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 15 毫秒
1.
Novel C-6 substituted pyrimidine derivatives are good substrates of herpes simplex virus type 1 thymidine kinase (HSV1-TK). Enzyme kinetic experiments showed that our lead compound, N-methyl DHBT (N-methyl-6-(1,3-dihydroxyisobutyl) thymine; N-Me DHBT), is phosphorylated at a similar rate compared to "gold standard" 9-[4-fluoro-3-(hydroxymethyl)butyl]guanine, FHBG, (K(m) = 10 ± 0.3 μM; k(cat) = 0.036 ± 0.015 sec(-1)). Additionally, it does not show cytotoxic properties on B16F1 cells up to a concentration of 10 mM. The x-ray analysis of the crystal structures of HSV1-TK with N-Me DHBT and of HSV1-TK with the fluorinated derivative N-Me FHBT confirmed the binding mode predicted by docking studies and their substrate characteristics. Moreover, the crystal structure of HSV1-TK with N-Me DHBT revealed an additional water-mediated H-bond interesting for the design of further analogues.  相似文献   

2.
P Manavalan  F A Momany 《Biopolymers》1980,19(11):1943-1973
Empirical conformational energy calculations have been carried out for N-methyl derivatives of alanine and phenylalanine dipeptide models and N-methyl-substituted active analogs of three biologically active peptides, namely thyrotropin-releasing hormone (TRH), enkephalin (ENK), and luteinizing hormone-releasing hormone (LHRH). The isoenergetic contour maps and the local dipeptide minima obtained, when the peptide bond (ω) preceding the N-methylated residue is in the trans configuration show that (1) N-methylation constricts the conformational freedom of both the ith and (i + 1)th residues; (2), the lowest energy position for both residues occurs around ? = ?135° ± 5° and ψ = 75° ± 5°, and (3) the αL conformational state is the second lowest energy state for the (i + 1)th residue, whereas for the ith residue the C5 (extended) conformation is second lowest in energy. When the peptide bond (ωi) is in the cis configuration the ith residue is energetically forbidden in the range ? = 0° to 180° and ψ = ?180° to +180°. Conformations of low energy for ωi = 0° are found to be similar to those obtained for the trans peptide bond. In all the model systems (irrespective of cis or trans), the αR conformational state is energetically very high. Significant deviations from planarity are found for the peptide bond when the amide hydrogen is replaced by a methyl group. Two low-energy conformers are found for [(N-Me)His2]TRH. These conformers differ only in the ? and ψ values at the (N-Me)His2 residue. Among the different low-energy conformers found for each of the ENK analogs [D -Ala2,(N-Me)Phe4, Met5]ENK amide and [D -Ala2,(N-Me)Met5]ENK amide, one low-energy conformer was found to be common for both analogs with respect to the side-chain orientations. The stability of the low-energy structures is discussed in the light of the activity of other analogs. Two low-energy conformers were found for [(N-Me)Leu7]LHRH. These conformations differ in the types of bend around the positions 6 and 7 of LHRH. One bend type is eliminated when the active analog [D -Ala6,(M-Me)Leu7]LHRH is considered.  相似文献   

3.
The development and regional distribution of ribonucleotide reductase (EC 1.17.4.1) were determined in rat brain. Ribonucleotide reductase was partially purified by ammonium sulfate fractionation (20-40% saturation). Enzyme activity was measured by a specific radiochemical assay. This method involved the reduction of [14C]cytidine diphosphate (CDP) to [14C]deoxy-cytidine diphosphate with subsequent hydrolysis and separation of the product ([14C]deoxycytidine) from substrate ([14C]cytidine) by Dowex-1-borate ion-exchange chro-matography. The specific activity of ribonucleotide reductase in whole brain of newborn rats was 3.78 ± 0.55 units (pmol/h)/mg protein (SEM; n = 6) and declined to 0.17 ± 0.01 units/mg protein (n = 7) at 10-12 weeks of age, with a further decline to 0.11 ± 0.01 units/mg protein (n = 3) at 1 year. Ribonucleotide reductase activity in rat liver decreased from 4.58 ± 0.62 units/mg protein (n = 3) in newborn animals to 0.06 ± 0.01 units/mg protein (n = 7) at 10-12 weeks and was present at trace levels at 6 months of age. The decline in specific activity with age was not due to a change in the Km for CDP. The Km for CDP in brain of newborn and adult rats was 80-90 μM. In 10- to 12-week-old rats, the specific activity of ribonucleotide reductase was similar in the various regions of the brain tested except for the brainstem, which had 50% lower specific activity than the whole brain. These results indicate that ribonucleotide reductase activity is present and widely distributed in adult rat brain.  相似文献   

4.
Abstract: We identified and characterized 125I-endothelin-1 (125I-ET-1) binding sites in tumor capillaries isolated from human glioblastomas, using the quantitative receptor autoradiographic technique with pellet sections. Quantification was done using the computerized radioluminographic imaging plate system. High-affinity ET receptors were localized in capillaries from glioblastomas and the surrounding brain tissues (KD = 4.7 ± 1.0 × 10?10 and 1.6 ± 0.3 × 10?10M, respectively; Bmax = 161 ± 38 and 140 ± 37 fmol/mg, respectively; mean ± SEM, n = 5). BQ-123, a selective antagonist for the ETA receptor, potently competed for 125I-ET-1 binding to sections of the microvessels with IC50 values of 5.1 ± 0.3 and 5.1 ± 1.5 nM, and 10?6M BQ-123 displaced 84 and 58% of ET binding to capillaries from tumors and brains, respectively. In addition, competition curves obtained in the presence of increasing concentrations of ET-3 showed two components (IC50 = 5.7 ± 2.5 × 10?10 and 1.4 ± 0.2 × 10?6M for tumor microvessels, 1.8 ± 0.6 × 10?10 and 1.1 ± 0.3 × 10?6M for brain microvessels, respectively). Our results indicate that (a) the method we used is simple and highly sensitive for detecting and characterizing various receptors in tumor capillaries, especially in the case of a sparse specimen, and (b) capillaries in glioblastomas express specific high-affinity ET binding sites, candidates for biologically active ET receptors, which predominantly belong to the ETA subtype.  相似文献   

5.
Aims: To investigate the in vitro antiviral activity of Distictella elongata (Vahl) Urb. ethanol extracts from leaves (LEE), fruits (FEE), stems and their main components. Methods and Results: The antiviral activity was evaluated against human herpesvirus type 1 (HSV‐1), murine encephalomyocarditis virus (EMCV), vaccinia virus Western Reserve (VACV‐WR) and dengue virus 2 (DENV‐2) by the 3‐(4, 5‐dimethylthiazol‐2‐yl)‐2,5‐diphenyltetrazolium bromide (MTT) colorimetric assay. LEE presented anti‐HSV‐1 [EC50 142·8 ± 5·3 μg ml?1; selectivity index (SI) 2·0] and anti‐DENV‐2 activity (EC50 9·8 ± 1·3 μg ml?1; SI 1·5). The pectolinarin ( 1 ) isolated from LEE was less active against HSV‐1 and DENV‐2. A mixture of the triterpenoids ursolic, pomolic and oleanolic acids was also obtained. Ursolic and oleanolic acids have shown antiviral activity against HSV‐1. A mixture of pectolinarin ( 1 ) and acacetin‐7‐O‐rutinoside ( 2 ) was isolated from FEE and has presented anti‐DENV‐2 activity (EC50 11·1 ± 1·6 μg ml?1; SI > 45). Besides the antiviral activity, D. elongata has disclosed antioxidant effect. Conclusions: These data shows that D. elongata has antiviral activity mainly against HSV‐1 and DENV‐2, besides antioxidant activity. These effects might be principally attributed to flavonoids isolated. Significance and Impact of the Study: Distictella elongata might be considered a promising source of anti‐dengue fever phytochemicals.  相似文献   

6.
Isradipine (PN 200–110) is a highly potent calcium entry blocker with an asymmetrically substituted dihydropyridine ring (methyl- and isopropylester, respectively). The binding of the (+)-(S)-isradipine and (?)-(R)-isradipine to isolated human serum albumin (HSA, 30 μmol/l) and α1-acid glycoprotein (AAG, 10 μmol/l) has been studied in vitro over a wide range of isradipine concentrations (0.06–20 μmol/l) using high-performance liquid chromatography (HPLC). HPLC experiments revealed that both isradipine enantiomers were bound to one class of high-affinity binding sites on the AAG molecule (n(S) = 0.83 ± 0.05, Ka(S) = (1.33 ± 0.25) × 106 1/mol, n(R) = 0.85 ± 0.07, Ka(R) = (1.17 ± 0.44) × 107 l/mol). The (R)-enantiomer also exhibited an interaction with the secondary low-affinity binding sites (n′K′a (R) = (2.66 ± 0.65) × 104 l/mol). In contrast, the pharmacologically more potent (+)-(S)-enantiomer was more strongly bound to HSA than its optical antipode (n(S) = 1.07 ± 0.07, Ka(S) = (1.76 ± 0.26) × 105 l/mol, nKa(R) = (3.62 ± 0.06) × 104 l/mol). In general, the resulting binding characteristics of individual isradipine enantiomers showed stereoselectivity, but this was opposite for the two most important plasma binding proteins. The process of accumulation of isradipine by human platelets in the therapeutically relevant range (10–80 ng/ml) at 37°C was devoid of stereoselectivity. © 1995 Wiley-Liss, Inc.  相似文献   

7.
ATP-Activated Nonselective Cation Current in NG108-15 Cells   总被引:5,自引:0,他引:5  
Abstract: ATP (1 mM) induced a biphasic increase in intracellular Ca2+ concentration ([Ca2+]i), i.e., an initial transient increase decayed to a level of sustained increase, in NG108-15 cells. The transient increase was inhibited by a phospholipase C inhibitor, 1-[6-[[17β-3-methoxyestra-1,3,5(10)-trien-17-yl]amino]hexyl]-1H-pyrrole-2,5-dione (U73122), whereas the sustained increase was abolished by removal of external Ca2+. We examined the mechanism of the ATP-elicited sustained [Ca2+]i increase using the fura-2 fluorescent method and the whole-cell patch clamp technique. ATP (1 mM) induced a membrane current with the reversal potential of 12.5 ± 0.8 mV (n = 10) in Tyrode external solution. The EC50 of ATP was ~0.75 mM. The permeability ratio of various cations carrying this current was Na+ (defined as 1) > Li+ (0.92 ± 0.01; n = 5) > K+ (0.89 ± 0.03; n = 6) > Rb+ (0.55 ± 0.02; n = 6) > Cs+ (0.51 ± 0.01; n = 5) > Ca2+ (0.22 ± 0.03; n = 3) > N-methyl-d -glucamine (0.13 ± 0.01; n = 5), suggesting that ATP activated a nonselective cation current. The ATP-induced current was larger at lower concentrations of external Mg2+. ATP analogues that induced the current were 2-methylthio-ATP (2MeSATP), benzoylbenzoic-ATP, adenosine 5′-thiotriphosphate (ATPγS), and adenosine 5′-O-(2-thiodiphosphate), but not adenosine, ADP, α,β-methylene-ATP (AMPCPP), β,γ-methylene-ATP (AMPPCP), or UTP. Concomitant with the current data, 2MeSATP and ATPγS, but not AMPCPP or AMPPCP, increased the sustained [Ca2+]i increase. We conclude that ATP activates a class of Ca2+-permeable nonselective cation channels via the P2z receptor in NG108-15 cells.  相似文献   

8.
Abstract: High-affinity μ-opioid receptors have been solubilized from rat brain membranes. In most experiments, rats were treated for 14 days with naltrexone to increase the density of opioid receptors in brain membranes. Occupancy of the membrane-associated receptors with morphine during solubilization in the detergent 3-[(3-cholamidopropyl)dimethyl]-1-propane sulfonate appeared to stabilize the μ-opioid receptor. After removal of free morphine by Sephadex G50 chromatography and adjustment of the 3-[(3-cholamidopropyl)dimethyl]-1-propane sulfonate concentration to 3 mM, the solubilized opioid receptor bound [3H][d -Ala2,N-Me-Phe4,Gly-ol5]-enkephalin ([3H]DAMGO), a μ-selective opioid agonist, with high affinity (KD = 1.90 ± 0.93 nM; Bmax = 629 ± 162 fmol/mg of protein). Of the membrane-associated [3H]-DAMGO binding sites, 29 ± 7% were recovered in the solubilized fraction. Specific [3H]DAMGO binding was completely abolished in the presence of 10 µM guanosine 5′-O-(3-thiotriphosphate). The solubilized receptor also bound [3H]diprenorphine, a nonselective opioid antagonist, with high affinity (KD = 1.4 ± 0.39 nM, Bmax = 920 ± 154 fmol/mg of protein). Guanosine 5′-O-(3-thiotriphosphate) did not diminish [3H]diprenorphine binding. DAMGO at concentrations between 1 nM and 1 µM competed with [3H]diprenorphine for the solubilized binding sites; in contrast, [d -Pen2,d -Pen5]-enkephalin, a δ-selective opioid agonist, and U50488H, a κ-selective opioid agonist, failed to compete with [3H]diprenorphine for the solubilized binding sites at concentrations of <1 µM. In the absence of guanine nucleotides, the DAMGO displacement curve for [3H]diprenorphine binding sites better fit a two-site than a one-site model with KDhigh = 2.17 ± 1.5 nM, Bmax = 648 ± 110 fmol/mg of protein and KDlow = 468 ± 63 nM, Bmax = 253 ± 84 fmol/mg of protein. In the presence of 10 µM guanosine 5′-O-(3-thiotriphosphate), the DAMGO displacement curve better fit a one- than a two-site model with KD = 815 ± 33 nM, Bmax = 965 ± 124 fmol/mg of protein.  相似文献   

9.
A Malvern laser light-scattering instrument has been modified for use at scattering angles down to 5° and both total intensity and quasi-elastic scattering experiments. A sample of sheared, length-fractionated calf-thymus DNA was characterized by sedimentation, viscosity and electron microscopy. Quasi-elastic scattering and absolute intensity determinations were performed with the laser instrument and intensity determinations only with a Fica conventional light-scattering photometer. The total intensity experiments gave M?w = (3.75 ± 0.15) × 106 and 〈R21/2z = (206.9 ± 10.3) nm which yielded a value for the persistence length, allowing for polydispersity, of 66 ± 6nm. The quasi-elastic experiments at scattering angles below 20° gave D020, w = (2.23 ± 0.06) × 10?8 cm2/sec which combined with S020, w = 15.6 in the Svedberg equation gave M?w = (3.73 ± 0.18) × 106. In addition, from the higher angle data we extracted a value of the longest intramolecular relaxation time, τ1 of 17.5 msec. This is not in particularly good agreement with τ1 predicted by the Zimm–Rouse theory using our other experimental parameters. The disagreement may be due to the restricted applicability of the Zimm–Rouse spring-bead model as a quantitative representation of DNA molecules. Alternatively, it may be due to present difficulties in the unambiguous interpretation of molecular motions from the experimental autocorrelation functions.  相似文献   

10.
The effects of (±)-, (+)-, and (?)-atenolol, sotalol, and amosulalol alone on the rat left atria and portal vein and on the respective β1- and β2-adrenoceptor-mediated responses to isoprenaline have been determined. (±)-Atenolol at 10?6 M had no effect whereas high concentrations of (+)- and (?)-sotalol, 10?5–10?4 M, and (±)-, (+)-, and (?)-amosulalol depressed the response of the rat left atria to cardiac stimulation which indicates membrane stabilizing activity. None of the drugs tested had any effect alone on the rat portal vein. The order of potency as antagonists was (±)-amosulalol > (±)-atenolol > (±)-sotalol at β1-adrenoceptors and (±)-amosulalol > (±)-sotalol > (±)-atenolol at β2-adrenoceptors. (±)-Atenolol and (±)-amosulalol are β1-selective whereas (±)-sotalol is β2-selective. For each of the racemic β-blockers, the β1- and β2-adrenoceptor blocking activity was predominantly due to the (?)-enantiomer. © 1993 Wiley-Liss, Inc.  相似文献   

11.
The synthesis and biological evaluation of '6-(1,3-dihydroxyisobutyl)thymine' (DHBT; 1), which corresponds to 6-[3-hydroxy-2-(hydroxymethyl)propyl]-5-methylpyrimidine-2,4(1H,3H)-dione, is reported. DHBT (1) was designed as a new substrate for herpes simplex virus type-1 thymidine kinase (HSV1 TK). The compound was found to be exclusively phosphorylated by HSV1 TK, and to exhibit good binding affinity (Ki = 35.3+/-1.3 microM). Cell-proliferation assays with HSV1-TK-transduced human osteosarcoma cells (143B-TK+-HSV1-WT) and with both human-thymidine-kinase-1-negative (143B-TK-) and non-transduced parental (MG-63) cells indicate that 1 is less cytotoxic than the standard drug Ganciclovir. Thus, DHBT (1) represents a promising precursor of a nontoxic reporter probe for the monitoring of HSV1 TK gene expression by means of positron-emission tomography (PET).  相似文献   

12.
Laser light-scattering has been used to investigate the size of native proteoglycan aggregates (PGA-aA1) from day-8 chick limb-bud chondrocyte cultures isolated under associative extraction and purification conditions in 0.4M guanidinium chloride (GdnHCl) solution. Dynamic light-scattering measurements yielded a hydrodynamic radius, Rs, of 244 ± 10 nm for PGA-aA1 in 0.4M GdnHCl, and a weight-average molecular weight (M w) of 150 ± 50 × 106 was obtained from a Zimm plot. Disaggregation in 4.0M GdnHCl aqueous solution yielded proteoglycan subunits (PGS) with Rs = 39 ± 2 nm, M w = 1.6 ± 0.3 × 106, which reassembled in 0.4M GdnHCl to form “reconstituted native” aggregates (PGA-raA1) with Rs = 121 ± 6 nm, M w = 17 ± 3 × 106. A second specimen of PGA-aA1 had Rs = 192 ± 10 nm, M w = 100 ± 10 × 106. The latter value was estimated from an empirical relationship between M w and Rs. After dissociation, this specimen reassembled to form PGA-raA1 with Rs = 85 ± 5 nm, M w = 12 ± 1 × 106. These data are compared with those for a specimen of reconstituted aggregate (PGA-A1) that had been extracted under dissociative conditions and then reaggregated by dialysis to 0.4M GdnHCl aqueous solution, for which Rs = 138 ± 9 nm, M w = 45 ± 8 × 106. From these values, we have calculated the weight-average number of subunits per aggregate Nw: 111 for PGA-aA1 and 12 for raA1 (70 and 7 for the second PGA-aA1 and PGA-raA1 specimen, respectively) as compared to 32 for PGA-A1. The numbers of subunits per aggregate were also determined from electron micrographs of spread specimens. The latter results show the same trends as those obtained by light scattering, but lead in each case to lower numbers of subunits per aggregate. These data demonstrate conclusively that PGA samples exhibit a higher degree of aggregation in solution than visualized in typical electron microscopy (EM) preparations, probably due to disaggregation during EM specimen preparation. Since Nw determined both by light scattering (LS) and by EM are larger for native versus reconstituted aggregate samples, our data point to a more compact aggregation of subunits along the hyaluronic acid (HA) chains in the former.  相似文献   

13.
A series of experiments were carried out to construct an energy budget for juvenile thick lipped mullet, Crenimugil labrosus Risso. A partial factorial experimental design was used to examine the effects of temperature, fish size and meal size on growth. The maximum ration that the fish were able to ingest completely per day was found to be 0·8, 1·4 and 2·3% wet body weight (b.w.) at 13,18 and 23°C, respectively. Ingested maintenance requirements (M.R.) were estimated to be 137, 205 and 288 cal fish-1 day-1 at 13, 18 and 23°C, respectively. At 18deg; C, M.R. varied as 25 W1.04 cal day-1, where W= fish weight (g). Growth rate increased with increasing temperature. Maximal conversion efficiency was 21–24% and was achieved closer to the maximum ingested ration with increasing temperature. The relationship between respiration rate and W at 18deg; C for 3-20 g fish is described by: respiration rate (ml O2 h-1) = 0·128 W0.976 The energy cost of apparent specific dynamic action at 18deg; C was found to vary between 5·1% and 23·6% of the calorific value of the ingested meal (1% wet b.w.) , mean (± S.E.)=10·2 ± 2·0%. Post mortem analyses of groups of fish fed 0·2, 0·8 and 1·5% wet b.w. meals showed a significant increase in total lipid and a significant decrease in water content with increasing ratio size. A negative correlation was found between body water content and total lipid (and calories). The mean assimilation efficiency (±s.e.) for 5–10 g mullet at 18deg; C was 73·9 ± 3·6%. The observations reported in this study were brought together to construct an energy budget for juvenile C. labrosus which was found to give a reliable prediction (within 10%) of energy demand and growth under the prevailing experimental conditions. Both gross (K1) and net (K2) growth efficiencies, based on energy values, increased with increasing ratio size up to satiation and were independent of temperature. The maximum values of K1 and K2 observed were 0·33 and 0·46, respectively. The third order efficiency (K3) appeared to be independent of temperature and ration size; mean values ranged between 0·66 and 0·84.  相似文献   

14.
The patch clamp K+-conductance G of the nicotinic acetylcholine receptor (AcChoR) dimer (Mr≈ 590 000) of Torpedo californica, reconstituted in lipid vesicles, which decreases with increasing Ca2+-concentration in the range 0.1≤[Ca2+]/mM≤2, can be quantitatively rationalized by Ca2+-binding to negatively charged sites, causing charge reversal reducing the normal K+-accumulation in the channel vestibules. Cleavage of the sialic acid residues (up to 20±2 per dimer) reduces the K+-accumulation factor α = G0/G from α = 3±0.8 of the normal AcChoR to α = 2±0.7 for the desialyated AcChoR. Desialysation also decreases the Ca2+-sensitivity of the conductance from G0 = 96.6±6 pS at [Ca2+]→0 of the normal AcChoR to G0 = 84.2±6 pS. Endogenous hyperphosphorylation (to up to 28±4 phosphates per dimer) enhances the vestibular K+-accumulation to α = 3.6±0.7, without affecting the Ca2+-dissociation equilibrium constant KCa = 0.34± 0.05 mM at 295 K (22 °C). Most interestingly, even in the absence of AcCho, the hyperphosphorylated AcChoR dimer exhibits spontaneously long-lasting open channel events (τ = 200±50 ms). At [AcCho] = 2 μM there are two open states (τ 1 = 20±10 ms, τ 2 = 140±60 ms) whereas the normal AcChoR dimer has only one open state (τ = 6±4 ms). – Physiologically important is that (i) the sialic acid and phosphate residues render the AcChoR conductance sensitive to control by divalent ions and (ii) the channel behavior of the hyperphosphorylated AcChoR without AcCho appears to indicate pathophysiologically high phosphorylation activity of the cell leading, among others, to myasthenic syndromes. Received: 10 November 1997 / Revised version: 12 January 1998 / Accepted: 7 March 1998  相似文献   

15.
Corticosteroid Modulation of Signal Transduction in the CATH.a Cell Line   总被引:1,自引:0,他引:1  
Abstract: Noradrenergic neuronal networks originating in the locus coeruleus have been implicated in the stress response. In order to study this system in vitro, we have employed a locus coeruleus-like cell line, CATH.a, and have determined the effect of dexamethasone on receptor-mediated second messenger responses. The CATH.a cell line produced increases in intracellular cyclic AMP conversion in response to corticotrophin-releasing factor (EC50 = 6.93 ± 1.26 nM, maximum conversion = 4.11 ± 0.20%) and vasoactive intestinal polypeptide (EC50 = 240 ± 40 nM, maximum conversion = 8.92 ± 1.24%). Forskolin (10 µM) increased conversion from 0.48 ± 0.05 to 6.39 ± 0.38%. The α2-adrenoceptor agonist 5-bromo-N-(4,5-dihydro-1H-imidazol-2-yl)-6-quinoxalinamine (UK14304) inhibited the forskolin response with an IC50 of 6.76 ± 0.11 nM. Carbachol increased total 3H-labelled inositol phosphate accumulation to a maximum of 3.01 ± 0.79 fold basal (EC50 = 7.94 ± 0.14 µM). Bradykinin produced a maximum 1.81 ± 0.05 fold basal stimulation of phosphoinositide hydrolysis (EC50 = 9.12 ± 0.16 nM). Both carbachol and bradykinin increased intracellular Ca2+ concentration probably via a combination of mobilisation of intracellular stores and gating of extracellular Ca2+. Incubation for 24 h with the glucocorticoid receptor agonist, dexamethasone (1 µM), significantly potentiated the receptor-mediated phosphoinositide responses to all the agents tested; however, of the receptor-mediated increases in cyclic AMP conversion, only the vasoactive intestinal polypeptide response was potentiated. These results show that the CATH.a cell line displays some of the properties expected of locus coeruleus neurons and that glucocorticoid receptor stimulation selectively modulates receptor-mediated increases in second messenger formation.  相似文献   

16.
Objective: Higher levels of physical activity have been associated with body weight maintenance, but previous work in our laboratory suggests that this is not purely related to energy balance. We hypothesize that this may be related to the partitioning of dietary fat between oxidation and storage. Research Methods and Procedures: Healthy women (age 24 ± 1 years, BMI = 21.2 ± 0.4 kg/m2) were recruited to participate in rest (n = 10) or exercise sessions of light (n = 11), moderate (n = 10), and heavy (n = 7) exercise. All exercises (1250 kJ above rest) were performed on a stationary cycle inside of a whole‐body calorimeter. [1‐13C]oleate and [d31]palmitate were given in a liquid meal 30 minutes post‐exercise. An additional study was done with identical exercise sessions, but with administration of an oral dose of [1‐13C]acetate and [d3]acetate 30 minutes post‐exercise to determine label sequestration. Results: Cumulative oxidation of [1‐13C]oleate was significantly greater after light (45 ± 3%), moderate (54 ± 4%), and heavy (51 ± 4%) exercise than that with rest (33 ± 3%) (p = 0.0008). Cumulative oxidation of [d31]palmitate did not differ among trials (12 ± 2%, 14 ± 1%, 17 ± 2%, and 14 ± 2% for rest, light, moderate, and heavy, respectively; p = 0.30). Discussion: Exercise standardized for energy expenditure increases monounsaturated fat oxidation more than saturated fat oxidation and that the increase occurs regardless of intensity. Recommendations for physical activity for the purposes of weight control may be specific for dietary fat composition.  相似文献   

17.
The present study attempts to assess the potential of artificial substrates to enhance fish production in inland saline groundwater ponds through periphyton production. Grey mullet, Mugil cephalus, was cultured for 100 days in ponds with substrate (treatment ponds) and without substrate (control ponds). To enhance the surface area, bamboo poles were used as substrate. The periphyton population, pigment concentration and hydrobiological characteristics of pond water were monitored. The studies revealed little difference in most of the water quality parameters observed in the two treatments. However, turbidity (27.0 ± 0.1–35.0 ± 0.1 Nephalo Turbidity Unit (NTU)), chlorophyll ‘a’ (6.6 ± 0.6–7.6 ± 0.6 μg L?1), plankton population (phytoplankton 8.4 × 103–9.4 ×103 numbers L?1; zooplankton 4.0 × 103–5.1 × 103 numbers L?1) and NH4–N (2.0 ± 0.2–2.3 ± 0.1 mg L?1) were high in the treatment with no additional substrate; however, in the treatment with substrate the total Kjeldahl nitrogen (9.8 ± 0.8–10.8 ± 0.7 mg L?1) and o‐PO4 (0.1 ± 0.01–0.1 mg L?1) remained significantly (P < 0.05) higher. Highest periphyton biomass in terms of dry matter (DM) (0.8 ± 0.01–1.4 ±0.01 mg cm?2), ash free DM (0.4 ± 0.0–0.6 ± 0.01 mg cm?2), chlorophyll ‘a’ (3.1 ± 0.2–8.1 ± 0.8 μg cm?2) and pheophytin ‘a’ (1.9 ± 0.4–3.9 ± 0.5 μg cm?2) was observed at 50 cm depth in ponds provided with additional substrate. Fifteen plankton genera showing periphytic affinity colonized the bamboo substrates. Fish growth (mean fish weight 524.3 ± 8.7 g and SGR 2.5 ± 0.1) was significantly (P < 0.05) higher in ponds provided with additional substrate compared with control ponds (387.2 ± 6.0). Length–weight relationship (LWR) (W = cLn) also showed that the exponential value (‘n’) of length was high in substrate‐supported ponds (n = 2.36) in comparison with controls (n = 1.09). These studies suggest that a periphyton‐supported aquaculture system can be used successfully for the culture of herbivorous brackishwater fish species like M. cephalus in inland saline groundwaters and thus could contribute to the development of sound and sustainable aquaculture technology.  相似文献   

18.
Cannabinoid CB1 receptors are highly expressed in the striatum where they are known to be co‐localized with dopamine D2 receptors. There is now strong evidence that cannabinoids modulate dopamine release in the brain. Using fast cyclic voltammetry, single pulse stimulation (0.1 ms; 10 V) was applied every 5 min and peak dopamine release was measured with a carbon fibre microelectrode. Application of the D2 receptor agonist, quinpirole, inhibited single pulse dopamine overflow in a concentration‐dependent manner (IC50: 3.25 × 10?8 M). The CB1 receptor agonist WIN55212‐2 (WIN; 1 μM) had no effect on single pulse dopamine release (93.9 ± 6.6% at 60 min, n = 5) but attenuated the inhibitory effect of quinpirole (30 nM; quinpirole 39.0 ± 4.2% vs. quinpirole + WIN, 48.2 ± 3.7%, n = 5, p < 0.05). This affect was antagonized by the CB1 receptor anatgonist [N‐(Piperidin‐1‐yl)‐5‐(4‐iodophenyl)‐1‐(2,4‐dichlorophenyl)‐4‐methyl‐1H‐pyrazole‐3‐carboxamide] (AM‐251, 1 μM). Dopamine release evoked by four pulses delivered at 1 Hz (4P1Hz) and 10 pulses delivered at 5 Hz (10P5Hz) was significantly inhibited by WIN [72.3 ± 7.9% control (peak 4 to 1 ratio measurement) and 66.9 ± 3.8% control (area under the curve measurement), respectively, p < 0.05; n = 6 for both]. Prior perfusion of WIN significantly attenuated the effects of quinpirole on multiple pulse‐evoked dopamine release (4P1Hz: quinpirole, 28.4 ± 4.8% vs. WIN + quinpirole, 52.3 ± 1.2%; 10P5Hz: quinpirole, 29.5 ± 1.3% vs. WIN + quinpirole, 59.4 ±7.1%; p < 0.05 for both; n = 6). These effects were also antagonized by AM‐251 (1 μM). This is the first report demonstrating a functional, antagonistic interaction between CB1 receptors and D2 autoreceptors in regulating rat striatal dopamine release.  相似文献   

19.
A laboratory-made sample of the polysaccharide xylinan (acetan) has been further characterized with respect to (i) purity, (ii) molar mass and polydispersity, and (iii) gross conformation by a combination of hydrodynamic measurements (sedimentation velocity and equilibrium analytical ultracentrifugation, viscometry, and dynamic light scattering) in aqueous NaCl (I = 0.10 mol·L−1). Sedimentation velocity diagrams recorded using Schlieren optics revealed highly pure material sedimenting as a single boundary [so20.w = 9.5 ± 0.7) S; ks = (273 ± 112) mL/g]. The hypersharp nature of these boundaries is symptomatic of a polydisperse and highly nonideal (in the thermodynamic sense) system. Low speed sedimentation equilibrium in the analytical ultracentrifuge using Rayleigh interference optics and two different types of extrapolation procedure (involving point and whole-cell molar masses) gave a weight average molar mass Mw of (2.5 ± 0.5) × 10−6 g·mol−1 and also a second virial coefficient, B = (2.8 ± 0.7) × 10−4 mL·mol·g−2, both values in good agreement with those from light scattering-based procedures (Part II of this series). A dynamic Zimm plot from dynamic light scattering measurements gave a z-average translational diffusion coefficient Do20.w = (3.02 ± 0.05) × 10−8 cm2·s−1 and the concentration-dependence parameter kD = (370 ± 15) mL/g. Combination of so20.w with Do20.w via the Svedberg equation gave another estimate for Mw of ≅ 2.4 × 106 g/mol, again in good agreement. Both the Wales-van Holde ratio (ks/[η]) ≅ 0.4 (with [η] = (760 ± 77) mL/g) and the ρ-parameter (ratio of the radius of gyration from static light scattering to the hydrodynamic radius from dynamic light scattering) as ρ > 2.0 all indicate an extended conformation for the macromolecules in solution. These findings, plus Rinde-type simulations of the sedimentation equilibrium data are all consistent with the interpretation in terms of a unimodal wormlike coil model performed earlier. © 1996 John Wiley & Sons, Inc.  相似文献   

20.
The objective of this study was to determine the effect of freezing on the function in Atlantic salmon Salmo salar spermatozoa. The semen was frozen in Cortland's medium + 1.3M dimethyl sulphoxide + 0.3M glucose + 2% bovine serum albumin (final concentration) in a ratio of 1:3 (semen:cryoprotectant) as the treatment (T) and fresh semen as the control (F). Straws of 0·5 ml of sperm suspension were frozen in 4 cm of N2L. They were thawed in a thermoregulated bath (40° C). After thawing, the percentage of spermatozoa with fragmented DNA [transferase dUTP (deoxyuridine triphosphate) nick‐end labelling (TUNEL)], plasma membrane integrity (SYBR‐14/PI) and mitochondrial membrane potential (ΔΨMMit, JC‐1) were evaluated by flow cytometry and motility was evaluated by optical microscope under stroboscopic light. The fertilization rates of the control and treatment semen were tested at a sperm density of 1·5 × 107 spermatozoa oocyte?1, by observation of the first cleavages after 16 h incubation at 10° C. In the cryopreserved semen (T), the mean ± s.d . DNA fragmentation was 4·8 ± 2·5%; plasma membrane integrity 75·2 ± 6·3%; mitochondrial membrane potential 51·7 ± 3·6%; motility 58·5 ± 5·3%; curved line velocity (VCL) 61·2 ± 17·4 µm s?1; average‐path velocity (VAP) 50·1 ± 17·3 µm s?1; straight‐line velocity (VSL) 59·1 ± 18·4 µm s?1; fertilization rate 81·6 ± 1·9%. There were significant differences in the plasma membrane integrity, mitochondrial membrane potential, motility, fertilization rate, VCL, VAP and VSL compared with the controls (P < 0·05). Also the mitochondrial membrane potential correlated with motility, fertilization rate, VCL and VSL (r = 0·75; r = 0·59; r = 0·77 and r = 0·79, respectively; P < 0·05); and the fertilization rate correlated with VCL and VSL (r = 0·59 and r = 0·55, respectively).  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