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 125 毫秒
1.
Insight into the hyperthermostable endo-β-1,3-glucanase pfLamA from Pyrococcus furiosus is obtained by using NMR spectroscopy. pfLamA functions optimally at 104 °C and recently the X-ray structure of pfLamA has been obtained at 20 °C, a temperature at which the enzyme is inactive. In this study, near-complete (>99%) NMR assignments are presented of chemical shifts of pfLamA in presence and absence of calcium at 62 °C, a temperature at which the enzyme is biologically active. The protein contains calcium and the effects of calcium on the protein are assessed. Calcium binding results in relatively small chemical shift changes in a region distant from the active site of pfLamA and thus causes only minor conformational modifications. Removal of calcium does not significantly alter the denaturation temperature of pfLamA, implying that calcium does not stabilize the enzyme against global unfolding. The data obtained form the basis for elucidation of the molecular origins involved in conformational stability and biological activity of hyperthermophilic endo-β-1,3-glucanases at extreme temperatures.  相似文献   

2.
Sulfolobus solfataricus 5′-deoxy-5′-melthylthioadenosine phosphorylase II (SsMTAPII), is a hyperthermophilic hexameric protein with two intrasubunit disulfide bonds (C138–C205 and C200–C262) and a CXC motif (C259–C261). To get information on the role played by these covalent links in stability and folding, the conformational stability of SsMTAPII and C262S and C259S/C261S mutants was studied by thermal and guanidinium chloride (GdmCl)-induced unfolding and analyzed by fluorescence spectroscopy, circular dichroism, and SDS-PAGE. No thermal unfolding transition of SsMTAPII can be obtained under nonreducing conditions, while in the presence of the reducing agent Tris-(2-carboxyethyl) phosphine (TCEP), a Tm of 100 °C can be measured demonstrating the involvement of disulfide bridges in enzyme thermostability. Different from the wild-type, C262S and C259S/C261S show complete thermal denaturation curves with sigmoidal transitions centered at 102 °C and 99 °C respectively. Under reducing conditions these values decrease by 4 °C and 8 °C respectively, highlighting the important role exerted by the CXC disulfide on enzyme thermostability. The contribution of disulfide bonds to the conformational stability of SsMTAPII was further assessed by GdmCl-induced unfolding experiments carried out under reducing and nonreducing conditions. Thermal unfolding was found to be reversible if the protein was heated in the presence of TCEP up to 90 °C but irreversible above this temperature because of aggregation. In analogy, only chemical unfolding carried out in the presence of reducing agents resulted in a reversible process suggesting that disulfide bonds play a role in enzyme denaturation. Thermal and chemical unfolding of SsMTAPII occur with dissociation of the native hexameric state into denatured monomers, as indicated by SDS-PAGE.  相似文献   

3.
Cold denaturation is a phenomenon seen in many different proteins. However, there have been no reports so far of its occurrence in hyperthermophile proteins. Here, using a recombinant triosephosphate isomerase (PfuTIM) from the hyperthermophile archaeon, Pyrococcus furiosus, we show that the heating of this protein through the low temperature side of its thermal unfolding transition in the presence of guanidinium hydrochloride (GdmCl) results in the formation of partially-disordered conformational ensembles that retain considerable native-like secondary and tertiary structure. Unlike PfuTIM itself, these thermochemically obtained partially-disordered PfuTIM ensembles display cold denaturation as they are cooled to room temperature. The protein thus shows hysteresis, adopting different structural states in a manner dependent upon the nature of the heating and cooling treatment, rather than upon the initial and final conditions of temperature and GdmCl concentration, indicating that some sort of a kinetic effect influences structure adoption and retention. The structure lost through cooling of partially-disordered PfuTIM is found to be regained through heating. The ability of GdmCl to thus apparently destabilize the highly thermodynamically and kinetically stable structure of PfuTIM (sufficiently, to cause it to display observable cold-denaturation and heat-renaturation transitions, in real-time, with cooling and heating) offers support to current ideas concerning the how hyperthermophile proteins achieve their high kinetic stabilities, and suggests that desolvation-solvation barriers may be responsible for high kinetic stability.  相似文献   

4.
d-Arabinose isomerase (d-AI), also known as l-fucose isomerase (l-FI), catalyzes the aldose–ketose isomerization of d-arabinose to d-ribulose, and l-fucose to l-fuculose. Bacillus pallidus (B. pallidus) d-AI can catalyze isomerization of d-altrose to d-psicose, as well as d-arabinose and l-fucose. Three X-ray structures of B. pallidusd-AI in complexes with 2-methyl-2,4-pentadiol, glycerol and an inhibitor, l-fucitol, were determined at resolutions of 1.77, 1.60 and 2.60 Å, respectively. B. pallidusd-AI forms a homo-hexamer, and one subunit has three domains of almost equal size; two Rossmann fold domains and a mimic of the (β/α) barrel fold domain. A catalytic metal ion (Mn2+) was found in the active site coordinated by Glu342, Asp366 and His532, and an additional metal ion was found at the channel for the passage of a substrate coordinated by Asp453. The X-ray structures basically supported the ene-diol mechanism for the aldose–ketose isomerization by B. pallidusd-AI, as well as Escherichia coli (E. coli) l-FI, in which Glu342 and Asp366 facing each other at the catalytic metal ion transfer a proton from C2 to C1 and O1 to O2, acting as acid/base catalysts, respectively. However, considering the ionized state of Asp366, the catalytic reaction also possibly occurs through the negatively charged ene-diolate intermediate stabilized by the catalytic metal ion. A structural comparison with E. colil-FI showed that B. pallidusd-AI possibly interconverts between “open” and “closed” forms, and that the additional metal ion found in B. pallidusd-AI may help to stabilize the channel region.  相似文献   

5.
The hyperthermophilic archaeon Pyrococcus furiosus was grown on pyruvate as carbon and energy source. The enzymes involved in gluconeogenesis were investigated. The following findings indicate that glucose-6-phosphate formation from pyruvate involves phosphoenolpyruvate synthetase, enzymes of the Embden-Meyerhof pathway and fructose-1,6-bisphosphate phosphatase.Cell extracts of pyruvate-grown P.furiosus contained the following enzyme activities: phosphoenolpyruvate synthetase (0.025 U/mg, 50 °C), enolase (0.9 U/mg, 80 °C), phosphoglycerate mutase (0.13 U/mg, 55 °C), phosphoglycerate kinase (0.01 U/mg, 50 °C), glyceraldehyde-3-phosphate dehydrogenase reducing either NADP+ or NAD+ (NADP+: 0.019 U/mg, NAD+: 0.009 U/mg; 50 °C), triosephosphate isomerase (1.4 U/mg, 50 °C), fructose-1,6-bisphosphate aldolase (0.0045 U/mg, 55 °C), fructose-1,6-bisphosphate phosphatase (0.026 U/mg, 75 °C), and glucose-6-phosphate isomerase (0.22 U/mg, 50 °C). Kinetic properties (V max values and apparent K m values) of the enzymes indicate that they operate in the direction of sugar synthesis. The specific enzyme activities of phosphoglycerate kinase, glyceraldehyde-3-phosphate dehydrogenase (NADP+-reducing) and fructose-1,6-bisphosphate phosphatase in pyruvate-grown P. furiosus were by a factor of 3, 10 and 4, respectively, higher as compared to maltose-grown cells suggesting that these enzymes are induced under conditions of gluconeogenesis. Furthermore, cell extracts contained ferredoxin: NADP+ oxidoreductase (0.023 U/mg, 60 °C); phosphoenolpyruvate carboxylase (0.018 U/mg, 50 °C) acts as an anaplerotic enzyme.Thus, in P. furiosus sugar formation from pyruvate involves reactions of the Embden-Meyerhof pathway, whereas sugar degradation to pyruvate proceeds via a modified non-phosphorylated Entner-Doudoroff pathway.  相似文献   

6.
Several molecular methods have been employed for Borrelia species identification. Newly developed technology, real-time polymerase chain reaction (RT-PCR), combines simultaneous amplification, detection and differentiation of strains in one PCR run. The aim of the study was to perform and evaluate RT-PCR for Borreliaburgdorferi sensu lato species identification. Borrelia species identification was accomplished on 374 Borrelia strains using two approaches: 1.) MluI restriction of entire borrelial chromosome (MluI-large restriction fragment patterns, LRFP), and 2.) RT-PCR targeting hbb gene and specific melting temperature (Tm) detection. The results of the two molecular methods were compared. With MluI-RFLP we were able to differentiate all Borrelia species and their subtypes within particular species. RT-PCR based on Tm determination identified unique strains within the species Borreliaafzelii (Tm 66.11 °C), B. burgdorferi sensu stricto (Tm 68.18 °C), Borreliaspielmanii (Tm 59.45 °C) and Borreliavalaisiana (Tm 59.62 °C). We were not able to distinguish the last two species that shared almost identical Tm. The large majority of Borreliagarinii strains shared Tm 51.42 °C, while subtype Mlg4 was characterized by Tm 56.87 °C. Strains of Borrelialusitaniae species also were heterogeneous; human isolate had Tm 63.47 °C while two tick isolates shared Tm 61.77 °C. Differences inside hbb gene enabled differentiation of the majority of Borrelia species, and revealed two clusters within B. garinii and B. lusitaniae species, respectively, but it was not possible to distinguish B. spielmanii form B. valaisiana. The major advantage of RT-PCR was that it was easy to perform and that the results were obtained within a few hours.  相似文献   

7.
Three new electrophoretic variants of human erythrocyte triosephosphate isomerase (TPI) have been partially purified and compared with the normal isozyme with respect to stability, kinetics, and immunological properties. TPI 2HR1, an anodally migrating variant, was less stable than normal in guanidine denaturation and thermodenaturation tests, although it exhibited normal kinetic properties. The level of enzyme activity in erythrocytes from the proband with the phenotype TPI 1-2HR1 was about 60% of the normal mean. The variant allozyme TPI 2NG1, an anodally migrating allozyme associated with normal activity, was very thermolabile at 55 and 57°C. It was also much more labile than normal in stability tests in buffers at pH 5 and pH 10, although it exhibited normal kinetic and immunological properties. TPI 4NG1, a cathodally migrating variant associated with normal activity and normal kinetic as well as immunological properties, was more stable than normal in pH 5 buffer. Family studies demonstrated that the unique characteristics of these variants are genetically transmitted. In two-dimensional electrophoresis of purified isozymes the variant subunits were separated from the normal in the pI axis. However, there is no difference between the variants and the normal in the molecular weight axis, suggesting that the variants result from single amino acid substitutions.  相似文献   

8.
To understand the pressure-adaptation mechanism of deep-sea enzymes, we studied the effects of pressure on the enzyme activity and structural stability of dihydrofolate reductase (DHFR) of the deep-sea bacterium Moritella profunda (mpDHFR) in comparison with those of Escherichia coli (ecDHFR). mpDHFR exhibited optimal enzyme activity at 50 MPa whereas ecDHFR was monotonically inactivated by pressure, suggesting inherent pressure-adaptation mechanisms in mpDHFR. The secondary structure of apo-mpDHFR was stable up to 80 °C, as revealed by circular dichroism spectra. The free energy changes due to pressure and urea unfolding of apo-mpDHFR, determined by fluorescence spectroscopy, were smaller than those of ecDHFR, indicating the unstable structure of mpDHFR against pressure and urea despite the three-dimensional crystal structures of both DHFRs being almost the same. The respective volume changes due to pressure and urea unfolding were − 45 and − 53 ml/mol at 25 °C for mpDHFR, which were smaller (less negative) than the corresponding values of − 77 and − 85 ml/mol for ecDHFR. These volume changes can be ascribed to the difference in internal cavity and surface hydration of each DHFR. From these results, we assume that the native structure of mpDHFR is loosely packed and highly hydrated compared with that of ecDHFR in solution.  相似文献   

9.
Evaluation of the cytotoxicity of an ethanolic root extract of Sideroxylonfoetidissimum subsp. gaumeri (Sapotaceae) revealed activity against the murine macrophage-like cell line RAW 264.7. Systematic bioassay-guided fractionation of this extract gave an active saponin-containing fraction from which four saponins were isolated. Use of 1D (1H, 13C, DEPT135) and 2D (COSY, TOCSY, HSQC, and HMBC) NMR, mass spectrometry and sugar analysis gave their structures as 3-O-(β-d-glucopyranosyl-(1 → 6)-β-d-glucopyranosyl)-28-O-(α-l-rhamnopyranosyl-(1 → 3)[β-d-xylopyranosyl-(1 → 4)]-β-d-xylopyranosyl-(1 → 4)-α-l-rhamnopyranosyl-(1 → 2)-α-l-arabinopyranosyl)-16α-hydroxyprotobassic acid, 3-O-β-d-glucopyranosyl-28-O-(α-l-rhamnopyranosyl-(1 → 3)[β-d-xylopyranosyl-(1 → 4)]-β-d-xylopyranosyl-(1 → 4)-α-l-rhamnopyranosyl-(1 → 2)-α-l-arabinopyranosyl)-16α-hydroxyprotobassic acid, 3-O-(β-d-glucopyranosyl-(1 → 6)-β-d-glucopyranosyl)-28-O-(α-l-rhamnopyranosyl-(1 → 3)-β-d-xylopyranosyl-(1 → 4)[β-d-apiofuranosyl-(1 → 3)]-α-l-rhamnopyranosyl-(1 → 2)-α-l-arabinopyranosyl)-16α-hydroxyprotobassic acid, and the known compound, 3-O-β-d-glucopyranosyl-28-O-(α-l-rhamnopyranosyl-(1 → 3)[β-d-xylopyranosyl-(1 → 4)]-β-d-xylopyranosyl-(1 → 4)-α-l-rhamnopyranosyl-(1 → 2)-α-l-arabinopyranosyl)-protobassic acid. Two further saponins were obtained from the same fraction, but as a 5:4 mixture comprising 3-O-(β-d-glucopyranosyl)-28-O-(α-l-rhamnopyranosyl-(1 → 3)-β-d-xylopyranosyl-(1 → 4)[β-d-apiofuranosyl-(1 → 3)]-α-l-rhamnopyranosyl-(1 → 2)-α-l-arabinopyranosyl)-16α-hydroxyprotobassic acid and 3-O-(β-d-apiofuranosyl-(1 → 3)-β-d-glucopyranosyl)-28-O-(α-l-rhamnopyranosyl-(1 → 3)[β-d-xylopyranosyl-(1 → 4)]-β-d-xylopyranosyl-(1 → 4)-α-l-rhamnopyranosyl-(1 → 2)-α-l-arabinopyranosyl)-16α-hydroxyprotobassic acid, respectively. This showed greater cytotoxicity (IC50 = 11.9 ± 1.5 μg/ml) towards RAW 264.7 cells than the original extract (IC50 = 39.5 ± 4.1 μg/ml), and the saponin-containing fraction derived from it (IC50 = 33.7 ± 6.2 μg/ml).  相似文献   

10.
Two new electrophoretic variants of human triosephosphate isomerase (TPI) have been partially purified and characterized. The TPI Manchester variant, a cathodally migrating electrophoretic allozyme identified in an individual with the phenotype TPI 1-Manchester, is associated with a normal level of enzyme activity in erythrocytes and normal kinetic properties. It is very thermolabile at 55 and 57° C, although it is not uniquely sensitive to either guanidine-HCl or urea denaturation. The TPI Hiroshima-2 variant is an anodally migrating allozyme (the phenotype of proband is TPI 1-Hiroshima-2) with normal activity and kinetic properties and also normal stability characteristics. It is inactivated less by antisera raised against normal human TPI than either the normal or the Manchester allozyme. Dissociation-reassociation experiments utilizing these allozymes have confirmed that normal human red blood cell TPI isozymes are produced by a sequence of reactions (presumably deamidations) involving alternating subunits.Financial support was derived from Contract EY-77-C-02-2828 from the Department of Energy.  相似文献   

11.
The stabilities of myoglobin, apo-myoglobin, and of two myoglobins with chlorophyllous chromophores (Zn-pheophorbide a and Zn-bacteriopheophorbide a), have been studied by thermal and chemical denaturation. With guanidinium chloride, the stability order is myoglobin > Zn-pheophorbide-myoglobin > Zn-bacteriopheophorbide-myoglobin ∼ apo-myoglobin. The thermal behavior is more complex. The transition temperature of thermal unfolding of the apoprotein (62.4 °C) is increased by Zn-pheophorbide a (83.9 °C) and Zn-bacteriopheophorbide a (82.6 °C) to a similar degree as by the native chromophore, heme (83.5 °C). The recovery with Zn-pheophorbide (92-98%) is even higher than with heme (74-76%), while with Zn-bacteriopheophorbide (40%) it is as low as with the apoprotein (42%). Recovery also depends on the rates of heating, and in particular the time spent at high temperatures. It is concluded that irreversibility of unfolding is related to loss of the chromophores, which are required for proper re-folding.  相似文献   

12.
A dispersion of melting temperatures at pH 5.3 for individual residues of the BBL protein domain has been adduced as evidence for barrier-free downhill folding. Other members of the peripheral subunit domain family fold cooperatively at pH 7. To search for possible causes of anomalies in BBL's denaturation behavior, we measured the pH titration of individual residues by heteronuclear NMR. At 298 K, the pKa of His142 was close to that of free histidine at 6.47 ± 0.04, while that of the more buried His166 was highly perturbed at 5.39 ± 0.02. Protonation of His166 is thus energetically unfavorable and destabilizes the protein by ∼ 1.5 kcal/mol. Changes in Cα secondary shifts at pH 5.3 showed a decrease in helicity of the C-terminus of helix 2, where His166 is located, which was accompanied by a measured decrease of 1.1 ± 0.2 kcal/mol in stability from pH 7 to 5.3. Protonation of His166 perturbs, therefore, the structure of BBL. Only ∼ 1% of the structurally perturbed state will be present at the biologically relevant pH 7.6. Experiments at pH 5.3 report on a near-equal mixture of the two different native states. Further, at this pH, small changes of pH and pKa induced by changes in temperature will have near-maximal effects on pH-dependent conformational equilibria and on propagation of experimental error. Accordingly, conventional barrier-limited folding predicts some dispersion of measured thermal unfolding curves of individual residues at pH 5.3.  相似文献   

13.
We report the biochemical and structural characterization of the purine-specific ribonucleoside hydrolase from the archaeon Sulfolobus solfataricus (SsIAG-NH). SsIAG-NH is a homodimer of 70 kDa specific for adenosine, guanosine and inosine. SsIAG-NH is highly thermophilic and is characterized by extreme thermodynamic stability (Tm, 107 °C), kinetic stability and remarkable resistance to guanidinium chloride-induced unfolding. A disulfide bond that, on the basis of SDS-PAGE is positioned intersubunits, plays an important role in thermal stability. SsIAG-NH shares 43% sequence identity with the homologous pyrimidine-specific nucleoside hydrolase from S. solfataricus (SsCU-NH). The comparative sequence alignment of SsIAG-NH, SsCU-NH, purine non-specific nucleoside hydrolase from Crithidia fasciculata and purine-specific nucleoside hydrolase from Trypanosoma vivax shows that, only few changes in the base pocket are responsible for different substrate specificity of two S. solfataricus enzymes. The structure of SsIAG-NH predicted by homology modeling allows us to infer the role of specific residues in substrate specificity and thermostability.  相似文献   

14.
The (beta/alpha)(8) barrel is the most commonly occurring fold among enzymes. A key step towards rationally engineering (beta/alpha)(8) barrel proteins is to understand their underlying structural organization and folding energetics. Using misincorporation proton-alkyl exchange (MPAX), a new tool for solution structural studies of large proteins, we have performed a native-state exchange analysis of the prototypical (beta/alpha)(8) barrel triosephosphate isomerase. Three cooperatively unfolding subdomains within the structure are identified, as well as two partially unfolded forms of the protein. The C-terminal domain coincides with domains reported to exist in four other (beta/alpha)(8) barrels, but the two N-terminal domains have not been observed previously. These partially unfolded forms may represent sequential intermediates on the folding pathway of triosephosphate isomerase. The methods reported here should be applicable to a variety of other biological problems involving protein conformational changes.  相似文献   

15.
The effect of ions on the thermostability and unfolding of Na,K-ATPase from shark salt gland was studied and compared with that of Na,K-ATPase from pig kidney by using differential scanning calorimetry (DSC) and activity assays. In 1 mM histidine at pH 7, the shark enzyme inactivates rapidly at 20 °C, as does the kidney enzyme at 42 °C (but not at 20 °C). Increasing ionic strength by addition of 20 mM histidine, or of 1 mM NaCl or KCl, protects both enzymes against this rapid inactivation. As detected by DSC, the shark enzyme undergoes thermal unfolding at lower temperature (Tm ≈ 45 °C) than does the kidney enzyme (Tm ≈ 55 °C). Both calorimetric endotherms indicate multi-step unfolding, probably associated with different cooperative domains. Whereas the overall heat of unfolding is similar for the kidney enzyme in either 1 mM or 20 mM histidine, components with high mid-point temperatures are lost from the unfolding transition of the shark enzyme in 1 mM histidine, relative to that in 20 mM histidine. This is attributed to partial unfolding of the enzyme due to a high hydrostatic pressure during centrifugation of DSC samples at low ionic strength, which correlates with inactivation measurements. Addition of 10 mM NaCl to shark enzyme in 1 mM histidine protects against inactivation during centrifugation of the DSC sample, but incubation for 1 h at 20 °C prior to addition of NaCl results in loss of components with lower mid-point temperatures within the unfolding transition. Cations at millimolar concentration therefore afford at least two distinct modes of stabilization, likely affecting separate cooperative domains. The different thermal stabilities and denaturation temperatures of the two Na,K-ATPases correlate with the respective physiological temperatures, and may be attributed to the different lipid environments.  相似文献   

16.
The objective of this study was to verify the effect of different freezing curves, straw sizes, and thawing rates on the cryopreservation of collared peccary semen. Twelve ejaculates were obtained from captive adult males by electroejaculation, and evaluated for sperm motility, kinetic rating, viability, morphology, and functional membrane integrity. The ejaculates were diluted in a coconut water extender (ACP-116c) with egg yolk and glycerol, packaged into 0.25 mL or 0.50 mL plastic straws and cryopreserved in liquid nitrogen following a slow (−10 °C/min) or a fast (−40 °C/min) freezing curve. After one week, samples were thawed at 37 °C/1 min or 70 °C/8 s and evaluated as reported for fresh semen, and also for kinematic parameters (computerized analysis). A significant decrease in sperm motility and kinetic rating was observed after glycerol addition at 5 °C and also after thawing for all the treatments (P < 0.05). Regarding post-thaw semen variables, no differences were verified between freezing curves when the same straw size and thawing rate were taken as reference (P > 0.05). In general, values for sperm characteristics found after thawing at 37 °C were better preserved than at 70 °C (P < 0.05), both in the use of 0.25 mL or 0.50 mL straws, which were similar for semen packaging (P > 0.05). The evaluation of the kinematic parameters of sperm motility confirmed these results at values varying from 20% to 30% motile sperm for the samples thawed at 37 °C, and values fewer than 12% motile sperm for samples thawed at 70 °C (P < 0.05). In conclusion, we recommend the use of a fast freezing curve that reduces the time spent on the cryopreservation of collared peccary semen, which could be packaged both in 0.25 mL or 0.50 mL straws, but the thawing should be conducted at 37 °C/1 min.  相似文献   

17.
An ethanol extract of the aerial parts of Delphinium gracile DC. yielded five flavonol glycosides quercetin-3-O-{[β-d-xylopyranosyl (1 → 3)-4-O-(E-p-caffeoyl)-α-l-rhamnopyranosyl (1 → 6)][β-d-glucopyranosyl (1 → 2)]}-β-d-glucopyranoside (1), quercetin-3-O-{[β-d-xylopyranosyl (1 → 3)-4-O-(E-p-coumaroyl)-α-l-rhamnopyranosyl (1 → 6)][β-d-glucopyranosyl (1 → 2)]}-β-d-glucopyranoside (2), quercetin-3-O-{[β-d-xylopyranosyl (1 → 3)-4-O-(Z-p-coumaroyl)-α-l-rhamnopyranosyl (1 → 6)][β-d-glucopyranosyl (1 → 2)]}-β-d-glucopyranoside (3), kaempferol-3-O-{[β-d-glucopyranosyl (1 → 3)-4-O-(E-p-coumaroyl)-α-l-rhamnopyranosyl (1 → 6)][β-d-glucopyranoside-7-O-(4-O-acetyl)-α-l-rhamnopyranoside (4) kaempferol-3-O-{[β-d-glucopyranosyl (1 → 3)-4-O-(E-p-coumaroyl)-α-l-rhamnopyranosyl (1 → 6)][β-d-glucopyranoside-7-O-(4-O-acetyl)-α-l-rhamnopyranoside (5) in addition to 4-(β-d-glucopyranosyloxy)-6-methyl-2H-pyran-2-one (6) and rutin. Structures were elucidated by spectroscopic methods.  相似文献   

18.
Global warming and associated increases in the frequency and amplitude of extreme weather events, such as heat waves, may adversely affect tropical rainforest plants via significantly increased tissue temperatures. In this study, the response to two temperature regimes was assessed in seedlings of the neotropical pioneer tree species, Ficus insipida. Plants were cultivated in growth chambers at strongly elevated daytime temperature (39 °C), combined with either close to natural (22 °C) or elevated (32 °C) nighttime temperatures. Under both growth regimes, the critical temperature for irreversible leaf damage, determined by changes in chlorophyll a fluorescence, was approximately 51 °C. This is comparable to values found in F. insipida growing under natural ambient conditions and indicates a limited potential for heat tolerance acclimation of this tropical forest tree species. Yet, under high nighttime temperature, growth was strongly enhanced, accompanied by increased rates of net photosynthetic CO2 uptake and diminished temperature dependence of leaf-level dark respiration, consistent with thermal acclimation of these key physiological parameters.  相似文献   

19.
Hemocyanin (Hc) is a type-3 copper protein, containing dioxygen-binding active sites consisting of paired copper atoms. In the present study the thermal unfolding of the Hc from the marine mollusc Rapana thomasiana (RtH) has been investigated by combining differential scanning calorimetry, Fourier transform infrared (FTIR) and UV–vis absorption spectroscopy. Two important stages in the unfolding pathway of the Hc molecule were discerned. A first event, with nonmeasurable heat absorption, occurring around 60 °C, lowers the binding of dioxygen to the type-3 copper groups. This pretransition is reversible and is ascribed to a slight change in the tertiary structure. In a second stage, with midpoint around 80 °C, the protein irreversibly unfolds with a loss of secondary structure and formation of amorphous aggregates. Experiments with the monomeric structural subunits, RtH1 and RtH2, indicated that the heterogeneity in the process of thermal denaturation can be attributed to the presence of multiple 50 kDa functional units with different stability. In accordance, the irreversible unfolding of a purified functional unit (RtH2-e) occurred at a single transition temperature. At slightly alkaline pH (Tris buffer) the C-terminal β-sheet rich domain of the functional unit starts to unfold before the α-helix-rich N-terminal (copper containing) domain, triggering the collapse of the global protein structure. Even around 90 °C some secondary structure is preserved as shown by the FTIR spectra of all investigated samples, confirming the high thermostability of molluscan Hc.  相似文献   

20.
There is a limited understanding of the folding of multidomain membrane proteins. Lactose permease (LacY) of Escherichia coli is an archetypal member of the major facilitator superfamily of membrane transport proteins, which contain two domains of six transmembrane helices each. We exploit chemical denaturation to determine the unfolding free energy of LacY and employ Trp residues as site-specific thermodynamic probes. Single Trp LacY mutants are created with the individual Trps situated at mirror image positions on the two LacY domains. The changes in Trp fluorescence induced by urea denaturation are used to construct denaturation curves from which unfolding free energies can be determined. The majority of the single Trp tracers report the same stability and an unfolding free energy of approximately + 2 kcal mol− 1. There is one exception; the fluorescence of W33 at the cytoplasmic end of helix I on the N domain is unaffected by urea. In contrast, the equivalent position on the first helix, VII, of the C-terminal domain exhibits wild-type stability, with the single Trp tracer at position 243 on helix VII reporting an unfolding free energy of + 2 kcal mol− 1. This indicates that the region of the N domain of LacY at position 33 on helix I has enhanced stability to urea, when compared the corresponding location at the start of the C domain. We also find evidence for a potential network of stabilising interactions across the domain interface, which reduces accessibility to the hydrophilic substrate binding pocket between the two domains.  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