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 46 毫秒
1.
L-lactate transport in Ehrlich ascites-tumour cells.   总被引:10,自引:0,他引:10       下载免费PDF全文
Ehrlich ascites-tumour cells were investigated with regard to their stability to transport L-lactate by measuring either the distribution of [14C]lactate or concomitant H+ ion movements. The movement of lactate was dependent on the pH difference across the cell membrane and was electroneutral, as evidenced by an observed 1:1 antiport for OH- ions or 1:1 symport with H+ ions. 2. Kinetic experiments showed that lactate transport was saturable, with an apparent Km of approx. 4.68 mM and a Vmax. as high as 680 nmol/min per mg of protein at pH 6.2 and 37 degrees C. 3. Lactate transport exhibited a high temperature dependence (activation energy = 139 kJ/mol). 4. Lactate transport was inhibited competitively by (a) a variety of other substituted monocarboxylic acids (e.g. pyruvate, Ki = 6.3 mM), which were themselves transported, (b) the non-transportable analogues alpha-cyano-4-hydroxycinnamate (Ki = 0.5 mM), alpha-cyano-3-hydroxycinnamate (Ki = 2mM) and DL-p-hydroxyphenyl-lactate (Ki = 3.6 mM) and (c) the thiol-group reagent mersalyl (Ki = 125 muM). 5. Transport of simple monocarboxylic acids, including acetate and propionate, was insensitive to these inhibitors; they presumably cross the membrane by means of a different mechanism. 6. Experiments using saturating amounts of mersalyl as an "inhibitor stop" allowed measurements of the initial rates of net influx and of net efflux of [14C]lactate. Influx and efflux of lactate were judged to be symmetrical reactions in that they exhibited similar concentration dependence. 7. It is concluded that lactate transport in Ehrlich ascites-tumour cells is mediated by a carrier capable of transporting a number of other substituted monocarboxylic acids, but not unsubstituted short-chain aliphatic acids.  相似文献   

2.
Effect of glucagon on energy-metabolite transport into cardiac muscle was studied during a single transit through the isolated rabbit heart using a rapid paired-tracer dilution method. Kinetic experiments revealed that 1.5 microM glucagon stimulated the influx of palmitate bound to 30 g/litre albumin, by increasing the V 2.3 times and increasing the Km for transport 2.4 times. Tracer uptake of D-glucose, as the only exogenous substrate provided, was increased by 80% by 1.5 microM glucagon. Myocardial utilization of [3H]-or [14C]-labelled short-chain monocarboxylic acids (L-lactate, pyruvate and acetate) was significantly reduced by glucagon, to the same degree as their unidirectional sarcolemmal transport. Inhibition of L-[14C]lactate uptake was dose-dependent and in positive correlation with myocardial lactate production. It is concluded that glucagon may regulate sarcolemmal permeability and myocardial utilization for energy-metabolites from the coronary circulation.  相似文献   

3.
Stimulation of pyruvate dehydrogenase (PDH) improves functional recovery of postischemic hearts. This study examined the potential for a mechanism mediated by substrate-dependent proton production and intracellular pH. After 20 min of ischemia, isolated rabbit hearts were reperfused with or without 5 mM dichloroacetate (DCA) in the presence of either 5 mM glucose, 5 mM glucose + 2.5 mM lactate, or 5 mM glucose + 2.5 mM pyruvate. DCA inhibits PDH kinase, increasing the proportion of dephosphorylated, active PDH. Unlike pyruvate or glucose alone, lactate + glucose did not support the effects of DCA on the recovery of rate-pressure product (RPP) (without DCA, RPP = 14,000 +/- 1,200, n = 6; with DCA, RPP = 13,700 +/- 1,800, n = 9). Intracellular pH, from (31)P nuclear magnetic resonance spectra, returned to normal within 2.1 min of reperfusion with all substrates except for lactate + glucose + DCA or lactate + DCA, which delayed pH recovery for up to 12 min (at 2.1 min pH = 6. 00 +/- 0.08, lactate + glucose + DCA; pH = 6.27 +/- 0.34, for lactate + DCA). Hearts were also reperfused after 10 min of ischemia with 0.5 mM palmitate + 5 mM DCA and either 2.5 mM pyruvate or 2.5 mM lactate. Again, intracellular pH recovery was delayed in the presence of lactate. PDH activation in the presence of lactate also decreased coupling of oxidative metabolism to mechanical work. These findings have implications for therapeutic use of stimulated carbohydrate oxidation in stunned hearts.  相似文献   

4.
The characteristics of pyruvate transport across the plasma membrane in the bloodstream form of Trypanosoma brucei were studied using [14C]pyruvate in combination with the silicone-oil centrifugation technique. We present evidence for the existence of a facilitated diffusion carrier in the plasma membrane of T. brucei which specifically mediates the translocation of pyruvate. The uptake of pyruvate followed saturation kinetics (Km 1.96 +/- 0.28 mM; Cmax 36.61 +/- 1.15 nmol pyruvate/30 sec.mg protein), after correction of the data for a nonsaturable diffusion component. The uptake of pyruvate was competitively inhibited by a number of (oxo)monocarboxylic acids, including pyruvate analogs and metabolically related substances, but not by L-lactate. The transport exhibited the phenomenon of transacceleration, indicative for the involvement of a facilitated diffusion carrier. The carrier is highly specific for pyruvate and differs from other known monocarboxylate carriers present in the mitochondrial and/or plasma membrane of other eukaryotic cells in that it does not transport L-lactate.  相似文献   

5.
Initial velocity, product inhibition, and substrate inhibition studies suggest that the endogenous lactate dehydrogenase activity of duck epsilon-crystallin follows an order Bi-Bi sequential mechanism. In the forward reaction (pyruvate reduction), substrate inhibition by pyruvate was uncompetitive with inhibition constant of 6.7 +/- 1.7 mM. In the reverse reaction (lactate oxidation), substrate inhibition by L-lactate was uncompetitive with inhibition constant of 158 +/- 25 mM. The cause of these inhibitions may be due to epsilon-crystallin-NAD(+)-pyruvate and epsilon-crystallin-NADH-L-lactate abortive ternary complex formation as suggested by the multiple inhibition studies. Pyruvate binds to free enzyme very poorly, with a very large dissociation constant. Bromopyruvate, fluoropyruvate, pyruvate methyl ester, and pyruvate ethyl ester are alternative substrates for pyruvate. 3-Acetylpyridine adenine dinucleotide, nicotinamide 1,N6-ethenoadenine dinucleotide, and nicotinamide hypoxanthine dinucleotide serve as alternative coenzymes for epsilon-crystallin. All the above alternative substrates or coenzymes showed an intersecting initial-velocity pattern conforming to the order Bi--Bi kinetic mechanism. Nicotinic acid adenine dinucleotide, thionicotinamide adenine dinucleotide, and 3-aminopyridine adenine dinucleotide acted as inhibitors for this enzymatic crystallin. The inhibitors were competitive versus NAD+ and noncompetitive versus L-lactate. alpha-NAD+ was a noncompetitive inhibitor with respect to the usual beta-NAD+. D-Lactate, tartronate, and oxamate were strong dead-end inhibitors for the lactate dehydrogenase activity of epsilon-crystallin. Both D-lactate and tartronate were competitive inhibitors versus L-lactate while oxamate was a competitive inhibitor versus pyruvate. We conclude that the structural requirements for the substrate and coenzyme of epsilon-crystallin are similar to those of other dehydrogenases and that the carboxamide carbonyl group of the nicotinamide moiety is important for the coenzyme activity.  相似文献   

6.
1. Studies on the kinetics of pyruvate transport into mitochondria by an 'inhibitor-stop' technique were hampered by the decarboxylation of pyruvate by mitochondria even in the presence of rotenone. Decarboxylation was minimal at 6 degrees C. At this temperature the Km for pyruvate was 0.15 mM and Vmax. was 0.54nmol/min per mg of protein; alpha-cyano-4-hydroxycinnamate was found to be a non-competitive inhibitor, Ki 6.3 muM, and phenyl-pyruvate a competitive inhibitor, Ki 1.8 mM. 2. At 100 muM concentration, alpha-cyano-4-hydroxycinnamate rapidly and almost totally inhibited O2 uptake by rat heart mitochondria oxidizing pyruvate. Inhibition could be detected at concentrations of inhibitor as low as 1 muM although inhibition took time to develop at this concentration. Inhibition could be reversed by diluting out the inhibitor. 3. Various analogues of alpha-cyano-4-hydroxycinnamate were tested on rat liver and heart mitochondria. The important structural features appeared to be the alpha-cyanopropenoate group and the hydrophobic aromatic side chain. Alpha-Cyanocinnamate, alpha-cyano-5-phenyl-2,4-pentadienoate and compound UK 5099 [alpha-cyano-beta-(2-phenylindol-3-yl)acrylate] were all more powerful inhibitors than alpha-cyano-4-hydroxycinnamate showing 50% inhibition of pyruvate-dependent O2 consumption by rat heart mitochondria at concentrations of 200, 200 and 50 nM respectively. 4. The specificity of the carrier for its substrate was studied by both influx and efflux experiments. Oxamate, 2-oxobutyrate, phenylpyruvate, 2-oxo-4-methyl-pentanoate, chloroacetate, dichloroacetate, difluoroacetate, 2-chloropropionate, 3-chloropropionate and 2,2-dichloropropionate all exchanged with pyruvate, whereas acetate, lactate and trichloroacetate did not. 5. Pyruvate entry into the mitochondria was shown to be accompanied by the transport of a proton (or by exchange with an OH-ion). This proton flux was inhibited by alpha-cyano-4-hydroxycinnamate and allowed measurements of pyruvate transport at higher temperatures to be made. The activation energy of mitochondrial pyruvate transport was found to be 113 kJ (27 kcal)/mol and by extrapolation the rate of transport of pyruvate at 37 degrees C to be 42 nmol/min per mg of protein. The possibility that pyruvate transport into mitochondria may be rate limiting and involved in the regulation of gluconegenesis is discussed. 6. The transport of various monocarboxylic acids into mitochondria was studied by monitoring proton influx. The transport of dichloroacetate, difluoroacetate and oxamate appeared to be largely dependent on the pyruvate carrier and could be inhibited by pyruvate-transport inhibitors. However, many other halogenated and 2-oxo acids which could exchange with pyruvate on the carrier entered freely even in the presence of inhibitor.  相似文献   

7.
Time courses of L-lactate and pyruvate uptake into isolated rat hepatocytes were measured in a citrate-based medium to generate a pH gradient (alkaline inside), by using the silicone-oil-filtration technique at 0 degrees C to minimize metabolism. At low concentrations of lactate and pyruvate (0.5 mM), transport was inhibited by over 95% by 5 mM-alpha-cyano-4-hydroxycinnamate, whereas at higher concentrations (greater than 10 mM) a significant proportion of transport could not be inhibited. The rate of this non-inhibitable transport was linearly related to the substrate concentration, was less with pyruvate than with L-lactate, and appeared to be due to diffusion of undissociated acid. Uptake of D-lactate was not inhibited by alpha-cyano-4-hydroxycinnamate and occurred only by diffusion. Kinetic parameters for the carrier-mediated transport process were obtained after correction of the initial rates of uptake of lactate and pyruvate in the absence of 5 mM-alpha-cyano-4-hydroxycinnamate by that in the presence of inhibitor. Under the conditions used, the Km values for L-lactate and pyruvate were 2.4 and 0.6 mM respectively and the Ki for alpha-cyano-4-hydroxycinnamate as a competitive inhibitor was 0.11 mM. Km values for the transport of L-lactate and pyruvate into rat erythrocytes under similar conditions were 3.0 and 0.96 mM. The Vmax. of lactate and pyruvate transport into hepatocytes at 0 degrees C was 3 nmol/min per mg of protein. Carrier-mediated transport of 0.5 mM-L-lactate was inhibited by 0.2 mM-p-chloromercuribenzenesulphonate (greater than 90%), 0.5 mM-quercetin (80%), 0.6 mM-isobutylcarbonyl-lactyl anhydride (70%) and 0.5 mM-4,4'-di-isothiocyanostilbene-2,2'-disulphonate (50%). A similar pattern of inhibition of lactate transport is seen in erythrocytes. It is suggested that the same or a similar carrier protein exists in both tissues. The results also show that L-lactate transport into rat hepatocytes is very rapid at physiological temperatures and is unlikely to restrict the rate of its metabolism. Differences between our results and those of Fafournoux, Demigne & Remesy [(1985) J. Biol. Chem. 260, 292-299] are discussed.  相似文献   

8.
The substrate specificity of a Na+-dependent transport pathway for L-lactate was studied in rabbit renal brush border membrane vesicles. Jmax for L-lactate transport was unaffected by the presence of a fixed concentration of two different short-chain monocarboxylic acids, while the apparent Kt(Ka) for L-lactate increased, and this is compatible with competitive inhibition. The inhibitor constants ("Ki"'s) for the transport pathway for the two solutes examined closely corresponded to the respective "Ki"'s derived from a Dixon plot. A broad range of compounds were then tested as potential inhibitors of L-lactate transport, and the "Ki"'s thereby derived yielded specific information regarding optimal substrate recognition by the carrier. A single carboxyl group is an absolute requirement for recognition, and preference is given to 3 to 6 C chain molecules. Addition of ketone, hydroxyl and, particularly, amine groups at any carbon position, diminishes substrate-carrier interaction. Intramolecular forces, notably the inductive effects of halogens, may play a role in enhancing substrate-carrier interaction; however, no correlation was found between pKa and "Ki" for the substrates examined. We conclude that a separate monocarboxylic acid transport pathway, discrete from either the D-glucose, alpha or beta neutral amino-acid, or dicarboxylic acid carriers, exists in the renal brush border, and this handles a broad range of monocarboxylates.  相似文献   

9.
L-Serine alone is not gluconeogenic in isolated rabbit hepatocytes, whereas in rat liver this amino acid has been reported to yield as much glucose as does L-lactate itself. The current study has been an investigation into the explanation of the difference between the two species. Hepatocytes were isolated from 48-h-starved, 750- to 1000-g male rabbits, and the viability of each preparation was judged by ATP levels (2.4 +/- 0.2 mumol/g wet wt) at the beginning and end of the incubation as well as gluconeogenesis from 10 mM L-lactate (0.83 +/- 0.08 mumol/min/g wet wt). L-Serine alone produced virtually no glucose or pyruvate accumulation above baseline. Hydroxypyruvate, however, did appear in the incubation mixture. When L-serine and pyruvate were combined to test the functional activity of L-serine:pyruvate aminotransferase (EC 2.6.1.51), however, gluconeogenesis remained at the rate produced by pyruvate alone (0.61 +/- 0.04 mumol/min/g wet wt). On the other hand, the combination of L-serine and L-lactate produced rates of glucose accumulation 35% above that of L-lactate alone. The combination of L-lactate plus hydroxypyruvate produced nearly maximal rates (1.39 +/- 0.08 mumol/min/g wet wt), approaching those achieved by a physiologic ratio (10:1) of L-lactate and pyruvate. Hydroxypyruvate itself was only moderately gluconeogenic (0.44 +/- 0.04 mumol/min/g wet wt). That a reduction of the cytoplasmic free [NAD+]/[NADH] ratio by L-lactate was not its only contribution to L-serine utilization was suggested by the fact that ethanol completely eliminated gluconeogenesis from virtually all precursors (or combinations) tested, with the exception of hydroxypyruvate. It has been concluded from the data that, probably in contrast to the rat, the major pathway for the entrance of L-serine into gluconeogenesis in rabbit hepatocytes is through the pathway initiated by L-serine: pyruvate aminotransferase and that L-lactate is an important participant (i) by generating cytoplasmic reducing equivalents (NADH), (ii) by supplying pyruvate for the transaminating reaction itself, and, perhaps, (iii) by preventing hydroxypyruvate from being reduced by L-lactate dehydrogenase (EC 1.1.1.27) to L-glycerate.  相似文献   

10.
Tauropine dehydrogenase (tauropine:NAD oxidoreductase) was purified from the shell adductor muscle of the ormer, Haliotis lamellosa. The enzyme was found to utilize stoichiometrically NADH as co-enzyme and pyruvate and taurine as substrates producing tauropine [rhodoic acid; N-(D-1-carboxyethyl)-taurine]. The enzyme was purified to a specific activity of 463 units/mg protein using a combination of ammonium sulphate fractionation, ion-exchange and affinity chromatography. The relative molecular mass was 38,000 +/- 1000 when assessed by gel filtration on Ultrogel AcA 54 and 42,000 +/- 150 by electrophoresis on 5-10% polyacrylamide gels in the presence of 1% sodium dodecyl sulphate; the data suggest a monomeric structure. Tauropine and pyruvate were found to be the preferred substrates. Among the amino acids tested for activity with the enzyme, only alanine is used as an alternative substrate, but with a rate less than 6% of the enzyme activity with taurine. Of the oxo acids tested, 2-oxobutyrate and 2-oxovalerate were also found to be substrates. Apparent Km values for the substrates NADH, pyruvate and taurine are 0.022 +/- 0.003 mM, 0.64 +/- 0.07 mM and 64.7 +/- 5.4 mM, respectively, at pH 7.0 and for the products, NAD+ and tauropine, are 0.29 +/- 0.01 mM and 9.04 +/- 1.27 mM, respectively, at pH 8.3. Apparent Km values for both pyruvate and taurine decrease with increasing co-substrate (taurine or pyruvate) concentration. NAD+ and tauropine were found to be product inhibitors of the forward reaction. NAD+ was a competitive inhibitor of NADH, whereas tauropine gave a mixed type of inhibition with respect to pyruvate and taurine. Succinate was found to inhibit non-competitively with respect to taurine and pyruvate with an apparent Ki value in the physiological range of this anaerobic end product. The inhibition by L-lactate, not an end product in the ormer, was competitive with respect to pyruvate. The physiological role or tauropine dehydrogenase during anaerobiosis is discussed.  相似文献   

11.
Inhibition of gluconeogenesis and glycogenolysis by 2,5-anhydro-D-mannitol   总被引:1,自引:0,他引:1  
2,5-Anhydro-D-mannitol (100 to 200 mg/kg) decreased blood glucose by 17 to 58% in fasting mice, rats, streptozotocin-diabetic mice, and genetically diabetic db/db mice. Serum lactate in rats was elevated 56% by 2,5-anhydro-D-mannitol, but this could be prevented by dichloroacetate (200 mg/kg) or thiamin (200 mg/kg). In hepatocytes from fasted rats, 1 mM 2,5-anhydro-D-mannitol inhibited gluconeogenesis from a mixture of alanine, lactate, and pyruvate. It also inhibited glucose production and stimulated lactate formation from glycerol or dihydroxyacetone. Glycogenolysis in hepatocytes from fed rats was markedly inhibited by 1 mM 2,5-anhydro-D-mannitol both in the presence or absence of 1 microM glucagon. 2,5-Anhydro-D-mannitol can be phosphorylated by fructokinase or hexokinase to the 1-phosphate and then by phosphofructokinase to the 1,6-bisphosphate. Rat liver glycogen phosphorylase was inhibited by 2,5-anhydro-D-mannitol 1-phosphate (apparent Ki = 0.66 +/- 0.09 mM) but was little affected by 2,5-anhydro-D-mannitol 1,6-bisphosphate. Rat liver phosphoglucomutase was inhibited by 2,5-anhydro-D-mannitol 1-phosphate (apparent Ki = 2.8 +/- 0.2 mM), whereas 2,5-anhydro-D-mannitol 1,6-bisphosphate served as an alternative activator (apparent K alpha = 7.0 +/- 0.5 microM). Rabbit liver pyruvate kinase was activated by 2,5-anhydro-D-mannitol 1,6-bisphosphate (apparent K alpha = 9.5 +/- 0.9 microM), whereas rabbit liver fructose 1,6-bisphosphatase was inhibited by 2,5-anhydro-D-mannitol 1,6-bisphosphate (apparent Ki = 3.6 +/- 0.3 microM). The phosphate esters of 2,5-anhydro-D-mannitol would, therefore, be expected to inhibit glycogenolysis and gluconeogenesis and stimulate glycolysis in liver.  相似文献   

12.
Myocardial ischemia-reperfusion is associated with bursts of reactive oxygen species (ROS) such as superoxide radicals (O(2)(-).). Membrane-associated NADH oxidase (NADHox) activity is a hypothetical source of O(2)(-)., implying the NADH concentration-to-NAD(+) concentration ratio ([NADH]/[NAD(+)]) as a determinant of ROS. To test this hypothesis, cardiac NADHox and ROS formation were measured as influenced by pyruvate or L-lactate. Pre- and postischemic Langendorff guinea pig hearts were perfused at different pyruvate/L-lactate concentrations to alter cytosolic [NADH]/[NAD(+)]. NADHox and ROS were measured with the use of lucigenin chemiluminescence and electron spin resonance, respectively. In myocardial homogenates, pyruvate (0.05, 0.5 mM) and the NADHox blocker hydralazine markedly inhibited NADHox (16 +/- 2%, 58 +/- 9%). In postischemic hearts, pyruvate (0.1-5.0 mM) dose dependently inhibited ROS up to 80%. However, L-lactate (1.0-15.0 mM) stimulated both basal and postischemic ROS severalfold. Furthermore, L-lactate-induced basal ROS was dose dependently inhibited by pyruvate (0.1-5.0 mM) and not the xanthine oxidase inhibitor oxypurinol. Pyruvate did not inhibit ROS from xanthine oxidase. The data suggest a substantial influence of cytosolic NADH on cardiac O(2)(-). formation that can be inhibited by submillimolar pyruvate. Thus cytotoxicities due to cardiac ischemia-reperfusion ROS may be alleviated by redox reactants such as pyruvate.  相似文献   

13.
Short- and medium-chain monocarboxylic acids showed an inhibitory effect on enzymatic activity of beta-hexosaminidase B (Hex B) when 4-methylumbelliferyl-2-acetamido-2-deoxyglucopyranoside (MU-GlcNAc) was used as substrate: 1. Two groups were distinguished according to the chain length of the monocarboxylic acids: the first was only constituted by acetic acid (C2) whereas the second group exhibited a broader chain length specificity for medium-chain monocarboxylic acids (between C6 and C9). 2. Both groups were reversible competitive inhibitors (Km = 0.52 +/- 0.15 mM; KiC2 = 21.4 +/- 3.0 mM; KiC7 = 3.4 +/- 0.5 mM). Competition experiments between C2 and C7 (as representent of medium-chain monocarboxylic acids group) demonstrated that these inhibitors were bound to different subsites. 3. Competition experiments between C2 and 2-acetamido-2-deoxy-D-galactonolactone (GalNAcLone) (a competitive inhibitor of lysosomal hexosaminidases) demonstrated that these two inhibitors were mutually exclusive, i.e. they were probably bound at the same subsite. This feature and the structural analogy of C2 with the acetyl residue of GalNAcLone (and of the saccharidic part of the substrate) suggested that C2 bound to the substrate site where the N-acetyl residue of the beta-N-acetyl hexosaminide was positioned. 4. The inhibitory effect of medium-chain monocarboxylic acids (C6 and C9) was dependent on their physical state. Below the critical micellar concentration (CMC), detected by a dye spectral shift method, no significant inhibition was detected, but as extensively reported using C7, an obvious inhibitory effect occurred at concentrations higher than CMC.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

14.
The properties of pyruvate kinase (PK) and phosphoenol pyruvate carboxykinase (PEP CK), two enzymes that determine the preferrential accumulation of either succinate or lactate as endproducts of carbohydrate metabolism, are described in adult Hymenolepis diminuta. PK activity at Vmax and Km levels of PEP was unaffected by ATP, alanine, FDP4, OR H+ ions, but was inhibited by 50% at 6.3 mM L-lactate and 30 mM HCO3. The addition of 30 mM HCO3 increased the Km(PEP) by 6-fold but did not alter the Vmax. The inhibition of PK by HCO3 cannot be explained entirely by an effect of ionic strength, but probably represents a specific modulator-enzyme interaction. Under similar conditions PEP CK was maximally activated. Although L-lactate inhibited PEP CK (Ki(lac) = 1.8 mM), this effector may play a minor role in regulation of PEP flux. These results implicate the poise of the HCO3-:CO2 system as a major determiner of endproduct accumulation in H. diminuta.  相似文献   

15.
We set out to study the pentose phosphate pathway (PPP) in isolated rat hearts perfused with [5-3H]glucose and [1-14C]glucose or [6-14C]glucose (crossover study with 1- then 6- or 6- then 1-14C-labeled glucose). To model a physiological state, hearts were perfused under working conditions with Krebs-Henseleit buffer containing 5 mM glucose, 40 microU/ml insulin, 0.5 mM lactate, 0.05 mM pyruvate, and 0.4 mM oleate/3% albumin. The steady-state C1/C6 ratio (i.e., the ratio from [1-14C]glucose to [6-14C]glucose) of metabolites released by the heart, an index of oxidative PPP, was not different from 1 (1.06 +/- 0.19 for 14CO2, and 1.00 +/- 0.01 for [14C]lactate + [14C]pyruvate, mean +/- SE, n = 8). Hearts exhibited contractile, metabolic, and 14C-isotopic steady state for glucose oxidation (14CO2 production). Net glycolytic flux (net release of lactate + pyruvate) and efflux of [14C]lactate + [14C]pyruvate were the same and also exhibited steady state. In contrast, flux based on 3H2O production from [5-3H]glucose increased progressively, reaching 260% of the other measures of glycolysis after 30 min. The 3H/14C ratio of glycogen (relative to extracellular glucose) and sugar phosphates (representing the glycogen precursor pool of hexose phosphates) was not different from each other and was <1 (0.36 +/- 0.01 and 0.43 +/- 0.05 respectively, n = 8, P < 0.05 vs. 1). We conclude that both transaldolase and the L-type PPP permit hexose detritiation in the absence of net glycolytic flux by allowing interconversion of glycolytic hexose and triose phosphates. Thus apparent glycolytic flux obtained by 3H2O production from [5-3H]glucose overestimates the true glycolytic flux in rat heart.  相似文献   

16.
The kinetics of pyruvate transport across the isolated red blood cell membrane were studied by a simple and precise spectrophotometric method: following the oxidation of NADH via lactate dehydrogenase trapped within resealed ghosts. The initial rate of pyruvate entry was linear. Influx was limited by saturation at high pyruvate concentration. Pyruvate influx was greatly stimulated by increasing ionic strength in the outer but not the inner aqueous compartment. The Km ranged from 15.0 mM at mu = 0.05 to 3.7 mM at mu = 0.01, while the V went from 0.611 - 10(15) to 0.137 - 10(-15) mol - min-1 - ghost-1. Ionic strength was shown to affect the translocation step and not pyruvate binding. The energy of activation of pyruvate flux into resealed ghosts was 25 kcal/mol, similar to that found in intact red blood cells. Inhibitors of pyruvate influx included such anions as thiocyanate, chloride, bicarbonate, alpha-cyanocinnamate, salicylate and ketomalonate (but not acetate); noncompetitive inhibitors were phloretin, 1-fluoro-2,4-dinitrobenzene, 4-acetamido-4'-isothiocyanate-stilbene-2,2'-disulfonic acid and o-phenanthroline/CuSO4 mixtures. The last reagent, known to induce disulfide links in certain membrane proteins, blocked the ionic strength stimulation of pyruvate influx in this study.  相似文献   

17.
Control of oxidative metabolism was studied using 13C NMR spectroscopy to detect rate-limiting steps in 13C labeling of glutamate. 13C NMR spectra were acquired every 1 or 2 min from isolated rabbit hearts perfused with either 2.5 mM [2-13C]acetate or 2.5 mM [2-13C]butyrate with or without KCl arrest. Tricarboxylic acid cycle flux (VTCA) and the exchange rate between alpha-ketoglutarate and glutamate (F1) were determined by least-square fitting of a kinetic model to NMR data. Rates were compared to measured kinetics of the cardiac glutamate-oxaloacetate transaminase (GOT). Despite similar oxygen use, hearts oxidizing butyrate instead of acetate showed delayed incorporation of 13C label into glutamate and lower VTCA, because of the influence of beta-oxidation: butyrate = 7.1 +/- 0.2 mumol/min/g dry wt; acetate = 10.1 +/- 0.2; butyrate + KCl = 1.8 +/- 0.1; acetate + KCl = 3.1 +/- 0.1 (mean +/- SD). F1 ranged from a low of 4.4 +/- 1.0 mumol/min/g (butyrate + KCl) to 9.3 +/- 0.6 (acetate), at least 20-fold slower than GOT flux, and proved to be rate limiting for isotope turnover in the glutamate pool. Therefore, dynamic 13C NMR observations were sensitive not only to TCA cycle flux but also to the interconversion between TCA cycle intermediates and glutamate.  相似文献   

18.
This study investigated 1) red blood cells (RBC) rigidity and 2) lactate influxes into RBCs in endurance-trained athletes with and without exercise-induced hypoxemia (EIH). Nine EIH and six non-EIH subjects performed a submaximal steady-state exercise on a cyclo-ergometer at 60% of maximal aerobic power for 10 min, followed by 15 min at 85% of maximal aerobic power. At rest and at the end of exercise, arterialized blood was sampled for analysis of arterialized pressure in oxygen, and venous blood was drawn for analysis of plasma lactate concentrations and hemorheological parameters. Lactate influxes into RBCs were measured at three labeled [U-14C]lactate concentrations (1.6, 8.1, and 41 mM) on venous blood sampled at rest. The EIH subjects had higher maximal oxygen uptake than non-EIH (P < 0.05). Total lactate influx was significantly higher in RBCs from EIH compared with non-EIH subjects at 8.1 mM (1,498.1 +/- 87.8 vs. 1,035.9 +/- 114.8 nmol.ml(-1).min(-1); P < 0.05) and 41 mM (2,562.0 +/- 145.0 vs. 1,618.1 +/- 149.4 nmol.ml(-1).min(-1); P < 0.01). Monocarboxylate transporter-1-mediated lactate influx was also higher in EIH at 8.1 mM (P < 0.05) and 41 mM (P < 0.01). The drop in arterial oxygen partial pressure was negatively correlated with total lactate influx measured at 8.1 mM (r = -0.82, P < 0.05) and 41 mM (r = -0.84, P < 0.05) in the two groups together. Plasma lactate concentrations and hemorheological data were similar in the two groups at rest and at the end of exercise. The results showed higher monocarboxylate transporter-1-mediated lactate influx in the EIH subjects and suggested that EIH could modify lactate influx into erythrocyte. However, higher lactate influx in EIH subjects was not accompanied by an increase in RBC rigidity.  相似文献   

19.
The effect of a single bout of exhaustive exercise on muscle lactate transport capacity was studied in rat skeletal muscle sarcolemmal (SL) vesicles. Rats were assigned to a control (C) group (n = 14) or an acutely exercised (E) group (n = 20). Exercise consisted of treadmill running (25 m/min, 10% grade) to exhaustion. SL vesicles purified from C and E rats were sealed because of sensitivity to osmotic forces. The time course of 1 mM lactate uptake in zero-trans conditions showed that the equilibrium level in the E group was significantly lower than in the C group (P < 0.05). The initial rate of 1 mM lactate uptake decreased significantly from 2.44 +/- 0.22 to 1.03 +/- 0.08 nmol. min(-1). mg protein(-1) (P < 0.05) after exercise, whereas that of 50 mM lactate uptake did not differ significantly between the two groups. For 100 mM external lactate concentration ([lactate]), exhaustive exercise increased initial rates of lactate uptake (219.6 +/- 36.3 to 465.4 +/- 80.2 nmol. min(-1). mg protein(-1), P < 0.05). Although saturation kinetics were observed in the C group with a maximal transport velocity of 233 nmol. min(-1). mg protein(-1) and a Michealis-Menten constant of 24.5 mM, saturation properties were not seen after exhaustive exercise in the E group, because initial rates of lactate uptake increased linearly with external [lactate]. We conclude that a single bout of exhaustive exercise significantly modified SL lactate transport activity, resulting in a decrease in 1 mM lactate uptake and was associated with alterations in the saturable properties at [lactate] above 50 mM. These results suggest that changes in sarcolemmal lactate transport activity may alter lactate and proton exchanges after exhaustive exercise.  相似文献   

20.
In this study, the mechanism of ischaemia-induced increased sarcolemmal permeability, as manifested by release of intracellular enzymes, was investigated. The role of changes in the sarcolemmal phospholipid bilayer in this process was evaluated by experimental modulation of the phospholipid fatty acid composition. The isolated perfused rat heart subjected to low-flow hypoxia, was used as a model of global ischaemia. Glucose as well as saturated (palmitate) and unsaturated (linoleate) long-chain fatty acids were used as substrates. Hearts perfused with palmitate or linoleate (1.5 mM, fatty acid/albumin ratio, 3.4) showed a significantly higher rate of lactate dehydrogenase release in both control and ischaemic conditions than hearts perfused with glucose (10 mM). Lactate dehydrogenase release in the fatty acid-perfused hearts was associated with a significant increase in the percentage unsaturation of the sarcolemmal phospholipid fatty acids. Glucose-perfused hearts, on the other hand, showed only minor changes in the sarcolemmal phospholipid fatty acid composition. Attempts to correlate enzyme release directly with an increase in the percentage unsaturation of phospholipid fatty acids failed, since enzyme release was also stimulated in control fatty-acid-perfused hearts which (when compared with glucose) contained a higher percentage saturated phospholipid fatty acids. The results suggest that myocardial ischaemia, apart from changes in the sarcolemmal phospholipid fatty acid composition, also induces several other changes in sarcolemmal composition (e.g., cholesterol loss) which may affect is permeability for macromolecules.  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