Você está na página 1de 10

REVIEWS

NEURODEGENERATIVE DISEASES
AND OXIDATIVE STRESS
Kevin J. Barnham*, Colin L. Masters* and Ashley I. Bush*‡
Oxidative stress has been implicated in the progression of Alzheimer’s disease, Parkinson’s
disease and amyotrophic lateral sclerosis. Oxygen is vital for life but is also potentially dangerous,
and a complex system of checks and balances exists for utilizing this essential element. Oxidative
stress is the result of an imbalance in pro-oxidant/antioxidant homeostasis that leads to the
generation of toxic reactive oxygen species. The systems in place to cope with the biochemistry
of oxygen are complex, and many questions about the mechanisms of oxygen regulation remain
unanswered. However, this same complexity provides a number of therapeutic targets, and
different strategies, including novel metal–protein attenuating compounds, aimed at a variety of
targets have shown promise in clinical studies.

HIPPOCAMPUS
Neurodegenerative diseases, such as Alzheimer’s disease unsaturated fatty acids. The chain reaction leads to the
A region of the brain consisting (AD), Parkinson’s disease (PD) and amyotrophic lateral formation of breakdown products including 4-hydroxy-
of the grey matter at the bottom sclerosis (ALS), are defined by the progressive loss of 2,3-nonenal (HNE), acrolein, malondialdehyde and
of the lateral ventricle that is specific neuronal cell populations and are associated F2-isoprostanes. Elevated HNE levels have been observed
involved in motivation, emotion
with protein aggregates. A common feature of these in AD1,2 and PD3 brain tissue, whereas increased HNE
and the formation of memory.
diseases is extensive evidence of oxidative stress, which has been observed in the cerebrospinal fluid (CSF) of
might be responsible for the dysfunction or death of ALS patients4. Acrolein, thiobarbituric acid-reactive
neuronal cells that contributes to disease pathogenesis. substances (TBARs, the most prevalent substrate of
Oxidative stress is the result of unregulated production which is malondialdehyde) and F2-isoprostanes are
of reactive oxygen species (ROS), such as hydrogen per- all increased in AD brains relative to age-matched
oxide, nitric oxide, superoxide and the highly reactive controls5. Malondialdehyde is increased in PD brains3,
hydroxyl radicals. High oxygen consumption, relatively whereas increased TBARs have been observed in the
low antioxidant levels and low regenerative capacity plasma of ALS patients6.
*Department of Pathology, result in brain tissue being susceptible to oxidative All four bases of DNA are susceptible to oxidative
The University of Melbourne, damage. Although a contentious idea, many believe that damage involving hydroxylation7, and significant pro-
and The Mental Health
Research Institute of Victoria,
oxidative stress contributes to the neurodegeneration tein carbonylation (for example, of creatine kinase and
Victoria 3010, Australia. associated with these diseases; as a result, a variety of β-actin2) and nitration is observed in AD brains,

Laboratory for Oxidation therapeutic strategies target this phenomena. whereas increased levels of 8-hydroxyguanine and
Biology, Genetics and Aging 8-hydroxy-2-deoxyguanosine are observed in PD brains;
Research Unit and Evidence of oxidative stress the selective attack on guanine bases implies OH radicals
Department of Psychiatry,
Harvard Medical School, Unsaturated lipids are particularly susceptible to oxida- as the oxidative species8.
Massachusetts General tive modification and lipid peroxidation is a sensitive Cells have endogenous defence mechanisms against
Hospital East, Charlestown, marker of oxidative stress. Lipid peroxidation is the oxidative stress and changes in these are also used as
Massachusetts 02129, USA. result of attack by radicals on the double bond of unsatu- markers of oxidative stress. In the AD brain, the activity
Correspondence to A.I.B.
e-mail: bush@helix.mgh.
rated fatty acids, such as linoleic acid and arachidonic of the antioxidant proteins catalase, superoxide dismutase
harvard.edu acid, to generate highly reactive lipid peroxy radicals (SOD), glutathione peroxidase and glutathione reductase
doi:10.1038/nrd1330 that initiate a chain reaction of further attacks on other are increased in the HIPPOCAMPUS and amygdala9,10.

NATURE REVIEWS | DRUG DISCOVERY VOLUME 3 | MARCH 2004 | 2 0 5


REVIEWS

neurological conditions, including epilepsy and stroke,


Box 1 | Friedreich’s ataxia
as well as the neurodegenerative diseases AD, PD, ALS
An abnormal GAA trinucleotide expansion within the first intron of the gene encoding and Huntington’s disease (reviewed in REF. 21).
frataxin, a mitochondrial protein, leads to a deficiency of frataxin and is the major cause of
Friedreich’s ataxia (FA)118. One consequence of this deficiency is iron accumulation within How are ROS generated?
the mitochondria, which leads to increased oxidative stress resulting in cardiomyopathy The generation of ROS requires the activation of molecu-
and neurodegeneration. Another consequence of the iron overload is a breakdown in the lar oxygen. At the quantum level, O2 is in a triplet spin-
mitochondrial respiratory chain due to a deficiency in Fe-S enzymes119. Two therapeutic state and consequently interactions with most organic
strategies are now being investigated that target the Fe-mediated oxidative stress. The first molecules are spin-forbidden (that is, they are energeti-
involves idebenone, a free-radical scavenger analogue of coenzyme Q10, although results
cally unfavourable). As utilization of O2 is a prerequisite
from trials with this compound have been mixed. Idebenone has been shown to reduce
for most life forms, this intrinsic feature of O2 must be
oxidative damage (as measured by urinary 8-hydroxy-2’-deoxyguanosine levels) and
overcome. Organisms have evolved a range of metallo-
rescue respiratory chain function120. Several pilot clinical studies have shown that
enzymes to take advantage of the interactions between
idebenone has significant benefit with respect to the hypertrophic cardiomyopathy
associated with FA; however, there were no indications of improved neurological O2 and metal ions to activate molecular oxygen as ROS;
functions120–122. An alternative strategy is to target the excess Fe3+ via the use of metal the subsequent free radicals are an intrinsic part of
chelators. Fibroblasts isolated from FA patients are particularly susceptible to oxidative normal metabolism. As ROS are also toxic, cells have
insults, and in a cell-culture model using these cells the iron chelator desferrioxamine developed highly elaborate means of regulating both
(DFO) was found to be protective123. The iron overload associated with FA is a specific metal ion interactions and the generation of ROS. The
overload of the mitochondria, not a gross overload; chelation strategies therefore need same properties that cells harness for beneficial means
to target this iron specifically. DFO has poor bioavailability and is unlikely to have a become destructive when the regulatory processes
significant effect on the iron levels within the mitochondria. A new class of iron breakdown.
chelators based on 2-pyridylcarboxaldehyde isonicotinoyl hydrazones that can target ROS are produced by a number of different pathways,
mitochondrial iron has recently been developed, but the applicability of these molecules including direct interactions between REDOX-active metals
as potential therapeutics for FA is still being assessed124. and oxygen species via reactions such as the FENTON and
Haber–Weiss reactions, or via indirect pathways involving
the calcium activation of metallo-enzymes such as
How do ROS kill cells? phospholipases, nitric oxide synthase and xanthine
As their name suggests, ROS are very reactive and will dehydrogenase (also known as xanthine oxidase)22.
react with a multitude of different molecules to initiate Calcium is crucial to signal transduction and as such is
neuronal cell death and hence neurodegeneration both sensitive to a number of different stimuli and able
through an array of different pathways. Oxidative stress to elicit a variety of different cellular responses. There is a
(that is, oxidation of lipids, proteins and DNA) results in large body of evidence documenting disruption of
impaired cellular functions and the formation of toxic calcium homeostasis in neurodegenerative diseases,
species, such as peroxides, alcohols, aldehydes, ketones leading to a breakdown in a large number of cellular
and cholesterol oxide. Cholesterol oxide is toxic to processes. However, most of these different signalling
lymphocytes and blood vessel macrophages11. pathways rely on feedback signalling and it is conse-
The oxidatively modified lipids acrolein and HNE quently difficult to differentiate cause and effect. An
induce toxicity by crosslinking to cystine, lysine and example of this is the interplay between calcium sig-
histidine residues via a Michael addition. Acrolein down- nalling and ROS generation: although increases in
regulates the uptake of glutamate and glucose from cell intracellular calcium have been reported to induce the
culture12, whereas HNE modifies proteins resulting in a production of ROS22, at subtoxic levels ROS have been
multitude of effects, including inhibition of the neuronal shown to be important cell signallers and will induce an
glucose transporter type-3, the glutamate transporter increase in cytosolic calcium23,24. This has lead to much
GLT-1 (REF. 13), as well as the Na++K+ ATPases14. HNE debate as to whether the observed disruption of calcium
activates c-Jun aminoterminal kinases and mitogen- homeostasis is a cause or consequence of ROS generation.
activated protein kinase-1 (also known as p38), thereby As a general principle, the chemical origin of the
stimulating an apoptotic cascade15. Modifications to majority of ROS is the reaction of molecular oxygen
proteins result in the impairment of enzymes (for with the redox-active metals copper and iron25. The
example, glutamine synthase, superoxide dismutase), ability of these metal ions to occupy multiple valence
whereas ROS interactions with DNA lead to mutations. states and undertake facile redox cycling, thereby acti-
The excessive generation of ROS leads to dysregula- vating molecular oxygen, has been utilized by a variety
tion of intracellular calcium signalling, and such dysreg- of enzymes. However, unregulated redox-active metals
ulation has been widely observed in neurodegenerative will inappropriately react with oxygen to generate ROS.
REDOX
diseases in which aberrant calcium levels stimulate We have proposed that the proteins implicated in several
A reversible chemical reaction multiple pathways that ultimately induce an apoptotic age-dependent neurodegenerative disorders (Aβ in AD,
in which one reaction is an cascade (reviewed in REFS 16–19). α-synuclein in PD, SOD1 in ALS, frataxin in Friedreich’s
oxidation reaction and the One of the downstream events that occurs in ataxia (BOX 1) and α-B-crystallin in cataracts), might
reverse a reduction.
response to an ROS-induced calcium influx is an excito- abnormally present Cu2+ or Fe3+ ligands for inappropriate
FENTON REACTION toxic response19,20, that is, activation of glutamate recep- reaction with O2 (FIG. 1). These proteins might have some
Mn+ + H2O2 → M(n+1)+ + OH– tors that trigger a cascade of events leading to cell death. aspect of their function subserved by these metal ions,
+ OH•. Excitotoxic responses have been implicated in several which normally occupy higher-affinity, embedded,

206 | MARCH 2004 | VOLUME 3 www.nature.com/reviews/drugdisc


REVIEWS

In vitro studies have shown that low µM levels of Zn2+


O2 O2–/H2O2
will induce protease-resistant aggregation and precipita-
Catechols tion of Aβ39. Cu2+ and Fe3+ also induce peptide aggrega-
Vitamin C e– tion that is exaggerated at acidic pH40. These metals are
Cholesterol, Fe2+/Cu+
arachidonic Cu+ Cu2+ normally found at high concentrations in the region of
acid the brain most susceptible to AD neurodegeneration.
and so on
Drugs OH• During neurotransmission, high concentrations of Zn
(300 µM) and Cu (30 µM) are released, which might
explain why Aβ precipitation into amyloid commences
in the synapse36,41.
The cause of the neuronal cell loss in AD might be
Aβ Lipid peroxidation
related to oxidative stress from excessive free-radical
SOD Protein oxidation generation29,30,38,42,43. We have proposed that the major
α-synuclein Protein aggregation source of oxidative stress and free-radical production in
α-crystallin DNA/RNA adducts
the brain in AD is the transition metals Cu and Fe29,43
Figure 1 | ROS generation by abnormal reaction of O2 with protein-bound Fe or Cu. One of when bound to Aβ.
the consequences of normal aging is that the levels of the redox-active metals copper and iron There is a large body of evidence indicating that the
in the brain increase. This increase could lead to hypermetallation of proteins that normally bind homeostasis of Zn, Cu and Fe, and their respective
redox-active metals at shielded sites. Adventitial binding — for example; at a loading site — will
increase the likelihood that ROS are generated inappropriately as illustrated, leading to the
binding proteins, are significantly altered in the AD
oxidative stress that is observed in neurodegenerative diseases. Aβ, amyloid-β; ROS, reactive brain (reviewed in REF. 44). A microparticle-induced
oxygen species; SOD, superoxide dismutase. X-ray emission analysis of the cortical and accessory
basal nuclei of the amygdala indicated that these metals
accumulate in the NEUROPILE of the AD brain in concen-
trations that are three- to fivefold increased compared
redox-shielded binding sites. As metal concentration with age-matched controls37. Evidence for abnormal Cu
rises in the brain with age, the probability increases that homeostasis in AD includes a 2.2-times increase in the
a redox-competent, low-affinity metal-binding site will concentration of Cu in CSF45, and an accompanying
recruit a metal ion from the normally redox-silent cellular increase of plasma Cu in AD46. There is an extensive lit-
pool. In this manner, proteins such as Aβ can harness erature describing abnormal levels of Fe and Fe-binding
endogenous biometals to foster the release of inappro- proteins in AD47. Notably, the Fe that is found within
priate redox activity and ROS generation. the amyloid deposits of human brain and in amyloid-
bearing APP transgenic mice is redox-active38,48. Raman
Metal ions and neurodegenerative diseases spectroscopy studies have demonstrated that Zn2+ and
The dominant risk factor associated with the neuro- Cu2+ are coordinated to the histidine residues of the
degenerative diseases is increasing age. Several studies in deposited Aβ in the senile plaque cores from diseased
mice have shown that one of the consequences of normal brain tissue and that the sulphur atom of methionine 35
aging is a rise in the levels of copper and iron in brain of Aβ is oxidized, which is indicative of a pro-oxidant
tissue (FIG. 1)26–28. The brain is an organ that concentrates environment49.
metal ions and recent evidence suggests that a break-
down in metal homeostasis is a key factor in a variety of Toxic mechanism(s) of Aβ. Synthetic Aβ is toxic to cells
age-related neurodegenerative diseases29,30. in the presence of Cu2+, but this toxicity is inhibited by
extracellular catalase, which implicates H2O2 in the toxic
Alzheimer’s disease. Genetic evidence from cases of pathway50,51. When Cu2+ or Fe3+ coordinate Aβ, exten-
familial AD indicates that Aβ metabolism is linked to sive redox chemical reactions take place that reduce the
the disease31,32. AD is characterized by the deposition oxidation state of both metals and produce H2O2 from
of AMYLOID plaques, the major constituent being the amy- O2 in a catalytic manner51–54. The formal reduction
loid-β peptide (Aβ) that is cleaved from the membrane- potential of Cu2+ to Cu+ by Aβ42 is highly positive
bound amyloid precursor protein (APP)33–35. Although (approximately +500–550 mV versus Ag/AgCl) and
the function of APP is unknown, recent evidence suggests characteristic of strongly reducing cupro-proteins52.
it functions in maintaining copper homeostasis (BOX 2). The generation of H2O2 in the presence of the reduced
form of the metal creates conditions in which Fenton
Interactions of Aβ and metals. Recent results have high- chemistry occurs with the generation of highly toxic
lighted the importance of Zn2+ in amyloid plaque for- OH• radicals53.
mation; for example, age- and female-sex-related plaque The toxicity of Aβ to neuronal cells is enhanced by
AMYLOID
Protein/peptide deposited formation in Tg2576 transgenic mice was reduced by the presence of Cu, which is usually present in culture
in diseased tissue, with high genetic ablation of the zinc transporter 3 protein, which media51. Aβ possesses histidine residues at positions 6,
β-sheet structure. is required for zinc transport into synaptic vesicles36. 13 and 14, which form a structural element that enables
The plaques could be described as metallic sinks Aβ to coordinate transition metal ions. A variety of
NEUROPILE
The mass of closely packed nerve
because remarkably high concentrations of Cu (400 spectroscopic studies have confirmed that the histidine
cell processes comprising the µM), Zn (1 mM) and Fe (1 mM) have been found residues constitute the principal site(s) of metal coordi-
central part of a ganglion. within the amyloid deposits in AD-affected brains37,38. nation. The interactions of synthetic Aβ with copper,

NATURE REVIEWS | DRUG DISCOVERY VOLUME 3 | MARCH 2004 | 2 0 7


REVIEWS

Box 2 | Amyloid precursor protein and copper homeostasis


To prevent transition-metal-mediated oxidative stress, cells have
evolved complex metal-transport systems that deliver Cu and
Fe to metallo-enzymes and proteins. These include the Cu
chaperone for superoxide dismutase 1 (SOD1), which facilitates
insertion of Cu into the active site of SOD125, and the Cu ATPase
that is mutated in Wilson’s disease126. The chaperones direct the
Cu atoms to specific intracellular proteins, which essentially
results in an absence of unbound Cu in the intracellular
environment127. Therefore, cuproproteins have an important
role in maintaining cellular Cu metabolism128. Amyloid
precursor protein (APP) is a ubiquitously expressed, high-
turnover protein and recent data suggest that APP and Aβ could
have some role in metal homeostasis. Utilizing transgenic mouse models it has been shown that overexpression of the
carboxyl-terminal fragment of APP, which contains Aβ, results in significantly reduced copper and iron levels in transgenic
mouse brain; overexpression of APP in Tg2576 transgenic mice results in significantly reduced copper, but not iron28,
whereas APP knockout mice have increased copper levels in the brain and liver129. Copper levels can modulate APP
processing130,131, with higher copper levels resulting in a reduction in Aβ production and a consequential increase in the
non-amyloidogenic p3 form of the peptide132. The effect of increasing brain Cu levels in reducing cerebral Aβ burden in
APP transgenic mice130,131 might seem to superficially contradict findings that link the toxicity of the peptide as well as
limited aggregation to binding Cu. However, at neutral pH Cu2+ displaces Zn2+ from Aβ, and induces far less precipitation
than Zn2+, which might also contribute to the Aβ-reducing effects of increasing Cu levels in the cerebrospinal fluid.
Independent copper-binding sites have been identified on both Aβ and APP, which might subserve a physiological
relationship with Cu. The structure of the APP copper-binding domain has recently been solved (Protein Data Bank
accession number: 1OWT; see figure) and found to contain a novel tetrahedral copper-binding site consisting of two
histidine residues (147, 151), a tyrosine (168) and methionine (170) that favours Cu(I) coordination133. The solvent
accessibility of this site, structural homology to copper chaperones, and the role of APP in neuronal copper homeostasis are
consistent with APP acting as a neuronal metallotransporter. Intriguingly, it appears as though messenger RNA of APP is
iron regulated through the APP 5’-untranslated region and this also points to a role for iron in the metabolism of APP134.

zinc and iron have been probed using nuclear magnetic model of high-affinity Cu/Zn binding to Aβ indicates
resonance and electron paramagnetic resonance spec- that Aβ will form a membrane-embedded hexamer
troscopy55,56. Both in aqueous solution and in membrane- when bound in that form55,56,63. We have developed the
mimetic environments, coordination of the metal ion is hypothesis that when Aβ becomes hyper-metallated
consistent, establishing that the three histidine residues by redox-active metal ions as a consequence of age-
are all involved in the coordination, along with an oxygen dependent elevations in tissue metal concentrations28,
ligand. Metal coordination is a highly cooperative event, subsequent oxidation releases soluble oxidized forms of
and the cooperativity is abolished by methylation of the the peptide that resist clearance. This might explain how
histidine side chains; this indicates that the second coop- zinc originating from the synapse becomes so enriched
erative site would result in a histidine residue(s) bridging in amyloid in AD. Our model suggests that in health
between metal ions with the resulting coordination soluble Aβ is not present in the cortical synapse. In AD,
sphere reminiscent of that observed in SOD1 (FIG. 2). soluble oxidized Aβ accumulates within the synapse, at
In addition to this site, Aβ is able to bind another 2.5 which the high Zn2+ concentrations precipitate the
equivalents of metal57, and although the nature of these copper/iron-metallated Aβ, creating a reservoir of
other lower-affinity sites has yet to be characterized at potentially toxic Aβ that is in dissociable equilibrium
least one of them is redox active51 and might be impor- with the soluble pool. The Zn2+ in the amyloid mass
tant in ROS generation. partially quenches H2O2 production, which explains
Reactions of Cu2+ with Aβ result in the oxidative why plaque amyloid burden correlates poorly with
modification of the sulphur atom of Met35 (REF. 58) and clinical dementia54, whereas soluble Aβ levels correlate
can cause covalent crosslinking of Aβ that yields soluble well with clinical severity61.
oligomers and other adducts57,59. The various forms of Although most researchers consider Aβ to be toxic
Aβ in the AD brain are usually oxidatively modified60. junk, data are emerging which suggest that Aβ could
DEMENTIA
We have proposed that such oxidative modification have an important physiological role as an antioxidant,
Mental deterioration of organic
or functional origin. might contribute to the release of abnormally soluble and that this function is corrupted as a result of the
forms of Aβ from the membrane; it is these soluble forms aging process; evidence supporting this concept has
SUBSTANTIA NIGRA that correlate with DEMENTIA61. Aβ with sulphur of Met35 recently been reviewed30,64.
A small area of the brain oxidized is more soluble and has a lower affinity for
containing a cluster of dark-
pigmented nerve cells that
membranes than the reduced form of Aβ, but is still Parkinson’s disease. PD is characterized by the loss of
produces dopamine for toxic to neurons58. In the normal brain, most of the Aβ dopaminergic neurons of the SUBSTANTIA NIGRA and the
neurotransmission. is associated with membranes62, and our structural deposition of intracellular inclusion bodies. The principal

208 | MARCH 2004 | VOLUME 3 www.nature.com/reviews/drugdisc


REVIEWS

Deposition

Aβ + M Aβ/M (Aβ/M)2 (Aβ/M)n

Excitotoxicity Memantine

MPAC

Extracellular H2O2 Calcium


dysregulation

•OH– modified Aβ ROS Intracellular H2O2

Antioxidants

Figure 2 | Oxidative stress in Alzheimer’s disease. Aβ coordinates the metal ions Zn, Fe and Cu, which induces aggregation
and, in the case of Fe and Cu, the generation of H2O2. The H2O2 can initiate a number of different events, including Fenton reactions
to form toxic hydroxyl radicals and calcium dysregulation. As calcium is pivotal in signal transduction, it can induce further
production of ROS and elicit an excitotoxicity response. Therapeutic strategies either in the clinic or undergoing clinical development
are noted in yellow boxes. Memantine, which is presently used in the clinic, inhibits excitotoxicity by targeting the N-methyl-D-
aspartate receptor. Antioxidants deactivate the generated ROS and MPACs seek to inhibit metal interactions with Aβ and prevent
the subsequent formation of ROS. MPACs, metal–protein attenuating compound; ROS, reactive oxygen species.

protein component of these deposits is α-synuclein65, and that the iron bound to neuromelanin is redox
which is ubiquitously expressed in the brain; mutations active76. The oxidative stress associated with PD could be
of α-synuclein (A30P and A53T) contribute to familial the result of a breakdown in the regulation of dopamine
forms of the disease66. (neuromelanin)/iron biochemistry (FIG. 3).
A characteristic feature of the neurons within the sub- A diverse array of evidence is emerging that α-synu-
stantia nigra is the age-dependent accumulation of neuro- clein has a role in modulating the activity of dopamine.
melanin67. In PD, these neuromelanin-containing cells The A53T mutation associated with familial PD impairs
are most likely to be lost68. Neuromelanin is a dark brown vesicular storage of dopamine77,78, which leads to the
pigment that accumulates metal ions, particularly iron. accumulation of dopamine in the cytoplasm and subse-
Although the composition of neuromelanin has not been quent generation of ROS through its interaction with
rigorously characterized, it is known that it consists pri- iron, a process that increases with age. The mutations in
marily of the products of dopamine redox chemistry69,70. α-synuclein have been shown to alter the expression of
Dopamine is an essential neurotransmitter, but as it is dihydropteridine reductase, which indirectly regulates
a catechol it is also a good metal chelator, and a potential the synthesis of dopamine79. Co-immunoprecipitation
electron donor (that is, a metal reductant). Dopamine experiments have shown that α-synuclein forms stable
coordinates metals such as Cu2+ and Fe3+ (REF. 71), reduces complexes with the human dopamine transporter,
the oxidation state of the metal, and subsequently engen- thereby inhibiting uptake of dopamine by its trans-
ders production of H2O2, setting up conditions for porter80 and that α-synuclein can regulate dopamine
Fenton chemistry. Synthetic melanins are produced by synthesis by inhibiting tyrosine hydroxylase81. The link
incubating dopamine with Cu2+ and Fe3+ (REF. 72). The between α-synuclein and redox chemistry associated
purpose (if any) of neuromelanin is unknown, but it has with iron-bound dopamine/neuromelanin has been
been postulated that it protects against dopamine- given further credence by a study showing that initiation
induced redox-associated toxicity73,74. At low iron con- of Lewy body formation coincides with α-synuclein
centrations, melanins are known to have antioxidant deposition exclusively within lipofuscin and neuro-
properties, but at higher metal loads melanins are pro- melanin deposits82; in addition, α-synuclein crosslinked
oxidant75. Another postulated role for neuromelanin is as to neuromelanin has been reported83. The breakdown in
an iron-storage molecule. Double et al.72 have shown α-synuclein-modulated dopamine homeostasis is con-
that neuromelanin isolated from human substantia sistent with the recent observation that the pathogenicity
nigra has both high- and low-affinity Fe3+-binding sites, of mutant α-synuclein is dopamine dependent84.

NATURE REVIEWS | DRUG DISCOVERY VOLUME 3 | MARCH 2004 | 2 0 9


REVIEWS

Mutations of SOD, a cupro-enzyme that detoxifies


the ROS superoxide, can convert the protein from an
Aggregation
anti-oxidant to a pro-oxidant capable of causing oxida-
tive insults. Evidence that inappropriate metal-mediated
Fe/ROS redox chemistry is central to the progression of ALS
includes the observation that copper chelators inhibit
α-synuclein
the course of the disease in both cell culture and mouse
Vesicular DA DA models92–95. These initial observations suggested that the
Fe toxicity associated with mutant SOD is the result of a
Fe corruption of the active site of this enzyme96. However,
Neuron
stimulation if the copper at the active site of SOD is the culprit
behind the toxicity, then knocking out the SOD copper
MPP+ ROS NM chaperone (CCS), which loads copper into the active
Fe site, should abolish the observed toxicity. To test this
hypothesis, Subramaniam97 crossed CCS knockout mice
DA
release with an SOD mutant ALS mouse model; the phenotype
of this cross showed reduced SOD activity, which is con-
Oxidative stress sistent with a low copper load in the active site, but the
mice still developed the ALS phenotype. These results
indicate that ALS is not due to a corruption of the
active site. However, these experiments did not examine
DA the possibility that other redox-active, lower-affinity
Figure 3 | Oxidative stress in Parkinson’s disease. Dopamine (DA) required for neuronal
metal-binding sites exist on SOD.
signalling is vesicle bound and redox inert; when DA is released from the vesicle into the Two such sites have been identified: copper can be
cytoplasm it is able to coordinate Fe and undergo redox reactions that result in the formation of incorrectly incorporated into the zinc site98, and in vitro
neuromelanin (NM) and reactive oxygen species (ROS). Neuromelanin will also coordinate Fe studies99 with an H46R mutant SOD linked to familial
and produce ROS. The equilibrium between vesicle bound dopamine and cytoplasmic ALS, and which has no SOD activity, have shown that a
dopamine is regulated by α-synuclein; mutations in this protein shift the dopamine equilibrium surface-exposed cysteine residue in SOD is also capable
in favour of the cytoplasm. In the presence of Fe and under conditions of oxidative stress
of coordinating copper and is redox active. As the con-
α-synuclein will aggregate and form deposits. MPP, 1-methyl-4-phenyl pyridine.
centration of copper rises with age28, the possibility that
these lower-affinity metal-binding sites are occupied is
increased and the hypothesis that these sites might be
In addition to the regulation of dopamine by α-syn- responsible for aberrant redox chemistry (FIG. 4), and
uclein, studies have shown a direct interaction of the generation of ROS and subsequent toxicity,
α-synuclein with metal ions, leading to protein aggre- remains untested100.
gation85–87. Methionine oxidation inhibits α-synuclein
aggregation; however, in the presence of certain metal Therapeutic options
ions aggregation and fibrillization of α-synuclein still For drugs to be effective treatments of neurodegenera-
occurs88, which highlights a potential role of metals and tive diseases they must be capable of penetrating the
oxidative stress in the deposition of α-synuclein. blood–brain barrier (BBB) (BOX 3).

Amyotrophic lateral sclerosis. ALS is distinguished by the Antioxidants. Antioxidants are molecules that react
loss of the lower motor neurons of the spinal cord and preferentially with ROS to inactivate them and have
upper motor neurons in the cerebral CORTEX; as with AD excited great interest because of their therapeutic
and PD there are both sporadic and familial forms of the potential. A number of studies have shown beneficial
disease. Like the other neurodegenerative diseases, ALS is effects of antioxidants in cell culture toxicity studies
characterized by the deposition of a misfolded protein in (reviewed in REF. 101). Antioxidants are part of our regular
neural tissue, in this instance copper/zinc SOD89. There dietary requirements and numerous claims have been
are more than 100 mutations of SOD associated with the made for a variety of antioxidant supplements, including
familial forms of the disease. Through transgenic mouse vitamin C, ubiquinone lipoic acid, β-carotene, creatine,
studies it has been shown that these mutations lead to a melatonin, curcumin and the red-wine micronutrients,
toxic gain of function by SOD90. The nature of this gain which have a high content of phenolic and flavonoid
of function is widely debated, and there are two main compounds.
theories: one suggests that the toxicity is due to misfolded Despite the large volume of positive cell culture data,
aggregated forms of SOD, whereas the other proposes there has been limited clinical evaluation of antioxidants;
that SOD becomes a pro-oxidant protein generating α-tocopherol (vitamin E) has been evaluated in both AD
ROS. The merits or otherwise of these two hypotheses and PD, and was found to have no beneficial effects in
have recently been reviewed91, and as the aggregation PD102. However, positive effects were noted in the AD
CORTEX
The unmyelinated neurons (the
mechanism lies outside the scope of this review it will trial103, with an increase in the median survival time of
grey matter) forming the outer not be discussed here, except to note that the two theories 230 days for the treated group and a significant delay
layer of the cerebrum. might not be mutually exclusive. in institutionalization, but there was no improvement in

210 | MARCH 2004 | VOLUME 3 www.nature.com/reviews/drugdisc


REVIEWS

2H+ severe range of the disease105. (Recent developments


Cu/ZnSOD H2O2 with memantine have been reviewed in REFS 106,107.)
2O2• – Amantadine, a partial NMDA antagonist, has been used
in the treatment of PD for some time with modest
Age-dependent clinical gains, including increased survival times108.
increase in Cu Riluzole is the only drug presently used to treat ALS
with any proven efficacy — the effects of Riluzole are
O2
also modest but it is able to extend mean survival time
Cu/ZnSOD H2O2 + OONO– of patients by several months109. The mode of action of
NO Riluzole is not fully understood but it is believed to
Cu inhibit glutamate release (and therefore to reduce excito-
toxicity) by interfering with sodium channels.
Neurotrophins have protective effects against oxida-
tive stress110; however, therapeutic applications are limited
by a lack of BBB permeability. Although efforts are con-
Low-affinity bound Cu ROS oxidative stress tinuing with regard to the design of small-molecule
Figure 4 | Oxidative stress in amyotrophic lateral sclerosis. stimulators of endogenous neurotrophins, an alterna-
The normal function of superoxide dismutase (SOD) is to tive approach to increasing endogenous neurotrophins
convert toxic superoxide radicals into H2O2 that are subse- — for example, brain-derived neurotrophic factor,
quently inactivated by catalase. In amyotrophic lateral sclerosis which is known to enhance learning and memory, and
this antioxidant protein is converted into a pro-oxidant protein.
also protect against oxidative stress — is caloric restric-
With age-dependent increases in copper levels, low-affinity
copper sites on SOD such as Cys111 are occupied; these sites
tion. It has been proposed that a restricted diet with
are redox active and give rise to aberrant redox chemistry and limited calorie intake induces a mild cellular stress
subsequent reactive oxygen species (ROS) generation. response, which in turn upregulates proteins designed
to promote growth and survival of neurons111.

Modulation of metal–protein interactions. The modest


cognitive test scores. A further trial to determine whether effects attained with antioxidants and therapies targeting
vitamin E can delay the progression of mild cognitive downstream effects of ROS generation, such as those
impairment in AD is presently ongoing104. targeting excitotoxicity, are not too surprising when one
One consequence of ROS generation is the initiation considers that they do not target the underlying cause of
of excitotoxicity; this is modulated through the over- the oxidative stress, but rather the consequences. A treat-
activation of glutamate receptors, and a number of ment strategy for oxidative stress is likely to be more
drugs that target these receptors have shown some effi- effective if it targets the origin of ROS generation. A
cacy in treating neurodegenerative diseases. Memantine, breakdown in metal homeostasis, corruption of metallo-
which targets the N-methyl-D-aspartate (NMDA) enzymes or a breakdown in endogenous antioxidant
receptor, has received FDA approval for use in AD. defences are the underlying causes of oxidative stress.
Memantine slows the development of the disease and is A pharmacological therapy targeting these interac-
of modest benefit to patients in the moderately severe to tions requires the identification of molecules that will
inhibit the deleterious effects of aberrant metal interac-
tions. One approach is to design molecules that com-
Box 3 | The blood–brain barrier pete with the target protein for the metal ions, that is, a
For a potential therapy targeting a neurological condition, the blood–brain barrier metal–protein attenuating compound (MPAC). This
(BBB) is a major hurdle that must be overcome. The BBB represents the junction at strategy should not be confused with the concept of
which the bloodstream meets the neuronal environment and is relatively impermeable ‘chelation therapy’, a term associated with the removal
to fluctuations in blood Cu, Fe and Zn levels (for example, post-prandially). A barrier of bulk metals, such as in Wilson’s disease (Cu) and
consisting of brain microvessel endothelial cells ensures that neurotoxic metabolites are β-thallesemia (Fe). The breakdown in metal homeo-
excluded from the brain, but allows essential nutrients to pass across. The prevailing stasis in these diseases leads to tissue saturation with
opinion is that small (<500 Da), compact hydrophobic molecules have the best chance metal. The mechanism of action of traditional chela-
of crossing the BBB by passive diffusion135. A number of predictors of BBB penetration tors is to sequester peripherally and clear by excretion.
have been developed but the two most routinely used measures for predicting the The intention of the MPAC is to disrupt an abnormal
ability of a compound to penetrate the BBB are octanol partition coefficients and polar metal–protein interaction to achieve a subtle reparti-
surface areas (PSA). Octanol/water partition coefficients (log P), either computationally tioning of metals and a subsequent normalization of
or experimentally determined, are used as a measure of lipophilicity of a compound; metal distribution. Once the toxic cycle is inhibited,
a rough guide suggests that a log P of at least 1.5 is required for adequate penetration of endogenous clearance processes can cope more effec-
the BBB, and the larger log P the greater the chance of a molecule passing the BBB136. tively with the accumulated protein.
There is clearly a practical limit to how hydrophobic a molecule can be before decreases The first attempt to clinically target metals in neuro-
in general solubility make a molecule unsuitable for therapeutic application. PSA is a degenerative diseases was the use of desferrioxamine in
measure of a molecule’s compactness, a smaller PSA being more favourable for passage
AD. In this trial, patients were given intramuscular
across the BBB. Molecules with a PSA of 30–40 Å2 are ideal, whereas molecules with a
injections twice daily, five days per week, for 24 months,
PSA of greater than 70 Å2 do not readily cross the BBB137.
a treatment regime that resulted in a significant reduction

NATURE REVIEWS | DRUG DISCOVERY VOLUME 3 | MARCH 2004 | 2 1 1


REVIEWS

in the rate of decline of daily living skills when com- The success of CQ as a potential approach for AD
pared with control patients112. encouraged the testing of this molecule in animal
In light of the interactions between Aβ and metals, models of PD. The 1-methyl-4-phenyl-1,2,3,6-
we have investigated the potential use of MPACs in treat- tetrapyridine (MPTP) model of PD is a toxicity para-
ing AD. We identified a hydrophobic moderate metal digm, in which PD-like symptoms are induced by a
chelator (Clioquinol (CQ, 5-chloro-7-iodo-8-hydroxy- toxin that targets the dopaminergic pathways (FIG. 3),
quinoline)) that is capable of crossing the BBB as a and cells with high neuromelanin content are more
prototypic MPAC. CQ is able to bind a range of metal susceptible to the toxin115. The toxicity induced by
ions with moderate affinity, that is, in the nM range. MPTP was inhibited by CQ and by overexpression of
CQ was given orally in a blinded study to Tg2576 the iron-binding protein ferritin116. These results are
transgenic mice113. The results showed a 49% decrease in consistent with the notion that 8-hydroxyquinolines
brain Aβ burden compared with non-treated controls are higher-affinity ligands for iron than dopamine71.
after nine weeks, with no evidence of toxicity; in addi- Similar results have been reported in the 6-hydroxy-
tion, the general health and body weight parameters dopamine-induced lesion model of PD that was treated
were more stable in the treated animals. Treatment with with an unspecified lipophilic derivative of 1,2-bis-
CQ did not lead to a systematic decrease in metal levels, (2-amino-phenoxy)ethane-N,N,N’,N’-tetraacetic
which is most probably due to the drug’s moderate acid, although no details of these experiments have
binding affinities. Interestingly, in the same study, TETA, been published117.
a hydrophilic high-affinity metal chelator that is inca-
pable of crossing the BBB and which has been used to Summary
treat Wilson’s disease, did not inhibit Aβ deposition, The brain — an organ that requires high metal ion
indicating that systemic depletion of metal-ions (chela- concentrations to maintain many of its functions —
tion therapy) is unlikely to be an effective therapeutic has a poor capacity to cope with oxidative stress, and
strategy for the treatment of AD. demonstrates little regenerative capacity. Data are
The success of CQ in the transgenic mouse trials emerging that implicate aberrant redox metal interac-
encouraged the testing of this molecule in clinical tions with key proteins in the neurodegenerative dis-
trials. The effects of oral CQ treatment in a random- eases AD, PD and ALS, and the subsequent induction of
ized, double-blind, placebo-controlled pilot Phase II oxidative stress leading to neurodegeneration. Oxidative
clinical trial of moderately severe AD patients were stress manifests its toxic effects through a variety of
evaluated114. The effect of treatment was statistically different pathways, which provide a number of potential
significant in preventing cognitive deterioration during therapeutic strategies. Antioxidants targeting ROS, and
a 36-week period in the more severely affected patients MPACs targeting the upstream metal–protein interac-
(baseline Alzheimer’s Disease Assessment Scale — tions that generate ROS, have shown promise in pre-
cognitive subscale ≥ 25). There was also a significant liminary clinical studies. Drugs that target excitotoxicity,
decline in plasma Aβ42 in the CQ group compared with a downstream consequence of oxidative stress, are now
an increase in the placebo group. in clinical use for all three diseases.

1. Selley, M. L., Close, D. R. & Stern, S. E. The effect of 10. Pappolla, M. A., Omar, R. A., Kim, K. S. & Robakis, N. K. 19. Mattson, M. P. & Chan, S. L. Neuronal and glial calcium
increased concentrations of homocysteine on the Immunohistochemical evidence of oxidative stress in signaling in Alzheimer’s disease. Cell Calcium 34, 385–397
concentration of (E)-4-hydroxy-2-nonenal in the plasma Alzheimer’s disease. Am. J. Pathol. 140, 621–628 (1992). (2003).
and cerebrospinal fluid of patients with Alzheimer’s 11. Ferrari, C. K. B. Free radicals, lipid peroxidation and 20. Yamamoto, K. et al. The hydroxyl radical scavenger
disease. Neurobiol. Aging 23, 383–388 (2002). antioxidents in apoptosis: implications in cancer, Nicaraven inhibits glutamate release after spinal injury in
2. Butterfield, D. A., Castegna, A., Lauderback, C. M. & Drake, J. cardiovascular and neurological diseases. Biologia 55, rats. Neuroreport 9, 1655–1659 (1998).
Evidence that amyloid β-peptide-induced lipid peroxidation 581–590 (2000). 21. Mattson, M. P. Excitotoxic and excitoprotective
and its sequelae in Alzheimer’s disease brain contribute to 12. Lovell, M. A., Xie, C. & Markesbery, W. R. Acrolein, a product mechanisms: abundant targets for the prevention and
neuronal death. Neurobiol. Aging 23, 655–664 (2002). of lipid peroxidation, inhibits glucose and glutamate uptake treatment of neurodegenerative disorders. Neuromolecular
3. Dexter, D. T. et al. Basal lipid peroxidation in substantia in primary neuronal cultures. Free Radic. Biol. Med. 29, Med. 3, 65–94 (2003).
nigra is increased in Parkinson’s disease. J. Neurochem. 714–720 (2000). 22. Lewen, A., Matz, P. & Chan, P. H. Free radical pathways in
52, 381–389 (1989). 13. Keller, J. N. et al. Impairment of glucose and glutamate CNS injury. J. Neurotrauma 17, 871–890 (2000).
4. Pedersen, W. A. et al. Protein modification by the lipid transport and induction of mitochondrial oxidative stress 23. Suzuki, Y. J., Forman, H. J. & Sevanian, A. Oxidants as
peroxidation product 4-hydroxynonenal in the spinal cords and dysfunction in synaptosomes by amyloid β-peptide: stimulators of signal transduction. Free Radic. Biol. Med.
of amyotrophic lateral sclerosis patients. Ann. Neurol. 44, role of the lipid peroxidation product 4-hydroxynonenal. 22, 269–285 (1997).
819–824 (1998). J. Neurochem. 69, 273–284 (1997). A seminal work describing a potential physiological
5. Arlt, S., Beisiegel, U. & Kontush, A. Lipid peroxidation in 14. Mark, R. J., Hensley, K., Butterfield, D. A. & Mattson, M. P. role for ROS.
neurodegeneration: new insights into Alzheimer’s disease. Amyloid β-peptide impairs ion-motive ATPase activities: 24. Neill, S., Desikan, R. & Hancock, J. Hydrogen peroxide
Curr. Opin. Lipidol. 13, 289–294 (2002). evidence for a role in loss of neuronal Ca2+ homeostasis and signalling. Curr. Opin. Plant Biol. 5, 388–395 (2002).
6. Sayre, L. M., Smith, M. A. & Perry, G. Chemistry and cell death. J. Neurosci. 15, 6239–6249 (1995). 25. Halliwell, B. & Gutteridge, J. Free Radicals in Biology and
biochemistry of oxidative stress in neurodegenerative 15. Tamagno, E. et al. H2O2 and 4-hydroxynonenal mediate Medicine (Oxford Univ. Press, Oxford, 1999).
disease. Curr. Med. Chem. 8, 721–738 (2001). amyloid β-induced neuronal apoptosis by activating JNKs The definitive textbook on ROS.
7. Gabbita, S. P., Lovell, M. A. & Markesbery, W. R. Increased and p38MAPK. Exp. Neurol. 180, 144–155 (2003). 26. Morita, A., Kimura, M. & Itokawa, Y. The effect of aging on
nuclear DNA oxidation in the brain in Alzheimer’s disease. 16. Ermak, G. & Davies, K. J. Calcium and oxidative stress: the mineral status of female mice. Biol. Trace Elem. Res. 42,
J. Neurochem. 71, 2034–2040 (1998). from cell signaling to cell death. Mol. Immunol. 38, 713–721 165–177 (1994).
8. Alam, Z. I. et al. Oxidative DNA damage in the parkinsonian (2002). 27. Takahashi, S. et al. Age-related changes in the
brain: an apparent selective increase in 8-hydroxyguanine 17. LaFerla, F. M. Calcium dyshomeostasis and intracellular concentrations of major and trace elements in the brain of
levels in substantia nigra. J. Neurochem. 69, 1196–1203 signalling in Alzheimer’s disease. Nature Rev. Neurosci. 3, rats and mice. Biol. Trace Elem. Res. 80, 145–158 (2001).
(1997). 862–872 (2002). 28. Maynard, C. J. et al. Overexpression of Alzheimer’s disease
9. Zemlan, F. P., Thienhaus, O. J. & Bosmann, H. B. 18. Gibson, G. E. Interactions of oxidative stress with cellular amyloid-β opposes the age-dependent elevations of brain
Superoxide dismutase activity in Alzheimer’s disease: calcium dynamics and glucose metabolism in copper and iron. J. Biol. Chem. 277, 44670–4476 (2002).
possible mechanism for paired helical filament formation. Alzheimer’s disease. Free Radic. Biol. Med. 32, References 26–28 describe the rise in metal
Brain Res. 476, 160–162 (1989). 1061–1070 (2002). concentration within the brain as a consequence of

212 | MARCH 2004 | VOLUME 3 www.nature.com/reviews/drugdisc


REVIEWS

normal aging, age being the major risk factor for the 53. Huang, X. et al. The Aβ peptide of Alzheimer’s disease 76. Faucheux, B. A. et al. Neuromelanin associated redox-active
neurodegenerative diseases. directly produces hydrogen peroxide through metal ion iron is increased in the substantia nigra of patients with
29. Bush, A. I. Metals and neuroscience. Curr. Opin. Chem. Biol. reduction. Biochemistry 38, 7609–7616 (1999). Parkinson’s disease. J. Neurochem. 86, 1142–1148 (2003).
4, 184–191 (2000). References 52 and 53 demonstrated that Aβ will 77. Lotharius, J. et al. Effect of mutant α-synuclein on
30. Bush, A. I. The metallobiology of Alzheimer’s disease. generate H2O2 when Aβ coordinates copper. dopamine homeostasis in a new human mesencephalic cell
Trends Neurosci. 26, 207–214 (2003). 54. Cuajungco, M. P. et al. Evidence that the β-amyloid plaques line. J. Biol. Chem. 277, 38884–38894 (2002).
31. Hardy, J. Amyloid, the presenilins and Alzheimer’s disease. of Alzheimer’s disease represent the redox-silencing and 78. Lotharius, J. & Brundin, P. Impaired dopamine storage
Trends Neurosci. 20, 154–159 (1997). entombment of Aβ by zinc. J. Biol. Chem. 275, resulting from α-synuclein mutations may contribute to the
32. Price, D. L., Tanzi, R. E., Borchelt, D. R. & Sisodia, S. S. 19439–19442 (2000). pathogenesis of Parkinson’s disease. Hum. Mol. Genet. 11,
Alzheimer’s disease: genetic studies and transgenic models. 55. Curtain, C. C. et al. Alzheimer’s disease amyloid-β binds 2395–2407 (2002).
Annu. Rev. Genet. 32, 461–493 (1998). copper and zinc to generate an allosterically ordered 79. Baptista, M. J. et al. Co-ordinate transcriptional regulation of
33. Masters, C. L. et al. Amyloid plaque core protein in membrane-penetrating structure containing superoxide dopamine synthesis genes by α-synuclein in human
Alzheimer disease and Down syndrome. Proc. Natl Acad. dismutase-like subunits. J. Biol. Chem. 276, 20466–20473 neuroblastoma cell lines. J. Neurochem. 85, 957–968 (2003).
Sci. USA 82, 4245–4249 (1985). (2001). 80. Wersinger, C., Prou, D., Vernier, P. & Sidhu, A. Modulation of
34. Glenner, G. G. & Wong, C. W. Alzheimer’s disease: initial This paper shows that metals can affect Aβ dopamine transporter function by α-synuclein is altered by
report of the purification and characterization of a novel membrane-bound structures with a metal impairment of cell adhesion and by induction of oxidative
cerebrovascular amyloid protein. Biochem. Biophys. Res. coordination site similar to the active site of SOD. stress. FASEB. J. (2003).
Commun. 120, 885–890 (1984). 56. Curtain, C. C. et al. Metal ions, pH, and cholesterol regulate 81. Perez, R. G. et al. A role for α-synuclein in the regulation of
References 33 and 34 represent the first characteri- the interactions of Alzheimer’s disease amyloid-β peptide dopamine biosynthesis. J. Neurosci. 22, 3090–3099 (2002).
zation of the deposited amyloid within the AD brain. with membrane lipid. J. Biol. Chem. 278, 2977–2982 (2003). 82. Braak, E. et al. α-synuclein immunopositive Parkinson’s
35. Kang, J. et al. The precursor of Alzheimer’s disease amyloid 57. Atwood, C. S. et al. Characterization of copper interactions disease-related inclusion bodies in lower brain stem nuclei.
A4 protein resembles a cell-surface receptor. Nature 325, with alzheimer amyloid β peptides: identification of an Acta Neuropathol. (Berl) 101, 195–201 (2001).
733–736 (1987). attomolar-affinity copper binding site on amyloid β 1–42. 83. Fasano, M., Giraudo, S., Coha, S., Bergamasco, B. &
36. Lee, J. Y., Cole, T. B., Palmiter, R. D., Suh, S. W. & Koh, J. Y. J. Neurochem. 75, 1219–1233 (2000). Lopiano, L. Residual substantia nigra neuromelanin in
Contribution by synaptic zinc to the gender-disparate 58. Barnham, K. J. et al. Neurotoxic, redox-competent Parkinson’s disease is cross-linked to α-synuclein.
plaque formation in human Swedish mutant APP transgenic Alzheimer’s β-amyloid is released from lipid membrane by Neurochem. Int. 42, 603–606 (2003).
mice. Proc. Natl Acad. Sci. USA 99, 7705–7710 (2002). methionine oxidation. J. Biol. Chem. 278, 42959–42965 84. Xu, J. et al. Dopamine-dependent neurotoxicity of α-
This paper highlights the pivotal role zinc plays in the (2003). synuclein: a mechanism for selective neurodegeneration in
deposition of Aβ. A demonstration that a non-fibrillar, highly soluble Parkinson disease. Nature Med. 8, 600–606 (2002).
37. Lovell, M. A., Robertson, J. D., Teesdale, W. J., form of Aβ is still neurotoxic. References 77–84 show links between α-synuclein,
Campbell, J. L. & Markesbery, W. R. Copper, iron and zinc 59. Atwood, C. S. et al. Copper catalyzed oxidation of Alzheimer which genetics demonstrates is important in the
in Alzheimer’s disease senile plaques. J. Neurol. Sci. 158, Aβ. Cell. Mol. Biol. (Noisy-Le-Grand) 46, 777–783 (2000). progression of PD, and dopamine, which is required
47–52 (1998). 60. Head, E. et al. Oxidation of Aβ and plaque biogenesis in for the redox chemistry that is responsible for the
This paper illustrates that plaques associated with AD Alzheimer’s disease and Down syndrome. Neurobiol. Dis. oxidative stress.
are metal sinks. 8, 792–806 (2001). 85. Paik, S. R., Shin, H. J., Lee, J. H., Chang, C. S. & Kim, J.
38. Smith, M. A., Harris, P. L., Sayre, L. M. & Perry, G. Iron 61. McLean, C. A. et al. Soluble pool of Aβ amyloid as a Copper(II)-induced self-oligomerization of α-synuclein.
accumulation in Alzheimer disease is a source of redox- determinant of severity of neurodegeneration in Alzheimer’s Biochem. J. 340, 821–828 (1999).
generated free radicals. Proc. Natl Acad. Sci. USA 94, disease. Ann. Neurol. 46, 860–866 (1999). 86. Uversky, V. N., Li, J. & Fink, A. L. Metal-triggered structural
9866–9868 (1997). A demonstration that AD progression correlates transformations, aggregation, and fibrillation of human
39. Huang, X. et al. Zinc-induced Alzheimer’s Aβ 1–40 with soluble Aβ, not the deposited material found α-synuclein. A possible molecular NK between Parkinson’s
aggregation is mediated by conformational factors. J. Biol. within plaques. disease and heavy metal exposure. J. Biol. Chem. 276,
Chem. 272, 26464–26470 (1997). 62. Cherny, R. A. et al. Aqueous dissolution of Alzheimer’s disease 44284–44296 (2001).
40. Atwood, C. S. et al. Dramatic aggregation of Alzheimer Aβ Aβ amyloid deposits by biometal depletion. J. Biol. Chem. 87. Lee, E. N., Lee, S. Y., Lee, D., Kim, J. & Paik, S. R. Lipid
by Cu(II) is induced by conditions representing physiological 274, 23223–23228 (1999). interaction of α-synuclein during the metal-catalyzed
acidosis. J. Biol. Chem. 273, 12817–12826 (1998). 63. Lau, T. L., Barnham, K. J., Curtain, C. C., Masters, C. L. & oxidation in the presence of Cu2+ and H2O2. J. Neurochem.
41. Terry, R. D. The pathogenesis of Alzheimer disease: an Separovic, F. Magnetic resonance studies of β-amyloid 84, 1128–1142 (2003).
alternative to the amyloid hypothesis. J. Neuropathol. Exp. peptides. Aust. J. Chem 56, 349–356 (2003). 88. Yamin, G., Glaser, C. B., Uversky, V. N. & Fink, A. L. Certain
Neurol. 55, 1023–1025 (1996). 64. Atwood, C. S. et al. Amyloid-β: a chameleon walking in two metals trigger fibrillation of methionine-oxidized α-synuclein.
42. Martins, R. N., Harper, C. G., Stokes, G. B. & Masters, C. L. worlds: a review of the trophic and toxic properties of J. Biol. Chem. 278, 27630–27635 (2003).
Increased cerebral glucose-6-phosphate dehydrogenase amyloid-β. Brain Res. Brain Res. Rev. 43, 1–16 (2003). 89. Bruijn, L. I. et al. Aggregation and motor neuron toxicity of
activity in Alzheimer’s disease may reflect oxidative stress. 65. Spillantini, M. G. et al. α-synuclein in Lewy bodies. Nature an ALS-linked SOD1 mutant independent from wild-type
J. Neurochem. 46, 1042–1045 (1986). 388, 839–840 (1997). SOD1. Science 281, 1851–1854 (1998).
43. Sayre, L. M. et al. In situ oxidative catalysis by neurofibrillary Identifies α-synuclein as the major protein deposited 90. Gurney, M. E. et al. Motor neuron degeneration in mice that
tangles and senile plaques in Alzheimer’s disease: a central in the Lewy bodies associated with PD. express a human Cu,Zn superoxide dismutase mutation.
role for bound transition metals. J. Neurochem. 74, 66. Gasser, T. Genetics of Parkinson’s disease. J. Neurol. 248, Science 264, 1772–1775 (1994).
270–279 (2000). 833–840 (2001). This paper reports that ALS mutant SOD is toxic —
44. Atwood, C. S., Huang, X., Moir, R. D., Tanzi, R. E. & 67. Marsden, C. D. Neuromelanin and Parkinson’s disease. that is, there is a toxic gain of function.
Bush, A. I. Role of free radicals and metal ions in the J. Neural. Transm. Suppl. 19, 121–141 (1983). 91. Valentine, J. S. & Hart, P. J. Misfolded CuZnSOD and
pathogenesis of Alzheimer’s disease. Met. Ions Biol. Syst. 68. Hirsch, E., Graybiel, A. M. & Agid, Y. A. Melanized amyotrophic lateral sclerosis. Proc. Natl Acad. Sci. USA
36, 309–364 (1999). dopaminergic neurons are differentially susceptible to 100, 3617–3622 (2003).
45. Basun, H., Forssell, L. G., Wetterberg, L. & Winblad, B. degeneration in Parkinson’s disease. Nature 334, 92. Wiedau-Pazos, M. et al. Altered reactivity of superoxide
Metals and trace elements in plasma and cerebrospinal fluid 345–348 (1988). dismutase in familial amyotrophic lateral sclerosis. Science
in normal aging and Alzheimer’s disease. J. Neural Transm. Identifies the selective destruction of specific cell 271, 515–518 (1996).
Park. Dis. Dement. Sect. 3, 231–258 (1991). types in the progression of PD. 93. Ghadge, G. D. et al. Mutant superoxide dismutase-1-linked
46. Squitti, R. et al. Elevation of serum copper levels in 69. Wakamatsu, K., Fujikawa, K., Zucca, F. A., Zecca, L. & Ito, S. familial amyotrophic lateral sclerosis: molecular mechanisms
Alzheimer’s disease. Neurology 59, 1153–1161 (2002). The structure of neuromelanin as studied by chemical of neuronal death and protection. J. Neurosci. 17,
47. Bishop, G. M. et al. Iron: a pathological mediator of degradative methods. J. Neurochem. 86, 1015–1023 8756–8766 (1997).
Alzheimer disease? Dev. Neurosci. 24, 184–187 (2002). (2003). 94. Hottinger, A. F., Fine, E. G., Gurney, M. E., Zurn, A. D. &
48. Smith, M. A. et al. Cytochemical demonstration of oxidative Identifies the chemical nature of neuromelanin Aebischer, P. The copper chelator D-penicillamine delays
damage in Alzheimer disease by immunochemical thereby providing further evidence as to why onset of disease and extends survival in a transgenic
enhancement of the carbonyl reaction with 2,4- neuromelanin-containing cells are susceptible in PD. mouse model of familial amyotrophic lateral sclerosis.
dinitrophenylhydrazine. J. Histochem. Cytochem. 46, 70. Zecca, L. et al. The neuromelanin of human substantia nigra: Eur. J. Neurosci. 9, 1548–1551 (1997).
731–735 (1998). structure, synthesis and molecular behaviour. J. Neural. 95. Azzouz, M. et al. Prevention of mutant SOD1 motoneuron
49. Dong, J. et al. Metal binding and oxidation of amyloid-β Transm. Suppl. 145–155 (2003). degeneration by copper chelators in vitro. J. Neurobiol. 42,
within isolated senile plaque cores: Raman microscopic 71. Gerard, C., Chehhal, H. & Hugel, R. P. Complexes of 49–55 (2000).
evidence. Biochemistry 42, 2768–2773 (2003). iron(III) with ligands of biological interest: dopamine and References 92–95 show that copper chelators inhibit
50. Behl, C., Davis, J. B., Lesley, R. & Schubert, D. Hydrogen 8-hydroxyquinoline-5-sulfonic acid. Polyhedron 13, the mutant SOD toxicity, implying that copper has a
peroxide mediates amyloid β protein toxicity. Cell 77, 591–597 (1994). pivotal role in the progression of ALS.
817–827 (1994). 72. Double, K. L. et al. Iron-binding characteristics of 96. Estevez, A. G. et al. Induction of nitric oxide-dependent
Initial identification of H2O2 as a key participant in the neuromelanin of the human substantia nigra. Biochem. apoptosis in motor neurons by zinc-deficient superoxide
toxicity of Aβ. Pharmacol. 66, 489–494 (2003). dismutase. Science 286, 2498–500 (1999).
51. Opazo, C. et al. Metalloenzyme-like activity of Alzheimer’s 73. Smythies, J. On the function of neuromelanin. Proc. R. Soc. 97. Subramaniam, J. R. et al. Mutant SOD1 causes motor
disease β-amyloid. Cu-dependent catalytic conversion of Lond. B Biol. Sci. 263, 487–489 (1996). neuron disease independent of copper chaperone-mediated
dopamine, cholesterol, and biological reducing agents to 74. Sulzer, D. et al. Neuromelanin biosynthesis is driven by copper loading. Nature Neurosci. 5, 301–307 (2002).
neurotoxic H2O2. J. Biol. Chem. 277, 40302–40308 excess cytosolic catecholamines not accumulated by Reports that the copper at the active site of SOD is
(2002). synaptic vesicles. Proc. Natl Acad. Sci. USA 97, not responsible for the observed toxicity.
52. Huang, X. et al. Cu(II) potentiation of alzheimer Aβ 11869–11874 (2000). 98. Goto, J. J. et al. Loss of in vitro metal ion binding specificity
neurotoxicity. Correlation with cell-free hydrogen peroxide 75. Ben-Shachar, D., Riederer, P. & Youdim, M. B. Iron-melanin in mutant copper-zinc superoxide dismutases associated
production and metal reduction. J. Biol. Chem. 274, interaction and lipid peroxidation: implications for with familial amyotrophic lateral sclerosis. J. Biol. Chem.
37111–37116 (1999). Parkinson’s disease. J. Neurochem. 57, 1609–1614 (1991). 275, 1007–1014 (2000).

NATURE REVIEWS | DRUG DISCOVERY VOLUME 3 | MARCH 2004 | 2 1 3


REVIEWS

99. Liu, H. et al. Copper2+ binding to the surface residue cysteine Proof-of-concept Phase II clinical trial that 130. Bayer, T. A. et al. Dietary Cu stabilizes brain superoxide
111 of His46Arg human copper-zinc superoxide dismutase, demonstrates that inhibiting Aβ metal interactions has dismutase 1 activity and reduces amyloid Aβ production in
a familial amyotrophic lateral sclerosis mutant. Biochemistry positive clinical results. APP23 transgenic mice. Proc. Natl Acad. Sci. USA 100,
39, 8125–8132 (2000). 115. Dauer, W. & Przedborski, S. Parkinson’s disease: 14187–14192 (2003).
Identification of a low-affinity, surface-located mechanisms and models. Neuron 39, 889–909 (2003). 131. Phinney, A. L. et al. In vivo reduction of amyloid-β by a
copper-binding site on SOD that is redox active and 116. Kaur, D. et al. Genetic or pharmacological iron chelation mutant copper transporter. Proc. Natl Acad. Sci. USA
therefore could play some role in mediating redox- prevents MPTP-induced neurotoxicity in vivo: a novel 100, 14193–14198 (2003).
associated oxidative stress. therapy for Parkinson’s disease. Neuron 37, 899–909 132. Borchardt, T. et al. Copper inhibits β-amyloid production
100. Bush, A. I. Is ALS caused by an altered oxidative activity of (2003). and stimulates the non-amyloidogenic pathway of
mutant superoxide dismutase? Nature Neurosci. 5, 919; A demonstration that targeting metals in a mouse amyloid-precursor-protein secretion. Biochem. J. 344 Pt
author reply 919–920 (2002). model of PD has positive effects. 2, 461–467 (1999).
101. Moosmann, M. & Behl, C. Antioxidants as treatment for 117. Angel, I. et al. Metal ion chelation in neurodegenerative References 130–132 Illustrate how copper can
neurodegenerative disorders. Exp. Opin. Invest. Drugs 11, disorders. Drug Dev. Res. 56, 300–309 (2002). mediate APP processing to reduce Aβ production.
1407–1435 (2002). 118. Campuzano, V. et al. Friedreich’s ataxia: autosomal 133. Barnham, K. J. et al. Structure of the Alzheimer’s disease
102. Effects of tocopherol and deprenyl on the progression of recessive disease caused by an intronic GAA triplet repeat amyloid precursor protein copper binding domain.
disability in early Parkinson’s disease. The Parkinson Study expansion. Science 271, 1423–1427 (1996). A regulator of neuronal copper homeostasis. J. Biol.
Group. N. Engl. J. Med. 328, 176–183 (1993). 119. Bradley, J. L. et al. Clinical, biochemical and molecular Chem. 278, 17401–17407 (2003).
103. Sano, M. et al. A controlled trial of selegiline, α-tocopherol, genetic correlations in Friedreich’s ataxia. Hum. Mol. Genet. 3D Structure of the copper–binding domain of APP
or both as treatment for Alzheimer’s disease. The 9, 275–282 (2000). showing where and how APP coordinates copper.
Alzheimer’s Disease Cooperative Study. N. Engl. J. Med. 120. Rustin, P. et al. Effect of idebenone on cardiomyopathy in 134. Rogers, J. T. et al. An iron-responsive element type II in
336, 1216–1222 (1997). Friedreich’s ataxia: a preliminary study. Lancet 354, the 5’-untranslated region of the Alzheimer’s amyloid
This study showed that the antioxident vitamin E has 477–479 (1999).
precursor protein transcript. J. Biol. Chem. 277,
protective effects in an AD clinical trial. 121. Hausse, A. O. et al. Idebenone and reduced cardiac
45518–45528 (2002).
104. Grundman, M. Vitamin E and Alzheimer’s disease: the basis hypertrophy in Friedreich’s ataxia. Heart 87, 346–349
135. Levin, V. A. Relationship of octanol/water partition
for additional clinical trials. Am. J. Clin. Nutr. 71, (2002).
coefficient and molecular weight to rat brain capillary
S630–S636 (2000). 122. Buyse, G. et al. Idebenone treatment in Friedreich’s ataxia:
permeability. J. Med. Chem. 23, 682–684 (1980).
105. Reisberg, B. et al. Memantine in moderate-to-severe Alz- neurological, cardiac, and biochemical monitoring.
136. Habgood, M. D. et al. Investigation into the correlation
heimer’s disease. N. Engl. J. Med. 348, 1333–1341 (2003). Neurology 60, 1679–1681 (2003).
between the structure of hydroxypyridinones and
106. Mobius, H. J. Memantine: update on the current evidence. 123. Wong, A. et al. The Friedreich’s ataxia mutation confers
blood–brain barrier permeability. Biochem. Pharmacol.
Int. J. Geriatr. Psychiatry 18, S47–S54 (2003). cellular sensitivity to oxidant stress which is rescued by
57, 1305–1310 (1999).
107. Winblad, B. & Jelic, V. Treating the full spectrum of dementia chelators of iron and calcium and inhibitors of apoptosis.
Hum. Mol. Genet. 8, 425–430 (1999). 137. Iyer, M., Mishru, R., Han, Y. & Hopfinger, A. J. Predicting
with memantine. Int. J. Geriatr. Psychiatry 18, S41–S46 (2003).
124. Richardson, D. R., Mouralian, C., Ponka, P. & Becker, E. blood–brain barrier partitioning of organic molecules using
108. Hely, M. A., Fung, V. S. & Morris, J. G. Treatment of
Parkinson’s disease. J. Clin. Neurosci. 7, 484–494 (2000). Development of potential iron chelators for the treatment membrane–interaction QSAR analysis. Pharm Res 19,
109. Bensimon, G., Lacomblez, L. & Meininger, V. A controlled of Friedreich’s ataxia: ligands that mobilize mitochondrial 1611–1621 (2002).
trial of riluzole in amyotrophic lateral sclerosis. ALS/Riluzole iron. Biochim. Biophys. Acta 1536, 133–40 (2001).
Study Group. N. Engl. J. Med. 330, 585–591 (1994). 125. Culotta, V. C. et al. The copper chaperone for superoxide Competing interests statement
110. Doraiswamy, P. M. Non-cholinergic strategies for treating dismutase. J. Biol. Chem. 272, 23469–23472 (1997). The authors declare competing financial interests: see Web version
and preventing Alzheimer’s disease. CNS Drugs 16, Identification of the protein responsible for loading for details.
811–824 (2002). copper into the active site of SOD.
111. Mattson, M. P. Will caloric restriction and folate protect 126. Waggoner, D. J., Bartnikas, T. B. & Gitlin, J. D. The role of
against AD and PD? Neurology 60, 690–695 (2003). copper in neurodegenerative disease. Neurobiol. Dis. 6, Online links
112. Crapper McLachlan, D. R. et al. Intramuscular 221–230 (1999).
desferrioxamine in patients with Alzheimer’s disease. 127. Rae, T. D., Schmidt, P. J., Pufahl, R. A., Culotta, V. C. & DATABASES
Lancet 337, 1304–1308 (1991). O’Halloran, T. V. Undetectable intracellular free copper: The following terms in this article are linked online to:
113. Cherny, R. A. et al. Treatment with a copper-zinc chelator the requirement of a copper chaperone for superoxide LocusLink: http://www.ncbi.nlm.nih.gov/LocusLink/
markedly and rapidly inhibits β-amyloid accumulation in dismutase. Science 284, 805–808 (1999). β-actin | catalase | creatine kinase | frataxin | glucose transporter
Alzheimer’s disease transgenic mice. Neuron 30, 665–676 This paper illustrates the elaborate mechanisms that type-3 | glutathione peroxidase | glutathione reductase |
(2001). nature has evolved to tightly regulate metal ions. mitogen-activated protein kinase-1 | SOD | α-synuclein |
Inhibiting metal Aβ interactions have positive 128. Andrews, N. C. Mining copper transport genes. Proc. Natl xanthine dehydrogenase
outcomes in a transgenic mouse model. Acad. Sci. USA 98, 6543–6545 (2001). Online Mendelian Inheritance in Man:
114. Ritchie, C. W. et al. Metal–protein attenuation with 129. White, A. R. et al. Copper levels are increased in the cerebral http://www.ncbi.nlm.nih.gov/Omim/
iodochlorhydroxyquin (clioquinol) targeting Aβ amyloid cortex and liver of APP and APLP2 knockout mice. Brain Alzheimer’s disease | amyotrophic lateral sclerosis |
deposition and toxicity in Alzheimer’s disease: biochemical Res. 842, 439–444 (1999). Friedreich’s ataxia | Huntington’s disease | nitric oxide synthase |
and clinical responses in a pilot phase 2 clinical trial. Arch. A demonstration that APP has a role in copper Parkinson’s disease | stroke | Wilson’s disease
Neurol. 60, 1685–1691 (2003). homeostasis. Access to this interactive links box is free online.

214 | MARCH 2004 | VOLUME 3 www.nature.com/reviews/drugdisc

Você também pode gostar