Redox-Active Metals: Iron and Copper

1 downloads 0 Views 883KB Size Report
In: King TE, Mason HS, Morrison M, eds. Oxidases and Related Redox ..... [90] Barb WG, Baxendale JH, George P, Hargrave KR. Reactions of ferrous and ferric.
In: Principles of Free Radical Biomedicine. Volume 1 ISBN: 978-1-61209-773-2 Editors: K. Pantopoulos and H. M. Schipper © 2012 Nova Science Publishers, Inc. The exclusive license for this PDF is limited to personal website use only. No part of this digital document may be reproduced, stored in a retrieval system or transmitted commercially in any form or by any means. The publisher has taken reasonable care in the preparation of this digital document, but makes no expressed or implied warranty of any kind and assumes no responsibility for any errors or omissions. No liability is assumed for incidental or consequential damages in connection with or arising out of information contained herein. This digital document is sold with the clear understanding that the publisher is not engaged in rendering legal, medical or any other professional services.

Chapter 5

Redox-Active Metals: Iron and Copper Willem H. Koppenol* and Patricia L. Bounds Institute of Inorganic Chemistry, Department of Chemistry and Applied Biosciences Eidgenössische Technische Hochschule, CH-8093 Zürich, Switzerland

1. Introduction Dioxygen† [1, 2] (O2) is simultaneously necessary for life and a source of potential harm [3, 4]. Harmful reactions, in which partially reduced oxygen species (PROS)‡ such as superoxide (O2 ) and hydrogen peroxide (H2O2) may be formed, are, in part, mediated by redox-active metals, the most physiologically relevant of which are iron and copper. These metals are essential for the transport, storage, and activation of O2, and for electron transfer, e.g., during respiration (see also Vol. II, Chapters 19-20). Along the reductive pathway of O2 to water, partially reduced and potentially reactive intermediates are sequentially made and managed by metalloproteins [5]. Although the redox reactivity of the metals in these proteins *

Email: [email protected] Nomenclature. Formula, systematic name and still allowed trivial name in italics: O 2, dioxygen, oxygen; O2 , dioxide(•1 ) or dioxidanidyl, superoxide; HO2 , hydridodioxygen(•), dioxidanyl, or hydrogen dioxide (hydroperoxyl or perhydroxyl are obsolete); H2O2, dioxidane, hydrogen peroxide; HO , hydridooxygen(•) or oxidanyl, hydroxyl; O , oxide(• ) or oxidanidyl; OCl , oxidochlorate(1 ), hypochlorite; HOCl, hydroxidochlorine, hypochlorous acid; NO , oxidonitrogen(•) or nitrogen monoxide (nitric oxide is obsolete); ONOO , oxidodioxidonitrate(1 ), peroxynitrite; ONOOH, hydroperoxidooxidonitrogen, peroxynitrous acid; Fe3+ or Fe(III), iron(3+) or iron(III) (ferric is obsolete); Fe2+ or Fe(II), iron(2+) or iron(II) (ferrous is obsolete); Cu+ or Cu(I), copper(+) or copper(I) (cuprous is obsolete); Cu2+ or Cu(II), copper(2+) or copper(II) (cupric is obsolete); FeO2+, oxidoiron(2+) or oxidoiron(IV) (ferryl is obsolete)[1,2]. The abbreviations edta, dtpa and nta, and atp refer to the metal chelators ethylenediaminetetraacetate, diethylenetriaminepentaacetate, nitrilotriacetate and adenosinetriposphate, respectively; these abbreviations are not capitalised.[2] NTBI is non-transferrinbound iron; cp20 is 3-hydroxy-1,2-dimethylpyridin-4(1H)-one, and icl670, 4-[3,5-bis(2hydroxyphenyl)-1H-1,2,4-triazol-1-yl]benzoic acid. ‡ We use the acronym PROS rather than ROS (for “reactive oxygen species”); PROS is not generally meant to include O2, which, strictly speaking, is also a reactive species. Analogously, we apply the acronym PONS for “partially oxidized nitrogen species” instead of RNS for “reactive nitrogen species”. †

92

Willem H. Koppenol and Patricia L. Bounds

is harnessed and directed by the protein and/or co-factor to which it is bound, there can be some non-productive release of O2 or H2O2. One approach that nature uses to circumvent non-productive reactions of oxygen is to employ more than one redox-active metal ion, sometimes stabilized by one or more non-redox-active metal ions, for the transfer of multiple electrons over a short time frame, e.g. a copper and an iron ion for the reduction of O2 by cytochrome c oxidase [6], and four manganese ions for the oxidation of water in photosystem II [7]. Even so, oxyhemoglobin [8, 9] and oxymyoglobin [10] reportedly release O2 , and xanthine oxidase [11], and cytochromes P450 [12] are known to produce H2O2. Another source of O2 in vivo is the reduction of O2 by electrons that leak from the mitochondrial electron transport chain; in fact, 1 – 4% of all O2 consumed during respiration may be reduced to O2 in this fashion [13, 14] (Vol. II, Chapter 15). It is now well established that PROSs, including peroxynitrite (ONOO ), play roles in a stunning array of diseases [15, 16]. One function of the protein environment of a metalloprotein is to modulate the reactivity of the metal ion; the potential for harmful reactivity likely to be catalyzed by free metal ions such as iron and copper outside of the protein milieu is the topic of this review. In vivo, excess free metal ions are sequestered in specialized storage proteins ― iron in ferritin [17] (Vol. II, Chapter 19) and copper possibly in metallothioneins (Vol. II, Chapter 20), while circulating iron is captured by transferrin [18] and copper by ceruloplasmin [19]. However, transfer of metals from storage to functional metalloproteins and the turnover of these proteins likely gives rise to low concentrations of low-molecular weight metal ion complexes [20, 21]; this pool of iron and copper may increase as a consequence of oxidative stress.

2. Redox Active Iron and Copper In Vivo The role played by iron and copper in diseases of ageing and oxidative stress has been recently reviewed [22]. Both iron and copper ions “redox cycle”, i.e. shuttle between two oxidation states via transfer of a single electron (Figure 1), as part of their normal functions in vivo. Redox cycling is also widely believed to be involved in the generation of damaging PROS; “free” iron and copper, actually free in solution or non-specifically bound to albumin or lowmolecular-weight ligands, are invoked as being particularly damaging. However, the link between pools of free transition metals, PROS, and disease is tenuous, and the evidence in the literature is largely circumstantial. Further, the speciation, that is the ligands, of “free” iron and copper ions involved in disease states have not been precisely identified. Diseases in which excess iron plays a role include hemochromatosis and iron-loading anemias, such as thalassemia, sickle cell disease and myelodysplasia [23] (Vol. III, Chapter 2). When normal mechanisms of iron intake and elimination become imbalanced, the binding capacity of ferritin and transferrin are exceeded, and pools of free, or non-transferrin-bound, iron (NTBI) increase (Vol. II, Chapter 19). The NTBI, which has been detected in the serum of hemochromatosis [24], sickle cell anemia and thalassemia patients [25], may consist in part of iron(III)-citrate complexes [26]. Although citrate has been identified as a ligand in NTBI, the exact composition of the complex or complexes existing in vivo remain elusive [27]. Citrate is present in the serum at a concentration of 0.1 mM, and the clinical range of NTBI concentrations varies between 0 to 10 M [28]. The aqueous speciation of iron(III) citrate is highly dependent on the iron/citrate molar ratio and, under the clinical conditions described,

Redox-Active Metals: Iron and Copper

93

several complexes have been shown to be possible [29]. The mononuclear Fe(cit)25 complex predominates over the clinically observed iron/citrate molar ratios, and, additionally, the multinuclear complexes Fe3(cit)33 and Fe3(cit)47 become relevant at higher NTBI concentrations [30]. Thus, the term NTBI may encompass several different species, of which one or more could be redox-active. Altered concentrations of copper, zinc and iron ions in the brain are a feature of Alzheimer‟s disease, and copper is widely thought to contribute to the oxidative stress model of Alzheimer pathophysiology [31] (see also Vol. III, Chapter 10). Copper has been shown to bind tightly to amyloid- peptide (K = 10–11 M), and the copper-amyloid- complex oxidizes cholesterol to cytotoxic products [32, 33]. Copper also plays a role in Parkinson‟s disease, atherosclerosis, diabetes, and other age-related diseases, as reviewed by Brewer [22].

3. Biological Oxidants Reactive species formed in vivo that are thought to be physiologically relevant include O2 , nitrogen monoxide (NO ), the hydroxyl radical (HO ), the trioxocarbonyl radical (CO3 ), nitrogen dioxide (NO2 ), H2O2, ONOO , and hypochlorite (OCl ) (see also Chapters 2-4). The first four species in the list each possess an unpaired electron and are, therefore, called radicals; H2O2, ONOO and OCl have no unpaired electrons and are, therefore, not radicals§. It is important to recognise that radical species are also functionally produced in vivo, e.g., the nitric oxide synthases generate NO , a diatomic radical species that plays vital roles in neuronal cell communication, endothelial vasodilation, and immune response [34, 35]. Of all the biological oxidants listed above, HO is the most destructive: HO is a promiscuous oxidant that reacts with small organic molecules, proteins, nucleic acids, and lipids by abstraction of hydrogen or addition to a double bond to form a radical species that can undergo further transformations. The HO has a high positive electrode potential|| (E° (HO ,H+/H2O) = +2.31 V at pH 7) [36], and the second-order rate constants (kobs = 109 – 1010 M 1s 1) [37] are in a range that indicates diffusion-controlled reactions. In contrast, H2O2 does not react directly with most organic compounds, and any toxicity attributed to H2O2 §

||

The term “radical”, which was used in the past to refer to a group that is part of an organic molecule, e.g., “methyl radical,” has been replaced with substituent; in former usage, the term “free radical” referred to an unattached group, usually containing an unpaired electron. In current usage, the word “radical” indicates that the group possesses an unpaired electron, and the word “free” is no longer included. The values given are relative to the normal hydrogen electrode, a platinum electrode in a solution that is 1 molal in H+ (1 gram H+ per kg of water, essentially the same as 1 molar) and in equilibrium with an atmosphere of 100 kPa of H2, that, by definition, has a value of 0 V. If we use O 2 and O2 as an example, the electrode potential of the couple O2/O2 refers to the reaction: O2 + e O2 , with a value of 0.35 V, relative to the normal hydrogen electrode. The electrode potential was formerly called a reduction potential. This value would be measured if one could connect a stable solution of O 2 (1 molal) in equilibrium with an atmosphere of O2 (100 kPa) with the normal hydrogen electrode by way of two platinum electrodes connected to a volt meter, and a salt bridge between the solutions to allow ions to flow between the solutions. Electrons would flow from the negatively charged electrode in the O2/O2 solution – producing O2 – to the positive electrode in the H+/H2 solution where H2 would be formed. It should be clear, though, that a stable solution of O2 does not exist; instead, the electrode potential is measured by indirect means. An oxidation potential refers to the reaction O2 O2 + e and the sign is reversed: +0.35 V.

94

Willem H. Koppenol and Patricia L. Bounds

would depend on the presence of redox-active transition metal ions, iron and copper, the topic of this Chapter. It is known that NO , O2 , and H2O2 are produced enzymatically, and the remaining species listed are formed from these (Figure 2). In phagocytic cells, i.e., neutrophils and macrophages, O2 generated by NADPH oxidase reacts to produce H2O2 (Reaction 1) and ONOO (Reaction 2) [38]. Also in neutrophils, H2O2 is a substrate in the reaction of myeloperoxidase to generate OCl (Reaction 3). The reactive ONOO and OCl synthesized in phagocytic cells are part of the host organism‟s defence arsenal to combat pathogenic organisms [39] and destroy foreign objects [40]. Reaction 4, the reaction of ONOO with carbon dioxide (CO2), yields mostly NO3 and CO2, but also the oxidizing radicals CO3 and NO2 [41, 42]. 2 O2 + 2H+

O2 + H2O2

(1)

O2 + NO

ONOO

(2)

H2O2 + Cl

OCl + H2O

(3)

ONOO + CO2

CO3 + NO2

(4)

Hasc–

asc – + H+

NTBI-Fe3+

NTBI-Fe2+

H2O2 + H+

H2O + HO Figure 1. Redox cycling by iron.

+

L-Arginine + 1.5 (NADPH + H ) + 2 O2

PROS

+

L-Citrulline + 1.5 (NADP ) + 2 H2O

NOS

CO2

NO

CO3 – + NO2

ONOOCO2–

ONOO–

CO2 + NO3–

NADPH + e– NADP– + H+

O2

NOX

O2



HO2 H

Figure 2. Generation and interconnectedness of PROS.

H2O2

SOD

+

O2



O2

OCl–

MPO

Cl–

H2 O

Redox-Active Metals: Iron and Copper

95

The electrode potential of O2 , E°(O2/O2 ) = –(0.18 + 0.02) V, or –(0.35 + 0.02) V (pO2 = 0.100 Mpa) indicates that O2 is only mildly reducing; O2 is, indeed, not very damaging [43]. Reaction 1, the disproportionation of O2 is catalyzed by superoxide dismutases [44] (SODs; see Vol. II, Chapter 5). At pH 7, O2 is 0.5% protonated (HO2 ), and the protonated form also acts as an oxidant: E° (HO2 ,H+/H2O2) = 1.04 V [45]. The electrode potentials indicate that the oxidation of O2 – by HO2 is thermodynamically likely, and the reaction indeed proceeds rapidly (k = 1.0 • 108 M 1s 1) [46] even in the absence of SOD. The importance of NO as a biological signaling molecule [47] has emerged in the past two decades. In our view, NO itself is not damaging per se aside from its affinity for iron(II) which may lead to inhibition of hemoproteins with an open coordination site. The very rapid reactions of NO with oxymyoglobin and oxyhemoglobin (k = 4.4 • 107 M 1s 1 and 8.9 • 107 M 1s 1, respectively, relative to heme concentration) produce the corresponding metmyoglobin and methemoglobin (and nitrate) [48] in which the heme iron is oxidized and cannot bind O2. In theory, NO2 can be formed via Reaction 5, the autoxidation of NO ; this reaction, however, requires two NO and one O2 [49-53], and the mechanism is complex [54]. Reaction 5 is likely to be slow (hours to days) in vivo, given the estimated physiological concentrations of NO and O2 [55]. In contrast, NO diffuses from tissue to the nearest red blood cell, where it is scavenged by hemoglobin more rapidly than it can react with O2 [56]. Thus, the lifetime of NO in vivo is determined by reactions with O2 and hemoglobin, and Reaction 5 can be neglected. 2 NO + O2

2NO2

(5)

The reaction of NO with O2 to form ONOO– (Reaction 2) is diffusion-controlled (k = 1.6 • 1010 M 1s 1) [57], and ONOO– is recognized as an important biological oxidant. At physiological pH, ONOO– is largely protonated (pKa = 6.8) [58]; ONOOH reacts much more slowly and selectively than HO [59] with biological targets. Evidence for formation of ONOOH in vivo is largely based on immunodetection of nitrotyrosine residues [60]; it should, however be noted that nitrotyrosine may also be formed by peroxidase-catalyzed reactions of nitrite and H2O2. Tyrosine nitration blocks phosphorylation and, thus, impacts on the signaling functions of tyrosine residues [61]. Furthermore, nitrotyrosine can, at least in vitro, accept an electron from ascorbate and then reduce O2 [62], thereby contributing to the formation of O2 –. The main reaction of ONOOH is isomerisation to NO3–, but it is widely believed that ONOOH undergoes homolysis to at least some extent to yield NO2 and HO [63-65]. Estimates of the yield of homolysis range from as low as ca. 1% [66] to as high as ca. 40% [67]. We have sought but failed to find evidence to support a mechanism in which homolysis accounts for more than a very low percentage of the decay of ONOOH [68-72]. The radicals produced in the reaction of ONOO– with CO2 [73], NO2 and CO3 [41, 42] are strongly oxidizing species: (E° (NO2 /NO2 ) = +1.04 V) [74] and (E° (CO3 /CO32 ) = +1.58 V) [75, 76]. These radicals have been implicated in ONOO–-mediated nucleic acid oxidation [77] and tyrosine nitration [78]. Finally, the OCl– produced by myeloperoxidase in activated phagocytic cells reacts with many biomolecules, including nucleotides, DNA and low-molecular-weight thiols [79], as well as protein thiol [80] and amino residues [81].

96

Willem H. Koppenol and Patricia L. Bounds

4. Redox Chemistry and Biochemistry of Iron: The Historical Context The involvement of transition metal ions in oxidative stress pathologies is often attributed to “Fenton chemistry”, and the Fenton reaction, reduction of H2O2 by iron(II) to generate HO (Reaction 6), has long been of interest to the free radical community. H. J. H. Fenton discovered the reaction that bears his name serendipitously [82]. While still an undergraduate student, Fenton was shown a violet coloured solution that had been obtained by a fellow student who was mixing reagents, including tartaric acid, at random, and later reproduced the reaction and published it as a test for tartaric acid in 1876 [83]. A full account, including the structure of the oxidized product, 2,3-dihydroxymaleic acid, was published in 1896 [84]. Although Fenton used iron(II) and H2O2 to modify organic compounds, he was not aware that HO• is formed in the reaction. Two years after Fenton‟s death, Fritz Haber and Richard M. Willstätter [85] published a paper on radical chain reactions in organic chemistry and biochemistry, in which they attributed the action of catalase on H2O2 to the initiation of a radical reaction (Reaction 6, the Fenton reaction) [86]. Later, Haber and his assistant Joseph Weiss [87, 88] investigated the decay of H2O2 by iron salts at low pH and concluded that, since more than 1 H2O2 per 2 Fe2+ is consumed when H2O2 is present in excess, the mechanism involves a radical chain reaction (Reactions 7 and 8), with chain termination by Reaction 9. Fe2+ + H2O2

Fe3+ + HO + HO•

(6)

HO• + H2O2

H2O + O2• + H+

(7)

O2• + H+ + H2O2 Fe2+ + HO• + H+

O2 + HO• + H2O Fe3+ + H2O

(8) (9)

By the late 1940s, the mechanism proposed by Haber and Weiss for the iron-catalyzed decomposition of H2O2 had been criticised by Philip George, who showed that O2•– does not react with H2O2 [89]. An alternative mechanism at low pH proposed by George and co-workers in 1949 [90] on the basis of product analyses is summarised in Table 1. Note that O2 occurs in these reactions as HO2 (pKa = 4.8) [43]. More extensive proposals for the mechanism were described in two publications in 1951 [91, 92], the first of which contains a reference to Fenton's first full report of the oxidation of tartaric acid [93]. The rate constants were determined later. The mechanism published by George and coworkers was corroborated in 1985 [94]. It should be noted that all of these investigations of iron-catalyzed decomposition of H2O2 were carried out at low pH and are generally not physiologically significant. A new field of study, free radical biochemistry, was created in the wake of the discovery in 1969 of an enzymatic function for hemocuprein [95], namely catalysis of the disproportionation of O2 (Reaction 1) [96]. The rate constant for the copper/zinc SOD (CuZnSOD) reaction, which is governed by electrostatic guidance of O2 to the active site [97, 98], indicates that the reaction is close to diffusion-controlled (kcat 1 2 109 M 1s 1) [99, 100]. The majority of

97

Redox-Active Metals: Iron and Copper

CuZnSOD, ca. 5–10 M, is found in the cytosol of the cell; some is also found in the mitochondrial intermembrane space [101]. In 1973, Fe- and Mn-containing SODs were discovered [102, 103]: the mitochondrial matrix contains manganese SOD (MnSOD), and bacteria contain a structurally related Fe-enzyme. A tetrameric CuZnSOD manages O2 in the extracellular space [104]. Given that Reaction 1 uncatalyzed is rapid (k = ca. 106 M 1s 1 at neutral pH), and that SODs are widespread, the conclusion is warranted that O2 is reactive and, therefore, must be dangerous. Unfortunately, George‟s work from the 1940s was not consulted, and, in the 1970s, reduction of H2O2 by O2 (Reaction 8) from the Haber-Wilstätter mechanism was proposed [105] as a possible harmful reaction from which the organism needed to be protected. At this point in history, Reaction 8, in which the relatively innocuous O2 is converted to the far more reactive HO radical, became known as the Haber-Weiss reaction#. Again, it had to be pointed out in other studies that Reaction 8 is too slow to be significant [106-111]. When it was realized that O2 does not reduce H2O2, iron complexes were invoked to act as catalysts, as in the mechanism of George et al., and Reactions 6 and 9 became known as the “Fenton-catalyzed Haber-Weiss reaction.” Later, it was realised and accepted that other reductants, e.g., monohydrogen ascorbate, are more likely candidates to reduce Fe3+ complexes in vivo.

5. Biological Relevance of the Fenton Reaction The toxicity of iron, and, by implication, copper, originates from its ability to reduce peroxides, via Fenton chemistry. Although George and coworkers provided a mechanism for the decay of H2O2 catalyzed by iron at low pH, it is not clear that the mechanism has any meaning at neutral pH where aqueous Fe3+ ions are not available for reaction because they have formed hydroxide, phosphate, or other complexes. If such Fe3+ complexes are present in solution, are they reduced by O2 , or monohydrogen ascorbate, and are the corresponding Fe2+ complexes likely to be oxidized by H2O2 or organic peroxides? Table 1. Rate constants for the reactions in the mechanism of iron-catalyzed decomposition of H2O2 at low pH [89] Reactions Fe2+ + H2O2 + H+ Fe3+ + H2O + HO• HO• + H2O2 H2O + HO2• Fe2+ + HO• + H+ Fe3+ + H2O 2+ + Fe + HO2 + H Fe3+ + H2O2 Fe3+ + HO2 Fe2+ + O2 + H+

#

Reaction No. 9 7 10 11 12

k (M 1s 1) 41.5 2.7 • 107 4.3 • 108 1.2 • 106 2.0 • 104 (pH 1)

It is not clear why Reactions 7 and 8 have become known as the Haber-Weiss cycle rather than the HaberWillstätter cycle; it may be because the often-quoted paper by Haber and Weiss is in English and, thus, more accessible to the international research community, while the language of the original paper of Haber and Willstätter is German. Haber and Willstätter 87 is cited by Haber and Weiss 88 but neither paper refers to Fenton 84, although Reaction 6 is the Fenton reaction.

98

Willem H. Koppenol and Patricia L. Bounds

The thermodynamic requirements for a metal complex in aqueous solution at neutral pH to redox-cycle are simple: when O2 – is the reductant, the metal-chelate must have a electrode potential between E°‟(O2/O2 –), –0.35 V and E°‟(H2O2/ OH,H2O), +0.39 V [5, 112]. The window of redox-opportunity (0.74 V) is therefore wide. For both iron and copper, redox cycling involves transfer of a single electron; thus, only one-electron electrode potentials may be considered. Potentials for two-electron oxidations, e.g., two-electron electrode potential of the NAD+, H+/NADH couple (–0.32 V) [113, 114], should never be considered in discussions of redox-cycling. Such a reaction is kinetically not feasible, as it would require the simultaneous collision of three reactants, NADH and two iron complexes, which is unlikely under the dilute concentrations found physiologically. Although O2 – has been shown to support hydroxylation of benzoic acid catalyzed by iron chelates [115], these studies were carried out with chelating agents that, with the exception of citrate [116], are not physiologically relevant. It is also important to recognize that, in most tissue compartments, the steady-state concentration of O2 – is maintained at a vanishingly low level by SODs [117]. The rate constant for reduction of iron(III)-edta† by O2 –, determined directly by pulse radiolysis, is ca. 1 • 106 M 1s 1 at physiological pH [118], three orders of magnitude smaller than the catalytic rate constant of CuZnSOD; rate constants for the reaction of iron(III)-dtpa† and -atp were much less than 1 • 106 M 1s 1 and could not be determined. Not surprisingly, it is now felt that monohydrogen ascorbate is a more likely physiological reductant of iron(III) complexes. Given that the electrode potential of the asc /Hasc couple is +0.28 V [119], the redox window for redox cycling involving ascorbate is much smaller, i.e., 0.11 V. However, for a typical serum concentration of 50 M monohydrogen ascorbate, the concentration of the ascorbyl radical is likely to be in the nanomolar range. Thus, the electrode potential for the ascorbyl/monohydrogen ascorbate couple is, according to the Nernst equation, ca. 0.18 V less than +0.28 V, which would widen the redox window to 0.29 V. Similar considerations of the H2O2,H+/HO , H2O couple increases the electrode potential from +0.39 V to ca. +0.9 V, an adjustment that results in a much larger window of 0.8 V, from +0.1 V to +0.9 V [120]. A number of studies with iron complexed to citrate, atp or aminopolycarbonates, such as edta, dtpa and nta†, suggest that the Fenton reaction, and its equivalent with organic peroxides, is thermodynamically feasible [121] and does proceed. Rate constants of the order of 102 to 105 M 1s 1 have been reported for the Fenton reactions of these various iron(II) complexes [122131]; there is, however, no good agreement between the published rate constants for a given complex. The reaction, which requires binding of H2O2 to iron(II), generally proceeds faster when more exchangeable water molecules are bound to iron(II) [132], i.e., when more ligand sites on the iron are available. However, all these reactions are much slower than that of H2O2 with catalase, and, thus, not likely to be physiologically significant.**

**

It is not sufficient to simply compare rate constants; the relative concentrations of the Fe2+ complex and catalase must be considered as well. It is the product of the concentration of the iron complexes or catalase with the respective rate constants for the reaction with H2O2 that must be compared. Catalase is present at ca. 2 M, and the concentration of redox-active iron is, at most, a few M. In the following, the concentration of the Fe2+ complex 5 M is used as an example to demonstrate that the Fenton reaction with Fe2+ complexes is outcompeted by catalase. This example is, of course, valid only for homogeneous solution: kFenton[Fe2+] = (1 • 104 M 1s 1)(5 • 10 6 M) = 0.05 s 1

Redox-Active Metals: Iron and Copper

99

In vitro studies performed under physiological conditions indicate that HO• is not the only oxidant formed during the Fenton reaction [121, 133-135]. Depending on the conditions, the oxidant may be (1) oxidoiron(IV), a higher oxidation state of iron, as suggested during the 1930s (Reaction 13) [136], (2) H2O2 bound to iron(III), or (3) HO•. The pattern of hydroxylation of salicylate by iron(II)-edta and H2O2 compared to that by HO• generated by radiolysis of water suggests that HO• formation is likely [137]. Isotope/ESR studies established that the oxygen in either the HO• or oxidoiron(IV) is from H2O2 [138]. Fe2+ + H2O2

FeO2+ + H2O

(13)

What is the likelihood that the Fenton reaction occurs in vivo? Assuming that monohydrogen ascorbate is the source of electrons in vivo, there are number of problems with the hypothesis that iron is a catalyst in oxidative processes: (1) We do not know the composition/structure of the iron complex (or complexes) in vivo reduced by monohydrogen ascorbate and subsequently oxidized during the Fenton reaction, nor do we know any rate constants for the reactions. (2) The rate constants for the Fenton reaction (102 105 M 1s 1) that have been determined are remarkably unimpressive when compared to that for the reaction of catalase with H2O2 (3.5 107 M 1s 1) [139]. Additionally, H2O2 is scavenged by glutathione peroxidase [140] and peroxiredoxin [141]. (3) The Fenton reaction is even slower in the presence the NO , which is present in vivo and which binds to iron(II) [142]. (4) The hypothesis does not include a protective role for SOD, but only for proteins that remove H2O2, which is contrary to experimental observations that SOD alone can protect tissues from oxyradical damage. Given that most biomolecules are unreactive toward O2 –, concentrations of ca. 5–10 M CuZnSOD in the cytosol would seem higher than necessary to protect against damage from O2 –. Clearly, SOD does protect, but not from Fenton chemistry. Beckman et al. [143] wrote in 1990 that “generation of strong oxidants by the iron-catalyzed Haber-Weiss reaction is not an entirely satisfactory explanation for SOD-inhibitable injury in vivo” and proposed that NO reacts with O2 – to form ONOO–, and that formation of ONOOH is kinetically far more feasible than the “iron-catalyzed Haber-Weiss reaction.” Earlier in this Chapter, we described ONOO as an important and selective oxidizing and nitrating species. Thus, SOD protects by eliminating O2 – and prevents, thereby, formation of ONOO . The diffusion-controlled rate constant for the reaction of O2 – with NO [57] necessitates the micromolar concentration of SOD observed in vivo. Where, then, may the Fenton reaction play a role? In the acidic environment of the lysosome, where ferritin is decomposed, the pH is between 4 and 5; here, the concentration of chelatable iron is relatively high (ca. 15 M), catalase is absent, the coordination of iron is altered and the iron may be reduced due to the presence of cysteine. Under conditions of oxidative stress, lysosomes may be damaged and ruptured, which would release relatively high concentrations of redox-active iron [144, 145] into the cytosol (Vol. II, Chapter 16). Ferritins released into the bloodstream during inflammation and autoimmune diseases may contribute to locally high iron concentrations that lead to apoptosis [146]. Fenton chemistry-derived oxidative damage can be prevented by chelation of the iron with desferrioxamine (dfo) [146-148].

kcatalase[catalase]

= (3.5 • 107 M 1s 1)(2 • 10 6 M) = 70 s

1

100

Willem H. Koppenol and Patricia L. Bounds

The extracellular NTBI levels may contribute to intracellular labile iron pool (LIP) levels (Vol. II, Chapter 19) and play a role in oxyradical damage [149], or be a symptom of iron overload. Any form of mobile iron poses a threat to the liver and the heart, where excess iron accumulates, and iron chelation therapy is employed to promote excretion. The „wish list‟ for an effective iron chelator has been described as: „stable in vivo, non-toxic, rapidly excreted, preferably orally-active, and produced with a low synthetic cost‟ [150]. Toxicity is ascribed to Fenton chemistry, thus, the chelate ought not be redox-active: the three chelating agents used clinically, dfo, cp20 and icl670, bind iron(III) far more tightly than iron(II), which results in very negative standard electrode potentials ( 0.45 to 0.60 V [120, 151-153]) well outside the window of redox opportunity, such that monohydrogen ascorbate cannot reduce these iron(III) complexes. The low electrode potentials also imply that dfo, cp20 and icl670 bind iron(II) at micromolar concentrations only in part or not at all, which has been experimentally verified for cp20 [154]: when iron(II) is oxidized, it is complexed by cp20, thereby preventing reduction and redox cycling. Although copper is less ubiquitous than iron in living systems, copper and iron proteins have similar functions, including electron transfer and O2 binding and activation. In humans, copper is used for electron transfer reactions in the enzymes CuZnSOD and cytochrome c oxidase. During the 1980s, it was recognized and demonstrated that free copper ions can, in theory, redox cycle and promote reactions analogous to those catalyzed by free iron [155158]. In in vitro studies under acidic conditions, copper(I) was shown to form an intermediate Cu+H2O2, which reacts with organic scavengers, however, not by a Fenton mechanism [159]. Other investigators [160] concluded on the basis of pH and substrate concentration dependence experiments that no HO is formed during the reaction of copper(I) with H2O2. Studies with red blood cells show that monohydrogen ascorbate can reduce copper complexes to initiate reactions similar to those described for iron [161]. However, the concentration of free copper ions in the cell is extremely low, only ca. 1 copper ion per cell [162], and the possibility that copper participates in oxyradical damage reactions in vivo is vanishingly small. The disproportionation of O2 – catalyzed by free copper ions is faster by a factor of 4 than that catalyzed by SOD [163, 164]; the inclusion of edta to bind free copper(II) prevents the reaction with O2 –. It is important to note that edta, nta and dtpa, common metal chelation agents that are often used as models for physiological low-molecular-weight iron complexes in biochemical studies, may enhance the Fenton activity of iron at neutral pH. Depending on the complex, one or more coordination sites of iron are occupied by water, which may give way to H2O2 [132]. In the case of copper, the complexes formed with these ligands are essentially inert at physiological pH, and redox chemistry is prevented. Thus, addition of chelators can produce differential effects on redox behavior depending on the target metal, e.g., edta and nta enhance iron-mediated damage to chromatin but protect against copper-mediated damage [165]. Similar differential effects of edta on oxidative reactivity in cerebrospinal fluid and brain tissue was recently reported [166], and these authors correctly point out that different chelation strategies are required for sequestration of iron – “the main characteristics of an iron sequestrant should be its ability to compete with metal ligands naturally present in CSF” – and copper – “copper requires application of a strong and efficient chelator”.

Redox-Active Metals: Iron and Copper

101

6. Cautionary Remarks Given the low micromolar concentrations of iron and H2O2 found under normal physiological conditions and the rate constants for Fenton reactions with iron bound to physiologically relevant ligands, it is kinetically unlikely that iron causes damage, and the same may hold for copper. Iron and copper may contribute to oxidative processes under conditions of higher local concentrations of these ions. The destruction of ferritin (and by extension ceruloplasmin) in lysosomes may provide such concentrations. NTBI in blood is an indicator of excess iron; this form of iron does not appear to be very redox-active, but can be considered dangerous since it is a form of mobile iron that may be deposited in the liver or in the heart. To determine the rate constant for a reaction of physiological interest, it is usually necessary to use reagents at concentrations that are significantly higher than physiological conditions; the rate constant thus determined can be extrapolated to the dilute conditions found in vivo to estimate how fast the process is in reality. However, a rate constant determined in vitro is valid for homogeneous solution, whereas the many reactions going on inside the cell take place under more heterogeneous conditions. Thus, the observation that common HO∙ scavengers fail to protect against oxidative damage in vivo has been attributed to formation of HO∙ at locations within cellular components that are poorly accessible to the scavenger. Alternatively, formation of oxidoiron(IV), which might react less rapidly than HO∙ with scavengers, has been postulated. Results obtained from experiments performed with cells to which relatively high concentrations of metal ions and/or H2O2 have been added cannot be realistically extended to prove that oxidative damage occurs under normal physiological conditions, where the antioxidant defence system consisting of monohydrogen ascorbate, glutathione, vitamin E, SODs, catalase, glutathione peroxidase, peroxiredoxin, etc. is intact. Some in vitro experiments can be additionally criticized for having used simple iron salt solutions: e.g., iron(III) salt solutions are very acidic, and, at neutral pH, colloids of iron(III) hydroxide/phosphate and iron(III)oxide, which may simply be toxic rather than catalysts of redox reactions, are formed. Iron(II) autoxidizes rapidly at pH 7 to generate O2∙– and H2O2, which are likely to complicate the interpretation of findings. In short, the speciation of a metal ion determines its thermodynamic and kinetic properties and, by extension, its reactivity. One obstacle to progress in the field of redox metal toxicity is the inability to follow oxidative stress in cells in a time-resolved fashion. Although widely used fluorescence techniques have been touted to do so, the addition of the fluorescent indicator itself may lead to the release of PROS [167, 168].

Conclusions Are redox-active metal ions causative agents in disease, or do diseases cause redox-active metal ions to accumulate? Although metals, especially iron and copper, are clearly implicated in a variety of oxidative-stress-related pathologies, it has not yet been definitively demonstrated whether low-molecular-weight metal ion complexes act to initiate damage or

102

Willem H. Koppenol and Patricia L. Bounds

are simply liberated as a sequela of oxyradical damage, and it is possible that metals, once released as a result of an oxidative insult, act in a cascade fashion to propagate injury. It is the findings from in vivo studies that most clearly support a role for involvement of redox-active metals in oxidative stress-related pathologies; results from in vitro mechanistic studies based on kinetics have sometimes called those findings into question. Both approaches to the study of free radical biology and medicine have merit and contribute to the ultimate goal of elucidating mechanisms that make sense at the molecular, cellular and organismic levels alike.

References [1] [2]

[3] [4]

[5] [6] [7]

[8] [9]

[10] [11]

[12] [13] [14]

Koppenol WH. Names for inorganic radicals (IUPAC recommendations 2000). Pure Appl Chem 2000; 72:437-446. Connelly NG, Damhus T, Hartshorn RM, Hutton AT. Nomenclature of Inorganic Chemistry. IUPAC Recommendations 2005. Cambridge: Royal Society of Chemistry, 2005. George P. The fitness of oxygen. In: King TE, Mason HS, Morrison M, eds. Oxidases and Related Redox Systems. New York: Wiley, 1965: 3-36. Koppenol WH. The paradox of oxygen: Thermodynamics versus Toxicity. In: King T, Mason HS, Morrison M, eds. Oxidases and Related Redox Systems. New York: A. Liss, 1988: 93-109. Koppenol WH. Oxygen activation by cytochrome P450: A thermodynamic analysis. J Am Chem Soc 2007; 129:9686-9690. Ferguson-Miller S, Babcock GT. Heme/copper terminal oxidases. Chemical Reviews 1996; 96:2889-2907. Pushkar Y, Yano J, Sauer K, Boussac A, Yachandra VK. Structural changes in the Mn4Ca cluster and the mechanism of photosynthetic water splitting. Proc Natl Acad Sci 2008; 105:1879-1884. Misra HP, Fridovich I. The generation of superoxide radical during the autoxidation of hemoglobin. J Biol Chem 1972; 247:6960-6962. Wever R, Oudega B, van Gelder BF, Amsterdam GU. Generation of superoxide radicals during the autoxidation of mammalian oxyhemoglobin. Biochim Biophys Acta 1973; 302:457-458. Gotoh T, Shikama K. Generation of the superoxide radical during autoxidation of oxymyoglobin. J Biochem 1976; 80:397-399. Kelley EE, Khoo NKH, Hundley NJ, Malik UZ, Freeman BA, Tarpey MM. Hydrogen peroxide is the major oxidant product of xanthine oxidase. Free Radic Biol Med 2010; 48:493-498. Zhang Z, Sibbesen O, Johnson RA, Ortiz de Montellano PR. The substrate specificity of cytochrome P450cam. Bioorg Med Chem Lett 1998; 6:1501-1508. Loschen G, Azzi A, Richter C, Flohé L. Superoxide radicals as precursors of mitochondrial hydrogen peroxide. FEBS Lett 1974; 42:68-72. Chance B, Sies H, Boveris A. Hydroperoxide metabolism in mammalian organs. Physiol Rev 1979; 59:527-605.

Redox-Active Metals: Iron and Copper

103

[15] Halliwell B, Gutteridge JMC. Free Radicals in Biology and Medicine. 4th ed. Oxford: Oxford University Press, 2007. [16] Pacher P, Beckman JS, Liaudet L. Nitric oxide and peroxynitrite in health and disease. Physiol Rev 2007; 87:315-424. [17] Harrison PM, Arosio P. Ferritins: molecular properties, iron storage function and cellular regulation. Biochim Biophys Acta 1996; 1275:161-203. [18] Mackenzie EL, Iwasaki K, Tsuji Y. Intracellular iron transport and storage: From molecular mechanisms to health implications. Antioxid Redox Signal 2008; 10:9971030. [19] Gutteridge JMC, Stocks J. Ceruloplasmin - Physiological and Pathological Perspectives. CRC Crit Rev Clin Lab Sci 1981; 14:257-329. [20] Puntarulo S. Iron, oxidative stress and human health. Mol Asp Med 2005; 26:299312. [21] Squitti R, Quattrocchi CC, Salustri C, Rossini PM. Ceruloplasmin fragmentation is implicated in 'free' copper deregulation of Alzheimer's disease. Prion 2008; 2:23-27. [22] Brewer GJ. Risks of copper and iron toxicity during aging in humans. Chem Res Toxicol 2010; 23:319-326. [23] Wood JC. Diagnosis and management of transfusion iron overload: The role of imaging. J Hematol 2007; 82:1132-1135. [24] de Valk B, Addicks MA, Gosriwatana I, Lu S, Hider RC, Marx JJM. Nontransferrin-bound iron is present in serum of hereditary haemochromatosis heterozygotes. Eur J Clin Invest 2000; 30:248-251. [25] Ahmed NK, Hanna M, Wang W. Nontransferrin-bound serum iron in thalassemia and sickle cell patients. Int J Biochem 1986; 18:953-956. [26] Grootveld M, Bell JD, Halliwell B, Aruoma OI, Bomford A, Sadler PJ. Nontransferrin-bound iron in plasma or serum from patients with idiopathic hemochromatosis. J Biol Chem 1989; 264:4417-4429. [27] Evans RW, Rafique R, Zarea A, Rapisarda C, Cammack R, Evans PJ et al. Nature of non-transferrin-bound iron: studies on iron citrate complexes and thalassemic sera. J Biol Inorg Chem 2008; 13:57-74. [28] Breuer W, Hershko C, Cabantchik ZI. The importance of non-transferrin bound iron in disorders of iron metabolism. Transf Sci 2000; 23:185-192. [29] Gautier-Luneau I, Merle C, Phanon D, Lebrun C, Biaso F, Serratrice G et al. New trends in the chemistry of iron(III) citrate complexes: Correlations between X-ray structures and solution species probed by electrospray mass spectrometry and kinetics of iron uptake from citrate by iron chelators. Chem Eur J 2005; 11:22072219. [30] Silva AMN, Kong XL, Parkin MC, Cammack R, Hider RC. Iron(III) citrate speciation in aqueous solution. Dalton 2009;8616-8625. [31] Varadarajan S, Yatin S, Aksenova M, Butterfield DA. Review: Alzheimer's amyloid b-peptide-associated free radical oxidative stress and neurotoxicity. J Struct Biol 2000; 130:184-208. [32] Lovell MA, Robertson JD, Teesdale WJ, Campbell JL, Markesbery WR. Copper, iron and zinc in Alzheimer's disease senile plagues. J Neurol Sci 1998; 158:47-52. [33] Nelson TJ, Alkon DL. Oxidation of cholesterol by amyloid precursor protein and bamyloid peptide. J Biol Chem 2005; 280:7377-7387.

104

Willem H. Koppenol and Patricia L. Bounds

[34] Wink DA, Mitchell JB. Chemical biology of nitric oxide: Insights into regulatory, cytotoxic, and cytoprotective mechanisms of nitric oxide. Free Radic Biol Med 1998; 25:434-456. [35] Bredt DS. Endogenous nitric oxide synthesis: Biological functions and pathophysiology. Free Rad Res 1999; 31:577-596. [36] Schwarz HA, Dodson RW. Equilibrium between hydroxyl radicals and thallium (II) and the oxidation potential of OH(aq). J Phys Chem 1984; 88:3643-3647. [37] Buxton GV, Greenstock CL, Helman WP, Ross AB. Critical review of rate constants for reactions of hydrated electrons, hydrogen atoms and hydroxyl radicals (•OH/•O ) in aqueous solution. J Phys Chem Ref Data 1988; 17:513-886. [38] Bastian NR, Hibbs JB, Jr. Assembly and regulation of NADPH oxidase and nitric oxide synthase. Curr Opin Immunol 1994; 6:131-139. [39] Hurst JK, Lymar SV. Cellularly generated inorganic oxidants as natural microbicidal agents. Acc Chem Res 1999; 32:520-528. [40] Sutherland K, Mahoney JR, Coury AJ, Eaton JW. Degradation of biomaterials by phagocyte-derived oxidants. J Clin Invest 1993; 92:2360-2367. [41] Meli R, Nauser T, Koppenol WH. Direct observation of intermediates in the reaction of peroxynitrite with carbon dioxide. Helv Chim Acta 1999; 82:722-725. [42] Bonini MG, Radi R, Ferrer-Sueta G, Ferreira AMD, Augusto O. Direct EPR detection of the carbonate radical anion produced from peroxynitrite and carbon dioxide. J Biol Chem 1999; 274:10802-10806. [43] Bielski BHJ, Cabelli DE, Arudi RL, Ross AB. Reactivity of HO 2/O2 radicals in aqueous solution. J Phys Chem Ref Data 1985; 14:1041-1100. [44] McCord JM, Fridovich I. Superoxide dismutase. An enzymic function for erythrocuprein (hemocuprein). J Biol Chem 1969; 244:6049-6055. [45] Koppenol WH, Stanbury DM, Bounds PL. Electrode potentials of partially reduced oxygen species, from dioxygen to water. Free Radic Biol Med. In press. [46] Bielski BHJ. Re-evaluation of the spectral and kinetic properties of HO2 and O2 free radicals. Photochem Photobiol 1978; 28:645-649. [47] Moncada S, Palmer RMJ, Higgs EA. Nitric oxide: Physiology, pathophysiology, and pharmacology. Pharmacol Rev 1991; 43:109-142. [48] Herold S, Exner M, Nauser T. Kinetic and mechanistic studies of the NO•-mediated oxidation of oxymyoglobin and oxyhemoglobin. Biochemistry 2001; 40:3385-3395. [49] Bodenstein M. Die Geschwindigkeit der Reaktion zwischen Stickoxyd und Sauerstoff. Z Elektrochem 1918; 24:183-220. [50] Zang V, Kotowski M, van Eldik R. Kinetics and mechanism of the formation of FeII(edta)NO in the system FeII(edta)/NO/HONO/NO2- in aqueous solutions. Inorg Chem 1988; 27:3279-3283. [51] Awad HH, Stanbury DM. Autoxidation of NO in aqueous solution. Int J Chem Kinet 1993; 25:375-381. [52] Kharitonov VG, Sundquist AR, Sharma VS. Kinetics of nitric oxide autoxidation in aqueous solution. J Biol Chem 1994; 269:5881-5883. [53] Pires M, Rossi MJ, Ross DS. Kinetics and mechanistic aspects of the NO oxidation by O2 in aqueous phase. Int J Chem Kinet 1994; 26:1207-1227.

Redox-Active Metals: Iron and Copper

105

[54] Galliker B, Kissner R, Nauser T, Koppenol WH. Intermediates in the autoxidation of nitrogen monoxide. Chem Eur J 2009; 15:6161-6168. [55] Koppenol WH. The chemistry and biochemistry of oxidants derived from nitric oxide. In: Weir EK, Archer SL, Reeves JT, eds. Nitric Oxide and Radicals in the Pulmonary Vasculature. Armonk, NY: Futura Publishing Co., Inc., 1996: 355-362. [56] Beckman JS, Koppenol WH. Nitric oxide, superoxide, and peroxynitrite: The good, the bad, and the ugly. Am J Physiol Cell Physiol 1996; 271:C1424-C1437. [57] Nauser T, Koppenol WH. The rate constant of the reaction of superoxide with nitrogen monoxide: Approaching the diffusion limit. J Phys Chem A 2002; 106:4084-4086. [58] Koppenol WH, Kissner R. Can O=NOOH undergo homolysis? Chem Res Toxicol 1998; 11:87-90. [59] Koppenol WH, Moreno JJ, Pryor WA, Ischiropoulos H, Beckman JS. Peroxynitrite, a cloaked oxidant formed by nitric oxide and superoxide. Chem Res Toxicol 1992; 5:834-842. [60] Ye YZ, Strong M, Huang ZQ, Beckman JS. Antibodies that recognize nitrotyrosine. Methods Enzymol 1996; 269:201-209. [61] Mondoro TH, Shafer BC, Vostal JG. Peroxynitrite-induced tyrosine nitration and phosphorylation in human platelets. Free Radic Biol Med 1997; 22:1055-1063. [62] Pruetz WA. Nitro-tyrosine as promoter of free radical damage in a DNA model system. Free Radic Res Commun 1986; 2:77-83. [63] Coddington JW, Hurst JK, Lymar SV. Hydroxyl radical formation during peroxynitrous acid decomposition. J Am Chem Soc 1999; 121:2438-2443. [64] Gerasimov OV, Lymar SV. The yield of hydroxyl radical from the decomposition of peroxynitrous acid. Inorg Chem 1999; 38:4317-4321. [65] Goldstein S, Lind J, Merényi G. Chemistry of peroxynitrites as compared to peroxynitrates. Chem Rev 2005; 105:2457-2470. [66] Van der Vliet A, O'Neill CA, Halliwell B, Cross CE, Kaur H. Aromatic hydroxylation and nitration of phenylalanine and tyrosine by peroxynitrite. FEBS Lett 1994; 339:89-92. [67] Yang G, Candy TEG, Boaro M, Wilkin HE, Jones P, Nazhat NB et al. Free radical yields from the homolysis of peroxynitrous acid. Free Radic Biol Med 1992; 12:327330. [68] Padmaja S, Kissner R, Bounds PL, Koppenol WH. Hydrogen isotope effect on the isomerization of peroxynitrous acid. Helv Chim Acta 1998; 81:1201-1206. [69] Kissner R, Nauser T, Kurz C, Koppenol WH. Peroxynitrous acid – Where is the hydroxyl radical? IUBMB Life 2003; 55:567-572. [70] Kurz C, Zeng X, Hannemann S, Kissner R, Koppenol WH. On the chemical and electrochemical one-electron reduction of peroxynitrous acid. J Phys Chem A 2005; 109:965-969. [71] Maurer P, Thomas CF, Kissner R, Rüegger H, Greter O, Röthlisberger U et al. Oxidation of nitrite by peroxynitrous acid. J Phys Chem A 2003; 107:1763-1769. [72] Gupta D, Harish B, Kissner R, Koppenol WH. Peroxynitrate is rapidly formed during decompostion of peroxynitrite at neutral pH. Dalton 2009;(29):5730-5736. [73] Lymar SV, Hurst JK. Rapid reaction between peroxonitrite ion and carbon dioxide: Implications for biological activity. J Am Chem Soc 1995; 117:8867-8868.

106

Willem H. Koppenol and Patricia L. Bounds

[74] Stanbury DM. Reduction potentials involving inorganic free radicals in aqueous solution. Adv Inorg Chem 1989; 33:69-138. [75] Huie RE, Clifton CL, Neta P. Electron transfer reaction rates and equilibria of the carbonate and sulfate radical anions. Radiat Phys Chem 1991; 38:477-481. [76] Lymar SV, Schwarz HA, Czapski G. Medium effects on reactions of the carbonate radical with thiocyanate, iodide, and ferrocyanide ions. Radiat Phys Chem 2000; 59:387-392. [77] Lee YA, Yun BH, Kim SK, Margolin Y, Dedon PC, Geacintov NE et al. Mechanisms of oxidation of guanine in DNA by carbonate radical anion, a decomposition product of nitrosoperoxycarbonate. Chem Eur J 2007; 4571:45714581. [78] Gow A, Duran D, Thom SR, Ischiropoulos H. Carbon dioxide enhancement of peroxynitrite-mediated protein tyrosine nitration. Arch Biochem Biophys 1996; 333:42-48. [79] Pruetz WA. Hypochlorous acid interactions with thiols, nucleotides, DNA, and other biological substrates. Arch Biochem Biophys 1996; 332:110-120. [80] 80. Carr AC, Winterbourn CC. Oxidation of neutrophil glutathione and protein thiols by myeloperoxidase-derived hypochlorous acid. Biochem J 1997; 327:275-281. [81] Hawkins CL, Davies MJ. Hypochlorite-induced oxidation of proteins in plasma: formation of chloramines and nitrogen-centred radicals and their role in protein fragmentation. Biochem J 1999; 340:539-548. [82] Koppenol WH. The centennial of the Fenton reaction. Free Radic Biol Med 1993; 15:645-651. [83] Fenton HJH. On a new reaction of tartaric acid. Chem News 1876; 33:190. [84] Fenton HJH. Constitution of a new dibasic acid, resulting from the oxidation of tartaric acid. J Chem Soc Trans 1896; 69:546-562. [85] Haber F, Willstätter R. Unpaarigheit und Radikalketten im Reaktion-Mechanismus organischer und enzymatischer Vorgänge. Chem Ber 1931; 64:2844-2856. [86] Koppenol WH. The Haber-Weiss cycle - 70 years later. Redox Rep 2001; 6(4):229234. [87] 87, Haber F, Weiss J. Über die Katalyse des Hydroperoxydes. Naturwiss 1932; 51:948-950. [88] Haber F, Weiss J. The catalytic decomposition of hydrogen peroxide by iron salts. Proc R Soc London 1934; 147:332-351. [89] George P. Some experiments on the reactions of potassium superoxide in aqueous solutions. Disc Faraday Soc 1947; 2:196-205. [90] Barb WG, Baxendale JH, George P, Hargrave KR. Reactions of ferrous and ferric ions with hydrogen peroxide. Nature 1949; 163:692-694. [91] Barb WG, Baxendale JH, George P, Hargrave KR. Reactions of ferrous and ferric ions with hydrogen peroxide. Part I. The ferrous ion reaction. Trans Faraday Soc 1951; 47:462-500. [92] Barb WG, Baxendale JH, George P, Hargrave KR. Reactions of ferrous and ferric ions with hydrogen peroxide. II. The ferric ion reaction. Trans Faraday Soc 1951; 47:591-616. [93] Fenton HJH. Oxidation of tartaric acid in the presence of iron. J Chem Soc Trans 1894; 65:899-910.

Redox-Active Metals: Iron and Copper

107

[94] Rush JD, Bielski BHJ. Pulse radiolytic studies of the reactions of hydrodioxyl/ peroxide with Fe(II)/Fe(III) ions. The reactivity of hydrodioxyl/superoxide with ferric ions and its implication on the occurrence of the Haber-Weiss reaction. J Phys Chem 1985; 89:5062-5066. [95] Mann T, Keilin D. Haemocuprein, a copper-protein compound of red blood corpuscles. Nature 1938; 142:148. [96] McCord JM, Fridovich I. Superoxide dismutase: The first twenty years (1968-1988). Free Radic Biol Med 1988; 5:363-369. [97] Koppenol WH. The physiological role of the charge distribution on superoxide dismutase. In: Rodgers MAJ, Powers EL, eds. Oxygen and Oxyradicals in Chemistry and Biology. New York: Academic Press, 1981: 671-674. [98] Getzoff ED, Tainer JA, Weinter PK, Kollman PA, Richardson JS, Richardson C. Electrostatic recognition between superoxide and copper, zinc superoxide dismutase. Nature 1983; 306:287-290. [99] Klug-Roth D, Fridovich I, Rabani J. Pulse radiolysis investigations of superoxide catalyzed disproportionation. Mechanism for bovine superoxide dismutase. J Am Chem Soc 1973; 95:2786-2791. [100] Fielden EM, Roberts PB, Bray RC, Lowe DJ, Mautner GN, Rotilio G et al. The mechanism of action of superoxide dismutase from pulse radiolysis and electron paramagnetic resonance. Biochem J 1974; 139:49-60. [101] Okado-Matsumoto A, Fridovich I. Subcellular distribution of superoxide dismutases (SOD) in rat liver - Cu,Zn-SOD in mitochondria. J Biol Chem 2001; 276:3838838393. [102] Weisiger RA, Fridovich I. Superoxide dismutase: organelle specificity. J Biol Chem 1973; 248:3582-3592. [103] Yost Jr FJ, Fridovich I. An iron-containing superoxide dismutase from Escherichia coli. J Biol Chem 1973; 248:4905-4908. [104] Marklund SL. Human copper-containing superoxide dismutase of high molecular weight. Proc Natl Acad Sci USA 1982; 79:7634-7638. [105] Beauchamp C, Fridovich I. A mechanism for the production of ethylene from methional. J Biol Chem 1970; 245:4641-4646. [106] Ferradini C, Seide C. Radiolyse de solutions acides et aérées de peroxyde d'hydrogène. Int J Radiat Phys Chem 1969; 1:219-224. [107] Rigo A, Stevanato R, Finazzi-Agro A, Rotilio G. An attempt to evaluate the rate of the Haber Weiss reaction by using hydroxyl radical scavengers. FEBS Lett 1977; 80:130-132. [108] Ferradini C, Foos J, Houée C, Pucheault J. The reaction between superoxide anion and hydrogen peroxide. Photochem Photobiol 1978; 28:697-700. [109] Koppenol WH, Butler J, van Leeuwen JW. The Haber-Weiss cycle. Photochem Photobiol 1978; 28:655-660. [110] Melhuish WH, Sutton HC. Study of the Haber-Weiss reaction using a sensitive method for detection of OH radicals. J Chem Soc Chem Commun 1978;970-971. [111] Weinstein J, Bielski BHJ. Kinetics of the interaction of HO2 and O2- radicals with hydrogen peroxide; the Haber-Weiss reaction. J Am Chem Soc 1979; 101:58-62.

108

Willem H. Koppenol and Patricia L. Bounds

[112] Koppenol WH. Thermodynamics of Fenton-driven Haber-Weiss and related reactions. In: Cohen G, Greenwald RA, eds. Oxy Radicals and Their Scavenging Systems. Vol. I. Molecular Aspects. New York: Elsevier Biomedical, 1983: 84-88. [113] Pierre JL, Fontecave M. Iron and activated oxygen species in biology: The basic chemistry. BioMetals 1999; 12:195-199. [114] Pierre JL, Fontecave M, Crichton RR. Chemistry for an essential biological process: the reduction of ferric ion. BioMetals 2002; 15:341-346. [115] Baker MS, Gebicki JM. The effect of pH on the conversion of superoxide to hydroxyl free radicals. Arch Biochem Biophys 1984; 234:258-264. [116] Baker MS, Gebicki JM. The effect of pH on yields of hydroxyl radicals produced from superoxide by potential biological iron chelators. Arch Biochem Biophys 1986; 246:581-588. [117] Koppenol WH, Bounds PL. Hormesis. Science 1989; 246:311. [118] Butler J, Halliwell B. Reaction of iron-EDTA chelates with the superoxide radical. Arch Biochem Biophys 1982; 218:174-178. [119] Williams MH, Yandell JK. Outer-sphere electron transfer reactions of ascorbate anions. Aust J Chem 1982; 35:1133-1144. [120] Merkofer M, Kissner R, Hider RC, Brunk UT, Koppenol WH. Fenton chemistry and iron chelation under physiological relevant conditions: Electrochemistry and kinetics. Chem Res Toxicol 2006; 19:1263-1269. [121] Koppenol WH. Chemistry of iron and copper in radical reactions. In: Rice-Evans CA, Burdon RH, eds. Free Radical Damage and its Control. Amsterdam: Elsevier Science B.V., 1994: 3-24. [122] Sutton HC, Winterbourn CC. Chelated iron-catalyzed OH. formation from paraquat radicals and H2O2:mechanism of formate oxidation. Arch Biochem Biophys 1984; 235:106-115. [123] Rahhal S, Richter HW. Reduction of hydrogen peroxide by the ferrous iron chelate of diethylenetriamine-N,N,N',N",N"-pentaacetate. J Am Chem Soc 1988; 110:31263133. [124] Wink DA, Nims RW, Desrosiers MF, Ford PF, Keefer LK. A kinetic investigation of intermediates formed during the Fenton reagent mediated degradation of Nnitrosodimethylamine: Evidence for an oxidative pathway not involving hydroxyl radical. Chem Res Toxicol 1991; 4:510-512. [125] Yamazaki I, Piette LH. ESR spin-trapping studies on the reaction of Fe2+ ions with H2O2-reactive species in oxygen toxicity in biology. J Biol Chem 1990; 265:1358913594. [126] Rush JD, Koppenol WH. The reaction between ferrous aminopolycarboxylate complexes and hydrogen peroxide. An investigation of the reaction intermediates by stopped-flow spectrophotometry. J Inorg Biochem 1987; 29:199-215. [127] Rush JD, Koppenol WH. Reactions of FeIInta and FeIIIedda with hydrogen peroxide. J Am Chem Soc 1988; 110:4957-4963. [128] Rush JD, Maskos Z, Koppenol WH. Reactions of iron(II)nucleotide complexes with hydrogen peroxide. FEBS Lett 1990; 261:121-123. [129] Gilbert BC, Jeff M. Oxidative damage and radical repair: One-electron transfer reactions involving metal ions, peroxides and free radicals. In: Rice-Evans C,

Redox-Active Metals: Iron and Copper

[130]

[131] [132] [133]

[134]

[135] [136] [137]

[138] [139] [140]

[141]

[142] [143]

[144]

[145]

109

Dormandy T, eds. Free Radicals: Chemistry, Pathology and Medicine. London: The Richelieu Press, 1988: 25-49. Croft S, Gilbert BC, Lindsay Smith JR, Whitwood AC. An E.S.R. investigation of the reactive intermediate generated in the reaction between Fe(II) and hydrogen peroxide in aqueous solution. Direct evidence for the formation of the hydroxyl radical. Free Radic Res Commun 1992; 17:21-39. Rush JD, Koppenol WH. Reactions of Fe(II)-ATP and Fe(II)-citrate complexes with tert-butyl hydroperoxide and cumylhydroperoxide. FEBS Lett 1990; 275:114-116. Graf E, Mahoney JR, Bryant RG, Eaton JW. Iron-catalyzed hydroxyl radical formation. J Biol Chem 1984; 259:3620-3624. Rush JD, Maskos Z, Koppenol WH. Distinction between hydroxyl radical and ferryl species. In: Packer L, Glazer AN, eds. Methods in Enzymology, Vol. 186. San Diego: Academic Press, 1990: 148-156. Yamazaki I, Piette LH. EPR spin-trapping study on the oxidizing species formed in the reaction of the ferrous ion with hydrogen peroxide. J Am Chem Soc 1991; 113:7588-7593. Wardman P, Candeias LP. Fenton chemistry: An introduction. Radiat Res 1996; 145:523-531. Bray WC, Gorin MH. Ferryl ion, a compound of tetravalent iron. J Am Chem Soc 1932; 54:2124-2125. Maskos Z, Rush JD, Koppenol WH. The hydroxylation of the salicylate anion by a Fenton reaction and gamma-radiolysis: A consideration of the respective mechanisms. Free Radic Biol Med 1990; 8:153-162. Lloyd RV, Hanna PM, Mason RP. The origin of the hydroxyl radical oxygen in the Fenton reaction. Free Radic Biol Med 1997; 22:885-888. Chance B, Greenstein DS, Roughton FJW. The mechanism of catalase action. I. Steady state analysis. Arch Biochem Biophys 1952; 37:301-321. Esworthy RS, Chu F-F, Geiger P, Girotti AW, Doroshow JH. Reactivity of plasma glutathione peroxidase with hydroperoxide substrates and glutathione. Arch Biochem Biophys 1993; 307:29-34. Peskin AV, Low FM, Paton LN, Maghzal GJ, Hampton MB, Winterbourn CC. The high reactivity of peroxiredoxin 2 with H2O2 is not reflected in its reaction with other oxidants and thiol reagents. J Biol Chem 2007; 282:11885-11892. Lu C, Koppenol WH. Inhibition of the Fenton reaction by nitrogen monoxide. J Biol Inorg Chem 2005; 10:732-738. Beckman JS, Beckman TW, Chen J, Marshall PA, Freeman BA. Apparent hydroxyl radical production by peroxynitrite: Implications for endothelial injury from nitric oxide and superoxide. Proc Natl Acad Sci USA 1990; 87:1620-1624. Yu ZQ, Persson HL, Eaton JW, Brunk UT. Intra-lysosomal iron: A major determinant of oxidant-induced cell death. Free Radic Biol Med 2003; 34:12431252. Persson HL, Yu ZQ, Tirosh O, Eaton JW, Brunk UT. Prevention of oxidant-induced lysosomal cell death by lysosomotropic iron-chelators. Free Radic Biol Med 2003; 34:1295-1305.

110

Willem H. Koppenol and Patricia L. Bounds

[146] Bresgen N, Jaksch H, Lacher H, Ohlenschläger I, Uchida K, Eckl PM. Iron-mediated oxidative stress plays an essential role in ferritin-induced cell death. Free Radic Biol Med 2010; 48(10):1347-1357. [147] Kurz T, Leake A, Von Zglinicki T, Brunk UT. Relocalized redox-active lysosomal iron is an important mediator of oxidative-stress-induced DNA damage. Biochem J 2004; 378(3):1039-1045. [148] Kurz T, Gustafsson B, Brunk UT. Intralysosomal iron chelation protects against oxidative stress-induced cellular damage. FEBS J 2006; 273:3106-3117. [149] Pietrangelo A. Iron, firend or foe? "Freedom" makes the difference. J Hepatol 2000; 32:862-864. [150] O'Sullivan B, Xu J, Raymond KN. New multidentate chelators for iron. In: Badman DG, Bergeron RJ, Brittenham GM, eds. Iron Chelators. New Development Strategies. Ponte Vedra Beach, FL: The Saratoga Group, 2000: 177-208. [151] Cooper SR, McArdle JV, Raymond KN. Siderophore electrochemistry: Relation to intracellular iron release mechanism. Proc Natl Acad Sci USA 1978; 75:3551-3554. [152] Steinhauser S, Heinz U, Bartholomä M, Weyhermüller T, Nick H, Hegetschweiler K. Complex formation of ICL670 and related ligands with Fe(III) and Fe(II). Eur J Inorg Chem 2004;4177-4192. [153] Merkofer M, Kissner R, Koppenol WH. Redox properties of the iron complexes of CP20, CP502, CP509 and ICL670. Helv Chim Acta 2004; 87:3021-3034. [154] Merkofer M, Domazou AS, Nauser T, Koppenol WH. Dissociation of CP20 from iron(II)(cp20)3; A pulse radiolysis study. Eur J Inorg Chem 2006;671-675. [155] Marshall LE, Graham DR, Riech KA, Sigman DS. Cleavage of deoxyribonucleic acid by the 1,10-phenanthroline-cuprous complex. Hydrogen peroxide requirement and primary and secondary structure specificity. Biochemistry 1981; 20:244-250. [156] Gutteridge JMC. Tissue damage by oxy-radicals: The possible involvement of iron and copper complexes. Med Biol 1984; 62:101-104. [157] Aust SD, Morehouse LA, Thomas CE. Role of metals in oxygen radical reactions. J Free Radic Biol Med 1985; 1:3-25. [158] Stoewe R, Prütz WA. Copper-catalyzed DNA damage by ascorbate and hydrogen peroxide: kinetics and yield. J Free Radic Biol Med 1987; 3:97-105. [159] Masarwa M, Cohen H, Meyerstein D, Hickman DL, Bakac A, Espenson JH. Reactions of low-valent transition-metal complexes with hydrogen peroxide. Are they "Fenton-like" or not? 1. The case of Cu+aq and Cr2+aq. J Am Chem Soc 1988; 110:4293-4297. [160] Johnson GRA, Nazhat NB, Saadalla-Nazhat RA. Reaction of aquacopper(I) ion with hydrogen peroxide. J Chem Soc Faraday Trans I 1988; 84:501-510. [161] Shinar E, Rachmilewitz EA, Shifter A, Rahamim E, Saltman P. Oxidative damage to human red cells induced by copper and iron complexes in the presence of ascorbate. Biochim Biophys Acta 1989; 1014:66-72. [162] Rae TD, Schmidt PJ, Pufahl RA, Culotta VC, O'Halloran TV. Undetectable intracellular free copper: The requirement of a copper chaperone for superoxide dismutase. Science 1999; 284:805-808. [163] Klug-Roth D, Rabani J. Pulse radiolytic studies on reactions of aqueous superoxide radicals with copper(II) complexes. J Phys Chem 1976; 80:588-591.

Redox-Active Metals: Iron and Copper

111

[164] Butler J, Koppenol WH, Margoliash E. Kinetics and mechanisms of the reduction of ferricytochrome c by the superoxide anion. J Biol Chem 1982; 257:10747-10750. [165] Dizdaroglu M, Rao G, Halliwell B, Gajewski E. Damage to the DNA bases in mammalian chromatin by hydrogen peroxide in the presence of ferric and cupric ions. Arch Biochem Biophys 1991; 285:317-324. [166] Spasojevic I, Mojovic M, Stevic Z, Spasic SD, Jones DR, Morina A et al. Bioavaliability and catalytic properties of copper and iron for Fenton chemistry in human cerebrospinal fluid. Redox Rep 2010; 15:29-35. [167] Zielonka J, Kalyanaraman B. "ROS-generating mitochondrial DNA mutations can regulate tumor cell metastasis"-a critical commentary. Free Radic Biol Med 2008; 45(9):1217-1219. [168] Muller FL. A critical evaluation of cpYFP as a probe for superoxide. Free Radical Biol Med 2009; 47:1779-1780.