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 281 毫秒
1.
The aims of this study were to design the formulation of curcumin (CUR) liposomes coated with N-trimethyl chitosan chloride (TMC) and to evaluate in vitro release characteristics and in vivo pharmacokinetics and bioavailability of TMC-coated CUR liposomes in rats. The structure of synthesized TMC was examined by infrared spectroscopy, with the presence of trimethyl groups, and by proton nuclear magnetic resonance spectroscopy, indicating the high degree of substitution quaternization (65.6%). Liposomes, composed of soybean phosphotidylcholine, cholestrol, and D-α-tocopheryl polyethylene glycol 1000 succinate, were prepared by a thin-film dispersion method. Characteristics of the CUR liposomes, including entrapment efficiency (86.67%), drug-loading efficiency (2.33%), morphology, particle size (221.4?nm for uncoated liposomes and 657.7?nm for TMC-coated liposomes), and zeta potential (-9.63 mV for uncoated liposomes and +15.64 mV for TMC-coated liposomes) were investigated. Uncoated CUR liposomes and TMC-coated CUR liposomes showed a similar in vitro release profile. Nearly 50% of CUR was released from liposomes, whereas 80% of CUR was released from CUR propylene glycol solution. CUR incorporated into TMC-coated liposomes exhibited different pharmacokinetic parameters and enhanced bioavailability (C(max)?=?46.13 μg/L, t(1/2)?=?12.05 hours, AUC?=?416.58 μg/L·h), compared with CUR encapsulated by uncoated liposomes (C(max)?=?32.12 μg/L, t(1/2)?=?9.79 hours, AUC?=?263.77 μg/L·h) and CUR suspension (C(max)?=?35.46 μg/L, t(1/2)?=?3.85 hours, AUC?=?244.77 μg/L·h). In conclusion, oral delivery of coated CUR liposomes is a promising strategy for poorly water-soluble CUR.  相似文献   

2.
《Chronobiology international》2013,30(9-10):1843-1853
The human circadian system is known to affect the pharmacokinetics and pharmacodynamics of several classes of respiratory disease medications. The current study involving 16 healthy adults investigated if the time-of-day of dosing of roflumilast, a novel phosphodiesterase-4 inhibitor, affects its pharmacokinetics. The rate of drug absorption (tmax: 1.50 versus 2.00 h) and peak concentration at tmax (Cmax: 3.79 versus 3.06 µg/L) was slightly greater with morning than evening administration, but without clinical significance. The extent of drug absorption (AUC) and drug elimination (t1/2) did not differ between the two dosing times. The pharmacokinetics of the active main metabolite, roflumilast N-oxide, also was not affected by the time of drug administration. Finally, the safety and tolerability of roflumilast did not differ between the two different times of administration. (E-mail correspondence: thomas.bethke@nycomed.com)  相似文献   

3.
The pharmacokinetics of ibuprofen enantiomers were investigated in a crossover study in which seven healthy male volunteers received single oral doses of 800 mg racemic ibuprofen as a soluble granular formulation (sachet) containing L-arginine (designated trade name: Spedifen®), 400 mg (-)R-ibuprofen arginine or 400 mg (+)S-ibuprofen arginine. Plasma levels of both enantiomers were monitored up to 480 minutes after drug intake using an enantioselective analytical method (HPLC with ultraviolet detection) with a quantitation limit of 0.25 mg/l. Substantial inter-subject variability in the evaluated pharmacokinetic parameters was observed in the present study. After (+)S-ibuprofen arginine, the following mean pharmacokinetic parameters ±SD were calculated for (+)S-ibuprofen: tmax 28.6 ± 28.4 min; Cmax 36.2 ± 7.7 mg/l; AUC 86.4 ± 14.9 mg · h/l; t½ 105.2 ± 20.4 min. After (-)R-ibuprofen arginine, the following mean pharmacokinetic parameters were calculated for (+)S-ibuprofen and (-)R-ibuprofen, respectively: tmax 90.0 ± 17.3 and 50.5 ± 20.5 min; Cmax 9.7 ± 3.0 and 35.3 ± 5.0 mg/l; AUC 47.0 ± 17.2 and 104.7 ± 27.7 mg · h/l; t½ 148.1 ± 63.6 and 97.7 ± 23.3 min. After racemic ibuprofen arginine, the following mean pharmacokinetic parameters were calculated for (+)S- and (-)R-ibuprofen, respectively: tmax 30.7 ± 29.1 and 22.9 ± 29.8 min.; Cmax 29.9 ± 5.6 and 25.6 ± 4.4 mg/l; AUC 105.1 ± 23.0 and 65.3 ± 15.0 mg · h/l; t½ 136.6 ± 20.7 and 128.6 ± 45.0 min. Tmax values of S(+)- and (-)R-ibuprofen after a single dose of 400 mg of each enantiomer did not differ significantly from the corresponding parameters obtained after a single dose of 800 mg of racemic ibuprofen arginine, indicating that the absorption rate of (-)R- and (+)S-ibuprofen is not different when the two enantiomers are administered alone or as a racemic compound. An average of 49.3 ± 9.0% of a dose of the (-)R-ibuprofen arginine was bioinverted into its antipode during the study period (480 minutes post-dosing). The percent bioinversion during the first 30 minutes after (-)R-ibuprofen arginine intake averaged 8.1 ± 3.9%. The mean AUC of (+)S-ibuprofen calculated after 800 mg racemic ibuprofen arginine (105.1 ± 23.0 mg · h/l) was lower than the mean AUC value obtained by summing the AUCs of (+)S-ibuprofen after administration of 400 mg (+)S-ibuprofen arginine and 400 mg (-)R-ibuprofen arginine (133.4 ± 26.6 mg · h/l). In conclusion, the administration of Spedifen® resulted in very rapid absorption of the (+)S-isomer (eutomer) with tmax values much lower than those observed for this isomer when conventional oral solid formulations such as capsules or tablets of racemic ibuprofen are administered. This characteristic is particularly favourable in those conditions in which a very rapid analgesic effect is required. Chirality 9:297–302, 1997. © 1997 Wiley-Liss, Inc.  相似文献   

4.
An enantioselective HPLC bioassay has been developed relying on extraction of (R)- and (S)-atenolol from alkalinized plasma or serum (pH > 12) into dichloromethane containing 5% (v/v) 1-butanol followed by an achiral derivatization of the drug with phosgene leading to (R)- and (S)-oxazolidine-2-one derivatives. Under these conditions there was quantitative conversion of the acetamido group to the corresponding nitrile. These stable derivatives were separated on a (R,R)-diaminocylohexane-dinitrobenzoyl chiral stationary phase [(R,R)-DACH-DNB] using dichloromethane/methanol 98/2 as mobile phase. Determination limits of 0.5 ng for (R)- and 0.6 ng for (S)-atenolol could be achieved using fluorimetric detection. The assay was applied to a human pharmacokinetic study which was performed in a randomized cross-over, double-blind fashion in 12 healthy volunteers, administering single oral doses of 100 mg (R,S)-, 50 mg (R)-, and 50 mg (S)-atenolol AUC0–24 and Cmax values of (R)-atenolol were slightly but significant higher than those of (S)-atenolol. The R/S ratios were 1.09 for AUC(R)/AUC(S) and 1.03 for Cmax (R)/Cmax(S) (P < 0.01) respectively after administration of the racemic drug. However, there were no differences between AUC, Cmax, and t½ values of each enantiomer, whether they were administered as single enantiometers or in the form of its racemic mixture. © 1993 Wiley-Liss, Inc.  相似文献   

5.
There are no studies indicating a possible modification of imipenem pharmacokinetics related to the hour (i.e., circadian time) of its administration. The aim of this study was to evaluate the influence of different times of intramuscular imipenem administration on its disposition in Wistar AF EOPS rats. Four groups of eight animals were given a single intramuscular injection of 140 mg/kg of imipenem either at 10∶00, 16∶00, 22∶00, or 04∶00 h. Blood samples were collected 0.5, 1, 2, 3, 4, 6, and 8 h after drug injection, and the main pharmacokinetic parameters determined were Cmax, Tmax, elimination half‐life (t1/2), area under the concentration‐versus‐time curve (AUC), total serum clearance (CL/F), and volume of distribution (V/F). Circadian variation of Cmax (49%), Tmax (92%), and AUC (19%) was observed leading to variability of imipenem exposure. Clearance and volume of distribution were modified according to the circadian time of drug injection but did not reach statistical significance. The results suggest that varying the time of administration induces intra‐individual variability.  相似文献   

6.
Photosynthesis and respiration of three Alaskan Porphyra species, P. abbottiae V. Krishnam., P. pseudolinearis Ueda species complex (identified as P. pseudolinearis” below), and P. torta V. Krishnam., were investigated under a range of environmental parameters. Photosynthesis versus irradiance (PI) curves revealed that maximal photosynthesis (Pmax), irradiance at maximal photosynthesis (Imax), and compensation irradiance (Ic) varied with salinity, temperature, and species. The Pmax of Porphyra abbottiae conchocelis varied between 83 and 240 μmol O2 · g dwt?1 · h?1 (where dwt indicates dry weight) at 30–140 μmol photons · m?2 · s?1 (Imax) depending on temperature. Higher irradiances resulted in photoinhibition. Maximal photosynthesis of the conchocelis of P. abbottiae occurred at 11°C, 60 μmol photons · m?2·s?1, and 30 psu (practical salinity units). The conchocelis of P. “pseudolinearis” and P. torta had similar Pmax values but higher Imax values than those of P. abbottiae. The Pmax of P. “pseudolinearis” conchocelis was 200–240 μmol O2 · g dwt?1 · h?1 and for P. torta was 90–240 μmol O2 · g dwt?1 · h?1. Maximal photosynthesis for P. “pseudolinearis” occurred at 7°C and 250 μmol photons · m?2 · s?1 at 30 psu, but Pmax did not change much with temperature. Maximal photosynthesis for P. torta occurred at 15°C, 200 μmol photons · m?2 · s?1, and 30 psu. Photosynthesis rates for all species declined at salinities <25 or >35 psu. Estimated compensation irradiances (Ic) were relatively low (3–5 μmol · photons · m?2 · s?1) for intertidal macrophytes. Porphyra conchocelis had lower respiration rates at 7°C than at 11°C or 15°C. All three species exhibited minimal respiration rates at salinities between 25 and 35 psu.  相似文献   

7.
A simple and sensitive flow injection–chemiluminescence (FI–CL) method has been developed for the determination of puerarin, based on the fact that puerarin can greatly inhibit CL of the luminol–H2O2–haemoglobin system. The inhibition of CL intensity was linear to the logarithm of the concentration of puerarin in the range 0.08–10.0 μg/mL (r2 = 0.9912). The limit of detection was 0.05 μg/mL (3σ) and the relative standard deviation (RSD) for 1.0 μg/mL (n = 11) of puerarin solution was 1.4%. Coupled with solid‐phase extraction (SPE) as the sample pretreatment, the determination of puerarin in biological samples and a preliminary pharmocokinetic study of puerarin in rats were performed. The recoveries for plasma and urine at three different concentrations were 89.2–110.0% and 91.4–104.8%, respectively. The pharmacokinetics of puerarin in plasma of rat coincides with the two‐compartment open model. The T1/2α, T1/2β, CL/F, VZ/F, AUC(0 – t), MRT(0 – ∞), Tmax and Cmax were 0.77 ± 0.21 h, 7.55 ± 2.64 h, 2.43 ± 1.02 L/kg/h, 11.40 ± 3.45 L/kg, 56.67 ± 10.65 mg/h/L, 5.04 ± 2.78 h, 1.00 ± 0.35 h and 19.70 ± 4.67 μg/mL, respectively. Copyright © 2011 John Wiley & Sons, Ltd.  相似文献   

8.
马肝醇脱氢酶催化有机硅酮不对称还原反应动力学   总被引:2,自引:0,他引:2  
探讨了马肝醇脱氢酶(HLADH)催化三甲基硅乙酮及其碳结构类似物不对称还原反应动力学.结果表明,在酶浓度低于150 mg/L时,底物浓度与反应初速度的关系符合米氏动力学方程;HLADH催化三甲基硅乙酮不对称还原反应的KmvmaxEa分别为2.67 mmol/L、0.118 mmol/(L·min·mg)和37 kJ/mol, 其碳结构类似物的相应值分别为3.56 mmol/L、0.084 mmol/(L·min·mg)和61 kJ/mol.  相似文献   

9.
The purpose was to assess the impact of the use of a chiral bioanalytical method on the conclusions of a bioequivalence study that compared two ibuprofen suspensions with different rates of absorption. A comparison of the conclusion of bioequivalence between a chiral method and an achiral approach was made. Plasma concentrations of R‐ibuprofen and S‐ibuprofen were determined using a chiral bioanalytical method; bioequivalence was tested for R‐ibuprofen and for S‐ibuprofen separately and for the sum of both enantiomers as an approach for an achiral bioanalytical method. The 90% confidence interval (90% CI) that would have been obtained with an achiral bioanalytical method (90% CI: Cmax: 117.69–134.46; AUC0t: 104.75–114.45) would have precluded the conclusion of bioequivalence. This conclusion cannot be generalized to the active enantiomer (90% CI: Cmax: 103.36–118.38; AUC0t: 96.52–103.12), for which bioequivalence can be concluded, and/or the distomer (90% CI: Cmax: 132.97–151.33; AUC0t: 115.91–135.77) for which a larger difference was observed. Chiral bioanalytical methods should be required when 1) the enantiomers exhibit different pharmacodynamics and 2) the exposure (AUC or Cmax) ratio of enantiomers is modified by a difference in the rate of absorption. Furthermore, the bioequivalence conclusion should be based on all enantiomers, since the distomer(s) might not be completely inert, in contrast to what is required in the current regulatory guidelines. In those cases where it is unknown if the ratio between enantiomers is modified by changing the rate of absorption, chiral bioanalytical methods should be employed unless enantiomers exhibit the same pharmacodynamics. Chirality 28:429–433, 2016. © 2016 Wiley Periodicals, Inc.  相似文献   

10.
The ground-state tautomerization of the G·C Watson–Crick base pair by the double proton transfer (DPT) was comprehensively studied in vacuo and in the continuum with a low dielectric constant (??=?4), corresponding to a hydrophobic interface of protein–nucleic acid interactions, using DFT and MP2 levels of quantum-mechanical (QM) theory and quantum theory “Atoms in molecules” (QTAIM). Based on the sweeps of the electron-topological, geometric, polar, and energetic parameters, which describe the course of the G·C???G?·C? tautomerization (mutagenic tautomers of the G and C bases are marked with an asterisk) through the DPT along the intrinsic reaction coordinate (IRC), it was proved that it is, strictly speaking, a concerted asynchronous process both at the DFT and MP2 levels of theory, in which protons move with a small time gap in vacuum, while this time delay noticeably increases in the continuum with ??=?4. It was demonstrated using the conductor-like polarizable continuum model (CPCM) that the continuum with ??=?4 does not qualitatively affect the course of the tautomerization reaction. The DPT in the G·C Watson–Crick base pair occurs without any intermediates both in vacuum and in the continuum with ??=?4 at the DFT/MP2 levels of theory. The nine key points along the IRC of the G·C base pair tautomerization, which could be considered as electron-topological “fingerprints” of a concerted asynchronous process of the tautomerization via the DPT, have been identified and fully characterized. These key points have been used to define the reactant, transition state, and product regions of the DPT reaction in the G·C base pair. Analysis of the energetic characteristics of the H-bonds allows us to arrive at a definite conclusion that the middle N1H?N3/N3H?N1 and the lower N2H?O2/N2H?O2 parallel H-bonds in the G·C/G?·C? base pairs, respectively, are anticooperative, that is, the strengthening of the middle H-bond is accompanied by the weakening of the lower H-bond. At that point, the upper N4H?O6 and O6H?N4 H-bonds in the G·C and G?·C? base pairs, respectively, remain constant at the changes of the middle and the lower H-bonds at the beginning and at the ending of the G·C???G?·C? tautomerization. Aiming to answer the question posed in the title of the article, we established that the G?·C? Löwdin’s base pair satisfies all the requirements necessary to cause point mutations in DNA except its lifetime, which is much less than the period of time required for the replication machinery to forcibly dissociate a base pair into the monomers (several ns) during DNA replication. So, from the physicochemical point of view, the G?·C? Löwdin’s base pair cannot be considered as a source of point mutations arising during DNA replication.  相似文献   

11.
The energy cost per unit of distance (C s, kilojoules per metre) of the front-crawl, back, breast and butterfly strokes was assessed in 20 elite swimmers. At sub-maximal speeds (v), C s was measured dividing steady-state oxygen consumption (O2) by the speed (v, metres per second). At supra-maximal v, C s was calculated by dividing the total metabolic energy (E, kilojoules) spent in covering 45.7, 91.4 and 182.9 m by the distance. E was obtained as: E = E an+O2max t pO2max(1−e−( t p/)), where E an was the amount of energy (kilojoules) derived from anaerobic sources, O2max litres per second was the maximal oxygen uptake, α (=20.9 kJ · l O2 −1) was the energy equivalent of O2, τ (24 s) was the time constant assumed for the attainment of O2max at muscle level at the onset of exercise, and t p (seconds) was the performance time. The lactic acid component was assumed to increase exponentially with t p to an asymptotic value of 0.418 kJ · kg−1 of body mass for t p ≥ 120 s. The lactic acid component of E an was obtained from the net increase of lactate concentration after exercise (Δ[La]b) assuming that, when Δ[La]b = 1 mmol · l−1 the net amount of metabolic energy released by lactate formation was 0.069 kJ · kg−1. Over the entire range of v, front crawl was the least costly stroke. For example at 1 m · s−1, C s amounted, on average, to 0.70, 0.84, 0.82 and 0.124 kJ · m−1 in front crawl, backstroke, butterfly and breaststroke, respectively; at 1.5 m · s−1, C s was 1.23, 1.47, 1.55 and 1.87 kJ · m−1 in the four strokes, respectively. The C s was a continuous function of the speed in all of the four strokes. It increased exponentially in crawl and backstroke, whereas in butterfly C s attained a minimum at the two lowest v to increase exponentially at higher v. The C s in breaststroke was a linear function of the v, probably because of the considerable amount of energy spent in this stroke for accelerating the body during the pushing phase so as to compensate for the loss of v occurring in the non-propulsive phase. Accepted: 14 April 1998  相似文献   

12.
Context: The acylglycerides of lipoic and dihydrolipoic acids may serve as slow-release sources for cutaneous delivery of these antioxidants when formulated in a liposomal vehicle.

Objective: Testing was conducted to determine the storage stability of lipoyl glycerides in phospholipid-based liposomes.

Materials and methods: Lipoyl glycerides prepared by transesterification of lipoic acid with high oleic sunflower oil were incorporated into unilamellar liposomes comprised of soy phosphatidylcholine (soyPC) or dioleoylphosphatidylcholine (DOPC).

Results: Lipoyl glycerides were stable in soyPC at 4?°C (90% remaining after five weeks) and decayed with a half-life (t½) of 14?d at 40?°C. In contrast, lipoyl glycerides embedded in DOPC were completely stable for four weeks at 40?°C. Dihydrolipoyl glycerides in soyPC converted to lipoyl glycerides at 4?°C (t½?=?14?d) over four weeks, and much more rapidly so at 40?°C (t½?=?1?d). A hydroperoxide accumulation analysis indicated that lipoyl glycerides and dihydrolipoyl glycerides were modified or degraded while suppressing autoxidation of the polyunsaturated fatty acids present in soyPC. Dynamic light scattering measurements found that liposomes containing lipoyl glycerides or dihydrolipoyl glycerides did not undergo significant size changes for at least 48?d, indicating that inclusion of the lipoic acid derivatives did not induce vesicle aggregation.

Discussion/Conclusion: Substitution of the soyPC with DOPC, which is not readily subject to autoxidation, provided a much more stable storage environment for lipoyl glycerides. These findings confirm the expectation that phospholipid liposomes need to be oxidatively stable vehicles for dermal delivery of lipoic acid derivatives.  相似文献   

13.
Emiliania huxleyi (strain L) expressed an exceptional P assimilation capability. Under P limitation, the minimum cell P content was 2.6 fmol P·cell?1, and cell N remained constant at all growth rates at 100 fmol N·cell?1. Both, calcification of cells and the induction of the phosphate uptake system were inversely correlated with growth rate. The highest (cellular P based) maximum phosphate uptake rate (VmaxP) was 1400 times (i.e. 8.9 h?1) higher than the actual uptake rate. The affinity of the P‐uptake system (dV/dS) was 19.8 L·μmol?1·h?1 at μ = 0.14 d?1. This is the highest value ever reported for a phytoplankton species. Vmax and dV/dS for phosphate uptake were 48% and 15% lower in the dark than in the light at the lowest growth rates. The half‐saturation constant for growth was 1.1 nM. The coefficient for luxury phosphate uptake (Qmaxt/Qmin) was 31. Under P limitation, E. huxleyi expressed two different types of alkaline phosphatase (APase) enzyme kinetics. One type was synthesized constitutively and possessed a Vmax and half‐saturation constant of 43 fmol MFP·cell?1·h?1 and 1.9 μM, respectively. The other, inducible type of APase expressed its highest activity at the lowest growth rates, with a Vmax and half‐saturation constant of 190 fmol MFP·cell?1·h?1 and 12.2 μM, respectively. Both APase systems were located in a lipid membrane close to the cell wall. Under N‐limiting growth conditions, the minimum N quotum was 43 fmol N·cell?1. The highest value for the cell N‐specific maximum nitrate uptake rate (VmaxN) was 0.075 h?1; for the affinity of nitrate uptake, 0.37 L·μmol?1·h?1. The uptake rate of nitrate in the dark was 70% lower than in the light. N‐limited cells were smaller than P‐limited cells and contained 50% less organic and inorganic carbon. In comparison with other algae, E. huxleyi is a poor competitor for nitrate under N limitation. As a consequence of its high affinity for inorganic phosphate, and the presence of two different types of APase in terms of kinetics, E. huxleyi is expected to perform well in P‐controlled ecosystems.  相似文献   

14.
Ritonavir is an antiretroviral drug characterized by low solubility and high permeability which corresponds to BCS class II drug. The purpose of the study was to develop solid dispersion by different methods and investigate them for in vitro and in vivo performance for enhancing dissolution and bioavailability, respectively. Since the drug possesses food-related absorption, the effect of biorelevant media (FaSSIF and FeSSIF state) on dissolution behavior was also studied. The solid dispersion was prepared using Gelucire as carrier in 1:4 ratio by different methods and were characterized for differential scanning calorimetry (DSC), X-ray diffractometry, scanning electron microscopy, and FT-IR. Oral bioavailability of 10 mg of ritonavir in solid dispersion prepared by solvent evaporation (SE1) and melt method (MM1) was compared with pure drug after oral administration of solid dispersion and pure drug to Albino Wistar rats of either sex. The results suggested formation of eutectic solid dispersion. In vitro dissolution studies was performed in 0.1 N HCl and biorelevant media showed enhanced dissolution rate as compared to pure drug in both FeSSIF media and 0.1 N HCl. The apparent rate of absorption of ritonavir from SE1 (C max 20221.37 ng/ml, t max 0.5 h) was higher than that of MM1 (C max 2,462.2, t max 1 h) and pure drug (C max 1,354.8 ng/ml, t max 0.5 h). On the basis of the result obtained, it was concluded that solid dispersion is a good approach to enhance solubility and bioavailability of poorly water-soluble ritonavir.  相似文献   

15.
This randomized and controlled trial investigated whether the increase in elite training at different altitudes altered the oxidative stress biomarkers of the nervous system. This is the first study to investigate four F4-neuroprostanes (F4-NeuroPs) and four F2-dihomo-isoprostanes (F2-dihomo-IsoPs) quantified in 24-h urine. The quantification was carried out by ultra high pressure liquid chromatography-triple quadrupole-tandem mass spectrometry (UHPLC-QqQ-MS/MS). Sixteen elite triathletes agreed to participate in the project. They were randomized in two groups, a group submitted to altitude training (AT, n?=?8) and a group submitted to sea level training (SLT) (n?=?8), with a control group (Cg) of non-athletes (n?=?8). After the experimental period, the AT group triathletes gave significant data: 17-epi-17-F2t-dihomo-IsoP (from 5.2?±?1.4?μg/mL 24?h?1 to 6.6?±?0.6?μg/mL 24?h?1), ent-7(RS)-7-F2t-dihomo-IsoP (from 6.6?±?1.7?μg/mL 24?h?1 to 8.6?±?0.9?μg/mL 24?h?1), and ent-7-epi-7-F2t-dihomo-IsoP (from 8.4?±?2.2?μg/mL 24?h?1 to 11.3?±?1.8?μg/mL 24?h?1) increased, while, of the neuronal degeneration-related compounds, only 10-epi-10-F4t-NeuroP (8.4?±?1.7?μg/mL 24?h?1) and 10-F4t-NeuroP (5.2?±?2.9?μg/mL 24?h?1) were detected in this group. For the Cg and SLT groups, no significant changes had occurred at the end of the two-week experimental period. Therefore, and as the main conclusion, the training at moderate altitude increased the F4-NeuroPs- and F2-dihomo-isoPs-related oxidative damage of the central nervous system compared to similar training at sea level.  相似文献   

16.
An easy, rapid and selective method for the determination of rofecoxib in human plasma is presented. The analytical technique is based on reversed-phase high-performance liquid chromatography coupled to atmospheric pressure chemical ionisation mass spectrometry (Finnigan Mat LCQ ion trap). The retention time of rofecoxib was 1.2 min. The method has been validated over a linear range from 1 to 500 μg/l using celecoxib as internal standard. After validation, the method was used to study the pharmacokinetic profile of rofecoxib in 12 healthy volunteers after administration of a single oral dose (12.5 mg). The presented method was sufficient to cover more than 95% of the area under the curve. The pharmacokinetic characteristics (mean±SD) were tmax: 2.4±1.0 h, cmax: 147±34 μg/l, AUC: 2038±581 μg h/l and t1/2: 11.3±2.1 h.  相似文献   

17.
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   

18.
Digoxin, frequently used in the treatment of congestive heart failure, has a very narrow therapeutic index. We studied the differences in digoxin pharmacokinetics when ingested in the morning versus evening. A single digoxin (0.25 mg) dose was given orally to the same group of 10 diurnally active healthy (6 male and 4 female) volunteers in the morning at 08:00 and evening at 20:00 in separate experiments scheduled 2 weeks apart. Blood samples were collected at specific times for 48h after each timed dose; digoxin was determined by radioimmunoassay (RIA). Maximum plasma concentration Cmax; Tmax, the time to reach Cmax; area under plasma concentration curve AUC; and elimination half-time T1/2 of digoxin were determined. Tmax was statistically significantly shorter (54 min) following 08:00 dosing compared to 20:00 dosing (96 min). Although the Cmax was higher after morning than evening dosing, it was not significantly so. No other parameter of digoxin pharmacokinetics except Tmax exhibited administration time dependency. (Chronobiology International, 18(5), 841–849, 2001)  相似文献   

19.
The effect of general anaesthetics, alcohols and hydrostatic pressure on the thermal transition in dipalmitoyl phosphatidylcholine multilayer liposomes has been measured using dilatometry. The volume increasse at the transition (ΔVt) is 0.0350 ± 0.0003 ml/g. the transition temperature (Tt) 41.84 ± 0.09°C and the width of the transition 1.025 ± 0.18°C. ΔH calculated by the Clapeyron-Clausius equation is 8.4 kcal/mol. The n-alcohols C3C5 reduced the transition temperature without affecting the transition width which was however, increased by n-hexanol. Trichloroethylene, the fluorescent probe N-phenyl-1-naphthyl-amine, and methoxyflurane all increased the transition width (reduced the cooperativity of the transition) with a simultaneous depression of Tt. Methoxyflurane caused a two-stage transition expansion. Diethyl ether's effect has similarities with both the C3 and C6 alcohols. Generally ΔVt was unaffected by the agents.Pressure increased Tt by 0.0238°C/atm linearly over the range 1–300 atm in both treated and untreated liposomes, and therefore cannot be said to antagonize anaesthetics. In both treated and untreated liposomes ΔVt and the width of the transition were unaffected by pressure. Pressure thus reverses the effects of anaesthetics on Tt but not their spread of the transition width.  相似文献   

20.
We compared autotrophic growth of the dinoflagellate Karlodinium micrum (Leadbeater et Dodge) and the cryptophyte Storeatula major (Butcher ex Hill) at a range of growth irradiances (Eg). Our goal was to determine the physiological bases for differences in growth–irradiance relationships between these species. Maximum autotrophic growth rates of K. micrum and S. major were 0.5 and 1.5 div.·d?1, respectively. Growth rates were positively correlated with C‐specific photosynthetic performance (PPC, g C·g C?1·h?1) (r2=0.72). Cultures were grouped as light‐limited (LL) and high‐light (HL) treatments to allow interspecific comparisons of physiological properties that underlie the growth–irradiance relationships. Interspecific differences in the C‐specific light absorption rate (EaC, mol photons·g C?1·h?1) were observed only among HL acclimated cultures, and the realized quantum yield of C fixation (φC(real.), mol C·mol photons?1) did not differ significantly between species in either LL or HL treatments. The proportion of fixed C that was incorporated into new biomass was lower in K. micrum than S. major at each Eg, reflecting lower growth efficiency in K. micrum. Photoacclimation to HL in K. micrum involved a significant loss of cellular photosynthetic capacity (Pmaxcell), whereas in S. major, Pmaxcell was significantly higher in HL acclimated cells. We conclude that growth rate differences between K. micrum and S. major under LL conditions relate primarily to cell metabolism processes (i.e. growth efficiency) and that reduced chloroplast function, reflected in PPC and photosynthesis–irradiance curve acclimation in K. micrum, is also important under HL conditions.  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