首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
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   

2.
In this work, we examined photoreactivity of synthetic eumelanins, formed by autooxidation of DOPA, or enzymatic oxidation of 5,6‐dihydroxyindole‐2‐carboxylic acid and synthetic pheomelanins obtained by enzymatic oxidation of 5‐S‐cysteinyldopa or 1:1 mixture of DOPA and cysteine. Electron paramagnetic resonance oximetry and spin trapping were used to measure oxygen consumption and formation of superoxide anion induced by irradiation of melanin with blue light, and time‐resolved near‐infrared luminescence was employed to determine the photoformation of singlet oxygen between 300 and 600 nm. Both superoxide anion and singlet oxygen were photogenerated by the synthetic melanins albeit with different efficiency. At 450‐nm, quantum yield of singlet oxygen was very low (~10?4) but it strongly increased in the UV region. The melanins quenched singlet oxygen efficiently, indicating that photogeneration and quenching of singlet oxygen may play an important role in aerobic photochemistry of melanin pigments and could contribute to their photodegradation and photoaging.  相似文献   

3.
The human skin and eye melanin is commonly viewed as an efficient photoprotective agent. To elucidate the molecular mechanism of the melanin-dependent photoprotection, we studied the interaction of two synthetic melanins, dopa-melanin and cysteinyldopa-melanin, with a wide range of oxidising and reducing free radicals using the pulse radiolysis technique. We have found that although both types of free radicals could efficiently interact with the synthetic melanins, their radical scavenging properties depended, in a complex way, on the redox potential, the electric charge and the lifetime of the radicals. Repetitive pulsing experiments, in which the free radicals, probing the polymer redox sites, were generated from four different viologens, indicated that the eumelanin model had more reduced than oxidised groups accessible to reaction with the radicals. Although with many radicals studied, melanin interacted via simple one-electron transfer processes, the reaction of both melanins with the strongly oxidising peroxyl radical from carbon tetrachloride, involved radical addition. Our study suggests that the free radical scavenging properties of melanin may be important in the protection of melanotic cells against free radical damage, particularly if the reactive radicals are generated in close proximity to the pigment granules.  相似文献   

4.
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   

5.
Although pigment melanin has long been though of as "inert," recent work has attested to its chemical reactivity. In this communication, we report that either commercial synthetic melanin prepared by persulfate oxidation of tyrosine ("Sigma melanin") or sepia melanin extracted from cuttlefish markedly accelerates the in vitro oxygenation of p-hydroxyanisole (MMEH), catalyzed by mushroom or B-16 melanoma tyrosinase. Kinetics of 4-methoxy-1,2-benzoquinone formation (lambda max = 413 nm) or of molecular O2 uptake were biphasic, with an initial slow rate ("lag time") followed by a fast linear increase. The biphasic response reflects an initial slow hydroxylation followed by a fast dehydrogenation. Added melanin markedly decreased the lag time but had little effect on subsequent dehydrogenation. Similar effects were observed for tyrosine itself. A complex between MMEH and melanin appears to be the "active" species in these reactions. The results indicate that melanin acts as an electron conduit, which accepts electrons from the substrate and transfers them to tyrosinase. The magnitude of the effect depends on the type of melanin as well as on its oxidation state. Kinetic analysis indicates that both melanins are very efficient at transferring electron to tyrosinase, and that Sigma melanin is roughly threefold more efficient than sepia melanin. The qualitative similarity of reaction between the synthetic and "natural" melanins suggests that the former may serve as a first approximation to the in vivo situation. On the other hand, the observed quantitative differences and the sensitivity of these results to the chemical state of melanin suggests that this methodology might eventually be adapted as a non-destructive probe of melanin in situ.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

6.
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   

7.
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   

8.
Formation of light-induced free radicals from natural eumelanin (from bovine eyes) and synthetic melanin (from oxidation of 3,4-dihydroxyphenylalanine) has been studied by electron spin resonance spectroscopy. Action spectra measured for natural melanins are very similar to that found for synthetic melanin, and are unaffected by the removal of associated protein. A comparison of action spectra with optical absorbance spectra shows that the former has a more marked wavelength dependence, suggesting that the chromophore that is most active in free-radical production is not the major melanin chromophore that absorbs visible light. Measurements of quantum yields for freeradical production have been made over a wavelength range from 600 to 230 nm. The efficiency of radical production from natural eumelanin is about three times greater than from the synthetic material. Although production of the melanin radicals detected is independent of oxygen, some correlation with oxygen consumption is evident; quantum yields for radical production are approximately three times those for oxygen consumption obtained under similar conditions. Possible reasons for this are discussed.  相似文献   

9.
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   

10.
Comparative studies of fungal melanin and two preparations of the high-molecular-weight humin-like substances formed during a solid-phase cultivation of the basidiomycete Cerrena maxima 0275 for 45 and 70 days were performed. The fungal melanin from Aspergillus niger and the humin-like substances synthesized by the basidiomycete C. maxima 0275 are similar in their physicochemical properties (elemental composition and behavior in acids and alkalis) and auxin-like activities. However, these biopolymers differ, essentially, at the structural level. According to IR spectroscopy data, the obtained humin-like substances display a higher similarity to natural humic acids and are more diverse in their functional groups compared with fungal melanins. Presumably, this is connected with the fact that laccase is involved in formation of humin-like substances; moreover, this enzyme is involved not only in the synthesis of these polymers, but also in their modification and degradation.  相似文献   

11.
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   

12.
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   

13.
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   

14.
Ceriporiopsis sp. strain MD-1, isolated from forest soil, produced several extracellular enzymes that decolorized human hair melanin. Among them, three enzymes (E1, E2-1, and E2-2) were purified to homogeneity and characterized. The enzymes required hydrogen peroxide in their enzyme reactions and, typical of other fungal peroxidases, oxidized various phenol compounds such as guaiacol, but not 3,4-dimethoxybenzyl alcohol. The spectra of the three enzymes showed an absorption maximum at 406 nm, indicating that they were heme proteins. However, the A406/A280 values of the enzymes were below 0.4, which was lower than those of other peroxidases. E2-1 and E2-2 were similar to each other in their molecular and catalytic properties, and they possibly represent products of posttranslational modifications and/or allelic variants of the same gene, mdcA. The corresponding cDNA was cloned and sequenced; the deduced amino acid sequence showed high identities to the manganese peroxidases from other microorganisms. The specific activities and Km values of E2-1 and E2-2 for synthetic and human hair melanins were much higher than those of Phanerochaete chrysosporium manganese peroxidase and lignin peroxidase.  相似文献   

15.
Melanins are the most prevalent pigments in animals and are involved in visual communication by producing colored traits that often evolve as intraspecific signals of quality. Identifying and quantifying melanins are therefore essential to understand the function and evolution of melanin‐based signals. However, the analysis of melanins is difficult due to their insolubility and the lack of simple methods that allow the identification of their chemical forms. We recently proposed the use of Raman spectroscopy as a simple, noninvasive technique that can be used to identify and quantify melanins in feathers and hairs. Contrarily, other authors later stated that melanins are characterized by a lack of defined Raman signals. Here, we use confocal Raman microscopy to confirm previous analyses showing that the two main chemical forms of melanins (eumelanin and pheomelanin) exhibit distinct Raman signal and compare different excitation wavelengths to analyze synthetic pheomelanin and natural melanins in feathers of different species of birds. Our analyses indicate that only laser excitation wavelengths below 1064 nm are useful for the analysis of melanins by Raman spectroscopy, and only 780‐nm laser in the case of melanins in feathers. These findings show that the capacity of Raman spectroscopy to distinguish different chemical forms of melanins depends on laser power and integration time. As a consequence, Raman spectroscopy should be applied after preliminar analyses using a range of these parameters, especially in fragile biological tissues such as feathers.  相似文献   

16.
Comparative studies of fungal melanin and two preparations of the high-molecular-weight humin-like substances formed during a solid-phase cultivation of the basidiomycete Cerrena maxima 0275 for 45 and 70 days were performed. The fungal melanin from Aspergillus niger and the humin-like substances synthesized by the basidiomycete C. maxima 0275 are similar in their physicochemical properties (elemental composition and behavior in acids and alkalis) and auxin-like activities. However, these biopolymers differ, essentially, at the structural level. According to IR spectroscopy data, the obtained humin-like substances display a higher similarity to natural humic acids and are more diverse in their functional groups compared with fungal melanins. Presumably, this is connected with the fact that laccase is involved in formation of humin-like substances; moreover, this enzyme is involved not only in the synthesis of these polymers, but also in their modification and degradation.  相似文献   

17.
The fungusInonotus obliquus (Pers.) Pil. synthesized high-molecular-weight phenolic pigments that were assigned to melanins according to their physicochemical properties. It was shown that copper ions (0.008%), pyrocatechol (1.0 mM), and tyrosine (20.0 mM) stimulated melanogenesis. The production of melanin correlated with the synthesis ofo- andp-diphenoloxidases. The fungal melanin had strong antioxidant and genoprotective effects.  相似文献   

18.
Electron Spin Resonance Studies on Melanin   总被引:2,自引:0,他引:2       下载免费PDF全文
Electron spin resonance (e.s.r.) observations of squid melanin have been conducted over the temperature range 500°K to 4.2°K, and the effect of various chemical treatments of the melanin upon the e.s.r. spectrum has been studied. The findings have shown that the paramagnetism of this melanin follows the Curie Law from 500°K to 4.2°K, that the spin signal can be eliminated by the addition of Cu++ to the melanin, and that the optical and e.s.r. absorptions of melanin are independent since either can be reduced or eliminated without affecting the other. Similar studies on synthetic melanins produced by autoxidation or by enzymatic oxidation of a number of biphenols were carried out. It was found that the e.s.r. signals of these synthetic melanins were strikingly similar (with respect to line width, line shape, and g-value) with those of squid melanin. It is concluded that the unpaired electrons observed are associated with trapped free radicals in the melanin polymer, that the biosynthesis of melanin may involve a free radical mechanism, and that these physical data are in accord with the concept of Nicolaus that melanin is a highly irregular, three-dimensional, polymer.  相似文献   

19.
Melanin complex of the fungus Inonotus obliquus   总被引:6,自引:0,他引:6  
The fungus Inonotus obliquus (Pers.) Pil. synthesised high-molecular-weight phenolic pigments that were assigned to melanins according to their physicochemical properties. It was showed that copper ions (0.008%), pyrocatechol (1.0 mM), and tyrosine (20.0 mM) stimulated the melanogenesis. The production of melanin correlated with the synthesis of o- and p-diphenoloxidases. The fungal melanin had strong antioxidant and genoprotective effects.  相似文献   

20.
Mammalian melanins exist in two chemically distinct forms: the brown to black eumelanins and the yellow to reddish-brown pheomelanins. They can be quantified by HPLC analysis of pyrrole-2,3,5-tricarboxylic acid (PTCA) and aminohydroxyphenylalanine (AHP). We recently developed a spectrophotometric method for assaying the total amount of eu- and pheomelanins by dissolving melanins in Soluene-350 plus water. In this study, we examined whether absorbance at 500 nm (A500) of the Soluene-350 solution reflects the total amount of melanins obtained by the HPLC methods, and whether the ratio of absorbances between 650 and 500 nm reflects the eumelanin/total melanin ratio in mouse hair, sheep wool, and human hair. Our findings were as follows: (1) Total melanin levels calculated from A500 values correlate well with those obtained from PTCA and AHP values by multiplying with the following factors: for mice, PTCA × 45 + AHP × 2.5; for sheep, PTCA × 40 + AHP × 15; and for humans, PTCA × 160 + AHP × 10. (2) The A650/A500 ratios were higher (0.25–0.33) in black to brown hair while they were significantly lower (0.10–0.14) in yellow to red hair. These results indicate that (1) the A500 value can be used to quantify the total combined amount of eu- and pheomelanins, and (2) the A650/A500 ratio can serve as a parameter to estimate the eumelanin/total melanin ratio. The present method provides a convenient way to qualitatively characterize eu- and pheomelanins in melanins produced in follicular melanocytes.  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