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 312 毫秒
1.
The hepta-, hexa- and penta-carboxylic porphyrins found in the faeces of rats poisoned with hexachlorobenzene have been separated by high-pressure liquid chromatography and characterized largely by spectroscopie methods. Their structures were confirmed by total synthesis, as part of a programme in which eleven of the fourteen hepta-, hexa- and penta-carboxylic porphyrins derived from uroporphyrin III have now been synthesized as their methyl esters. The four isomeric heptacarboxylic and three of the pentacarboxylic porphyrinogens have been incubated with haemolysates of chicken erythrocytes, and they are all converted into protoporphyrin IX but at different rates. On the basis of this and other evidence we conclude that the decarboxylation of uroporphyrinogen III to coproporphyrinogen III is a stepwise process taking place by a preferred pathway (both in normal and abnormal metabolism); the acetic acid groups are decarboxylated in a sequential clockwise fashion starting with that on the D ring and followed by those on the A, B and C rings. In the poisoned rats the uroporphyrinogen decarboxylase enzyme (or group of enzymes) is probably partially inhibited and the pentacarboxylic porphyrinogen with an acetic acid group on ring C accumulates. The latter is then transformed by a side pathway into dehydroisocoproporphyrinogen and thence into dehydroisocoproporphyrin and its congeners.  相似文献   

2.
F Li  C K Lim    T J Peters 《The Biochemical journal》1987,243(2):621-623
A reversed-phase h.p.l.c. system is described for the separation of all five naturally occurring pentacarboxylic porphyrinogen isomers. The compounds are detected electrochemically with high sensitivity. The peaks are positively identified by h.p.l.c. analysis of the pentacarboxylic porphyrinogens from reduction of pentacarboxylic porphyrins prepared by partial decarboxylation of hexa- and hepta-carboxylic porphyrin III of known structures. The resolution of pentacarboxylic porphyrinogens is superior to that of the porphyrins and the method is applicable to the small-scale preparative isolation of pure isomers.  相似文献   

3.
The action of porphyrins, uroporphyrin I and III (URO I and URO III), pentacarboxylic porphyrin I (PENTA I), coproporphyrin I and III (COPRO I and COPRO III), protoporphyrin IX (PROTO IX) and mesoporphyrin (MESO), on the activity of human erythrocytes delta-aminolevulinic acid dehydratase, porphobilinogenase, deaminase and uroporphyrinogen decarboxylase in the dark and under UV light was investigated. Both photoinactivation and light-independent inactivation was found in all four enzymes using URO I as sensitizer. URO III had a similar action as URO I on porphobilinogenase and deaminase and PROTO IX exerted equal effect as URO I on delta-aminolevulinic acid dehydratase and uroporphyrinogen decarboxylase. Photodynamic efficiency of the porphyrins was dependent on their molecular structure. Selective photodecomposition of enzymes by URO I, greater specificity of tumor uptake by URO I and enhanced porphyrin synthesis by tumors from delta-aminolevulic acid, with predominant formation of URO I, underline the possibility of using URO I in detection of malignant cells and photodynamic therapy.  相似文献   

4.
High performance liquid chromatography (HPLC) has been used to separate porphyrin methyl esters isolated from porphyric patients. Fecal porphyrins were esterified directly with boron trifluoride: urinary porphyrins were absorbed on talcum and then treated with boron trifluoride. HPLC permits rapid, efficient, and quantitative detection and identification of porphyrins in complex natural mixtures.  相似文献   

5.
Sixteen porphyrins, including neutral, anionic and cationic meso-(aryl)porphyrins and meso-(1-methyl-4-pyridinium)porphyrins were herein evaluated in terms of their photosensitizing properties against HaCaT keratinocytes. After an initial screening, the cationic porphyrins were studied in more details, by both determining their log POW and performing PDT assays in lower porphyrin concentrations. Porphyrins presenting two or more adjacent positively charged groups, directly linked to the macrocycle meso positions, appeared to be the most effective photosensitizers. The present study also included the dicationic 5,10-diphenyl-15,20-di(1-methylpyridinium-4-yl)porphyrin (14b), which has previously shown promising results on a psoriasis-like in vivo model. Overall results indicated that the beneficial effect related to porphyrins on psoriasis can be related to the decreasing of keratinocyte viability. Furthermore, some of the cationic porphyrins studied appeared as candidates to be utilized as photosensitizers for psoriasis treatment.  相似文献   

6.
Porphyrin methyl esters and the isomers of uroporphyrin and heptacarboxylic porphyrin were separated by high-performance liquid chromatography. Isocoproporphyrin was also separated from coproporphyrin. By slight modifications to the solvent mixture, the separation of all biological polycarboxylic porphyrins was achieved. These separations were made possible through the high efficiency of 10- or 5-μm particle-size Radial-PAK cartridges, which have been used in the separation of porphyrins in various excreta and tissues in a number of porphyrias.  相似文献   

7.
The separation and quantitation of plasma free acid porphyrins by high-pressure liquid chromatography and fluorescence is described. Porphyrins were extracted from plasma in a simple manner with a recovery >90%. They were separated by high-pressure liquid chromatography on a silica gel (10 μm) column, using a gradient of acetone:dilute acetic acid. Resolution of seven free acid porphyrin standards including coproporphyrins I and III, but not uroporphyrins I and III, was achieved in 12 min at picomolar concentrations. Plasma of patients with erythropoietic protoporphyria displayed protoporphyrin. Uroporphyrin was the only porphyrin found in plasma of eight patients with porphyria cutanea tarda. Normal plasma contained small amounts of uroporphyrin and/or traces of protoporphyrin.  相似文献   

8.
A diploid yeast strain D61.M was used to study induction of mitotic chromosomal malsegregation, mitotic recombination and point mutation. Several ketones (including acetone and methyl ethyl ketone) and some organic acid esters (including the methyl, ethyl and 2-methoxyethyl esters of acetic acid) and acetonitrile strongly induced aneuploidy but not recombination or point mutation. Only diethyl ketone induced low levels of recombination and point mutation in addition to aneuploidy. Related compounds were weak inducers of aneuploidy: methyl n-propyl ketone, the methyl esters of propionic and butyric acid, acetic acid esters of n- and iso-propanol and ethyl propionate. No mutagenicity was found with n-butyl and isoamyl acetate, ethyl formate, acetyl acetone (2,5-dipentanone) and dioxane. Methyl isopropyl ketone induced only some recombination and point mutation but no aneuploidy. Efficient induction was only observed with a treatment protocol in which growing cells were exposed to the chemicals during a growth period of 4 h at 28 degrees C followed by incubation in ice for more than 90 min, usually overnight for 16-17 h. Aneuploid cells could be detected in such cultures during a subsequent incubation at growth temperature if the chemical was still present. Detailed analysis showed that there was a high incidence of multiple events of chromosomal malsegregation. It is proposed that the mutagenic agents act directly on tubulin during growth so that labile microtubules are formed which dissociate in the cold. When cells are brought back to temperatures above the level critical for reassembly of tubulin and allowed to grow, faulty microtubules are formed.  相似文献   

9.
Uroporphyrinogen decarboxylase (EC 4.1.1.37) which converts uroporphyrinogen I or III into coproporphyrinogen I or III, respectively, was purified about 5,500-fold from chicken erythrocytes. Purification was accomplished by chromatography on DEAE-cellulose, ammonium sulfate fractionation, chromatography on Sephadex G-100, and chromatofocusing. The most purified preparation was homogeneous on polyacrylamide gel electrophoresis and had a specific activity of 1,420 units/mg of protein, the highest value so far reported. The molecular weight, as determined by Sephadex G-150 gel chromatography, is 79,000. The subunit molecular weight, as determined by sodium dodecyl sulfate-polyacrylamide gel electrophoresis, is 39,700, suggesting that uroporphyrinogen decarboxylase is dimeric in form. The purified enzyme had an isoelectric point of 6.2 and a pH optimum of 6.8. The SH reagents inhibited the enzyme activity, but neither metal ions nor cofactor requirements could be demonstrated. A new and simple method for the separation of free uroporphyrin, hepta-, hexa-, and pentacarboxylic porphyrins and coproporphyrin was developed using a high pressure liquid chromatograph equipped with a spectrofluorometric detector. Kinetic studies of the sequential decarboxylation of uroporphyrinogen with purified enzyme were performed. 3,4,3',4'-Tetrachlorobiphenyl and 3,4,5,3',4'5'-hexachlorobiphenyl which specifically induce delta-aminolevulinic acid synthetase also strongly inhibit uroporphyrinogen decarboxylase directly at two steps, i.e. first in the formation of hexacarboxylic porphyrinogen III from heptacarboxylic porphyrinogen III and second in the formation of heptacarboxylic porphyrinogen III from uroporphyrinogen III.  相似文献   

10.
An attractant for Mexican fruit fly, Anastrepha ludens (Loew) (Diptera: Tephritidae), was developed from a commercial product called Sabor Uva containing processed Concord grape juice. The principal volatile components of Sabor Uva aroma were identified and an aqueous mixture of 15 components that was gas chromatographically similar to Sabor Uva was prepared. This mixture was equivalent to Sabor Uva in attractiveness by using wind-tunnel bioassays. After deleting chemicals that did not contribute to attractiveness, and increasing the concentrations of the remaining chemicals, the final attractant contained propylene glycol (90,000 ppm, vol/vol), acetic acid (4500), methyl anthranilate (1800), ethyl 2-methylpropionate (670), and one or both of the esters ethyl 3-methylbutyrate (44) and 2-methylbutyl propionate (44), in aqueous solution. This mixture was approximately 1.8X as attractive as Sabor Uva by indirect comparison. Deletion of propylene glycol, acetic acid, methyl anthranilate, or ethyl 2-methylpropionate from the mixture significantly decreased attractiveness. Deletion of either of the other two esters seemed to diminish attractiveness although effects were not statistically significant. Deletion of water from the mixture significantly decreased attractiveness. We conclude that propylene glycol, acetic acid, methyl anthranilate, water, and at least one or as many as all three of the methyl-branched esters are essential for complete attractiveness.  相似文献   

11.
A series of meso-5,10,15-tris(N-methyl-4-pyridiniumyl)-20-(4-alkylamidophenyl) porphyrins were synthesized by derivatizing the amino group on the phenyl ring with the following hydrophobic groups: –C(O)C7F15, –C(O)CHCH2, C(O)CH3, –C(O)C7H15, and –C(O)C15H31. The cationic tris-pyridiumyl porphyrin core serves as a DNA binding motif and a photosensitizer to photomodify DNA molecules. The changes of the UV–Vis absorption spectra during the titration of these porphyrins with calf thymus DNA revealed a large bathochromic shift (up to 14 nm) and a hypochromicity (up to 55%) of the porphyrins Soret bands, usually considered as proof of porphyrin intercalation into DNA. Association constants (K) calculated according to the McGhee and von Hippel model, were in the range of 106–107 M−1. An increase in hydrophobicity of the substituents at the 20−meso-position produced higher binding affinity. These porphyrins caused photomodification of the supercoiled plasmid DNA when a green laser beam at 532 nm was applied. Those with higher surface activity acted more efficiently as DNA photomodifiers. The porphyrin with a perfluorinated alkyl chain (–COC7F15) at the meso-20-position inhibited the growth of gram-positive bacteria (S. aureus, or S. epidermidis). Other porphyrins exhibited moderate activity against both gram-negative and gram-positive organisms.  相似文献   

12.
A study is described of the regulation of porphyrin synthesis in Escherichia coli using a heme-permeable, hemH deletion mutant, designated VS212. This strain utilizes only exogenous hemin that is supplied in the medium and accumulates porphyrins since the final step in the synthesis of heme is genetically blocked. It is possible, therefore, to monitor the rate of synthesis of heme by examining the accumulation of porphyrins. Using this system, we found that the rate of production of porphyrins depended on the availability of heme. The lower the concentration of hemin in the medium, the higher the level of porphyrins that accumulated. We next examined the mechanism responsible for the activation of porphyrin synthesis upon starvation for heme. The main activation occurred at the step that leads to the synthesis of 5-aminolevulinic acid (ALA). Starvation for heme induced the expression of a hemA-lacZ fusion gene, as previously reported, but an activation pathway that is independent of the hemA promoter was also identified. We found that starvation for heme caused the stringent response, and such starvation promoted the synthesis of porphyrins without having any effect on the expression of the hemA-lacZ fusion gene. We suggest a model for the regulation of porphyrin synthesis whereby the synthesis of porphyrins is coordinated with that of proteins.  相似文献   

13.
Thin-layer chromatography on silica gel 60 plates in the solvent N,N-dimethylformamide/methanol/ethylene glycol/glacial acetic acid/1-chlorobutane/chloroform (4/35/6/0.4/18/20 by volume) separates porphyrin carboxylic acids by the number of free carboxyl groups. Coproporphyrins I and III and isocoproporphyrin are separated in 30 min, other porphyrins in 15 min. The N,N-dimethylformamide and acetic acid in the solvent strongly increase porphyrin fluorescence on the plates. Fading and diffusion of the fluorescent patterns is prevented by storage of the plates in the cold and dark without oxygen and with desiccant. In a preliminary step, porphyrins are purified in high yields, concentrated, and deacidified rapidly (2 min) by reversephase chromatography on cartridges containing a C18 spacer or on Amberlite XAD-2 columns. The methods are applied to urines of porphyria patients and for following porphyrin ester hydrolysis.  相似文献   

14.
The type III heptacarboxylic porphyrinogens derived from enzymic decarboxylation of an acetic acid substituent on uroporphyrinogen III to a methyl group by human hepatic uroporphyrinogen decarboxylase has been analysed by reversed-phase high-performance liquid chromatography with electrochemical detection. The results showed that all four possible heptacarboxylic acid porphyrinogen isomers, with the methyl group attached to rings A, B, C and D of the tetrapyrrole macrocycle, respectively, were formed in almost equal proportions. It was concluded that the normal pathway of uroporphyrinogen III decarboxylation in human liver follows a random mechanism.  相似文献   

15.
C K Lim  F Li    T J Peters 《The Biochemical journal》1987,247(1):229-232
A reversed-phase h.p.l.c. system is described for the separation of the four type-III heptacarboxylic porphyrinogen isomers. The effects of buffer concentration, pH and type and proportion of organic modifier in the mobile phase on retention and resolution of isomers were studied. Optimum separation on an ODS-Hypersil column was by elution with a ternary mobile phase of acetonitrile, methanol and 1 M-ammonium acetate, pH 5.16 (7:3:90, by vol.). Isomer identification was based on a comparison of their retention times with those of authentic standards, and was further confirmed by h.p.l.c. analysis of the characteristic mixture of three pentacarboxylic porphyrins formed after partial decarboxylation of individual isomers in 0.3 M-HCl at 160 degrees C.  相似文献   

16.
We have already reported in Balb C mouse transplantable mammary carcinoma, that uroporphyrin I and III are superior as tumour localizers when compared to hematoporphyrin derivative and a derivative thereof, photofrin II. This study compares the binding of porphyrins to proteins which may be found in tumour cells or stroma to investigate whether there is a common binding determinant. Coproporphyrin III and deuteroporphyrin IX which are non-tumour localizing porphyrins, were also part of the comparative study. The interaction of these porphyrins with acid soluble collagen and acid insoluble collagen, elastin, and fibrin was evaluated, and the binding of uroporphyrin isomers I and III and deuteroporphyrin IX to gelatin and fibrinogen, was also determined. The results suggest that collagen, especially the acid soluble form, and gelatin preferentially bind the four porphyrins which localize in mammary carcinoma tissue. The well reported observations that malignant epithelial cells, including breast cancer, produce collagen and contain a rate-limiting enzyme in collagen biosynthesis would support the notion that de novo synthesis of this protein may in part govern the tumour uptake and retention of porphyrins. Elastin, fibrinogen and fibrin showed non-discriminant binding to the porphyrins under study.  相似文献   

17.
Total phospholipids were extracted from the heart, hepatopancreas, and hemolymph of the Indian horseshoe crab Carcinoscorpius rotundicauda by the conventional method. Characteristic group reaction and 2-dimensional thin-layer chromatography on silica gel were used for identification of different phospholipids. The phospholipid profile obtained from hemolymph and 2 major organs are comparable and show phosphatidyl choline (PC) and phosphatidyl ethanolamine to be the major phospholipids. A phospholipid has been consistently detected migrating immediately below the PC in the thin-layer chromatogram of lipids extracted from the hepatopancreas. When mixed methyl esters of this slower moving PC are resolved on a silica gel plate ran in hexane ether:acetic acid 80:20:1, with appropriate controls, an additional spot is seen just below the normal methyl ester, indicating a difference between the fatty acid compositions of 2 PC (e.g., regular and slower). The slower mixed methyl esters were found to comprise mainly the 4 saturated fatty acids: lauric, myristic, palmitic, and stearic. The slow moving PC seems to consist mainly of molecular species with the above-mentioned saturated fatty acids at both Sn 1 and Sn 2 positions.  相似文献   

18.
We describe here a sensitive method for the purification and analysis of porphyrins present in hematoporphyrin derivative. Hematoporphyrin derivative is a solution containing a complex mixture of dicarboxylic porphyrins such as hematoporphyrin IX, monohydroxyethyl monovinyl deuteroporphyrin isomers, and protoporphyrin IX in addition to porphyrin aggregates of variable molecular sizes. This mixture is known for its ability to be selectively retained by tumor cells and for its cytotoxicity in the presence of light. In order to study the mechanisms of hematoporphyrin derivative uptake and its cellular metabolism, extraction methods are required that combine high recoveries with minimum changes of very labile components. Extraction with perchloric acid: methanol mixtures recovered only some 60% of the porphyrins taken up by tumor cells and artifactual fluorescent spots were seen on thin-layer chromatograms. Improved yields were obtained upon extraction with dimethyl sulfoxide or Triton X-100:4-(2-hydroxyethyl)-1-piperazineethanesulfonic acid (Hepes) buffer mixture, but the extracts were not suitable for reverse-phase thin-layer chromatography (RTLC). The procedure described here consists of extracting porphyrins from cultured tumor cells with a buffered detergent followed by sequential chromatography on DEAE-cellulose columns and on reverse-phase octadecylsilyl cartridges. Identification of the isolated free dicarboxylic porphyrins is conveniently done by RTLC.  相似文献   

19.
One of the major issues with heterologous production of proteins in Pichia pastoris X33 under AOX1 promoter is repeated methanol induction. To obviate repeated methanol induction, methyl esters were used as a slow release source of methanol in lipase expressing mut+ recombinant. Experimental design was based on the strategy that in presence of lipase, methyl esters can be hydrolysed to release their products as methanol and fatty acid. Hence, upon break down of methyl esters by lipase, first methanol will be used as a carbon source and inducer. Then P. pastoris can switch over to fatty acid as a carbon source for multiplication and biomass maintenance till further induction by methyl esters. We validated this strategy using recombinant P. pastoris expressing Lip A, Lip C from Trichosporon asahii and Lip11 from Yarrowia lipolytica. We found that the optimum lipase yield under repeated methanol induction after 120 h was 32866 U/L, 28271 U/L and 21978 U/L for Lip C, Lip A and Lip 11 respectively. In addition, we found that a single dose of methyl ester supported higher production than repeated methanol induction. Among various methyl esters tested, methyl oleate (0.5%) caused 1.2 fold higher yield for LipA and LipC and 1.4 fold for Lip11 after 120 h of induction. Sequential utilization of methanol and oleic acid by P. pastoris was observed and was supported by differential peroxisome proliferation studies by transmission electron microscopy. Our study identifies a novel strategy of using methyl esters as slow release methanol source during lipase expression.  相似文献   

20.
Mixed esters of ascorbic acid were synthesized using methyl esters of palm and soybean oils as acyl donors, in acetone at 50 degrees C, and catalyzed by Novozym 435. A conversion of 62% was obtained with palm oil methyl ester at an ascorbic acid to acyl donor molar ratio of 1:4; the mixed ester contained 45.89% ascorbyl palmitate, 42.59% ascorbyl oleate and 10.1% ascorbyl linoleate. Acylation with soybean oil methyl ester resulted in 17% conversion, yielding a mixed ester containing 10.08% ascorbyl palmitate, 20.68% ascorbyl oleate, and 64.96% of ascorbyl linoleate. The mixed esters of ascorbic acid can find direct use in food and cosmetics.  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