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.
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   

2.
Prolonged stability is a desired property for the biotechnological application of enzymes since it allows its reutilization, contributing to making biocatalytic processes more economically competitive with respect to chemical synthesis. In this study, we have applied selection by folding interference at high temperature in Thermus thermophilus to obtain thermostable variants of the esterase I from Pseudomonas fluorescens (PFEI). The most thermostable variant (Q11L/A191S) showed a melting temperature (Tm) of 77.3 ± 0.1°C (4.6°C higher than the wild-type) and a half-life of over 13 hr at 65°C (7.9-fold better than the wild-type), with unchanged kinetic parameters. Stabilizing mutations Q11L and A191S were incorporated into PFEI variant L30P, previously described to be enantioselective in the hydrolysis of the (−)-enantiomer of the Vince lactam. The final variant Q11L/L30P/A191S showed a significant improvement in thermal stability (Tm of 80.8 ± 0.1°C and a half-life of 65 min at 75°C), while retaining enantioselectivity (E > 100). Structural studies revealed that A191S establishes a hydrogen bond network between a V-shaped hairpin and the α/β hydrolase domain that leads to higher rigidity and thus would contribute to explaining the increase in stability.  相似文献   

3.
Ca2+/calmodulin‐dependent protein kinase II (CaMKII) is a Ser/Thr kinase necessary for long‐term memory formation and other Ca2+‐dependent signaling cascades such as fertilization. Here, we investigated the stability of CaMKIIα using a combination of differential scanning calorimetry (DSC), X‐ray crystallography, and mass photometry (MP). The kinase domain has a low thermal stability (apparent Tm = 36°C), which is slightly stabilized by ATP/MgCl2 binding (apparent Tm = 40°C) and significantly stabilized by regulatory segment binding (apparent Tm = 60°C). We crystallized the kinase domain of CaMKII bound to p‐coumaric acid in the active site. This structure reveals solvent‐exposed hydrophobic residues in the substrate‐binding pocket, which are normally buried in the autoinhibited structure when the regulatory segment is present. This likely accounts for the large stabilization that we observe in DSC measurements comparing the kinase alone with the kinase plus regulatory segment. The hub domain alone is extremely stable (apparent Tm ~ 90°C), and the holoenzyme structure has multiple unfolding transitions ranging from ~60°C to 100°C. Using MP, we compared a CaMKIIα holoenzyme with different variable linker regions and determined that the dissociation of both these holoenzymes occurs at a higher concentration (is less stable) compared with the hub domain alone. We conclude that within the context of the holoenzyme structure, the kinase domain is stabilized, whereas the hub domain is destabilized. These data support a model where domains within the holoenzyme interact.  相似文献   

4.
5.
The Gp2 domain is a 45 amino-acid scaffold that has been evolved for specific, high-affinity binding towards multiple targets and was proven useful in molecular imaging and biological antagonism. It was hypothesized that Gp2 may benefit from increased hydrophilicity for improved physiological distribution as well as for physicochemical robustness. We identified seven exposed hydrophobic sites for hydrophilic mutations and experimentally evaluated single mutants, which yielded six mutations that do not substantially hinder expression, binding affinity or specificity (to epidermal growth factor receptor), and thermal stability. Eight combinations of these mutations improved hydrophilicity relative to the parental Gp2 clone as assessed by reverse-phase high-performance liquid chromatography (p < 0.05). Secondary structures and refolding abilities of the selected single mutants and all multimutants were unchanged relative to the parental ligand. A variant with five hydrophobic-to-hydrophilic mutations was identified with enhanced solubility as well as reasonable binding affinity ( K d = 53–63 nM), recombinant yield (1.3 ± 0.8 mg/L), and thermal stability ( T m = 53 ± 3°C). An alternative variant with a cluster of three leucine-to-hydrophilic mutations was identified with increased solubility, nominally increased binding affinity ( K d = 13–28 nM) and reasonable thermal stability ( T m = 54.0 ± 0.6°C) but reduced yield (0.4 ± 0.3 mg/L). In addition, a ≥7°C increase in the midpoint of thermal denaturation was observed in one of the single mutants (T21N). These mutants highlight the physicochemical tradeoffs associated with hydrophobic-to-hydrophilic mutation within a small protein, improve the solubility and hydrophilicity of an existent molecular imaging probe, and provide a more hydrophilic starting point for discovery of new Gp2 ligands towards additional targets.  相似文献   

6.
We examined the biophysical properties of cuticular lipids isolated from the housefly, Musca domestica. Melting temperatures (Tm) of surface lipids isolated from female houseflies decreased from 39.3 °C to 35.3 °C as the females attained sexual maturity and produced sex pheromone, whereas those prepared from males did not change with age. Lipids melted over a 10–25 °C temperature range, and their physical properties were a complex function of the properties of the component lipids. The Tm of total cuticular lipids was slightly below that of cuticular hydrocarbons (HC), the predominant lipid fraction. Hydrocarbons were further fractionated into saturated, unsaturated, and methyl-branched components. The order of decreasing Tm was total alkanes > total HCs > methyl-branched alkanes > alkenes. For 1-day-old flies, measured Tms of hydrocarbons were 1.3–5.5 °C lower than Tms calculated from a weighted average of Tms for saturated and unsaturated components. For 4-day-old flies, calculated Tms underestimated Tm by 11–14 °C. © 1995 Wiley-Liss, Inc.  相似文献   

7.
Two different types of approach were taken to improve the hydrolytic activity towards crystalline cellulose at elevated temperatures of Melanocarpus albomyces Cel7B (Ma Cel7B), a single-module GH-7 family cellobiohydrolase. Structure-guided protein engineering was used to introduce an additional tenth disulphide bridge to the Ma Cel7B catalytic module. In addition, a fusion protein was constructed by linking a cellulose-binding module (CBM) and a linker from the Trichoderma reesei Cel7A to the C terminus of Ma Cel7B. Both approaches proved successful. The disulphide bridge mutation G4C/M70C located near the N terminus, close to the entrance of the active site tunnel of Ma Cel7B, led to improved thermostability (ΔT m = 2.5°C). By adding the earlier found thermostability-increasing mutation S290T (ΔT m = 1.5°C) together with the disulphide bridge mutation, the unfolding temperature was increased by 4°C (mutant G4C/M70C/S290T) compared to that of the wild-type enzyme, thus showing an additive effect on thermostability. Both disulphide mutants had increased activity towards microcrystalline cellulose (Avicel) at 75°C, apparently solely because of their improved thermostability. The addition of a CBM also improved the thermostability (ΔT m = 2.5°C) and caused a clear (sevenfold) increase in the hydrolysis activity of Ma Cel7B towards Avicel at 70°C.  相似文献   

8.
Abstract The germination of Sorghum bicolor seeds of 9 genotypes was tested at temperatures between 8°C and 48°C on a thermal gradient plate. Samples were tested from three regions of the panicle expected to differ in temperature during grain filling. Seeds of a tenth genotype, SPV 354, produced in controlled-environment glasshouses at different panicle temperatures, were tested similarly. In addition, the emergence of SPV 354 was measured from planting depths of 2 and 5 cm at mean soil temperatures of 15, 20 and 25°C. Four methods of calculating mean germination rate for the nine genotypes were compared. Germination characters like base, optimum and maximum temperature (Tb, To, Tm), thermal time (θ)and the germination rate at To(Rmax showed only small differences between methods. There was a range of genotypic variation in all characters: Tb 8.5–11.9°C; To, 33.2–37.5°C; Tm, 46.8–49.2°C; θ, 23.4–38.0°Cd; Rmax, 0.69–1.14-d-1. In contrast, mean germinability (G) was between 90% and 100% over the temperature range 13–40°C. Panicle temperature had no effect on any germination character in SPV 354. However, deeper burial increased θ for emergence and decreased G, irrespective of soil temperature except at 5 cm. Increasing panicle temperature, by reducing seed size, reduced G and increased θ by about 10% only at 15°C and 5 cm depth.  相似文献   

9.
Principal component analysis (PCA) of published DNA-relatedness data showed the usefulness of this method in displaying relationships among closely related bacteria. Very similar ordinations were obtained when relative binding ratios (RBR) at 60°C or 75°C or ΔT m values were used to form the data matrix. A curvilinear relationship and a (quasi) linear relationship were found, respectively, between 75°C and 60°C RBR and ΔT m and 60°C RBR. These statistical relationships explain the similarity of PCA results using either measurement (60°C RBR, 75°C RBR, or ΔT m). Use of PCA is suggested to delineate groups within a complex set of DNA-relatedness data. The level of ΔT m within groups and between groups should help decide whether these groups are genospecies.  相似文献   

10.
An overview is presented of some of the major insights that have come from studies of the structure, stability, and folding of T4 phage lysozyme. A major purpose of this review is to provide the reader with a complete tabulation of all of the variants that have been characterized, including melting temperatures, crystallographic data, Protein Data Bank access codes, and references to the original literature. The greatest increase in melting temperature (Tm) for any point mutant is 5.1°C for the mutant Ser 117 → Val. This is achieved in part not only by hydrophobic stabilization but also by eliminating an unusually short hydrogen bond of 2.48 Å that apparently has an unfavorable van der Waals contact. Increases in Tm of more than 3–4°C for point mutants are rare, whereas several different types of destabilizing substitutions decrease Tm by 20°C or thereabouts. The energetic cost of cavity creation and its relation to the hydrophobic effect, derived from early studies of “large‐to‐small” mutants in the core of T4 lysozyme, has recently been strongly supported by related studies of the intrinsic membrane protein bacteriorhodopsin. The L99A cavity in the C‐terminal domain of the protein, which readily binds benzene and many other ligands, has been the subject of extensive study. Crystallographic evidence, together with recent NMR analysis, suggest that these ligands are admitted by a conformational change involving Helix F and its neighbors. A total of 43 nonisomorphous crystal forms of different monomeric lysozyme mutants were obtained plus three more for synthetically‐engineered dimers. Among the 43 space groups, P212121 and P21 were observed most frequently, consistent with the prediction of Wukovitz and Yeates.  相似文献   

11.
L H Chang  A G Marshall 《Biopolymers》1986,25(7):1299-1313
The unfolding of B. subtilis 5S RNA is examined by direct calorimetric measurement in the presence of various concentrations of Na+ and Mg2+. The composite differential scanning calorimetry (DSC) curve is analyzed into 3–5 individual two-state melting transitions. In the absence of added Na+ or Mg2+, the 5S RNA segments melt together at Tm = 40°C. Addition of Na+ stabilizes the molecular structure (Tm = 56°C) and widens the melting temperature range, so that up to five component transitions are observed. Addition of Mg2+ alone produces a very stable structure (Tm = 75°C) with highly cooperative melting. Finally, addition of both Na+ and Mg2+ produces the highest stability (Tm = 76°C). The results are interpreted according to hypothetical secondary and tertiary base-pairing schemes. The conformational changes demonstrated here may facilitate the movement of the protein synthesis machinery during RNA translation.  相似文献   

12.
The collagen triple helix has a larger accessible surface area per molecular mass than globular proteins, and therefore potentially more water interaction sites. The effect of deuterium oxide on the stability of collagen model peptides and Type I collagen molecules was analyzed by circular dichroism and differential scanning calorimetry. The transition temperatures (Tm) of the protonated peptide (Pro‐Pro‐Gly)10 were 25.4 and 28.7°C in H2O and D2O, respectively. The increase of the Tm of (Pro‐Pro‐Gly)10 measured calorimetrically at 1.0°C min?1 in a low pH solution from the protonated to the deuterated solvent was 5.1°C. The increases of the Tm for (Gly‐Pro‐4(R)Hyp)9 and pepsin‐extracted Type I collagen were measured as 4.2 and 2.2°C, respectively. These results indicated that the increase in the Tm in the presence of D2O is comparable to that of globular proteins, and much less than reported previously for collagen model peptides [Gough and Bhatnagar, J Biomol Struct Dyn 1999, 17, 481–491]. These experimental results suggest that the interaction of water molecules with collagen is similar to the interaction of water with globular proteins, when the ratio of collagen to water is very small and collagen is monomerically dispersed in the solvent. © 2009 Wiley Periodicals, Inc. Biopolymers 93: 93–101, 2010. This article was originally published online as an accepted preprint. The “Published Online” date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com  相似文献   

13.
Glutathione S-transferases (GSTs) are multifunctional enzymes and play an important role in cellular detoxification. Besides this, GSTs act as cytosolic carrier proteins that bind hydrophobic compounds such as heme, bilirubin, steroids, and polycyclic hydrocarbons. GST has great importance in biotechnology, as it is a target for vaccine and drug development and biosensors development for xenobiotics. Moreover, the GST tag has been extensively used for protein expression and purification. Until now, biophysical properties of camel liver GST have not been characterized. In the present study we have purified camel (Camelus dromedarius) liver GST to homogeneity in a single step by affinity chromatography with 23.4-fold purification and 60.6% yield. Our results showed that maximal activity of GST was at pH 6.5 and it was stable in the pH range of 5 to 10. The optimum temperature was 55°C and the Tm was 57°C. The chemical chaperone glycerol (3.3 M) was able to protect GST activity and aggregation against thermal denaturation by stabilizing the protein structure at 50 and 57°C, respectively. However, L-arginine (125 mM) did not protect GST against thermal stress. Far-ultraviolet circular dichroism (CD) spectra showed that glycerol protected the secondary structure of GST while L-arginine induced conformational changes under thermal stress. In conclusion, our studies on the GST stability suggest that glycerol works as a stabilizer and L-arginine acts as a destabilizer.  相似文献   

14.
15.
Abstract

Thermodynamic parameters of melting process (δHm, Tm, δTm) of calf thymus DNA, poly(dA)poly(dT) and poly(d(A-C))·poly(d(G-T)) were determined in the presence of various concentrations of TOEPyP(4) and its Zn complex. The investigated porphyrins caused serious stabilization of calf thymus DNA and poly poly(dA)poly(dT), but not poly(d(A-C))poly(d(G-T)). It was shown that TOEpyp(4) revealed GC specificity, it increased Tm of satellite fraction by 24°C, but ZnTOEpyp(4), on the contrary, predominately bound with AT-rich sites and increased DNA main stage Tm by 18°C, and Tm of poly(dA)poly(dT) increased by 40 °C, in comparison with the same polymers without porphyrin. ZnTOEpyp(4) binds with DNA and poly(dA)poly(dT) in two modes—strong and weak ones. In the range of r from 0.005 to 0.08 both modes were fulfilled, and in the range of r from 0.165 to 0.25 only one mode—strong binding—took place. The weak binding is characterized with shifting of Tm by some grades, and for the strong binding Tm shifts by ~ 30–40°C. Invariability of ΔHm of DNA and poly(dA)poly(dT), and sharp increase of Tm in the range of r from 0.08 to 0.25 for thymus DNA and 0.01–0.2 for poly(dA)poly(dT) we interpret as entropic character of these complexes melting. It was suggested that this entropic character of melting is connected with forcing out of H2O molecules from AT sites by ZnTOEpyp(4) and with formation of outside stacking at the sites of binding. Four-fold decrease of calf thymus DNA melting range width ΔTm caused by increase of added ZnTO- Epyp(4) concentration is explained by rapprochement of AT and GC pairs thermal stability, and it is in agreement with a well-known dependence, according to which ΔT~TGC-TAT for DNA obtained from higher organisms (L. V. Berestetskaya, M. D. Frank-Kamenetskii, and Yu. S. Lazurkin. Biopolymers 13, 193–205 (1974)). Poly (d(A-C))poly(d(G-T)) in the presence of ZnTOEpyp(4) gives only one mode of weak binding. The conclusion is that binding of ZnTOEpyp(4) with DNA depends on its nucleotide sequence.  相似文献   

16.
The present study investigated whether increasing morning rectal temperatures (Trec) to evening levels, or increasing morning and evening Trec to an “optimal” level (38.5°C), resulting in increased muscle temperatures (Tm), would offset diurnal variation in repeated sprint (RS) performance in a causal manner. Twelve trained males underwent five sessions [age (mean ± SD) 21.0 ± 2.3 years, maximal oxygen consumption (V?O2max) 60.0 ± 4.4 mL.kg–1 min–1, height 1.79 ± 0.06 m, body mass 78.2 ± 11.8 kg]. These included control morning (M, 07:30 h) and evening (E, 17:30 h) sessions (5-min warm-up), and three further sessions consisting of a warm-up morning trial (ME, in 39–40°C water) until Trec reached evening levels; two “optimal” trials in the morning and evening (M38.5 and E38.5, in 39–40°C water) respectively, until Trec reached 38.5°C. All sessions included 3 × 3-s task-specific warm-up sprints, thereafter 10 × 3-s RS with 30-s recoveries were performed a non-motorised treadmill. Trec and Tm measurements were taken at the start of the protocol and following the warm-up periods. Values for Trec and Tm at rest were higher in the evening compared to morning values (0.48°C and 0.69°C, p < 0.0005). RS performance was lower (7.8–8.3%) in the M for distance covered (DC; p = 0.002), average power (AP; p = 0.029) and average velocity (AV; p = 0.002). Increasing Trec in the morning to evening values or optimal values (38.5°C) did not increase RS performance to evening levels (p = 1.000). However, increasing Trec in the evening to “optimal” level through a passive warm-up significantly reduced DC (p = 0.008), AP (p < 0.0005) and AV (p = 0.007) to values found in the M condition (6.0–6.9%). Diurnal variation in Trec and Tm is not wholly accountable for time-of-day oscillations in RS performance on a non-motorised treadmill; the exact mechanism(s) for a causal link between central temperature and human performance are still unclear and require more research.  相似文献   

17.
Although the dot-blot-SNP technique is a simple cost-saving technique suitable for genotyping of many plant individuals, optimization of hybridization and washing conditions for each SNP marker requires much time and labor. For prediction of the optimum hybridization conditions for each probe, we compared T m values estimated from nucleotide sequences using the DINAMelt web server, measured T m values, and hybridization conditions yielding allele-specific signals. The estimated T m values were comparable to the measured T m values with small differences of less than 3°C for most of the probes. There were differences of approximately 14°C between the specific signal detection conditions and estimated T m values. Change of one level of SSC concentrations of 0.1, 0.2, 0.5, and 1.0× SSC corresponded to a difference of approximately 5°C in optimum signal detection temperature. Increasing the sensitivity of signal detection by shortening the exposure time to X-ray film changed the optimum hybridization condition for specific signal detection. Addition of competitive oligonucleotides to the hybridization mixture increased the suitable hybridization conditions by 1.8. Based on these results, optimum hybridization conditions for newly produced dot-blot-SNP markers will become predictable.  相似文献   

18.
Contributed equally to this work. To further understand the origin of the double thermal transitions of collagen in acidic solution induced by heating, the denaturation of acidic soluble collagen was investigated by micro-differential scanning calorimeter (micro-DSC), circular dichroism (CD), dynamic laser light scattering (DLLS), transmission electron microscopy (TEM), and two-dimensional (2D) synchronous fluorescence spectrum. Micro-DSC experiments revealed that the collagen exhibited double thermal transitions, which were located within 31–37?°C (minor thermal transition, T s?~?33?°C) and 37–55?°C (major thermal transition, T m?~?40?°C), respectively. The CD spectra suggested that the thermal denaturation of collagen resulted in transition from polyproline II type structure to unordered structure. The DLLS results showed that there were mainly two kinds of collagen fibrillar aggregates with different sizes in acidic solution and the larger fibrillar aggregates (T p2?=?40?°C) had better heat resistance than the smaller one (T p1?=?33?°C). TEM revealed that the depolymerization of collagen fibrils occurred and the periodic cross-striations of collagen gradually disappeared with increasing temperature. The 2D fluorescence correlation spectra were also applied to investigate the thermal responses of tyrosine and phenylalanine residues at the molecular level. Finally, we could draw the conclusion that (1) the minor thermal transition was mainly due to the defibrillation of the smaller collagen fibrillar aggregates and the unfolding of a little part of triple helices; (2) the major thermal transition primarily arose from the defibrillation of the larger collagen fibrillar aggregates and the complete denaturation of the majority part of triple helices.  相似文献   

19.
We have used differential scanning calorimetry (DSC) to determine the unfolding properties of commercial products of human serum albumin (HSA) prepared from pooled human blood, transgenic yeast, and transgenic rice. The initial melting temperatures (Tm1) for the unfolding transitions of the HSA products varied from 62°C to 75°C. We characterized the samples for purity, fatty acid content, and molecular weight. The effects of adding fatty acids, heat pasteurization, and a low pH defatting technique on the transition temperatures were measured. Defatted HSA has a structure with the lowest stability (Tm of ~62°C). When fatty acids are bound to HSA, the structure is stabilized (Tm of ~64–72°C), and prolonged heating (pasteurization at 60°C) results in a heat‐stabilized structural form containing fatty acids (Tm of ~75–80°C). This process was shown to be reversible by a low pH defatting step. This study shows that the fatty acid composition and bioprocessing history of the HSA commercial products results in the large differences in the thermal stability. © 2014 American Institute of Chemical Engineers Biotechnol. Prog., 31:62–69, 2015  相似文献   

20.
The advent of metagenomics has greatly facilitated the discovery of enzymes with useful biochemical characteristics for industrial and biomedical applications, from environmental niches. In this study, we used sequence‐based metagenomics to identify two antibiotic resistance enzymes from the secluded, lower convective layer of Atlantis II Deep Red Sea brine pool (68°C, ~2200 m depth and 250‰ salinity). We assembled > 4 000 000 metagenomic reads, producing 43 555 contigs. Open reading frames (ORFs) called from these contigs were aligned to polypeptides from the Comprehensive Antibiotic Resistance Database using BLASTX. Two ORFs were selected for further analysis. The ORFs putatively coded for 3′‐aminoglycoside phosphotransferase [APH(3′)] and a class A beta‐lactamase (ABL). Both genes were cloned, expressed and characterized for activity and thermal stability. Both enzymes were active in vitro, while only APH(3′) was active in vivo. Interestingly, APH(3′) proved to be thermostable (Tm = 61.7°C and ~40% residual activity after 30 min of incubation at 65°C). On the other hand, ABL was not as thermostable, with a Tm = 43.3°C. In conclusion, we have discovered two novel AR enzymes with potential application as thermophilic selection markers.  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