Substituted porphyrins

Information

  • Patent Grant
  • 7485721
  • Patent Number
    7,485,721
  • Date Filed
    Monday, June 9, 2003
    21 years ago
  • Date Issued
    Tuesday, February 3, 2009
    15 years ago
Abstract
A series of ortho isomers of meso tetrakis N-alkylpyridylporphyrins (alkyl being methyl, ethyl, n-propyl, n-butyl, n-hexyl, and n-octyl) and their Mn(III) complexes were synthesized and characterized by elemental analysis, uv/vis spectroscopy, electrospray ionization mass spectrometry and electrochemistry. An increase in the number of carbon atoms in the alkyl chains from 1 to 8 is accompanied by an increase in: (a) lipophilicity measured by the chromatographic retention factor, Rf; (b) metal-entered redox potential, E1/2 from +220 to +367 mV vs NHE, and (c) proton dissociation constant, pKa2 from 10.9 to 13.2. A linear correlation was found between E1/2 and Rf of the Mn(III) porphyrins and between the pKa2 and Rf of the metal-free compounds. As the porphyrins become increasingly more lipophilic, the decrease in hydration disfavors the separation of charges, while enhancing the electron-withdrawing effect of the positively charged pyridyl nitrogen atoms. Consequently, the E1/2 increases linearly with the increase in pKa2, a trend in porphyrin basicity opposite from the one we previously reported for other water-soluble Mn(III) porphyrins. All of these Mn(III) porphyrins are potent catalysts for superoxide dismutation (disproportionation). Despite the favorable increase of E1/2 with the increase in chain length, the catalytic rate constant decreases from methyl (log kcat=7.79) to n-butyl, and then increases such that the n-octyl is as potent an SOD mimic as are the methyl and ethyl compounds. The observed behavior originates from an interplay of hydration and steric effects that modulate electronic effects.
Description
INTRODUCTION

Low-molecular weight catalytic scavengers of reactive oxygen and nitrogen species, aimed at treating oxidative stress injuries, have been actively sought. Three major groups of manganese complexes have been developed and tested in vitro and in vivo; Mn porphyrins,1-9 Mn cyclic polyamines10 and Mn salen derivatives.11 Based on a structure-activity relationships that we developed for water-soluble MN(III) and Fe(III) porphyrins,2-4 Mn(III) meso tetrakis(N-methylpyridinium-2-yl)porphyrin (MnIIITM-2-PyP5+, AEOL-10112) and meso tetrakis(N-ethylpyridinium-2-yl)porphyrins (MnIIITE-2-PyP5+, AEOL-10113) were proposed and then shown to be potent catalysis for superoxide dismutation.4.12 The alkyl substitutions at the ortho positions restrict the rotation of the pyridyl rings with respect to the porphyrin plane. Consequently both compounds exist as mixtures of four atropoisomers, all of which were shown to be equally potent catalysts for O2 dismutation.13 These Mn porphyrins also allow SOD-deficient Escherchia coli to grow under aerobic conditions,4,12 and offer protection in rodent models of oxidative stress such as stroke,14 diabetes,15 sickle cell disease,16 and cancer/radiation.17 The high formal +5 charge of these metalloporphyrins could influence their tissue distribution, transport across biological membranes, and binding to other biomolecules and their low lipophilicities may restrict their protective effects. With the aim of modulating metalloporphyrin subcellular distribution, higher N-alkylpyridylporphyrin analogues (Scheme I) with increased lipophilicity were synthesized. We anticipate that their comparative kinetic and thermodynamic characterization will deepen our insight into the modes of action of porphyrin-based catalytic antioxidants and the mechanisms of oxidative stress injuries.





BRIEF DESCRIPTION OF THE DRAWINGS


FIG. 1. Structures of the most hydrophilic (MnIIITM-2-PyP5+) and the most lipophilic (MnIIITnOct-2-PyP5+) members of the series studied. The αβαβ atropoisomers are shown.



FIG. 2. The lipophilicity, Rf of H2T(alky)-2-PyP4+ (A) and MnIIIT(alkyl)-2-PyP5+ compounds (B) vs the number of CH2 groups.



FIG. 3. Proton dissociation constants pKa2 of the metal-free porphyrins, H2T(alkyl)-2-PyP4+ (A), and the metal-centered redox potentials E1/2 for the Mn(III)/Mn(II) couple of MnIII(alkyl)-2-PyP5+ porphyrins (B) as a function of Rf. Inserts: pKa2 (FIG. 3A) and E1/2(FIG. 3B) vs the number of CH2 groups.



FIG. 4. The reactivity of water-soluble Mn(III) porphyrins (A) (ref 4) and MnIIIT(alkyl)-2-PyP5+ porphyrins (B) as catalysts for O2 dismutation, expressed in terms of log kcat vs E1/2.



FIG. 5. E1/2 for the Mn(III)/Mn(II) couple of MnIIIT(alkyl)-2-PyP5+ porphyrins vs pKa2 of the corresponding metal-free ligands. Insert: E1/2 of water-soluble Mn(III) porphyrins vs pKa3 (data from ref 4); MnIIITE-2-PyP5+ (1), MnTM-2-PyP5+ (2), MnPTrM-2-PyP4+ (3), MnIIITM-4-PyP5+ (4), MnIIITM-3-PyP5+ (5), MnIIIT(2,6-Cl2-3-SO3-P)P3− (6), MnIIIT(TFTMA)P5+ (7), MnIIIT(αααα-2-MINP)P5+ (8), MnIIIT(2,6-Cl2-3-SO3-P)P3− (9), MNIIIT(2,4,6-Me3-3,5-(SO3)2-P)P7− (10), MnIII(TMA)P5+ (11), MnTSPP3− (12), MnTCPP3− (13), MnIIIhematoP (14).



FIG. 6. Cyclic voltammetry of 0.5 mM MnIIITE-2-PyP5+ and MnIIITnHex-2-PyP5+ porphyrins in a 0.05 M phosphate buffer (pH 7.8, 0.1 M NaCl) at a scan rate of 0.1 V/s.



FIG. 7. Electrospray mass spectrometry of 0.5 mM solutions of H2T(alkyl)2-PyP4+ compounds in 1:1 water: acetonitrile at a cone voltage of 20 V.



FIG. 8. Electrospray mass spectrometry of 0.5 mM solutions of MnIIIT(alkyl)2-PyP5+compounds in 1:1 water: acetonitrile at a cone voltage of 20 V.



FIG. 9. As the slkyl chains of Mn(III) porphyrins lengthen, the favorable increase in E1/2 overcomes the unfavorable steric/electrostatic effects such that the n-octyl is as potent a catalyst of O2 dismutation as are the methyl and ethyl compounds.





DETAILED DESCRIPTION OF THE INVENTION

The present invention relates to a compound of formula




embedded image


wherein


each R is, independently, an C1-C12 alkyl (preferably, a C8 to C12 alkyl),


each A is, independently, hydrogen or an electron withdrawing group,


M is a metal selected from the group consisting of manganese, iron, copper, cobalt, nickel and zinc, and


Z is a counterion. In one embodiment, at least one A is a halogen.


The invention further relates to a method of protecting cells (eg mammalian cells) from oxidant-induced toxicity comprising contacting the cells with a protective amount of a compound as described above. The invention further relates to a method of treating a pathological condition of a patient resulting from oxidant-induced toxicity comprising administering to the patient an effective amount of such a compound. The invention also relates to a method of treating a pathological condition of a patient resulting from degradation of NO., comprising administering to the patient an effective amount of a compound as described above. Additionally, the invention relates to a method of treating a patient for inflammatory lung disease comprising administering to the patient an effective amount of a compound as described above. The inflammatory lung disease can be a hyper-reactive airway disease. The disease can be asthma.


The entire content of all documents cited herein are incorporated herein by reference. Also incorporated herein by reference is Batinic-Haberle et al, J. Chem. Soc., Dalton Trans. 2002, 2689-2696.


Also incoprporated by reference is U.S. application Ser. No. 09/880,125, filed Jun. 14, 2001.


EXAMPLE
Experimental

Materials and Methods


General. MnCl2×4 H2O, and Baker-flex silica gel IB TLC plates were purchased from J.T. Baker. N,N′-dimethylformamide, ethyl p-toluenesulfonate, 2-propanol (99.5+%), NH4PF6(99.99%), NaCl, sodium L-ascorbate (99+% ) and tetrabutylammoniurn chloride were from Aldrich, while xanthine, and ferricytochrome c were from Sigma. The n-propyl, n-butyl, n-hexyl and n-octyl esters of p-toluenesulfonic acid were from TCI America. Methanol (anhydrous, absolute), ethanol (absolute), acetone, ethyl ether (anhydrous), chloroform, EDTA and KNO3 were from Mallinckrodt and acetonitrile was from Fisher Scientific. Xanthine oxidase was prepared by R. Wiley and was supplied by K.V. Rajagopalan.18 Catalase was from Boehringer, ultrapure argon from National Welders Supply Co., and tris buffer (ultrapure) was from ICN Biomedicals, Inc.


H2T(alkyl)-2-PyP4+. Tetrakis(2-pyridyl)porphyrin, H2T-2-PyP was purchased from Mid-Century Chemicals, Chicago, Ill. The increased lipophilicity of the n-propyl, n-butyl, n-hexyl, and n-octyl analogues required a slight modification of the synthetic approach used for methyl and ethyl compounds.4,12 Typically, 100 mg of H2T-2-PyP was dissolved in 20 mL of DMF at 100° C., followed by the addition of 4 mL of the corresponding p-coluenesulfonate. The course of N-alkylation was followed by thin-layer chromatography on silica gel TLC plates using 1:1:8 KNO3-saturated H2O:H2O: acetonitrile as a mobile phase. While complete N-alkylation is achieved within a few hours for the methyl analogue, the required time gradually increases and it took three and five days to prepare the n-hexyl and n-octyl compounds, respectively. Upon completion, for the methyl, ethyl and n-propyl compounds, the reaction mixture was poured into a separatory funnel containing 200 mL each of water and chloroform and shaken well. The chloroform layer was discarded and the extraction with CHCl3 was repeated several times. The n-butyl, n-hexyl and n-octyl analogues are more lipophilic and tended to remain in the chloroform layer. Therefore, increasing amounts of methanol were added to the water/CHCl3 mixture in order to force the porphyrin into the aqueous/methanol layer. This layer was filtered and the porphyrin was precipitated as the PF6 salt by the addition of a concentrated aqueous solution of NH4PF6. The precipitate was thoroughly washed with 1:1 2-propanol:diethylether in the case of methyl and ethyl compounds and with pure diethylether for the others. The precipitate was then dissolved in acetone, filtered and precipitated as the chloride salt by the addition of tetrabutylammonium chloride dissolved in acetone. The precipitate was washed thoroughly with acetone, and dried in vacuo at room temperature. Elemental analysis: H2TnPr-2-PyPCl4×12.5 H2O (C52H71N8O12.5Cl4): Found: C, 54.2; H, 6.42; N, 9.91; Cl, 12.04. Calculated: C, 54.20; H, 6.18; N, 9.68; Cl, 12.25. H2TnBut-2-PyPCl4×10.5 H2O (C56H75N8O10.5Cl4): Found: C, 57.16; H, 6.94; N, 9.513; Cl, 1.77. Calculated: C, 57.10; H, 6.41; N, 9.51; Cl, 12.03. H2TnHex-2-PyPCl4×11 H2O (C64H100N8O11Cl4): Found: C, 59.19; H, 7.31; N, 8.61; Cl, 11.09. Calculated: C, 59.16; H, 7.751; N, 8.60; Cl, 10.91. H2TnOct-2-PyPCl4×13.5 H2O (C64H121N8O13.5Cl4): Found: C, 59.37; H, 7.41; N, 7.73. Calculated: C, 59.37; H, 8.37; N, 7.69.


MnIIIT(alkyl)-2-PyP5+. Metalation of the N-alkylated porphyrins was achieved as described previously for the methyl and ethyl compounds.4.12 Metal incorporation became slower as the alkyl chains lengthened. Under same conditions (20-fold excess metal, 25° C., pH 12.3) it occurs almost instantaneously for methyl and ethyl, within minutes for n-propyl, in ˜30 minutes for n-butyl, in ˜1 hour with the n-hexyl, and took several hours at 100° C. for the n-octyl porphyrin. The formation of the Mn(I) porphyrin and its oxidation to Mn(III) were clearly distinguishable steps when the n-hexyl and n-octyl analogues were metalated. As was the case with the metal-free ligands, the PF6 salts of Mn(III) n-propyl, n-butyl, n-hexyl and n-octyl compounds were washed only with diethylether. Elemental analysis: MnIIITnPr-2-PyPCl5×11.5 H2O (MnC52H75N8O11.5Cl5: Found: C, 50.90; H, 6.07; N, 9.27; Cl, 13.48. Calculated: C, 50.85; H, 6.16; N, 9.12; Cl, 14.43. MnIIITnBut-2-PyPCl5×12.5 H2O (MnC56H85N8O12.5Cl5): Found: C, 51.58; H, 6.33; N, 9.55; Cl, 15.53. Calculated: C, 51.64; H, 6.58; N, 8.60; Cl, 13.61. MnIIITnHex-2-PyPCl5×10.5 H2O (MnC64H97N8O12.5Cl5): Found: C, 55.64; H, 7.14; N, 8.23; Cl, 12.60. Calculated: C, 55.76; H, 7.09; N, 8.13; Cl, 12.86. MnIIITnOct-2-PyPCl5×10 H2O×2.5 NH4Cl (MnC64H122N10.5O10Cl7.5): Found: C, 53.56; H, 7.13; N, 9.12; Cl, 16.84. Calculated: C, 53.53; H, 7.60; N, 9.10; Cl, 16.46.


Thin-layer chromatography. All ligands and their Mn(III) complexes were chromatographed on silica gel TLC plates using 1:1:8 KNO3-saturated H2O:H2O: acetonitrile. The atropoisomers could not be separated for the methyl19 and ethyl analogues,24 they begin to separate for the n-propyl and n-butyl species and were clearly resolved with the n-hexyl and n-octyl compounds.


Uv/vis spectroscopy. The uv/vis spectra were taken on a Shimadzu UV-2501 PC spectrophotometer at 25° C. The proton dissociation constants (pKa2), were determined spectrophotometrically at 25° C., at an ionic strength of 0.1 M (NaOH/NaNO3), as previously described.4


Electrochemistry. Measurements were performed on a CH Instruments Model 600 Voltammetric Analyzer.3.4 A three-electrode system in a small volume cell (0.5 mL to 3 mL), with a 3 mm-diameter glassy carbon button working electrode (Bioanalytical Systems), plus the Ag/AgCl reference and Pt auxilliary electrodes was used. Solutions contained 0.05 M phosphate buffer, pH 7.8, 0.1 M NaCl, and 0.5 m/M metalloporphyrin. The scan rates were 0.01-0.5 V/s, typically 0.1 V/s. The potentials were standardized against the potassium ferrocyanide/ferricyanide20 and/or against MnIIITE-2-PyP5+. All voltammograms were reversible.


Electrospray mass spectrometry. ESMS measurements were performed on a Micromass Quattro LC triple quadrupole mass spectrometer equipped with a pneumatically assisted electrostatic ion source operating at atmospheric pressure. Typically, the 0.5 mM 50% aqueous acetonitrile solutions of chloride salts of metal-free porphyrins or their Mn(III) complexes were introduced by loop injection into a stream of 50% aqueous acetonitrile flowing at 8 μL/min. Mass spectra were acquired in continuum mode, scanning from 100-500 m/z in 5 s, with cone voltages of 20 V and 24 V. The mass scale was calibrated using polyethylene glycol.


Catalysis of O2dismutation. We have previously shown that the O2/cytochrome c reduction assay gives the same catalytic rate constants as does pulse radiolysis for MnIIITE-2-PyP5+, {MnIIIBVDME}2, {MnIIIBV}2 and MnCl2.21 Therefore the convenient cytochrome c assay was used to characterize the series of Mn(III) N-alkylpyridylporphyrins. The xanthine/xanthine oxidase reaction was the source of O2 and ferricytochrome c was used as the indicating scavenger for O2.22 The reduction of cytochrome c was followed at 550 nm. Assays were conducted at (25±1) ° C., in 0.05 M phosphate buffer, pH 7.8, 0.1 mM EDTA, in the presence and absence of 15 μg/mL of catalase. Rate constants for the reaction of metalloporphyrins with O2′ were based upon competition with 10 μM cytochrome c, kcyt c=2.6×105 M−1 s−1 as described elsewhere.21 The O2 was produced at the rate of 1.2 μM per minute. Any possible interference through inhibition of the xanthine/xanthine oxidase reaction by the test compounds was examined by following the rate of urate accumulation at 295 nm in the absence of cytochrome c. No reoxidation of cytochrome c by the metalloporphyrins was observed


Results


Thin Layer Chromatography. The increase in the length of the alkyl chains is accompanied by an increase in the lipophilicity of the compounds as indicated by the increase in the retention factor Rf (porphyrin path/solvent path) (Table 1, FIG. 2). The apparent lag that was observed in the case of shorter chains with Mn(III) complexes (FIG. 2B), is presumably due to their higher overall formal charge (+5 for the Mn(III) complexes, +4 for the ligand). As the chains lengthen, their contribution to the overall lipophilicity increases, and eventually the n-octyl porphyrin and its Mn(III) complex are more alike in Rf than are methyl analogues.


Uv/vis spectroscopy. Molar Absorptivies. The porphyrins obeyed the Beer-Lambert law from 10−7 M to 10−5 M, and the uv/vis data are given in Table 2. As the length of alkyl chains increased from methyl to n-butyl a red shift of the Soret absorption maxima was generally observed, as well as an increase in the molar absorptivities, and these effects plateau beyond buryl compound. Such trends may be understood in terms of the interplay of porphyrin nucleus distortion (red shifts) and the electron-withdrawing (blue shifts) effect of the N-alkylpyridyls groups.12,23


Metalation behavior and proton dissociation constants. The rates of Mn2+ incorporation at pH ˜12.3 decreased with an increase in chain length. The same was found for the kinetics of Zn2+ and Cu2+ insertion into these compounds below pH 7, where the kinetics were first order in metal and porphyrin concentration.24 Since the free-base porphyrin H2P4+ reactants were mixtures of the four atropoisomers, each isomer has a similar metalation rate constant. As noted before for both water soluble and insoluble porphyrins, compounds with substituents in the ortho positions tend to metalate more slowly than derivatives with the same groups in the meta orpara positions. 25-34


The proton dissociation constants, Ka2 and Ka3 are defined as follows:

H2P4+custom character HP3++H+Ka2  [1]
H3P5+custom character H2P4++H+ Ka3  [2]

The pKa2 values for the N-alkylpyridyl series are given in Table 1. As the alkyl chains lengthen the porphyrins become less hydrated and the separation of charges (eq [1]) becomes less favorable, ie. pKa2 increases (FIG. 3 insert). FIG. 3A shows the linear relationship between pKa2 and Rf.


Equilibrium constants pKa3 for reaction [2] are 1.8 for the meta H2TM-3-PyP4+, 1.4 for the para H2TM-4-PyP4+, and −-0.9 for ortho H2TM-2-PyP4+.4,25 While the meta and para N-methylpyridylporphyrins are mixtures of protonated H3P5+ and H4P6+ species in 1.0 M HCl, the ortho substituted H2TM-2-PyP4+ to H2TnOct-2-PyP4+ compounds remain as the unprotonated free base H2P4+ in 1.0 M HCl and in 1.0 M HNO3. With ortho, meta and para N-methylpyridylporphyrins the pKa2 increases as the pKa3 increases.


The half-lives for the acid and anion-catalyzed removal of zinc from Zn N-methylated derivatives35 in 1.0 M HNO3 were 89 s for the meta, 165 s for the para, and 19 hours for the ortho ZnTM-2-PyP4+. No indication of zinc loss was found within a week for the ZnTnHex-2-PyP4+ compound.36 Similar behavior is found in 1.0 M HCl, with t1/2 ranging from 21 s for the meta methyl to 76 hours for ZnTnOct-2-PyP4+.24 In accord are the observations that when solid MnTnHex-2-PyP5+ was dissolved in 12 M HCl, the spectra did not change within 3 months, while over 50% of the Mn from MnIIITM-2-PyP5+ species was lost within a month. In addition to porphyrin ring distortion,29-32 the steric hindrance and solvation effects imposed by the progressively longer alkyl chains may also contribute to the differences in metaladon/demetalation behavior.


Due to their high metalcentered redox potentials, the Mn(III) meso tetrakis ortho N-alkylpyridylporphyrins in vivo will be readily reduced with cell reductants such as ascorbic acid.2,3,12 The reduced Mn(II) porphyrins will also be transiently formed in the catalysis of O2 dismutation. Therefore, we also examined the behavior of the reduced and more biologically relevant MnIIT(alkyl)-2-PyP4+ compounds. We compared the methyl, n-hexyl and n-octyl derivatives (6 μM) aerobically and anaerobically in the presence of a 70-fold excess of ascorbic acid (pH 7.8, 0.1 M tris buffer) and in the presence and absence of a 150-fold excess of EDTA. Under anaerobic conditions both Mn(II) porphyrins were stable to Mn loss and porphyrin decomposition inside 24 hours. Aerobically, −40% of Mn methyl but none of the Mn n-hexyl and n-octyl compounds underwent degradation within 125 min. The absorption spectral changes indicate that the degradation occurred through the Mn porphyrin catalyzed reduction of oxygen by ascorbate resulting in the formation of H2O2. The peroxide in turn causes porphyrin destruction. These observations are consistent with previous results which indicate that a more electron rich compound (MnIITM-2-PyP4+) reduces O2 faster than does a more electron deficient species (MnIITnOct-2-PyP4+).2,3 EDTA did not significantly influence porphyrin degradation or Mn loss.


Electrochemistry. Cyclic voltammetry of the Mn(III) porphyrins shows a reversible voltammogram that we ascribe to the Mn(III)/Mn(II) redox couple. The metal-centered redox potentials, E1/2 are in Table 1 and the representative voltammograms of the MnIII/IITE-2-PyP5+/4+ and MnIII/IITnHex-2-PyP5+/4+ compounds are shown in the Supporting Material, FIG. 6. Both lipophilicity (FIG. 2B) and E1/2 (FIG. 3B, insert) increase exponentially with the number of CH2 groups in the alkyl chains. Consequently, the increase in E1/2 is a linear function of Rf (FIG. 3B).


Electrospray mass spectrometry. The ESMS proved to be a valuable tool for accessing the properties of the free base porphyrins and their Mn complexes whereby the impact of structure on salvation, ion-pairing, redox properties, protonation/deprotonation, dealkylation, and catalytic properties are clearly depicted.


H2T(alkyl)-2-PyP4+. The ESMS of the metal-free porphyrins obtained at the low cone voltage of 20 V showed dominant molecular ions assigned to H2P4+/4 and/or its mono-deprotonated analogue, H2-P4+-H+/3 (Table 3, FIG. 7). Negligible double deprotonation (H2P4+-2H+/2) was noted. Only H2TM-2-PyP4+ gave rise to a high-intensity H2P4++H+/5 peak.


The ESMS shows a pronounced decrease in solvation by acetonitrile as the alkyl chains lengthen. Compared to the base peak, the relative intensities of the mono-solvated molecular ions range from 40% for methyl, 15% for ethyl, and <10% for the higher analogues. Only with the n-hexyl and n-octyl porphyrins are small peaks (<5%) from ions associated with chloride found.


From methyl to n-butyl, the ratio of the molecular ion to mono-deprotonated ion peaks decreases, consistent with the trend in pKa2.Thus, the base peak for methyl is that of the molecular ion, while the base peak for the n-propyl and n-butyl porphyrins is the mono-deprotonated ion. This pka2 trend is overcome by the higher lipophilicities of the n-hexyl and n-octyl compounds, where roughly equal-intensity molecular ion (100%) and mono-deprotonated ion (98%) peaks are observed. The loss of one alkyl group (H2P4+-a+/3) was noted for all derivatives (except for the methyl), and either no or negligible loss of a second alkyl group (H2P4+-2a+/2) was found


MnIIIT(alkyl)-2-PyP5+. The ESMS of the Mn(III) complexes was done at a lower cone voltage (20 V) than in our previous study (30-58 V).37 Therefore, less fragmentation occurs and more solvent-associated and ion-paired species could be observed (Table 4, FIG. 8). Solvation and ion pairing are more pronounced when compared with the metal-free ligands. The more lipophilic Mn(C) porphyrins are more easily desolvated in the electrospray ionization source. In accordance with our previous observations, the ESMS also clearly reflects the redox properties of these compounds.37,38 The higher the E1/2 the more reduced porphyrins are noted. Species solvated with acetonitrile or associated with chloride were observed with both Mn(III) and Mn(II) compounds. Two chlorides were associated only with Mn(III) porphyrins.


In the ESMS of the n-hexyl and noctyl porphyrins we observed strong signals at m/z 337 and 375 that are assigned to compounds doubly reduced either at the metal (MnIP3+/3) or at both the metal and porphyrin ring (MnIIP3+/3). Such doubly reduced manganese porphyrins should have a higher tendency to lose the metal, and indeed peaks for the metal-free species were found for the n-hexyl and n-octyl derivatives, while only traces of doubly reduced and demetalated species were found for n-butyl.


The ESMS behavior of Mn porphyrins changes sharply once the alkyl chains lengthen beyond butyl, as observed with corresponding metal-free analogues. No loss of methyl groups was detected.37 As the chains lengthen up to butyl the loss of an alkyl group from Mn(III) and Mn(II) porphyrins becomes more pronounced and then the tendency decreases with n-hexyl and n-octyl. The same trend, but of lower intensity was noted for the loss of two alkyl groups. The ratio of mono-chlorinated Mn(III) to mono-chlorinated Mn(II) species decreases from methyl to n-butyl and then increases up to n-octyl. Thus the base peak of the methyl and ethyl porphyrins relates to MnIIIP5++Cl/4, while for the n-propyl and n-butyl derivatives it relates to MnIIP4++Cl/3. Yet, with the n-hexyl, the MnIIIP5++Cl/4 and MnIIP4++Cl/3 peaks are both of 100% intensity, and the di-chlorinated species (MnIIIP5++2Cl/3) is of 86% intensity. With the n-octyl analogue, the mono- and di-chlorinated species give rise to 100% MnIIIP5++Cl/4 and 89% MnIIIP5++2Cl/3 peaks, and the third most intense (59%) signal relates to MnIIP4++Cl/3. The lack of significant association of metal-free porphyrins with chloride observed here and elsewhere,37 strongly supports the idea that chloride is bound to the metal. Furthermore, at the same cone voltage, the base peak of ortho MnTM-2-PyP5+ is the mono-chlorinated species, which was only 35% for para isomer. This suggests that the longer the chains, the more defined the cavity, which can hold up to two chloride ions, and the more stable is the Mn(III) state. While a species bearing two chlorides is hardly noted in MnIIITM-2-PyP5+, it is the second major peak in the ESMS of MnIIITnOct-2-PyP5+.


Catalysis of O2 dismutation. None of the parent metal-free porphyrins exhibit any O2 dismuting activity. All of the manganese compounds are potent catalysts of O2 dismutation with log kcat between 7.79 and 7.25. As shown in Table 1, log kcat decreases from methyl to n-butyl and then increases, making n-octyl and methyl of comparable antioxidant potency.


Discussion


When designing metalloporphyrin SOD mimics we are aiming at approximating the redox properties of the enzyme active site. Superoxide dismutases catalyse the dismutation (disproportionation) of O2 to H2O2 and O2 at ˜+300 V vs NHE (pH 7.0).39.40 This potential is roughly midway (+360 mV vs NHE) between the potential for the reduction (+890 V vs NHE)41 and the oxidation of O2 (−160 V vs )41 thus providing an equal driving force for both half-reactions in the catalytic cycle. The O2 dismutation by CuZn—SOD occurs with catalytic rate constant, kcat=kred=kox=2×109 M−1 s−1 (log kcat=9.3).42-44


We previously demonstrated a structure-activity relationship between log kcat and the metal-centered E1/2 of the Mn(III)/Mn(II) couple for a variety of water-soluble meso substituted porphyrins (FIG. 4A).2-4 Electron-withdrawing substituents on the porphyrin ring shift E1/2 towards more positive values resulting in higher values for kcat.2-4 Each 120 mV increase in E1/2 gave a 10-fold increase in kcat, 4 consistent with the Marcus equation45 for outer-sphere electron transfer reactions (FIG. 4A). The Marcus equation is valid as long as one of the two steps in the catalytic dismutation cycle is

MnIIIP+O2custom characterMnIIP+O2, kred  [2]
MnIIP+O2+2H+custom characterMnIIIP+H2O2, kox   [3]

rate-limiting.


On the basis of such structure-activity relationships, the ortho isomers of Mn(III) meso tetrakis N-methyl- and N-ethylpyridylporphyrins were tested and proved to be potent catalysts of O2 dismutation. Their log kcat values are 7.79 and 7.76 and they operate at potentials (+220 and +228 V) similar to the potential of the enzyme itself. These two metalloporphyrins also exhibit protection in in vivo models of oxidative stress injuries.14-17 We have now extended our work to a series of MnIIIT(alkyl)-2-PyP5+ compounds where alkyl is methyl, ethyl, n-propyl, n-butyl, n-hexyl, and n-octyl (FIG. 1). The significant differences in lipophilicity along the series (FIG. 2A), with retention of catalytic potency (Table 1), might lead to favorably selective subcellular distributions of these new MnIIIT(alkyl)-2-PyP5+ compounds and hence broader their utility.


E1/2 vs pKa2. We did not expect a profound change in E1/2 along the series based on the fact that the increase in alkyl chain length from methyl to n-hexyl is without effect on the basicity of alkylamines.46 However, we found that the metal-centered redox potentials varied from +220 mV for methyl to +367 mV (vs NHE) for the n-octyl compound. Such an increase in E1/2 may originate from progressively unshielded positive charges at pyridyl nitrogens which would then exert stronger electron-withdrawing effect on the coordinated Mn as the compounds increase in lipophilicity. This reasoning is supported by the ESMS data (Table 4, FIG. 8) which show that the susceptibility to desolvation is accompanied by a greater preponderance of reduced Mn(II) porphyrin ions as the alkyl chains of the Mn complexes lengthen. We have previously reported4 that mainly electronic effects determine the relation between the pKa of the metal-free porphyrin and the E1/2 of the corresponding metal complex such that the decrease in pKa3 is accompanied by a linear increase in E1/2 (FIG. 5, insert). However as the compounds become increasingly more lipophilic, the lack of solvation disfavors separation of charges (higher pKa2 values), while the electron-withdrawing effects of the positively charged pyridyl nitrogens are enhanced. Thus the electronic pKa2 effects are overcome by solvation/steric effects resulting in an inverted trend, i. e. the E1/2 now increases in a linear fashion with an increase in pKa2 (FIG. 5).


Log kcat vs E1/2. Based on a previously established structure-activity relationship for water-soluble Mn(III) porphyrins,4 we expected the 147 mV increase in E1/2 to be accompanied by a ˜12-fold increase in kcat (FIG. 4A).4 We actually found that kcat decreased from methyl to n-butyl, and then increased by the same factor of ˜3 to n-octyl (Table 1, FIG. 4B). One explanation is that the Mn porphyrins are solvated to different extents, as indicated by the ESMS data, and this in turn affects the magnitude of kcat. The trend in kcat may also be influenced by the electrostatic/steric effects originating from the shielding of the single positive charge on the Mn(III) center. Thus the difference in the magnitude of lipophilicity between the metal-free ligands (formally +4) and the Mn(III) complexes (formally +5) becomes less noticeable as the alkyl chains get longer (Table 1). These H2P4+ compounds of formal +4 charge behave in solution kinetically as +1.6 to +1.8 electrolytes.24 From methyl to n-butyl, log kcat decreases almost linearly (FIG. 4B, insert). Due to the exponential increase in E1/2 along the series of Mn porphyrins (FIG. 3B, insert), the unfavorable electrostatic/steric effects are in part opposed and finally overcome by the progressively more favorable redox potentials that originate from increased desolvation (lipophilicity). Consequently, the very lipophilic n-octyl compound is essentially as potent an SOD mimic as the less lipophilic methyl and ethyl derivatives.


Regan et al47 were able to uncouple the steric and solvation effects in reactions of chloride ions with methyl- and tert-butyl-substituted chloroacetonitrile, and showed that both were of comparable magnitudes. Similarly, the reactivity of N-alkylpyridylporphyrins are the result of the interplay of electronic, steric and salvation effects, the latter dominating with the more lipophilic members of the series.


Recent findings indicate that biologically relevant reactions, other than O2 dismutation, can occur at the metal center in Mn porphyrins.2,3,5,7,8,48-52 The same has been reported for the enzyme active site,20,53-57 thus raising the complexity of the free radical chemistry and biology of the enzymes and their mimics. Reactive oxygen and nitrogen species are involved in direct damage of key biological targets such as nucleic acids, proteins and fatty acids, and there is an increasing amount of evidence that such species are also involved in the modulation of signaling processes.14,58,59 Thus, it is important to understand the mechanisms of action of Mn porphyrins and related compounds. Based on the electrostatic, steric, solvation, and lipophilic effects observed in this study, we expect the members of N-alkylpyridyl series to differ one from another in in vivo models of oxidative stress injuries with respect to their specificity towards reactive oxygen and nitrogen species as well as with regard to their pharmacokinetics. Such work is in progress.


Abbreviations


SOD, superoxide dismutase; AN, acetonitrile; DMF, N,N′-dimethylformamide; NHE, normal hydrogen electrode; TLC, thin-layer chromatography; H2P4+, any meso tetrakis N-alkylpyridylporphyrin ligand; MnIII/IIP4+/5+ any Mn(III/II) meso tetrakis N-alkylpyridylporphyrin; meso refers to the substituents at the 5,10,15, and 20 (meso carbon) position of the porphyrin core. MnIIIT(alkyl)-2(3,4)-PyP5+, manganese(III) meso tetrakis(N-methyl, N-ethyl, N-n-propyl, N-n-butyl, N-n-hexyl, N-n-octyl)pyridinium-2(3,4)-yl)porphyrin; alkyl is M, methyl; E, ethyl; nPr, n-propyl; nBu, n-butyl; nHex, n-hexyl; nOct, n-octyl on the pyridyl ring; 2 is the ortho, 3, the meta and 4 the para isomer: MnIIITM-2-PyP5+ is AEOL-10112, and MnIIITE-2-PyP5+ is AEOL-10113; MnIIIPTr(M-2-PyP4+, manganesc(III) 5-phenyl-10,15,20-tris(N-methylpyridinum-2-yl)porphyrin; MnIIIBM-2-PyP3+, manganese(III) meso bis(2-pyridyl)-bis(N-methylpyridinium-2-yl)porphyrin; MnIIITrM-2-PyP4+, 5-(2-pyridyl)-10,15,20-tris(N-methylpyridinium-2-yl)porphyrin; MnIIIT(TMA)P5+, manganese(III) meso tetrakis(N, N, N-trimethylanilinium-4-yl)porphyrin; MnIIIT(TFTMA)P5+, manganese(III) mesa tetrakis(2,3,5,6-tetrafluoro-N, N, N-trimethylanilinium-4-yl)poprhyrin; MnIIITCPP3−, manganese meso tetrakis(4-carboxylatophenyl)porphyrin; MnTSPP3−, manganese(III) meso tetrakis(4-sulfonatophenyl)porphyrin; MnIIIT(2,6-Cl4-3-SO3-P)P3−manganese (III) meso tetrakis(2,6-dichloro-3-sulfonatophenyl)porphyrin; MnIIIT(2,6-F2-3-SO3-P)P3−, manganese (III) meso tetrakis(2,6-difluoro-3-sulfonatophenyl)porphyrin; MnIIIT(2,4,6-Me3-3-(SO3)4-P)P7−, manganese(III) 5,10,15,20-tetrakis(2,4,6,-trimethyl-3,5-disulfonatophenyl)porphyrin; MnIIIhematoP, manganese(III) hematoporphyrin IX.


REFERENCES



  • 1. R. F. Pasternack, A. Banth, J. M. Pasternack and C. S. Johnson, J. Inorg. Biochem. 1981, 15, 261 (b) R. F. Pasternack and B. J. Halliwell, J. Am. Chem. Soc. 1979,101, 1026.

  • 2. I. Batinić-Haberle, Methods Enzymol. 2002, 349, 223.

  • 3. I. Spasojević and I. Batinić-Haberle, Inorg. Chim. Acta, 2001, 317, 230.

  • 4. I. Batinić-Haberle, I. Spasojević, P. Hambright, L. Benov, A. L. Crumbliss and I. Fridovich, Inorg. Chem. 1999, 38, 4011.

  • 5. G. Ferrer-Sueta, I. Batinić-Haberle, I. Spasojević, I. Fridovich and R. Radi, Chem. Res. Toxicol. 1999, 12, 42.

  • 6. R. Kachadourian, I. Batinić-Haberle and I. Fridovich, Inorg. Chem. 1999, 38, 391.

  • 7. J. Lee, J. A. Hunt and J. T. Groves, J. Am. Chem. Soc. 1998, 120, 6053.

  • 8. J. P. Crow, Arch. Biochem. Biophys. 1999, 371, 41.

  • 9. M. Patel and B. J. Day, Trends Pharmacol. 1999, 20, 359.

  • 10. (a) K. Aston, N. Rath, A. Naik, U. Slomczynska, O. F. Schall and D. P. Riley, Inorg. Chem. 2001, 40, 1779. (b) S. Cuzzocrea, E. Mazzon, L. Dugo, A. P. Caputi, K. Aston, D. P. Riley and D. Salvemini, Br. J. Pharmacol. 2001, 132, 19.

  • 11. (a) S. Melov, J. Ravenscroft, S. Malik, M. S. Gill, D. W. Walker, P. E. Clayton, D. C. Wallace, B. Malfroy, S. R. Doctrow and G. J. Lithgow, Science, 2000, 289, 1567. (b) K. Baker, C. Bucay Marcus, K. Huffman, H. Kruk, B. Malfroy and S. R. Doctrow, J. Pharmacol. Exp. Ther., 1998, 284, 215.

  • 12. I. Batinić-Haberle, L. Benov, I. Spasojević and I. Fridovich, J. Biol ChewM, 1998, 273, 24521.

  • 13. I. Spasojević, R. Menzeleev, P. S. White and I. Fridovich, Inorg. Chem., 2002, submitted.

  • 14. G. B. Mackensen, M. Patel, H. Sheng, C. C. Calvi, L Batinić-Haberle, B. J. Day, L. P. Liang, I. Fridovich, J. D. Crapo, R. D. Pearlstein and D. S. Warner, J. Neurosci., 2001, 21, 4582.

  • 15. J. D. Piganelli, S. C. Flores, C. Cruz, J. Koepp, I. Batinić-Haberie J. Crapo, B. J. Day, R. Kachadourian, R. Young, B. Bradley and K. Haskins, Diabetes, 2002, 51, 347.

  • 16. M. Aslan, T. M. Ryan, B. Adler, T. M. Townes, D. A. Parks, J. A. Thompson, A. Tousson, M. T. Gladwin, M. M. Tarpey, M., R. P. Patel, I. Batinić-Haberle, C. R. White and B. A. Freeman, Proc. Natl. Acad Sci. USA, 2001, 98, 15215.

  • 17. (a) I. Batinić-Haberle, I. Spasojević, I. Fridovich, M. S. Anscher and {hacek over (Z)}. Vujaskovic, Proc of the 43rd Annual Meeting of American Society for Therapeutics in Radiation Onciology, San Francisco 2001, 235-236. (b) Z. Vuja{hacek over (s)}ković, I. Batinić-Haberle, I. Spasojević, T. V. Samulski, M. W. Dewhirst and M. S. Anscher, Annual Meeting of Radiation Research Society, San Juan, Puerto Rico 2001. (c) {hacek over (Z)}. Vuja{hacek over (s)}ković, I. Batinić-Haberle, I. Spasojević, Irwin Fridovich, M. S. Anscher and M. W. Dewhirst, Free Rad. Biol. Med. 2001, S128.

  • 18. W. R. Waud, F. O. Brady, R. D. Wiley and K. V. Rajagopalan, Arch. Biochem. Biophys. 1975, 19, 695.

  • 19. T. Kaufmann, T., B. Shamsai, R. S. Lu, R. Bau and G. M. Miskelly, Inorg. Chem. 1995, 34, 5073.

  • 20. I. M. Kolthof and W. J. Tomsicek, W. J., J. Phys. Chem. 1935, 39, 945.

  • 21. L Batinić-Haberle, I. Spasojević, R. D. Stevens, P. Hambright, A. N. Thorpe, J. Grodkowski, P. Neta and I. Fridovich, Inorg. Chem. 2001, 40, 726.

  • 22. J. M. McCord and I. Fridovich, J. Biol. Chem. 1969, 244, 6049.

  • 23. I. Batinić-Haberle, S. I. Liochev, I. Spasojević and I. Fridovich, Arch. Biochem. Biophys. 1997, 343, 225.

  • 24. P. Hambright, I. Spasojević, I. Fridovich and I. Batinić-Haberle, in preparation.

  • 25. P. Hambright, Water-Soluble Metalloporphyrins in The Porphyrin Handbook, K. M. Kadish, K. M. Smith, R. Guillard, Eds. Academic Press, N.Y. 2000, Chapter 18.

  • 26. T. P. G. Sutter and P. Hambright, J. Coord. Chem. 1993, 30, 317.

  • 27. L. R. Robinson and P. Hambright, Inorg. Chem., 1992, 31, 652.

  • 28. J. B. Reid and P. Hambright, Inorg. Chem. 1977, 16, 968.

  • 29. M. Inamo, N. Kamiya, Y. Inada, M. Nomura and S. Funahashi, Inorg. Chem., 2001, 40, 5636.

  • 30. P. B. Chock and P. Hambright, J. Am. Chem. Soc. 1974, 96, 3123.

  • 31. S. Funahashi, Y. Inada and M. Inamo, Anal. Sci. 2001, 17, 917.

  • 32. T. P. G. Sutter, R. Rahimi, P. Hambright, J. Bommer, M. Kumar and P. Neta, J. Chem. Soc. Faraday Trans. 1993, 84, 495.

  • 33. B. Cheng, O. Q. Munro, H. M. Marques and W. R. Scheidt, J. Am. Chem. Soc. 1997,119, 10732.

  • 34. R. F. Pasternack, N. Sutin and D. H. Turner, J. Am. Chem. Soc. 1976, 98, 1908.

  • 35. P. Hambright, T. Gore and K. Burton, Inorg. Chem. 1976, 15, 2314.

  • 36. J. Davilla, A. Harriman, M. -G. Richoux and L. R. Milgrom, J. Chem. Soc. Chem. Commun. 1987, 525.

  • 37. I. Batinić-Haberle, R. D. Stevens and I. Fridovich, J. Porphyrins Phthalocyanines, 2000, 4, 217.

  • 38. R. Kachadourian, N. Srinivasan, C. A. Haney and R. D. Stevens, J. Porphyrins Phthalocyanines, 2001, 5, 507.

  • 39. Vance, C. K and Miller, A. -F., Biochemistry, 2001, 40, 13079.

  • 40. (a) G. D. Lawrence and D. T. Sawyer, Biochemistry, 1979, 18, 3045. (b) W. C. Jr. Barrette, D. T. Sawyer, J. A. Free and K. Asada, Biochemistry 1983, 22, 624.

  • 41. Wood, P. M., Biochem. J., 1988, 253, 287.

  • 42. Vance, C. K. and Miller, A.-F., J. Am. Chem Soc., 1998, 120, 461.

  • 43. R. M. Ellerby, D. E. Cabelli, J. A. Graden and J. S. Valentine, J. Am. Chem. Soc., 1996, 118, 6556.

  • 44. D. Klug-Roth, I. Fridovich and J. Rabani, J. Am. Chem. Soc., 1973, 95, 2786.

  • 45. R. A. Marcus, Annu. Rev. Phys. Chem., 1964, 15, 155.

  • 46. CRC Handbook of Chemistry and Physics, D. R. Lide, Editor-in-Chief, 74th Edition, 1993-1994, CRC Press, Boca Raton.

  • 47. I. Spasojević, I. Batinić-Haberle and I. Fridovich, Nitric Oxide: Biology and Chemistry 2000, 4, 526.

  • 48. (a) C. K. Regan, S. L. Craig and J. I. Brauman, Science, 2002, 295, 2245.

  • 49. R. Shimanovich and J. T. Groves, Arch. Biochem. Biophys., 2001, 387, 307.

  • 50. N. Jin, J. L. Bourassa, S. C. Tizio and J. T. Groves, Angew. Chem. Int. Ed., 2000, 39,3849.

  • 51. H. Zhang, J. Joseph, M. Gurney, D. Becker and B. Kalyanaraman, J. Biol. Chem., 2002, 277, 1013.

  • 52. N. Motohashi and Y. Saito, Chem. Pharm. Bull., 1995, 43, 505.

  • 53. C. Quijano, D. Hernandez-Saavedra, L. Castro, J. M. McCord, B. A. Freeman and R. Radi, J. Biol. Chem., 2001, 276, 11631.

  • 54. (a) S. L. Jewer, A. M. Rocklin, M. Ghanevati, J. M. Abel and J. A. Marach, Free Rad. Biol. Med., 1999, 26, 905. (b) S. P. A. Goss, R. J. Singh and B. Kalyanaraman, J. Biol. Chem., 1999, 274, 28233. (c) S. I Liochev and I. Fridovich, Free Rad. Biol. Med., 199, 27, 1444.

  • 55. (a) A. G. Estevez, J. P. Crow, J. B. Sampson, L Reither, J. Zhuang, G. J. Richardson, M. M. Tarpey, L. Barbeito and J. S. Beckman, Science, 1999, 286, 2498. (b) S. I. Liochev and I. Fridovich, J. Biol. Chem., 2001, 276, 35253.

  • 56. S. I. Liochev and I. Fridovich J. Biol. Chem., 2000, 275, 38482.

  • 57. E. D. Coulter, J. P. Emerson, D. M. Jr., Kurtz and D. E. Cabelli, J. Am. Chem. Soc., 2000, 122, 11555.

  • 58. B. M. Matata and M. Galinanes, J. Biol. Chem., 2002, 277, 2330.

  • 59. (a) {hacek over (Z)}. Vuja{hacek over (s)}ković, I. Batinić-Haberle, M. S. Anscher, Z. N. Rabbani, T. V. Samulski, K. Amin, M. W. Dewhirst and Z. Haroon, Proc. of the 43rd Annual Meeting of American Society for Therapeutics in Radiation Onciology, San Francisco 2001, 88-89. (b) {hacek over (Z)}. Vuja{hacek over (s)}ković, I. Batinić-Haberle, Z. N. Rabbani, Q.-F. Feng, S. K Kang, L Spasojević, T. V. Samulski, I. Fridovich, M. W. Dewhirst, M. S. Anscher, Free Rad. Biol. Med. In press.










TABLE 1







Metal-Centered Redox Potentials E1/2, log kcat for O2 Dismutation, and


Chromatographic Rf values.














E1/2c



Porphyrin
Rfa
pKa2b
mV vs NHE
log kcatd














MnIIITM-2-PyP5+
0.09 (0.13)
10.9
+220
7.79


MnIIITE-2-PyP5+
0.13 (0.21)
10.9
+228
7.76


MnIIITnPr-2-PyP5+
0.20 (0.31)
11.4
+238
7.38


MnIIITnBut-2-PyP5+
0.33 (0.46)
11.7
+254
7.25


MnIIITnHex-2-PyP5+
0.57 (0.63)
12.2
+314
7.48


MnIIITnOct-2-PyP5+
0.80 (0.86)
13.2
+367
7.71






aRf (compound path/solvent path) on silica gel TLC plates in 1:1:8 KNO3-saturated H2O:H2O:acetonitrile. Rf for the metal-free porphyrins are in parentheses.




bpKa2 determined at 25° C. ionic strength 0.10 (NaNO3/NaOH).




cE1/2 determined in 0.05 M phosphate buffer (pH 7.8, 0.1 M NaCl).




dKcat determined using the cytochrome c assay. in 0.05 M phosphate buffer, pH 7.8, at (25 ± 1) ° C.














TABLE 2







Molar Absorptivities of Tetrakis (N-alkylpyridinium-2-yl)porphyrin


chlorides and their Mn(III) Complexes.








Porphyrin
λnm(log ε)a





H2TM-2-PyP4+
413.2(5.32); 510.4(4.13); 544.4(3.49); 581.4(3.72);



634.6(3.13)


H2TE-2-PyP4+
414(5.33); 511(4.20); 545(3.58); 582(3.80);



635(3.38);


H2TnPr-2-PyP4+
415(5.38); 511.5(4.24); 545(3.62); 583(3.84);



635(3.37)


H2TnBut-2-PyP4+
415(5.37); 511(4.24); 544(3.60); 583(3.84);



636(3.39)


H2TnHex-2-PyP4+
415.5(5.34); 510.5(4.24); 543(3.62); 584.5(3.84);



638(3.43)


H2TnOct-2-PyP4+
416.5(5.31); 510(4.25); 542(3.59); 585(3.82);



639.5(3.43)


MnIIITM-2-PyP5+
363.5(4.64); 411(4.27); 453.4(5.11); 499(3.66);



556(4.03); 782(3.15)


MnIIITE-2-PyP5+
363.5(4.68); 409(4.32); 454(5.14); 499(3.75);



558(4.08); 782(3.26)


MnIIITnPr-2-PyP5+
363(4.70); 411(4.37); 454(5.21); 498(3.81);



559(4.12); 782(3.35)


MnIIITnBut-2-PyP5+
364(4.70); 410(4.35); 454(5.23); 498(3.83);



559(4.14); 781(3.33)


MnIIITnHex-2-PyP5+
364.5(4.70); 415(4.57); 454.5(5.21); 507(3.85);



560(4.12); 780(3.30)


MnIIITnOct-2-PyP5+
364(4.72); 414(4.44); 454.5(5.24); 500.5(3.84);



559.5(4.14); 781(3.25)






aThe molar absorptivities were determined in water at room temperature.














TABLE 3







Electrospray Mass Spectrometry Results for H2T(alkyl)-2-PyP4+


Compounds.a









m/z













Speciesb
M
E
nPr
nBu
nHex
nOct
















H2P4+/4
169
184
198
212
239
268


H2P4+ + AN/4
180
194
208
222
250


H2P4+ + 2AN/4
190


H2P4+ − H+/3
226
245
263
282
319
357


H2P4+ − H+ + AN/3
240
258
278


H2P4+ − H+ + H2O/3



288


H2P4+ − H+ + Cl/2




496


H2P4+ − a+/3

235
249
263
291
319


H2P4+ − a+ − H+/2

352
374
394
436


H2P4+ − a+ + H2O/3


255


H2P4+ − 2a+/2


352
366


H2P4+ + H+/5
136


H2P4+ + H+ + AN/5
143


H2P4+ + H+ + 2AN/5
152


H2P4+ + H+ + 2Cl/3




343
381


H2P4+ + 2H+ + 2Cl/4





286


H2P4+ − 2H+/2
339
367
395
423
479






a0.5 mM solutions of H2P4+ in 1:1 acetonitrile:water, 20 V cone voltage.




bAN denotes acetonitrile and a is an alkyl group.














TABLE 4







Electrospray Mass Spectrometry for MnIIIT(alkyl)-2-PyP5+ Porphyrins.a









m/z













Speciesb
M
E
nPr
nBu
nHex
nOct
















MnIIIP5+/5
146
157






MnIIIP5+ + AN/4
155
166
177
188


MnIIIP5+ + 2AN/5
163
174
185
196


MnIIIP5+ + 3AN/5
171
182
193
205


MnIIIP5+ + 4AN/5
179
190

213


MnIIIP5+ + 5AN/5
187
198


MnIIIP5+ + 6AN/5
195


MnIIIP5+ + H2O/5
150


MnIIIP5+ + Cl/4
192
206

234
262
290


MnIIIP5+ + 2Cl/3
267
286
305
323
361
398


MnIIIP5+ + Cl + AN/4
202
216
230
244
272


MnIIIP5+ − a/4


200


MnIIIP5+ − a + AN/4

200

221
242


MnIIIP5+ − a + Cl/3

264
279
293
321
349


MnIIIP5+ − 2a/3

243
252
262

299


MnIIIP5+ − 2a + AN/3



275
294


MnIIP4+/4
183
197
211


281


MnIIP4+ + AN/4
193
207
221
235
263


MnIIP4+ + 2AN/4
204


MnIIP4+ + Cl/3
255
274
293
312
349
387


MnIIP4+ − a/3

253
266
281
309
337


MnIIIP5+ − Mn3+ + H+/3



281
319
357


MIIP−3+/3 or MnIP3+/3



294
337
375






a0.5 mM solutions of MnIIIP5+ in 1:1 acetonitrile:water, 20 V cone voltage.




bAN denotes acetonitrile and a is an alkyl group.






Claims
  • 1. A compound of formula
  • 2. The compound according to claim 1 wherein at least one A is a halogen.
  • 3. The compound according to claim 1 wherein said compound is of Formula I or III.
  • 4. The compound according to claim 3 wherein said compound is of Formula I and M is manganese.
  • 5. A compound of formula
  • 6. The compound according to claim 5 wherein at least one A is ahalogen.
Parent Case Info

This application claims priority from Provisional Application No. 60/386,454, filed Jun. 7, 2002, the content of which is incorporated herein by reference.

US Referenced Citations (76)
Number Name Date Kind
2951799 Sharp Sep 1960 A
4614723 Schmidt Sep 1986 A
4657902 Kappas et al. Apr 1987 A
4746735 Kruper, Jr. et al. May 1988 A
4758422 Quay Jul 1988 A
4829984 Gordon May 1989 A
4837221 Bonnett Jun 1989 A
4851403 Picker et al. Jul 1989 A
4866054 Dori et al. Sep 1989 A
4885114 Gordon et al. Dec 1989 A
4892941 Dolphin et al. Jan 1990 A
4895719 Radhakrishnam Jan 1990 A
4963367 Ecanow Oct 1990 A
5010073 Kappas et al. Apr 1991 A
5051337 Sakoda et al. Sep 1991 A
5087438 Gordon Feb 1992 A
5109016 Dixon et al. Apr 1992 A
5130245 Marklund et al. Jul 1992 A
5162519 Bonnett Nov 1992 A
5169630 Okaya et al. Dec 1992 A
5171680 Mullenbach et al. Dec 1992 A
5192757 Johnson et al. Mar 1993 A
5192788 Dixon et al. Mar 1993 A
5202317 Bruice Apr 1993 A
5217966 Bruice Jun 1993 A
5223538 Fridovich Jun 1993 A
5227405 Fridovich Jul 1993 A
5236914 Meunier Aug 1993 A
5236915 Fiel Aug 1993 A
5248603 Marklund et al. Sep 1993 A
5262532 Tweedle et al. Nov 1993 A
5277908 Beckman et al. Jan 1994 A
5281616 Dixon et al. Jan 1994 A
5284647 Niedballa Feb 1994 A
5366729 Marklund et al. Nov 1994 A
5403834 Malfroy-Camine et al. Apr 1995 A
5405369 Selman et al. Apr 1995 A
5472691 Marklund et al. Dec 1995 A
5493017 Thieren et al. Feb 1996 A
5563132 Bodaness Oct 1996 A
5599924 Therien et al. Feb 1997 A
5604199 Funanage Feb 1997 A
5610293 Riley et al. Mar 1997 A
5637578 Riley et al. Jun 1997 A
5674467 Maier et al. Oct 1997 A
5747026 Crapo May 1998 A
5767272 Wijesekera et al. Jun 1998 A
5834509 Malfroy-Camine et al. Nov 1998 A
5874421 Riley et al. Feb 1999 A
5948771 Danziger Sep 1999 A
5976498 Neumann et al. Nov 1999 A
5976551 Mottez et al. Nov 1999 A
5994339 Crapo et al. Nov 1999 A
5994410 Chiang et al. Nov 1999 A
6013241 Marchal et al. Jan 2000 A
6046188 Malfroy-Camine et al. Apr 2000 A
6060467 Buelow et al. May 2000 A
6084093 Riley et al. Jul 2000 A
6087493 Wheelhouse et al. Jul 2000 A
6103714 Fridovich et al. Aug 2000 A
6127356 Crapo et al. Oct 2000 A
6180620 Salvemini Jan 2001 B1
6204259 Riley et al. Mar 2001 B1
6214817 Riley et al. Apr 2001 B1
6245758 Stern et al. Jun 2001 B1
6372727 Crow et al. Apr 2002 B1
6395725 Salvemini May 2002 B1
6403788 Meunier et al. Jun 2002 B1
6417182 Abrams et al. Jul 2002 B1
6548045 Sakata et al. Apr 2003 B2
6566517 Miura et al. May 2003 B2
6573258 Bommer et al. Jun 2003 B2
6602998 Kobuke et al. Aug 2003 B2
6624187 Pandey et al. Sep 2003 B1
20020042407 Fridovich et al. Apr 2002 A1
20020058643 Cherian et al. May 2002 A1
Foreign Referenced Citations (45)
Number Date Country
0 127 797 Dec 1984 EP
0 186 962 Jul 1986 EP
0 282 899 Sep 1988 EP
0 284 645 Oct 1988 EP
0 336 879 Oct 1989 EP
0 337 601 Oct 1989 EP
0 345 171 Dec 1989 EP
0 414 915 Mar 1991 EP
0 462 836 Dec 1991 EP
0 524 161 Jan 1993 EP
0 532 327 Mar 1993 EP
2 676 738 Nov 1992 FR
02289844 Nov 1989 JP
03273082 Dec 1991 JP
WO 9104315 Apr 1991 WO
WO 9119977 Dec 1991 WO
9207935 May 1992 WO
WO 9208482 May 1992 WO
WO 9215099 Sep 1992 WO
WO 9302090 Feb 1993 WO
WO 9404614 Mar 1994 WO
WO 9405285 Mar 1994 WO
WO 9510185 Apr 1995 WO
WO 9531197 Nov 1995 WO
WO 9609038 Mar 1996 WO
WO 9609053 Mar 1996 WO
WO 9640148 Dec 1996 WO
WO 9640223 Dec 1996 WO
WO 9706824 Feb 1997 WO
WO 9706830 Feb 1997 WO
WO 9706831 Feb 1997 WO
WO 9733588 Sep 1997 WO
WO 9733877 Sep 1997 WO
WO 9833503 Jun 1998 WO
WO 9858636 Dec 1998 WO
WO 9923097 May 1999 WO
WO 9955388 Nov 1999 WO
WO 0004868 Feb 2000 WO
WO 0019993 Apr 2000 WO
WO 0023568 Apr 2000 WO
WO 0043395 Jul 2000 WO
WO 0072893 Dec 2000 WO
WO 0075144 Dec 2000 WO
WO 0126655 Apr 2001 WO
WO 0196345 Dec 2001 WO
Related Publications (1)
Number Date Country
20040058902 A1 Mar 2004 US
Provisional Applications (1)
Number Date Country
60386454 Jun 2002 US