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 diversity of pigmentation in the skin, hair, and eyes of humans has been largely attributed to the diversity of pH in melanosomes with acidic pH being proposed to suppress melanin production. Tyrosinase has an optimum pH of 7.4 and its activity is suppressed greatly at lower pH values. The first step of eumelanogenesis is the oxidation of tyrosine to dopachrome (DC) via dopaquinone. However, how eumelanogenesis is controlled by pH beyond this stage is not known. In this study, we examined the effects of pH (5.3–7.3) on the conversion of DC to 5,6‐dihydroxyindole (DHI) and 5,6‐dihydroxyindole‐2‐carboxylic acid (DHICA) and the subsequent oxidation of DHI and DHICA to form eumelanin. The effects of Cu2+ ions on those reactions were also compared. The results indicate that an acidic pH greatly suppresses the late stages of eumelanogenesis and that Cu2+ ions accelerate the conversion of DC to DHICA and its subsequent oxidation.  相似文献   

2.
An enzyme-free amperometric method was established for glucose detection using a nanoporous gold film (NPGF) electrode prepared by a rapid one-step anodic potential step method within 5 min. The prepared NPGF had an extremely high roughness and was characterized by scanning electron microscopy (SEM) and cyclic voltammetry. Electrochemical responses of the as-prepared NPGF to glucose in 0.1M phosphate buffer solution (PBS, pH 7.4) with or without Cl(-) were discussed. In amperometric studies carried out at -0.15 V in the absence of Cl(-), the NPGF electrode exhibited a high sensitivity of 232 μA mM(-1)cm(-2) and gave a linear range from 1mM up to 14 mM with a detection limit of 53.2 μM (with a signal-to-noise ratio of 3). In addition, the oxidation of ascorbic acid (AA) and uric acid (UA) can be completely eliminated at such a low applied potential. On the other hand, the quantification of glucose in 0.1M PBS (pH 7.4) containing 0.1M NaCl offered an extended linear range from 10 μM to 11 mM with a sensitivity of 66.0 μA mM(-1)cm(-2) and a low detection limit of 8.7 μM (signal-to-noise ratio of 3) at a detection potential of 0.2V.  相似文献   

3.
《Nitric oxide》2000,4(1):4-14
Under aerobic conditions, exposure of peroxidized lipids to nitric oxide (NO) was found to result in a rapid decrease in the levels of thiobarbituric acid-reactive substances (TBARS). Addition of 10–100 μM NO to rat brain homogenates preincubated for 2 h at 37°C caused up to a 20% decrease in the levels of TBARS compared to controls. A similar inhibitory effect was observed on TBARS produced by Fe2+-induced decomposition of 15-hydroperoxyeicosatetraenoic acid (15-HPETE), due apparently to NO-induced decomposition of the hydroperoxide (ferrous oxidation/xylenol orange assay). Prostaglandin G2 (PGG2, 35 μM), as a model bicyclic endoperoxide, and malondialdehyde (MDA, 20 μM), the main component of TBARS, proved also susceptible to degradation by NO or NO donors (diethylamine NONOate, DEA/NO) at concentrations of 100 μM or higher in 0.05 M phosphate buffer, pH 7.4, and at 37°C, as indicated by the reduced response to the TBA assay. No significant effect on TBARS determination was caused by nitrite ions. These and other data indicate that NO can inhibit TBARS formation by decomposing primary lipid peroxidation products, chiefly 15-HPETE and related hydroperoxides, and, to a lesser extent, later stage TBARS precursors, including bicyclic endoperoxides and MDA, via nitrosation and other oxidative routes, without however affecting chromogenic reactions during the assay.  相似文献   

4.
Shelkovnikov S  Gonick HC 《Life sciences》2004,75(23):2765-2773
Nitric oxide (NO) and peroxynitrite (ONOO) have been reported to destroy catecholamines. We compared the ability of NO donors and peroxynitrite to decompose epinephrine in both chemical and pharmacological experiments. Epinephrine (1 microM) was incubated with NO donors (SNAP and MAHMA NONOate) and ONOO at a concentration of 0.1 mM in phosphate buffer (pH 7.4; 0.1 M) or Krebs solution for 10 minutes at 37 degrees C. HPLC revealed that the concentration of epinephrine in the presence of NO donors was unaltered. In contrast, peroxynitrite decreased epinephrine concentration more than 20 fold. Similar relationships were obtained in the study of rat thoracic aorta ring contraction. The contractile activity (EC50) of epinephrine in control solutions and after incubation of NE with NO donors did not change. EC50 was measured at 8-10 nM in control solutions and after preincubation with NO donors. However when epinephrine was preincubated with peroxynitrite, no contractile effect was evoked. Therefore, under these experimental conditions peroxynitrite, but not NO donors, was capable of destroying epinephrine.  相似文献   

5.
Eumelanins in animals are biosynthesized by the combined action of tyrosinase, 3,4-dihydroxyphenylalanine (DOPA)chrome isomerase, and other factors. Two kinds of eumelanins were characterized from mammalian systems; these are 5,6-dihydroxyindole (DHI)-melanin and 5,6-dihydroxyindole-2-carboxylic acid (DHICA)-melanin. In insects, melanin biosynthesis is initiated by phenoloxidase and supported by DOPAchrome isomerase (decarboxylating). Based on the facts that DOPA is a poor substrate for insect phenoloxidases and DHI is the sole product of insect DOPAchrome isomerase reaction, it is proposed that insects lack DHICA-melanin. Accordingly, the phenoloxidase isolated from the hemolymph of Manduca sexta failed to oxidize DHICA. Control experiments reveal that mushroom tyrosinase, as well as laccase, which is a contaminant in the commercial preparations of mushroom tyrosinase, are capable of oxidizing DHICA. Neither the whole hemolymph nor the cuticular extracts of M. sexta possessed any detectable oxidase activity towards this substrate. Thus, insects do not seem to produce DHICA-eumelanin. A useful staining procedure to localize DHICA oxidase activity on gels is also presented.  相似文献   

6.
Several studies have confirmed that a melanocyte-specific enzyme, dopachrome tautomerase (EC 5.3.2.3), catalyzes the isomerization of dopachrome to 5,6-dihydroxyindole-2-carboxylic acid (DHICA) (Pawelek, 1991). Here we report that DHICA, produced either enzymatically with dopachrome tautomerase or through chemical synthesis, spontaneously polymerized to form brown melanin that was soluble in aqueous solutions above pH 5. Under the same reaction conditions, solutions of either DOPA, DOPAchrome, or 5,6-dihydroxyindole (DHI) formed black, insoluble melanin precipitates. When DHICA and DHI were mixed together, with DHICA in molar excess, little or no precipitation of DHI-melanin occurred and the rate and extent of soluble melanin formation was markedly enhanced over that achieved with DHICA alone, suggesting co-polymerization of DHICA and DHI. With or without DHI, DHICA-melanins absorbed throughout the ultraviolet and visible spectra (200-600 nm). The DHICA-melanins precipitated below pH 5, at least in part because of protonation of the carboxyl groups. DHICA-melanins could be passed through 0.22 μm filters but could not be dialyzed through semi-permeable membranes with exclusion limits of 12,000-14,000 daltons. HPLC/molecular sieve analyses revealed apparent molecular weights ranging from 20,000 to 200,000 daltons, corresponding to 100-1,000 DHICA monomers per molecule of melanin. DHICA-melanins were stable to boiling, lyophilization, freezing and thawing, and incubation at room temperature for more than 1 year. The natural occurrence of oligomers of DHICA was first reported by Ito and Nichol (1974) in their studies of the brown tapetal pigment in the eye of the sea catfish (Arius felis L.). In experiments reported here, brown, but not black, melanins from mouse hairs, human melanoma cells, and peacock feathers were soluble in aqueous buffers. Since DHICA-melanins are both soluble and brown, the results raise the possibility that they are determinants of brown colors in the animal kingdom.  相似文献   

7.
Eumelanin is a brown‐black pigment comprising 5,6‐dihydroxyindole (DHI) and its 2‐carboxy derivative (DHICA), but the detailed structure of eumelanin is unclear. Chemical degradation is a powerful tool for analyzing melanin. H2O2 oxidation degradation of eumelanin affords pyrrole‐2,3,5‐tricarboxylic acid (PTCA) and pyrrole‐2,3‐dicarboxylic acid (PDCA). The ratio of PDCA to PTCA provides information about the eumelanin structure. In this article, we propose simple equations on the basis of previous experimental results on dimer yields for evaluating the yields of PTCA and PDCA from any DHI oligomers. Assuming the chemical disorder model of DHI‐melanin, we solve an equation where a theoretical expression for the ratio of PDCA to PTCA is set to the corresponding experimental value to obtain a plausible Poisson distribution of DHI oligomers. The results demonstrate that the main contributors to DHI‐melanin are tetramers and pentamers as shown by the mass spectrometry.  相似文献   

8.
To inquire into the role of the carboxyl group as determinant of the properties of 5,6‐dihydroxyindole melanins, melanins from aerial oxidation of 5,6‐dihydroxyindole‐2‐carboxylic acid (DHICA) and its DHICA methyl ester (MeDHICA) were comparatively tested for their antioxidant activity. MALDI MS spectrometry analysis of MeDHICA melanin provided evidence for a collection of intact oligomers. EPR analysis showed g‐values almost identical and signal amplitudes (ΔB) comparable to those of DHICA melanin, but spin density was one order of magnitude higher, with a different response to pH changes. Antioxidant assays were performed, and a model of lipid peroxidation was used to compare the protective effects of the melanins. In all cases, MeDHICA melanin performed better than DHICA melanin. This capacity was substantially maintained following exposure to air in aqueous buffer over 1 week or to solar simulator over 3 hr. Different from DHICA melanin, MeDHICA melanin was proved to be fairly soluble in different water‐miscible organic solvents, suggesting its use in dermocosmetic applications.  相似文献   

9.
Direct microcalorimetric measurements were made of the reaction between acetylcholine chloride and acetylcholinesterase (EC 3.1.1.7) that was extracted from electric eel (Electrophorus electricus) and purified by affinity chromatography. Tris-HCl, sodium phosphate and potassium phosphate were used as buffers and sources of ions for the reaction. At pH 7.2 and in 0.1-0.2 M phosphate buffer, the delta H for acetylcholine hydrolysis was found to be -0.107 kcal/mol (under buffered conditions) and -0.931 kcal/mol under unbuffered conditions (water). At pH 8.0 in 0.1 M Tris-HCl buffer, values greater than -2.5 kcal/mol were obtained, with the highest value of -9.2 kcal/mol being seen with bovine erythrocyte acetylcholinesterase. Tris-HCl buffer at 4 X 10(-2) M enhanced the reaction velocity by 51.2% over that of 4 X 10(-3) M buffer. Enzyme purity, pH and ionic milieu of reaction mixture, and substrate concentration affected the measured delta H value.  相似文献   

10.
A pulse radiolytic investigation has been conducted to establish whether a redox reaction takes place between dopaquinone and 5,6-dihydroxyindole (DHI) and its 2-carboxylic acid (DHICA) and to measure the rate constants of the interactions. To obviate possible confounding reactions, such as nucleophilic addition, the method employed to generate dopaquinone used the dibromide radical anion acting on dopa to form the semiquinone which rapidly disproportionates to dopaquinone. In the presence of DHI the corresponding indole-5,6-quinone (and/or tautomers) was also formed directly but, by judicious selection of suitable relative concentrations of initial reactants, we were able to detect the formation of additional indolequinone from the redox exchange reaction of DHI with dopaquinone which exhibited a linear dependency on the concentration of DHI. Computer simulation of the experimental time profiles of the absorption changes showed that, under the conditions chosen, redox exchange does proceed but not quite to completion, a forward rate constant of 1.4 x 10(6)/M/s being obtained. This is in the same range as the rate constants previously established for reactions of dopaquinone with cyclodopa and cysteinyldopa. In similar experiments carried out with DHICA, the reaction more obviously does not go to completion and is much slower, k (forward) =1.6 x 10(5)/M/s. We conclude that, in the eumelanogenic pathway, DHI oxidation may take place by redox exchange with dopaquinone, although such a reaction is likely to be less efficient for DHICA.  相似文献   

11.
A highly purified trypsin inhibitor was obtained from Echinodorus paniculatus when an extract prepared from E. paniculatus seed flour (25 gl(-1), with 0.1 M ammonium acetate buffer, pH 8.3, under agitation for 6 min at 28 degrees C) was chromatographed on Sephadex G-25 (12 mlh(-1)), followed by affinity chromatography on immobilized Cratylia mollis isolectins (Cra Iso 1,2,3-Sepharose). The column chromatography was performed at 24 degrees C; the matrix was washed (30 mlh(-1)) with 0.1 M sodium phosphate buffer, pH 7.4 or with the same buffer containing 0.2 M glucose, followed by application of inhibitor sample and elution with 0.015 M sodium borate buffer, pH 7.4, or 1.0 M NaCl. A purified fraction of inhibitor was obtained by gel filtration chromatography (GF-450/HPLC column). Trypsin inhibitory activity was eliminated when the inhibitor was treated with metaperiodate showing that the carbohydrate moiety was important for trypsin inhibition. Binding of inhibitor was also evaluated on immobilized concanavalin A (Con A-Sepharose) using previously described chromatographic conditions with results similar to Cra Iso 1,2,3-Sepharose chromatography.  相似文献   

12.
Several studies have confirmed that a melanocyte-specific enzyme, dopachrome tautomerase (EC 5.3.2.3), catalyzes the isomerization of dopachrome to 5,6-dihydroxyindole-2-carboxylic acid (DHICA) (Pawelek, 1991). Here we report that DHICA, produced either enzymatically with dopachrome tautomerase or through chemical synthesis, spontaneously polymerized to form brown melanin that was soluble in aqueous solutions above pH 5. Under the same reaction conditions, solutions of either DOPA, DOPAchrome, or 5,6-dihydroxyindole (DHI) formed black, insoluble melanin precipitates. When DHICA and DHI were mixed together, with DHICA in molar excess, little or no precipitation of DHI-melanin occurred and the rate and extent of soluble melanin formation was markedly enhanced over that achieved with DHICA alone, suggesting co-polymerization of DHICA and DHI. With or without DHI, DHICA-melanins absorbed throughout the ultraviolet and visible spectra (200-600 nm). The DHICA-melanins precipitated below pH 5, at least in part because of protonation of the carboxyl groups. DHICA-melanins could be passed through 0.22 micron filters but could not be dialyzed through semi-permeable membranes with exclusion limits of 12,000-14,000 daltons. HPLC/molecular sieve analyses revealed apparent molecular weights ranging from 20,000 to 200,000 daltons, corresponding to 100-1,000 DHICA monomers per molecule of melanin. DHICA-melanins were stable to boiling, lyophilization, freezing and thawing, and incubation at room temperature for more than 1 year. The natural occurrence of oligomers of DHICA was first reported by Ito and Nichol (1974) in their studies of the brown tapetal pigment in the eye of the sea catfish (Arius felis L.).(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

13.
It was shown that ceruloplasmin, apart from the known oxidative conversion of dopamine into melanin, can also produce (DHI)-melanin from 5,6-dihydroxyindole and THP-melanin from tetrahydropapaveroline. Ceruloplasmin acts as an oxidase and the kinetic parameters for these oxidative reactions are reported. Since these ceruloplasmin-catalyzed reactions occur also at pH 7.4, they could have a significant physiological impact. This ceruloplasmin-oxidasic activity is enhanced by copper ions and inhibited by chelators, such as ethylenediaminetetraacetic acid (EDTA) and desferoxamine (DEF). Some possible implication of melanin production in blood are discussed.  相似文献   

14.
Dopachrome conversion, in which dopachrome is converted into 5,6‐dihydroxyindole (DHI) or 5,6‐dihydroxyindole‐2‐carboxylic acid (DHICA) upstream of eumelanogenesis, is a key step in determining the DHI/DHICA monomer ratio in eumelanin, which affects the antioxidant activity. Although the ratio of DHI/DHICA formed and the conversion rate can be regulated depending on pH, the mechanism is still unclear. To clarify the mechanism, we carried out first‐principles calculations. The results showed the kinetic preference of proton rearrangement to form quinone methide intermediate via β‐deprotonation. We also identified possible pathways to DHI/DHICA from the quinone methide. The DHI formation can be achieved by spontaneous decarboxylation after proton rearrangement from carboxyl group to 6‐oxygen. α‐Deprotonation, which leads to DHICA formation, can also proceed with a significantly reduced activation barrier compared with that of the initial dopachrome. Considering the rate of the proton rearrangements in a given pH, we conclude that the conversion is suppressed at acidic pH.  相似文献   

15.
Eumelanin photoprotects pigmented tissues from ultraviolet (UV) damage. However, UVA‐induced tanning seems to result from the photooxidation of preexisting melanin and does not contribute to photoprotection. We investigated the mechanism of UVA‐induced degradation of 5,6‐dihydroxyindole‐2‐carboxylic acid (DHICA)‐melanin taking advantage of its solubility in a neutral buffer and using a differential spectrophotometric method to detect subtle changes in its structure. Our methodology is suitable for examining the effects of various agents that interact with reactive oxygen species (ROS) to determine how ROS is involved in the UVA‐induced oxidative modifications. The results show that UVA radiation induces the oxidation of DHICA to indole‐5,6‐quinone‐2‐carboxylic acid in eumelanin, which is then cleaved to form a photodegraded, pyrrolic moiety and finally to form free pyrrole‐2,3,5‐tricarboxylic acid. The possible involvement of superoxide radical and singlet oxygen in the oxidation was suggested. The generation and quenching of singlet oxygen by DHICA‐melanin was confirmed by direct measurements of singlet oxygen phosphorescence.  相似文献   

16.
The regulation of the final steps of the melanogenesis pathway, after L-2-carboxy-2,3-dihydroindole-5,6-quinone (dopachrome) formation, is studied. It is shown that both tyrosinase and dopachrome tautomerase are involved in the process. In vivo, it seems that tyrosinase is involved in the regulation of the amount of melanin formed, whereas dopachrome tautomerase is mainly involved in the size, structure and composition of melanin, by regulating to the incorporation of 5,6-dihydroxyindole-2-carboxylic acid (DHICA) into the polymer. Moreover, using L-3,4-dihydroxyphenylalanine (dopa) and related compounds, it was shown that the presence of dopachrome tautomerase mediates an initial acceleration of melanogenesis since L-dopachrome is rapidly transformed to DHICA, but that melanin formation is inhibited because of the stability of this carboxylated indole compared to 5,6-dihydroxyindole (DHI), its decarboxylated counterpart obtained by spontaneous decarboxylation of L-dopachrome. Using L-dopa methyl ester as a precursor of melanogenesis, it is shown that this carboxylated indole does not polymerize in the absence of DHI, even in the presence of tyrosinase. However, it is incorporated into the polymer in the presence of both tyrosinase and DHI. Thus, this study suggests that DHI is essential for melanin formation, and the rate of polymerization depends on the ratio between DHICA and DHI in the medium. In the melanosome, this ratio should be regulated by the ratio between the activities of dopachrome tautomerase and tyrosinase.  相似文献   

17.
Multiplicity of platelet MAO was studied. Multiple forms of enzyme was separated with the use of 1.5% Tritone x-100 and, 1.3 M urea at pH 7.4 in, 0.01 M K-Na phosphate buffer. Three forms of MAO--MAO-I, MAO-II, and MAO-III were obtained under separation of solubilized proteins by affinity chromatography on AN-Sepharose 4B. Deprenyl in 10(-5) M inhibited all forms of the activity completely. Relation between multiple forms of the brain and platelet was discussed.  相似文献   

18.
It is shown that the composition of fixatives (2.5% glutaraldehyde in 0.1 M phosphate buffer with pH 7.2-7.4, and the mixture of 2.5% glutaraldehyde with 2% paraformaldehyde in 0.1 M phosphate buffer with pH 7.2-74) and duration of fixation (30 minutes, 24 hours, 7 days, and 30 days) under the room temperature exerts no influence on preservation of HeLa cultured cells. The ultrastructure of all the organelles of these cells is similar in any cases examined. All the membrane structures are well preserved; no condensation of chromatin is observed; the widths of the canals of endoplasmic reticulum, and of the intracristal and lateral spaces of mitochondria are invariable. Polysomes are present in the cytoplasm throughout the period of fixation.  相似文献   

19.
Lecithin-cholesterol acyltransferase was purified from rat plasma and the properties of this enzyme during the purification procedures and those of the purified enzyme were investigated in comparison with the human enzyme. The rat enzyme was not adsorbed on hydroxyapatite, which was employed for the purification of the human enzyme. When purified human enzyme was incubated at 37 degrees C in 0.1 mM phosphate buffer (pH 7.4; ionic strength, 0.00025), no alteration of enzyme activity was observed for up to 6 h. In the case of the rat enzyme, however, approximately 40% of the enzyme activity was lost under the same conditions. The human enzyme and rat enzyme were both retained on a Sepharose 4B column to which HDL3 was covalently linked, in 39 mM phosphate buffer, pH 7.4. Although the human enzyme was eluted from the column in 1 mM phosphate buffer, the rat enzyme was dissociated from the column at a lower buffer concentration (0.1 mM phosphate buffer). These findings indicate that the rat enzyme effectively associated with HDL3 in 39 mM phosphate buffer, pH 7.4, but the association was more sensitive to increase of ionic strength compared with that of the human enzyme.  相似文献   

20.
Nitric oxide (NO) and peroxynitrite (ONOO) are said to destroy norepinephrine (NE). We studied the role of NE decomposition by NO donors and ONOO as they affect the contractile activity of NE in rat denuded thoracic aorta. First, we determined the relaxing effect of NO donors (SNAP, PROLI/NO, Sodium nitrite, SIN-1) and ONOO after precontraction by NE (1 microM). SNAP and SIN-1 (EC(50) 50-110 nM) were more active than PROLI/NO, Sodium nitrite or ONOO (EC(50) 19-30 microM). The relaxing effect of NO donors and ONOO were decreased by ODQ (10 microM), a guanylate cyclase inhibitor. Second, we compared the contractile activity of NE before and after preincubation with NO donors or ONOO in presence of ODQ. NE (1 microM) was incubated with NO donors or ONOO at the concentrations of 0.1 mM in both Krebs solution or phosphate buffer (pH 7.4; 0.1 M) for 10 minutes at 37 degrees C. NE evoked the aorta contraction in the same concentrations before and after preincubation with NO donors. In contrast, ONOO decreased effect of NE, EC(50) was measured at 4.3+/-0.3 nM and 13.4+/-1.6 nM, before and after preincubation of NE with ONOO respectively. Third, we measured the NE concentration using the HPLC method. We revealed that the concentration of NE after preincubation with NO donors was unaltered. However HPLC measurement revealed that NE concentration after preincubation with ONOO was reduced 2-3-fold. Therefore, under these experimental conditions ONOO, but not NO donors, was capable of destroying NE.  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