) with increasing salt concentration, suggesting that there was no damage to PSII in salt‐adapted plants. On the other hand, there was a striking difference in the response of PSII and CO2 assimilation capacity to heat stress in non‐salt‐adapted and salt‐adapted leaves. Leaves from salt‐adapted plants maintained significantly higher Fv/Fm values than those from non‐salt‐adapted leaves at temperatures higher than 42 °C. The Fv/Fm differences between non‐salt‐adapted and salt‐adapted plants persisted for at least 24 h following heat stress. Leaves from salt‐adapted plants also maintained a higher net CO2 assimilation rate than those in non‐salt‐adapted plants at temperatures higher than 42 °C. This increased thermotolerance was independent of the degree of salinity since no significant changes in Fv/Fm and net CO2 assimilation rate were observed among the plants treated with different concentrations of NaCl. The increased thermotolerance of PSII induced by salinity was still evident when heat treatments were carried out under high light. Given that photosynthesis is considered to be the physiological process most sensitive to high temperature damage, increased thermotolerance of photosynthesis may be of significance since A. centralasiatica, a typical halophyte, grows in the high salinity regions in the north of China, where the temperature in the summer is often as high as 45 °C.  相似文献   

13.
Larvae of the goldenrod gall moth, Epiblema scudderiana, use a freeze avoidance strategy of cold hardiness to survive the winter. A key metabolic adaption that supports subzero survival is the accumulation of large amounts of glycerol as a colligative antifreeze. Production of glycerol relies on polyol dehydrogenase (PDH) which catalyzes the NADPH‐dependent conversion of glyceraldehyde into glycerol. Kinetic analysis of PDH from E. scudderiana revealed significant changes in properties as a result of subzero temperature acclimation; the Km for glyceraldehyde in 5°C‐acclimated larvae was 7.0 mM and doubled in ? 15°C‐exposed larvae. This change suggested that PDH is regulated by a state‐dependent covalent modification. Indeed, high and low Km forms could be interconverted by incubating larval extracts in vitro under conditions that stimulated either endogenous protein kinases or protein phosphatases. Protein kinase incubations doubled the Km glyceraldehyde of the 5°C enzyme, whereas protein phosphatase incubations decreased the Km of the ? 15°C enzyme by about 50%. PDH was purified by ion exchange and affinity chromatography steps and then subjected to electrophoresis. Staining with ProQ Diamond phosphoprotein stain showed a much higher phosphate content of PDH from ? 15°C‐acclimated larvae, a result that was further confirmed by immunoblotting that showed a much greater phosphoserine content on the ? 15°C enzyme. These experiments established that PDH is regulated by state‐dependent reversible phosphorylation in E. scudderiana and suggest that this regulatory mechanism makes a significant contribution to controlling the synthesis, maintenance, and degradation of glycerol pools over the winter months. © 2011 Wiley Periodicals, Inc.  相似文献   

14.
The methylotrophic yeast Hansenula polymorpha has the potential to be used in the process of simultaneous saccharification and fermentation (SSF) of xylan derived xylose at elevated temperatures. To improve parameters of high‐temperature resistance and high‐temperature fermentation of H. polymorpha, strains carrying deletion of acid trehalase gene (ATH1) and overexpressing genes coding for heat‐shock proteins Hsp16p and Hsp104p were constructed. Results indicate that the corresponding recombinant strains have up to 12‐fold increased tolerance to heat‐shock treatment. The deletion of ATH1 gene and constitutive expression of HSP16 and HSP104 resulted in up to 5.8‐fold improvement of ethanol production from xylose at 50°C. Although the maximum ethanol concentration achieved from xylose was 0.9 g L−1, our model H. polymorpha strains with elevated thermotolerance can be further modified by metabolic engineering to construct improved high‐temperature ethanol producers from this pentose. Biotechnol. Bioeng. 2009; 104: 911–919. © 2009 Wiley Periodicals, Inc.  相似文献   

15.
We investigated the heat tolerance of adults of three replicated lines of Drosophila melanogaster that have been evolving independently by laboratory natural selection for 15 yr at “nonextreme” temperatures (18°C, 25°C, or 28°C). These lines are known to have diverged in body size and in the thermal dependence of several life-history traits. Here we show that they differ also in tolerance of extreme high temperature as well as in induced thermotolerance (“heat hardening”). For example, the 28°C flies had the highest probability of surviving a heat shock, whereas the 18°C flies generally had the lowest probability. A short heat pretreatment increased the heat tolerance of the 18°C and 25°C lines, and the threshold temperature necessary to induce thermotolerance was lower for the 18°C line than for the 25°C line. However, neither heat pretreatment nor acclimation to different temperatures influenced heat tolerance of the 28°C line, suggesting the loss of capacity for induced thermotolerance and for acclimation. Thus, patterns of tolerance of extreme heat, of acclimation, and of induced thermotolerance have evolved as correlated responses to natural selection at nonextreme temperatures. A genetic analysis of heat tolerance of a representative replicate population each from the 18°C and 28°C lines indicates that chromosomes 1, 2, and 3 have significant effects on heat tolerance. However, the cytoplasm has little influence, contrary to findings in an earlier study of other stocks that had been evolving for 7 yr at 14°C versus 25°C. Because genes for heat stress proteins (hsps) are concentrated on chromosome 3, the potential role of hsps in the heat tolerance and of induced thermotolerance in these naturally selected lines is currently unclear. In any case, species of Drosophila possess considerable genetic variation in thermal sensitivity and thus have the potential to evolve rapidly in response to climate change; but predicting that response may be difficult.  相似文献   

16.
Mutualistic organisms can be particularly susceptible to climate change stress, as their survivorship is often limited by the most vulnerable partner. However, symbiotic plasticity can also help organisms in changing environments by expanding their realized niche space. Coral–algal (Symbiodinium spp.) symbiosis exemplifies this dichotomy: the partnership is highly susceptible to ‘bleaching’ (stress‐induced symbiosis breakdown), but stress‐tolerant symbionts can also sometimes mitigate bleaching. Here, we investigate the role of diverse and mutable symbiotic partnerships in increasing corals' ability to thrive in high temperature conditions. We conducted repeat bleaching and recovery experiments on the coral Montastraea cavernosa, and used quantitative PCR and chlorophyll fluorometry to assess the structure and function of Symbiodinium communities within coral hosts. During an initial heat exposure (32 °C for 10 days), corals hosting only stress‐sensitive symbionts (Symbiodinium C3) bleached, but recovered (at either 24 °C or 29 °C) with predominantly (>90%) stress‐tolerant symbionts (Symbiodinium D1a), which were not detected before bleaching (either due to absence or extreme low abundance). When a second heat stress (also 32 °C for 10 days) was applied 3 months later, corals that previously bleached and were now dominated by D1a Symbiodinium experienced less photodamage and symbiont loss compared to control corals that had not been previously bleached, and were therefore still dominated by Symbiodinium C3. Additional corals that were initially bleached without heat by a herbicide (DCMU, at 24 °C) also recovered predominantly with D1a symbionts, and similarly lost fewer symbionts during subsequent thermal stress. Increased thermotolerance was also not observed in C3‐dominated corals that were acclimated for 3 months to warmer temperatures (29 °C) before heat stress. These findings indicate that increased thermotolerance post‐bleaching resulted from symbiont community composition changes, not prior heat exposure. Moreover, initially undetectable D1a symbionts became dominant only after bleaching, and were critical to corals' resilience after stress and resistance to future stress.  相似文献   

17.
A thermotolerant Saccharomyces cerevisiae yeast strain, YK60‐1, was bred from a parental strain, MT8‐1, via stepwise adaptation. YK60‐1 grew at 40°C, a temperature at which MT8‐1 could not grow at all. YK60‐1 exhibited faster growth than MT8‐1 at 30°C. To investigate the mechanisms how MT8‐1 acquired thermotolerance, DNA microarray analysis was performed. The analysis revealed the induction of stress‐responsive genes such as those encoding heat shock proteins and trehalose biosynthetic enzymes in YK60‐1. Furthermore, nontargeting metabolome analysis showed that YK60‐1 accumulated more trehalose, a metabolite that contributes to stress tolerance in yeast, than MT8‐1. In conclusion, S. cerevisiae MT8‐1 acquired thermotolerance by induction of specific stress‐responsive genes and enhanced intracellular trehalose levels. © 2013 American Institute of Chemical Engineers Biotechnol. Prog., 29:1116–1123, 2013  相似文献   

18.
The terrestrial cyanobacterium Nostoc commune Vaucher ex Bornet et Flahault occurs worldwide, including in Japan and on the Antarctic continent. The terrestrial green alga Prasiola crispa (Lightf.) Kütz. is also distributed in Antarctica. These two species need to acclimate to the severe Antarctic climate including low ambient temperature and desiccation under strong light conditions. To clarify this acclimation process, the physiological characteristics of the photosynthetic systems of these two Antarctic terrestrial organisms were assessed. The relative rate of photosynthetic electron flow in N. commune collected in Japan and in Antarctica reached maxima at 900 and 1,100 μmol photons · m?2 · s?1, respectively. The difference seemed to reflect the presence of high amounts of UV‐absorbing substances within the Antarctic cyanobacterium. On the other hand, the optimal temperatures for photosynthesis at the two locations were 30°C–35°C and 20°C–25°C, respectively. This finding suggested a decreased photosynthetic thermotolerance in the Antarctic strain. P. crispa exhibited desiccation tolerance and dehydration‐induced quenching of PSII fluorescence. Re‐reduction of the photooxidized PSI reaction center, P700, was also inhibited at fully dry states. Photosynthetic electron flow in P. crispa reached a maximum at 20°C–25°C and at a light intensity of 700 μmol photons ? m?2 ? s?1. Interestingly, the osmolarity of P. crispa cells suggested that photosynthesis is performed using water absorbed in a liquid form rather than water absorbed from the air. Overall, these data suggest that these two species have acclimated to optimally photosynthesize under conditions of the highest light intensity and the highest temperature for their habitat in Antarctica.  相似文献   

19.
The mechanosensitive channel of small conductance (MscS) contributes to the survival of bacteria during osmotic downshock by transiently opening large diameter pores for the efflux of cellular contents before the membrane ruptures. Two crystal structures of the Escherichia coli MscS are currently available, the wild type protein in a nonconducting state at 3.7 Å resolution (Bass et al., Science 2002; 298:1582–1587) and the Ala106Val variant in an open state at 3.45 Å resolution (Wang et al., Science 2008; 321:1179–1183). Both structures used protein solubilized in the detergent fos‐choline‐14. We report here crystal structures of MscS from E. coli and Helicobacter pylori solubilized in the detergent β‐dodecylmaltoside at resolutions of 4.4 and 4.2 Å, respectively. While the cytoplasmic domains are unchanged in these structures, distinct conformations of the transmembrane domains are observed. Intriguingly, β‐dodecylmaltoside solubilized wild type E. coli MscS adopts the open state structure of A106V E. coli MscS, while H. pylori MscS resembles the nonconducting state structure observed for fos‐choline‐14 solubilized E. coli MscS. These results highlight the sensitivity of membrane protein conformational equilibria to variations in detergent, crystallization conditions, and protein sequence.  相似文献   

20.
The effects of heat-stress kinetics on the viability of Escherichia coli were investigated. Cells were exposed to heat-stress treatments extending from 30 to 50°C, with either a slope (40 min) or a shock (10 s), both followed by a 1-h plateau at 50°C in nutritive medium. A higher survival rate was observed after the slope than after the shock, when both were followed by a plateau, so the heat slope induced a certain degree of thermotolerance. This tolerance was partly (i) linked to de novo protein synthesis during the subsequent plateau phase, and (ii) abolished after rapid cooling from 50 to 30°C, which means that cellular components with rapidly reversible thermal properties are involved in this type of thermotolerance. The heat-slope-induced thermotolerance was chiefly linked to the maintenance of the plasma membrane integrity (preservation of structure, fluidity, and permeability), and not to GroEL or DnaK overexpression. Moreover, the high level of cell mortality induced by the heat shock could be related to changes in the membrane integrity.  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