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.
A comparison of the thermoregulation of water foraging wasps (Vespula vulgaris, Polistes dominulus) under special consideration of ambient temperature and solar radiation was conducted. The body surface temperature of living and dead wasps was measured by infrared thermography under natural conditions in their environment without disturbing the insects’ behaviour. The body temperature of both of them was positively correlated with Ta and solar radiation. At moderate Ta (22–28 °C) the regression lines revealed mean thorax temperatures (Tth) of 35.5–37.5 °C in Vespula, and of 28.6–33.7 °C in Polistes. At high Ta (30–39 °C) Tth was 37.2–40.6 °C in Vespula and 37.0–40.8 °C in Polistes. The thorax temperature excess (TthTa) increased at moderate Ta by 1.9 °C (Vespula) and 4.4 °C (Polistes) per kW−1 m−2. At high Ta it increased by 4.0 °C per kW−1 m−2 in both wasps. A comparison of the living water foraging Vespula and Polistes with dead wasps revealed a great difference in their thermoregulatory behaviour. At moderate Ta (22–28 °C) Vespula exhibited distinct endothermy in contrast to Polistes, which showed only a weak endothermic activity. At high Ta (30–39 °C) Vespula reduced their active heat production, and Polistes were always ectothermic. Both species exhibited an increasing cooling effort with increasing insolation and ambient temperature.  相似文献   

2.
Standard maize starch was hydrothermally treated by Instantaneous Controlled Pressure Drop (DIC) process at three pressure levels (1, 2 and 3 bar) corresponding to the temperatures of 100, 122 and 135 °C (at 13–27% moisture), respectively. The structural effects of various hydrothermal conditions were examined with differential scanning calorimetry (DSC) and wide-angle X-ray diffraction. In order to understand the changes that occur during DIC treatment, melting endotherms of native maize starch at various moisture contents were determined. The gelatinization temperatures of DIC treated standard maize starch increased with DIC treatment. The transition temperatures (To, Tp) are closely related to the combined effect of pressure and processing time. At approximately 10 min of processing time, To and Tp were 65.7 and 72.3, 68.8 and 73.6 °C, 74.8 and 79.8 °C for pressure levels of 1, 2 and 3 bar, respectively (against 63.1 and 69.6 °C for native starch). DIC treatment narrowed the gelatinization temperature range and decreased gelatinization enthalpy (ΔH), as the severity of processing conditions increased. ΔH decreased from 11.4 J g−1 (native) to 11.0 (1 bar), 9.0 (2 bar) and 1.7 J g−1 (3 bar) for treated maize starch during approximately 10 min. Relative crystallinity of hydrothermally treated starch decreased with increasing DIC conditions. The A-type crystalline pattern was progressively lost (at pressure level 2 bar) and substituted by the Vh-type X-ray diffraction pattern, corresponding to the formation of amylose–lipid complexes. For severe DIC conditions (pressure level of 3 bar), the substitution was completed. Microscopic observations revealed progressive loss of the birefringence of DIC treated starch granules except at low pressure (1 bar), while the integrity of starch granules was preserved for all the conditions. These modifications that reveal important changes in the crystalline organization of the starch granules are related to their functional properties.  相似文献   

3.
The phase transition of a complex, biological membrane was studied in relation to two variables: high pressure and the alcohols, pentanol and benzyl alcohol, in order to determine whether earlier studies of defined phospholipid bilayers may be applied to natural membranes. The isolated membrane of A choleplasma laidlawii was used and three independent methods were chosen to detect the transition; optical transmission (giving Tt, an index of the end-of-melting temperature); fluorescence polarisation (giving Tp, the temperature midway through the change in polarisation which characterises the transition) and differential thermal analysis giving a record of the temperature range occupied by the endothermic process. Pressure increased Tt by 0.017 K·atm−1 and Tp by 0.016 K·atm−1, consistent with dT/dP = T·ΔV/ΔH. Pentanol (and benzyl alcohol) lowered Tt, Tp and the temperature of the endotherm seen with differential thermal analysis. Thus the membrane transition responds to pressure and alcohols in agreement with thermodynamic theory.  相似文献   

4.
Pigeon flight in a wind tunnel   总被引:2,自引:0,他引:2  
Summary Core temperatureT c, breast temperatureT s–br and leg temperatureT s–1 were measured simultaneously in pigeons during rest and flight in a wind tunnel, using thermistors.MeanT c at rest is 39.8±0.7°C and is independent of ambient temperatureT a (10–30°C). In the first minutes of flight,T c increases to 1.5–3.0°C above resting level and remains at this higher level. This hyperthermia increases withT a (v=const.). It is±constant in the lowT a range (10.6–13.9°C) at flight speeds v ranging from 10–18 m s–1 and normal body mass, but increases with v and elevated body mass in the highT a range (23.7–28.8°C). T s–1 is adapted toT a at rest and increases in flight up to 3–4°C belowT c. This increase inT s–1 is linear toT a. T s–br is always lower thanT c, in extreme cases reaching restingT c in flight.Supported by the Deutsche Forschungsgemeinschaft  相似文献   

5.
Crystal structures of cyclomaltohexaose (α-cyclodextrin) complexes with p-chlorophenol and p-cresol have been determined by single-crystal X-ray diffraction studies. The space group of the α-cyclodextrin–p-chlorophenol complex is P212121 with unit cell dimensions of a=15.299(3), b=24.795(5), c=13.447(5) Å, and that of the α-cyclodextrin–p-cresol complex is P21 with unit cell dimensions of a=7.927(7), b=13.568(7), c=24.54(1) Å, β=90.41(8)°. In spite of the similar structures of guest molecules, both complexes have different inclusion modes and packing structures.  相似文献   

6.
This study examined the significance of ECG-derived indexes in quantifying ventricular repolarization dispersion (VRD) given its value as a risk marker for severe myocardial arrhythmia. Multilead ECG recordings from an isolated rabbit heart model, including control and globally increased VRD (IVRD) beats, were studied. The IVRD was induced by supplying d-Sotalol (DS) or premature ventricular stimulation (PVS). ECG indexes came from (a) the absolute ECG summation signal, from which we obtained the amplitude and area of the T-wave, and the T-wave width (TW), which we consider as IVRD indexes, and (b) the Singular Value Decomposition (SVD) of the ECG, from which the θPT (angle between the first SVD principal axis and the repolarization axis), T-wave residuum (TWR), T-wave morphology dispersion (TMD), unnormalized TMD (UTMD), and θRT (the angle between the depolarization and the repolarization vectors) were estimated as IVRD indexes. Results were compared with the classical QT-based VRD indexes (σQTe, standard deviation of QT end). The main results are TW: 78.0±10.3 vs. 133.6±29.6 ms, for control vs. IVRD generated using DS, p<0.005 and 95.2±7.9 vs. 118.5±15.7 ms when PVS was used, p<0.007; σQTe: gives 6.5±1.4 vs. 11.6±1.9 ms, for DS p<0.007 and 7.6±2.2 vs. 13.0±3.4 ms for PVS, p<0.007; respectively. θPT: 35±51° vs. 117±49°, p<0.009 in DS. We concluded that globally induced IVRD is well reflected by the TW parameter, being the most sensitive of the studied ones. The IVRD can also be quantified by using the θPT index.  相似文献   

7.
The molecular conformation of the monoclinic crystalline polymorph of prostaglandin A1 has been determined by X-ray diffraction techniques. The space group is P21 with a = 13.637 (2), b = 7.567 (1), I c = 10.576 (2) Å, β = 107.37 (3)°; Dc = 1.073 g·cm−3 for Z = 2. The molecular conformation is characterized by the nearly parallel arrangement of the C1–C7 and C13–C20 side chains, with a general flattening of the overall structure when compared with the orthorhombic polymorph. The cyclopentenone moiety assumes a C8 envelope conformation with C8 and O9 displaced +0.29 Å and −0.18 Å from the C9–C10=C11–C12 plane respectively. Concerted, small variations of the torsion angles, primarily about the C8–C12, C14–C15 and C16–C17 bonds, bring the monoclinic and orthorhombic conformations into coincidence.  相似文献   

8.
Galactomannans isolated from legume seed endosperms, including those of commercial interest, have been characterized by multidetection aqueous SEC. Galactomannans derived from seeds of the Faboideae subfamily had substantially higher Mw than those from Caesalpinioideae seeds (Mw,Fab = 2.4–3.1 × 106 g/mol, Mw,Caes. = 0.86–2.1 × 106 g/mol) and within the latter botanical subfamily, an apparent correlation between Mw and the degree of galactose substitution DG was found. The molar mass distributions were unimodal and differed primarily by a scale factor, with distributional widths narrower than a true Flory ‘most-probable distribution’; good fits to Schulz–Zimm model were obtained. Across subfamilies no differences were found in the exponents of [η]–M and RvM relationships (0.61 ± 0.02, 0.54 ± 0.01, respectively), the Flory chain stiffness ratio (C = 20 ± 1 (BSF analysis)), or the persistence length (Lp = 5.5 ± 0.2 nm) obtained from SEC fraction data. However, it was found that prefactors in the [η]–M and RvM relationships as well as the unperturbed parameter KΘ decrease in proportion to DG and therefore chain density. Generalized relationships incorporating galactose-dependent prefactors were therefore developed to model SEC fraction data of native galactomannans ([η]GM = (1800 ± 200) × Mo−1.61 × M0.61±0.02, Rv,GM = 0.63 ± 0.05 × Mo−0.54 × M0.54±0.01) as well as lower-M fractions obtained by ultrasonication ([η]GM = (730 ± 100) × Mo−1.71 × Mw0.71±0.02, Rv,GM = 0.49 ± 0.05 × Mo−0.57 × Mw0.57±0.01, M ≈ 1 × 105-native). As a consequence of this dependence and the observed patterns in molar mass variation, [η] varies within a narrow range for galactomannans as a whole despite substantial Mw differences.  相似文献   

9.
1. The fat mouse Steatomys pratensis natalensis (mean body mass 37.4±0.43 (se)) has a low euthermic body temperature Tb=30.1–33.8 °C and a low basal metabolic rate (BMR)=0.50 ml O2 g−1 h−1.
2. Below an ambient temperature (Ta)=15 °C, the mice were hypothermic.
3. The lowest survivable Ta=10 °C.
4. Torpor is efficient in conserving energy between Ta=15–30 °C, below Ta=15 °C, the mice arouse.
5. Euthermic and torpid mice were hyperthermic at Ta=35 °C.
6. Thermal conductance was 0.159 ml O2 g−1 h−1 °C−1, 98.8% of the expected value.
7. Non-shivering thermogenesis (NST) was 2.196 ml O2 g−1 h−1 (3.69×BMR).
8. Maximal oxygen consumption, however, was 3.83 ml O2 g−1 h−1 (6.44×BMR), indicating that other methods of heat production are additive.
9. Because fat mice conserve energy by torpor only between Ta=15–30 °C, we suggest that torpor may be a more important mechanism for surviving food shortages than for surviving cold weather.
Keywords: Steatomys pratensis natalensis; Metabolism; Torpor; Fat mouse  相似文献   

10.
11.
The objective of this study was to examine the composition and molecular structure of the endosperm cell walls (CW) derived from barley grain grown in three environments in Canada, and differing in grain hardness, protein, and total β-glucan contents. The endosperm CW were isolated from barley, cv. Metcalfe, grown in Davidson, SK (Sample A), Hythe, AB (sample B), and Hamiota, MB (sample C). The CW were sequentially extracted with water at 65 oC, saturated Ba(OH)2, again with water at 25 oC, and 1 M NaOH, resulting in fractions designated WE65, BaE, Ba/WE, and NaE, respectively. The monosaccharide analysis indicated the presence of β-glucans, arabinoxylans, and small amounts of arabinogalactans, glucomannans, and xyloglucans. Cellulose was detected in the CW remnants. The CW of sample A, exhibiting a lower grain hardness than sample B, contained the lowest amount of β-glucans, but the highest amount of arabinoxylans and the mannose-containing polysaccharides. The CW of sample C, characterized by very high protein content in the grain, contained the highest amount of β-glucans and the lowest amount of other polysaccharides. Polysaccharides in the CW of sample B, exhibiting the highest grain hardness, were characterized by the highest weight average molecular weights (Mw). β-Glucans in the CW of Sample B showed the highest ratio of DP3/DP4 and the longest cellulosic fragments in the polymeric chains. Of the three barley samples, arabinoxylans in the endosperm CW of sample A exhibited the lowest degree of branching, the highest amount of unsubstituted Xyl residues, and the highest ratio of singly to doubly substituted Xylp. The highest water solubility of the CW of sample C was associated with the highest concentration of β-glucans, the lowest DP3/DP4 ratio, and the lowest Mw of the polymeric constituents. Arabinoxylans with the lowest amount of doubly substituted but the highest amount of unsubstituted xylose residues and long sequences of unsubstituted xylan regions were found in the NaE fractions. The NaE fractions showed a high ratio of →4)-Glcp-(1→ to →3)-Glcp-(1→ linkages and some →4)-Manp-(1→ linkages, indicating a high level of long cellulosic regions in β-glucan chains and the presence of glucomannans.  相似文献   

12.
Summary Reaction kinetic analysis of the electrical properties of the electrogenic Cl pump inAcetabularia has been extended from steady-state to nonsteady-state conditions: electrical frequency responses of theAcetabularia membrane have been measured over the range from 1 Hz to 10 kHz at transmembrane potential differences across the plasmalemma (V m ) between –70 and –240 mV using voltage-clamp techniques. The results are well described by an electrical equivalent circuit with three parallel limbs: a conventional membrane capacitancec m , a steadystate conductanceg o (predominantly of the pump pathway plus a minor passive ion conductance) and a conductanceg s in series with a capacitancec p which are peculiar to the temporal behavior of the pump. The absolute values and voltage sensitivities of these four elements have been determined:c m of about 8 mF m–2 turned out to be voltage insensitive; it is considered to be normal.g o is voltage sensitive and displays a peak of about 80 S m–2 around –180 mV. Voltage sensitivity ofg s could not be documented due to large scatter ofg s (around 80 S m–2).c p behaved voltage sensitive with a notch of about 20 mF m–2 around –180 mV, a peak of about 40 mF m–2 at –120 mV and vanishing at –70 mV. When these data are compared with the predictions of nonsteady-state electrical properties of charge transport systems (U.-P. Hansen, J. Tittor, D. Gradmann, 1983,J. Membrane Biol. in press), model A (redistribution of states within the reaction cycle) consistently provides magnitude and voltage sensitivity of the elementsg o ,g s andc p of the equivalent circuit, when known kinetic parameters of the pump are used for the calculations. This analysis results in a density of pump elements in theAcetabularia plasmalemma of about 50 nmol m–2. The dominating rate constants for the redistribution of the individual states of the pump in the electric field turn out to be in the range of 500 sec–1, under normal conditions.  相似文献   

13.
The reproductive and developmental biology of Gonatocerus ashmeadi Girault, a parasitoid of the glassy-winged sharpshooter Homalodisca coagulata (Say), was determined at five constant temperatures in the laboratory: 15; 20; 25; 30; 33 °C. At 30 °C, G. ashmeadi maintained the highest successful parasitism rates with 46.1% of parasitoid larvae surviving to adulthood. Lifetime fecundity was greatest at 25 °C and fell sharply as temperature either increased or decreased around 25 °C. Temperature had no effect on sex ratio of parasitoid offspring. Mean adult longevity was inversely related to temperature with a maximum of 20 days at 15 °C to a minimum of eight days at 33 °C. Developmental rates increased nonlinearly with increasing temperatures. Developmental rate data were fitted with the modified Logan model for oviposition to adult development times across each of the five experimental temperatures to determine optimal and upper lethal temperature thresholds. The lower developmental threshold estimated by the Logan model and linear regression were 1.10 and 7.16 °C, respectively. Linear regression of developmental rate for temperatures 15–30 °C indicated that 222 degree-days were required above a minimum threshold of 7.16 °C to complete development. A temperature of 37.6 °C was determined to be the upper development threshold with optimal development occurring at 30.5 °C. Demographic parameters were calculated and pseudo-replicates for intrinsic rate of increase (rm), net reproductive rates (Ro), generation time (Tc), population doubling time (Td), and finite rate of increase (λ) were generated using the bootstrap method. Mean bootstrap estimates of demographic parameters were compared across temperatures using ANOVA and nonlinear regression.  相似文献   

14.
Thermoregulatory sweating [total body (m sw,b), chest (m sw,c) and thigh (m sw,t) sweating], body temperatures [oesophageal (T oes) and mean skin temperature (T sk)] and heart rate were investigated in five sleep-deprived subjects (kept awake for 27 h) while exercising on a cycle (45 min at approximately 50% maximal oxygen consumption) in moderate heat (T air andT wall at 35° C. Them sw,c andm sw,t were measured under local thermal clamp (T sk,1), set at 35.5° C. After sleep deprivation, neither the levels of body temperatures (T oes,T sk) nor the levels ofm sw, b,m sw, c orm sw, t differed from control at rest or during exercise steady state. During the transient phase of exercise (whenT sk andT sk,1 were unvarying), them sw, c andm sw, t changes were positively correlated with those ofT oes. The slopes of them sw, c versusT oes, orm sw, t versusT oes relationships remained unchanged between control and sleep-loss experiments. Thus the slopes of the local sweating versusT oes, relationships (m sw, c andm sw, t sweating data pooled which reached 1.05 (SEM 0.14) mg·cm–2·min–1°C–1 and 1.14 (SEM 0.18) mg·cm–2·min–1·°C–1 before and after sleep deprivation) respectively did not differ. However, in our experiment, sleep deprivation significantly increased theT oes threshold for the onset of bothm sw, c andm sw, t (+0.3° C,P<0.001). From our investigations it would seem that the delayed core temperature for sweating onset in sleep-deprived humans, while exercising moderately in the heat, is likely to have been due to alterations occurring at the central level.  相似文献   

15.
Pulsed NMR techniques have been applied to the study of the relaxation parameters characterizing 23Na within frog striated muscle. Experiments were performed at 3°C, 22–24°C and 39°C at a Larmor frequency of 15.7 MHz; at 22–24°C, measurements were obtained both at 15.7 MHz and at 7.85 MHz.As previously reported, only a single spine-lattice relaxation time (T1) was observed, but both slow (T2)I and fast (T2)II components of the spin-spin relaxation time were measured. The effect of temperature (θ) upon (1/T1) was qualitatively similar to that reported for 23Na in free solution; (θ) did not significantly affect (1/T2) over the range of temperatures studied. (1/T2)I, and to a lesser degreee, (1/T1) exhibited a modest inverse dependence of doubtful significance on the Larmor frequency.The data are examined within the framework of a simple specific model; a conservative values in assumed for the quadrupolar coupling constant characterizing immobilized intracellular Na+. Within this framework, the results suggest that the fraction of bound ions whose molecular tumbling is severely restricted does not exceed some few percent of the total sodium population.  相似文献   

16.
The crystal structure of 1,6-anhydro-β-d-mannopyranose, C6H10O5, is orthorhombic, P212121, with a = 10.971(2), b = 13.935(3), c = 9.012(1) Å, V = 1377.76 »3 (MoKα, λ = 0.7107 Å), Z = 8, Dx = 1.563 M.gm−3, Dm = 1.565 M.gm−3. the structure was solved by MULTAN and refined to R(F) = 0.043 for 2355 reflections. The two symmetry-independent molecules in the unit cell have similar conformations, except for the orientation of one of the three hydroxyl groups. The conformation of the pyranose rings is 1C4 distorted towards Eo, and that of the anhydro rings is E. There are significant differences between the two molecules in two of the four C---O bond-lengths. These occur where there are important differences in the hydrogen-bonding environment of the oxygen atoms. The molecules are hydrogen-bonded by three linear and three bifurcated O---H···O interactions which form four-membered loops linked into infinite chains. Empirical force-field calculations with MMI-CARB reproduced the geometry of the molecules within the variations observed experimentally between the two molecules, except for a C---O bond in one of the molecules. The effect of excluding the anomeric effect from the theoretical calculations was not significant. Calculations for an intramolecularly hydrogen-bonded molecule were also carried out as a model for the molecules in a non-polar solvent.  相似文献   

17.
The structural characteristics of citric acid-treated rice starch, which was subjected to autoclaving and stored at a high temperature under acidic conditions, were investigated by high performance anion-exchange chromatography, multi-angle laser-light scattering, X-ray diffraction, differential scanning calorimetry (DSC), nuclear magnetic resonance, and scanning electron microscopy. The citric acid-treated rice starch contained chains of different lengths (DP 7−70) and had an A + V type polymorphism. The DSC thermogram of acid-treated rice starch showed a broad endothermic peak, which indicated that the structure of the acid-treated rice starch contained a number of double helices with various melting temperatures. Microscopic observation showed that the internal structure of the acid-treated starch displayed more or less spherical lumps that could be composed of short chains and amylose–lipid complex. The digestive properties of acid-treated rice starch were altered by heat processing such as cooking and autoclaving because crystalline regions were converted to amorphous regions or lamellae.  相似文献   

18.
Summary H+-coupled transport in plant and fungal cells is relatively insensitive to external pH (pH o ). H+-coupled Cl transport at the plasma membrane ofChara corallina was studied to explore the phenomena responsible for this insensitivity. Raising pH o from a control value of 7.5 to 9.0 results in a modest (2.5-fold) decline inJ max and increase inK m . Further increase in pH o results in a selective increase inJ max, in accordance with predictions from a reaction kinetic model of the transport system (Sanders, D., Hansen, U.-P., 1981.J. Membrane Biol. 58:139–153). Increase in cytosolic Cl concentration ([Cl] c ) also results in a selective decrease inJ max at pH o =7.5.Quantitative kinetic modeling of the results is not possible if it is assumed that the sole effect of pH o isvia mass action on the binding of external H+ to a transport site. If, instead, the dependence of cytosolic pH (pH c ) on pH o (Smith, F.A., 1984,J. Exp. Bot. 35:1525–1536) is taken into account along with the dependence of Cl influx on pH c (Sanders, D., 1980,J. Membrane Biol. 53:129–141), then the observed modest changes in Michaelis parameters can be accommodated by a reaction kinetic model. The quantitative parameters of the model yield respective pK a s of the internal and external H+-binding sites=7.85 and 7.2, respective dissociation constants of the internal and external Cl-binding sites=160 and 40 m, and an additional, kinetically transparent, H+-binding site with a pK a >8.0. The quantitative model independently predicts the response ofJ max andK m to acidic conditions.The results are discussed in terms of the general physiological requirement that fluxes through H+-coupled transport systems are relatively insensitive to environmental variation in pH o . It is proposed that (i) the weak (but finite) dependence of pH c on pH o , coupled with (ii) the strong dependence of H+-coupled transport on pH c are instrumental in endowing H+-coupled transport systems with a relative insensitivity to variation in pH o . This hypothesis might also explain why pH c in plants and fungi is not acutely controlled with respect to variation of pH o .  相似文献   

19.
Summary A stopped-flow rapid reaction apparatus was used for measuring changes in extracellular pH (pH o ) of red cell suspensions under conditions wheredpH o /dt was determined by the rate of HCO 3 /X exchange across the membrane (X =Cl, Br, F, I, NO 3 or SCN). The rate of the exchange at 37°C decreased forX in the order: Cl>Br>F>I>NO 3 >SCN, with rate constants in the ratios 10.860.770.550.520.31. When HCO 3 is exchanged for Cl, Br, F, NO 3 or SCN, a change in the rate-limiting step of the process takes place at a transition temperature (T T ) between 16 and 26°C. In I medium, however, no transition temperature is detected between 3 and 42°C. AlthoughT T varies withX , the activation energies both above and belowT T are similar for Cl, Br, NO 3 and F. The values of activation energy are considerably higher whenX =I or SCN. The apparent turnover numbers calculated for HCO 3 /X exchange (except forX =I) at the correspondingT T ranged from 140 to 460 ions/site ·sec for our experimental conditions. These findings suggest that: (i) HCO 3 /X exchange for allX studied takes place via the rapid anion exchange pathway; (ii) the rate of HCO 3 /X exchange is influenced by the specific anions involved in the 11 obligatory exchange; and (iii) the different transition temperatures in the Arrhenius diagrams of the HCO 3 /X exchange do not seem to be directly related to a critical turnover number, but may be dependent upon the influence ofX on protein-lipid interactions in the red blood cell membrane.  相似文献   

20.
A sensitive, robust gas chromatographic–mass spectrometric assay suitable for use in pharmacokinetic or bioequivalence studies is presented for the selective serotonin reuptake inhibitor, fluoxetine, and its major metabolite, norfluoxetine (N-desmethylfluoxetine). This method employs solid-phase extraction followed by acetylation with trifluoroacetic anhydride and analysis of the derivatives using selected ion monitoring. The lower limit of quantification was 1.0 ng/ml, and the assay was linear for both analytes from 1 to 100 ng/ml. Mean recoveries following solid-phase extraction at concentrations of 5.0, 20 and 100 ng/ml were 91% (fluoxetine) and 87% (norfluoxetine). Assay precision (as mean RSD) and accuracy (as mean relative error) for both analytes were tested at the same three nominal concentrations and were found to be within 10% in all cases. Analysis of fluoxetine concentrations in plasma samples from 18 volunteers following administration of a single 40 mg dose of fluoxetine provided the following pharmacokinetic data (mean±SD): Cmax, 32.73±9.21 ng/ml; AUC0–∞, 1627±1372 ng/ml h; Tmax, 3.08 h (median); ke, 0.022±0.007 h−1; elimination half-life, 37.69±21.70 h.  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