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 31 毫秒
1.
 Direct cyclic voltammetry and 1H NMR spectroscopy have been combined to investigate the electrochemical and spectroscopic properties of cytochrome c 553 isolated from the alkaliphilic soil bacterium Bacillus pasteurii. A quasi-reversible diffusion-controlled redox process is exhibited by cytochrome c 553 at a pyrolitic graphite edge microelectrode. The temperature dependence of the reduction potential, measured using a non-isothermal electrochemical cell, revealed a discontinuity at 308 K. The thermodynamic parameters determined in the low-temperature range (275–308 K;ΔS°′=–162.7±1.2 J mol–1 K–1, ΔH°′=–53.0±0.5 kJ mol–1, ΔG°′=–4.5±0.1 kJ mol–1, E°′=+47.0±0.6 mV) indicate the presence of large enthalpic and entropic effects, leading, respectively, to stabilization and destabilization of the reduced form of cytochrome c 553. Both effects are more accentuated in the high-temperature range (308–323 K;ΔS°′=–294.1±8.4 J mol–1 K–1, ΔH°′=–93.4±3.1 kJ mol–1, ΔG°′=–5.8±0.6 kJ mol–1, E°′=+60.3±5.8 mV), with the net result being a slight increase of the standard reduction potential. These thermodynamic parameters are interpreted using the compensation theory of hydration of biopolymers as indicating the extrusion, upon reduction, of water molecules from the hydration sphere of the cytochrome. The low-T and high-T conformers differ by the number of water molecules in the solvation sphere: in the high-T conformer, the number of water molecules extruded upon reduction increases, as compared to the low-T conformer. The ionic strength dependence of the reduction potential at 298 K, treated within the frame of extended Debye-Hückel theory, yields values of E °′ (I=0) =–25.4±1.4 mV, z red=–11.3, and z ox=–10.3. The pH dependence of the reduction potential at 298 K shows a plateau in the pH range 7–10 and an increase at more acidic pH, allowing the calculation of pK O=5.5 and pK R=5.7, together with the estimate of the reduction potentials of completely protonated (+71 mV) and deprotonated (+58 mV) forms of cytochrome c 553. 1H NMR spectra of the oxidized paramagnetic cytochrome c 553 indicate the presence of a His-Met axial coordination of the low-spin (S=1/2) heme iron, which is maintained in the temperature interval 288–340 K at pH 7 and in the pH range 4.8–10.0 at 298 K. The temperature dependence of the hyperfine-shifted signals shows both Curie-type and anti-Curie-type behavior, with marked deviations from linearity, interpreted as indicating the presence of a fast equilibrium between the low-T and high-T conformers, having slightly different heme electronic structures resulting from the T-induced conformational change. Increasing the NaCl concentration in the range 0–0.2 M causes a slight change of the 1H NMR chemical shifts of the hyperfine-shifted signals, with no influence on their linewidth. The calculated lower limit value of the apparent affinity constant for specific ion binding is estimated as 5.2±1.1 M–1. The pH dependence of the isotropically shifted 1H NMR signals of the oxidized cytochrome displays at least one ionization step with pK O=5.7. The thermodynamic and spectroscopic data indicate a large solvent-derived entropic effect as the main cause for the observed low reduction potential of B. pasteurii cytochrome c 553. Received: 9 January 1998 / Accepted: 8 April 1998  相似文献   

2.
A superoxide dismutase (SOD) was characterized from Beauveria bassiana, a fungal entomopathogen widely applied to insect control. This 209-aa enzyme (BbSod2) showed no more than 71% sequence identity to other fungal Mn-SODs, sharing all conserved residues with the Mn-SOD family and lacking a mitochondrial signal. The SOD activity of purified BbSod2 was significantly elevated by Mn2+, suppressed by Cu2+ and Zn2+ but inhibited by Fe3+. Overexpressing the enzyme in a BbSod2-absent B. bassiana strain enhanced its SOD activity (107.2 ± 6.1 U mg−1 protein) by 4–10-fold in different transformants analyzed. The best BbSod2-transformed strain with the SOD activity of 1,157.9 ± 74.7 U mg−1 was 93% and 61% more tolerant to superoxide-generating menadione in both colony growth (EC50 = 2.41 ± 0.03 versus 1.25 ± 0.01 mM) and conidial germination (EC50 = 0.89 ± 0.06 versus 0.55 ± 0.07 mM), and 23% more tolerant to UV-B irradiation (LD50 = 0.49 ± 0.02 versus 0.39 ± 0.01 J cm−2). Its virulence to Spodoptera litura larvae was enhanced by 26% [LT50 = 4.5 (4.2–4.8) versus 5.7 (5.2–6.4) days]. Our study highlights for the first time that the Mn2+-cofactored, cytosolic BbSod2 contributes significantly to the virulence and stress tolerance of B. bassiana and reveals possible means to improving field persistence and efficacy of a fungal formulation by manipulating the antioxidant enzymes of a candidate strain.  相似文献   

3.
 Dicyanoaurate(I), Au(CN)2 , an important metabolite of chrysotherapy agents (anti-arthritic gold drugs), contains two tightly bound cyanide ligands which render it relatively unreactive toward ligand exchange reactions with potential gold-binding ligands. The extent and nature of its binding to bovine serum albumin (BSA), which may modulare the in vivo activity of Au(CN)2 , were investigated to determine whether Au(CN)2 might be more bioavailable than other gold complexes. 13C NMR spectroscopy, radioisotope tracers, chromatography, ultrafiltration, and atomic spectroscopy, employing Au(13CN)2 or Au(14CN)2 as appropriate, revealed two distinct binding mechanisms. The dominant reaction is reversible association (non-specific binding) of intact Au(CN)2 ions to form BSA·[Au(CN)2 ] n adducts. Approximately one equivalent binds with an equilibrium binding constant (pH 7.4, 25  °C) of K 1=5.5 (±1.1)×104, and three additional equivalents bind with a constant of 7.0 (±0.1)×103. Au(13CN)2 associated with albumin is characterized by a broad 13C NMR resonance at δC=154.7 ppm compared to the sharp resonance of the free complex at 156.4 ppm. The BSA·[Au(CN)2 ] n adducts readily dissociate during gel exclusion chromatography and are therefore underestimated, but are retained and accurately quantitated by ultrafiltration methods. The second binding mechanism is a ligand exchange reaction at Cys-34, to form AlbSAuCN, which accounts for only a small fraction (≤11%) of the bound gold. The small extent of the latter interaction differentiates Au(CN)2 from the gold drugs such as auranofin, aurothiomalate (Myochrysin) and aurothioglucose (Solganol), which undergo ligand exchange at Cys-34 of albumin to form tightly bound gold-protein complexes. The weak interaction at Cys-34 and the facile dissociation of bound, intact Au(CN)2 are consistent with its putative role as a gold metabolite that can be accumulated intracellularly. Received: 2 July 1997 / Accepted: 24 September 1997  相似文献   

4.
Isotherms of the EtBr adsorption on native and denatured poly(dA)poly(dT) in the temperature interval 20–70°C were obtained. The EtBr binding constants and the number of binding sites were determined. The thermodynamic parameters of the EtBr intercalation complex upon changes of solution temperature 20–48°C were calculated: 1.0·106 M−1K≤1.4·106 M−1, free energy ΔG o=−8.7±0.3 kcal/mol, enthalpy ΔH o≅0, and entropy ΔS o=28±0.5 cal/(mol deg). UV melting has shown that the melting temperature (T m) of EtBr-poly(dA)poly(dT) complexes (μ=0.022,4.16·10−5 M EtBr) increased by 17°C as compared with the ΔT m of free homopolymer, whereas the half-width of the transition (T m) is not changed. It was shown for the first time that EtBr forms complexes of two types on single-stranded regions of poly(dA)poly(dT) denatured at 70°C: strong (K 1=1.7·105 M−1; ΔG o=−8.10±0.03 kcal/mol) and weak (K 2=2.9·103 M−1; ΔG o=−6.0±0.3 kcal/mol).The ΔG o of the strong and weak complexes was independent of the solution ionic strength, 0.0022≤μ≤0.022. A model of EtBr binding with single-stranded regions of poly(dA)poly(dT) is discussed.  相似文献   

5.
The metal–thiolate connectivity of recombinant Cd7-MT10 metallothionein from the sea mussel Mytilus galloprovincialis has been investigated for the first time by means of multinuclear, multidimensional NMR spectroscopy. The internal backbone dynamics of the protein have been assessed by the analysis of 15N T 1 and T 2 relaxation times and steady state {1H}–15N heteronuclear NOEs. The 113Cd NMR spectrum of mussel MT10 shows unique features, with a remarkably wide dispersion (210 ppm) of 113Cd NMR signals. The complete assignment of cysteine Hα and Hβ proton resonances and the analysis of 2D 113Cd–113Cd COSY and 1H–113Cd HMQC type spectra allowed us to identify a four metal–thiolate cluster (α-domain) and a three metal–thiolate cluster (β-domain), located at the N-terminal and the C-terminal, respectively. With respect to vertebrate MTs, the mussel MT10 displays an inversion of the α and β domains inside the chain, similar to what observed in the echinoderm MT-A. Moreover, unlike the MTs characterized so far, the α-domain of mussel Cd7-MT10 is of the form M4S12 instead of M4S11, and has a novel topology. The β-domain has a metal–thiolate binding pattern similar to other vertebrate MTs, but it is conformationally more rigid. This feature is quite unusual for MTs, in which the β-domain displays a more disordered conformation than the α-domain. It is concluded that in mussel Cd7-MT10, the spacing of cysteine residues and the plasticity of the protein backbone (due to the high number of glycine residues) increase the adaptability of the protein backbone towards enfolding around the metal–thiolate clusters, resulting in minimal alterations of the ideal tetrahedral geometry around the metal centres.  相似文献   

6.
Water status and gas exchange of beech (Fagus sylvatica L.) and sessile oak [Quercus petraea (Mattuschka) Liebl.] were studied in a mixed stand in the Montejo de la Sierra forest (central Iberian Peninsula), one of the southernmost locations of both species in Europe. Gas exchange and water potential were measured in leaves at different canopy levels over several days in two growing seasons. The daily variation pattern was established with the measurements of three selected dates per year, representative of the soil moisture content situations in early, mid- and late summer. A similar daily time course of leaf water potential was found for the two species. Nevertheless, beech showed a most noticeable decrease of water potential at midmorning and maintained lower leaf water potential than oak in the early afternoon. In 1994 the sessile oak saplings showed higher values of predawn water potential (Ψpd) than beech at the end of summer, when soil moisture content was lowest (20 cm depth). Beech showed a significantly lower net assimilation rate (A) than sessile oak for leaves under the same PPFD. Maximum net photosynthesis values (A max) for beech and sessile oak on sunny leaves were 10.1±0.4 μmol m–2 s–1 and 17.8±1.7 μmol m–2 s–1 respectively, and those for water vapour stomatal conductance (g wv) were 265±31 mmol m–2 s–1 and 438±74 mmol m–2 s–1. Differences in A and g wv between the two species were maintained throughout the day on all measurement dates. No clear relationship was found between water status of saplings and stomata performance; there was only a negative correlation between Ψpd and g wvmid in beech. Nevertheless, a significant response to the air vapour pressure gradient between leaf and air was translated into stomata closure on an hourly basis, more intensively in beech. Received: 4 March 1999 / Accepted: 21 December 1999  相似文献   

7.
Two moderately halophilic, Gram-negative, rod-shaped bacteria, designated YIM 93003T and YIM 94343T, were isolated from a salt lake in Xinjiang province, north-west China. The two strains YIM 93003T and YIM 94343T grew at 20–40°C, pH 6–9, 0.5–24% (w/v) NaCl and at 20–40°C, pH 6–9, 0.5–23% (w/v) NaCl, respectively. No growth occurred in absence of NaCl. Phylogenetic analyses based on 16S rRNA gene sequences showed that strains YIM 93003T and YIM 94343T were phylogenetically affiliated to the genus Halomonas and exhibited sequence similarity of 97.5% and 97.4% to the type strain Halomonas anticariensis DSM 16096T, respectively. The strains possessed chemotaxonomic markers that were consistent with their classification in the genus Halomonas (Q-9 as predominant respiratory quinine; C18:1ω7c, C16:0 and C16:1 ω7c/iso-C15:02-OH as the major fatty acids). The DNA–DNA hybridization values for strains YIM 93003T and YIM 94343T, YIM 93003T and DSM 16096T, YIM 94343T and DSM 16096T were 38.1 ± 3.0, 18.3 ± 4.7, and 20.8 ± 4.6%, respectively. The G+C contents of the strains YIM 93003T and YIM 94343T were 63.4 and 64.0 mol%, respectively. Based on comparative analysis of physiological, biochemical and chemotaxonomic data, including low DNA–DNA hybridization results, two novel species, Halomonas qijiaojingensis sp. nov., and Halomonas flava sp. nov., are proposed. The type strains are YIM 93003T (=CCTCC AB 208133T =KCTC 22228T) and YIM 94343T (=CCTCC AB 2010382T =KCTC 23356T), respectively.  相似文献   

8.
Data on thermal energetics for vespertilionid bats are under-represented in the literature relative to their abundance, as are data for bats of very small body mass. Therefore, we studied torpor use and thermal energetics in one of the smallest (4 g) Australian vespertilionids, Vespadelus vulturnus. We used open-flow respirometry to quantify temporal patterns of torpor use, upper and lower critical temperatures (T uc and T lc) of the thermoneutral zone (TNZ), basal metabolic rate (BMR), resting metabolic rate (RMR), torpid metabolic rate (TMR), and wet thermal conductance (C wet) over a range of ambient temperatures (T a). We also measured body temperature (T b) during torpor and normothermia. Bats showed a high proclivity for torpor and typically aroused only for brief periods. The TNZ ranged from 27.6°C to 33.3°C. Within the TNZ T b was 33.3±0.4°C and BMR was 1.02±0.29 mlO2 g−1 h−1 (5.60±1.65 mW g−1) at a mean body mass of 4.0±0.69 g, which is 55 % of that predicted for a 4 g bat. Minimum TMR of torpid bats was 0.014±0.006 mlO2 g−1 h−1 (0.079±0.032 mW g−1) at T a=4.6±0.4°C and T b=7.5±1.9. T lc and C wet of normothermic bats were both lower than that predicted for a 4 g bat, which indicates that V. vulturnus is adapted to minimising heat loss at low T a. Our findings support the hypothesis that vespertilionid bats have evolved energy-conserving physiological traits, such as low BMR and proclivity for torpor.  相似文献   

9.
A Gram-stain-negative, non-motile, rod-shaped bacterial strain, JW-64-1T, capable of degrading methamidophos was isolated from a methamidophos-manufacturing factory in China, and was subjected to a polyphasic taxonomic investigation. Strain JW-64-1T produced circular, smooth, transparent, yellow-colored colonies (1.0–2.0 mm) on LB agar after 2 days incubation. It grew optimally at 25–30°C and pH 7.0 without the presence of NaCl. The G+C content of the total DNA was 63.6 mol%. A phylogenetic analysis based on 16S rRNA gene sequences showed that strain JW-64-1T fell within the cluster comprising Luteibacter species. The 16S rRNA gene sequence of strain JW-64-1T was most closely related to Luteibacter rhizovicinus DSM 16549T (98.6%), followed by Luteibacter yeojuensis DSM 17673T (98.4%) and L. anthropi CCUG 25036T (98.2%). The major cellular fatty acids of strain JW-64-1T were iso-C15:0 (24.1%), iso-C17:0 (20.2%) and summed feature 9 comprising iso-C17:1 ω9c and/or C16:0 10-methyl (20.3%). The major isoprenoid quinine was Q-8 (98%), and the major polar lipids were diphosphatidylglycerol, phosphatidylglycerol, phosphoaminolipid, aminolipids-1, aminolipids-2, and phospholipids. The values for DNA–DNA relatedness between strain JW-64-1T and the closest phylogenetic relatives of L. rhizovicinus and Luteibacter yeojuensis were 34.8 ± 2.6 and 25.6 ± 3.1%, respectively. On the basis of the phenotypic, chemotaxonomic, DNA–DNA relatedness and phylogenetic analysis based on the 16S rRNA gene sequences, strain JW-64-1T represents a novel species of the genus Luteibacter, for which the name Luteibacter jiangsuensis sp. nov. is proposed. The type strain is JW-64-1T (=CGMCC 1.10133T = DSM 22396T).  相似文献   

10.
A novel strictly anaerobic bacterium designated SPDX02-08T was isolated from a deep terrestrial geothermal spring located in southwest France. Cells (1–2 × 2–6 μm) were non-motile, non sporulating and stained Gram negative. Strain SPDX02-08T grew at a temperature between 40 and 60°C (optimum 55°C), pH between 6.3 and 7.3 (optimum 7.2) and a NaCl concentration between 0 and 5 g/l (optimum 2 g/l). Sulfate, thiosulfate and sulfite were used as terminal electron acceptors, but not elemental sulfur, nitrate, nitrite, Fe (III) or fumarate. In the presence of sulfate, strain SPDX02-08T completely oxidized pyruvate, propionate, butyrate, isobutyrate, valerate, isovalerate and hexadecanoate. Stoichiometric measurements revealed a complete oxidation of part of lactate (0.125 mol of acetate produced per mole lactate oxidized). Strain SPDX02-08T required yeast extract to oxidize formate and H2 but did not grow autotrophically on H2. Among the substrates tested, only pyruvate was fermented. The G+C content of the genomic DNA was 57.6 mol%. Major cellular fatty acids of strain SPDX02-08T were iso-C15:0, C15:0, and C16:0. Phylogenetic analysis of the 16S small-subunit (SSU) ribosomal RNA gene sequence indicated that strain SPDX02-08T belongs to the genus Desulfosoma, family Syntrophobacteraceae, having Desulfosoma caldarium as its closest phylogenetic relative (97.6% similarity). The mean DNA/DNA reassociation value between strain SPDX02-08T and Desulfosoma caldarium was 16.9 ± 2.7%. Based on the polyphasic differences, strain SPDX02-08T is proposed to be assigned as a new species of the genus Desulfosoma, Desulfosoma profundi sp. nov. (DSM 22937T = JCM 16410T). GenBank accession number for the 16S rRNA gene sequence of strain SPDX02-08T is HM056226.  相似文献   

11.
A wet suit may not provide adequate thermal protection when diving in moderately cold water (17–18°C), and any resultant mild hypothermia may impair performance during prolonged diving. We studied heat exchange during a dive to a depth of 5 m in sea water (17–18.5°C) in divers wearing a full wet suit and using closed-circuit oxygen breathing apparatus. Eight fin swimmers dived for 3.1 h and six underwater scooter (UWS) divers propelled themselves through the water for 3.7 h. The measurements taken throughout the dive were the oxygen pressure in the cylinder and skin and rectal temperatures (T re). Each subject also completed a cold score questionnaire. The T re decreased continuously in all subjects. Oxygen consumption in the fin divers (1.40 l · min−1) was higher than that of the UWS divers (1.05 l · min−1). The mean total insulation was 0.087°C · m2 · W−1 in both groups. Mean body insulation was 37% of the total insulation (suit insulation was 63%). The reduction in T re over the 1st hour was related to subcutaneous fat thickness. There was a correlation between cold score and T re at the end of 1 h, but not after that. A full wet suit does not appear to provide adequate thermal protection when diving in moderately cold water. Accepted: 21 January 1997  相似文献   

12.
Forests with nitrogen-fixing trees (N–fixers) typically accumulate more carbon (C) in soils than similar forests without N–fixing trees. This difference may develop from fundamentally different processes, with either greater accumulation of recently fixed C or reduced decomposition of older soil C. We compared the soil C pools under N–fixers with Eucalyptus (non–N–fixers) at four tropical sites: two sites on Andisol soils in Hawaii and two sites on Vertisol and Entisol soils in Puerto Rico. Using stable carbon isotope techniques, we tracked the loss of the old soil organic C from the previous C4 land use (SOC4) and the gain of new soil organic C from the C3, N–fixer, and non–N–fixer plantations (SOC3). Soils beneath N–fixing trees sequestered 0.11 ± 0.07 kg m−2 y−1 (mean ± one standard error) of total soil organic carbon (SOCT) compared with no change under Eucalyptus (0.00 ± 0.07 kg m−2 y−1; P = 0.02). About 55% of the greater SOCT sequestration under the N–fixers resulted from greater retention of old SOC4, and 45% resulted from greater accretion of new SOC3. Soil N accretion under the N–fixers explained 62% of the variability of the greater retention of old SOC4 under the N–fixers. The greater retention of older soil C under N–fixing trees is a novel finding and may be important for strategies that use reforestation or afforestation to offset C emissions. Received 12 March 2001; accepted 5 October 2001.  相似文献   

13.
We aimed to investigate the pattern of utilisation of torpor and its impact on energy budgets in free-living grey mouse lemurs (Microcebus murinus), a small nocturnal primate endemic to Madagascar. We measured daily energy expenditure (DEE) and water turnover using doubly labelled water, and we used temperature-sensitive radio collars to measure skin temperature (T sk) and home range. Our results showed that male and female mouse lemurs in the wild enter torpor spontaneously over a wide range of ambient temperatures (T a) during the dry season, but not during the rainy season. Mouse lemurs remained torpid between 1.7–8.9 h with a daily mean of 3.4 h, and their T sk s fell to a minimum of 18.8 °C. Mean home ranges of mouse lemurs which remained normothermic were similar in the rainy and dry season. During the dry season, the mean home range of mouse lemurs showing daily torpor was significantly smaller than that of animals remaining normothermic. The DEE of M. murinus remaining normothermic in the rainy season (122 ± 65.4 kJ day−1) was about the same of that of normothermic mouse lemurs in the dry season (115.5 ± 27.3 kJ day−1). During the dry season, the mean DEE of M. murinus that utilised daily torpor was 103.4 ± 32.7 kJ day−1 which is not significantly different from the mean DEE of animals remaining normothermic. We found that the DEE of mouse lemurs using daily torpor was significantly correlated with the mean temperature difference between T sk and T a (r 2=0.37) and with torpor bout length (r 2 =0.46), while none of these factors explained significant amounts of variation in the DEE of the mouse lemurs remaining normothermic. The mean water flux rate of mouse lemurs using daily torpor (13.0 ± 4.1 ml day−1) was significantly lower than that of mouse lemurs remaining normothermic (19.4 ± 3.8 ml day−1), suggesting the lemurs conserve water by entering torpor. Thus, this first study on the energy budget of free-ranging M. murinus demonstrates that torpor may not only reflect its impact on the daily energy demands, but involve wider adaptive implications such as water requirements. Accepted: 29 August 2000  相似文献   

14.
 The Gd(III) complex of 4-pentylbicyclo[2.2.2]octane-1-carboxyl-di-l-aspartyl-lysine-derived DTPA, [GdL(H2O)]2–, binds to serum albumin in vivo, through hydrophobic interaction. A variable temperature 17O NMR, EPR, and Nuclear Magnetic Relaxation Dispersion (NMRD) study resulted in a water exchange rate of k 298 ex=4.2×106 s–1, and let us conclude that the GdL complex is identical to [Gd(DTPA)(H2O)]2– in respect to water exchange and electronic relaxation. The effect of albumin binding on the water exchange rate has been directly evaluated by 17O NMR. Contrary to expectations, the water exchange rate on GdL does not decrease considerably when bound to bovine serum albumin (BSA); the lowest limit can be given as k ex, GdL-BSA=k ex, GdL / 2. In the knowledge of the water exchange rate for the BSA-bound GdL complex, the analysis of its NMRD profile at 35  °C yielded a rotational correlation time of 1.0 ns, one order of magnitude shorter than that of the whole protein. This value is supported by the longitudinal 17O relaxation rates. This indicates a remarkable internal flexibility, probably due to the relatively large distance between the protein- and metal-binding moieties of the ligand. Received: 25 June 1998 / Accepted: 11 August 1998  相似文献   

15.
Studied was the effect of temperature in the range 12–46 °C on the rate of bacterial decolorization of the mono-azo dye Acid Orange 7 by Alcaligenes faecalis 6132 and Rhodococcus erythropolis 24. With both strains the raise of temperature led to a corresponding raise of decolorization rate better manifested by R. erythropolis. The analysis of the Arrhenius plot revealed a break near the middle of the temperature range. The regression analysis showed practically complete identity of the observed break point temperatures (T BP): 20.7 °C for Alc. faecalis and 20.8 °C for R. erythropolis. The values of the activation energy of the decolorization reaction (E a) were found to depend on both the organism and the temperature range. In the range below T BP the estimated values of E a were 138 ± 7 kJ mol−1 for Alc. faecalis and 160 ± 8 kJ mol−1 for R. erythropolis. In the range above T BP they were 54.2 ± 1.8 kJ mol−1 for Alc. faecalis and 37.6 ± 4.1 kJ mol−1 for R. erythropolis. Discussed are the possible reasons for the observed abrupt change of the activation energy.  相似文献   

16.
When aspirating human red blood cells (RBCs) into 1.3 μm pipettes (ΔP = −2.3 kPa), a transition from blocking the pipette below a critical temperature T c = 36.3 ± 0.3°C to passing it above the T c occurred (micropipette passage transition). With a 1.1 μm pipette no passage was seen which enabled RBC volume measurements also above T c. With increasing temperature RBCs lost volume significantly faster below than above a T c = 36.4 ± 0.7 (volume transition). Colloid osmotic pressure (COP) measurements of RBCs in autologous plasma (25°C ≤ T ≤ 39.5°C) showed a T c at 37.1 ± 0.2°C above which the COP rapidly decreased (COP transition). In NMR T1-relaxation time measurements, the T1 of RBCs in autologous plasma changed from a linear (r = 0.99) increment below T c = 37 ± 1°C at a rate of 0.023 s/K into zero slope above T c (RBC T1 transition). In conclusion: An amorphous hemoglobin–water gel formed in the spherical trail, the residual partial sphere of the aspirated RBC. At T c, a sudden fluidization of the gel occurs. All changes mentioned above happen at a distinct T c close to body temperature. The T c is moved +0.8°C to higher temperatures when a D2O buffer is used. We suggest a mechanism similar to a “glass transition” or a “colloidal phase transition”. At T c, the stabilizing Hb bound water molecules reach a threshold number enabling a partial Hb unfolding. Thus, Hb senses body temperature which must be inscribed in the primary structure of hemoglobin and possibly other proteins. This article is dedicated to Ludwig Artmann who died on July 21, 2001 on a beautiful summer day during which we performed experiments far away. Ludwig Artmann was a man who encouraged us to be strong and to study hard no matter what were the costs.  相似文献   

17.
 The ligand DOTASA was designed and synthesized in the aim of obtaining a kinetically and thermodynamically stable Gd(III) chelate which, through its uncoordinated carboxylate function, will provide an efficient pathway to couple the complex to bio- or macromolecules without affecting the coordination pattern of DOTA. Furthermore, it allows us to study the influence of an extra carboxylate arm on the parameters determining proton relaxivity in comparison to the commercial agent [Gd(DOTA)(H2O)]. A combined variable-temperature 17O NMR, EPR and nuclear magnetic relaxation dispersion study on the Gd(III) chelate resulted in k 298 ex=(6.3±0.2)×106 s–1 for the water exchange rate and τ298 R=125±2 ps for the rotational correlation time. The slight increase in both k 298 ex and τ298 R, as compared to those for [Gd(DOTA)(H2O)], is attributed to the presence of the extra negative charge. The longer rotational correlation time results in a proton relaxivity of 5.03 mM–1 s–1 for [Gd(DOTASA)(H2O)]2–, which is approximately 30% higher than that for [Gd(DOTA)(H2O)]. The increased water exchange rate of [Gd(DOTASA)(H2O)]2– has no consequence for proton relaxivity since this latter is exclusively limited by fast rotation for both complexes. However, for slowly rotating macromolecular agents, which contain a covalently coupled DOTASA unit instead of a coupled DOTA, this increased exchange rate will have a significant positive effect. Received: 31 December 1998 / Accepted: 4 March 1999  相似文献   

18.
 The purpose of this investigation was to examine the effects of aging and aerobic fitness on exercise- and methylcholine-induced sweating responses during heat acclimation. Five younger [Y group – age: 23±1 (SEM) years; maximal oxygen consumption (V.O2max): 47±3 ml·kg–1·min–1], four highly fit older (HO group – 63±3 years; 48±4 ml·kg–1·min–1) and five normally fit older men (NO group – 67±3 years; 30±1 ml·kg–1·min–1) who were matched for height, body mass and percentage fat, were heat acclimated by daily cycle exercise (≈35% V.O2max for 90 min) in a hot (43°C, 30% RH) environment for 8 days. The heat acclimation regimen increased performance time, lowered final rectal temperature (T re) and percentage maximal heart rate (%HRmax), improved thermal comfort and decreased sweat sodium concentration similarly in all groups. Although total body sweating rates (M.sw) during acclimation were significantly greater in the Y and HO groups than in the NO group (P<0.01) (because of the lower absolute workload in the NO group), the M.sw did not change in all groups with the acclimation sessions. Neither were local sweating rates (m. sw) on chest, back, forearm and thigh changed in all groups by the acclimation. The HO group presented greater forearm m. sw (30–90 min) values and the Y group had greater back and thigh m. sw (early in exercise) values, compared to the other groups (P<0.001). In a methylcholine injection test on days immediately before and after the acclimation, the order of sweat output per gland (SGO) on chest, back and thigh was Y>HO>NO, and on the forearm Y=HO>NO. No group differences were observed for activated sweat gland density at any site. The SGO at the respective sites increased in the post-acclimation test regardless of group (P<0.01), but on the thigh the magnitude of the increase was lower in the NO (P<0.02) and HO (P=0.07) groups than in the Y group. These findings suggest that heat tolerance and the improvement with acclimation are little impaired not only in highly fit older but also normally fit older men, when the subjects exercised at the same relative exercise intensity. Furthermore, the changes induced by acclimation appear associated with an age-related decrease in V.O2max. However methylcholine-activated SGO and the magnitude of improvement of SGO with acclimation are related not only to V.O2max but also to aging, suggesting that sensitivity to cholinergic stimulation decreases with aging. Received: 8 May 1998/Accepted: 5 October 1998  相似文献   

19.
A novel thermophilic anaerobic and microaerophilic bacterium (optimal growth in the presence of 5–10% O2), strain Nad S1T was isolated from the terrestrial hot spring of Hammam Sidi Jdidi, Nabeul, Tunisia. Cells were motile rods having a Gram-positive cell wall structure. Strain Nad S1T grew optimally at 55°C (range 37–70°C). Optimum pH for growth was 6.5–7.0. It was halotolerant growing with NaCl up to 7% (optimum concentration 1.5–3.0%). It grew chemoorganotrophically on various carbohydrates, organic-acids and amino-acids as energy sources, or chemolithotrophically on H2 using nitrate, as terminal electron acceptor. Beside oxygen (under microaerobic conditions) and nitrate, nitrite was also used. Nitrate was completely reduced to N2. No fermentation occurred. The genomic DNA G + C content was 41.8 mol%. Based on 16S rRNA gene sequence analysis, strain Nad S1T belongs to the Bacillaceae family within the class ‘Bacilli’. Because of its phylogenetic and phenotypic characteristics, we propose this isolate to be assigned as a novel genus and a novel species within the domain Bacteria, Microaerobacter geothermalis gen. nov., sp. nov. The type strain is Nad S1T (=DSM 22679T =JCM 16213T).  相似文献   

20.
A novel Gram-positive, aerobic, rod-shaped and mycelia-producing bacterial strain, designated KLBMP 1050T, was isolated from the stem of the oil-seed plant Jatropha curcas L. collected from Sichuan Province, south-west China. Phylogenetic analysis based on the 16S rRNA gene sequence revealed that the isolate KLBMP 1050T belonged to the genus Nocardioides, with the highest sequence similarity to Nocardioides albus KCTC 9186T (99.38 %) and Nocardioides luteus KCTC 9575T (99.03 %). However, the DNA–DNA relatedness of isolate KLBMP 1050T to these two type strains were 37.5 ± 3.5 and 33 ± 2.3 %, respectively. Strain KLBMP 1050T grew at the pH range 6–11, temperature range 10–32 °C and with 0–12 % NaCl. The physiological properties of strain KLBMP 1050T differ from those of N. albus KCTC 9186T and N. luteus KCTC 9575T. The cell-wall peptidoglycan contained ll-diaminopimelic acid and MK-8(H4) was the major respiratory quinone. The predominant cellular fatty acid of strain KLBMP 1050T was iso-C16:0 (23.3 %). The total DNA G+C content was 70.1 mol%. On the basis of the phenotypic, chemotaxonomic and phylogenetic data, strain KLBMP 1050T represents a novel species of the genus Nocardioides, for which the name Nocardioides panzhihuaensis sp. nov. is proposed. The type strain is KLBMP 1050T (= KCTC 19888T = NBRC 108680T).  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