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 109 毫秒
1.
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   

2.
The absorption spectra of melanins isolated from some black ascomycetes, as well as of synthetic melanin and natural melanin from Sepia officinalis, were recorded in the long-wavelength ultraviolet region A (320 nm < lambda < 400 nm) and in the blue-violet region of the electromagnetic spectrum at illumination intensities varying from 0.02 to 1 mW/cm2. The photochemical properties of fungal melanins were found to be dependent on both the producing strain and the conditions of its cultivation. The fungal melanins are more susceptible to photomodification and more biologically active than the synthetic melanin, indicating that these properties may be related. The data obtained suggest that the fungal melanins susceptible to photomodification possess higher biological activity than commercial melanins.  相似文献   

3.
The infrared and electron spin resonance spectra of synthetic 3,4-dihydroxyphenylalanine (DOPA) and tyrosine melanins and chemically modified melanin samples were determined, and it was shown that unmodified and reduced DOPA melanins exhibited similar ir spectra. Oxidized DOPA melanins showed a higher number of carboxy groups in the sample. A significant increase of free radical content in reduced DOPA melanin and a decrease of free radical content in oxidized DOPA melanin in comparison to unmodified samples were demonstrated by the use of ESR methodology. Methylation of tyrosine melanin with an excess of diazomethane gave very rich ir spectra as compared to melanins methylated with methanol saturated by gaseous HCl. In tyrosine melanin samples the esterification of carboxy groups with methanol caused a decrease in the free radical content. When diazomethane was used, the methylated melanin samples had free radical levels reduced to only about 4% of the total observed for unmodified tyrosine melanin.  相似文献   

4.
The protective role of melanin, either synthetic or derived from a metastatic lung melanoma nodule, was studied in terms of its ability to interact with active oxygen species (O2, H2O2, RO, ROO, etc.). Both melanins showed the ability to react with O2. The superoxide dismutase-like activity corresponds to 21 and 10 U/mg for synthetic and tumor melanin, respectively. The latter value accounts for about 8% of the superoxide dismutase activity of cultured melanoma cells. Neither type of melanin showed catalase-like or glutathione peroxidase-like activity. Both types of melanin reacted with RO and ROO radicals as determined by inhibition of the lipid peroxidation reaction of rat liver homogenates. The spontaneous lipid peroxidation of rat liver homogenate was inhibited up to 90% and 80% by synthetic and tumor melanin with half-maximal effects at 2.5 and 5.5 μg melanin/ml, respectively. The 2,2-azobis-(2 amidino propane) (AAPH)-initiated lipid peroxidation of rat liver homogenate was inhibited up to 3% and 20% by synthetic and tumor melanin, with half maximal effect at 120 and 500 μg melanin/ml, respectively. Both types of melanin were able to protect the in vitro inactivation of glucose oxidase, which occurs in the presence of AAPH-generated radicals.  相似文献   

5.
Melanins are widely used in medicine, pharmacology, cosmetics and other fields. Although several technologies for the purification of water-insoluble dioxyphenylalanine (DOPA) melanins have been described, a source of water-soluble melanin is highly desirable. Here we describe an effective procedure for the isolation and purification of water-soluble melanin using the culture medium of Bacillus thuringiensis subsp. galleriae strain K1. Water-soluble melanin from this organism has an isoelectric point (pI=3.0-3.2) and was purified optimally by adsorbtion using the IA-1r resin and elution as a concentrated solution. The purified melanin obtained exhibited a similar infra-red absorbtion spectrum to synthetic melanin and contained quinolic and phenolic structures and an amino acid content of around 20% after acid hydrolysis. The molecular weight of the purified melanin determined by SDS-PAGE was 4 kDa and the electromagnetic spin resonance spectrum of the purified microbial melanin was a slightly asymmetric singlet without hyperfine structure with about 7 Gauss width of the line between points of the maximum incline and g=2.006. The concentration of paramagnetic centers in melanin is 0.21x10(18) spin/g. The results obtained provide a rapid, simple and inexpensive method for the large scale purification of water soluble melanin that may have widespread applications.  相似文献   

6.
Using scanning tunneling microscopy (STM), we have imaged two types of mildly-bleached, synthetic tyrosine-derived melanins for comparison with the unbleached melanin from which they were prepared. These mildly-bleached melanins were generated by mild oxidation of the unbleached melanin, using either basic hydrogen peroxide or air/light. The unbleached melanin, and two mildly-bleached melanins, were independently deposited from very dilute tetrahydrofuran (THF) solutions onto highly oriented pyrolytic graphite (HOPG) substrate for STM imaging. Lateral dimensions (23 A, average of two directions) of structures from each of the three samples showed no differences. However, structures from both mildly-bleached melanins showed similar dramatic decreases (from approximately 15 A to approximately 5 A) in their STM-measured apparent heights, compared with structures from the unbleached melanin sample. These STM observations are compatible with structural models for unbleached and mildly-bleached melanins, incorporating a three-dimensional structure for unbleached melanin composed of multi-layered, pi-pi-stacked, carboxylic and amino variants of polyaromatic polymeric sheets. The STM-observed decrease in apparent heights after mild oxidation, which we associate with a change in stack height, has been confirmed by experiments using tapping mode atomic force microscopy (TM-AFM) for the unbleached and mildly-hydrogen-peroXide-bleached melanins (from approximately 14 A to approximately 6 A). In these TM-AFM experiments, the melanins were deposited directly onto magnesium cation-treated glass substrates in contact with methanolic solutions of each of the melanins. We interpret our mild-bleaching results as an oxidative conversion of the multi-layered, stacked sheets of mainly carboxylic and amino variants of polyquinhydrone-like moieties, to largely de-stacked, mildly-bleached melanin sheets. These oxidized and, hence, electron-deficient sheets should not readily form multi-layered, pi-pi interacting stacks, but instead appear to be either single-layer polyquinone sheets or, at most, double-layer polyquinhydrone sheets. The effects of such de-stacking on in vivo melanin photoprotection, and structural similarities between melanin derived from natural sources and the synthetic melanin samples used in this work are discussed.  相似文献   

7.
Fungal melanins differ in planar stacking distances   总被引:1,自引:0,他引:1  
Melanins are notoriously difficult to study because they are amorphous, insoluble and often associated with other biological materials. Consequently, there is a dearth of structural techniques to study this enigmatic pigment. Current models of melanin structure envision the stacking of planar structures. X ray diffraction has historically been used to deduce stacking parameters. In this study we used X ray diffraction to analyze melanins derived from Cryptococcus neoformans, Aspergillus niger, Wangiella dermatitides and Coprinus comatus. Analysis of melanin in melanized C. neoformans encapsulated cells was precluded by the fortuitous finding that the capsular polysaccharide had a diffraction spectrum that was similar to that of isolated melanin. The capsular polysaccharide spectrum was dominated by a broad non-Bragg feature consistent with origin from a repeating structural motif that may arise from inter-molecular interactions and/or possibly gel organization. Hence, we isolated melanin from each fungal species and compared diffraction parameters. The results show that the inferred stacking distances of fungal melanins differ from that reported for synthetic melanin and neuromelanin, occupying intermediate position between these other melanins. These results suggest that all melanins have a fundamental diffracting unit composed of planar graphitic assemblies that can differ in stacking distance. The stacking peak appears to be a distinguishing universal feature of melanins that may be of use in characterizing these enigmatic pigments.  相似文献   

8.
Eumelanin is photoprotective while pheomelanin is phototoxic to pigmented tissues. Ultraviolet A (UVA)-induced tanning seems to result from the photooxidation of pre-existing melanin and contributes no photoprotection. However, data available for melanin biodegradation remain limited. In this study, we first examined photodegradation of eumelanin and pheomelanin in human black hairs and found that the ratio of Free (formed by peroxidation in situ) to Total (after hydrogen peroxide oxidation) pyrrole-2,3,5-tricarboxylic acid (PTCA) increases with hair aging, indicating fission of the dihydroxyindole moiety. In red hair, the ratio of thiazole-2,4,5-tricarboxylic acid (TTCA) to 4-amino-3-hydroxyphenylalanine (4-AHP) increases with aging, indicating the conversion from benzothiazine to benzothiazole moiety. These photodegradation of melanins were confirmed by UVA (not UVB) irradiation of melanins from mice and human hairs and synthetic eumelanin and pheomelanin. These results show that both eumelanin and pheomelanin degrade by UVA and that Free/Total PTCA and TTCA/4-AHP ratios serve as sensitive indicators of photodegradation.  相似文献   

9.
Individuals with black skin rarely get skin cancer, and melanomas, tumors arising from pigmented cells, are generally resistant to radiation therapy. The role of melanin in these two phenomena has not been defined, but oxygen-radical species have been implicated in both effects. These studies were undertaken to determine the ability of various melanins to compete for ionizing radiation-produced radicals which destroy nucleic acid bases. The ability of Sigma eumelanin (S-eumelanin) to protect against the radiolysis of thymidine in buffered solutions was compared to the protective ability of seven amino acids, including melanin precursors; bovine serum albumin, as a model protein; ficoll, as a model polysaccharide; and DNA. Both proteins and polysaccharides are known to scavenge hydroxyl radicals in cells. The concentration of thymidine after exposure to gamma radiation was determined by High Performance Liquid Chromatography (HPLC) analysis after removal of insoluble melanin by acid precipitation. S-eumelanin was more effective at competing with thymidine for free radicals than bovine serum albumin, Ficoll, or DNA, but less effective than certain of the small molecules. Several of the above compounds were also examined for ability to protect against thymine radiolysis. In addition, melanins from other sources were compared to S-eumelanin. Of these, enzymatically synthesized phaeomelanin was the most effective. The results indicate that melanins can compete for base- and nucleoside-damaging free radicals more effectively than other cellular macromolecules. Of the small molecules, the phenolic compounds had the greatest scavenging ability. In vivo, melanins are found in melanosomes bound to protein. Therefore, the relevance of these findings to the photo- and radiobiology of melanins in vivo has yet to be determined.  相似文献   

10.
Melanin standard method: particle description.   总被引:1,自引:0,他引:1  
Melanin isolated from the ink sac of Sepia officinalis (Sepia melanin) has been proposed as a standard for natural eumelanin. There are no standard methods for the isolation, purification, and storage of melanins. Mild methods designed to preserve the native composition and structure of melanin are needed. The specific aim of the present work, using Sepia melanin, was to develop a mild and generally applicable protocol for the isolation and purification of melanins. It is well established that melanin polymers contain a large number of free carboxylic acid residues. These anionic residues are responsible for the cation exchange properties observed for melanins. Heating melanins with hydrochloric acid at reflux has been demonstrated to lead to extensive decarboxylation. Indeed, heat alone has been shown to cause decarboxylation, and care must be exercised to avoid such conditions. By analogy with cation exchange resins, melanins should be isolated and named according to the associated counterion (e.g., Sepia melanin--K+ form). The method reported here avoided extremes in pH and temperature, and was designed to yield melanin in the K+ form. Physical disaggregation of particulate melanin using a wet milling step was also found to facilitate removal of significant quantities of adsorbed protein. The following physical parameters were used to monitor the purification and to characterize the resultant melanin: pH, conductance, particle size, and diffuse reflectance spectroscopy.  相似文献   

11.
Peroxynitrite-mediated linoleic acid oxidation and tyrosine nitration were analysed in the presence of synthetic model neuromelanins: dopamine (DA) -melanin, cysteinyldopamine (CysDA) -melanin and various DA/CysDA copolymers. The presence of melanin significantly decreased the amount of 3-nitrotyrosine formed. This inhibitory effect depended on the type and concentration of melanin polymer. It was found that incorporation of CysDA-derived units into melanin attenuated its protective effect on tyrosine nitration induced by peroxynitrite. In the presence of bicarbonate, the melanins also inhibited 3-nitrotyrosine formation in a concentration dependent manner, although the extent of inhibition was lower than in the absence of bicarbonate. The tested melanins inhibited peroxynitrite-induced formation of linoleic acid hydroperoxides, both in the absence and in the presence of bicarbonate. In the presence of bicarbonate, among the oxidation products appeared 4-hydroxynonenal (HNE). CysDA-melanin inhibited the formation of HNE, while DA-melanin did not affect the aldehyde level. The results of the presented study suggest that neuromelanin can act as a natural scavenger of peroxynitrite.  相似文献   

12.
Melanin isolated from the ink sac of Sepia officinalis (Sepia melanin) has been proposed as a standard for natural eumelanin. There are no standard methods for the isolation, purification, and storage of melanins. Mild methods designed to preserve the native composition and structure of melanin are needed. The specific aim of the present work, using Sepia melanin, was to develop a mild and generally applicable protocol for the isolation and purification of melanins. It is well established that melanin polymers contain a large number of free carboxylic acid residues. These anionic residues are responsible for the cation exchange properties observed for melanins. Heating melanins with hydrochloric acid at reflux has been demonstrated to lead to extensive decarboxylation. Indeed, heat alone has been shown to cause decarboxylation, and care must be exercised to avoid such conditions. By analogy with cation exchange resins, melanins should be isolated and named according to the associated counterion (e.g., Sepia melanin—K+ form). The method reported here avoided extremes in pH and temperature, and was designed to yield melanin in the K+ form. Physical disaggregation of particulate melanin using a wet milling step was also found to facilitate removal of significant quantities of adsorbed protein. The following physical parameters were used to monitor the purification and to characterize the resultant melanin: pH, conductance, particle size, and diffuse reflectance spectroscopy.  相似文献   

13.
Turkovskii  I. I.  Yurlova  N. A. 《Microbiology》2002,71(4):410-416
The absorption spectra of melanins isolated from some black ascomycetes, as well as of synthetic melanin and natural melanin from Sepia officinalis, were recorded in the long-wavelength ultraviolet region A (320 nm < < 400 nm) and in the blue–violet region of the electromagnetic spectrum at illumination intensities varying from 0.02 to 1 mW/cm2. The photochemical properties of fungal melanins were found to be dependent on both the producing strain and the conditions of its cultivation. The fungal melanins are more susceptible to photomodification and more biologically active than the synthetic melanin, indicating that these properties may be related. The data obtained suggest that the fungal melanins susceptible to photomodification possess higher biological activity than commercial melanins.  相似文献   

14.
Using scanning tunneling microscopy (STM), we have imaged two types of mildly‐bleached, synthetic tyrosine‐derived melanins for comparison with the unbleached melanin from which they were prepared. These mildly‐bleached melanins were generated by mild oxidation of the unbleached melanin, using either basic hydrogen peroxide or air/light. The unbleached melanin, and two mildly‐bleached melanins, were independently deposited from very dilute tetrahydrofuran (THF) solutions onto highly oriented pyrolytic graphite (HOPG) substrate for STM imaging. Lateral dimensions (23 Å, average of two directions) of structures from each of the three samples showed no differences. However, structures from both mildly‐bleached melanins showed similar dramatic decreases (from ~15 Å to ~5 Å) in their STM‐measured apparent heights, compared with structures from the unbleached melanin sample. These STM observations are compatible with structural models for unbleached and mildly‐bleached melanins, incorporating a three‐dimensional structure for unbleached melanin composed of multi‐layered, Π–Π‐stacked, carboxylic and amino variants of polyaromatic polymeric sheets. The STM‐observed decrease in apparent heights after mild oxidation, which we associate with a change in stack height, has been confirmed by experiments using tapping mode atomic force microscopy (TM‐AFM) for the unbleached and mildly‐hydrogen‐peroxide‐bleached melanins (from ~14 Å to ~6 Å). In these TM‐AFM experiments, the melanins were deposited directly onto magnesium cation‐treated glass substrates in contact with methanolic solutions of each of the melanins. We interpret our mild‐bleaching results as an oxidative conversion of the multi‐layered, stacked sheets of mainly carboxylic and amino variants of polyquinhydrone‐like moieties, to largely de‐stacked, mildly‐bleached melanin sheets. These oxidized and, hence, electron‐deficient sheets should not readily form multi‐layered, Π–Π interacting stacks, but instead appear to be either single‐layer polyquinone sheets or, at most, double‐layer polyquinhydrone sheets. The effects of such de‐stacking on in vivo melanin photoprotection, and structural similarities between melanin derived from natural sources and the synthetic melanin samples used in this work are discussed.  相似文献   

15.
Melanins are widely used in medicine, pharmacology, cosmetics and other fields. Although several technologies for the purification of water‐insoluble dioxyphenylalanine (DOPA) melanins have been described, a source of water‐soluble melanin is highly desirable. Here we describe an effective procedure for the isolation and purification of water‐soluble melanin using the culture medium of Bacillus thuringiensis subsp. galleriae strain K1. Water‐soluble melanin from this organism has an isoelectric point (pI = 3.0–3.2) and was purified optimally by adsorbtion using the IA‐1r resin and elution as a concentrated solution. The purified melanin obtained exhibited a similar infra‐red absorbtion spectrum to synthetic melanin and contained quinolic and phenolic structures and an amino acid content of around 20% after acid hydrolysis. The molecular weight of the purified melanin determined by SDS‐PAGE was 4 kDa and the electromagnetic spin resonance spectrum of the purified microbial melanin was a slightly asymmetric singlet without hyperfine structure with about 7 Gauss width of the line between points of the maximum incline and g = 2.006. The concentration of paramagnetic centers in melanin is 0.21 × 1018 spin/g. The results obtained provide a rapid, simple and inexpensive method for the large scale purification of water soluble melanin that may have widespread applications.  相似文献   

16.
Scavenging of superoxide radicals by melanin is a possible factor in the photoprotection afforded by melanin pigments. The reaction between superoxide anions and melanins has been studied by electron spin resonance and spin trapping methods. It was found that superoxide anions react to produce melanin free radicals in a reaction inhibited by superoxide dismutase but not by catalase. The rate of radical formation depends on the concentration of melanin and superoxide, the pH of the medium and the presence of diamagnetic metal ions. The melanin pigment competes with the enzyme superoxide dismutase for removal of superoxide radicals. It was found that the xanthine-xanthine oxidase system is not suitable for studying the reaction of superoxide with melanin, as the enzymatic activity of xanthine oxidase is considerably inhibited by melanin.  相似文献   

17.
A number of purified natural and synthetic melanins have been examined by X-ray diffraction. A consistent finding with all samples was the lack of structure in the diffraction pattern corresponding to any significant crystallinity in these melanin preparations. A diffuse ring, centered at a Bragg spacing of 3.4 A was consistently found in samples of melanin from animal sources, and a similar ring at 4.2 A in all melanins obtained from plants. Models for these two polymer types, based upon the current concept that they primarily involve indole and catechol monomeric units respectively, were then evaluated by a Monte Carlo method. From the comparison of the observed spacings with the calculated ones it was concluded that the 4.2 A spacing in the catechol melanins is probably related to the average interaction between adjacent monomeric units, with mutually random orientations. The 3.4 A spacing observed in indole melanins appears to derive from the tendency of indole monomers (probably of adjacent chains) tending to aggregate in near parallel stacks. Some randomness in the form of translations and rotations parallel to the planar groups is consistent with the diffraction patterns. An interesting finding was that the diffraction pattern of synthetic melanin prepared by the alkaline auto-oxidation of catechol gave the 3.4 A spacing found in the indole melanins of natural origin.  相似文献   

18.
Electron spin lattice relaxation times (T1) and the phase memory times (Tm) were obtained for the synthetic melanin system from 3-hydroxytyrosine (dopa) by means of electron spin echo spectroscopy at 77 degrees K. Saturation behavior of the ESR spectra of melanins in melanin-containing tissue and of the synthetic melanin was also determined at the same temperature. The spin lattice relaxation time and the spectral diffusion time of the synthetic melanin are very long (4.3 ms and 101 microseconds, respectively, in the solid state), and the ESR signal saturates readily at low microwave powers. On the other hand, ESR spectra of natural melanins from the tissues chosen for this study, as well as those of synthetic melanins which contain Fe3+ of g = 4.3 and Mn2+ of g = 2, are relatively difficult to saturate compared with samples without such metal ions. These results show clearly that a large part of those two metal ions in sites responsible for the ESR spectral components with these particular g values are coordinated to melanin in melanin-containing tissue, and modify the magnetic relaxation behavior of the melanin. Accumulations of these metal ions in melanins are different from system to system, and they increase in the order: hair (black), retina and choroid (brown), malignant melanoma of eye and skin, and lentigo and nevus of skin.  相似文献   

19.
Microsclerotia of three melanin-deficient mutants of Verticillium dahliae formed malanin from (+)-scytalone, 1,8-dihydroxynaphthalene, catechol, and L-3,4-dihydroxyphenylalanine. The melanins formed from (+)-scytalone or 1,8-dihydroxynaphthalene resembled wild-type melanin chemically and ultrastructurally, whereas the melanins formed from catechol and L-3,4-dihydroxyphenlalanine were different. This suggests that scytalone and 1,8-dihydroxynaphthalene but no catechol or L-3,4-dihydroxyphenylalanine are natural intermediates of melanin biosynthesis in V. dahliae.  相似文献   

20.
Natural and synthetic melanins catalyze the conversion of glycine to glyoxylate and formic acid in vitro. The conversion depends upon the concentration of melanin.  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