首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
The increase of sorbitol and fructose levels caused by aldose reductase activation and sorbitol dehydrogenase inhibition were observed in sciatic nerve of streptozotocin-diabetic rats. Elevated polyol pathway activity has been implicated in the development of diabetic complications such as neuropathy. The regulation of polyol pathway enzymes is based on the changes of redox state of free nicotinamide nucleotides. The decrease of the NADP+/NADPH ratio in cytosolic compartment of sciatic nerve cells activated aldose reductase and the decrease of the NAD+/NADH ratio inhibited sorbitol dehydrogenase. Nicotinamide as a precursor of NAD+ biosynthesis increased the free NADP+/NADPH and NAD+/NADH ratios and inhibited the activity of polyol pathway. The sorbitol level decreased in sciatic nerve of nicotinamide-treated streptozotocin-diabetic rats as compared to non-treated ones. Thus, the data provide evidence for important role of nicotinamide, as an antidiabetic drug, in prevention or correction of diabetic neuropathy.  相似文献   

2.
One of the major mechanisms involved in diabetic microangiopathy is considered to be an altered polyol pathway. However, clarifying the pathophysiology is difficult due to the lack of a sensitive method for measuring the reduction of glucose to sorbitol in tissue. Here we report a sensitive and selective method for polyol measurement using trifluoroacetyl (TFA) derivatives of polyols and stable isotope-labeled D-sorbitol (U-[13C]sorbitol, 13C6H14O6, 98.7%) as an internal standard. Gas chromatography—mass spectrometry (GC—MS) using an SE-30 capillary column gave elution of TFA derivatives of sugars, polyols and U-[13C]sorbitol within 8 min, with clear separation of sorbitol. In the calibration study, the coefficients of correlation between the amount of sorbitol added and that determined in standard solutions containing 0.1–8.0 nmol sorbitol, erythrocyte mixture and liver cytosol mixture were r=0.999, r=0.997 and r=0.997, respectively. The precision of the GC—MS measurement of standard solution was C.V.=4.3%. Because glucose is used as a substrate, the method can clarify the polyol pathway under physiological conditions. With this method, Km and Vmax values of the reductase in erythrocytes were 115±19 mmol/l and 4.42±0.26 nmol/min/g of hemoglobin. In human liver, on the other hand, they were 75±132 mmol/l and 0.77±0.090 nmol/min/mg of protein, respectively. This difference of Km values suggested that aldehyde reductase rather than aldose reductase is mainly responsible for reducing glucose to sorbitol in the liver. In conclusion, this newly developed method offers a highly sensitive and selective procedure for measuring low concentrations of sorbitol in various tissues and cells and should enable clarification of the kinetics of glucose reduction to sorbitol, which in turn can be used to evaluate the role of an altered polyol pathway in the pathophysiology of diabetic microangiopathy.  相似文献   

3.
NADPH and NADP+ levels were measured in rat lens from normal controls, from galactose-fed and diabetic rats during the first week of cataract formation.The level of NADPH in normal rat lens was determined to be 12.3 ± 0.4 nmol/g wet weight, and that of NADP+ 4.6 ± 0.2 nmol/g wet weight. In early cataract formation NADPH levels decreased rapidly during the first 2 days and then remained stable at 76% of control for galactose-fed and 84% for diabetic rats. NADP+ levels increased by 38% of control for galactose-fed and 54% for diabetic rats. Calculated NADPH/NADP+ ratios dropped from 3.36 ± 0.21 to 1.86 ± 0.16 in galactose fed rats, and from 2.81 ± 0.15 to 1.61 ± 0.16 in diabetic rats (P < 0.001 for both experimental groups). These data are consistent with rapid NADPH oxidation during onset of lens cataracts. No significant changes in aldose reductase enzymatic activity levels were observed in either the galactosemic or the diabetic rats during the times measured.  相似文献   

4.
We wished to examine the effects of diabetes on muscle glutamine kinetics. Accordingly, female Wistar rats (200 g) were made diabetic by a single injection of streptozotocin (85 mg/kg) and studied 4 days later; control rats received saline. In diabetic rats, glutamine concentration of gastrocnemius muscle was 33% less than in control rats: 2.60 ± 0.06 μmol/g vs. 3.84 ± 0.13 μmol/g (P < 0.001). In gastrocnemius muscle, glutamine synthetase activity (Vmax) was unaltered by diabetes (approx. 235 nmol/min per g) but glutaminase Vmax increased from 146 ± 29 to 401 ± 94 nmol/min per g; substrate Km values of neither enzyme were affected by diabetes. Net glutamine efflux (AZ concentration difference × blood flow) from hindlimbs of diabetic rats in vivo was greater than control values (?30.0 ± 3.2 vs. ?1.9 ± 2.6 nmol/min per g (P < 0.001) and hindlimb NH3 uptake was concomitantly greater (about 27 nmol/min per g). The glutamine transport capacity (Vmax) of the Na-dependent System Nm in perfused hindlimb muscle was 29% lower in diabetic rats than in controls (820 ± 50 vs. 1160 ± 80 nmol/min per g (P < 0.01)), but transporter Km was the same in both groups (9.2 ± 0.5 nM). The difference between inward and net glutamine fluxes indicated that glutamine efflux in perfused hindlimbs was stimulated in diabetes at physiological perfusate glutamine (0.5 mM); ammonia (1 mM in perfusate) had little effect on net glutamine flux in control and diabetic muscles. In Intramuscular Na+ was 26% greater in diabetic (13.2 μmol/g) than control muscle, but muscle K+ (100 μmol/g) was similar. The accelerated rate of glutamine release from skeletal muscle and the lower muscle free glutamine concentration observed in diabetes may result from a combination of; (i), a diminished Na+ electrochemical gradient (i.e., the net driving force for glutamine accrual in muscle falls); (ii), a faster turnover of glutamine in muscle and (iii), an increased Vmax/Km for sarcolemmal glutamine efflux.  相似文献   

5.
A yeast growing at 48°C was isolated from soil and the strain was identified as Cryptococcus lactativorus. The aldose reductase which the strain produced was purified 114-fold with an overall recovery of 36%. The stability of the enzyme was higher than that of other aldose reductases. The half life of the enzyme was 800 h and 14 h at 30°C and 50°C, respectively. The enzyme showed the best activity with d-xylose. l-Sorbose and d-fructose were also reduced by the enzyme. The enzyme was active with both NADPH and NADH as a conenzyme, and the activity with NADH was 1.25 times higher than that with NADPH. The Kmapp value for d-xylose was 8.6 mM and the Vmaxapp was 20.8 units/mg NADH was used as a coenzyme. The Kmapp values for NADPH and NADH were 6μM and 170 μM, respectively, when d-glucose was used as a substrate.  相似文献   

6.
ω-Hydroxyfatty acid:NADP oxidoreductase, an enzyme involved in suberin biosynthesis, is induced by wounding potato tubers. Initial velocity and product inhibition studies with the purified enzyme suggested an ordered sequential mechanism, where NADPH is added first, followed by 16-oxohexadecanoate, and NADP is released after 16-hydroxyhexadecanoate. Substrate inhibition by NADPH was observed at concentrations higher than 0.2 mm. The inhibitory NADPH molecule competes with 16-oxohexadecanoate, indicating that it forms a dead-end complex with the E-NADPH form of the enzyme. The kinetics for the NADPH inhibition suggested that n > 1 in the rate equation v = V[NADPH](Km + [NADPH]+ [NADPH]n+1Ki); i.e., more than two NADPH molecules bind to enzyme. The Km for 16-oxohexadecanoate did not change from pH 7.5 to 9.0 but increased about 10-fold from pH 9.0 to 10.0, whereas the Km for NADPH and hexadecanal did not vary significantly in this pH range. Phenylglyoxal inactivated the enzyme; NADPH and AMP (which competes with NADPH; Ki = 1.1 mM) provided protection against such inactivation. Diethylpyrocarbonate also caused inactivation which was reversed by hydroxylamine; NADPH but not AMP protected the enzyme from this inhibition. Pyridoxal-5′-phosphate reversibly inactivated the enzyme and NaBH4 reduction of the pyridoxal phosphate-treated enzyme resulted in irreversible inhibition; a combination of NADPH and ω-oxo C16 acid provided protection against such inactivation. As the chain length of alkanals increased from C3 to C8, the Km for the substrate decreased drastically from 7000 to 90μm and a further increase in chain length from C8 to C20 resulted in only a small decrease in Km. The Km and V for 8-oxooctanoate and 10-oxodecanoate are compared with the values obtained for 16-oxohexadecanoate. Based on these results, it is proposed that arginine acts as the binding site for NADPH, a hydrophobic crevice with lysine at the bottom forms the binding site for 16-oxohexadecanoate and histidine participates in the reaction as the proton donor.  相似文献   

7.
Aldose reductase inhibitors have considerable potential for the treatment of diabetic complications, without increased risk of hypoglycemia. Search for components inhibiting aldose reductase led to the discovery of active compounds contained in Evodia rutaecarpa Bentham (Rutaceae), which is the one of the component of Kampo-herbal medicine. The hot water extract from the E. rutaecarpa was subjected to distribution or gel filtration chromatography to give an active compound, N2-(2-methylaminobenzoyl)tetrahydro-1H-pyrido[3,4-b]indol-1-one (rhetsinine). It inhibited aldose reductase with IC50 values of 24.1 μM. Furthermore, rhetsinine inhibited sorbitol accumulation by 79.3% at 100 μM. These results suggested that the E. rutaecarpa derived component, rhetsinine, would be potentially useful in the treatment of diabetic complications.  相似文献   

8.
NADPH and NADP+ levels were measured in rat lens from normal controls, from galactose-fed and diabetic rats during the first week of cataract formation. The level of NADPH in normal rat lens was determined to be 12.3 +/- 0.4 nmol/g wet weight, and that of NADP+ 4.6 +/- 0.2 nmol/g wet weight. In early cataract formation NADPH levels decreased rapidly during the first 2 days and then remained stable at 76% of control for galactose-fed and 84% for diabetic rats. NADP+ levels increased by 38% of control for galactose-fed and 54% for diabetic rats. Calculated NADPH/NADP+ ratios dropped from 3.36 +/- 0.21 to 1.86 +/- 0.16 in galactose fed rats, and from 2.81 +/- 0.15 to 1.61 +/- 0.16 in diabetic rats (P less than 0.001 for both experimental groups). These data are consistent with rapid NADPH oxidation during onset of lens cataracts. No significant changes in aldose reductase enzymatic activity levels were observed in either the galactosemic or the diabetic rats during the times measured.  相似文献   

9.
D.K. Srivastava  L.E. Anderson 《BBA》1983,724(3):359-369
Glucose-6-phosphate dehydrogenase (D-glucose-6-phosphate: NADP oxidoreductase, EC 1.1.1.49) has been purified to electrophoretic homogeneity from pea chloroplasts. The enzyme, which has a Stokes radius of 52 Å, is a tetramer made up of four 56000 Da monomers. The pH optimum is around 8.2. The enzyme is absolutely specific for NADP. The apparent Km(NADP) is 2.4 ± 0.1 μM. NADPH inhibition of the enzyme is competitive with respect to NADP (mean Ki, 18 ± 5 μM) and is mixed (Kp >Km, Vmax >Vp) with respect to glucose 6-phosphate (mean crossover point, 0.5 ± 0.1 mM). The apparent Km(glucose 6-phosphate) is 0.37 ± 0.01 mM. The purified enzyme is inactivated in the light in the presence of dilute stroma and washed thylakoids, and by dithiothreitol. Enzyme which has been partially inactivated by treatment with dithiothreitol can be further inactivated in the light in the presence of dilute stroma and washed thylakoids and reactivated in the dark, but only to the extent of the reverse of light inactivation. Dithiothreitol-inactivated enzyme is not reactivated further by addition of crude stroma or oxidized thioredoxin. Dithiothreitol-dependent inactivation of the enzyme follows pseudo-first-order kinetics and shows rate saturation. The enzyme which has been partially inactivated by treatment with dithiothreitol does not differ from the untreated control with respect to thermal and tryptic inactivation. However, enzyme which has been partially light inactivated shows different thermal and tryptic inactivation patterns as compared to the dark control. These observations suggest that the changes in the enzyme brought about by light modulation are not necessarily identical with those brought about by dithiothreitol inactivation.  相似文献   

10.
  • 1.1. Aldose reductase, aldehyde reductase and high-Km, aldose reductase were purified from the inner medulla of dog kidney.
  • 2.2. Compared with aldose reductase, high-Km aldose reductase had a lower isoelectric point, a lower activity for aldo-sugars and a lower sensitivity for aldose reductase inhibitors, and it was not activated by sulfate ions. Both reductases had the same molecular weight (38,500) and immunochemical properties.
  • 3.3. High-Km aldose reductase was easily converted into an aldose reductase-like enzyme, namely a generated reductase upon incubation in neutral buffer solution.
  • 4.4. The generated reductase was identical with aldose reductase with respect to the isoelectric point, substrate specificity, activation by sulfate ions and IC50 values for aldose reductase inhibitors. The generated reductase revealed immunochemical identity with aldose reductase as well as high-Km aldose reductase.
  相似文献   

11.
Targeting long-term diabetic complications, as well as inflammatory pathologies, aldose reductase inhibitors (ARIs) have been gaining attention over the years. In the present work, in order to address the poor membrane permeation of previously reported ARIs, derivatives of N-phenylpyrrole, bearing groups with putative pKa  7.4, were synthesized and evaluated for aldose reductase inhibitory activity. The 2-fluorophenol group proved the most promising moiety, and further modifications were explored. The most active compound (31), identified as a submicromolar inhibitor (IC50 = 0.443 μM), was also selective against the homologous enzyme aldehyde reductase. Cross-docking revealed that 31 displays a peculiar interaction network that may be responsible for high affinity. Physicochemical profiling of 31 showed a pKa of 7.64, rendering it less than 50% ionized in the physiological pH range, with potentially favorable membrane permeation. The latter was supported from the successful inhibition of sorbitol formation in rat lenses and the ability to permeate rat jejunum.  相似文献   

12.
13.
6-Phosphogluconate dehydrogenase (6PG) was purified from rat small intestine with 36% yield and a specific activity of 15 U/mg. On SDS/PAGE, one band with a mass of 52 kDa was found. On native PAGE three protein and two activity bands were observed. The pH optimum was 7.35. Using Arrhenius plots, Ea, ΔH, Q10 and Tm for 6PGD were found to be 7.52 kcal/mol, 6.90 kcal/mol, 1.49 and 49.4°C, respectively. The enzyme obeyed “Rapid Equilibrium Random Bi Bi” kinetic model with Km values of 595 ± 213 μM for 6PG and 53.03±1.99 μM for NADP. 1/Vm versus 1/6PG and 1/NADP plots gave a Vm value of 8.91±1.92 U/mg protein. NADPH is the competitive inhibitor with a Ki of 31.91±1.31 μM. The relatively small Ki for the 6PGD:NADPH complex indicates the importance of NADPH in the regulation of the pentose phosphate pathway through G6PD and 6PGD.  相似文献   

14.
It has been suggested (Ueda, T., Otsuka, H. and Goda, K. (1978) J. Biochem. 84, 687–696) that direct cleavage of kynurenine, catalysed by kynureninase, followed by microsomal hydroxylation of the resultant anthranilic acid, may provide an alternative to the established pathway of kynurenine metabolism that involves direct hydroxylation followed by cleavage to 3-hydroxyanthranilic acid. To test this suggestion, anthranilic acid was administered to rats; there was no increase in either the concentration of nicotinamide nucleotides in the liver or the urinary excretion of N1-methyl nicotinamide. However, injection of either kynurenine or 3-hydroxyanthranilic acid did increase the concentration of nicotinamide nucleotides in the liver. The kinetics of kynurenine hydroxylase (Km = 1.8±0.6·10?5 mol/l) and kynureninase (Km = 2.5±0.8·10?4 mol/l, liver steady-state kynurenine = 4.9±0.9 μmol/kg) are such that the preferred route of kynurenine metabolism is probably by way of hydroxylation rather than cleavage.  相似文献   

15.
Numerous physiological aldehydes besides glucose are substrates of aldose reductase, the first enzyme of the polyol pathway which has been implicated in the etiology of diabetic complications. The 2-oxoaldehyde methylglyoxal is a preferred substrate of aldose reductase but is also the main physiological substrate of the glutathione-dependent glyoxalase system. Aldose reductase catalyzes the reduction of methylglyoxal efficiently (kcat=142 min−1 and kcat/Km=1.8×107 M−1 min−1). In the presence of physiological concentrations of glutathione, methylglyoxal is significantly converted into the hemithioacetal, which is the actual substrate of glyoxalase-I. However, in the presence of glutathione, the efficiency of reduction of methylglyoxal, catalyzed by aldose reductase, also increases. In addition, the site of reduction switches from the aldehyde to the ketone carbonyl. Thus, glutathione converts aldose reductase from an aldehyde reductase to a ketone reductase with methylglyoxal as substrate. The relative importance of aldose reductase and glyoxalase-I in the metabolic disposal of methylglyoxal is highly dependent upon the concentration of glutathione, owing to the non-catalytic pre-enzymatic reaction between methylglyoxal and glutathione.  相似文献   

16.
Methylglyoxal reductase was purified from Hansenula mrakii IFO 0895 to a homogenous state on polyacrylamide gel electrophoresis. The enzyme consisted of a single polypeptide chain with a molecular weight of 34,000. The enzyme was specific to methylglyoxal (Km = 1.92 mM) and NADPH (Km = 40.8 μM). The activity of the enzyme was inhibited by p-chloromercuribenzoate and HgCl2. NADP also inhibited the activity of the enzyme, and the Ki value was calculated to be 0.25 mM.  相似文献   

17.
Effect of oral administration of sodium orthovanadate for three weeks on polyol pathway in renal cortex and medulla was studied in control and alloxan diabetic rats. An enhancement in aldose reductase in cortex and medulla and sorbitol dehydrogenase in cortex was observed in alloxan diabetic rats. Despite depressed insulin secretion, vanadate treatment to diabetic rats counteracted hyperglycemia, normalized elevated enzyme activities and glucose level, prevented medullary sorbitol accumulation and markedly checked increase in kidney weight. These results show that vanadate causes marked improvement in renal hypertrophy and has an antidiabetogenic effect on polyol pathway in diabetic kidney.  相似文献   

18.
5,10-Methylenetetrahydrofolate reductase (EC 1.1.1.68) was purified from the cytosolic fraction of sheep liver by (NH4)2 SO4 fractionation, acid precipitation, DEAE-Sephacel chromatography and Blue Sepharose affinity chromatography. The homogeneity of the enzyme was established by sodium dodecyl sulphate-polyacrylamide gel electrophoresis, ultracentrifugation and Ouchterlony immunodiffusion test. The enzyme was a dimer of molecular weight 1,66,000 ± 5,000 with a subunit molecular weight of 87,000 ±5,000. The enzyme showed hyperbolic saturation pattern with 5-methyltetrahydrofolate.K 0.5 values for 5-methyltetrahydrofolate menadione and NADPH were determined to be 132 ΜM, 2.45 ΜM and 16 ΜM. The parallel set of lines in the Lineweaver-Burk plot, when either NADPH or menadione was varied at different fixed concentrations of the other substrate; non-competitive inhibition, when NADPH was varied at different fixed concentrations of NADP; competitive inhibition, when menadione was varied at different fixed concentrations of NADP and the absence of inhibition by NADP at saturating concentration of menadione, clearly established that the kinetic mechanism of the reaction catalyzed by this enzyme was ping-pong.  相似文献   

19.
Salicylate is an important intermediate in the bacterial degradation of polycyclic aromatic hydrocarbons and salicylate hydroxylases play essential roles in linking the peripheral and ring-cleavage catabolic pathways. Unlike the well-characterized salicylate 1-hydroxylases, the rarely occurred salicylate 5-hydroxylase (S5H) has not been characterized in detail. In this study, the three-component Fe-S protein complex (NagAaGHAb) of S5H from Ralstonia sp. strain U2 was purified, and its biochemical and catalytic properties were characterized. The oxygenase component NagGH exhibited an α3β3 heterohexameric structure and contained one Rieske-type [2Fe-2S] cluster and one mononuclear iron per α subunit. NagAa is the ferredoxin-NADP+ reductase component containing flavin and plant type [2Fe–2S] cluster. The ferredoxin component NagAb was characterized as a [2Fe-2S] dimer which remains remarkably stable in denaturing gel electrophoresis after being heated at 100 °C for 1 h. Purified NagAa and NagAb, NagGH catalyzed the hydroxylation of salicylate to gentisate with a specific activity of 107.12?±?14.38 U/g and showed an apparent K m for salicylate of 102.79?±?27.20 μM and a similar K m value for both NADH and NADPH (59.76?±?7.81 μM versus 56.41?±?12.76 μM). The hydroxylase exhibited different affinities for two hydroxysalicylates (2,4-dihydroxybenzoate K m of 93.54?±?18.50 μM versus 2,6-dihydroxybenzoate K m of 939.80?±?199.46 μM). Interestingly, this S5H also showed catalytic activity to the pollutant 2-nitrophenol and exhibited steady-state kinetic data of the same order of magnitude as those for salicylate. This study will allow further comparative studies of structure–function relationships of the ring hydroxylating mono- and di-oxygenase systems.  相似文献   

20.
L-929 cells acclimated to media made hyperosmotic (600 mosmol/kgH2O) by addition of NaCl, sorbitol, or mannitol show, on SDS-polyacrylamide gels, a markedly enhanced protein band at 40 kDa, most likely corresponding to the enzyme aldose reductase. The effect was not observed in cells acclimated to a medium rendered hyperosmotic by addition of proline. The major organic osmolyte accumulated is sorbitol in cells acclimated to high-sorbitol or high-NaCl medium, proline in cells acclimated to high-proline medium. Cells acclimated to any of these hyperosmotic media display unaltered Na+ levels and similarly increased K+ levels and decreased Cl levels. These results are interpreted in terms of the mechanisms involved in aldose reductase induction and in regulation of the enzyme activity in long-term acclimation to hyperosmotic media. © 1996 Wiley-Liss, Inc.  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