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.
1. Factorial designs were used to investigate the spatial and temporal components of temperature sampling error in two medium-sized (maximum fetch = 9 and 11 km; maximum depth = 24 and 70m) Wisconsin lakes (Mendota and Green). Sampling designs were then optimized for estimating temperature distributions, thermocline migration, lake heat content, and vertical eddy conductivities in these and similar stratified lakes. 2. Errors were at a minimum for: (i) isotherms positioned 1/3 - 1/2 of the way down through the seasonal thermocline; (ii) at lake stations located near the node of the principal internal seiche; (iii) after an extended interval of low windpower. If seven spatially distributed lake stations were sampled n1, n2 times during the two low-power intervals bracketing a weather front, and the rime delay between revisits was randomized with respect to the period of the uninodal temperature seiches, the resulting standard error (cm) in ΔZ?17 (depth of 17° isotherm; an unbiased, minimum-variance estimator for main thermocline migration) was (144/n1+ 144/n2)1/2. If n1=n2= 2 ‘revisits’, the resulting CV (coefficient of variation) was 5–10% for ΔZ?17 accompanying major individual cold fronts in early summer. 3. When estimating Kz, the vertical eddy conductivity, the most important component of error relates to ΔHz, the change in heat content below depth z. The error in ΔHz is minimized in the same manner as for lake-mean isotherm depth. Using the Mendota sampling design described above, the RMS error in ΔHz decreases from ~90 cal cm?2 in the upper metalimnion to ~35calcm?2 near the base of the metalimnion. For seven take stations, and n1= n2= 2, Kz can be estimated with a CV ~10% bracketing a single major cold front. The CV decreases approximately as ΔHz?1, hence is roughly proportional to Δt?1 or to (cumuiative windpower)?1.  相似文献   

2.
Carlos Alemn 《Biopolymers》1994,34(7):841-847
A quantum mechanical study to compare the ability of α-aminoisobutyric acid (Aib), de-hydroalanine (ΔAla), and alanine (Ala) residues to stabilize helical conformations has been performed. To address the study, the oligopeptides Xn (X = Aib, ΔAla and Ala), where n varies from 1 to 6, were computed with the AM1 semiempirical method. The results show that the residues modified at the Cα carbon atom, Aib and ΔAla, are better helical formers than Ala. Thus, a cooperative energy effect was found for both residues, and especially for ΔAla. These terms permit the understanding the different conformational behaviors between Ala and its Cα-modified residues Aib and ΔAla. This trend is important for de novo protein design, where Aib and ΔAla must be considered useful residues in the design of synthetic helical motifs. © 1994 John Wiley & Sons, Inc.  相似文献   

3.
N‐alkylated trans‐diaziridines are an intriguing class of compounds with two stereogenic nitrogen atoms which easily interconvert. In the course of our investigations of the nature of the interconversion process via nitrogen inversion or electrocyclic ring opening ring closure, we synthesized and characterized the three constitutionally isomeric diaziridines 1,2‐di‐n‐propyldiaziridine 1 , 1‐isopropyl‐2‐n‐propyldiaziridine 2 , and 1,2‐diisopropyldiaziridine 3 to study the influence of the substituents on the interconversion barriers. Enantiomer separation was achieved by enantioselective gas chromatography on the chiral stationary phase Chirasil‐β‐Dex with high separation factors α (1‐isopropyl‐2‐n‐propyldiaziridine: 1.18; 1, 2‐diisopropyldiaziridine: 1.24; 100°C 50 kPa He) for the isopropyl substituted diaziridines. These compounds showed pronounced plateau formation between 100 and 150°C, and peak coalescence at elevated temperatures. The enantiomerization barriers ΔG? and activation parameters ΔH? and ΔS? were determined by enantioselective dynamic gas chromatography (DGC) and direct evaluation of the elution profiles using the unified equation implemented in the software DCXplorer. Interestingly, 1‐isopropyl‐2‐n‐propyldiaziridine and 1,2‐diisopropyldiaziridine exhibit similar high interconversion barriers ΔG? (100°C) of 128.3 ± 0.4 kJ mol?1 and 129.8 ± 0.4 kJ mol?1, respectively, which indicates that two sterically demanding substituents do not substantially increase the barrier as expected for a distinct nitrogen inversion process. Chirality, 2010. © 2009 Wiley‐Liss, Inc.  相似文献   

4.
The thermodynamic parameters, ΔH′, ΔG′, and ΔS′, and the stoichiometry for the binding of the substrate 2′-deoxyuridine-5′-phosphate (dUMP) and the inhibitor 5-fluoro-2′-deoxyuridine-5′-phosphate (FdUMP) to Lactobacillus casei thymidylate synthetase (TSase) have been investigated using both direct calorimetric methods and gel filtration methods. The data obtained show that two ligand binding sites are available but that the binding of the second mole of dUMP is extremely weak. Binding of the first mole of dUMP can best be illustrated by dUMP + TSase + H+?(dUMP-TSase-H+). [1] The enthalpy, ΔH1′, for reaction [1] was measured directly on a flow modification of a Beckman Model 190B microcalorimeter. Experiments in two different buffers (I = 0.10 m) show that ΔH1′ = ?28 kJ mol?1 and that 0.87 mol of protons enters into the reaction. Analysis of thermal titrations for reaction [1] indicates a free energy change of ΔG1′ = ?30 kJ mol?1 (K1 = 1.7 × 105 m?1). From these parameters, ΔS1′ was calculated to be +5 J mol?1 degree?1, showing that the reaction is almost totally driven by enthalpy changes. Gel filtration experiments show that at very high substrate concentrations, binding to a second site can be observed. Gel filtration experiments performed at low ionic strength (I = 0.05 m) reveal a stronger binding, with ΔG1′ = ?35 kJ mol?1 (K1 = 1.2 × 106 m?1), suggesting that the forces driving the interaction are, in part, electrostatic. Addition of 2-mercaptoethanol (0.10 m) had the effect of slightly increasing the dUMP binding constant. Binding of FdUMP to TSase is best illustrated by 2FdUMP + TSase + nHH+?FdUMP2 ? TSase ? (H+)nH. [2] The enthalpy for this reaction, ΔH2, was also measured calorimetrically and found to be ?30 kJ mol?1 with nH = 1.24 at pH 7.4 Assuming two FdUMP binding sites per dimer as established by Galivan et al. [Biochemistry15, 356–362 (1976)] our calorimetric results indicate different binding energies for each site. Based on the binding data, a thermodynamic model is presented which serves to rationalize much of the confusing physical and chemical data characterizing thymidylate synthetase.  相似文献   

5.
Δ53β hydroxysteroid dehydrogenase activity transforms biologically inactive Δ53β hydroxy steroids into the active Δ43-keto products (e.g. pregnenolone to progesterone). Using a cytochemical procedure which allows for the continuous microdensitometric monitoring of an enzyme reaction as it proceeds and a well described cytochemical assay for Δ53β HSD we have analysed the initial velocity rates (Vo) for dehydroepiandrosterone (DHEA) binding to this enzyme in regressing (i.e. 20α hydroxy steroid dehydrogenase positive) corpus luteum (CL) cells in unfixed tissue sections (5 μm) of the dioestrous and proestrous rat ovary. The results are mean ± S.E.M. The relationship between DHEA concentration (0 to 50 μM) and Δ53β HSD activity in the dioestrous corpora lutea was sigmoidal and had an atypical 1/Vo versus 1/S plot, the x intercept being positive. Using a 1/Vo versus 1/S2 plot the Vmax was determined to be 1·0 ± 0·08 μmol min?1 mg?1 CL (n = 6). The Hill constant was 2·7 ± 0·02 (n = 6) suggesting a high degree of positive co-operativity for DHEA binding. The S concentration for half maximal activity was 17 ± 1 μmoles (n = 6). In the corpora lutea cells of the proestrous ovary, the Vmax for DHEA transformation was unchanged (0·95 ± 0·04 μmol min?1 mg?1, n = 3) whilst the S0·5 was significantly increased to 27 ± 0·1 (p < 0·01, n = 3). The Hill constant remained positive being 2·9 ± 0·2 (n = 3). NAD+ binding to 3β HSD in regressing corpora lutea of the proestrous ovary has been demonstrated previously to be hyperbolic and fit the classical Michaelis-Menten model.1 Extending the analysis of NAD+ binding to the regressing corpus luteum of the dioestrous rat ovary revealed similar kinetic characteristics to that seen with the proestrous enzyme, the apparent Vmax and Km being 0·84 ± 0·04 μmol min?1 mg?1 CL (n = 3) and 27 ± 7 μmol 1?1 (n = 3) respectively. The Hill constant was 1·1 ± 0·03 (n = 3), indicating no co-operativity of co-factor binding.  相似文献   

6.
Z-Dehydrophenylalanine (ΔzPhe) possessing four oligopeptides, Boc-(L -Ala-ΔzPhe-Aib)n-OCH3 (n = 1–4: Boc, t-butoxycarbonyl; Aib, α-aminoisobutyric acid), were synthesized, and their solution conformations were investigated by 1H-nmr, ir, uv, and CD spectroscopy and theoretical CD calculation. 1H-nmr (the solvent accessibility of NH groups) and ir studies indicated that all the NH groups except for those belonging to the N-terminal L -Ala-ΔzPhe moiety participate in intramolecular hydrogen bonding in chloroform. This suggests that the peptides n = 2–4 have a 4 → 1 hydrogen-bonding pattern characteristic of 310-helical structures. The uv spectra of all these peptides recorded in chloroform and in trimethyl phosphate showed an intense maximum around 276 nm assigned to the ΔzPhe chromophores. The corresponding CD spectra of the peptides n = 2–4 showed exciton couplets with a negative peak at longer wavelengths, whereas that of the peptide n = 1 showed only weak signals. Theoretical CD spectra were calculated for the peptides n = 2–4 of several helical conformations, on the basis of exciton chirality method. This calculation indicated that the three peptides form a helical conformation deviating from the perfect 310-helix that contains three residues per turn, and that their side chains of Δz Phe residues are arranged regularly along the helix. The center-to-center distance between the nearest phenyl pair(s) was estimated to be ~ 5.5 Å. The chemical shifts of the ΔzPhe side-chain protons (Hβ and aromatic H) for the peptides n = 2–4 indicated anisotropic shielding effect of neighboring phenyl group(s); the effect also supports a regular arrangement of the Δz Phe side chains along the helical axis. © 1993 John Wiley & Sons, Inc.  相似文献   

7.
Feces are a treasure trove in the study of animal behavior and ecology. Stable carbon and nitrogen isotope analysis allows to assess the dietary niches of elusive primate species and primate breastfeeding behavior. However, some fecal isotope data may unwillingly be biased toward the isotope ratios of undigested plant matter, requiring more consistent sample preparation protocols. We assess the impact of this potential data skew in 114 fecal samples of wild bonobos (Pan paniscus) by measuring the isotope differences (Δ13C, Δ15N) between bulk fecal samples containing larger particles (>1 mm) and filtered samples containing only small particles (<1 mm). We assess the influence of fecal carbon and nitrogen content (ΔC:N) and sample donor age (subadult, adult) on the resulting Δ13C, Δ15N values (n = 228). Additionally, we measure the isotope ratios in three systematically sieved fecal samples of chimpanzees (Pan troglodytes verus), with particle sizes ranging from 20 μm to 8 mm (n = 30). We found differences in fecal carbon and nitrogen content, with the smaller fecal fraction containing more nitrogen on average. While the Δ13C values were small and not affected by age or ΔC:N, the Δ15N values were significantly influenced by fecal ΔC:N, possibly resulting from the differing proportions of undigested plant macroparticles. Significant relationships between carbon stable isotope ratios (δ13C) values and %C in large fecal fractions of both age groups corroborated this assessment. Δ15N values were significantly larger in adults than subadults, which should be of concern in isotope studies comparing adult females with infants to assess breastfeeding. We found a random variation of up to 3.0‰ in δ13C and 2.0‰ in nitrogen stable isotope ratios within the chimpanzee fecal samples separated by particle sizes. We show that particle size influences isotope ratios and propose a simple, cost-effective filtration method for primate feces to exclude larger undigested food particles from the analysis, which can easily be adopted by labs worldwide.  相似文献   

8.
The thermal triple helix–coil transition of covalently bridged collagenlike peptides with repeating sequences of (Ala-Gly-Pro)n, n = 5–15, was studied optically. The peptides were soluble in water/acetic acid (99:1) and were found to form triple-helical structures in this solvent system beginning with n = 8. The thermodynamic analysis of the transition equilibrium curves for n = 9–13 yielded the parameters ΔH°s = ?7.0 kJ per tripeptide unit, ΔS°s = ?23.1 J deg?1 mol?1 per tripeptide unit for the coil-to-helix transition, and the apparent nucleation parameter σ ? 5 × 10?2. It was suggested through double-jump temperature experiments that the rate-limiting step during refolding is not only influenced by the difficulties of nucleation, but also by cistrans isomerization of the Gly-Pro peptide bond.  相似文献   

9.
The thermodynamic parameters, log β, ΔH and ΔS, for formation of lanthanide-1-hydroxy-4,7- disulfo-2-naphthoic acid complexes have been determined at 25 °C in 0.10 M NaClO4 solutions by potentiometric and calorimetric titrations. Under the experimental conditions, the data can be explained with the formation of LnL, LnL25− and LnHL complexes (H2L2− = 1-hydroxy-4,7-disulfo-2- naphthoic acid anion). At pH < 3 the LnHL complex is the major species, whereas by increasing pH the formation of LnLn3−4n complexes becomes more important. The data are compared to the comparable data for complexing by aromatic carboxylic acids.  相似文献   

10.
1. Stable isotopes of nitrogen are useful for quantifying the trophic structure of food webs, but only if the variation in trophic enrichment (ΔN), which is the difference in δ15N between a consumer and its food, is small relative to the value of ΔN itself. 2. We examined the sources of variation in zooplankton ΔN by measuring the trophic enrichment (ΔN) of seven species of freshwater cladocerans, and by testing for an effect of age and temperature on the ΔN of Daphnia pulicaria. 3. We found that ΔN was similar among Cladocera and was not correlated with body size. Overall, the ΔN for D. pulicaria was 1.4‰ (SE = 0.69, n = 57), as was expected for the detritus diet that we used in our experiments. We found no effect of temperature (15–25 °C) on ΔN, but found that ΔN of D. pulicaria increased with increasing age (10–30 days). 4. We developed a new method to test for trophic‐level variation in a group of consumers that explicitly accounts for the uncertainty in ΔN. Using this approach, we confirmed that natural assemblages of zooplankton feed at several trophic levels in lake food webs.  相似文献   

11.
Nine sterols, most showing Δ5- or Δ5,22-unsaturation, were identified in the marine diatom Biddulphia sinensis. One sterol, cholesta-5,22E-dien-3β-ol, comprised 70–80% of the total sterols which is the first such predominance noted in a diatom. The only Δ7-sterol detected was cholest-7-en-3β-ol and this was a very minor component. A sterol showing unusual side-chain alkylation,23,24-dimethylcholesta-5,22E-dien-3β-ol, was identified for the first time in a diatom. Total fatty acids exhibited a predominance of Δ9- 16:1, 14:0, 20:5 and 16:0, typical of diatoms, although the proportions of these acids were found to vary with culture maturity. n-Heneicosahexaene was the major hydrocarbon together with a small amount of squalene.  相似文献   

12.
Alkyl gallate, which is known as an antioxidant, intensively inhibited Δ5 and Δ6 desaturation in both rat liver microsomes and an arachidonic acid-producing fungus Mortierella alpina 1S-4. The rat liver microsomal Δ5 and Δ6 desaturases were inhibited by gallic acid esterified with alcohols with various numbers of carbons, suggesting that the necessary structure in an esterified alcohol for the inhibition is not so strict. Among the three hydroxy groups in gallic acid, the m-hydroxy group was shown to be the necessary structure. Kinetic analyses revealed that propyl gallate is a noncompetitive inhibitor of Δ5 desaturase (Ki = 2.6 · 10−5 M) and Δ6 desaturase (Ki = 1.7 · 10−4 M). These data indicate that alkyl gallate is a new type of desaturase inhibitor and different from known natural inhibitors, i.e., sesamin and curcumin.  相似文献   

13.
The desmethyl sterol composition of the oomycete Dictyuchus monosporus is unusual in that it is a mixture of 56.9 % Δ5-sterols and 42.6 % Δ7-sterols. The Δ5-sterols are cholesterol, 24 methylenecholesterol and fucosterol; the Δ7-sterols are cholest-7-enol, ergosta-7,24(28)-dienol and stigmasta-7,E-24(28)-dienol. Stigmasta-7,E-24(28)-dienol, is identified for the first time from natural sources. In addition, traces of lanosterol are present.  相似文献   

14.
Application of stable isotope data to trophic studies requires understanding of factors influencing the isotopic discrimination factor (Δ) between consumers and their prey resources. This is missing for many omnivorous species, despite their diet and environment potentially impacting Δ. The effects of temperature, diet (including formulated feeds) and tissue type on Δ13C and Δ15N were thus tested experimentally. A temperature experiment exposed three species to identical diets at 18 and 23°C, whereas a diet experiment exposed one species to four diets at 18°C. At 23°C, C:N ratios, Δ13C and Δ15N were generally elevated versus 18°C. After lipid correction, tissue/species-specific differences at 23°C in Δ13C and Δ15N were up to 0.73 and 0.54‰ higher, respectively. Across the four diets, there were also significant differences in Δ13C and Δ15N between a natural diet and diets based on formulated feeds. Δ13C and Δ15N of muscle were 1.51 to 2.76‰ and 3.13 to 5.44‰, respectively. Highest Δ for both isotopes was from a formulated feed based on plant material that resulted in lower dietary protein content and quality. Thus, diet and environment fundamentally affected the isotopic discrimination factors and these factors require consideration within trophic studies based on stable isotopes.  相似文献   

15.
The mutant STE 1 was isolated by screening an ethylmethane sulfonate (EMS)-mutagenized population of Arabidopsis thaliana which consisted of 22 000 M2 plants divided into 1100 pools of 20 plants by gas chromatography of sterols extracted from small leaf samples. STE 1 was characterized by the accumulation of three Δ7-sterols concomitantly with the decrease of the three corresponding Δ5-sterols which are the end products of the sterol pathway in wild-type leaves. The structure of these Δ7-sterols was determined after two steps of purification on HPLC, by gas chromatography coupled with mass spectrometry (GC-MS) and proton nuclear magnetic resonance spectrometry (1H-NMR). The accumulation of Δ7-sterols suggested that the mutant is deficient in the activity of the Δ7-sterol-C-5-desaturase. Genetic analysis showed that the accumulation of Δ7-sterols was due to a single recessive nuclear mutation. The mutant line STE 1 was backcrossed four times to the wild-type. The resulting STE 1 plants had wild-type morphology and set seeds normally, suggesting that the Δ7-sterols in STE 1 are good surrogates of physiologically active Δ5-sterols to sustain normal development. STE 1 roots were transformed with the Saccharomyces cerevisiae ERG 3 gene encoding the Δ7-sterol-C-5-desaturase under the control of the CaMV 35S promoter. Seven transgenic STE 1 root-derived calli showed an increase in Δ5-sterols and a concomitant decrease in Δ7-sterols in comparison with STE 1 untransformed root-derived calli. Northern blot analysis using the ERG 3 probe showed a strong expression of ERG 3 in three of the seven transgenic calli. These results suggest that the accumulation of Δ7-sterols in the STE 1 mutant is due to a deficiency of the Δ7-sterol-C-5-desaturation step in the plant sterol biosynthesis pathway.  相似文献   

16.
Nitrogen stable isotope natural abundance data are often used in trophodynamic research. The assumed nitrogen diet-tissue fractionation (Δδ15N) determines conclusions about trophic level, potential food sources and ontogenetic diet shifts. Δδ15N is usually assumed to be 3.0-3.4‰ per trophic level and unaffected by the size or age of animals or their environment. To assess the effects of body size, experimental duration and environmental conditions on fish tissue Δδ15N, two populations of European sea bass (Dicentrarchus labrax) were reared on constant diets of dab (Limanda limanda) muscle or sandeel (Ammodytes marinus) for 2 years under natural light and temperature regimes. Bass were sampled at approximately monthly intervals to determine Δδ15N for muscle, heart and liver tissue. Mean values of Δδ15N were 3.83‰, 3.54‰, 2.05‰ (sandeel diet) and 3.98‰, 3.32‰, 1.95‰ (dab diet) for muscle, heart and liver tissue respectively. The assumption that fractionation was independent of body mass was upheld for muscle and heart tissue, but not for liver. Time effects on muscle Δδ15N were explainable by a sinusoidal function with a period of 1 year and wave height ∼ 0.3‰. Time resulted in increases in heart δ15N and decreases in liver δ15N which were small compared to background variation, equating to 1/6 of a trophic level over 2 years, and unlikely to have great significance in ecological studies. Heart and liver δ15N were also affected by temperature probably reflecting the metabolic functions of these tissues and their associated rates of turnover. However in heart the explanatory power of temperature appeared tied to that of time. Although the Δδ15N for bass muscle on both diets approached 4‰, the Δδ15N values from this study, when combined with those from the literature, suggest that where fish species specific data are not available, a mean Δδ15N for fish muscle of 3.2‰ should be applied (mean white muscle Δδ15N = 3.15). The literature based mean Δδ15N for whole fish was lower than that of white muscle suggesting that a separate Δδ15N (2.9‰) should be applied when sampling whole fish.  相似文献   

17.
An early step in the morphogenesis of the double-stranded DNA (dsDNA) bacteriophage HK97 is the assembly of a precursor shell (prohead I) from 420 copies of a 384-residue subunit (gp5). Although formation of prohead I requires direct participation of gp5 residues 2-103 (Δ-domain), this domain is eliminated by viral protease prior to subsequent shell maturation and DNA packaging. The prohead I Δ-domain is thought to resemble a phage scaffolding protein, by virtue of its highly α-helical secondary structure and a tertiary fold that projects inward from the interior surface of the shell. Here, we employ factor analysis of temperature-dependent Raman spectra to characterize the thermostability of the Δ-domain secondary structure and to quantify the thermodynamic parameters of Δ-domain unfolding. The results are compared for the Δ-domain within the prohead I architecture (in situ) and for a recombinantly expressed 111-residue peptide (in vitro). We find that the α-helicity (∼ 70%), median melting temperature (Tm = 58 °C), enthalpy (ΔHm = 50 ± 5 kcal mol− 1), entropy (ΔSm = 150 ± 10 cal mol− 1 K− 1), and average cooperative melting unit (〈nc〉 ∼ 3.5) of the in situ Δ-domain are altered in vitro, indicating specific interdomain interactions within prohead I. Thus, the in vitro Δ-domain, despite an enhanced helical secondary structure (∼ 90% α-helix), exhibits diminished thermostability (Tm = 40 °C; ΔHm = 27 ± 2 kcal mol− 1; ΔSm = 86 ± 6 cal mol− 1 K− 1) and noncooperative unfolding (〈nc〉 ∼ 1) vis-à-vis the in situ Δ-domain. Temperature-dependent Raman markers of subunit side chains, particularly those of Phe and Trp residues, also confirm different local interactions for the in situ and in vitro Δ-domains. The present results clarify the key role of the gp5 Δ-domain in prohead I architecture by providing direct evidence of domain structure stabilization and interdomain interactions within the assembled shell.  相似文献   

18.
δ13C data are often used in trophodynamic research where diet-tissue fractionation (Δδ13C) is assumed to be 0-1‰ per trophic level and unaffected by the size of animals or their environment. Variation in Δδ13C will influence conclusions about food sources, energy pathways and trophic level. To assess the effects of body size, age and environmental conditions on Δδ13C, European sea bass (Dicentrarchus labrax) were reared on constant diets of dab (Limanda limanda) or (Ammodytes marinus) for 2years under natural environmental regimes. Bass were sampled approximately monthly to determine Δδ13C for muscle, heart and liver tissue and were 1.66‰, − 0.18‰, − 1.77‰ (sandeel diet) and 1.34‰, − 1.18‰, − 1.75‰ (dab diet) respectively. Arithmetic lipid correction increased Δδ13C to > 2‰ for muscle and liver. Δδ13C was dependent on body mass and experimental duration (age) and generally declined with weight or time even after correction for lipid content. For liver, increasing temperature increased Δδ13C. The Δδ13C estimates from this study were compared with all available published Δδ13C estimates for fish. Bass muscle Δδ13C was similar to previous estimates for fish white muscle Δδ13C (1.56 ± 1.10‰) and whole body Δδ13C (1.52 ± 1.13‰). Fractionations derived in this study, combined with those from the literature, support the use of diet-tissue fractionation values of between 1‰-2‰ for δ13C, rather than the commonly used 0‰ − 1‰. For muscle Δδ13C, 1.5‰ is appropriate.  相似文献   

19.
Heat conduction solution enable rapid determination of the heats of aerobic and anaerobic metabolism of substrate by microorganisms. Aliquots of 1.0 ml cell suspension, 5 × 109 cell/ml, were mixed with a few dozen nmol substrate contained in 0.5 ml, under a controlled atmosphere of air, O2, or N2. At these substrate concentration, with adapted microorganisms, metabolism and its heat generation are usually complete within 300 to 600 sec. The raw data yield ΔHapp values. The ΔHapp were determined in the range 0.001 to 0.010% substrate, and extrapolated (limit substrate concentration →0), to yield Δ0H?, the limiting differential molar heat of metabolism. The Δ0H? values express the heat generated when there is rapid metabolism but little new growth, minimal contribution by H+ transfer from metabolites, and maintenance of aerobicity or anaerobicity as specified. Escherichiacoli B/5 was used for aerobic and anaerobic combustion of eight sugars. Pseudomonas multivorans, and an Acinetobacter, strain B-1, were used for aerobic metabolism of benzene, toluene, naphthalene, and a methylnaphthalene. The larger heats of combustion of the hydrocarbons enable the use of aqueous solutions of hydrocarbons well below their solubility limits. The quotient Δ0H?/n (n = atoms carbon/molecule substrate) varies from (-)36 to (-)67 kcal/mol carbon for the sugars. The most reduced sugar yields the largest exothermic heats. The quotient varies from (-)27 to (-)81 kcal/mol carbon for the aromatic hydrocarbons. Comparison of the calorimetric heats of metabolism of those from total aerobic combustion in aquo (where available) give measure of the efficiencies with which the heat contents of the aqueous substrate are used by the bacteria.  相似文献   

20.
To further understand the mechanism of action and pharmacokinetics of medroxyprogesterone acetate (MPA), the binding interaction of MPA with bovine serum albumin (BSA) under simulated physiological conditions (pH 7.4) was studied using fluorescence emission spectroscopy, synchronous fluorescence spectroscopy, circular dichroism and molecular docking methods. The experimental results reveal that the fluorescence of BSA quenches due to the formation of MPA–BSA complex. The number of binding sites (n) and the binding constant for MPA–BSA complex are ~1 and 4.6 × 103 M?1 at 310 K, respectively. However, it can be concluded that the binding process of MPA with BSA is spontaneous and the main interaction forces between MPA and BSA are van der Waals force and hydrogen bonding interaction due to the negative values of ΔG0, ΔH0 and ΔS0 in the binding process of MPA with BSA. MPA prefers binding on the hydrophobic cavity in subdomain IIIA (site II′′) of BSA resulting in a slight change in the conformation of BSA, but BSA retaining the α‐helix structure. Copyright © 2016 John Wiley & Sons, Ltd.  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