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.
Effect of ATP on the Calcium Efflux in Dialyzed Squid Giant Axons   总被引:12,自引:9,他引:3       下载免费PDF全文
Dialysis perfusion technique makes it possible to control the internal composition of squid giant axons. Calcium efflux has been studied in the presence and in the virtual absence (<5 µM) of ATP. The mean calcium efflux from axons dialyzed with 0.3 µM ionized calcium, [ATP]i > 1,000 µM, and bathed in artificial seawater (ASW) was 0.24 ± 0.02 pmol·cm-2·s-1 (P/CS) (n = 8) at 22°C. With [ATP]i < 5 µM the mean efflux was 0.11 ± 0.01 P/CS (n = 15). The curve relating calcium efflux to [ATP]i shows a constant residual calcium efflux in the range of 1–100 µM [ATP]i. An increase of the calcium efflux is observed when [ATP]i is >100 µM and saturates at [ATP]i > 1,000 µM. The magnitude of the ATP-dependent fraction of the calcium efflux varies with external concentrations of Na+, Ca++, and Mg++. These results suggest that internal ATP changes the affinity of the calcium transport system for external cations.  相似文献   

2.
1. Superovulated rat ovary slices from rats treated with 20μg. of luteininzing hormone/100g. body wt. 2hr. before death and from control animals have been incubated in vitro. Output of Δ4-3-oxo steroids (0·2μmole/g. wet wt./hr. in control tissue) was linear for 4hr., and was increased by approx. 70% in slices from luteinizing hormone-treated rats. Rate of oxygen consumption (90·0±4·6μmoles/g. wet wt./hr.) was linear for 3hr. and unaltered by luteinizing hormone treatment or addition of glucose (1mg./ml.) to the medium. 2. In slices from control animals, steady-state rate of glucose uptake was 78·0±2·9μg. atoms of carbon/g. wet wt./hr.; steady-state rates of lactate output, pyruvate output and incorporation of [U-14C]-glucose carbon atoms into carbon dioxide and total lipid extract were 60·7±0·9, 2·4±0·1, 18·0±1·1 and 0·7±0·1μg. atom of carbon/g. wet wt./hr. and accounted for 104·5±1·9% of the glucose uptake. In slices from luteinizing hormone-treated rats, glucose uptake and outputs of lactate, pyruvate and [14C]carbon dioxide were increased by approx. 25%, and 108·4±3·2% of the glucose uptake could be accounted for. 3. The total lipid extract was separated by thin-layer chromatography and saponification. Of the 14C incorporated into this fraction during incubation with [U-14C]glucose 97% was found in the fractions containing glyceride glycerol and less than 3% in the fractions containing sterols, steroids or fatty acids. Appreciable quantities of 14C were incorporated into these lipid fractions from [1-14C]acetate. 4. From a consideration of the tissue glycogen content, the specific activities of [14C]lactate and glucose 6-phosphate (C-1) derived from [1-14C]-, [6-14C]- and [U-14C]-glucose, and the ratio of [14C]carbon dioxide yields from [1-14C]glucose and [6-14C]glucose, it was concluded that there was no appreciable glycogenolysis or flow through the pentose phosphate cycle. 5. In ovary slices from both control and luteinizing hormone-treated animals, glucose in vitro raised the incorporation rate of 14C from [1-14C]acetate into sterols and steroids. Luteinizing hormone in vivo stimulated the incorporation rate in vitro but only in the presence of glucose. 6. In slices incubated in medium containing [3H]water, [14C]sorbitol and glucose (1mg./ml.), the total water space (865±7·1μl./g.) and the extracellular water space (581±22μl./g.) were unchanged by luteinizing hormone treatment in vivo but the glucose space was raised from 540±23·6μl./g. to 639±31·3μl./g. 7. Luteinizing hormone treatment was found to lower the tissue concentration of the hexose monophosphates and to increase the total activity of hexokinase, glucose 6-phosphate dehydrogenase and 6-phosphogluconate dehydrogenase and possibly of phosphofructokinase. 8. The kinetic properties of a partially purified preparation of phosphofructokinase were found to be qualitatively similar to those from other mammalian tissues. 9. The results are discussed with reference to both the role of glucose metabolism in steroidogenesis and the mechanism by which luteinizing hormone increases the rate of glucose uptake.  相似文献   

3.
The conversion factor for the calculation of bacterial production from rates of [3H]thymidine incorporation was examined with diluted batch cultures of freshwater bacteria. Natural bacterial assemblages were grown in aged, normal, and enriched media at 10 to 20°C. The generation time during 101 growth cycles covered a range from 4 to >200 h. The average conversion factor was 2.15 × 1018 cells mol-1 of thymidine incorporated into the trichloroacetic acid (TCA) precipitate (standard error = 0.29 × 1018; n = 54), when the generation time exceeded 20 h. At generation times of <20 h, the average conversion factor was 11.8 × 1018 cells mol-1 of thymidine incorporated into TCA precipitate (standard error = 1.72 × 1018; n = 47). The amount of radioactivity in purified DNA increased with decreasing generation time and increasing conversion factor (calculated from the TCA precipitate), corresponding to a decrease in the percentage in protein. The conversion factors calculated from purified DNA or from the TCA precipitate gave the same variability. Conversion factors did not change significantly with the medium, but were significantly higher at 20°C than at 15 and 10°C. A detailed examination of the [3H]thymidine concentrations that were needed to achieve maximum labeling in DNA was carried out 6 times during a complete growth cycle. During periods with low generation times and high conversion factors, 15 nM [3H]thymidine was enough for the maximum labeling of the TCA precipitate. This suggests that incorporation of [3H]thymidine into DNA is probably limited by uptake during periods with generation times of <20 h and that freshwater bacterioplankton cell production sometimes is underestimated when a conversion factor of 2.15 × 1018 cells mol-1 of thymidine incorporated is used.  相似文献   

4.
The metabolism of [2-14C]indole in the rat   总被引:3,自引:1,他引:2  
1. [2-14C]Indole has been synthesized from [14C]formate and o-toluidine via N[14C]-formyltoluidine. 2. When fed to rats, the 14C of [14C]indole (dose 70–80mg./kg. body wt.) is fairly rapidly excreted, and in 2 days an average of 81% appears in the urine, 11% in the faeces and 2·4% as carbon dioxide in the expired air. 3. Radioactivity is excreted in the urine as indoxyl sulphate (50% of the dose), indoxyl glucuronide (11%), oxindole (1·4%), isatin (5·8%), 5-hydroxyoxindole conjugates (3·1%), N-formylanthranilic acid (0·5%) and unchanged indole (0·07%). The faeces contain indoxyl sulphate (0·4% of the dose) and indole (0·2%), but the major metabolites have not been identified. 4. Fed to rats with biliary cannulae an average of 5·6% of a dose of [14C]indole (20–60mg./kg. body wt.) is excreted in the bile in 2 days. Radioactivity is present as indoxyl sulphate (0·8% dose) and 5-hydroxyoxindole conjugates (0·6%). 5. Rats further metabolize indoxyl into N-formylanthranilic acid and anthranilic acid, and oxindole into 5-hydroxyoxindole. 6. With rat-liver microsomes plus supernatant under aerobic conditions, indole gives indoxyl, oxindole, possibly isatin, N-formylanthranilic acid and anthranilic acid, but under anaerobic conditions gives only oxindole. Similarly, under aerobic conditions, oxindole gives 5-hydroxyoxindole, anthranilic acid and o-aminophenylacetic acid. 7. Indole is metabolized by two pathways, one via indoxyl to isatin, N-formylanthranilic acid and anthranilic acid, and the other via oxindole to 5-hydroxyoxindole and possibly to o-aminophenylacetic and anthranilic acid. 8. The following new compounds are described: 4-hydroxy-2-nitrophenylacetic acid, 3-, 4- and 5-benzyloxy-2-nitrophenylacetic acid, 5- and 7-hydroxyoxindole and 5-aminoacridine indoxyl sulphate.  相似文献   

5.
We investigated 3,3′,5-tri-iodo-l-thyronine transport by human erythrocytes and by `ghosts'' prepared from these cells. Uptake of tri-iodothyronine by erythrocytes at 37°C was time-dependent with a maximum reached after 60min. Tracer analysis after incubation for 1min revealed only one saturable binding site, with Km 128±19nm (mean±s.e.m.; n=7) and Vmax. 4.6±0.7pmol of tri-iodothyronine/min per 6×107 cells. After 10min incubation Km 100±16nm (n=10) was found with Vmax. 7.7±1.2pmol of tri-iodothyronine/10min per 6×107 cells. At 0°C the uptake system is still active, with Km 132±26nm and Vmax. 1.8±0.3pmol of tri-iodothyronine/10min per 6×107 cells. The Vmax. with intact cells is 5-fold greater than the Vmax. with membranes derived from the same amount of cells when uptake studies are performed in media with similar free tri-iodothyronine concentrations. This indicates that at least 80% of tri-iodothyronine taken up by the intact erythrocytes enters the cell. This saturable uptake system can be inhibited by X-ray-contrast agents in a dose-dependent fashion. (±)-Propranolol, but not atenolol, has the same effect, indicating that the membrane-stabilizing properties of (±)-propranolol are involved. Furthermore, there is no inhibition by ouabain or vanadate, which indicates that tri-iodothyronine uptake is not dependent on the activity of Na++K+-dependent adenosine triphosphatase. We have prepared erythrocyte `ghosts'', resealed after 2.5min with 0mm-, 2mm- or 4mm-ATP inside. Inclusion of ATP and integrity of the membrane of the erythrocyte `ghosts'' were verified on the basis of an ATP-concentration-dependent functioning of the Ca2+ pump. No difference was found in the uptake of tri-iodothyronine by erythrocyte `ghosts'' with or without ATP included, indicating that uptake of tri-iodothyronine is not ATP-dependent. The following conclusions are drawn. (1) Tri-iodothyronine enters human erythrocytes. (2) There is only one saturable uptake system present for tri-iodothyronine, which is neither energy (i.e. ATP)-dependent nor influenced by the absence of an Na+ gradient across the plasma membrane. This mode of uptake of tri-iodothyronine by human erythrocytes is in sharp contrast with that of rat hepatocytes, which uptake system is energy-dependent and ouabain-sensitive [Krenning, Docter, Bernard, Visser & Hennemann (1978) FEBS Lett. 91, 113–116; Krenning, Docter, Bernard, Visser & Hennemann (1980) FEBS Lett. 119, 279–282]. (3) X-ray-contrast agents inhibit tri-iodothyronine uptake by erythrocytes in a similar fashion to that by which they inhibit the uptake of tri-iodothyronine by rat hepatocytes [Krenning, Docter, Bernard, Visser & Hennemann (1982) FEBS Lett. 140, 229–233].  相似文献   

6.
Catalytic subunits (C) of uterine smooth-muscle adenylate cyclase were activated (C*) by incubating the enzyme with the GTP analogue guanosine 5′-[βγ-imido]triphosphate (p[NH]ppG), followed by treatment with GTP and washing at 2°C. Activation (C→C*) proceeded in a time- and temperature-dependent manner as disclosed by subsequent assay of the pretreated particles at 37°C. The properties of the activated subunits were a function of the pretreatment temperature and not those of the enzyme assay performed at 37°C. Over the range 6–24°C, activation by pretreatment with p[NH]ppG followed simple Michaelis–Menten kinetics, and increase in temperature increased the concentration of catalytic subunits in the C* state and decreased Km for the guanosine nucleotide. Characterization of the temperature-dependent effects of pretreatment with p[NH]ppG suggested that activation of the catalytic subunit at the temperature in situ (37°C) was moderately endergonic (ΔH0 ~8kJ·mol−1) and accompanied by an increase in entropy (ΔS0 ~146J·mol−1·K−1). The β-adrenergic catecholamine receptor, reflected by isoproterenol's effect on activation by pretreatment with p[NH]ppG, increased the concentration of catalytic subunits in the C* state but had an insignificant (P>0.05) effect on the Km at every temperature. This result suggested that formation of the receptor–hormone complex produced an increase in the first-order rate constant without an appreciable effect on the actual catalytic-subunit activation step. The primary function of the β-adrenergic catecholamine receptor under these conditions appeared to be regulation of the concentration of activation sites available for binding of p[NH]ppG.  相似文献   

7.

Purpose

Neuroinflammation is involved in several brain disorders and can be monitored through expression of the translocator protein 18 kDa (TSPO) on activated microglia. In recent years, several new PET radioligands for TSPO have been evaluated in disease models. [18F]DPA-714 is a TSPO radiotracer with great promise; however results vary between different experimental models of neuroinflammation. To further examine the potential of [18F]DPA-714, it was compared directly to [11C]PK11195 in experimental cerebral ischaemia in rats.

Methods

Under anaesthesia, the middle cerebral artery of adult rats was occluded for 60 min using the filament model. Rats were allowed recovery for 5 to 7 days before one hour dynamic PET scans with [11C]PK11195 and/or [18F]DPA-714 under anaesthesia.

Results

Uptake of [11C]PK11195 vs [18F]DPA-714 in the ischemic lesion was similar (core/contralateral ratio: 2.84±0.67 vs 2.28±0.34 respectively), but severity of the brain ischemia and hence ligand uptake in the lesion appeared to vary greatly between animals scanned with [11C]PK11195 or with [18F]DPA-714. To solve this issue of inter-individual variability, we performed a direct comparison of [11C]PK11195 and [18F]DPA-714 by scanning the same animals sequentially with both tracers within 24 h. In this direct comparison, the core/contralateral ratio (3.35±1.21 vs 4.66±2.50 for [11C]PK11195 vs [18F]DPA-714 respectively) showed a significantly better signal-to-noise ratio (1.6 (1.3–1.9, 95%CI) fold by linear regression) for [18F]DPA-714.

Conclusions

In a clinically relevant model of neuroinflammation, uptake for both radiotracers appeared to be similar at first, but a high variability was observed in our model. Therefore, to truly compare tracers in such models, we performed scans with both tracers in the same animals. By doing so, our result demonstrated that [18F]DPA-714 displayed a higher signal-to-noise ratio than [11C]PK11195. Our results suggest that, with the longer half-life of [18F] which facilitates distribution of the tracer across PET centre, [18F]DPA-714 is a good alternative for TSPO imaging.  相似文献   

8.
Accurate measures of plasma FA oxidation can improve our understanding of diseases characterized by impaired FA oxidation. We describe and compare the 24 h time-courses of FA oxidation using bolus injections of [1-14C]palmitate versus [9,10-3H]palmitate under postabsorptive, postprandial, and walking conditions. Fifty-one men and 95 premenopausal women participated in one condition (postabsorptive, postprandial, or walking), one tracer (14C- or 3H-labeled), and an acetate or palmitate study. Groups were matched for sex, age, and body mass index (BMI). At 24 h, cumulative [3H]acetate recovery as 3H2O was 80 ± 6%, 78 ± 2%, and 81 ± 6% in the postabsorptive, postprandial, and walking conditions, respectively (not significant). Model-predicted maximum [1-14C]acetate recovery as expired 14CO2 was 59 ± 12%, 52 ± 8%, and 65 ± 10% in the postabsorptive, postprandial, and walking condition, respectively (one way ANOVA, P = 0.12). When corrected with the corresponding acetate recovery factors, 24 h time-courses of FFA oxidation were similar between [1-14C]palmitate and [9,10-3H]palmitate in all three conditions. In contrast to previous meal ingestion studies, an acetate-hydrogen recovery factor was needed to achieve comparable oxidation rates using an intravenous bolus of [3H]palmitate. In conclusion, intravenous boluses of [9,10-3H]palmitate versus [1-14C]palmitate gave similar estimates of 24 h cumulative FFA oxidation in age-, sex- and BMI-matched individuals.  相似文献   

9.
The human pathogen Shigella flexneri subverts host function and defenses by deploying a cohort of effector proteins via a type III secretion system. The IpaH family of 10 such effectors mimics ubiquitin ligases but bears no sequence or structural homology to their eukaryotic counterpoints. Using rates of 125I-polyubiquitin chain formation as a functional read out, IpaH9.8 displays V-type positive cooperativity with respect to varying concentrations of its Ubc5B∼125I-ubiquitin thioester co-substrate in the nanomolar range ([S]½ = 140 ± 32 nm; n = 1.8 ± 0.1) and cooperative substrate inhibition at micromolar concentrations ([S]½ = 740 ± 240 nm; n = 1.7 ± 0.2), requiring ordered binding to two functionally distinct sites per subunit. The isosteric substrate analog Ubc5BC85S-ubiquitin oxyester acts as a competitive inhibitor of wild-type Ubc5B∼125I-ubiquitin thioester (Ki = 117 ± 29 nm), whereas a Ubc5BC85A product analog shows noncompetitive inhibition (Ki = 2.2 ± 0.5 μm), consistent with the two-site model. Re-evaluation of a related IpaH3 crystal structure (PDB entry 3CVR) identifies a symmetric dimer consistent with the observed cooperativity. Genetic disruption of the predicted IpaH9.8 dimer interface reduces the solution molecular weight and significantly ablates the kcat but not [S]½ for polyubiquitin chain formation. Other studies demonstrate that cooperativity requires the N-terminal leucine-rich repeat-targeting domain and is transduced through Phe395. Additionally, these mechanistic features are conserved in a distantly related SspH2 Salmonella enterica ligase. Kinetic parallels between IpaH9.8 and the recently revised mechanism for E6AP/UBE3A (Ronchi, V. P., Klein, J. M., and Haas, A. L. (2013) E6AP/UBE3A ubiquitin ligase harbors two E2∼ubiquitin binding sites. J. Biol. Chem. 288, 10349–10360) suggest convergent evolution of the catalytic mechanisms for prokaryotic and eukaryotic ligases.  相似文献   

10.
Kv7 K+-channel subunits differ in their apparent affinity for PIP2 and are differentially expressed in nerve, muscle, and epithelia in accord with their physiological roles in those tissues. To investigate how PIP2 affinity affects the response to physiological stimuli such as receptor stimulation, we exposed homomeric and heteromeric Kv7.2, 7.3, and 7.4 channels to a range of concentrations of the muscarinic receptor agonist oxotremorine-M (oxo-M) in a heterologous expression system. Activation of M1 receptors by oxo-M leads to PIP2 depletion through Gq and phospholipase C (PLC). Chinese hamster ovary cells were transiently transfected with Kv7 subunits and M1 receptors and studied under perforated-patch voltage clamp. For Kv7.2/7.3 heteromers, the EC50 for current suppression was 0.44 ± 0.08 µM, and the maximal inhibition (Inhibmax) was 74 ± 3% (n = 5–7). When tonic PIP2 abundance was increased by overexpression of PIP 5-kinase, the EC50 was shifted threefold to the right (1.2 ± 0.1 µM), but without a significant change in Inhibmax (73 ± 4%, n = 5). To investigate the muscarinic sensitivity of Kv7.3 homomers, we used the A315T pore mutant (Kv7.3T) that increases whole-cell currents by 30-fold without any change in apparent PIP2 affinity. Kv7.3T currents had a slightly right-shifted EC50 as compared with Kv7.2/7.3 heteromers (1.0 ± 0.8 µM) and a strongly reduced Inhibmax (39 ± 3%). In contrast, the dose–response curve of homomeric Kv7.4 channels was shifted considerably to the left (66 ± 8 nM), and Inhibmax was slightly increased (81 ± 6%, n = 3–4). We then studied several Kv7.2 mutants with altered apparent affinities for PIP2 by coexpressing them with Kv7.3T subunits to boost current amplitudes. For the lower affinity (Kv7.2 (R463Q)/Kv7.3T) or higher affinity (Kv7.2 (R463E)/Kv7.3T) channels, the EC50 and Inhibmax were similar to Kv7.4 or Kv7.3T homomers (0.12 ± 0.08 µM and 79 ± 6% [n = 3–4] and 0.58 ± 0.07 µM and 27 ± 3% [n = 3–4], respectively). The very low-affinity Kv7.2 (R452E, R459E, and R461E) triple mutant was also coexpressed with Kv7.3T. The resulting heteromer displayed a very low EC50 for inhibition (32 ± 8 nM) and a slightly increased Inhibmax (83 ± 3%, n = 3–4). We then constructed a cellular model that incorporates PLC activation by oxo-M, PIP2 hydrolysis, PIP2 binding to Kv7-channel subunits, and K+ current through Kv7 tetramers. We were able to fully reproduce our data and extract a consistent set of PIP2 affinities.  相似文献   

11.
In the intact eye, the transition from light to dark alters pH, [Ca2+], and [K] in the subretinal space (SRS) separating the photoreceptor outer segments and the apical membrane of the retinal pigment epithelium (RPE). In addition to these changes, oxygen consumption in the retina increases with a concomitant release of CO2 and H2O into the SRS. The RPE maintains SRS pH and volume homeostasis by transporting these metabolic byproducts to the choroidal blood supply. In vitro, we mimicked the transition from light to dark by increasing apical bath CO2 from 5 to 13%; this maneuver decreased cell pH from 7.37 ± 0.05 to 7.14 ± 0.06 (n = 13). Our analysis of native and cultured fetal human RPE shows that the apical membrane is significantly more permeable (≈10-fold; n = 7) to CO2 than the basolateral membrane, perhaps due to its larger exposed surface area. The limited CO2 diffusion at the basolateral membrane promotes carbonic anhydrase–mediated HCO3 transport by a basolateral membrane Na/nHCO3 cotransporter. The activity of this transporter was increased by elevating apical bath CO2 and was reduced by dorzolamide. Increasing apical bath CO2 also increased intracellular Na from 15.7 ± 3.3 to 24.0 ± 5.3 mM (n = 6; P < 0.05) by increasing apical membrane Na uptake. The CO2-induced acidification also inhibited the basolateral membrane Cl/HCO3 exchanger and increased net steady-state fluid absorption from 2.8 ± 1.6 to 6.7 ± 2.3 µl × cm−2 × hr−1 (n = 5; P < 0.05). The present experiments show how the RPE can accommodate the increased retinal production of CO2 and H2O in the dark, thus preventing acidosis in the SRS. This homeostatic process would preserve the close anatomical relationship between photoreceptor outer segments and RPE in the dark and light, thus protecting the health of the photoreceptors.  相似文献   

12.
Hypoxanthine (H), the deamination product of adenine, has been implicated in the high frequency of A to G transitions observed in retroviral and other RNA genomes. Although H·C base pairs are thermodynamically more stable than other H·N pairs, polymerase selection may be determined in part by kinetic factors. Therefore, the hypoxanthine induced substitution pattern resulting from replication by viral polymerases may be more complex than that predicted from thermodynamics. We have examined the steady-state kinetics of formation of base pairs opposite template H in RNA by HIV-RT, and for the incorporation of dITP during first- and second-strand synthesis. Hypoxanthine in an RNA template enhances the k2app for pairing with standard dNTPs by factors of 10–1000 relative to adenine at the same sequence position. The order of base pairing preferences for H in RNA was observed to be H·C >> H·T > H·A > H·G. Steady-state kinetics of insertion for all possible mispairs formed with dITP were examined on RNA and DNA templates of identical sequence. Insertion of dITP opposite all bases occurs 2–20 times more frequently on RNA templates. This bias for higher insertion frequencies on RNA relative to DNA templates is also observed for formation of mispairs at template A. This kinetic advantage afforded by RNA templates for mismatches and pairing involving H suggests a higher induction of mutations at adenines during first-strand synthesis by HIV-RT.  相似文献   

13.
Previous work has indicated that changes in gibberellin (GA) metabolism may be involved in chilling-induced release from dormancy in somatic embryos of grape (Vitis vinifera L. × V. rupestris Scheele). We have chilled somatic embryos of grape for 2, 4, or 8 weeks, then incubated them with [3H]GA4 (of high specific activity, 4.81 × 1010 becquerel per millimole) for 48 hours at 26°C. Chilling had little effect on the total amount of free [3H]GA-like metabolites formed during incubation at 26°C, but did change the relative proportions of individual metabolites. The amount of highly water-soluble [3H] metabolites formed at 26°C decreased in embryos chilled for 4 or 8 weeks. The concentration of endogenous GA precursors (e.g., GA12 aldehyde-, kaurene-, and kaurenoic acid-like substances) increased in embryos chilled for 4 or 8 weeks. Treatment with abscisic acid (ABA) (known to inhibit germination in grape embryos) concurrent with [3H]GA4 treatment at 26°C, reduced the uptake of [3H] GA4 but had little effect on the qualitative spectrum of metabolites. However, in the embryos chilled for 8 weeks and then treated with ABA for 48 hours at 26°C, there was a higher concentration of GA precursors than in untreated control embryos. Chilled embryos thus have an enhanced potential for an increase in free GAs through synthesis from increased amounts of GA precursors, or through a reduced ability to form highly water-soluble GA metabolites (i.e., GA conjugates or polyhydroxylated free GAs).  相似文献   

14.
The indole alkaloid gramine is toxic to animals and may play a defensive role in plants. Under certain conditions, shoots of barley cultivars such as `Arimar' and CI 12020 accumulate gramine (N,N-dimethyl-3-aminomethylindole) and lesser amounts of its precursors 3-aminomethylindole (AMI) and N-methyl-3-aminomethylindole (MAMI); other cultivars such as `Proctor' do not. When grown at optimal temperatures (21°C/16°C, day/night), Arimar contained a high level of gramine in the first leaf (approximately 6 milligrams per gram dry weight), but progressively less accumulated in successive leaves so that the gramine level in the shoot as a whole fell sharply with age. In Arimar and CI 12020 plants transferred at the two- to three-leaf stage from 21°C/16°C to supra-optimal temperatures (≥30°C/25°C), there was massive gramine accumulation in leaves which developed at high temperature, so that gramine level in the whole shoot remained high (about 3-8 milligrams per gram dry weight).

Proctor lacked both constitutive gramine accumulation in the first leaf and heat-induced gramine accumulation in later leaves. The following evidence indicates that this results from a lesion in the pathway of synthesis (tryptophan →→ AMI → MAMI → gramine) between tryptophan and AMI. (a) Proctor and Arimar leaves readily absorbed [14C]gramine, but neither cultivar degraded it extensively. (b) Arimar leaf tissue incorporated [14C]formate label into the N-methyl groups of gramine and MAMI, and converted [methylene-14C]tryptophan to AMI, MAMI, and gramine; Proctor leaf tissue did not, even when a trapping pool of unlabeled gramine was supplied. (c) Proctor converted [14C]MAMI to gramine as actively as Arimar. (d) Proctor incorporated [14C]formate label into gramine and MAMI when supplied with AMI; the ratio [14C]gramine/[14C]MAMI fell with leaf age, suggesting that the two N-methylations involve different enzymes. Inasmuch as Proctor leaf tissue did not methylate added tryptamine or tyramine, the N-methyltransferase(s) of gramine synthesis may be substrate specific.

In sterile culture at optimal temperatures, 10 millimolar gramine did not affect autotrophic growth of Arimar or Proctor plantlets or heterotrophic growth of callus. At supra-optimal temperature, plantlet growth was reduced by gramine although callus growth was not. We speculate that gramine-accumulating cultivars may suffer autotoxic effects at high leaf temperatures.

  相似文献   

15.
1. 26-Hydroxycholesterol was obtained by reducing the methyl ester of (±)-3β-hydroxycholest-5-en-26-oic acid, which was synthesized from 25-oxonorcholesterol. 2. Methods for preparing 7α-hydroxycholesterol and 7-dehydrocholesterol were modified to allow the micro-scale preparation of these [14C]sterols from [26-14C]-cholesterol. 3. 26-Hydroxycholesterol was oxidized more readily than 7α-hydroxycholesterol, 7-dehydrocholesterol or cholesterol by mitochondrial preparations from livers of mice, rats, guinea pigs, common toads (Bufo vulgaris) and Caiman crocodylus. 4. (±)-3β-Hydroxy[26-14C]cholest-5-en-26-oic acid was oxidized very rapidly to 14CO2 by mouse and guinea-pig mitochondria without evident discrimination between the two optical isomers. 5. An enzyme system that oxidizes 26-hydroxycholesterol to 3β-hydroxycholest-5-en-26-oic acid was identified in the soluble extract of rat-liver mitochondria. This enzyme could use NADP in place of NAD but was not identical with liver alcohol dehydrogenase (EC 1.1.1.1). 6. [26-14C]Cholesteryl 3β-sulphate was not oxidized by fortified mouse-liver preparations that oxidized [26-14C]cholesterol to 14CO2.  相似文献   

16.

Objective

Inflammation is an important contributor to atherosclerosis progression. A glucose analogue 18F-fluorodeoxyglucose ([18F]FDG) has been used to detect atherosclerotic inflammation. However, it is not known to what extent [18F]FDG is taken up in different stages of atherosclerosis. We aimed to study the uptake of [18F]FDG to various stages of coronary plaques in a pig model.

Methods

First, diabetes was caused by streptozotocin injections (50 mg/kg for 3 days) in farm pigs (n = 10). After 6 months on high-fat diet, pigs underwent dual-gated cardiac PET/CT to measure [18F]FDG uptake in coronary arteries. Coronary segments (n = 33) were harvested for ex vivo measurement of radioactivity and autoradiography (ARG).

Results

Intimal thickening was observed in 16 segments and atheroma type plaques in 10 segments. Compared with the normal vessel wall, ARG showed 1.7±0.7 times higher [18F]FDG accumulation in the intimal thickening and 4.1±2.3 times higher in the atheromas (P = 0.004 and P = 0.003, respectively). Ex vivo mean vessel-to-blood ratio was higher in segments with atheroma than those without atherosclerosis (2.6±1.2 vs. 1.3±0.7, P = 0.04). In vivo PET imaging showed the highest target-to-background ratio (TBR) of 2.7. However, maximum TBR was not significantly different in segments without atherosclerosis (1.1±0.5) and either intimal thickening (1.2±0.4, P = 1.0) or atheroma (1.6±0.6, P = 0.4).

Conclusions

We found increased uptake of [18F]FDG in coronary atherosclerotic lesions in a pig model. However, uptake in these early stage lesions was not detectable with in vivo PET imaging. Further studies are needed to clarify whether visible [18F]FDG uptake in coronary arteries represents more advanced, highly inflamed plaques.  相似文献   

17.
The serum decay of rat serum high-density lipoprotein (HD lipoprotein), labelled biosynthetically with 32P in the phospholipid or with 3H in the cholesteryl ester moiety, was measured in rats after partial hepatectomy or sham operation. The serum decay of 3H-labelled HD lipoprotein cholesteryl esters was biexponential. In sham-operated rats the t½ values for the rapid phase and the slow phase were 0.2±0.1h and 4.2±0.4h (means±s.e.m.) respectively. After removal of two-thirds of the liver the t½ value of the rapid phase did not change (0.1±0.1h), whereas the t½ value of the slow phase increased to 5.7±0.8h. Partial hepatectomy hardly changed extrahepatic tissue radioactivities, whereas the percentage of the injected dose recovered in the liver 6h after injection decreased from 34.0±1.9% before to 13.5±1.6% after partial hepatectomy. The 32P-labelled HD lipoprotein phospholipids showed a rapid monoexponential decay from serum with t½ values of 0.71±0.3h and 1.48±0.11h after sham operation or partial hepatectomy respectively. The tissue 32P radioactivities in the shamoperated rats, measured 1h after injection, were 46.0±1.7% (liver), 1.7±0.3% (adipose tissue), 3.7±1.2% (skeletal muscle) and 3.0±0.0% (erythrocytes) of the injected dose. Only the value for liver was affected by partial hepatectomy and decreased to 16.7±3.8%. In a previous publication [Van Tol, Van Gent, Van't Hooft & Vlaspolder (1978) Atherosclerosis 29, 439–448] we showed in a highly comparable experimental setting that the turnover rates of HD apolipoproteins A and C in vivo are not influenced by removal of two-thirds of the liver. From the present study it is clear that the removal rates of radioactive HD lipoprotein cholesteryl esters and HD lipoprotein phospholipids from serum in vivo are decreased by partial hepatectomy. The results indicate the possibility of partly separate metabolic pathways of HD apolipoproteins A and C, HD lipoprotein cholesteryl esters and HD lipoprotein phospholipids. The phospholipids and cholesteryl esters of HD lipoprotein are metabolized predominantly by the liver. Possible mechanisms for the hepatic uptake and metabolism of HD lipoprotein cholesteryl (esters) and phospholipids are discussed.  相似文献   

18.
Excised light-grown leaves and etiolated leaves of Hordeum vulgare L. cv Dyan catabolized applied (±)-[2-14C]abscisic acid ([±]-[2-14C]ABA) to phaseic acid (PA), dihydrophaseic acid (DPA), and 2′-hydroxymethyl ABA (2′-HMABA). Identification of these catabolites was made by microchemical methods and by combined capillary gas chromatographymass spectrometry (GC-MS) following high dose feeds of nonlabeled substrate to leaves. Circular dichroism analysis revealed that 2′-HMABA was derived from the (−) enantiomer of ABA. By selecting tissue samples in which endogenous catabolites were undetectable by gas chromatography, it was possible to identify unequivocally ABA catabolites by GC-MS without the need to employ deuteriated substrate to distinguish the (±)-ABA catabolites from the same endogenous compounds. Refeeding studies were used to confirm the catabolic route. The methyl ester of (±)-[214C]-ABA was hydrolyzed efficiently by light-grown leaves of H. vulgare. Leaf age played a significant role in (±)-ABA catabolism, with younger leaves being less able than their older counterparts to catabolize this compound. The catabolism of (±)-ABA was inhibited markedly in water-stressed Hordeum leaves which was characterized by a decreased incorporation of label into 2′-HMABA, DPA, and conjugates. The specific, mixed function oxidase inhibitor, ancymidol, did not inhibit, dramatically, (±)-ABA catabolism in light-grown leaves of Hordeum whereas the 80s ribosome, translational inhibitor, cycloheximide, inhibited this process markedly. The 70s ribosome translational inhibitors, lincomycin and chloramphenicol, were less effective than cycloheximide in inhibiting (±)-ABA catabolism, implying that cytoplasmic protein synthesis is necessary for the catabolism of (±)-ABA in Hordeum leaves whereas chloroplast protein synthesis plays only a minor role. This further suggests that the enzymes involved in (±)-ABA catabolism in this plant are cytoplasmically synthesized and are `turned-over' rapidly, although the enzyme responsible for glycosylating (±)-ABA itself appeared to be stable.  相似文献   

19.
X-prolyl-dipeptidyl aminopeptidase, which hydrolyzed Gly-Pro-p-nitroanilide (relative activity [RA] = 100%) and Arg-Pro-p-nitroanilide (RA, 130%), was purified to homogeneity from the cell extract of Lactobacillus helveticus CNRZ 32. The enzyme also hydrolyzed Ala-Pro-Gly (RA, 11%) and Ala-Ala-p-nitroanilide (RA, 2%) but was not active on Ala-Leu-Ala, dipeptides, and endopeptidase and carboxypeptidase substrates. The enzyme was purified 145-fold by streptomycin sulfate precipitation, ammonium sulfate fractionation, and a series of column chromatographies on DEAE-cellulose, arginine-Sepharose 4B, and glycyl-prolyl-AH-Sepharose 4B. The purified enzyme appeared as a single band on native polyacrylamide gel and sodium dodecyl sulfate-polyacrylamide gel electrophoreses and had a molecular weight of 72,000. Optima for activity by the purified enzyme were pH 7.0 and 40°C. The enzyme was incubated at 40°C for 15 min with various metal ions. It was activated by Mg2+ (2.5 mM), Ca2+ (0.1 to 2.5 mM), Na+ (10 to 50 mM), and K+ (10 to 50 mM) and was inhibited by Hg2+ (0.1 to 2.5 mM), Cu2+ (0.1 to 2.5 mM), and Zn2+ (0.1 to 2.5 mM). Enzyme activity was partially inhibited by EDTA (1.0 mM, 20 h at 40°C), 1,10-phenanthroline (1.0 mM, 15 min at 40°C), phenylmethylsulfonyl fluoride (1.0 mM), N-ethylmaleimide (1.0 mM), and iodoacetate (1.0 mM). It was completely inhibited by diisopropyl fluorophosphate (1.0 mM, 2 h at 40°C) and p-chloromercuribenzoate (1.0 mM, 15 min at 40°C). The enzyme was not affected by dithioerythritol (1.0 to 10 mM).  相似文献   

20.
Bromus inermis Leyss cell cultures treated with 75 micromolar abscisic acid (ABA) at both 23 and 3°C developed more freezing resistance than cells cultured at 3°C. Protein synthesis in cells induced to become freezing tolerant by ABA and low temperature was monitored by [14C]leucine incorporation. Protein synthesis continued at 3°C, but net cell growth was stopped. Most of the major proteins detected at 23°C were synthesized at 3°C. However, some proteins were synthesized only at low temperatures, whereas others were inhibited. ABA showed similar effects on protein synthesis at both 23 and 3°C. Comparative electrophoretic analysis of [14C]leucine labeled protein detected the synthesis of 19, 21 and 47 kilodalton proteins in less than 8 hours after exposure to exogenous ABA. Proteins in the 20 kilodalton range were also synthesized at 3°C. In addition, a 31 kilodalton protein band showed increased expression in freezing resistant ABA treated cultures after 36 hours growth at both 3 and 23°C. Quantitative analysis of [14C]leucine labeled polypeptides in two-dimensional gels confirmed the increased expression of the 31 kilodalton protein. Two-dimensional analysis also resolved a 72 kilodalton protein enriched in ABA treated cultures and identified three proteins (24.5, 47, and 48 kilodaltons) induced by low temperature growth.  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