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 synthesis of asymmetric meso-aryl-substituted porphyrins containing three 4-methoxycarbonylphenyl groups, and as a forth substituent 4-hydroxyphenyl or 4-hydroxy-3- methoxyphenyl radicals, or the isomeric 3- and 4-pyridyl substituents is described. O-alkyl derivatives of 4-hydroxyl residue are obtained. The ytterbium complexes ofthese porphyrins were synthesized and studied their luminescence spectral properties were studied. A significant difference in the lifetimes of the excited state ofytterbium complexes of esters and acids of asymmetric porphyrins is demonstrated.  相似文献   

2.
Radical production during the photolysis of deaerated aqueous alkaline solutions (pH 11) of some water-soluble porphyrins was investigated. Metal-free and metallo complexes of tetrakis (4-N-methylpyridyl)porphyrin (TMPyP) and tetra (4-sulphonatophenyl)porphyrin (TPPS4) were studied. Evidence for the formation of OH radicals during photolysis at 615, 545, 435, 408 and 335 nm of Fe(III) TPPS4 is presented. Fe(III) TMPyP, Mn(III) TPPS4 and Mn(III) TMPyP also gave OH radicals but only during photolysis at 335 nm. The method of spin trapping with 5,5-dimethyl-1-pyrroline-1-oxide (DMPO) and 4-pyridyl-1-oxide-N-tert-butylnitrone (POBN) combined with e.s.r. was used for the detection of OH, H and hydrated electrons. With the spin trap DMPO, photolysis generated DMPO-OH adducts under certain conditions but no DMPO-H adducts could be observed. With POBN, no POBN-H adducts were found. The formation of OH was confirmed by studying competition reactions for OH between the spin traps and OH scavengers (formate, isopropanol) and the concomitant formation of the CO-2 adduct and the (CH3)2COH adduct with both DMPO and POBN. The photochemical generation of OH radicals was pH dependent; at pH 7.5 no OH radicals could be detected. Photolysis (615-335 nm) of dicyanocomplexes of the Fe(III) porphyrins did not produce OH radicals. When corresponding Cu(II), Ni(II), Zn(II) and metal-free porphyrins were photolysed at 615 and 335 nm, no OH radicals could be spin trapped. These results tend to associate the well-known phenomenon of photoreduction of Fe(III) and Mn(III) porphyrins with the formation of OH radicals. This process is described mainly as the photoreduction of the metal ion by the ligand-bound hydroxyl ion via an intramolecular process.  相似文献   

3.
Meso-substituted porphyrins, ((4-N-methyl-pyridyl)n(Ph)4-n)PH2, n = 1 to 4, bearing between 1 and 4 positive charges have been synthetized and studied for their interaction with Calf Thymus DNA. Competition binding experiments using ethidium bromide or one of its dimers show that these porphyrins and some of their Cu(II) or Fe(III)Cl complexes have apparent binding constants between 3 10(5) and 5 10(7) M-1. Fluorescence energy transfer experiments show that not only the tetracationic previously described porphyrin but also the tri- and dicationic porphyrins are able to intercalate into DNA. These data indicate a greater importance of the polyaromatic porphyrin ring than of the number or position of the positive charges for meso-tetra-arylporphyrin interaction with DNA.  相似文献   

4.
Thirty-three porphyrins or metalloporphyrins corresponding to the general formula [meso-[N-methyl-4(or 3 or 2)-pyridiniumyl]n(aryl)4-nporphyrin]M (M = H2, CuII, or ClFeIII), with n = 2-4, have been synthesized and characterized by UV-visible and 1H NMR spectroscopy and mass spectrometry. These porphyrins differ not only in the number (2-4) and position of their cationic charges but also in the steric requirements to reach even temporarily a completely planar geometry. In particular, they contain 0, 1, 2, 3, or 4 meso-aryl substituents not able to rotate. Interaction of these porphyrins or metalloporphyrins with calf thymus DNA has been studied and their apparent affinity binding constants have been determined by use of a competition method with ethidium bromide which was applicable not only for all the free base porphyrins but also for their copper(II) or iron(III) complexes. Whatever their mode of binding may be, their apparent affinity binding constants were relatively high (Kapp between 1.2 x 10(7) and 5 x 10(4) M-1 under our conditions), and a linear decrease of log Kapp with the number of porphyrin charges was observed. Studies of porphyrin-DNA interactions by UV and fluorescence spectroscopy, viscosimetry, and fluorescence energy transfer experiments showed that not only the tetracationic meso-tetrakis[N-methyl-4(or 3)-pyridiniumyl]porphyrins, which both involved four freely rotating meso-aryl groups, but also the corresponding tri- and dicationic porphyrins were able to intercalate into calf thymus DNA. Moreover, the cis dicationic meso-bis(N-methyl-2-pyridiniumyl)diphenylporphyrin, which involved only two freely rotating meso-aryl groups in a cis position, was also able to intercalate. The other meso-(N-methyl-2-pyridiniumyl)n(phenyl)4-nporphyrins, which involved either zero, one, or two trans freely rotating meso-aryl groups, could not intercalate into DNA. These results show that only half of the porphyrin ring is necessary for intercalation to occur.  相似文献   

5.
Abstract

Interactions of meso-tetra-(4-N-oxyethylpyridyl) porphyrin (TOEPyP(4)), its 3-N analog (TOEPyP(3)) and their Co, Cu, Ni, Zn metallocomplexes with duplex DNA have been investigated by uv/visible absorbance and circular dichrosim spectroscopies. Results reveal the interactions of these complexes with duplex DNA are of two types. (1) External binding of duplex DNA by metalloporphyrins containing Zn and Co, and (2) Binding of duplex DNA both externally and internally (by intercalation) by porphyrins not containing metals, and metalloporphyrins containing Cu and Ni. Results indicate that (4N-oxyethylpyridyl) porphyrins intercalate more preferably in the structure of duplex DNA and have weaker external binding than 3N-porphyrins.  相似文献   

6.
The reactions of dilute solutions of octaethylporphyrin and its iron (II) and iron (III) complexes with methyl, 2-cyanopropyl, t-butoxy, and benzoyloxy radicals are described. The results are summarized: (i) The reactivity of the porphyrin and its high-spin iron (II) and iron (III) complexes toward alkyl and t-butoxy radicals stands in the order: FeII > FeIII ? free porphyrin. For benzoyloxy radicals the order is FeII > Porp > FeIII. (ii) The exclusive path of reaction of high-spin iron (II) porphyrin with radicals is the rapid reduction of the radical and generation of an iron (III) porphyrin. The dominant path of reaction of high-spin iron (III) porphyrin with alkyl and (presumably) t-butoxy radicals is a rapid axial inner sphere reduction of the porphyrin. An axial ligand of iron is transferred to the radical. (iv) The reaction of benzoyloxy radicals with high or low-spin iron (III) porphyrins occurs primarily at the meso position. With the low-spin dipyridyl complex in pyridine the attendant reduction to iron (II) can be observed spectrally. Methyl radicals also reduce this complex by adding to the meso position. (v) The reaction of a radical with either an iron (II) or an iron (III) porphyrin results in the generation of the other valence state of iron and consequently oxidation and reduction products emanating from both iron species are obtained. (vi) No evidence for an iron (IV) is intermediate is apparent. (vii) Iron (II) porphyrins in solvents that impart either spin state are easily oxidized by diacyl peroxides. The occurrence of both axial and peripheral redox reactions with the iron complexes supports an underlying premise of a recent theory of hemeprotein reactivity. The relevance of the work to bioelectron transfer and heme catabolism is noted.  相似文献   

7.
Electron Paramagnetic Resonance (EPR) spectroscopy was employed in the study of the interaction between L-3,4-dihydroxyphenylalanine (L-Dopa) melanin and the cat-ionic porphyrins meso-tetrakis(1-methylpyridinium-4yl)-porphyrin (TMPyP), meso-tetrakis-(1-benzylpyridinium-4-yl)-porphyrin (TBzPyP), and their respectives complexes ZnTMPyP and ZnTBzPyP. By monitoring signal intensities and progressive microwave power saturation it was shown that the interaction increases the equilibrium concentration of free radicals in L-Dopa melanin in the dark. The extent of increase is dependent on the presence of molecular oxygen and on the type of porphyrin. Not all interacting sites available for complexation in L-Dopa melanin are involved in the formation of free radicals. It was also observed that the interaction with porphyrins promotes an increase in the number of photoinduced free radicals in L-Dopa melanin during illumination with visible light.  相似文献   

8.
The DNA strand scission activities of three positional isomers of Fe(III) meso-tetra(N-methylpyridyl)porphine (Fe(III)TnMPyP, where n = 2, 3 or 4) have been investigated using PM2 DNA as a substrate. A significant degree of strand scission activity was noted in the presence of oxygen without the addition of a reducing agent. This activity was probably due to the presence of reducing agents in the agarose gels used to separate the DNA forms, as higher levels were recorded with reducing agents added to the strand scission mixture. The relative order of strand scission activity in the absence of added reducing agents was found to be Fe(III)T2MPyP greater than Fe(III)T4MPyP greater than Fe(III)T3MPyP. Comparative studies were also made with Fe(II)bleomycin. High concentrations of some reducing agents inhibited strand scission. Oxygen was required to produce optimal strand scission activity for all three porphyrins. It was also noted from spectroscopic measurements that the reduced porphyrins were degraded in the presence of oxygen. Studies with a series of potential strand scission inhibitors suggest that hydrogen peroxide and possibly peroxy radicals are intermediates in the reaction mechanism, while diffusible hydroxyl radicals appear to be excluded. However, superoxide radicals cannot be ruled out.  相似文献   

9.
The influence of water-soluble cationic 3N- and 4N-pyridyl porphyrins with different peripheral substituents (oxyethyl, buthyl, allyl, and metallyl) on melting parameters of DNA has been studied. Results indicate that the presence of porphyrin changes the shape and parameters of DNA melting curve. The increase of porphyrins concentration results in the increase of the melting temperature (Tm) and the melting interval (ΔT) of DNA. At the porphyrin-DNA concentration ratio r?=?0.01, changes in the melting temperature have not been observed. The melting intervals almost do not change upon adding of the 4N-porphyrins, while the decrease of ΔT, in the presence of 3N-porphyrins, is observed. Because the intercalation binding mechanism occurs in GC-rich regions of DNA, we assume that 3N-porphyrins, intercalated in GC-rich regions, reduce the thermal stability of these sites, bringing them closer to the thermal stability of the AT-sites, which is the reason for the decrease in the melting interval. While at the relative concentration r?=?0.01 for 4-N porphyrins, already the external binding mechanism “turns on” and the destabilizing effect of porphyrins on GC-pairs compensates stabilizing effect on AT-pairs, as a result of which change in the melting of DNA upon complexation with these porphyrins is not observed. The decrease of the hypochromic effect also indicates the intercalation of investigated porphyrins in the DNA structure, which weakens the staking interaction of base pairs of DNA. The increase of the hypochromic effect of DNA upon binding with porphyrin depends on the type of peripheral substituents of the porphyrin. The results show that porphyrins with butyl and allyl substituents weaken staking interaction of base pairs less than porphyrins with other substituents. The largest change was observed for metallyl porphyrins. It can be the result of bulky peripheral substituents, which make significant local changes in DNA structure.  相似文献   

10.
Reactive oxygen species (ROS) from both endogenous and exogenous sources can cause oxidative DNA damage and dysregulated cell signaling, which are involved in the multistage process of carcinogenesis such as tumor initiation, promotion and progression. A number of structurally different anticarcinogenic agents inhibit inflammation and tumor promotion as they reduce ROS production and oxidative DNA damage. Evidence suggests that porphyrins can interfere with the actions of various carcinogens and mutagens by forming face-to-face complexes and their antimutagenic or antigenotoxic effects may also be attributed to their antioxidant activities. However, little is known regarding the anti-tumor promoting potential and mechanism of the porphyrin compounds. Based on our previous results on the inhibitory effects of chlorophyllin (CHL), hemin and tetrakis(4-benzoic acid)porphyrin (TBAP) against two-stage mouse skin carcinogenesis, we have investigated their anti-tumor promoting mechanisms. In the present work, CHL, hemin and TBAP reduced superoxide anion generation by 12-O-tetradecanoylphorbol-13-acetate (TPA) in differentiated HL-60 cells and the production of hydroxyl radicals by Fenton reaction. Porphyrins exert a dose-related inhibition of his+ reversion in Salmonella typhimurium TA102 induced by tert-butylhydroperoxide (t-BOOH). DNA strand breaks by ROS derived from H2O2/Cu(II) and the formation of 8-hydroxydeoxyguanosine (8-OH-dG) in calf thymus DNA treated with H2O2/UV also were inhibited markedly by porphyrins in a concentration-dependent manner. Furthermore, CHL, hemin and TBAP decreased myeloperoxidase (MPO) activity and H2O2 formation as well as epidermal ornithine decarboxylase (ODC) activity in mouse skin treated with TPA. These results demonstrate that the antioxidative properties of porphyrins are important for inhibiting TPA-induced tumor promotion.  相似文献   

11.
The development of catalytic, asymmetric transformations in water is a challenging task. The lanthanides are becoming reagents of choice for many Lewis acid-catalyzed reactions in aqueous media as they are water tolerant. However, enantioselective reactions catalyzed by lanthanides are difficult to achieve in water due to the instability of the reported catalysts. Herein we report the development of stable, well-defined chiral lanthanide complexes and their effectiveness in the asymmetric reduction of alpha-keto acids in aqueous solution. This is the first example of asymmetric reduction by a chiral lanthanide complex in water. Although modest ees are obtained (40-50%) the ytterbium complexes offer a unique advantage as they have the ability to monitor, direct from the reaction mixture, the % ee for the reaction, by 1H NMR, through a dipolar analysis of the observed paramagnetic shift.  相似文献   

12.
Interactions of meso-tetra-(4-N-oxyethylpyridyl) porphyrin (TOEPyP(4)), its 3-N analog (TOEPyP(3)) and their Co, Cu, Ni, Zn metallocomplexes with duplex DNA have been investigated by uv/visible absorbance and circular dichrosim spectroscopies. Results reveal the interactions of these complexes with duplex DNA are of two types. (1) External binding of duplex DNA by metalloporphyrins containing Zn and Co, and (2) Binding of duplex DNA both externally and internally (by intercalation) by porphyrins not containing metals, and metalloporphyrins containing Cu and Ni. Results indicate that (4N-oxyethylpyridyl) porphyrins intercalate more preferably in the structure of duplex DNA and have weaker external binding than 3N-porphyrins.  相似文献   

13.
The π-cation radicals of the metalloporphyrins magnesium octaethylporphyrin (MgOEP), magnesium tetraphenylporphyrin (MgTPP), and zinc tetraphenylporphyrin (ZnTPP), as well as the free base porphyrins of tetratolylporphyrin (H2TTP) and tetraphenylporphyrin (H2TPP) have been formed at liquid nitrogen temperatures in a rigid matrix of alkyl chloride glasses containing CCl4 or 1,1,2,2-tetrachloroethane (TCE), following photolysis of the porphyrins with visible light. The reaction proceeds via electron transfer from the photoexcited porphyrin to the solvent molecules; the efficiency of thie electron transfer may be qualitatively evaluated in terms of electron tunneling in the solid matrices. This is the first report of the photochemical formation of a free base porphyrin π-cation radical species.  相似文献   

14.
The interaction of meso-tetra(4-N-hydroxyethylpyridyl)porphyrin, meso-tetra(3-N-hydroxyethylpyridyl)porphyrin, and their zinc complexes with bovine serum albumin (BSA) was studied by electronic spectroscopy, CD, and equilibrium dialysis at pH 7.2. The titration of the porphyrins with BSA was accompanied by a decrease in light absorption and a bathochromic shift of the Soret band, as well as by the appearance of an isobestic point. The porphyrin interaction with BSA also led to the induction of positive CD spectra in the visible region, which is explained by the porphyrin sorption on the protein globule. The equilibrium dialysis helped in determining the stoichiometry of binding and the binding constants of the porphyrins under study with BSA using Scatchard plots. This interaction is nonspecific and reversible.  相似文献   

15.
Solution properties of three manganese porphyrins, in monomeric form, were investigated. These were the 'picket-fence-like' porphyrin Mn(III)-alpha,alpha,alpha,beta- tetra-ortho(N-methylisonicotinamidophenyl)porphyrin (Mn(III)PFP) and two 'planar unhindered' porphyrins, the Mn(III)TMPyP (tetrakis (4-N-methylpyridyl)porphyrin) and Mn(III)TAP (tetra(4-N,N,N-trimethylanilinium)porphyrin). The porphyrin properties studied were: the absorption spectra in their manganic and manganous forms; acid/base properties of the aquo complexes; the effect of potential axial ligands (up to a concentration of 0.1 mol dm-3) and their one electron reduction potentials. Knowing these properties, the reaction of the Mn(III) porphyrins with the superoxide radical and other reducing radicals were studied using the pulse radiolysis technique. The second-order reaction rate constant of O2- with the Mn(III) porphyrins, which governs the catalytic efficiency of the metalloporphyrins upon the disproportionation of the superoxide radical, was 5.1 X 10(7) to 4.0 X 10(5) dm3 mol-1 s-1, depending on the pH and the nature of the metalloporphyrin. These values are at least one order of magnitude lower than found for Fe(III)TMPyP. One electron reduction of the three Mn(III) porphyrins by eaq-, CO2-, CH2OH and (CH3)2COH had similar second-order rate constants (10(9)-10(10) dm3 mol-1 s-1). That for (CH3)2(CH2)COH was about 10(5) dm3 mol-1 s-1. Reduction in all cases produced the corresponding Mn(II) porphyrin and no intermediate was found. The oxidation reaction of the Mn(II) porphyrins by O2- was approximately two orders of magnitude faster when compared to the reduction of Mn(III) porphyrins with the same radical. Since the reactivities of O2- towards the three manganese (III) compounds follow their reduction potentials, it is suggested that these reactions are governed by an outer-sphere mechanism. This suggestion is corroborated by the finding that water molecules acting as axial ligands, in these aqueous solution systems, are not replaced by another potential ligand when the latter is in the concentration range of 100 mM or less.  相似文献   

16.
DNA cleavage specificity of a group of cationic metalloporphyrins   总被引:9,自引:0,他引:9  
The ability of a group of water-soluble metalloporphyrins to cleave DNA has been investigated. Incubation of Mn3+, Fe3+, or Co3+ complexes of meso-tetrakis(N-methyl-4-pyridiniumyl)porphine (H2T4MPyP) with DNA in the presence of ascorbate, superoxide ion, or iodosobenzene results in DNA breakage. Comparisons between the rates of porphyrin autodestruction with the rates of strand scission of covalently closed circular PM2 DNA indicate that the porphyrins remain intact during the cleavage process. Analysis of the porphyrin-mediated strand scissions on a 139-base-pair restriction fragment of pBR322 DNA using gel electrophoresis/autoradiography/microdensitometry reveals that the minimum porphyrin cleavage site is (A X T)3. The cleavage pattern within a given site was found to be asymmetric, indicating that porphyrin binding and the strand scission process are highly directional in nature. In addition to an analysis of the mechanism of porphyrin-mediated strand breakage in terms of the DNA cleavage mechanism of methidium-propyl-iron-EDTA and Fe-bleomycin, the potential of the cationic metalloporphyrins as footprinting probes and as new "reporter ligands" for DNA is presented and discussed.  相似文献   

17.
Terpene ligands (1S,2S,5S)-3-[{2-[(2-hydroxybenzylidene)amino]ethyl}imino]-2,6,6-trimethylbicyclo[3.1.1.]heptane-2-ol and 3-({2-[(2-hydroxy-2,6,6-trimethylbicyclo[3.1.1.]hept-3-ilidene)amino]ethyl}imino)-2,6,6-trimethylbicyclo[3.1.1.]heptane-2-ol have been synthesized for the first time. The efficiency of complexes based on terpene and salen ligands in asymmetric sulfoxidation has been compared. Catalytic systems based on terpene ligands have been used for the first time in the asymmetric oxidation of phenylphenacyl sulfide with the formation of sulfoxide with an enantiomeric excess of 47%.  相似文献   

18.
The oxidative chemistry of nickel(II) porphyrins is reviewed. Whether electron abstraction occurs from the metal to yield Ni(III) or from the porphyrin to yield Ni(II) pi cation radicals is discussed in terms of the relative energy levels of the metal and porphyrin orbitals. The effects of axial ligands in further modulating this ordering as well as the orbital occupancy of Ni(III) are also reviewed. Structural considerations, based on existing stereochemical data for Ni(I), high spin Ni(II) and related Ni(III) tetraaza complexes, are used to predict the metrics of Ni(III) porphyrins for which no structural data are available.  相似文献   

19.
In this study, we performed an ultraviolet-visible (UV-Vis) and circular dichroism (CD) spectroscopic analysis of the binary and ternary supramolecular structures formed by self-assembling the following three water-soluble porphyrins with and without a chiral template: the negatively charged, meso-Tetra(4-sulfonatophenyl) porphine (H2TPPS4−); the positively charged meso-trans-(di(N-methyl-4-pyridyl)diphenyl) porphine (trans-DmPyDPP) and meso-cis-(di(N-methyl-4-pyridyl)diphenyl) porphine (cis-DmPyDPP). Polyglutamic acid (both L and D enantiomers) was selected as the chiral template due to its ability to change secondary structure with pH. The propensity for the porphyrins to show an induced CD in the presence of polyglutamic acid is demonstrated. The induced chirality of all supramolecular structures was found to depend on the pH of the solution, the chirality of the polymer, and the order of addition of the positively and negatively charged porphyrins (for ternary complexes). Of particular interest is that the interaction of H2TPPS4− with the chiral scaffold seems to undergo a dynamic rearrangement of the supramolecular structure as evident from the change in the CD spectrum over time. Moreover, experiments with ternary complexes suggest that the preferential interaction of trans-DmPyDPP with the random coil of the polymer shows promise as a sensor of protein secondary structure.  相似文献   

20.
Abstract

A series of novel cadmium(II) and mercury(II) halide and thiocyanate complexes with an asymmetric Schiff base ligand of 4-(3-(2-(4-(dimethyl aminophenyl allylidene aminopropyl-imino)prop-1-ethyl)-N,N-dimethyl benzene amine has been synthesised and characterised using spectral, physical and analytical data, such as 1H NMR, UV-Vis and FTIR spectroscopy, melting point, elemental analysis and molar conductivity measurements. The spectral and physical data proposed a pseudo-tetrahedral geometry around the metal centre in the metal complexes. Moreover, the in vitro antibacterial activity of all compounds was assayed against two gram-positive and two gram-negative bacterial strains by a disk diffusion method and the results showed that all compounds have antibacterial characteristics. Also, the minimum inhibitory concentration and minimum bactericidal concentration of each compound were determined.  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