You are on page 1of 11

FEBS Letters 586 (2012) 585–595

journal homepage: www.FEBSLetters.org

Review

Superoxide dismutases: Ancient enzymes and new insights


Anne-Frances Miller ⇑
Department of Chemistry, University of Kentucky, Lexington, KY 40506-0055, USA

a r t i c l e i n f o a b s t r a c t

Article history: Superoxide dismutases (SODs) catalyze the de toxification of superoxide. SODs therefore acquired
Received 10 October 2011 great importance as O2 became prevalent following the evolution of oxygenic photosynthesis. Thus
Revised 27 October 2011 the three forms of SOD provide intriguing insights into the evolution of the organisms and organ-
Accepted 30 October 2011
elles that carry them today. Although ancient organisms employed Fe-dependent SODs, oxidation
Available online 10 November 2011
of the environment made Fe less bio-available, and more dangerous. Indeed, modern lineages make
Edited by Miguel Teixeira and Ricardo Louro greater use of homologous Mn-dependent SODs. Our studies on the Fe-substituted MnSOD of Esch-
erichia coli, as well as redox tuning in the FeSOD of E. coli shed light on how evolution accommo-
Keywords:
dated differences between Fe and Mn that would affect SOD performance, in SOD proteins whose
Superoxide dismutase activity is specific to one or other metal ion.
Proton-coupled electron transfer Ó 2011 Federation of European Biochemical Societies. Published by Elsevier B.V. All rights reserved.
Redox tuning
Iron
Manganese

1. Introduction The most primitive versions of SOD employ Fe, consistent with
the prevalence of Fe and its ready availability in the reducing envi-
Superoxide dismutases, or SODs, are enzymes that play a piv- ronment in which life is believed to have begun. However more
otal role in metabolizing O2 , preempting oxidizing chain reac- modern organisms employ a version of this enzyme that requires
tions that cause extensive damage, and forestalling formation of Mn for activity, consistent with diminished bioavailability of Fe
a cascade of deleterious reactive oxygen species (ROS) including and increased Fe toxicity as O2 levels rose. Thus, changes in the
hydrogen peroxide (H2O2), hypochlorite (OCl), peroxynitrate inorganic chemical makeup of the environment appear to have dri-
(ONO2) and hydroxyl radical (HO). Aerobic metabolism makes ven evolution of SOD. In what follows, I discuss chemical chal-
some 18 times more energy available per glucose than does glycol- lenges inherent in the transition between catalysis based on Fe
ysis, making possible large and complex organisms. Thus, ability to and catalysis based on Mn. I then review what has been learned
employ O2 constitutes a decisive evolutionary advantage. Yet even about the occurrence and phylogeny of modern SODs. A growing
now, equipped as we are with multiple SODs and numerous sup- wealth of such information is beginning to reveal the paths by
porting antioxidant systems, it is inexorable damage due to ROS which we have acquired and inherited the SODs we have today,
that eventually limits our lives. from diverse ancestors.
Ability to survive O2 is such a stringent selection criterion that
very few organisms lacking SOD survived the transition from 2. Three distinct families of SODs
reducing to oxidizing environment brought about by the evolution
of oxygenic photosynthesis some 2.4 billion years ago [1]. The evo- The name SOD denotes not one, but three unrelated enzymes.
lutionary pressure to develop protection against superoxide was All three earned the name by virtue of ability to convert two mol-
sufficiently intense that SODs evolved on at least three separate ecules of superoxide to one each of dioxygen and hydrogen perox-
occasions. One of these was sufficiently ancient and important that ide, with consumption of two equivalents of H+.
this enzyme is found in all kingdoms of life, indicating that it One family of SODs employs a Ni ion to mediate the chemistry,
evolved even before differentiation of eubacteria from archaea. another uses a Cu ion complexed with a Zn to do the same. The
Thus, the enzyme that keeps us young has been a spectator to third family encompasses enzymes that use a Mn or an Fe as well
much of the evolution of life. as enzymes that can use either.1 The different families of SODs differ
not only with regard to the metal ion that supports activity, but also

⇑ Fax: +1 859 323 1069. 1


We refer to enzymes active only with Mn bound as MnSODs, SODs requiring
E-mail address: afm@uky.edu bound Fe for activity as FeSOD and SODs active with either Fe or Mn as Mn/FeSODs.

0014-5793/$36.00 Ó 2011 Federation of European Biochemical Societies. Published by Elsevier B.V. All rights reserved.
doi:10.1016/j.febslet.2011.10.048
586 A.-F. Miller / FEBS Letters 586 (2012) 585–595

Fig. 1. Comparison of the ribbon structures that characterize the three families of SOD. In FeSOD and Cu,ZnSOD one monomer is coloured, for NiSOD two monomers are
coloured. Each structure in the second row displays the view obtained by rotating the structure above by approximately 90° around a horizontal axis in the plane of the page,
tipping what was the top of the figure in the upper row towards the back. The NiSODs are homohexamers of four-helix bundles of total molecular weight  80 kDa [7]. Each
four-helix bundle binds a Ni ion (green sphere) at the N-terminus. However each Ni site also derives supporting hydrogen bonds from residues from the neighboring four-
helix bundle in a reciprocal arrangement, so two four-helix bundles are colored and the other four are shown in greys. The Cu,ZnSODs are generally 32 kDa homodimers (or
dimers of dimers), where each monomer is a flattened eight-strand beta barrel. The Cu and Zn ions (gold and silver spheres, respectively) are bound on the outside of the
barrel by two loops, including a short helix. The Fe or MnSODs are 45 kDa homodimers (or dimers thereof) where each monomer includes an alpha-helical N-terminal domain
and a C-terminal domain comprised of a three-stranded beta sheet surrounded by alpha helices. The Mn or Fe ion (red sphere) is bound between the two domains by two
amino acids from each and a solvent molecule (orange sphere). Cartoons were made using Pymol [8] and the coordinates 1Q0D.pdb for NiSOD [3], 1HL5.pdb chains F and M
[7,9] for Cu,ZnSOD, and 1ISB.pdb for FeSOD [10,11].

with regard to the protein fold (Fig. 1). Recent progress in under- In order to learn what distinctions might be involved in opti-
standing the mechanism of Ni-SODs has been described by [2,3], mizing Mn-dependent as opposed to Fe-dependent SOD activity,
and ongoing advances in understanding the biological significance we compared the MnSOD and FeSOD of Escherichia coli. As for FeS-
of Cu,Zn SODs are summarized in [4] while the mechanism is re- ODs and MnSODs in general, those of E. coli share a conserved pro-
viewed in [5,6]. The current review will concentrate initially on tein fold and a dimer interface that involves active site residues as
the Fe- and/or Mn-utilizing SODs and then place them in the context well as the funnel by which substrate gains access to the catalytic
of the occurrence and possible evolution of all three SODs. metal ion (Fig. 2A). Both are dimers; they share 43% amino acid
identity including the three His and one Asp that bind the metal
3. New SODs from old: differences required for Mn-based SOD ion in all Fe- and MnSODs, in addition to the conserved Tyr and a
activity vs. Fe-based activity nearby Gln2 (Fig. 2B). In both, the Fe or Mn ion also binds to a sol-
vent molecule that in turn engages in hydrogen bonds (H-bonds)
Diverse primitive organisms employ FeSOD, yet it is apparently with the second sphere Gln. Thus, comparison of the two active sites
absent from animals and fungi, and present in plants only in the does not suggest an obvious explanation for the different metal ion
chloroplasts. This suggests that ancient versions of this enzyme dependencies of activity. Moreover both sites are able to bind either
used Fe but that there was a switch to the use of Mn that occurred metal ion in similar coordination geometries3 [12] with similar elec-
concurrent with evolution of eukaryotes. The switch could have oc- tronic structures [13–15].
curred in response to diminished bioavailability of Fe once the The very similar sites provide a conserved context in which to
atmosphere became oxidizing and Fe’s tendency to engage in Fen- identify differences that could enable one protein to support Mn-
ton chemistry with superoxide, especially under conditions of oxi- based activity while the other supports only activity based on Fe.
dative stress. However the cost of Fe acquisition continues to be Fe and Mn have similar ionic radii and ligand preferences, and both
borne for production of hemes and numerous FexSx and Fe-contain- cycle between their 2+ and 3+ oxidation states in the course of SOD
ing enzymes. A second factor will be that for many Fe-dependent turnover. However the oxidation states with the same charge cor-
systems, the chemistry in question simply does not lend itself to respond to different d-electron configurations for Mn and Fe, and
Mn-mediated catalysis. Another factor was likely the ease with this has consequences for both the ease with which the 3+ ions
which evolution could execute a transition from Fe-based activity can be reduced to the 2+ state, and the tendency of each to coordi-
to activity based on Mn, without ‘interruption of service’. Most nate a sixth ligand. Thus, the different metal ions possess different
simply, for evolution to be able to effect change, the enzyme natural tendencies that may require different modulation and
should possess a selectable amount of the destination activity to complementation on the part of the protein.
serve as the raw material for the action of evolution. SOD meets
both criteria, as witnessed by bacterial and archaeal SODs that
can function with either Fe or Mn. Indeed, these ‘cambialistic’ SODs 2
This Gln is replaced by a His in Mycobacterium tuberculosis and its relatives.
are often found in anaerobic organisms that are relatively primi- 3
The active site pK of Fe3+(Mn)SOD is considerably lower than that of Fe3+SOD.
tive. Thus modern FeSODs and MnSODs could have evolved from Therefore the structure of Fe-substituted MnSOD (Fe(Mn)SOD) reveals a second
a common ancestor via SODs able to use either metal. coordinated OH- in one active site that is absent from the structure of FeSOD [12].
A.-F. Miller / FEBS Letters 586 (2012) 585–595 587

Fig. 2. (A) Overlay of the backbones of FeSOD (orange) and Fe-substituted MnSOD (Fe(Mn)SOD), magenta) and B: detail of the active sites of the overlain proteins. Only one
monomer is shown. Figure was made using Pymol [8] and the coordinates 1ISB.pdb and 1MMM.pdb respectively [10,12]. The amino acid numbering of E. coli FeSOD and
MnSOD are used throughout.

4. Differences between Fe and Mn to be accommodated by nor or acceptor. This allows them to function as stand-alone
different SOD proteins antioxidant defenses and also to consume two equivalents of
superoxide per turnover.
Fe- and MnSODs alternate between the 2+ and the 3+ state of The Em of hexaaquo Fe is 770 mV, between the Ems of the two
the metal ion in the course of disproportionating superoxide. In SOD reactions, but much higher than the midpoint, so the (Fe)SOD
the equations that follow, M indicates the metal ion and the pro- protein should depress Fe’s Em by some 400 mV for optimal theo-
tonation state of the coordinated solvent molecule is also indi- retical SOD activity (Fig. 3). This is likely accomplished to a signif-
cated, immediately following the metal ion. icant extent by the use of a negatively charged Asp ligand.

However hexaaquo Mn’s Em is much higher (1510 mV), so I pro-
O þ
2 þ H þ SOD  M  OH ! O2 þ SOD  M2þ  OH2 ð1aÞ posed that for optimal performance, a Mn-specific SOD protein
þ 2þ 3þ 
O
2 þ H þ SOD  M  OH2 ! H2 O2 þ þSOD  M  OH ð1bÞ would have to depress the metal ion’s Em significantly more than
3+
would an Fe-specific SOD protein (Fig. 3). Indeed, I proposed that
In the first half reaction (1a), reduction of Mn involves conver-
sion of a metal ion with 4 d-electrons to a 5 d-electron system, but
reduction of Fe3+ corresponds to accommodation of a sixth d-elec-
tron. A fifth electron is readily accommodated by an empty d-orbi-
tal4 whereas the sixth electron has to share a d-orbital with one of
the pre-existing d-electrons. Thus, reduction of Mn3+ to Mn2+ is
much more favourable than reduction of Fe3+ to Fe2+. The energetics
of reduction are embodied by the reduction midpoint potential
Em = DG°/nF, where F is Faraday’s constant (Coulombs per mole
of electrons) and n is the number of electrons being transferred
(=1 here). Thus, just as a large pKa is indicative of a base that readily
acquires a proton, a positive Em signifies a site that readily acquires
an electron and is a good oxidant. Sites with negative Ems tend to do-
nate electrons and act as reductants (just as compounds with low
pKas release a proton). The greater intrinsic affinity of Mn3+ for an
electron yields Ems that are 300 to 700 mV higher than those of
Fe3+ for analogous simple complexes5 [16,17].
The higher Em of high-spin Mn3+/Mn2+ explains its lower Fenton
reactivity than Fe and makes it a safer metal ion for use under con-
ditions of oxidative stress. However a high Em also makes it ineffi-
cient in SOD activity. The ideal Em for a SOD would be half-way
between the Ems of SOD’s two half-reactions, near 360 mV vs. the
normal hydrogen electrode (NHE) [18]. A metal site with too low
an Em will be able to reduce superoxide (Eq. (1b)) but will not be
able to extract an electron from superoxide (Eq. (1a)). A site with
too high an Em will be able to oxidize superoxide (Eq. (1a)), but will
not be able to pass the gained electron on to a second molecule of
superoxide (Eq. (1b)). Both types of site can consume superoxide
but only stoichiometrically, unless they are regenerated by other
electron transport systems. Yet part of the elegance and biochem-
ical economy of SODs is their independence from any electron do-

4
In both FeSOD and MnSOD, Fe and Mn are high-spin in both oxidation states.
5
When the Em is higher by 300 mV, reduction is more favourable by  30 kJ/mol or Fig. 3. cartoon of redox tuning of the Fe3+/2+ and Mn3+/2+ ions by the (Fe)SOD and
7 kcal/mol. (Mn)SOD proteins (the proteins of FeSOD and MnSOD, respectively).
588 A.-F. Miller / FEBS Letters 586 (2012) 585–595

greater redox depression on the part of the MnSOD protein could


explain the inactivity of Fe-substituted MnSOD [19,20].

4.1. Ems of metal-substituted SODs

In order to determine whether the (Fe)SOD and (Mn)SOD pro-


teins apply different redox tuning to their bound metal ions, we
compared the Ems of Fe in FeSOD with Fe bound in the (Mn)SOD
protein. We also assessed the Ems of Mn bound in each of the
two proteins [14,20]. (In what follows, ‘(Fe)SOD’ and ‘(Mn)SOD’ Scheme 1. Proton-coupled reduction of the metal ion of SODs, where M stands for
Fe or Mn, and the Gln H-bonding to the coordinated solvent is Gln69 of FeSOD or
stand for the proteins of the FeSOD and MnSOD, respectively, Gln146 of MnSOD (see Fig. 2B). Details of the coordination are omitted from the
Mn(Fe)SOD indicates Mn-substitutes FeSOD and Fe(Mn)SOD indi- right-hand side for the sake of clarity, since they do not change.
cates Fe-substituted MnSOD.)
We confirmed that the reduction midpoint potentials of FeSOD
(100 mV) and MnSOD (300 mV) are both intermediate between rate of turnover, the Em places a fundamental thermodynamic limit
the Ems of SOD’s two half-reactions [20,21]. However we found on what chemistry the enzyme can conduct. The Em measured
that the Em of Fe(Mn)SOD was much lower and that of Mn(Fe)SOD for Fe(Mn)SOD is too low for it to serve as an electron acceptor
was much higher. While it is difficult to obtain precise Em values from superoxide and thus suffices to explain the extremely low
for the SODs, due to the very slow equilibration of the active site turnover rate of Fe(Mn)SOD at neutral pH [20].
metal ion with mediators, experiments with redox-active dyes Our findings indicate that in order to optimize SOD’s perfor-
confirmed that when Fe is bound in the (Mn)SOD protein, its Em mance with Mn as the catalytic metal ion, evolution would have
is some 300 mV lower than when it is bound in (Fe)SOD protein had to identify ways of depressing the bound Mn’s Em more than
[20]. Thus, FeSOD could be reduced by ascorbate (Em  60 mV) had been necessary for Fe.
but Fe(Mn)SOD could not. Fe2+(Mn)SOD could reduce superoxide,
demonstrating that the reduced state could bind substrate and 5. A second challenge: different preferences regarding
moreover was competent for both the proton transfer and electron coordination number
transfer inherent to reaction 1b [20]. However Fe3+(Mn)SOD lacked
Fe3+SOD’s ability to accept an electron from superoxide despite its Besides the large difference between the intrinsic Ems of Mn3+/2+
greater-than-WT ability to bind small anions [15]. Moreover this and Fe3+/2+, the different electronic configurations of these ions in
was true even at lower pH, where competitive inhibition by OH the corresponding oxidation states are also expected to produce
is attenuated [20]. Thus, it appears that the (Mn)SOD protein de- different ligand-binding preferences. As noted above, Fe3+ is a d5
presses the Em of Fe much more than does (Fe)SOD (Fig. 3). ion (it has 5 d-electrons) and Fe2+ is not, whereas for Mn this prop-
We showed that the same holds true for Mn. We found that the erty is reversed with Mn2+ being a d5 ion and Mn3+ not. For Fe3+
Mn(Fe)SOD was fully reduced as-isolated and very difficult to oxi- and Mn2+ with one electron in each d-orbital, there is no orbital
dize (Fig. 3). Although the Mn3+ state could be generated using the with special priority. Ligand binding affects the metal ion’s d-elec-
very powerful oxidants KMnO4 and K2IrCl6 , milder oxidants able to trons to similar extents regardless of which orbitals the ligand
oxidize MnSOD were unable to oxidize Mn(Fe)SOD. We were not interacts with. For a d6 ion, one d-orbital has two electrons and
able to execute a titration, but based on the Ems of the oxidants the complex is more stable overall if this orbital does not interact
that failed to oxidize Mn(Fe)SOD, its Em must be >870 mV [14]. with as many or as strong ligands as the other (singly occupied)
Thus we found that for Fe, the (Mn)SOD protein produces an Em orbitals do. Thus, it is beneficial to avoid coordinating a sixth ligand
some 300 mV lower than that produced by (Fe)SOD, and for Mn, and retain an empty coordination site where ligand would have
the (Mn)SOD protein produces an Em at least 500 mV lower than interacted with the doubly-occupied orbital. By contrast a d4 ion
that produced by (Fe)SOD. These differences are consistent with benefits from asymmetry in the opposite way. Because one d-orbi-
the differences between model Mn3+/2+ and Fe3+/2+ complexes. tal is empty, strong ligands can bind the metal ion along the axis
Moreover our findings suffice to explain the inactivity of E. coli defined by the empty dz2 orbital without incurring a large cost to
Fe(Mn)SOD and Mn(Fe)SOD [22]. They also predict that SODs that the stability of the complex. Meanwhile the axes describing the
are active with either Fe or Mn will apply redox tuning intermedi- occupied d-orbitals are populated by less strong or fewer ligands
ate between the tuning applied by E. coli’s metal-specific SOD for maximum stability of the complex [26]. Thus, Mn3+ SOD can
proteins and that their Fe-versions will have lower Ems than their keep its four d-electrons in the four d-orbitals that have the least
Mn-versions. We are testing these predictions. contact with the strong OH ligand that defines the Z axis, and in-
It has been pointed out that Fe3+(Mn)SOD binds a second OH stead populate the orbitals in the X-Y plane defined by the weaker
more tightly than does Fe3+SOD and that this could be responsible two His and Asp ligands (1.5 ligands per direction vs. 2 along Z).
for the low activity [23,24]. However since Fe3+(Mn)SOD also binds This is an example of Jahn-Teller stabilization. d6 Fe2+ benefits from
other small anions much more tightly than does Fe3+SOD [15,24], it having only three ligands in the X-Y plane, as it can doubly popu-
remains unknown whether its affinity for substrate will also be late dYZ at lower cost than doubly populating the dz2 which is di-
elevated. Competitive inhibition by OH does not depend only on rected straight at two ligands [26]. Mn2+ or Fe3+ with one
the affinity for OH, but on the relative affinity for OH vs. sub- electron in each d orbital cannot exploit an asymmetric ligand
strate, which is not known. Moreover the restoration of a small environment and so tend not to favour one: they can accommodate
amount of activity at lower pH is exactly what is expected consid- a sixth ligand and benefit from the electrostatic advantages of
ering that reduction of FeSOD is accompanied by protonation of the doing so whereas Fe2+ or Mn3+ would sacrifice advantages stem-
coordinated solvent [25] (Scheme 1), because the Em of a proton- ming from a less symmetric coordination environment.
coupled-reduction will increase by 60 mV for each pH unit Consistent with the above, small anions bind directly to the Fe3+
decrease in pH. Thus lower pH would raise Fe(Mn)SOD’s low Em of Fe3+SOD, but not the Fe2+ of Fe2+SOD [27]. Indeed, NMR spectros-
towards a catalytically competent value. Thus, while the redox copy provided the first positive evidence for substrate analog bind-
tuning is related to proton transfer, H-bonding and likely substrate ing to the reduced state of Fe2+SOD, indicating that substrate
binding as well, and these features will surely all contribute to the analogs (and substrate) bind outside the coordination sphere of
A.-F. Miller / FEBS Letters 586 (2012) 585–595 589

Fe2+, in an outer-sphere binding site. Similarly, Fee showed that the because they also protonate the substrate in the second of these
pK near 9 of the Fe3+SOD active site corresponds to OH coordinat- two reactions (1b).
ing to Fe3+ [28]. However for Fe2+SOD we used 13C NMR and 13Cf- The proton taken up in conjunction with metal ion reduction is
Tyr labelled Fe2+SOD to show that the corresponding active site pK understood to be acquired by the coordinated solvent [25] so that
of Fe2+SOD represents an outer-sphere event: deprotonation of the redox reaction is correctly written M3+OH + e + H+ ?
Tyr34 [29]. M2+OH2. Because coordinated OH is a better ligand for Fe3+
In contrast with Fe3+, Mn3+ is a d4 ion and indeed, although (and Mn3+) than for Fe2+ (and Mn2+), I proposed that the active site
N3 binds to Mn3+SOD, the complex it forms is predominantly out- of (Mn)SOD might suppress metal ion reduction (and the Em) by
er-sphere at room temperature [30]. Thus, the (Mn)SOD protein disfavouring protonation of coordinated solvent [20,22,34].6
provides assistance to substrate analog binding, in the form of an This model moreover is consistent with phylogenetic analyses,
H-bond from Tyr34 [31]. In accordance with its d5 electronic con- which find that the largest conserved difference between FeSODs
figuration, Fe3+(Mn)SOD displays considerably higher affinity for and MnSODs is the origin of the active site Gln (occasionally His)
coordinating anions such as N3, F and OH than does Mn3+SOD which provides the only H-bond between the protein matrix and
and the ions coordinate Fe3+ rather than binding in the outer the coordinated solvent7 [39,40]. In dimeric FeSODs this Gln derives
sphere as in Mn3+SOD. The fact that Fe3+(Mn)SOD’s affinities for from an a helix in the N-terminal domain (positon 69 in E. coli)
small anions are also considerably higher than those of Fe3+SOD whereas in MnSODs it derives from a loop between b strands in
suggests that assistance provided by the (Mn)SOD protein is the C-terminal domain (position 146 in E. coli) [10,41,42]. Tetrameric
retained upon Fe-substitution. FeSODs employ a residue from the second of these positions but they
In Mn2+SOD, Un and Whittaker have shown that substrate ana- use a His instead of a Gln [40]. Similarly, mutational studies have
logs coordinate to Mn2+ consistent with this ion’s d5 configuration, demonstrated that residues that can tune the position and polariza-
and in contrast to Mn3+SOD and Fe2+SOD [32,33]. Tabares and Un tion of the active site Gln affect the Em of the bound metal ion as well
showed that Tyr34 favoured coordination of the anion to Mn2+ at as the protein’s relative activity with Mn bound vs. Fe bound [34,43].
higher temperatures approaching physiological values, whereas Thus the largest difference between FeSODs and MnSODs is in fact
for the Mn3+ state the native Tyr tends to suppress formation of the single residue with the largest influence over the protonation
an inner-sphere complex [24]. This would tend to augment rather state of coordinated solvent (Fig. 2B).
than compensate for the coordination preferences of Mn2+ vs. We proposed that the lower Ems produced by (Mn)SOD could be
Mn3+, however it would have the benefit to the enzyme of favour- explained by a stronger H-bond donated from Gln to coordinated
ing dissociation of peroxide formed in reaction 1b by emphasizing OH than in (Fe)SODs, as this would disfavour acquisition of the re-
outer-sphere interactions with Mn3+. dox-coupled proton and thus metal ion reduction [13,20]. This was
Mn2+ should be competent to bind substrate analogs in the in- borne out by 15N NMR direct observation of the H-bonding Gln
ner sphere, however Mn2+(Fe)SOD has much lower inner-sphere Side chain N [34]. We found that the active site Gln of the (Mn)SOD
anion binding ability than does Mn2+SOD, since 100 mM N3 or protein is roughly twice as strongly coupled to Fe2+ as the Gln of
F did not cause any discernable change in the Mn2+ EPR signal (Fe)SOD, consistent with a considerably stronger hydrogen bond
[14]. The (Fe)SOD protein environment therefore seems to be less in the (Mn)SOD protein.
supportive of inner sphere anion binding than the (Mn)SOD active Our model predicts that mutations that stabilize coordinated
site, for Mn2+ as for Fe3+. OH vs. H2O will lower the Em (and increase the site’s activity with
Thus, we find that the electronic configuration of the metal ion Mn vs. Fe). To test our proposal, we constructed mutants in which
is the dominant factor in the inner-sphere vs. outer-sphere nature Gln69 was replaced by either a His or a Glu residue [21]. Whereas
of anion binding, with the d5 state of both metal ions coordinating Gln69 of FeSOD is constrained to be an H-bond donor by its teth-
substrate analogs directly, although these correspond to different ering H-bond from Trp [10], we postulated that His could be either
oxidation states. It does not appear that the (Fe)SOD and (Mn)SOD a donor or an acceptor, depending on its tautomeric state. In con-
proteins compensate for each metal ion’s different preferences. trast, we proposed that Glu would be ionized at neutral pH and act
Rather, the mechanistic feature exerting the strongest selective as an obligate H-bond acceptor. Thus we hypothesized that if
pressure may be ability to dissociate from peroxide. replacement with His weakens H-bond donation to coordinated
solvent then coordinated H2O should no longer be destabilized
6. How the SOD proteins tune the Em and the Em would no longer be actively depressed. Furthermore,
we proposed that if the coordinated solvent were subject to H-
The too-low Em of Fe(Mn)SOD can explain its inactivity, but in bond acceptance by Glu, favouring coordinated H2O, this would
turn requires an explanation of its own. The differences between in turn stabilize the reduced state of the metal ion and raise the
the Ems for metal bound in (Mn)SOD vs. in (Fe)SOD of -300 mV Em in the Q69E mutant.
for Fe and -500 mV for Mn are very large, corresponding to 5 - 8 Our hypotheses were borne out on both counts [21]. The sur-
pH unit changes in the pKa of an acid. However (Fe)SOD is superfi- prise was in the large magnitude of the effect. While the Q69H mu-
cially very similar to (Mn)SOD (Fig. 2). Therefore, we proposed that tant had an Em some 250 mV above that of WT, the Em of the Q69E
the different Ems could reflect differences in the extent to which
the different active sites favour uptake of the proton that is cou-
pled to metal ion reduction [20,34] (Scheme 1). 6
Proton transfer is likely also to be coupled to substrate binding and product
Proton-coupled electron transfer is arguably the rule, rather release [35]Bull, C. and Fee, J.A. (1985). Steady-State Kinetic Studies of Superoxide
than the exception, in biochemical redox reactions. Acquisition of Dismutases: Properties of the Iron Containing Protein from Escherichia coli. J. Am.
a proton in conjunction with metal ion reduction results in conser- Chem. Soc. 107, 3295-3304, [36]Abreu, I.A., Rodriguez, J.A. and Cabelli, D.E. (2005).
Theoretical studies of manganese and iron superoxide dismutases: superoxide
vation of the same total charge [29,35] and in the case of SOD there binding and superoxide oxidation. J. Phys. Chem. B 109, 24502-24509, [37]Miller,
is the additional chemical incentive that two protons are co-sub- A.-F., Yikilmaz, E. and Vathyam, S. (2010). 15N-NMR characterization of His residues
strates in the disproportionation of superoxide. Indeed, reduction in and around the active site of FeSOD. Biochim. et Biophys. Acta 1804, 275-284,
of superoxide to peroxide can occur only in conjunction with or [38]Maliekal, J. et al. (2002). Comparison and Contrasts between the Active Site pKs of
Mn-Superoxide Dismutase and those of Fe-Superoxide Dismutase. J. Am. Chem. Soc.
following protonation, since it is extremely unfavourable to add
124, 15064-15075.
an electron to a negatively charged diatomic molecule. SODs are 7
We do not count the H-bond between coordinated solvent and the ligand Asp
able to both oxidize and reduce the same substrate precisely because this is internal to the metal site.
590 A.-F. Miller / FEBS Letters 586 (2012) 585–595

mutant was at least 660 mV higher than that of WT [21] and the critical: we showed that the strength and polarity of the H-bond
enzyme was stable in the Fe2+ state. The direction of the changes to coordinated solvent exerts a very strong effect on the Em of
supported H-bonding as the dominant mechanism of redox tuning the metal ion, which in turn is directly reflected in catalytic
over electrostatics, since installation of Glu, with its tendency to activity. The stronger hydrogen bond donation we observed in
negative charge, should have favoured Fe3+ over Fe2+ and lowered (Mn)SOD can account for the lowered Ems produced by (Mn)SOD,
the Em based on electrostatics. and in turn provide a mechanism by which evolution could have
Crystallography showed that the His69 was not constrained to optimized the SOD protein for function with Mn with relatively
donate an H-bond the way Gln69 is in WT [21], moreover in the re- minor perturbations to the structure, and none to the residues
duced state of Q69H FeSOD, 15N NMR spectroscopy revealed that directly responsible for binding the metal ion. Our chemical
H69 is an H-bond acceptor instead of a donor [37]. Thus it appears mechanism explains the conservation across species and even
that in this mutant, residue 69 is able to adapt to the protonation phyla of different placements of the active site Gln (His) in FeS-
state of the coordinated solvent as opposed to imposing on it an ODs vs. MnSODs [40].
H-bond whose polarity is determined by the rest of the protein
[44]. 7. Occurrences of SOD
In the Q69E mutant, a combination of density functional theory
(DFT) calculations and X-ray crystallographic structure determina- SODs are found in all the kingdoms of life (Fig. 4). The occur-
tion at 1.1 Å were used to probe the protonation state and H-bond- rences of the different SODs, as well as homologies and phyloge-
ing of residue 69 [44]. Crystallography showed that the overall netic trees provide evidence as to the evolutionary histories of
structure of the mutant enzyme was essentially superimposable the different SODs.
on that of the WT and active site atom positions differed by at most
0.5 Å [21]. Although the 1.1 Å crystal structure was refined two 7.1. NiSODs
ways, one assuming protonated Glu69 and one assuming ionized
Glu69, both converged on the same positions and distances. Simi- NiSOD was first discovered in Streptomyces [49]. Genes with
larly, DFT calculations were performed assuming three different strong homology to NiSOD have been identified in the genomes
possible distributions of labile protons among the ionizable active of additional genus’ in actinobacteria [50] and NiSOD has been
site residues. All three energy minimizations produced the same identified in cyanobacteria [51]. Schmidt et al. found putative Ni-
result: ionized Glu69 accepting H-bonds from Tyr34 and coordi- SOD genes in other families of bacteria and even a potential NiSOD
nated solvent in an arrangement with some sharing among all fusion gene in the eukaryotic green algae Ostreococcus tauri, the
three residues [44]. The distances obtained from the four crystallo- smallest free-living eukaryote known. Nonetheless, they found no
graphically independent active sites agreed very well with the dis- evidence for NiSODs in gram-positive bacteria, archaea or eukary-
tances obtained from DFT where the protonation states are known. otes other than the aforementioned green alga [50].8
Thus, although our crystal structure was not able resolve H-bond- Because NiSOD is restricted to relatively few groups, it appears
ing H atoms in the active site, by reference to the DFT we could to have evolved after differentiation of eukaryotes and substantial
conclude that in reduced Q69E-FeSOD, ionized Glu69 accepts an diversification among bacteria. Because NiSOD’s active site derives
H-bond from coordinated solvent. Moreover the H-bond is shorter mainly from nine N-terminal amino acids that are not integral to
than in WT FeSOD and therefore likely stronger [44]. Indeed the the secondary or tertiary structure [3,7], the Ni binding site might
Q69E Fe2+SOD is more stable that WT, indicating that this mutation have been a later addition to a pre-existing four-helix bundle
releases strain in the active site. Thus the reduced state is more protein (Fig. 1).
stable.
It also appears that the oxidized state of Q69E-FeSOD is less sta- 7.2. Cu,ZnSOD
ble in the WT-like conformation, as it is isolated with a sixth ligand
bound, most likely OH [44]. The implication is that in the oxidized Many Gram-negative bacterial pathogens possess periplasmic
state Glu69 is neutral, in a GluH + OH <=> Glu H2O equilibrium Cu,ZnSODs. The bacterial Cu,ZnSODs retain the same monomer
stabilized in the former form by OH‘s ability to coordinate Fe3+. structure as is found in eukaryotes, but different association be-
Thus we infer that Glu69 changes from being protonated in oxi- tween monomers. For example E. coli’s Cu,ZnSOD appears to be
dized Q69E-FeSOD to ionized in the reduced state [44]. Since the monomeric [52,53] and Salmonella enterica has two genes for
MCD and NMR spectra reveal WT-like coordination and electronic Cu,ZnSOD, one of which yields a dimeric Cu,ZnSOD while the other
state for Fe in both oxidation states [45], we infer that coordinated yields a monomeric Cu,ZnSOD [54].
solvent still acquires a proton upon reduction of Fe, but consider- 14 of 15 archaeal genomes screened lacked a Cu,ZnSOD homo-
ing that a proton is lost from Glu, the net reaction is transfer of a logue with over 20% homology to human Cu,ZnSOD, but Banci et al.
proton from Glu to coordinated solvent. This will be much less identified a gene with 32% identity to that of human Cu,ZnSOD in
energetically costly than the WT mechanism in which the proton Methanosarcina [55]. Unfortunately, catalytic activity was not
must be acquired from water. Thus, the greatly increased Em of tested.
Q69E-FeSOD can be explained by the combination of loss of im- So far, protists appear to lack Cu,ZnSOD [56]. This is fascinating
posed H-bond donation to coordinated H2O in the reduced state, if true, since plants, animals and fungi all possess Cu,ZnSOD. The
gain of a very favourable H-bond acceptance from coordinated sol- other eukaryotes’ Cu,ZnSOD could have been obtained by gene
vent instead in the reduced state, and a subtle but important transfer from the endosymbiont that gave rise to mitochondria.
mechanistic change affecting the origin of the proton gained by However since many endosymbiont genes were lost in the course
coordinated solvent [44]. Together, these three contributions are of gene redistribution to the nucleus, and some protists even lost
able to explain the surprisingly large increase in the Em of Q69E- mitochondria, the ancestor of protists could have lost more genes
FeSOD, and the combination of the Q69H- and Q69E-FeSOD mu- than other eukaryotes did and this could explain protists’ lack of
tants demonstrates that manipulation of the energies associated a Cu,ZnSOD.
with protons whose movements are coupled to electron transfer
can indeed alter the Em by several hundreds of mV [22].
Thus we have demonstrated that the position and identity of 8
Schmidt et al. nonetheless note the presence of a peptide with high homology to
the residue that H-bonds to the coordinating solvent molecule is NiSOD’s active site ’nickel hook’ in a Xenopus protein DUF1619.
A.-F. Miller / FEBS Letters 586 (2012) 585–595 591

Fig. 4. Distribution of SODs in cells and branches of life. Different colours indicate the different SODs: orange for FeSOD, magenta for MnSOD, blue for Cu,ZnSOD and green for
NiSOD. Quaternary states of SODs are indicated in the cell cartoon but not in the tree. This cartoon summarizes general features only, for details please see the text and Refs.
[46–48]. There is an older report of CuSOD in the nucleus.

Yeast Cu,ZnSOD has been found not only in the cytoplasm, but [59,70]. Given the relative portability of this gene, it is possible that
also in the mitochondrial intermembrane space [57]. Indeed, it has been transferred from bacteria to eukaryotes independently
sequencing of the Candida genome revealed two genes for Cu,Zn- of endosymbiotic events [47]. Virus-mediated gene transfer could
SOD (reviewed by [58]). also explain the presence of a Cu,ZnSOD gene in just one species
Plants contain Cu,ZnSOD in the cytosol and in the chloroplast. of archaeon (so far) [47,55], and would be consistent with this
There may moreover be multiple Cu,ZnSODs expressed from sev- gene’s occasional lack of amino acid residues that are normally
eral different genes [59]. Plant Cu,ZnSODs are a highly homologous responsible for metal ion binding, and thus critical for catalytic
group but there are a few features that distinguish cytosolic from activity [55]. Fink and Scandalios suggest that plant chloroplast
chloroplastic Cu,Zn-SODs, including the numbers and positions of and cytosolic Cu,ZnSODs evolved from a common ancestor [59].
introns. A notable exception is the (presumed) cytoplasmic The fact that Cu,ZnSOD is absent from protists as well as from such
Cu,ZnSOD of the liverwort (Marchantia paleacea) which more clo- primitive plants as Chlamydomonas and moss [48] raises the ques-
sely resembles chloroplast Cu,ZnSODs. This Cu,ZnSOD has been tion of how it was acquired by higher eukaryotes. The large differ-
proposed to resemble the last common ancestor of plant cytoplas- ences between eukaryotic and bacterial Cu,ZnSODs do not suggest
mic and chloroplast Cu,ZnSODs [59]. Cu,ZnSOD has also been found an obvious link between the two.
in plant peroxisomes [60].
Animal cells possess dimeric Cu,ZnSOD in the cytoplasm [61], 7.3. Fe-, Mn- and Fe/MnSODs
and also export a tetrameric glycosylated Cu,ZnSOD [62], each
based on a separate gene. The extracellular Cu,ZnSODs of MnSODs and FeSOD are so similar that they are generally ac-
Schistosoma improve the parasites’ ability to survive the oxidant cepted to have arisen from a common ancestor [47,59] (see
attacks of phagocytes [63]. Cytoplasmic Cu,ZnSOD has also been Fig. 4). Since then, FeSODs and MnSODs have significantly diverged
found in the nucleus [64], the intermembrane space of mitochon- from one another, as their amino acid sequences cluster in two
dria [57,65] and peroxisomes [66,67]. However since the precise separate groups in phylogenetic trees [47,59]. Thus the presence
intracellular location of Cu,ZnSOD is responsive to the metabolic of FeSOD and MnSOD in different organelles and in different clas-
state of cells and tissues [68] the situation is complicated and ses of organisms provides a fascinating perspective on the evolu-
our knowledge is incomplete. Analysis of the genes for Cu,ZnSODs tion of life [47].
indicates that the intracellular and extracellular Cu,ZnSODs should
collectively be regarded as distinct from the bacterial Cu,ZnSODs 7.3.1. Prokaryotes
[69]. The extracellular SODs from mammals appear to be more clo- FeSOD is found in both aerobic and anaerobic bacteria [71]. It is
sely related to fungal Cu,ZnSODs [59], raising the possibility that also found in oxygenic as well as anoxygenic photosynthetic bacte-
the extracellular version is the more ancient one. ria [72,73]. Pathogens such as Helicobacter pylorii have FeSOD,
Cu,ZnSOD’s distribution among organelles must be interpreted which may contribute to their ability to survive host defenses [74].
with caution, as this SOD has been found in the genomes of several MnSOD is found in aerobic bacteria and its transcription is up-
viruses where it could be attributed to gene transfer from a host regulated upon exposure to oxygen, suggesting that it represents a
592 A.-F. Miller / FEBS Letters 586 (2012) 585–595

later elaboration [75]. Cyanobacteria are shown to be unique in with changes in peroxisome activity. However regulation of mito-
containing membrane-bound manganese superoxide dismutases chondrial SOD activity has emerged as a very important aspect of
(MnSOD) [76]. regulation of the levels of redox signalling molecules such as hydro-
Some archaea such as Aeropyrum pernix and Pyrobaculum calid- gen peroxide (H2O2, a product of SOD activity) and nitric oxide (NO,
ifontis possess Mn/FeSODs that are active with either metal ion but which is consumed by superoxide) [94,95]. For example, elevated
tend to emphasize Mn incorporation more under aerobic condi- MnSOD levels have been related to the aggressiveness of some can-
tions and Fe under anaerobic conditions [75,77]. The aerobic archa- cers [96,97]. This is an exciting and rapidly developing area, which is
eon Sulfolobus solfataricus and facultative aerobe Acidianus unfortunately beyond the scope of this review.
ambivalens produce SODs that are FeSODs [78,79]). Even strictly
anaerobic methanogens have been found to have SODs. Methanob- 8. Insights into evolution of eukaryotes, based on the
acter bryantti and M. thermoautotrophicum have tetrameric FeSODs distribution of Fe and Mn SOD’s
whose amino acid sequence and susceptibility to inhibition bears
stronger similarity to the MnSODs’ than the bacterial FeSODs Because Fe and/or MnSODs can be found throughout all king-
[80,81]. doms of life, they can tell us about even very early events in its
The SODs of aerobic halophyles are MnSODs [53,82] which have evolution. Indeed, several valuable analyses of SOD phylogenies
greater resemblance to the MnSODs of mitochondria than the have appeared recently [47,59].
MnSODs of bacteria [59]. The FeSODs and MnSODs are understood to have arisen from
a common ancestor very early in the history of life, since they
7.3.2. Eukaryotes are represented in both eubacteria and archaea. Mn/FeSODs
FeSOD has been found in protists ranging from the amitochond- and FeSODs appear the more primitive based on their prevalence
riate Entamoeba histolica [83], trypanosomes [56], plasmodia [84] in more primitive prokaryotes and primitive eukaryotes (Fig. 4),
and Perkinsus marinus [85], and has been shown to provide these as well as the lower extent of regulation associated with their
parasites with protection against oxidative stress unleashed by expression, compared to that of the MnSOD gene. Features spe-
macrophages or produced by many anti-malarial and anti-trypan- cifically associated with Mn-based activity are common to bacte-
osomal drugs [56]. Some of these FeSODs are cytoplasmic and oth- rial and archaeal MnSODs suggesting optimization of Mn use
ers are mitochondrial, based on work in Plasmodium [86]. prior to divergence of these kingdoms.11 However Fe use is asso-
MnSOD has been reported tightly bound to the thylakoid mem- ciated with quite different amino acid signatures in dimeric vs. tet-
branes of the photosynthetic protist Euglena [87]. This is reminis- rameric FeSODs suggesting more time for divergence within this
cent of the membrane-bound MnSOD of filamentous group, and/or possibly a return to use of Fe from among a Mn-uti-
cyanobacteria. lizing tetrameric lineage.
Numerous fungi have been found to have MnSOD.9 Among It is now accepted that all modern mitochondria stem from
pathogens, mitochondrial MnSOD is important for extended survival endosymbiotic assimilation of a species of a-proteobacteria, which
and virulence [88]. Several fungi are reported to have cytosolic contributed many genes for energy metabolism and heterotrophy
MnSOD. For example, Candida albicans has a dimeric cytoplasmic to the host genome [98]. Even amitochondriate eukaryotes
MnSOD in addition to the expected tetrameric mitochondrial (protists) appear to be descended from an ancestor that once
MnSOD (reviewed by Fréalle [58]). possessed a mitochondrion. Nonetheless, the MnSOD of mitochon-
Plants have long been known to have FeSODs in their chloro- dria does not appear to be of bacterial origin.
plasts [89]. Recently however FeSOD has also been found in the The host organism that first acquired mitochondria appears to
cytoplasm of cowpea [90]. Some plants have several FeSOD genes have been of archaeal origin based on strong homologies between
[48]. In L. japonicus, one FeSOD is localized to the chloroplast eukaryotic and archaeal transcriptional and translational machin-
whereas the other appears to be cytoplasmic [91]. FeSOD in the ery (reviewed by Martin [98]). It seems simplest to propose that
chloroplast may associate with the plastid nucleoid and participate a cytoskeleton evolved first [98]. Once phagocytosis was possible,
in signalling or gene regulation [48]. the tremendous possibilities for accelerated acquisition of advan-
Plants have MnSOD in their mitochondria, as expected. It was tageous traits via mergers and acquisitions would have immedi-
also reported to be in chloroplasts of the unicellular alga ately favoured the first organisms able to assimilate bacteria
Chlamydomonas, but not in those of higher plants. Again, there [99]. Given the metabolic revolution precipitated by the transfer
may be several genes, possibly to provide different SODs for differ- of entire constellations of genes from the endosymbiont to the host
ent sub-compartments of the mitochondrion. Plant MnSOD is also genome, it is not difficult to imagine that a burst of diversification
found in peroxisomes, where its levels are regulated independently would follow, allowing rapid divergence and development of the
of the levels of mitochondrial MnSOD [92]. Plant MnSODs share different branches of eukaryotes [100], and also eclipsing the
some 70% homology, and are more distantly related to plant chances of any other branches of primitive life to compete and con-
FeSODs, suggesting separate evolutionary origins. fer similar status on their progeny to that acquired by eukaryotes.
In general, animals possess MnSOD in their mitochondria10 In summary it seems reasonable to consider that acquisition of
(Crustaceans that employ Cu heavily for oxygen transport lack cyto- mitochondria was very closely related to and even possibly insep-
plasmic Cu,ZnSOD and have a cytoplasmic MnSOD, in addition to the arable from the evolution of eukaryotes, requiring only the possi-
expected mitochondrial MnSOD [93]). Mitochondrial MnSOD serves bility of phagocytosis as a prerequisite [98,100].
complementary roles to those played by cytosolic and extracellular Homology among SOD sequences adds to this picture that the
Cu,ZnSOD, and accordingly, is regulated differently. In animals, MnSODs of modern eukaryotes’ mitochondria appears to be de-
MnSOD is localized where it can metabolize superoxide produced rived from an archaeal MnSOD. Specifically, BLASTp homology
as a byproduct of respiratory electron transfer. This can be viewed searches based on MnSOD from halobacterium identify mitochon-
as a chronic problem, in contrast to bursts of reactive oxygen species
(ROS) produced by macrophages or elevated levels of ROS associated
11
It is difficult to make firm conclusions because many SODs are identified as Fe- or
Mn- specific based on their amino acid sequences rather than measurements of the
activity of Fe and Mn forms, creating a risk of circular logic. In addition, specificity is
9
A FeSOD has not yet been identified in a fungus, to my knowledge. often not absolute, with many prokaryotic SODs showing some amount of activity
10
Animals seem to lack FeSOD. with each metal ion, even if one metal ion performs better than the other.
A.-F. Miller / FEBS Letters 586 (2012) 585–595 593

drial MnSODs as homologues, as well as SODs from other archaea. Acknowledgements


However they are more distantly related to bacterial SODs. In con-
trast, MnSODs from a-proteobacteria display much higher homol- This paper is dedicated to the memory and inspiring contribu-
ogy with SODs of other bacteria and not with eukaryotic tions to science of Prof. A.V. Xavier. The author is grateful to the
mitochondrial SODs or SODs from archaea. Thus the sequence N.I.H. for funding under 1R01GM085302-01A.
homologies of MnSODs suggest that -1- the host of the endosym-
biosis was of achaeal lineage, -2- the host’s MnSOD gene was re- References
tained and augmented with a presequence to produce a
mitochondrial targeting peptide, yielding modern mitochondrial [1] Blankenship, R.E. (2010) Early evolution of photosynthesis. Plant Physiol. 154,
434–438.
MnSOD whereas -3- endosymbiont MnSOD and FeSOD genes ap- [2] Herbst, R.W., Guce, A., Bryngelson, P.A., Higgins, K.A., Ryan, K.C., Cabelli,
pear to have been lost from animals and fungi. Many other genes D.E., Garman, S.C. and Maroney, M.J. (2009) Role of conserved tyrosine
were lost from the endosymbiont’s genome, so loss of the SOD residues in NiSOD catalysis: a case of convergent evolution. Biochem. 48,
3354–3369.
genes would not be remarkable. Proposal 1 reiterates conclusions [3] Wuerges, J., Lee, J.-W., Yim, Y.-I., Yim, H.-S., Kang, S.-O. and Carugo, K.D.
from much more extensive works [99]. However proposal 2 sug- (2004) Crystal structure of nickel-containing superoxide dismutase reveals
gests two-way traffic in that mitochondria acquired a protein of another type of active site. Proc. Natl. Acad. Sci. USA 101, 8569–8574.
[4] Potter, S.Z. and Valentine, J.S. (2003) The perplexing role of copper-zinc
host origin, though not the gene.
superoxide dismutase in amyotrophic lateral sclerosis (Lou Gehrig’s disease).
Protists appear to be an exception to the above in that they J. Biol. Inorg. Chem. 8, 373–380.
lack a MnSOD but possess a FeSOD instead. Phylogenetic analysis [5] Miller, A.-F. (2003) Superoxide processing in: Coordination Chemistry in the
does not support its acquisition from an a-proteobacterial endo- Biosphere and Geosphere (Que, L. Jr.Jr. and Tolman, W., Eds.), pp. 479–506,
Pergamon, Oxford.
symbiont that became the mitochondria. The protist mitochon- [6] Hart, P.J., Balbirnie, M.M., Ogihara, N.L., Nersissian, A.M., Weiss, M.S.,
drial FeSODA and FeSODC seem more closely related to Valentine, J.S. and Eisenberg, D. (1999) A structure-based mechanism for
e-proteobacteria instead, and appear to stem from a different copper–zinc superoxide dismutase. Biochem. 38, 2167–2178.
[7] Barondeau, D.P., Kassmann, C.J., Bruns, C.K., Tainer, J.A. and Getzoff, E.D.
gene acquisition event than the one in which the cytoplasmic (2004) Nickel superoxide dismutase structure and mechanism. Biochemistry
FeSODBs were obtained [101]. More information is needed to set- 43, 8038–8047.
tle this question. [8] DeLano, W.L. (2002). The PyMOL Molecular Graphics System. http://
www.pymol.org DeLano Scientific.
Regarding chloroplasts, plant FeSODs were found to be more [9] Strange, R.W. et al. (2003) The structure of holo and metal-deficient wild-
closely related to cyanobacterial FeSODs than to those of Archaea type human Cu, Zn superoxide dismutase and its relevance to familial
[47]. Thus, modern plant chloroplasts are believed to have retained amyotrophic lateral sclerosis. J. Mol. Biol. 328, 877–891.
[10] Lah, M.S., Dixon, M.M., Pattridge, K.A., Stallings, W.C., Fee, J.A. and Ludwig,
a SOD of endosymbiotic origin12 [47,48,59]. It may be beneficial for M.L. (1995) Structure-Function in E. coli Iron Superoxide Dismutase:
chloroplasts to employ a SOD that does not require Mn, since photo- Comparisons with the Manganese Enzyme from T. thermophilus.
system II requires Mn for function. Thus, use of a FeSOD may benefit Biochemistry 34, 1646–1660.
[11] Ramirez, D.C., Gomez-Mejiba, S.E., Corbett, J.T., Deterding, L.J., Tomer, K.B. and
chloroplasts by avoiding competition for the same resource between
Mason, R.P. (2009) Cu, Zn-superoxide dismutase-driven free radical
two valuable enzymes. modifications: copper- and carbonate radical anion-initiated protein radical
In summary, the path by which we inherited Cu,ZnSOD is still a chemistry. Biochem. J. 417, 341–353.
mystery, but MnSODs appear to have been inherited by eukaryotes [12] Edwards, R.A., Whittaker, M.M., Whittaker, J.W., Jameson, G.B. and Baker, E.N.
(1998) Distinct metal environment in Fe-substituted manganese superoxide
from an archaeal ancestor whereas the FeSODs of protists may be dismutase provides a structural basis of metal specificity. J. Am. Chem. Soc.
derived from lateral gene transfer and the FeSODs of chloroplasts 120, 9684–9685.
are likely of cyanobacterial origin. [13] Jackson, T.A. and Brunold, T.C. (2004) Combined spectroscopic/computational
studies on Fe- and Mn-dependent superoxide dismutases: insights into
Archaeal SODs frequently display activity with either Mn or Fe second-sphere tuning of active site properties. Acc. Chem. Res. 37, 461–470.
and could represent intermediates between FeSODs that preceded [14] Vance, C.K. and Miller, A.-F. (2001) Novel insights into the basis for E. coli
them and the MnSODs of modern mitochondria. It is interesting to SODs metal ion specificity, from Mn-substituted Fe–SOD and its very high Em.
Biochemistry 40, 13079–13087.
note therefore that archaeal SODs such as that of Pyrobaculum lack [15] Vance, C.K. and Miller, A.-F. (1998) Spectroscopic comparisons of the pH
the Gln at position 69 that is typical of bacterial FeSODs but pro- dependencies of Fe-substituted-(Mn) superoxide dismutase and Fe-
vide an H-bonding partner for coordinated solvent from position superoxide dismutase. Biochemistry 37, 5518–5527.
[16] Bull, C., McClune, G.J. and Fee, J.A. (1983) The mechanism of iron EDTA
146 (based on multiple sequence alignment). Our work indicates catalyzed superoxide dismutation. J. Am. Chem. Soc. 105, 5290–5300.
that this would produce a much lower Em [14,47]. However the [17] Gupta, R. and Borovik, A.S. (2003) Monomeric MnIII/II and FeIII/II complexes
side chain provided by position 146 is a His instead of a Gln, and with terminal hydroxo and oxo lignads: probing reactivity via O–H bond
dissociation energies. J. Am. Chem. Soc. 125, 13234–13242.
our work suggests that this would result in a raised Em which
[18] Barrette, J., Sawyer, W.C., Fee, D.T. and Asada, K. (1983) Potentiometric
would therefore mitigate the effect of the placement at position titrations and oxidation–reduction potentials of several iron superoxide
146 and produce an intermediate Em [20]. This would be consistent dismutases. Biochemistry 22, 624–627.
with this archaeal SOD’s ability to utilize either metal ion as a basis [19] Miller, A.-F. and Sorkin, D.L. (1997) Superoxide dismutases: a molecular
perspective. Comments in Molecular and Cellular Biophysics 9, 1–48.
for SOD activity [21], potentially providing an evolutionary bridge [20] Vance, C.K. and Miller, A.-F. (1998) A simple proposal that can explain the
to Mn use. Mutation of the His to Gln would have produced the inactivity of metal-substituted superoxide dismutases. J. Am. Chem. Soc. 120,
fully depressed Em and favoured Mn use to the point that Fe was 461–467.
[21] Yikilmaz, E., Rodgers, D.W. and Miller, A.-F. (2006) The crucial importance of
no longer active. chemistry in the structure-function link: manipulating hydrogen bonding in
Thus the different kinds of SOD provide complementary win- iron-containing superoxide dismutase. Biochemistry 45, 1151–1161.
dows on the evolution of metalloenzymes. Old as they are, the [22] Miller, A.-F. (2008) Redox tuning over almost 1 V in a structurally-conserved
active site: lessons from Fe-containing superoxide dismutase. Acc. Chem. Res.
superoxide dismutases are sure to yield more surprises, answers 41, 501–510.
and questions. [23] Yamakura, F., Kobayashi, K., Ue, H. and Konno, M. (1995) The pH-
dependent changes of the enzymic activity and spectroscopic properties
of iron-substituted manganese superoxide dismutase. Eur. J. Biochem.
227, 700–706.
[24] Whittaker, M.M. and Whittaker, J.W. (1997) Mutagenesis of a proton linkage
12
The FeSOD gene has apparently been lost from the chloroplasts of some types of pathway in Escherichia coli manganese superoxide dismutase. Biochemistry
algae [59]Fink, R.C. and Scandalios, J.G. (2002). Molecular evolution and structure- 36, 8923–8931.
function relationships of the superoxide dismutase gene families in angiosperms and [25] Stallings, W.C., Metzger, A.L., Pattridge, K.A., Fee, J.A. and Ludwig, M.L. (1991)
their relationship to other eukaryotic and prokaryotic superoxide dismutases. Arch. Structure–function relationships in iron and manganese superoxide
Biochem. Biophys. 399, 19-36. dismutases. Free Rad. Res. Comm. 12–13, 259–268.
594 A.-F. Miller / FEBS Letters 586 (2012) 585–595

[26] Jackson, T.A., Xie, J., Yikilmaz, E., Miller, A.-F. and Brunold, T.C. (2002) [52] Pesce, A., Capasso, C., Battistoni, A., Folcarelli, S., Rotilio, G., Desideri, A. and
Spectroscopic and computational studies on iron and manganese superoxide Bolognesi, M. (1997) Unique structural features of the menomeric Cu, Zn
dismutases: nature of the chemical events associated with active site pKs. J. superoxide dismutase from Escherichia coli, revealed by X-ray
Am. Chem. Soc. 124, 10833–10845. crystallography. J. Mol. Biol. 274, 408–420.
[27] Miller, A.-F., Sorkin, D.L. and Padmakumar, K. (2005) Anion-binding [53] Cannio, R., Fiorentino, G., Morana, A., Rossi, M. and Bartolucci, S. (2000)
properties of reduced and oxidized iron-containing superoxide dismutase Oxygen: friend or foe? Archaeal superoxide dismutases in the protection of
reveal no requirement for Tyrosine 34. Biochemistry 44, 5969–5981. intra- and extracellular oxidative stress. Frontiers in Biosci. 5, 768–779.
[28] Fee, J.A., McClune, G.J., Lees, A.C., Zidovetzki, R. and Pecht, I. (1981) The pH [54] Ammendola, S. et al. (2008) Regulatory and structural differences in the Cu,
dependence of the spectral and anion binding properties of iron containing Zn-superoxide dismutases of Salmonella enterica and their significance for
superoxide dismutase from E. coli B: an explanation for the azide inhibition of virulence. J. Biol. Chem. 283, 13688–13699.
dismutase activity. Israel J. Chem. 21, 54–58. [55] Banci, L. et al. (2005) A prokaryotic superoxide dismutase paralog lacking two
[29] Miller, A.-F., Padmakumar, K., Sorkin, D.L., Karapetian, A. and Vance, C.K. Cu ligands: from largely unstructured in solution to ordered in the crystal.
(2003) Proton-coupled electron transfer in Fe-superoxide dismutase and Mn- Proc. Natl. Acad. Sci. USA 102, 7541–7546.
superoxide dismutase. J. Inorg. Biochem. 93, 71–83. [56] Wilkinson, S.R., Prathalingam, S.R., Taylor, M.C., Ahmed, A., Horn, D. and
[30] Jackson, T.A., Karapetian, A., Miller, A.-F. and Brunold, T.C. (2004) Kelly, J.M. (2006) Functional characterization of the iron superoxide
Spectroscopic and computational studies of the azide-adduct of dismutase gene repertoire in Trypanosoma brucei. Free Radic. Biol. Med. 40,
manganese superoxide dismutase: definitive assignment of the ligand 193–195.
responsible for the low-temperature thermochromism. J. Am. Chem. Soc. [57] Sturtz, L.A., Diekert, K., Jensen, L.T., Lill, R. and Culotta, V.C. (2001) A fraction
126, 12477–12491. of yeast Cu, Zn-superoxide dismutase and its metallochaperone, CCS, localize
[31] Jackson, T.A., Karapetian, A., Miller, A.-F. and Brunold, T.C. (2005) Probing the to the intermembrane space of mitochondria. J. Biol. Chem. 276, 38084–
geometric and electronic structures of the low temperature azide adduct and 38089.
the product-inhibited form of oxidized manganese superoxide dismutase. [58] Fréalle, E., Noël, C., Viscogliosi, E., Camus, D., Dei-Cas, E. and Delhaes, L.q.
Biochemistry 44, 1504–1520. (2005) Manganese superoxide dismutase in pathogenic fungi: an issue with
[32] Whittaker, J.W. and Whittaker, M.M. (1991) Active site spectral studies on pathophysiological and phylogenetic involvements. FEMS Immun. Med.
manganese superoxide dismutase. J. Am. Chem. Soc. 113, 5528–5540. Micro. 45, 411–422.
[33] Tabares, L.C., Cortez, N. and Un, S. (2007) Role of tyrosine 34 in the anion [59] Fink, R.C. and Scandalios, J.G. (2002) Molecular evolution and structure-
binding equilibria in manganese(II) superoxide dismutases. Biochemistry 46, function relationships of the superoxide dismutase gene families in
9320–9327. angiosperms and their relationship to other eukaryotic and prokaryotic
[34] Schwartz, A.L., Yikilmaz, E., Vance, C.K., Vathyam, S., Koder Jr., R.L. and Miller, superoxide dismutases. Arch. Biochem. Biophys. 399, 19–36.
A.-F. (2000) Mutational and spectroscopic studies of the significance of the [60] Bueno, P., Varela, J., Giménez-Gallego, G. and del Río, L.A. (1995) Peroxisomal
active site Gln to metal ion specificity in superoxide dismutase. J. Inorg. copper, zing superoxide dismutase - characterization of the isoenzyme from
Biochem. 80, 247–256. watermelon cotyledons. Plant. Physiol. 108, 1151–1160.
[35] Bull, C. and Fee, J.A. (1985) Steady-state kinetic studies of superoxide [61] McCord, J.M. and Fridovich, I. (1969) Superoxide dismutase an enzymic
dismutases: properties of the iron containing protein from Escherichia coli. J. function for erythrocuprein (hemocuprein). J. Biol. Chem. 244, 6049–
Am. Chem. Soc. 107, 3295–3304. 6055.
[36] Abreu, I.A., Rodriguez, J.A. and Cabelli, D.E. (2005) Theoretical studies of [62] Antonyuk, S.V., Strange, R.W., Marklund, S.L. and Hasnain, S.S. (2009) The
manganese and iron superoxide dismutases: superoxide binding and structure of human extracellular copper-zinc superoxide dismutase at 1.7 Å
superoxide oxidation. J. Phys. Chem. B 109, 24502–24509. resolution: insights into heparin and collagen binding. J. Mol. Biol. 388, 310–
[37] Miller, A.-F., Yikilmaz, E. and Vathyam, S. (2010) 15N-NMR characterization 326.
of His residues in and around the active site of FeSOD. Biochim. Biophys. Acta [63] Hong, Z., Lo Verde, P.T., Thakur, A., Hammarskjold, M.L. and Rekosh, D. (1993)
1804, 275–284. Schistosoma mansoni: a Cu/Zn superoxide dismutase is glycosylated when
[38] Maliekal, J. et al. (2002) Comparison and contrasts between the active site expressed in mammalian cells and localizes to a subtegumental region in
pKs of Mn-superoxide dismutase and those of Fe-superoxide dismutase. J. adult schistosomes. Exp. Parasitol. 76, 101–114.
Am. Chem. Soc. 124, 15064–15075. [64] Chang, L.Y., Slot, J.W., Geuze, H.J. and Crapo, J.D. (1988) Molecular
[39] Parker, M.W., Blake, C.C.F., Barra, D., Bossa, F., Schinina, M.E., Bannister, W.H. immunocytochemistry of the CuZn superoxide dismutase in rat
and Bannister, J.V. (1987) Structural identity between the iron- and hepatocytes. J. Cell. Biol. 107, 2169–2179.
manganese-containing superoxide dismutases. Protein Eng. 1, 393–400. [65] Weisinger, R.A. and Fridovich, I. (1973) Mitochondrial superoxide dismutase.
[40] Wintjens, R., Gilis, D. and Rooman, M. (2008) Mn/Fe superoxide dismutase Site of synthesis and intramitochondiral localization. J. Biol. Chem. 248,
interaction fingerprints and prediction of oligomerization and metal cofactor 4793–4796.
from sequence. Proteins, Str. Func. Bioinformat. 70, 1564–1577. [66] Kira, Y., Sato, E.F. and Inoue, M. (2002) Association of Cu, Zn-type superoxide
[41] Carlioz, A., Ludwig, M.L., Stallings, W.C., Fee, J.A., Steinman, H.M. and Touati, dismutase with mitochondria and peroxisomes. Arch. Biochem. Biophys. 399,
D. (1988) Iron superoxide dismutase. nucleotide sequence of the gene from 96–102.
Escherichia coli K12 and correlations with crystal structures. J. Biol. Chem. [67] Keller, G.-A., Warner, T.G., Steimer, K.S. and Hallewell, R.A. (1991) Cu, Zn
263, 1555–1562. superoxide dismutase is a peroxisomal enzyme in human firboblasts and
[42] Stoddard, B.L., Howell, P.L., Ringe, D. and Petsko, G.A. (1990) The 2.1-Å hepatoma cells. Proc. Natl. Acad. Sci. USA 88, 7381–7385.
resolution structure of iron superoxide dismutase from Pseudomonas ovalis. [68] Geller, B.L. and Winge, D.R. (1982) Rat liver Cu, Zn-superoxide dismutase. J.
Biochemistry 29, 8885–8893. Biol. Chem. 257, 8945–8952.
[43] Yamakura, F., Sugio, S., Hiraoka, B.Y., Ohmori, D. and Yokota, T. (2003) [69] Bordo, D., Djinovic, K. and Bolognesi, M. (1994) Conserved patterns in the Cu,
Pronounced conversion of the metal-specific activity of superoxide Zn superoxide dismutase family. J. Mol. Biol. 238, 366–386.
dismutase from Porphyromonas gingivalis by the Mutation of a Single [70] Tomalski, M.D., Eldreidge, R. and Miller, L.K. (1991) A baculovirus homolog of
Amino Acid (Gly155Thr) Located Apart from the Active Site. Biochemistry a Cu/Zn superoxide dismutase gene. Virology 184, 149–161.
42, 10790–10799. [71] Dos Santos, W.G., Pacheco, I., Liu, M.Y., Teixeira, M., Xavier, A.V. and LeGall, J.
[44] Yikilmaz, E., Porta, J., Grove, L.E., Vahedi-Faridi, A., Bronshteyn, Y., Brunold, (2000) Purification and characterization of an iron superoxide dismutase and
T.C., Borgstahl, G.E.O. and Miller, A.-F. (2007) How can a single second sphere a catalase from the sulfate-reducing bacterium Desulfovibrio gigas. J.
amino acid substitution cause reduction midpoint potential changes of Bacteriol. 182, 769–804.
hundreds of millivolts? J. Am. Chem. Soc. 129, 9927–9940. [72] Regelsberger, G., Atzenhofer, W., Ruker, F., Peschek, G.A., Jakopitsch, C.,
[45] Yikilmaz, E., Xie, J., Miller, A.-F. and Brunold, T.C. (2002) Hydrogen-bond- Paumann, M., Furtmuller, P.G. and Obinger, C. (2002) Biochemical
mediated tuning of the redox potential of the non-heme Fe site of superoxide characterization of a membrane-bound manganese-containing superoxide
dismutase. J. Am. Chem. Soc. 124, 3482–3483. dismutase from the cyanobacterium Anabaena PCC 7120. J. Biol. Chem. 277,
[46] Bafana, A., Dutt, S., Kumar, A., Kumar, S. and Ahuja, P.S. (2011) The basic and 43615–43622.
applied aspects for superoxide dismutase. J. Mol. Cat. B: Enzymatic 68, 129– [73] Li, T., Huang, X., Zhou, R.B., Liu, Y.F., Li, B., Nomura, C. and Zhao, J.D. (2002)
138. Differential expression, localization of Mn, Fe superoxide dismutases in the
[47] Wolfe-Simon, F., Grzebyk, D., Schofield, O. and Flakowski, P.G. (2005) The role heterocystous cyanobacterium Anabaena sp. strain PCC 7120. J. Bact. 184,
and evolution of superoxide dismutases in algae. J. Phycol. 41, 453–465. 5096–5103.
[48] Pilon, M., Ravet, K. and Tapken, W. (2011) The biogenesis and physiological [74] Esposito, L., Seydel, A., Aiello, R., Sorrentino, G., Cendron, L., Zanotti, G. and
function of chloroplast superoxide dismutases. Biochim. Biophys. Acta 1807, Zagari, A. (2008) The crystal structure of the superoxide dismutase from
989–998. Helicobacter pylori reveals a structured C-terminal extension. Biochim.
[49] Youn, H.-D., Kim, E.-J., Roe, J.-H., Hah, Y.C. and Kang, S.-O. (1996) A novel Biophys. Acta 1784, 1601–1606.
nickel-containing superoxide dismutase from Streptomyces spp. Biochem. J. [75] Amo, T., Atomi, H. and Imanaka, T. (2003) Biochemical properties and
318, 889–896. regulated gene expression of the superoxide dismutase from the facultatively
[50] Schmidt, A., Gube, M., Schmidt, A. and Kothe, E. (2009) In silico analysis of aerobic hyperthermophile Pyrobaculum calidifontis. J. Bacteriol. 185, 6340–
nickel containing superoxide dismutase evolution and regulation. J. Basic 6347.
Microbiol. 49, 109–118. [76] Atzenhofer, W., Regelsberger, G., Jacob, U., Peschek, G.A., Furtmuller, P.,
[51] Eitinger, T. (2004) In vivo production of active nickel superoxide dismutase Huber, R. and Obinger, C. (2002) The 2.0 Å resolution structure of the
from Prochlorococcus marinus MIT9313 is dependent on its cognate catalytic portion of a cyanobacterial membrane-bound manganese
peptidase. J. Bateriol. 186, 7821–7825. superoxide dismutase. J. Mol. Biol. 321, 479–489.
A.-F. Miller / FEBS Letters 586 (2012) 585–595 595

[77] Nakamura, T. and Uegaki, K. (2011) Crystal structure of the cambialistic superoxide dismutase from cowpeak root nodules. Plant. Physiol. 133, 773–
superoxide dismutase from Aeropyrum pernix K1 - insights into the enzyme 782.
mechanism and stability. FEBS J. 278, 598–609. [91] Rubio, M.C., Becana, M., Sato, S., James, E.K., Tabata, S. and Spaink, H.P. (2007)
[78] Yamano, S. and Maruyama, T. (1999) An azide-insensitive superoxide Characterization and genomic clones and expression analysis of the three
dismutase from a hyperthermophilic archaeon, Sulfolobus solfataricus. J. types of superoxide dismutases during nodule development in Lotus
Biochem. 125, 186–193. japonicus. Molecular Plant-Microbe Interact. 20, 262–275.
[79] Kardinahl, S., Anemüller, S. and Schäfer, G. (2000) The hyperthermostable Fe- [92] del Río, L.A., Sandalio, L.M., Altomare, D.A. and Zilinskas, B.A. (2003)
superoxide dismutase from the archaeon Acidianus ambivalens: Mitochondrial and peroxisomal manganese superoxide dismutase:
characterization, recombinant expression, crystallization and effects of differential expression during leaf senescence. J. Exp. Bot. 54, 923–933.
metal exchange. J. Biol. Chem. 381, 1089–1101. [93] Brouwer, M., Brouwer, T.H., Grater, W. and Brown-Peterson, N. (2003)
[80] Kirby, T.W., Lancaster, J.R. and Fridovich, I. (1981) Isolation and Replacement of a cytosolic copper/zinc superoxide dismutase by a novel
characterization of the iron-containing superoxide dismutase of cytosolic manganese superoxide dismutase in curstaceans that use copper
Methanobacterium bryantii. Arch. Biochem. Biophys. 210, 140–148. (haemocyanin) for oxygen transport. Biochem. J. 374, 219–228.
[81] Takao, M., Yasui, A. and Oikawa, A. (1991) Unique characteristics of [94] Holley, A.K., Dhar, S.K. and St Clair, D.K. (2010) Manganese superoxide
superoxide dismutase of a strictly anaerobic archaebacterium dismutase vs P53 regulation of mitochondrial ROS. Mitochondrion 10, 649–
Meinanobacterium thermoautotrophicum. J. Biol. Chem. 266, 14151–14154. 661.
[82] May, B.P. and Dennis, P.P. (1989) Evolution and regulation of the gene [95] Lisanti, M.P., Martinez-Outschoorn, U.e., Lin, Z., Pavlides, S., Whitaker-
encoding superoxide dismutase from the archaebacterium Halobacterium Menezes, D., Pestell, R.G., Howell, A. and Sotgia, F. (2011) Hydrogen
cutirubrum. J. Biol. Chem. 264, 12253–12258. peroxide fuels aging, inflammation, cancer metabolism and metastasis: the
[83] Loftus, B. et al. (2005) The genome of the protist parasite Entamoeba seed and teh soil also needs ‘fertilizer’. Cell Cycle 10, 2440–2449.
histolytica. Nature 433, 865–868. [96] Quiros-Gonzalez, I., Sainz, R.M., Hevia, D. and Mayo, J.C. (2011) MnSOD drives
[84] Bécuwe, P. et al. (1996) Characterization of iron-dependent endogenous neuroendocrine differentiation, androgen independence, and cell survival in
superoxide dismutase of Plasmodium falciparum. Mol. Biochem. Parasitol. prostate cancer cells. Free Radic. Biol. Med. 50, 525–536.
76, 125–134. [97] Mohr, A., Buneker, C., Gough, R.p. and Zwacka, R.M. (2008) MnSOD protects
[85] Asojo, O.A., Schott, E.J., Vasta, G.R. and Silva, A.M. (2006) Structures of colorectal cancer cells from TRAIL-induced apoptosis by inhibition of Smac/
PmSOD1 and PmSOD2, two superoxide dismutases from the protozoan DIABLO release. Oncogene 27, 763–774.
parasite Perkinsus marinus. Acta Crystallogr. Sect. F 62, 1072–1075. [98] Martin, W., Hoffmeisster, M., Rotte, C. and Henze, K. (2001) An overview of
[86] Sienkiewicz, N. et al. (2004) Identification of a mitochondrial superoxide endosymbiotic models for the origins of eukaryotes, their ATP-producing
dismutase with an unusual targeting sequence in Plasmodium falciparum. organelles (mitochondria and hydrogenosomes), and their heterotrophic
Mol. Biochem. Parasitol. 137, 121–132. lifestyle. Biol. Chem. 382, 1521–1539.
[87] Kanematsu, S. and Asada, K. (1979) Arch. Biochem. Biophys. 195, 535–545. [99] Poole, A.M. and Neumann, N. (2011) Reconciling an archaeal origin of
[88] Narasipura, S.D., Chaturvedi, V. and Chaturvedi, S. (2005) Characterization of eukaryotes with engulfment: a biologically plausible update of the Eocyte
Cryptococcus neoformans variety fattii SOD2 reveals distinct roles of the two hypothesis. Res. Microbiol. 162, 71–76.
superoxide dismutases in fungal biology and virulence. Mol. Microbiol. 55, [100] Gray, M.W., Burger, G. and Lang, B.F. (1999) Mitochondrial evolution. Science
1782–1800. 283, 1476–1481.
[89] Van Camp, W., Bowler, C., Villarroel, R., Tsang, E.W.T., Van Montagu, M. and [101] Dufernez, F., Yernaux, C., Gerbod, D., Noël, C., Chauvenet, M., Wintjens, R.,
Inze, D. (1990) Characterization of iron superoxide dismutase cDNAs from Edgcomb, V.P., Capron, M., Opperdoes, F.R. and Viscogliosi, E. (2006) The
plants obtained by genetic complementation in Escherichia coli. Proc. Natl. presence of four iron-containing superoxide dismutase isozymes in
Acad. Sci. USA 87, 9903–9907. Trypanosomatidae: Characterization, subcellular localization, and
[90] Moran, J.F., James, E.K., Rubio, M.C., Sarath, G., Klucas, R.V. and Becana, M. phylogenetic origin in Trypanosoma brucei. Free Radic. Biol. Med. 40, 210–225.
(2003) Functional characterization and expression of a cytosolic iron-

You might also like