You are on page 1of 18

Progress in Neurobiology 92 (2010) 1–18

Contents lists available at ScienceDirect

Progress in Neurobiology
journal homepage: www.elsevier.com/locate/pneurobio

Biological metals and Alzheimer’s disease: Implications for therapeutics and


diagnostics
James A. Duce a,b, Ashley I. Bush a,c,*
a
The Mental Health Research Institute, Parkville, Victoria 3052, Australia
b
Centre for Neuroscience, The University of Melbourne, Victoria 3010, Australia
c
The Department of Pathology, The University of Melbourne, Victoria 3010, Australia

A R T I C L E I N F O A B S T R A C T

Article history: The equilibrium of metal ions is critical for many physiological functions, particularly in the central
Received 9 February 2010 nervous system, where metals are essential for development and maintenance of enzymatic activities,
Received in revised form 14 April 2010 mitochondrial function, myelination, neurotransmission as well as learning and memory. Due to their
Accepted 16 April 2010
importance, cells have evolved complex machinery for controlling metal-ion homeostasis. However,
disruption of these mechanisms, or absorption of detrimental metals with no known biological function,
Keywords: alter the ionic balance and can result in a disease state, including several neurodegenerative disorders
Alzheimer’s disease
such as Alzheimer’s disease. Understanding the complex structural and functional interactions of metal
Ageing
Copper
ions with the various intracellular and extracellular components of the central nervous system, under
Zinc normal conditions and during neurodegeneration, is essential for the development of effective therapies.
Iron Accordingly, assisting the balance of metal ions back to homeostatic levels has been proposed as a
Metal-protein attenuating compounds disease-modifying therapeutic strategy for Alzheimer’s disease as well as other neurodegenerative
diseases.
ß 2010 Elsevier Ltd. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .... . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .2
2. Metals. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .... . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .2
2.1. Toxicological metals . . . . . . . . . . . . . . . . . . . . . . . . . . . . .... . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .2
2.1.1. Lead . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .... . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .2
2.1.2. Aluminium . . . . . . . . . . . . . . . . . . . . . . . . . . . . .... . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .2
2.2. Biochemically functional metals (biometals). . . . . . . . . .... . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3
2.2.1. Copper . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .... . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3
2.2.2. Zinc . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .... . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .4
2.2.3. Iron. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .... . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .5
3. Metal involvement with AD related proteins . . . . . . . . . . . . . . .... . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .6
3.1. b-amyloid precursor protein . . . . . . . . . . . . . . . . . . . . . .... . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .6
3.1.1. APP binding to metal. . . . . . . . . . . . . . . . . . . . .... . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .6
3.1.2. Regulation of APP by metal. . . . . . . . . . . . . . . .... . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .6
3.1.3. Metal involvement in secretase processing of APP . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .6

Abbreviations: Ab, b-amyloid; AD, Alzheimer’s disease; ADAM, cell surface metalloproteinase family of disintegrin and metalloprotease; APP, b-amyloid precursor protein;
ApoE, apolipoprotein E; ATP7a, ATPase, Cu2+ transporting, alpha polypeptide (associated with Menkes syndrome); ATP7b, ATPase, Cu2+ transporting, beta polypeptide
(associated with Wilson syndrome); BACE, b-APP-site cleaving enzyme; BBB, blood brain barrier; CCS1, copper chaperone of superoxide dismutase 1; CNS, central nervous
system; CQ, clioquinol (5-chloro-7-iodo-8-hydroxyquinoline).; CSF, cerebral spinal fluid; Ctr1, high affinity copper uptake protein 1; Cu2+, oxidized copper; Cu+, reduced
copper; DFO, desferrioxamine; ERK, extracellular signal-regulated kinase; Fe3+, oxidized iron; Fe2+, reduced iron; GSK-3, glycogen synthase kinase 3; HFE, hemochromatosis
gene; HNE, 4-hydroxy-2-nonenal; H2O2, hydrogen peroxide; IDE, insulin-degrading enzyme; IRE, iron responsive element; JNK, c-Jun NH2-terminal kinase; MMP, matrix
metalloproteinase; MPAC, metal-protein attenuating compound; MT, metallothionein; NEP, neprilysin; NF, neurofilaments; NFT, neurofibrillary tangle; NMDA, N-methyl-D-
aspartic acid; PI3K, phosphoinositol-3-kinase; SOD1, superoxide dismutase 1; TACE, tumor necrosis factor-alpha converting enzyme (also called ADAM-17); ZIP, zinc-
importing protein; ZnT, zinc transporter.
* Corresponding author at: The Mental Health Research Institute, Parkville, Victoria 3052, Australia.
E-mail address: abush@mhri.edu.au (A.I. Bush).

0301-0082/$ – see front matter ß 2010 Elsevier Ltd. All rights reserved.
doi:10.1016/j.pneurobio.2010.04.003
2 J.A. Duce, A.I. Bush / Progress in Neurobiology 92 (2010) 1–18

3.2. b-amyloid . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .7
3.2.1. Ab binding to metal . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .7
3.2.2. Ab production modulated by metal . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .8
3.2.3. Metal modulation of Ab degradation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .8
3.2.4. Metal mediated production of reactive oxygen species. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .8
3.3. Tau. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .9
3.3.1. Tau binding to metals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .9
3.3.2. Metal modulation through kinases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .9
3.4. Apolipoprotein E . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .9
4. Therapeutics and diagnostic targeting of metal ions in AD . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .10
4.1. Antioxidants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .10
4.2. Metal chelators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .10
4.3. Metal complexes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .10
4.4. Metal-protein attenuating compounds (MPACs) or Ionophores . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .11
5. Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .12
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .12
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .12

1. Introduction 2.1.1. Lead


Lead, seems to particularly affect the developing CNS.
Since Alzheimer’s disease (AD) was first described over 100 Consequently, children are at a greater risk than adults in suffering
years ago, it is now thought to affect 2% of the industrial world’s from the neurotoxic effects of lead. One of the basic mechanisms
population and is the third leading cause of death in these proposed for leads neuronal toxicity is its substitution for calcium
countries (Mattson, 2004). As such, AD represents a huge in intra-cellular signal transduction. This substitution can affect
socioeconomic problem that requires better diagnostic tools, calcium influx, which is instrumental in the presynaptic release of
management and effective therapies to ease the burden of this neurotransmitters. As early as the second century B.C., mankind
disease. AD brains contain the characteristic pathological has related the use of lead with behavioural and CNS dysfunction.
markers of senile plaques and neurofibrillary tangles. Extracel- However, with the inclusion of lead in paints and fuel, more
lular senile plaques predominantly contain aggregated b- modern studies have directly associated lead with intellectual
amyloid (Ab), a small cleaved product from the b-amyloid impairment (Canfield et al., 2003).
precursor protein (APP) (Iwatsubo et al., 1994), while neurofi- Less well studied is the involvement of lead in AD. Recent
brillary tangles consist of insoluble intra-neuronal inclusions rodent and primate animal studies, however, have implicated a
formed from hyperphosphorylated tau (Clark et al., 1998). The LEARn (latent early-life associated regulation) alteration causing
major risk factor known for AD is age; 95% of all AD cases have no an elevation of APP and Ab in old age (Basha et al., 2005b; Wu et al.,
clear pattern of inheritance and it is believed that interactions 2008) as well as increased oxidative stress (Wu et al., 2008). This
with both genetic and environmental factors contribute to the elevation is not related to the secretases as neither expression
etiology of AD (Migliore and Coppede, 2009). However, since it profiles of the secretases were altered during the animals life post
was first proposed decades ago (Crapper et al., 1973), there has infant exposure to lead, nor did lead have an effect on the
been an expanding interest in both biochemically functional and proteolytic activity of these enzymes on brain extracts. However,
toxicological metals as contributing to the etiology and this metal was still able to promote the formation of Ab aggregates
pathogenesis of AD. at nanomolar concentrations (Basha et al., 2005a) as well as cause
the accumulation of intracellular Ab and increased dense-core
2. Metals plaques in the primate model (Wu et al., 2008).

Metals are widely distributed in nature and in biological 2.1.2. Aluminium


systems they can be broadly classified as either toxicological or Aluminium is the most widely distributed metal in the
biochemically functional. As with many proteins, the mis- environment and is extensively used in daily life, thus providing
metabolism of any metal ion within the body will result in easy exposure to human beings. The toxicity of aluminium is
obtaining levels outside the normal physiological range and can directly linked to its bioavailability. In biological fluids, this
result in biological damage (Frausto da Silva and Williams, 2001). It trivalent cation is rarely present as an ion because it complexes
is also evident that metals play a crucial role in functioning of the extensively with biologically available ligands such as phosphate,
central nervous system (CNS). hydroxide and citrate. Because of its high affinity, aluminium was
used as a phosphate-binding gel in patients with chronic renal
2.1. Toxicological metals failure to control hyperphosphotemia. It was in this population of
individuals that signs of aluminium neurotoxicity were first
Toxicological metals, such as mercury, cadmium, lead and recorded (Alfrey et al., 1976).
aluminium, have no known normal biological function and are One of the first reports linking aluminium to AD demonstrated
detrimental to any organism when absorbed. Interestingly, the that there is an elevated level of the metal in necropsy and biopsy
brain appears to be the primary target organ for such metals. While brain samples from histopathologically confirmed AD patients
biochemically functional metals are tightly homeostatically (Crapper et al., 1973). However, the linkage between aluminium
regulated, non-essential toxicological metals may be able to freely and AD has remained controversial since, as a relationship
gain access to an organism by virtue of the fact that they share between increased aluminium exposure and AD has been found
properties such as ionic charge or other characteristics with their in some studies (Xu et al., 1992) but not confirmed in others
biochemically functional counterparts. (McDermott et al., 1979). Further suggestion that aluminium may
J.A. Duce, A.I. Bush / Progress in Neurobiology 92 (2010) 1–18 3

play a role in the etiology of AD comes from it’s ability to increase; superoxide dismutase 1, SOD1) (Tapiero et al., 2003). Significant
oxidative stress that is induced by iron and other metals (Xie et al., phenotypes are known to result from a diminution in these
1996), Ab production (Clauberg and Joshi, 1993), plaque deposi- enzymatic activities.
tion (Kawahara et al., 1994), b-pleated sheet formation (Vyas and Recently the presence of the ionic species has been reported in
Duffy, 1995), as well as polymerization and aggregation of Ab the synapse at micromolar concentrations. In this environment
protein (Paik et al., 1997). But again controversy lies with many of postsynaptic N-methyl-D-aspartic acid (NMDA) neurites release
these studies as aluminium concentrations used are above those ionic copper upon NMDA activation (Schlief et al., 2005). Activation
that are physiologically relevant (for fuller review see (Gupta et al., of synaptic NMDA receptors in the hippocampal neurons results in
2005; Kawahara, 2005)). trafficking of ATP7a and an associated efflux of free copper (15 mM)
into the synaptic cleft (Fig. 1A) (Schlief et al., 2005). At present, this
2.2. Biochemically functional metals (biometals) is the only known microenvironment where free copper is known.
Catalytic amounts of copper can function as electron acceptors
The biochemically functional metal ion content of the brain is promoting the reaction of nitric oxide with thiols, and potentially
stringently regulated and there is no passive flux of metals from the release of this metal could function as a molecular switch to
the circulation to the brain: movement of metals across the blood control extracellular S-nitrosylaton of the NMDA receptor, a post-
brain barrier (BBB) is highly regulated. While iron, copper and zinc translational mechanism shown to be crucial for modulating
are being increasingly implicated in interactions with the major receptor function (Schlief et al., 2005). Copper has also been
protein components of neurodegenerative disease, this is not reported by the same group to specifically protect against NMDA-
merely due to increased (e.g., toxicological) exposure to these mediated excitotoxic cell death in primary hippocampal neurons,
metals, but rather due to a breakdown in the homeostatic and that this protective effect of copper depends on endogenous
mechanisms that compartmentalize and regulate metals. nitric oxide production in hippocampal neurons (Schlief et al.,
Often termed as trace metals, these ions are neither trace in 2006). ATP7a expression is developmentally regulated, peaking
concentration or irrelevant to functional biology. One estimate is during synaptogenesis, and plays a role in facilitating copper
that 30% of enzymes use metals, however this percentage may be transport into the brain through the endothelial cells that comprise
greater as this estimate has not been evaluated systematically the BBB (El Meskini et al., 2007; Niciu et al., 2007; Qian et al., 1998).
since the publishing of complete genomes (Guengerich, 2009). While the copper uptake protein, Ctr1, is present in neurons, the
These ions participate in many essential activities and deficiencies specific mechanism involved in reuptake of copper released into
can be lethal. In general, these ions are bound to ligands (e.g. the synapses is still not yet known but is likely to be energy-
transferrin or ceruloplasmin) and were for some time thought not dependent (Giese et al., 2005). This is of relevance as mitochondrial
to be found as free species (Rae et al., 1999). However with the energy failure is a feature of ageing and AD and may lead to a
advent of modern cellular and molecular techniques, ‘‘free’’ ionic pooling to synaptic free copper.
metals are now recognized in specific cellular compartments.
While it is unlikely that any ionic metal is truly free, this term is 2.2.1.1. Changes in copper homeostasis with ageing. As mentioned,
used in this manuscript to explain the pool of labile, exchangeable the dominant risk factor associated with all neurodegenerative
or loosely ligated metals that are accessible by chelator binding. diseases is increasing age. Levels of copper have been reported to
Here, we will focus on the three major biological transition change with age in plasma, cerebral spinal fluid (CSF) and brain
metals; copper, zinc and iron that make up a substantial tissue. Levels of plasma copper are low at birth and increase
endogenous pool of transition metals within the brain. Both free sharply after postnatal day 10, concomitantly with the predomi-
zinc and copper have been reported to act as secondary nant copper storage protein ceruloplasmin (Milne and Johnson,
messengers within the synaptic cleft and the intracellular pool 1993). In subjects greater than 75 years of age copper has
of labile iron has been well established to modulate the expression repeatedly been reported as elevated in serum (Madaric et al.,
of various proteins. 1994) and plasma (Ekmekcioglu, 2001; Iskra et al., 1993; McMaster
et al., 1992; Menditto et al., 1993; Milne and Johnson, 1993).
2.2.1. Copper Conversely, several animal and human studies have reported a
Copper is a redox-active metal that is predominantly used by rise in levels of brain copper from youth to adulthood. Analyses of
organisms living in oxygen-rich environments and exists either in normal mice (BL6/SJL) have demonstrated a 46% increase in copper
the oxidized (Cu2+) and reduced (Cu+) valence states (Ridge et al., content (Maynard et al., 2006, 2002). Interestingly, once in
2008). With these changes in redox state, copper can coordinate adulthood a marked drop in levels from middle age onwards
with a range of ligands that include carboxylate oxygen, imidazole occurs. However it is important to note that these gross changes in
nitrogen, cysteine thiolate, and methionine thioether groups and copper may be regional specific as electron paramagnetic imaging
engages in cation-interactions (Bertini, 2007; Franz, 2008). Copper has demonstrated an age-related increase in clusters of copper
may also bind to amides from the backbone (e.g. prion) (Viles et al., within the brain (Wender et al., 1992).
1999) and side chains (e.g. glutamine in stellacyanin) (Hart et al.,
1996) of proteins. Many enzymes harness the changes in bound 2.2.1.2. Changes in copper homeostasis in Alzheimer’s disease. Ho-
copper oxidation state, in the presence of oxygen, to catalyze redox meostatic alterations in brain copper levels have been implicated
chemistry for a wide range of chemical transformations that are in the pathogenesis of several neurological disorders including
central to biology. Because the free forms of the metal are Alzheimer’s, Parkinson’s and prion diseases (Brown and Kozlowski,
potentially damaging, absorption, distribution and excretion are 2004; Mercer, 2001; Torsdottir et al., 1999). The serum and CSF
tightly controlled, and orchestrated by a variety of proteins. In the levels of copper are significantly higher in patients with AD
plasma, the major copper-binding proteins are ceruloplasmin, compared to age match controls (Basun et al., 1991; Squitti et al.,
albumin, and transcuprein, which transport the absorbed copper to 2002a), and potentially correlates with ceruloplasmin protein
all tissues. Transport across cellular membranes is predominantly expression even though the excess pool of the metal is not bound
through the copper transporters Ctr1 for import and ATP7a to this carrier protein (Squitti et al., 2005), however these findings
(Menkes) or ATP7b (Wilson) for export. Once inside cells, cytosolic await replication. This increased serum copper was reported to
copper chaperones deliver the metal to specific molecules (e.g. correlate well with higher levels of serum peroxides in AD patients
copper chaperone of superoxide dismutase 1, CCS1, delivers to (Squitti et al., 2002b). Incidentally, copper mediates Low-density
[(Fig._1)TD$IG]
4 J.A. Duce, A.I. Bush / Progress in Neurobiology 92 (2010) 1–18

lipoprotein (LDL) oxidation by homocysteine (Nakano et al., 2004)


and plasma homocysteine levels are another risk factor for AD
(Seshadri et al., 2002).
Within the brain, copper has been shown to be associated with
senile plaques, significantly increasing copper content from 79 mM
(in the normal age-matched neuropil) to 390 mM within these
plaques (Lovell et al., 1998). Lovell and colleagues also went on to
report an increase in parenchymal copper levels (Lovell et al.,
1998), however their analysis was confined to the amygdala and
contradicts a larger collection of studies reporting a decrease in
bulk tissue copper levels in AD-affected cortical regions (Deibel
et al., 1996; Loeffler et al., 1996). Considering the increased copper
levels in senile plaques with the overall reduction in copper levels,
a complex picture emerges where copper seems to be abnormally
redistributed in AD and collects outside the cell, leaving the tissue
relatively deficient.
Interestingly, this abnormal redistribution of copper in the
brain may have a lipid component, as rabbits maintained on an
elevated cholesterol and copper diet demonstrated accelerated
plaque formation through aggregation of Ab and promotion of
oxidative events in the brain (Sparks and Schreurs, 2003). This is
supported by the observation that copper in combination with a
high fat diet increases the risk for AD (Morris et al., 2006).

2.2.2. Zinc
Zinc is another transition metal abundantly present in all
animal tissue. Diverse classes of proteins require bound zinc for
normal function, these include; zinc metalloenzymes (e.g. SOD1),
transcription factors containing zinc-binding motifs such as zinc
fingers (e.g. p53 and GAL4) (Vallee et al., 1991), signaling (e.g.
protein kinase C) (Szallasi et al., 1996) and storage to buffer
cytosolic zinc pools (e.g. metallothioneins) (Ebadi et al., 1994). Of
all organs, the brain probably contains the highest levels of zinc,
with a possible exception of pancreatic b islets (Frederickson,
1989). In its free ionic form zinc in the brain is highly enriched in
many of the glutamatergic nerve terminals (10–15%), where it is
released upon neuronal activation (Fig. 1A). Recent studies have
investigated in depth the mechanism involved in maintaining
intraneuronal zinc concentrations (for review see (Sensi et al.,
2009)). Although cytosolic free zinc is typically in the picomolar
range (Frederickson et al., 2005), this is substantially increased
within the synaptic bouton to micromolar concentrations. Upon
neuronal activity or membrane depolarization zinc is released and
the ionic concentration is further amplified to millimolar
concentrations within the synaptic cleft. As with copper (Section
2.2.1), synaptically released zinc interacts with neuronal receptors
such as NMDA (Smart et al., 2004) as well as various neuronal ion
channels and transporters. A multitude of zinc transporters (ZnTs),
Fig. 1. Models of zinc and copper in the glutamatergic synapse in health and
Alzheimer’s disease. (A) The healthy synapse. The vesicular zinc transporter ZnT3
transfers zinc into synaptic vesicles. Upon stimulated release zinc concentrations aggregated as it becomes loaded with zinc. While the soluble Ab monomers are
can achieve 300 mM within the synaptic cleft. Copper is released post-synaptically constitutively degraded, zinc-loaded Ab oligomers are resistant to degradation.
following NMDA-induced activation, which causes the translocation of ATP7a and MT3 is also decreased in AD, so promoting abnormal metal-Ab interaction and
its associated copper-laden vesicles to the synaptic cleft. Copper concentrations sequestered metal ions by Ab allows unopposed glutamate activation of the NMDA
may reach 15 mM in the cleft. Both copper and zinc are able to inhibit the NMDA receptor, which could lead to the increased release of post-synaptic copper. (C)
receptor response, which may feedback to prevent further copper from being Ionophore intervention. The metal-free ligands CQ and PBT2 enter the brain and are
released into the cleft. Upon cleavage of APP, Ab is also constitutively released into attracted to pools of extracellular metals, a feature unique to AD. By binding to
the synaptic cleft upon neuronal activity and may normally act as a ‘zinc sponge’ to copper and zinc and removing these metals from Ab, CQ and PBT2 facilitate the
lower synaptic levels. Ab would typically be cleared by movement into the dissolution of Ab to monomers. The dissociated metal ions may be in a ternary
periphery or degradation by extracellular proteases such as neprilysin and insulin complex with the drug itself or in a complex with dissociated Ab. These complexes
degrading enzyme (IDE). Despite high peak concentrations upon neuronal are taken up by neighbouring cells, where the elements are separated. While Ab
stimulation, the average concentrations of free synaptic copper and zinc is kept may be degraded by intracellular proteolysis and CQ/PBT2 is returned to the
low over time by a variety of other ways including putative energy-dependent extracellular space, the metal ions (Cu or Zn) activate PI3K and sets up a series of
reuptake mechanisms as well as buffering by metallothioneins (e.g. MT3) from kinase-mediated events. This leads to the phosphorylation and inhibition of GSK-3
neighbouring astrocytes. (B) Alzheimer’s disease. Decreased mitochondrial energy resulting in decreased tau phosphorylation as well as the activation of MMP2 and 3,
leads to reduced metal reuptake, which causes the average concentration of metals which in turn facilitates the extracellular break down and clearance of Ab. In
to rise over time. This allows copper and zinc to react with Ab released into the addition, the ionophoric properties of CQ/PBT2 lower average synaptic metal
synaptic cleft to form oxidized, cross-linked soluble oligomers and precipitated concentrations and so prevent Ab precipitation, covalent Ab oligomer formation,
amyloid. Ab can bind up to 2.5 moles of metal ions, but becomes more densely and toxic redox activity.
J.A. Duce, A.I. Bush / Progress in Neurobiology 92 (2010) 1–18 5

zinc-importing proteins (ZIPs) and buffering proteins such as the An average adult brain has approximately 60 mg of non-heme
metallothioneins bind cytosolic zinc to prevent free zinc becoming iron, with some cerebral regions such as the substantia nigra,
toxic (Sensi et al., 2009). The zinc transporter ZnT3 has been shown globus palladus, caudate nucleus and putamen retaining the
to be essential for loading zinc into synaptic vesicles (Cole et al., highest levels (Bartzokis et al., 1997; Martin et al., 1998). In the
1999; Linkous et al., 2008) and is likely to represent the major, brain iron is not only important for oxygen transport and electron
possibly sole, synaptic vesicle zinc transporter. ZnT3 is primarily transfer but also neurotransmitter synthesis and myelin produc-
co-localized with zinc in glutamatergic synapses (Palmiter et al., tion (Takeda, 2004). Iron transport into the brain across the BBB is a
1996) present in the hippocampus and neocortex. ZnT3 is likely to complex process still not fully understood, but requires the iron
govern the downstream effects of synaptic zinc on a variety of carrier transferrin. Circulating iron, once oxidized to its ferric state
signaling pathways that mediate higher cognitive function. using a ferroxidase (e.g. ceruloplasmin) cannot simply cross the
BBB (Morris et al., 1992) but requires an iron complexed with
2.2.2.1. Changes in zinc homeostasis with ageing. In contrast to transferrin to bind the transferrin receptor on the lumen side of
copper, zinc levels in rodent and human plasma has been reported endothelial cells lining the cerebrovasculature. The transferrin
to be highest at birth and steadily decrease with age (Bunker et al., complex then enters the cells via endocytosis and is transported
1987; Martinez Lista et al., 1993; Monget et al., 1996; Munro et al., further into the brain to be distributed to various cell types that
1987; Ravaglia et al., 2000). One report on healthy men aged 8–89 require the ion. This process is highly regulated by the abundance
years indicated that plasma zinc levels tended to remain constant or the deficiency in both transferrin (with or without iron
throughout life until the age of 75 where zinc levels suddenly incorporated) and its receptor. Translational regulation of these
decreased (Madaric et al., 1994). proteins as well as other iron binding proteins (e.g. ferritin) is
While these changes have not reflected an age-related controlled by iron regulatory proteins that bind to iron responsive
alternation in the global zinc levels within the brain (Maynard elements (IRE) on RNA to alter the expression levels (Caughman
et al., 2002), specific areas known to have high concentrations of et al., 1988; Kikinis et al., 1995).
zinc such as the hippocampus, are known to decrease with age
(Adlard et al., 2010) In accordance, ZnT3 is also decreased with age 2.2.3.1. Changes in iron homeostasis with ageing. In the healthy
in the same area (Adlard et al., 2010). state, there should never be an over-accumulation of free iron.
Unfortunately, although humans have an intricate mechanism for
2.2.2.2. Changes in zinc homeostasis in Alzheimer’s disease. The controlling intestinal absorption of iron, they lack a mechanism
presence of zinc ions in both plasma and CSF have been reported to (other than bleeding) for elimination of grossly excessive
further decrease in AD compared to age-matched control patients quantities. In the developed world, accumulation of excess iron
(Basun et al., 1991; Baum et al., 2010; Molina et al., 1998). While in a number of tissues from males can begin in early adulthood and
nutritional deficiency is common in advanced age, this further then increase almost linearly with age. Females delay over-
decrease in AD zinc levels may in turn affect brain levels of the ion accumulation through menstruation and/or pregnancy, but
and exaggerate amyloid pathology, through it’s entrapment within postmenopausal women attain parity with men in iron burden
the plaque. As previously described with copper, zinc is highly within a few decades. In contrast, plasma levels of iron are reduced
enriched within AD plaques (1055 mM) compared to normal age- with age and likewise, so are iron-containing proteins such as
matched neuropil (350 mM) (Lovell et al., 1998). Histochemically hemoglobin (Ahluwalia et al., 2000).
reactive zinc deposits are also found specifically localized to Age-related increases in brain iron have been well documented
cerebral amyloid angiopathy deposits and neurofibrillary tangle in all species examined, including humans (Bartzokis et al., 1994;
(NFT)-bearing neurons (Friedlich et al., 2004; Suh et al., 2000). Hallgren and Sourander, 1958), non-human primates (Hardy et al.,
Dyshomeostasis of zinc in the AD brain may arise from 2005) and rodents (Maynard et al., 2002; Roskams and Connor,
inhibition of zinc export by 4-hydroxynonenal, a peroxidation 1994; Sohal et al., 1999; Suh et al., 2005). This accumulated iron is
product of Ab:copper redox activity (see Section 3.2.4) that is also evident with increased ferritin expression; thought to occur as
elevated in AD tissue (Smith et al., 2006b). In addition, rodent an attempt to store surplus intracellular free iron (Connor et al.,
studies have indicated that systemic zinc deficiency increases 1990, 1995; Zecca et al., 2001). As with copper, these changes may
brain zinc retention (Takeda et al., 2001) by suppressing levels of be regional as electron paramagnetic imaging has shown that age-
the intracellular zinc export promoter ZnT1 (Chowanadisai et al., related increases in iron are clustered (Wender et al., 1992).
2005). Indeed, levels of ZnT1, ZnT4 and ZnT6 (Lyubartseva et al.,
2009) as well as ZnT3 (Adlard et al., 2010), are all altered in AD 2.2.3.2. Changes in iron homeostasis in Alzheimer’s disease. The
brain tissue. Interestingly, estrogen can also modulate levels of increase in iron with AD associated senile plaques was first
ZnT3 and thus synaptic zinc (Lee et al., 2004b), this is of particular described by Goodman in 1953 (Goodman, 1953). Since then his
significance as sex is another major risk factor for AD. observation has been repeatedly reinforced by others and
validated with the use of various specific and sensitive techniques
2.2.3. Iron (Bishop et al., 2002; Collingwood and Dobson, 2006; Collingwood
With the exception of a few bacterial species that use et al., 2005; Connor et al., 1992a,b). Within AD plaques, iron
manganese, all other forms of life are iron dependent for a variety accumulates to nearly 3 times that of normal neuropil levels
of essential biochemical processes. It is a vital component of the (940 mM compared to 340 mM) (Lovell et al., 1998). Deregulation
heme in hemoglobin, myoglobin, and cytochromes and is also an of brain iron metabolism is multifactorial and comprises non-
essential cofactor for non-heme enzymes such as ribonucleotide genetic and genetic factors. It may also occur at multiple levels,
reductase, the limiting enzyme for DNA synthesis. When in excess, including iron uptake and release, storage, intracellular metabo-
iron is toxic because it generates superoxide anions and hydroxyl lism and regulation (Ke and Qian, 2007; Zhu et al., 2007).
radicals that react readily with biological molecules, including The increased iron is found primarily complexed with ferritin in
proteins, lipids, and DNA. As a result, humans possess elegant the plaque-associated neuritic processes (Grundke-Iqbal et al.,
control mechanisms to maintain iron homeostasis by coordinately 1990) and within neurons and neurofibrillar tangles (Bouras et al.,
regulating iron absorption, iron recycling, and mobilization of 1997; LeVine, 1997; Morris et al., 1994). While ferritin levels are
stored iron. Disruption of these processes causes either iron- increased in AD, transferrin (involved in the physiological
deficient anemia or iron overload disorder (Zhang and Enns, 2009). transport and utilization of iron concentration) is decreased
6 J.A. Duce, A.I. Bush / Progress in Neurobiology 92 (2010) 1–18

(Connor et al., 1992b). Oxidative stress, potentially involving iron copper levels. Modulation of copper levels in fibroblasts from mice
redox state, is one of the earliest events in AD and seems to be with either chronic copper overload (caused by an over expression
involved in the onset, progression and pathogenesis of the disease of a copper efflux protein in ‘Mobr’ and ‘Modap’ mice) or copper
(Castellani et al., 2007; Zhu et al., 2007). Disruption in brain iron deficiency (caused by a mutation in the copper transporter, ATP7a),
homeostasis through alterations of iron regulatory proteins can can both up- and down-regulate APP mRNA expression (Armen-
increase the vulnerability of cells to oxidative stress (Connor et al., dariz et al., 2004; Bellingham et al., 2004b). A study of healthy
1992b; Pinero et al., 2000). Changes in superoxide levels due to postmenopausal women (n = 25) that consumed either a low
alteration of SOD1 activity also affect iron metabolism in glial and (1 mg/d) or a high (3 mg/d) copper content diet (also supplemen-
neuronal cells (Culotta et al., 2006; Danzeisen et al., 2006). ted with zinc) confirm that a low copper diet was associated with a
Lactoferrin, upregulated in AD, exerts an anti-inflammatory significant decrease in APP expression in platelets (Davis et al.,
function via its inhibitory effect on hydroxyl radical formation 2000).
and, through its anti-oxidative properties, prevents DNA damage Conversely, both in vivo and in vitro studies in APP and b-
(Sacharczuk et al., 2005). amyloid precursor-like protein 2 knockout mice have demonstrat-
Several genetic studies on increased risk factors of AD ed that there are significant increases in copper levels in brain and
demonstrate involvement in the genotype of the hemochromatosis liver tissue as well as cortical neuron and fibroblast primary
gene (HFE), (known to cause iron overload in hemochromatosis), cultures derived from these animals (Bellingham et al., 2004a;
and the transferrin C2 allele (Bertram et al., 2007; Robson et al., White et al., 1999). Furthermore, the over expression of APP in
2004). The combination of the C2 variant of the transferrin gene various transgenic mouse lines carrying a mutated APP gene
(TF-C2) and the C282Y allele of HFE gene (HFE C282Y), or the decreased copper levels both in vivo and in vitro at ages varying
combination of HFE C282Y and HFE H63D are risk factors for from 2.8 to 18 months of age (Bayer et al., 2003; Maynard et al.,
developing AD. Carriers of such combination were at five times 2002; Phinney et al., 2003).
greater risk for AD. The addition of the apolipoprotein E epsilon4 In addition to regulation by copper, APP translation is also
allele further increases the risk of AD (Connor and Lee, 2006; mediated by the IRE type II present in its 50 untranslated region
Lehmann et al., 2006; Namekata et al., 1997). Brain ferritin iron (Venti et al., 2004). The APP mRNA IRE is located immediately
may also influence age- and gender-related risk of neurodegenera- upstream of an interleukin-1 responsive acute box domain (+101
tion as there is a high correlation between brain ferritin iron levels, to +146). APP 50 -UTR conferred translation was selectively down
measured in vivo by MRI, and age (Bartzokis et al., 2007). Jefferies regulated in response to intracellular metal chelation and this is
and colleagues showed that the transferrin homologue; melano- reversed by iron influx (Rogers et al., 2002). Iron regulation of APP
transferrin (p97), that binds iron through a high affinity iron- mRNA through the APP 50 -UTR points to a role for iron in the
binding site, could be used as a potential early detection biomarker metabolism of APP.
for AD (Jefferies et al., 2001; Ujiie et al., 2002).
Interestingly, as with copper, elevated levels of cholesterol and 3.1.3. Metal involvement in secretase processing of APP
iron in combination have been observed to promote an AD-like Metals may also indirectly affect APP by altering APP
phenotype in transgenic animals (Ghribi et al., 2006) and to be a processing. The processing of APP is well documented and involves
risk factor in humans (Mainous et al., 2005). In a set of 6558 US a number of activities by alpha-, beta- and gamma-secretase (for
adults, followed from 1974 to 1992, an elevated risk of AD occurred review see (Gandy, 2005; Ling et al., 2003; Suh and Checler, 2002)).
if both dietary cholesterol and iron were above normal (Mainous Candidate alpha-secretases, tumor necrosis factor-alpha convert-
et al., 2005). ing enzyme (TACE or ADAM-17) and ADAM-10 (Buxbaum et al.,
1998), are members of the cell surface metalloproteinase family of
3. Metal involvement with AD related proteins disintegrin and metalloprotease (ADAM) proteins. Beta-secretase
has been identified as a membrane anchored aspartyl protease
3.1. b-amyloid precursor protein termed b-APP-site cleaving enzyme or BACE (Cai et al., 2001; Luo
et al., 2001; Sinha et al., 1999; Vassar et al., 1999; Yan et al., 1999)
3.1.1. APP binding to metal while the high molecular weight active gamma-secretase complex
There are putative metal binding sites for copper (Atwood et al., is comprised of at least four components, presenilin, APH-1, PEN-2
2000; Barnham et al., 2003; Hesse et al., 1994; Simons et al., 2002; and nicastrin (De Strooper, 2003). Beta- and gamma-secretase
Valensin et al., 2004) and zinc (Bush et al., 1993, 1994a,b,c) in the result in a hierarchical and site-specific cleavage of APP to generate
APP sequence. Copper binding sites are located at the amino- Ab peptides of varying length. All three secretases are now known
terminal ectodomain of APP and also within the Ab sequence of to have interactions with different metal species.
APP (described further in Section 3.2.1). The structure of the APP The alpha-secretase, TACE, is a multidomain protein that
copper binding domain consist of four ligands (His-147, His- contains a zinc ion within its catalytic domain (Cross et al., 2002).
151,Tyr-168 and Met-170 (Barnham et al., 2003), which show Following TACE synthesis, it is maintained in an inactive state
structural homology to other copper chaperones. mediated by an intramolecular bond between a cysteine in the
The binding of zinc to APP involves a conserved region of amino prodomain and a zinc atom in the catalytic site (referred to as the
acids between position 170 and 188 of APP (Bush et al., 1993, ‘‘cysteine-switch’’ motif) of the protein. It has subsequently been
1994a). This domain consists of two key cysteine ligands at shown, however, that TACE enzymatic activity can be mediated
position 186 and 187 which are crucial to the binding, as well as independently of this classic cysteine-switch motif, such that other
other potential ligands (e.g. Cys-174, Met-170, Asp-177 and Glu- regions of the prodomain can interact with the catalytic domain
184) (Ciuculescu et al., 2005). The binding of zinc may have an (and the zinc ion) to inhibit enzymatic activity (Buckley et al.,
important functional role in the dimerization of APP (Ciuculescu 2005; Gonzales et al., 2004). Similarly, the activity of ADAM-10 can
et al., 2005), which could occur in vivo (Scheuermann et al., 2001). be inhibited by a dominant negative form of this candidate alpha-
secretase that has a point mutation in its zinc-binding site
3.1.2. Regulation of APP by metal (Lammich et al., 1999). Iron may also alter APP cleavage by
APP is a ubiquitously expressed protein that potentially modulating furin, a proconvertase involved in the regulation of
functions to participate in metal ion homeostasis. There are a alpha-secretase-dependent processing (Silvestri and Camaschella,
number of in vivo studies that highlight the sensitivity of APP to 2008).
J.A. Duce, A.I. Bush / Progress in Neurobiology 92 (2010) 1–18 7

BACE1 binds copper in its C-terminal domain (via cysteine 2.2.2) and may account for the synaptic localization of Ab
linkages) and co-immunoprecipitates in brain homogenate with precipitation (Fig. 1B) (Terry, 1996). Iron will also precipitate Ab
the copper chaperone CCS (Angeletti et al., 2005). BACE1 interacts under acidic conditions, although only copper and zinc have been
with domain I of CCS (the domain that mediates copper binding) shown to be specifically associated with senile plaques and to
and may directly compete with SOD1, as BACE1 expression reduces copurify with Ab extracted from the human brain (Dong et al.,
the activity of SOD1. Overexpression of CCS in cells already 2003; Opazo et al., 2002). Of note, these affinities for metals are so
overexpressing BACE1 will restore SOD1 activity (Angeletti et al., low that the concentrations of zinc, copper and iron present in
2005). Accordingly, both BACE1 and CCS are transported together common laboratory buffers are sufficient to induce nucleation of
through the axon (Angeletti et al., 2005). Endocytosis of APP to Ab, leading to accelerated aggregation of the peptide (Huang et al.,
BACE1-rich endosomes is required to initiate b-cleavage of APP 2004).
(Vassar et al., 1999), and this event is sensitive to flotillin-2 and There are up to 2.5 metal binding sites on Ab, possibly
cholesterol depletion; both of which are known to be enriched coordinated by oligomers (Atwood et al., 2000). Binding of the
within lipid rafts along with a subset of cellular APP (Ehehalt et al., oxidized forms of both zinc and copper (Atwood et al., 2000; Bush
2003; Schneider et al., 2008). It has recently been observed that et al., 1994c; Danielsson et al., 2007; Karr and Szalai, 2008; Syme
elevated copper reduced flotillin-2 association with lipid rafts and et al., 2004; Syme and Viles, 2006) as well as the reduced form of
thereby reduced the endocytosis of APP via lipid rafts (Hung et al., copper (Himes et al., 2008; Hureau et al., 2009; Shearer and Szalai,
2009). 2008) to Ab in vitro (either in aqueous solution or membrane
Presenilin, a subunit of the gamma-secretase complex, is also mimetic environments) is mediated by nitrogen ligands from the
sensitive to metal levels. Neonatal mouse cortical cultures have three histidine residues at positions 6, 13 and 14, along with an
shown that exogenous zinc administration (but not copper or iron) oxygen ligand (Curtain et al., 2001, 2003). Rat/mouse Ab
increases the C-terminal fragments of presenilin 1 by enhancing de (containing amino acid substitutions for histidine at position 13,
novo synthesis of the protein (Park et al., 2001). More recent data arginine at position 5 and tyrosine at position 10) have significant
suggests that zinc chloride causes the oligomerisation of an APP structural changes in the metal ion coordination chemistry
gamma-secretase substrate and inhibits its cleavage in vitro, compared to human Ab (Gaggelli et al., 2008). This supports
supporting a role for zinc dysregulation in abnormal Ab processing histidine at position 13 as having a predominant role in metal
(Hoke et al., 2005). binding and may be responsible for the lower aggregation
A recent report determined APP from human fibroblasts to be propensity of rat Ab when complexed to copper or zinc. It may
processed simultaneously by both alpha- and beta-secretase, as also explain why these animals are exceptional among mammals
copper-deficient fibroblasts secrete beta-secretase cleaved soluble for not forming cerebral Ab deposits with age (Vaughan and Peters,
APP N-termini but also produce alpha-secretase cleaved APP C- 1981). However, a role for the two other substitutions in the rat
termini (C83) (Cater et al., 2008). Copper deficiency in these cells sequence in aggregation cannot yet be excluded.
also markedly reduced the steady-state level of APP mRNA while The oxygen ligand for metal coordination is most likely from the
APP protein level remained constant, indicating that copper aspartate at position 1 (Dorlet et al., 2009; Drew et al., 2009b;
deficiency may accelerate APP translation (Cater et al., 2008). Mekmouche et al., 2005) and alanine at position 2 (Dorlet et al.,
This may be cell-type specific though as copper deficiency in 2009; Drew et al., 2009a). While other oxygen ligands have been
human neuroblastoma cells did not affect the cleavage of APP proposed, such as tyrosine at position 10, glutamate at position 11
(Cater et al., 2008). or H2O, these have all been contested (Karr and Szalai, 2007; Syme
Together these data demonstrate that there may be a direct et al., 2004), particularly as peptides lacking the first one to three
interaction of metals on the secretases and, therefore, also on APP amino acids do not bind copper in the same fashion as the native
processing. In addition, the zinc binding domain in the Ab region of peptide (Karr et al., 2005). Recent in vitro analysis of the copper-
the APP sequence spans the alpha-secretase cleavage site and may, binding domain in the 3–42 truncated Ab (Ab3–42) sequence has
therefore, modulate the cleavage of Ab from APP and also protect demonstrated that it self-aggregates more rapidly than Ab1–42 and
Ab from proteolytic degradation (Bush et al., 1994b). is a more potent initiator of Ab1–40 aggregation, particularly in the
presence of copper (McColl et al., 2009). Clearly, while the
3.2. b-amyloid coordination of metal in Ab is generally accepted as a square-
planar configuration, involving three histidine imidazoles, the
3.2.1. Ab binding to metal fourth ligand is still yet to be fully elucidated. More detailed
Ab is rapidly precipitated by zinc over a broad pH range (pH reviews of the metal binding chemistry of Ab have been recently
5.5–7.5) (Atwood et al., 1998; Bush et al., 1994b,c; Yoshiike et al., published (Faller and Hureau, 2009; Rauk, 2009; Rozga and Bal,
2001), including physiological pH (pH 7.4) (Cherny et al., 1999), 2009).
whereas aggregation of Ab by copper and iron is only induced Ab binds equimolar amounts of copper and zinc at pH 7.4, but
under mildly acidic conditions (pH 6.8–7.0) (Atwood et al., 1998, under the conditions of acidosis that are thought to occur in the AD
2000; Ha et al., 2007). The Ab affinity for copper is well below brain (pH 6.6), copper displaces zinc from Ab (Atwood et al., 2000).
physiological concentrations achieved within the brain, and is As copper has been shown to promote the aggregation of Ab in
reported to be as high as attomolar for aggregated Ab1–42 (Atwood vitro and Ab is known to bind copper in AD brain (Dong et al., 2003;
et al., 2000), although some reports have lower apparent affinities Lovell et al., 1998; Opazo et al., 2002) it may seem surprising that
in the submicromolar range for soluble Ab (Garzon-Rodriguez increasing brain copper levels in transgenic models of the disease
et al., 1999; Syme et al., 2004). The differences between reported would inhibit Ab pathology (see Section 3.2.2). However, this is
affinities are likely due to the different techniques used to consistent with evidence that copper abnormally redistributes in
determine these values and to the conditions (e.g. buffer) of the AD and collects outside of the cell, leaving the tissue relatively
experiments (this is covered in more detail by (Faller and Hureau, deficient. This would also go someway to explain the deficiency of
2009; Rozga and Bal, 2009)). The affinity for zinc is also in the low copper-dependent enzymatic activities in AD such as cytochrome c
micromolar range (Bush et al., 1994c; Huang et al., 1997). These oxidase (Cardoso et al., 2004; Cottrell et al., 2001; Maurer et al.,
levels are of particular relevance to neurotransmission as such 2000; Sullivan and Brown, 2005) and SOD1 (De Deyn et al., 1998;
concentrations of free ionic zinc and copper are released at the Omar et al., 1999), as well as the reported elevation of serum
synapse (Lee et al., 2002; Schlief et al., 2005) (see Sections 2.2.1 and copper (Squitti et al., 2002a, 2005).
8 J.A. Duce, A.I. Bush / Progress in Neurobiology 92 (2010) 1–18

3.2.2. Ab production modulated by metal NEP is sufficient to reduce Ab levels, slow or completely prevent
In order to investigate if metal-mediated modulation of APP plaque formation and extend lifespan (Leissring et al., 2003).
processing could result in the observed physiological change in Ab Both these proteases are members of the zinc metallopeptidase
generation, a number of studies have been carried out to examine family of proteins that share a common primary structure in their
the effect copper and zinc may have on the brain from APP sequence that is involved in the binding of zinc. This structure
transgenic models. One of the simplest ways to do this is by dietary possesses the short consensus sequence (HEXXH) in which the two
supplementation. Copper given to APP23 mice (overexpressing histidines act as the ligand for the catalytic zinc. Some family
human APP with the AD-related Swedish mutation) (Bayer et al., members may also have an elongated C-terminal motif
2003) for 3 months was sufficient to significantly increase copper (HEXXHXXGXXH/D) in which there is an additional zinc-binding
levels within the brain compared to non-supplemented APP23 histidine (or aspartate) (Gomis-Ruth, 2003). This potentially
littermates and this was complemented with a normalization in means that both IDE and NEP activities could be influenced by
activity of the copper-dependent enzyme; SOD1. Relative to Ab, abnormal zinc metabolism. In addition, it has recently been
the increase in bioavailable copper in these mice resulted in a demonstrated that both enzymes are oxidatively modified in the
lowering of soluble and insoluble Ab (Bayer et al., 2003). Dietary AD brain (Caccamo et al., 2005; Shinall et al., 2005; Wang et al.,
zinc has also been administered to 2 transgenic models of AD, the 2003). There is a concentration dependent loss in activity of both
Tg2576 (overexpressing human APP with the AD-related Swedish IDE and NEP upon treatment with hydrogen peroxide. More
mutation) and TgCRND8 (overexpressing human APP with the AD- specific treatment with various reactive oxygen species differen-
related Swedish and Indiana mutations). The provision of a zinc- tially affected the activity of NEP and IDE. Likewise, treatment with
enriched diet also reduced amyloid-beta plaque deposits, but 4-hydroxy-2-nonenal (HNE), which occurs as a function of
potentiated Alzheimer-like spatial memory impairments (Linkous oxidative stress and is augmented in the AD brain, also
et al., 2009). Correspondingly, lowering dietary zinc in another differentially affected the activity of these metallopeptidases.
transgenic model of AD caused a significant increase in plaque These data suggest that the generation of reactive oxygen species,
volume (Stoltenberg et al., 2007). perhaps as a product of metal interactions with Ab (see Section
Similar copper related observations were described with the 3.2.4), may serve to inactivate the principal degradative enzymes
TgCRND8 when crossed with a transgenic model containing the involved in the regulation of Ab.
‘toxic-milk’ mutation in the gene encoding the copper transport An alternative potential mediator of Ab catabolism is plasmin.
protein, ATP7b (Phinney et al., 2003). The mutation in ATP7b Activation of plasmin can be induced by Ab through the
causes the inability of copper to be loaded into secretory vesicles upregulation of its precursor, tissue-plasminogen activator (tPA)
resulting in an accumulation of intracellular copper. TgCRND8 (see (Ling et al., 2003)). While plasmin is not a zinc-dependent
mice with the ATP7b mutation were confirmed to have elevated enzyme in the same fashion as NEP and IDE, it has been
copper within the brain and demonstrated a decrease in amyloid demonstrated that it too may be regulated by site-specific
plaque content as well as the soluble and insoluble Ab levels oxidation, perhaps via a modification of the histidine residue in
suggesting the intracellular shift in copper reduces Ab aggregation. its active site (Lind et al., 1993). Furthermore, inhibition of
Plasma levels of Ab were also reduced, including the endogenous substrate cleavage by tissue-plasminogen activator occurred when
murine Ab content, confirmed a decreased extracellular Ab co-incubated with copper and ascorbate (Lind et al., 1993). Lastly,
content. the tissue-plasminogen activator/plasmin system may also be
As previously explained, the highest pool of free zinc is present involved in the activation of matrix metalloproteinases (MMPs),
in synaptic vesicles that are released during synaptic transmission another family of zinc-dependent enzymes particularly relevant to
of neocortical glutamatergic fibers (Quinta-Ferreira and Matias, the degradation of extracellular Ab. Recent in vitro studies have
2005). The activity of the zinc transporter ZnT3 is required for the shown that MMP2 can produce site specific degradation of Ab that
passage and pooling of zinc within these pre-synaptic vesicles. The then becomes inaccessible when the peptide is incubated with
genetic ablation of ZnT3 in the Tg2576 mouse causes a profound zinc, but not copper, to form amyloid (Crouch et al., 2009b) (also
reduction in the cerebral plaque load (Lee et al., 2002). Further- see Sections 4.3 and 4.4 and (Yong et al., 1998) for review).
more, there is a depletion of an exchangeable zinc pool in the Together these studies provide considerable evidence that a
cerebrovascular wall, which is also associated with a significant mis-metabolism of metal ion homeostasis is sufficient to
reduction in cerebral amyloid angiopathy (Friedlich et al., 2004). It significantly alter the normal course of APP processing and Ab
is the high concentration of synaptically released zinc in the generation, deposition and degradation in biologically relevant
synaptic cleft that is believed to play an essential role in amyloid animal models. These animal studies are supported by the wealth
formation in transgenic models for AD (Fig. 1B). of in vitro data characterizing the interactions between metal ions
and Ab.
3.2.3. Metal modulation of Ab degradation
The involvement of metal ions in the aggregation of Ab within 3.2.4. Metal mediated production of reactive oxygen species
the brain only represents one part of the biological equation, as Overwhelming evidence in the literature describes oxidation
metal ions may also participate in the clearance/degradation of Ab injury as a major feature of AD. While under certain in vitro
(for review on Ab catabolism, see (Carson and Turner, 2002; Ling conditions Ab has been shown to act as an antioxidant (Baruch-
et al., 2003)). Two of the primary Ab degrading enzymes identified Suchodolsky and Fischer, 2009; Nadal et al., 2008; Shearer and
are insulin-degrading enzyme (IDE) and neprilysin (NEP), a neutral Szalai, 2008) with neurotrophic and neuroprotective effects
endopeptidase. IDE, which degrades a variety of substrates (e.g. (Whitson et al., 1989), compelling evidence has shown that the
insulin and amylin) including Ab, is ubiquitously expressed and redox activity for variants of Ab is greatest for Ab1–42 human > Ab1–
localized intra- and extracellularly within the CNS. With the use of 40 human » Ab1–40 mouse  0 (Huang et al., 1999a), corresponding to
mice deficient in either enzyme, Ab catabolism has been shown to the potentiation of Ab toxicity by copper (Huang et al., 1999b;
significantly decrease resulting in an increase of endogenous Ab Opazo et al., 2002) and the prevalence of Ab1–42 over Ab1–40 in
(Farris et al., 2003; Iwata et al., 2001). It has subsequently been both sporadic and familial AD. Such studies have shown the
demonstrated that NEP is capable of degrading both monomeric coordination of either oxidized copper (Cu2+) or iron (Fe3+) to Ab
and oligomeric forms of Ab1–40 and Ab1–42 (Kanemitsu et al., result in the reduction of the metal ions and is the subject of a
2003). Conversely, the transgenic overexpression of either IDE or focused recent review by Hureau and Faller (2009). Ab’s redox
J.A. Duce, A.I. Bush / Progress in Neurobiology 92 (2010) 1–18 9

activity leads to the generation of reactive oxygen species (ROS), neurochemistry is not as developed as APP’s or Ab’s, but there
when not accompanied by the oxidation of another moiety, are a growing number of in vitro reports that indicate that tau may
(Puglielli et al., 2005) and the production of H2O2, in the presence well participate in the metal-related abnormalities observed in AD.
of biological reducing agents used as electron donors (Dikalov However, it is still unclear how the relationship between Ab and
et al., 2004; Haeffner et al., 2005; Murray et al., 2005; Nelson and tau lesions may be linked by metal interactions. Various repeat
Alkon, 2005; Opazo et al., 2002; Puglielli et al., 2005; Smith et al., motifs on tau bind copper and iron in a pH- and stoichiometric-
2006a; Tabner et al., 2002). Possible biological reductants include dependent fashion but have yet to have an affinity determined
cholesterol, long-chain fatty acids, 3,4-dihydroxyphenylalanine under in vivo physiological conditions. Even so, the resulting metal
(DOPA), dopamine quinone, dihydroxyindol, isodityrosine and binding has been shown to result in a conformational change in
ascorbate (Dong et al., 2003; Haeffner et al., 2005; Huang et al., tau, promoting its phosphorylation (Malm et al., 2007) and
1999a; Jiang et al., 2007; Murray et al., 2005; Nelson and Alkon, potentially inducing aggregation (Ma et al., 2006; Ma et al.,
2005; Opazo et al., 2002; Puglielli et al., 2005; Smith et al., 2007; 2005; Zhou et al., 2007). Only Fe3+ induces aggregation of
Yoshimoto et al., 2005). This can lead to a variety of Ab side-chains hyperphosphorylated tau and this can be reversed by its reduction
being oxidized, and to covalent oligomerisation. One likely back to Fe2+ (Yamamoto et al., 2002). Indeed, the treatment of tau
candidate from the copper-mediated redox reaction is the sulphur aggregates from the AD brain with reducing agents result in the
atom of methionine-35 becoming methionine sulfoxide (Ali et al., resolubilization of tau and a release of Fe2+, further demonstrating
2005; Ciccotosto et al., 2004), however this remains in doubt as it the potential role of iron in NFT formation (Yamamoto et al., 2002).
has also been deemed unnecessary for toxicity (da Silva et al., Prior to these reports, NFTs were reported as capable of binding
2009; Hou and Zagorski, 2006). Other adducts can also be adventitious copper and iron in a redox-competent manner, acting
generated, including aldehyde adducts to the lysine residues as a source for reactive oxygen species within the neuron (Sayre
(Chen et al., 2007) and the tyrosine at position 10 (Barnham et al., et al., 2000). In support, recent data indicates that tau interaction
2004; Haeffner et al., 2005). with Cu2+ can mediate the generation of hydrogen peroxide in a
Tyrosine at position 10 is particularly susceptible to free radical manner analogous to Ab (Su et al., 2007).
attack, due to the conjugated aromatic ring, and in the presence of While tau is the principal component of NFT, there are other
Cu2+ and H2O2 forms dityrosine cross-linked oligomers (Atwood cytoskeletal components present that are also known to interact
et al., 2004). Both dityrosine and 3-nitrotyrosine are elevated with metals. Neurofilaments (NF), for example, show phosphory-
within the neuronal lesions in AD brain (Hensley et al., 1998). The lation dependent alterations very early in the AD cascade and are
formation of dityrosine linkage in Ab cannot be proteolytically also found within the NFT and in association with the dystrophic
degraded (Atwood et al., 1998) and facilitates further peptide neurites surrounding Ab plaques (King et al., 2000). Purification of
aggregation, leading to the formation of higher order oligomers this multi-subunit protein demonstrated that it stoichiometrically
(Barnham et al., 2004). In addition, Ab radicals formed after binds at least one copper and four zincs (Pierson and Evenson,
reduction of copper can form covalent adducts onto other proteins. 1988). The assembly of NFs may be mediated, in part, by metal ions
Peroxidases like cycloxygenase 2 (COX2) are particularly vulnera- such as copper, which foster the assembly of the light NF subunit.
ble because of the formation of dityrosine bridges, and cyclox-
ygenase 2:Ab covalent complexes are elevated in AD brain 3.3.2. Metal modulation through kinases
(Nagano et al., 2004). 2-oxo-histidine adducts of Ab have also Many of the kinases that phosphorylate tau (ultimately
been extracted from AD plaques (Metodiewa, 1998) as have N3- resulting in hyperphosphorylation and dissociation of tau from
pyroglutamate modified forms of Ab that are the main ligands for the microtubules) are induced by metal ions zinc, copper and iron.
the amyloid positron emission tomography (PET) ligand Pittsburgh Zinc is able to activate rac-beta serine/threonine protein kinase
Compound-B (PIB) (Maeda et al., 2007). (PKB), glycogen synthase kinase 3b (GSK-3b), extracellular signal-
The byproduct H2O2 is also freely permeable across all tissue regulated kinase (ERK1/2), c-Jun NH2-terminal kinase (JNK), p38
and cellular boundaries, and can thus further react with reduced and p70 S6 kinase, that are all consequently involved in tau
metal ions (Fe2+, Cu+) to generate hydroxyl radicals via Fenton hyperphosphorylation in SH-SY5Y and N2a cells (An et al., 2005;
chemistry. This in turn generates lipid peroxidation products such Bjorkdahl et al., 2005). Tangle-bearing neurons from AD brain have
as HNE, protein carbonyl modifications, and nucleic acid adducts also been observed to fill with zinc (Bjorkdahl et al., 2005; Suh
(e.g. 8-OH guanosine), all of which typify AD neuropathology et al., 2005). NFs are also phosphorylated at various sites by a
(Haeffner et al., 2005; Jiang et al., 2007; Markesbery and Lovell, number of the tau-kinases induced by zinc (Bjorkdahl et al., 2005).
2007; Murray et al., 2007; Nelson and Alkon, 2005; Opazo et al., Phosphoinositol-3-kinase (PI3K) is rapidly activated by copper
2002; Puglielli et al., 2005; Smith et al., 2006a, 1996, 1997). In AD, with subsequent downstream activation of serine/threonine
the H2O2 scavenging defenses, e.g., catalase and glutathione protein kinase (Akt) and phosphorylation of GSK3 (Ostrakhovitch
peroxidase, are likely to be taxed by the catalytic generation of et al., 2002). Copper may also be able to activate membrane
H2O2 from the Ab:metal complexes. receptors such as the epidermal growth factor receptor (EGFR) and
As a whole, metal dysregulation may not only affect the course result in downstream activation of mitogen-activated protein
of the disease by interaction with an amyloidogenic pathway but kinases (MAPK), including ERK1/2 (Wu et al., 1999).
also with a number of other AD-related pathways. There are other In contrast, the treatment of hippocampal neuron cultures with
metal-modulated cellular processes and downstream events that iron citrate results in a decrease in tau phosphorylation at AD-
must be considered. These also may interact with and mediate the related epitopes, possibly from a decrease in the activity of the
ageing process and subsequent disorders such as AD. Cdk5/p25 complex, where p25 may be a regulator of the activity of
protein kinase Cdk5 (Egana et al., 2003).
3.3. Tau
3.4. Apolipoprotein E
3.3.1. Tau binding to metals
The microtubule-associated protein tau is the principal Apolipoprotein E (ApoE) allelic status is another major risk
component of the paired helical filaments that comprise neurofi- factor for AD (Corder et al., 1993) and modulates zinc and copper
brillary tangles (NFTs) in a number of neurodegenerative disorders, induced Ab precipitation in accordance with the isoforms ability to
including AD. As yet the literature linking tau to metal bind these ions (Miyata and Smith, 1996; Moir et al., 1999). The
10 J.A. Duce, A.I. Bush / Progress in Neurobiology 92 (2010) 1–18

fourth isoform of ApoE (ApoE4) is the least protective for risk of AD. binding small-molecule, metallothionein-3 (MT-3) has chelating
Interestingly this isoform differs from the other ApoE alleles by properties and is important for copper and zinc homeostasis
containing a cysteine to arginine substitution at position 112 for within neurons, astrocytes and the extracellular space of the brain
isoform 3 and 158 for isoform 2. This cysteine is believed to be (Vasak and Hasler, 2000). It has been found to be downregulated in
involved in transition metal binding and may be related to the AD (Yu et al., 2001) and was recently used to protect cultured
diminished antioxidant effect of the ApoE4 allele (Moir et al., neurons from Ab toxicity (Meloni et al., 2008). MT-3’s mechanism
1999). appears to consist of reducing and sequestering the copper from
Interestingly, increased tau phosphorylation in apoE4 trans- Ab and swapping it for a zinc ion carried within MT-3. While MT-3
genic mice has been associated with ERK activation, already may be involved in the mechanism of action for other therapeutic
described to be modified by zinc, suggesting that apoE4 and zinc compounds (see Fig. 1), it has not to date been used as a therapeutic
may act in concert to contribute to the pathogenesis of AD (Harris for AD itself.
et al., 2004). Desferrioxamine (DFO), a preferential iron chelator able to also
bind zinc, copper and aluminium (Keberle, 1964), was the first
4. Therapeutics and diagnostic targeting of metal ions in AD medicinal small-molecule chelator tested for the treatment of AD.
Crapper-McLachlan and colleagues reported a significant improve-
One logical and increasingly popular theory for the use of ment of some cognitive functions in AD patients on a 2-year-long
neurotherapeutic small molecules in AD is to target the initiating placebo controlled phase II clinical trial of DFO injected intramus-
event in the generation of free radicals. This may be done by cularly (Crapper McLachlan et al., 1991). However, as with most
administering antioxidants or by employing metal complexing metal chelators, DFO is a large hydrophilic molecule, which is not
agents to prevent the metal ions from participating in redox orally bio-available and does not efficiently penetrate the BBB.
chemistry. This review will cover some of the more promising Hence, it is not clear whether the beneficial effect seen with DFO
agents, however Scott & Orvig provide a more extensive recent treatment was due to the drug’s interaction with metals, or due to a
review (Scott and Orvig, 2009). different mechanism all together (Cuajungco et al., 2000). A
A common mistaken belief in targeting metal ions for smaller clinical trial with d-penicillamine also produced largely
therapeutics is that chelation, meaning the removal of metal ions inconclusive cognition results due to the high drop-out rate of
from tissue, is the obvious intervention. While there are several patients but did show a decrease in serum oxidative markers after
medical chelators in use, their approved use is confined to genuine 6 months (Squitti et al., 2002b).
overexposure situations (e.g., Wilson’s disease, or lead toxicity), or Preclinically a number of metal chelators have been tested on
to rheumatoid arthritis. There have been a few placebo-controlled AD models with varying success. The hexadentate chelator, DP-109
clinical trials on AD with chelators reported, yet there is still no (DPharm, Rehovot, Israel), is a large synthetic pro-drug that
evidence that AD will respond to traditional medical chelation and becomes activated following the cleavage of its two long-chain
the greater risk is that the removal of essential metal ions will esters. A 3-month oral administration of DP-109 to Tg2576 mice
result in adverse effects (e.g. iron deficiency anemia). Another reduced levels of aggregated insoluble Ab in amyloid plaques and
important property of a potential neurodegenerative therapeutic is cerebral amyloid angiopathy and conversely increased Ab via
its ability to cross the BBB. This has previously limited a large resolubilization (Lee et al., 2004a). As with DFO, how DP-109 exerts
number of common metal chelators as possibilities due to their its anti-amyloidogenic effect is still unclear, as it is not expected to
hydrophilic nature. easily cross the BBB. Recently, DP-109 and DP-460 (another
While these problems can, to some extent, be overcome by lipophilic chelator that preferentially binds calcium, copper and
engineering small molecules that target specific compartments or zinc) were reported to have neuroprotective effects in a model of
organelles, complex situations such as the pooling of metals in amyotrophic lateral sclerosis (a transgenic mouse containing the
plaques and their relative deficiency within neighbouring cells, call G93A mutation in SOD1) (Petri et al., 2007), another neurodegen-
for the development of small molecules with more sophisticated erative disease associated with metal imbalance (Carri et al., 2003;
properties, e.g. ionophores. Watanabe et al., 2007). Other chelating agents have been reported
to have different effects in vitro, including reduced Ab1–42-induced
4.1. Antioxidants oxidative stress (Perluigi et al., 2006) as well as the solubilization of
both hypophosphorylated tau (Shin et al., 2003) and Ab from AD
As a preventative approach antioxidant molecules may be used brain (Cherny et al., 2000). Further in vivo studies with these
for their ability to neutralize free or incorrectly bound metals, chelators are required to advance this therapeutic route and to rule
thereby interfering with the ‘down-stream’ generation of reactive out any systemic effects.
oxygen species and other radicals (Behl and Moosmann, 2002).
Numerous molecules with antioxidant properties, such as oestro- 4.3. Metal complexes
gen, melatonin, vitamin C and E (L-ascorbate and - topopherol,
respectively), ginkgo bilboa extract, curcumin and flavonoids have Metallo-complexes are emerging as a new alternative thera-
neuroprotective effects against Ab-induced toxicity in cell- based peutic approach for AD. Unlike chelation, the rational is to remove
experiments (Shishodia et al., 2005; Zatta et al., 2003) and animal metals from biologically deleterious sites and potentially deliver
models (Cheng et al., 2006; Defeudis, 2002; Feng et al., 2004; them to areas of deficiency, thereby maintaining overall metal
Matsubara et al., 2003; Yang et al., 2005), but have shown homeostasis. Currently, metallo-complexes are able to complete
conflicting results in a clinical setting (Oken et al., 1998; Schneider this strategy either by delivering metals to the relevant cellular
et al., 2005; Zandi et al., 2004) and may even be detrimental (Lloret compartment using pyrrolidine dithiocarbamate (Metal-PDTC) or
et al., 2009). bis(thiosemicarbazone) (Metal-BTSC) or by preventing the harmful
binding of metal to Ab using platinum (Pt2+) complexed to 1,10
4.2. Metal chelators phenanthroline derivatives (L-PtCl2).
PDTC’s ability to interact with copper and zinc gives this metallo-
Metal chelation in an organism depletes the total pool of complex anti-inflammatory, antioxidant and anti-apoptotic prop-
bioavailable metals by binding two or more metal ions to form a erties (Hayakawa et al., 2003; Liu et al., 1999; Schreck et al., 1992). It
cyclic ring that immobilizes the metal ions. Though not a metal ion inhibits the transcription-factor regulator nuclear factor-kB (NF-kB)
J.A. Duce, A.I. Bush / Progress in Neurobiology 92 (2010) 1–18 11

in a neonatal hypoxia-ischaemia model, as well as activating Akt and generation (Reznichenko et al., 2006). A further beneficial effect of
inhibiting GSK-3b (Nurmi et al., 2006). It is suggested to have a iron depletion could arise from inhibition of hypoxia-inducible
similar role in an AD model (APP/PS1 double transgenic mice) factor; prolyl 4-hydroxylases, which is known to be neuroprotec-
(Zhang et al., 2009). Oral administration of PDTC was able to increase tive (Siddiq et al., 2005).
cerebral copper levels as well as decrease tau-phosphorylation via
down-regulation of the GSK-3 signaling cascade leading to an 4.4. Metal-protein attenuating compounds (MPACs) or Ionophores
improvement in spatial memory (Malm et al., 2007). While oral
administration of this drug led to no effects on amyloid burden, glial Similar to metal-complexes, MPACs are able to compete with
activation or oxidative stress (Malm et al., 2007; Zhang et al., 2009), endogenous ligands for metal ions. They target the harmful
daily intaperitoneal injection over 5 months of the same drug ‘upstream’ metal-protein reactions by using their moderate,
attenuated astrogliosis and markedly increased cerebral Ab1–42 reversible, affinity towards metals and restore normal metal
burden (Zhang et al., 2009). levels in specific cellular compartments (Masters and Beyreuther,
Metal-BTSCs are small, neutral compounds that cross mem- 2006). The first-generation molecule was clioquinol (CQ; 5-chloro-
branes and have been previously used for a range of different 7-iodo-8-hydroxyquinoline). Originally prescribed as an oral anti-
purposes, particularly as cancer imaging agents (Dearling et al., amebic in addition to being used for the treatment of dysentery
2002; Green et al., 1988; West et al., 1993). Metal-BTSCs and diarrhoea (Flagstad, 1977; Neldner and Hambidge, 1975), this
successfully cross the BBB and are relatively non-toxic (Dearling quinoline compound is absorbed quickly as it is highly lipophilic
et al., 2002; Green et al., 1988). Glyoxalbis(N(4)-methyl-3- and is excreted easily through its conversion to glucuronated and
thiosemicarbazonato)copper(II) (Cu-GTSM) is a metal-BTSC that sulphate metabolites (Bondiolotti et al., 2007; Kotaki et al., 1983;
binds and transports copper into cells and is engineered to release Liewendahl and Lamberg, 1967). However, its oral preparation was
the bound copper in response to the normal intracellular reducing withdrawn from the market during the 1960s to 1970s, as it was
environment (Dearling et al., 2002; Donnelly et al., 2008; suspected to be involved in the development of subacute myelo-
Fujibayashi et al., 1997). The significance of this is that rather optico-neuropathy (SMON) (Osterman, 1971; Shimada et al., 1971;
than just increasing cellular copper levels, Cu-GTSM increases Tsubaki et al., 1971), possibly confined to iron deficient individuals
cellular copper bioavailability. Importantly, this allows the who consumed overdoses for prolonged periods (Yassin et al.,
interaction of the copper with cellular components and subse- 1998; Yassin et al., 2000).
quent triggering of cell signal mechanisms. In contrast, alternative Despite these previous complications with the compound, oral
Metal-BTSCs such as diacetylbis(N(4)-methyl-3-thiosemicarbazo- treatment with CQ in an AD mouse model (Tg2576) resulted in a
nato) copper(II) (Cu-ATSM) also increase cellular copper but the reduction of cortical deposition of amyloid (49%), supporting initial
metal is not released from the organic scaffold and therefore is not studies showing CQ’s ability to dissolve synthetic Ab:copper/zinc
bioavailable, except under hypoxic conditions (Dearling et al., aggregates and amyloid deposits from post-mortem AD brain
2002; Donnelly et al., 2008). Treatment of CHO cells over- (Cherny et al., 2001). The 9-week trial led to an improvement in the
expressing APP with Cu-GTSM has shown that copper bioavail- general health and weight of these mice compared to untreated
ability is significantly increased and the levels of secreted Ab are littermates (Cherny et al., 2001) and an improved cognitive
reduced in a dose-dependent manner that correlated with copper performance in the Tg2576 and APP/PS1 mouse models (Adlard
uptake (Donnelly et al., 2008). This effect was mediated by et al., 2008). Results indicated CQ was an ionophore rather than a
activation of PI3K resulting in GSK3 inhibition, up-regulation of chelator as it was also able to efficiently cross the BBB, complex
JNK and ERK activity and finally activation of MMPs (White et al., with zinc-metallated Ab and increase both copper and zinc in the
2006). In contrast, no effect on cell signaling, the MMPs or Ab levels brains of treated mice (Adlard et al., 2008; Cherny et al., 2001;
were observed in Cu-ATSM treated cells supporting the hypothesis Opazo et al., 2006). CQ has only a moderate affinity for copper and
that the metal must be bioavailable (i.e. released from the organic zinc, through a square, planar arrangement (Di Vaira et al., 2004),
scaffold to induce therapeutic effects). Similar results were which is sufficient to facilitate dissociation of these metal ions
reported in a neuronal cell line as well as in an orally administered from the low affinity metal binding sites of Ab. While, subsequent
AD mouse model with Cu-GTSM mediated inhibition of GSK3b also studies have shown CQ to only reverse copper-suppressed, but not
leading to inhibited phosphorylation of tau (Crouch et al., 2009a). zinc-suppressed, Ab1–40 fibril formation (Raman et al., 2005), CQ is
Cu-GTSM restored cognitive performance in the APP/PS1 trans- able to rescue copper-mediated Ab toxicity in neuronal cell culture
genic mice to levels expected for cognitively normal animals (Abramov et al., 2003) and markedly reduce Ab oligomeric
(Crouch et al., 2009a). targeting to the synapse (Deshpande et al., 2009). However, other
In an alternative mechanism of treating AD with metallo- studies still highlight that CQ increases oxidative neurotoxicity
complexes, L-PtCl2 were designed to bind and alkylate the (Benvenisti-Zarom et al., 2005). A small phase I study showed CQ to
imidazole side chains on histidine residues 6, 13 and 14 of Ab. lower CSF tau but caused no change to CSF-Ab1–42 levels (Regland
Alkylation of Ab is proposed to prevent the detrimental binding of et al., 2001). In a subsequent pilot Phase II double-blind, placebo
copper to the metal binding site leading to inhibition of subsequent controlled trial, the oral administration of CQ in moderately severe
Ab-copper oligomerisation and binding to the cell membrane AD patients for 36 weeks slowed the rate of cognitive decline with
(Barnham et al., 2008). Pt(4,7-diphenyl-[1,10] phenanthroline)Cl2 no signs of severe adverse effects (Ritchie et al., 2003). It also
was identified as a compound that bound Ab, changing the peptide lowered plasma Ab1–42. Ensuing phase II/III studies, however, were
conformation and inhibiting aggregation (Barnham et al., 2008). In stalled by the difficulties encountered in preventing di-iodo-8-
addition, this complex inhibited Ab-related neurotoxicity and hydroxy quinoline contamination during the required larger scale
reversed synaptotoxicity caused by Ab-inhibition of long-term chemical synthesis for such trials.
potentiation on mouse hippocampal slices (Barnham et al., 2008). The second generation 8-hydroxy quinoline derivative of CQ,
Future evaluation of the compound’s ability to cross the BBB and PBT2 lacks iodine and is an easier chemical to synthesize. This
exert beneficial effects in animal models for AD need to be compound also has higher solubility and increased BBB perme-
performed. ability than CQ. In the APP/PS1 transgenic AD mouse model, PBT2
An additional theoretical metal complex based therapeutic for decreased soluble interstitial Ab within hours, and improved
AD would target the increase of brain iron in the disease. Such cognitive performance to levels equivalent to or greater than wild-
drugs may also attempt to decrease APP translation and hence Ab type controls within days of treatment (Adlard et al., 2008). In
12 J.A. Duce, A.I. Bush / Progress in Neurobiology 92 (2010) 1–18

addition, there was a significant decrease in insoluble Ab load and Adlard, P.A., Cherny, R.A., Finkelstein, D.I., Gautier, E., Robb, E., Cortes, M., Volitakis,
I., Liu, X., Smith, J.P., Perez, K., Laughton, K., Li, Q.X., Charman, S.A., Nicolazzo, J.A.,
the phosphorylation of tau, as well as a significant increase in Wilkins, S., Deleva, K., Lynch, T., Kok, G., Ritchie, C.W., Tanzi, R.E., Cappai, R.,
synaptophysin levels–suggesting that a number of primary indices Masters, C.L., Barnham, K.J., Bush, A.I., 2008. Rapid restoration of cognition in
used to characterize AD pathology within the brain had been Alzheimer’s transgenic mice with 8-hydroxy quinoline analogs is associated
with decreased interstitial Abeta. Neuron 59, 43–55.
successfully modulated (Adlard et al., 2008). Following a successful Adlard, P.A., Parncutt, J.M., Finkelstein, D.I., Bush, A.I., 2010. Cognitive loss in zinc
phase I study, PBT2 has also completed a randomized, double blind, transporter-3 knock-out mice: a phenocopy for the synaptic and memory
placebo controlled Phase II clinical trial in 78 subjects with early deficits of Alzheimer’s disease? J. Neurosci. 30, 1631–1636.
Ahluwalia, N., Gordon, M.A., Handte, G., Mahlon, M., Li, N.Q., Beard, J.L., Weinstock,
AD over a 12 weeks treatment period. This study demonstrated D., Ross, A.C., 2000. Iron status and stores decline with age in Lewis rats. J. Nutr.
safety and tolerability at 50 mg and 250 mg daily doses over the 130, 2378–2383.
full 12 weeks, and reduced CSF levels of Ab1–42 at the 250 mg dose. Alfrey, A.C., LeGendre, G.R., Kaehny, W.D., 1976. The dialysis encephalopathy
syndrome. Possible aluminum intoxication. N. Engl. J. Med. 294, 184–188.
Improved cognitive performance in patients taking PBT2 (com-
Ali, F.E., Separovic, F., Barrow, C.J., Cherny, R.A., Fraser, F., Bush, A.I., Masters, C.L.,
pared to placebo) was observed when tested on executive function Barnham, K.J., 2005. Methionine regulates copper/hydrogen peroxide oxidation
in the Neuropsychological Test Battery (NTB) at the end of the 12 products of Abeta. J. Pept. Sci. 11, 353–360.
week trial (Faux et al., in press; Lannfelt et al., 2008). These results An, W.L., Bjorkdahl, C., Liu, R., Cowburn, R.F., Winblad, B., Pei, J.J., 2005. Mechanism
of zinc-induced phosphorylation of p70 S6 kinase and glycogen synthase kinase
appear to be the basis for proceeding with further Phase IIb and 3beta in SH-SY5Y neuroblastoma cells. J. Neurochem. 92, 1104–1115.
Phase III trials of what may be a disease-modifying drug based Angeletti, B., Waldron, K.J., Freeman, K.B., Bawagan, H., Hussain, I., Miller, C.C., Lau,
upon the modulation of metals. K.F., Tennant, M.E., Dennison, C., Robinson, N.J., Dingwall, C., 2005. BACE1
cytoplasmic domain interacts with the copper chaperone for superoxide dis-
The exact mechanism of action of CQ and PBT2 is still uncertain. mutase-1 and binds copper. J. Biol. Chem. 280, 17930–17937.
However, in cell culture, CQ-copper complexes have been shown to Armendariz, A.D., Gonzalez, M., Loguinov, A.V., Vulpe, C.D., 2004. Gene expression
enter cells where they markedly inhibit the secretion of Ab profiling in chronic copper overload reveals upregulation of Prnp and App.
Physiol. Genomics 20, 45–54.
thought to occur through its degradation via the upregulation of Atwood, C.S., Moir, R.D., Huang, X., Scarpa, R.C., Bacarra, N.M., Romano, D.M.,
MMP-2 and MMP-3 (Fig. 1C). Similar to metal-BTSC, CQ-copper Hartshorn, M.A., Tanzi, R.E., Bush, A.I., 1998. Dramatic aggregation of Alzheimer
also promotes phosphorylation of GSK-3 and can thus potentiate abeta by Cu(II) is induced by conditions representing physiological acidosis. J.
Biol. Chem. 273, 12817–12826.
activation of JNK and degradation of Ab1–40 (White et al., 2006). Atwood, C.S., Perry, G., Zeng, H., Kato, Y., Jones, W.D., Ling, K.Q., Huang, X., Moir, R.D.,
This leads to a current working hypothesis in AD where CQ or PBT2 Wang, D., Sayre, L.M., Smith, M.A., Chen, S.G., Bush, A.I., 2004. Copper mediates
enters the brain and is attracted to the extracellular pool of metals dityrosine cross-linking of Alzheimer’s amyloid-beta. Biochemistry 43, 560–568.
Atwood, C.S., Scarpa, R.C., Huang, X., Moir, R.D., Jones, W.D., Fairlie, D.P., Tanzi, R.E.,
that are in a dissociable equilibrium within the amyloid plaque.
Bush, A.I., 2000. Characterization of copper interactions with alzheimer amyloid
The compound then binds zinc and copper in the amyloid, possibly beta peptides: identification of an attomolar-affinity copper binding site on
forming a ternary complex with Ab, and the drug-metal complex amyloid beta1-42. J. Neurochem. 75, 1219–1233.
then enters the cell (Deraeve et al., 2006; Rozga et al., 2009). This Barnham, K.J., Haeffner, F., Ciccotosto, G.D., Curtain, C.C., Tew, D., Mavros, C.,
Beyreuther, K., Carrington, D., Masters, C.L., Cherny, R.A., Cappai, R., Bush,
activates MMPs and facilitates the clearance of Ab from the A.I., 2004. Tyrosine gated electron transfer is key to the toxic mechanism of
synapse. In parallel, while bound to the drug, the oxidative Alzheimer’s disease beta-amyloid. FASEB J. 18, 1427–1429.
oligomerisation and the toxic redox activity of Ab oligomers are Barnham, K.J., Kenche, V.B., Ciccotosto, G.D., Smith, D.P., Tew, D.J., Liu, X., Perez, K.,
Cranston, G.A., Johanssen, T.J., Volitakis, I., Bush, A.I., Masters, C.L., White, A.R.,
both blocked. Smith, J.P., Cherny, R.A., Cappai, R., 2008. Platinum-based inhibitors of amyloid-
beta as therapeutic agents for Alzheimer’s disease. Proc. Natl. Acad. Sci. U.S.A.
105, 6813–6818.
5. Conclusion
Barnham, K.J., McKinstry, W.J., Multhaup, G., Galatis, D., Morton, C.J., Curtain, C.C.,
Williamson, N.A., White, A.R., Hinds, M.G., Norton, R.S., Beyreuther, K., Masters,
This review has attempted to illustrate the potentially C.L., Parker, M.W., Cappai, R., 2003. Structure of the Alzheimer’s disease amyloid
pleiotropic role that metals have in AD, with functional relevance precursor protein copper binding domain. A regulator of neuronal copper
homeostasis. J. Biol. Chem. 278, 17401–17407.
to a diverse number of mechanisms known to be disrupted within Bartzokis, G., Beckson, M., Hance, D.B., Marx, P., Foster, J.A., Marder, S.R., 1997. MR
the disease. In turn, biological metals represent potent targets for evaluation of age-related increase of brain iron in young adult and older normal
pharmacological manipulation to limit and perhaps even prevent males. Magn. Reson. Imaging 15, 29–35.
Bartzokis, G., Mintz, J., Sultzer, D., Marx, P., Herzberg, J.S., Phelan, C.K., Marder, S.R.,
the onset and progression of AD. As with any therapeutic target, 1994. In vivo MR evaluation of age-related increases in brain iron. AJNR Am. J.
issues of specificity and selectivity are important factors that Neuroradiol. 15, 1129–1138.
require careful consideration when testing potential drug candi- Bartzokis, G., Tishler, T.A., Lu, P.H., Villablanca, P., Altshuler, L.L., Carter, M., Huang,
D., Edwards, N., Mintz, J., 2007. Brain ferritin iron may influence age- and
dates. While there are promising preclinical and clinical data in the gender-related risks of neurodegeneration. Neurobiol. Aging 28, 414–423.
literature that show some effectiveness in reversing the clinical Baruch-Suchodolsky, R., Fischer, B., 2009. Abeta40, either soluble or aggregated, is a
syndrome in patients and models of the disease, large-scale clinical remarkably potent antioxidant in cell-free oxidative systems. Biochemistry 48,
4354–4370.
trial testing is still to be awaited. With a greater understanding of Basha, M.R., Murali, M., Siddiqi, H.K., Ghosal, K., Siddiqi, O.K., Lashuel, H.A., Ge, Y.W.,
AD pathogenesis and as the multitude of alternative drug trials Lahiri, D.K., Zawia, N.H., 2005a. Lead (Pb) exposure and its effect on APP
currently underway come to completion, Ab: metal -based proteolysis and Abeta aggregation. FASEB J. 19, 2083–2084.
Basha, M.R., Wei, W., Bakheet, S.A., Benitez, N., Siddiqi, H.K., Ge, Y.W., Lahiri, D.K.,
therapeutics remains a promising prospect for the treatment of AD.
Zawia, N.H., 2005b. The fetal basis of amyloidogenesis: exposure to lead and
latent overexpression of amyloid precursor protein and beta-amyloid in the
Acknowledgements aging brain. J. Neurosci. 25, 823–829.
Basun, H., Forssell, L.G., Wetterberg, L., Winblad, B., 1991. Metals and trace elements
in plasma and cerebrospinal fluid in normal aging and Alzheimer’s disease. J.
This work was supported by funds from the National Institute Neural. Transm. Park Dis. Dement. Sect. 3, 231–258.
on Aging (1RO1AG12686), the Australian Research Council, the Baum, L., Chan, I.H., Cheung, S.K., Goggins, W.B., Mok, V., Lam, L., Leung, V., Hui, E.,
Ng, C., Woo, J., Chiu, H.F., Zee, B.C., Cheng, W., Chan, M.H., Szeto, S., Lui, V., Tsoh,
Australian National Health & Medical Research Council, Opera- J., Bush, A.I., Lam, C.W., Kwok, T., 2010. Serum zinc is decreased in Alzheimer’s
tional Infrastructure Support (OIS) from the Victorian State disease and serum arsenic correlates positively with cognitive ability. Biometals
Government and the Alzheimer’s Association. 23, 173–179.
Bayer, T.A., Schafer, S., Simons, A., Kemmling, A., Kamer, T., Tepest, R., Eckert, A.,
Schussel, K., Eikenberg, O., Sturchler-Pierrat, C., Abramowski, D., Staufenbiel, M.,
References Multhaup, G., 2003. Dietary Cu stabilizes brain superoxide dismutase 1 activity
and reduces amyloid Abeta production in APP23 transgenic mice. Proc. Natl.
Abramov, A.Y., Canevari, L., Duchen, M.R., 2003. Changes in intracellular calcium Acad. Sci. U.S.A. 100, 14187–14192.
and glutathione in astrocytes as the primary mechanism of amyloid neurotox- Behl, C., Moosmann, B., 2002. Antioxidant neuroprotection in Alzheimer’s disease as
icity. J. Neurosci. 23, 5088–5095. preventive and therapeutic approach. Free Radic. Biol. Med. 33, 182–191.
J.A. Duce, A.I. Bush / Progress in Neurobiology 92 (2010) 1–18 13

Bellingham, S.A., Ciccotosto, G.D., Needham, B.E., Fodero, L.R., White, A.R., Masters, Treatment with a copper-zinc chelator markedly and rapidly inhibits beta-
C.L., Cappai, R., Camakaris, J., 2004a. Gene knockout of amyloid precursor protein amyloid accumulation in Alzheimer’s disease transgenic mice. Neuron 30, 665–
and amyloid precursor-like protein-2 increases cellular copper levels in primary 676.
mouse cortical neurons and embryonic fibroblasts. J. Neurochem. 91, 423–428. Cherny, R.A., Barnham, K.J., Lynch, T., Volitakis, I., Li, Q.X., McLean, C.A., Multhaup, G.,
Bellingham, S.A., Lahiri, D.K., Maloney, B., La Fontaine, S., Multhaup, G., Camakaris, J., Beyreuther, K., Tanzi, R.E., Masters, C.L., Bush, A.I., 2000. Chelation and interca-
2004b. Copper depletion down-regulates expression of the Alzheimer’s disease lation: complementary properties in a compound for the treatment of Alzhei-
amyloid-beta precursor protein gene. J. Biol. Chem. 279, 20378–20386. mer’s disease. J. Struct. Biol. 130, 209–216.
Benvenisti-Zarom, L., Chen, J., Regan, R.F., 2005. The oxidative neurotoxicity of Cherny, R.A., Legg, J.T., McLean, C.A., Fairlie, D.P., Huang, X., Atwood, C.S., Beyreuther,
clioquinol. Neuropharmacology 49, 687–694. K., Tanzi, R.E., Masters, C.L., Bush, A.I., 1999. Aqueous dissolution of Alzheimer’s
Bertini, I., 2007. Biological Inorganic Chemistry: Structure and Reactivity. University disease Abeta amyloid deposits by biometal depletion. J. Biol. Chem. 274,
Science Books, Sausalito, CA. 23223–23228.
Bertram, L., McQueen, M.B., Mullin, K., Blacker, D., Tanzi, R.E., 2007. Systematic Chowanadisai, W., Kelleher, S.L., Lonnerdal, B., 2005. Zinc deficiency is associated
meta-analyses of Alzheimer disease genetic association studies: the AlzGene with increased brain zinc import and LIV-1 expression and decreased ZnT-1
database. Nat. Genet. 39, 17–23. expression in neonatal rats. J. Nutr. 135, 1002–1007.
Bishop, G.M., Robinson, S.R., Liu, Q., Perry, G., Atwood, C.S., Smith, M.A., 2002. Iron: a Ciccotosto, G.D., Tew, D., Curtain, C.C., Smith, D., Carrington, D., Masters, C.L., Bush,
pathological mediator of Alzheimer disease? Dev. Neurosci. 24, 184–187. A.I., Cherny, R.A., Cappai, R., Barnham, K.J., 2004. Enhanced toxicity and cellular
Bjorkdahl, C., Sjogren, M.J., Winblad, B., Pei, J.J., 2005. Zinc induces neurofilament binding of a modified amyloid beta peptide with a methionine to valine
phosphorylation independent of p70 S6 kinase in N2a cells. Neuroreport 16, substitution. J. Biol. Chem. 279, 42528–42534.
591–595. Ciuculescu, E.D., Mekmouche, Y., Faller, P., 2005. Metal-binding properties of the
Bondiolotti, G., Sala, M., Pollera, C., Gervasoni, M., Puricelli, M., Ponti, W., Bareggi, peptide APP170-188: a model of the ZnII-binding site of amyloid precursor
S.R., 2007. Pharmacokinetics and distribution of clioquinol in golden hamsters. protein (APP). Chemistry 11, 903–909.
J. Pharm. Pharmacol. 59, 387–393. Clark, C.M., Ewbank, D., Lee, V.M.-., Trojanowski, J.Q., 1998. Molecular Pathology of
Bouras, C., Giannakopoulos, P., Good, P.F., Hsu, A., Hof, P.R., Perl, D.P., 1997. A laser Alzheimer’s Disease: Neuronal Cytoskeletal Abnormalities. Butterworth–Hei-
microprobe mass analysis of brain aluminum and iron in dementia pugilistica: nemann, Boston.
comparison with Alzheimer’s disease. Eur. Neurol. 38, 53–58. Clauberg, M., Joshi, J.G., 1993. Regulation of serine protease activity by aluminum:
Brown, D.R., Kozlowski, H., 2004. Biological inorganic and bioinorganic chemistry of implications for Alzheimer disease. Proc. Natl. Acad. Sci. U.S.A. 90, 1009–1012.
neurodegeneration based on prion and Alzheimer diseases. Dalton Trans. 1907– Cole, T.B., Wenzel, H.J., Kafer, K.E., Schwartzkroin, P.A., Palmiter, R.D., 1999. Elimi-
1917. nation of zinc from synaptic vesicles in the intact mouse brain by disruption of
Buckley, C.A., Rouhani, F.N., Kaler, M., Adamik, B., Hawari, F.I., Levine, S.J., 2005. the ZnT3 gene. Proc. Natl. Acad. Sci. U.S.A. 96, 1716–1721.
Amino-terminal TACE prodomain attenuates TNFR2 cleavage independently of Collingwood, J., Dobson, J., 2006. Mapping and characterization of iron compounds
the cysteine switch. Am. J. Physiol. Lung Cell. Mol. Physiol. 288, L1132–1138. in Alzheimer’s tissue. J. Alzheimers Dis. 10, 215–222.
Bunker, V.W., Hinks, L.J., Stansfield, M.F., Lawson, M.S., Clayton, B.E., 1987. Meta- Collingwood, J.F., Mikhaylova, A., Davidson, M., Batich, C., Streit, W.J., Terry, J.,
bolic balance studies for zinc and copper in housebound elderly people and the Dobson, J., 2005. In situ characterization and mapping of iron compounds in
relationship between zinc balance and leukocyte zinc concentrations. Am. J. Alzheimer’s disease tissue. J. Alzheimers Dis. 7, 267–272.
Clin. Nutr. 46, 353–359. Connor, J.R., Lee, S.Y., 2006. HFE mutations and Alzheimer’s disease. J. Alzheimers
Bush, A.I., Multhaup, G., Moir, R.D., Williamson, T.G., Small, D.H., Rumble, B., Dis. 10, 267–276.
Pollwein, P., Beyreuther, K., Masters, C.L., 1993. A novel zinc(II) binding site Connor, J.R., Menzies, S.L., St Martin, S.M., Mufson, E.J., 1990. Cellular distribution of
modulates the function of the beta A4 amyloid protein precursor of Alzheimer’s transferrin, ferritin, and iron in normal and aged human brains. J. Neurosci. Res.
disease. J. Biol. Chem. 268, 16109–16112. 27, 595–611.
Bush, A.I., Pettingell Jr., W.H., de Paradis, M., Tanzi, R.E., Wasco, W., 1994a. The Connor, J.R., Menzies, S.L., St Martin, S.M., Mufson, E.J., 1992a. A histochemical study
amyloid beta-protein precursor and its mammalian homologues. Evidence for a of iron, transferrin, and ferritin in Alzheimer’s diseased brains. J. Neurosci. Res.
zinc-modulated heparin-binding superfamily. J. Biol. Chem. 269, 26618–26621. 31, 75–83.
Bush, A.I., Pettingell Jr., W.H., Paradis, M.D., Tanzi, R.E., 1994b. Modulation of A beta Connor, J.R., Snyder, B.S., Arosio, P., Loeffler, D.A., LeWitt, P., 1995. A quantitative
adhesiveness and secretase site cleavage by zinc. J. Biol. Chem. 269, 12152–12158. analysis of isoferritins in select regions of aged, parkinsonian, and Alzheimer’s
Bush, A.I., Pettingell, W.H., Multhaup, G., d Paradis, M., Vonsattel, J.P., Gusella, J.F., diseased brains. J. Neurochem. 65, 717–724.
Beyreuther, K., Masters, C.L., Tanzi, R.E., 1994c. Rapid induction of Alzheimer A Connor, J.R., Snyder, B.S., Beard, J.L., Fine, R.E., Mufson, E.J., 1992b. Regional distri-
beta amyloid formation by zinc. Science 265, 1464–1467. bution of iron and iron-regulatory proteins in the brain in aging and Alzheimer’s
Buxbaum, J.D., Liu, K.N., Luo, Y., Slack, J.L., Stocking, K.L., Peschon, J.J., Johnson, R.S., disease. J. Neurosci. Res. 31, 327–335.
Castner, B.J., Cerretti, D.P., Black, R.A., 1998. Evidence that tumor necrosis factor Corder, E.H., Saunders, A.M., Strittmatter, W.J., Schmechel, D.E., Gaskell, P.C., Small,
alpha converting enzyme is involved in regulated alpha-secretase cleavage of G.W., Roses, A.D., Haines, J.L., Pericak-Vance, M.A., 1993. Gene dose of apolipo-
the Alzheimer amyloid protein precursor. J. Biol. Chem. 273, 27765–27767. protein E type 4 allele and the risk of Alzheimer’s disease in late onset families.
Caccamo, A., Oddo, S., Sugarman, M.C., Akbari, Y., LaFerla, F.M., 2005. Age- and Science 261, 921–923.
region-dependent alterations in Abeta-degrading enzymes: implications for Cottrell, D.A., Blakely, E.L., Johnson, M.A., Ince, P.G., Turnbull, D.M., 2001. Mitochon-
Abeta-induced disorders. Neurobiol. Aging 26, 645–654. drial enzyme-deficient hippocampal neurons and choroidal cells in AD. Neu-
Cai, H., Wang, Y., McCarthy, D., Wen, H., Borchelt, D.R., Price, D.L., Wong, P.C., 2001. rology 57, 260–264.
BACE1 is the major beta-secretase for generation of Abeta peptides by neurons. Crapper, D.R., Krishnan, S.S., Dalton, A.J., 1973. Brain aluminum distribution in
Nat. Neurosci. 4, 233–234. Alzheimer’s disease and experimental neurofibrillary degeneration. Science
Canfield, R.L., Henderson Jr., C.R., Cory-Slechta, D.A., Cox, C., Jusko, T.A., Lanphear, 180, 511–513.
B.P., 2003. Intellectual impairment in children with blood lead concentrations Crapper McLachlan, D.R., Dalton, A.J., Kruck, T.P., Bell, M.Y., Smith, W.L., Kalow, W.,
below 10 microg per deciliter. N. Engl. J. Med. 348, 1517–1526. Andrews, D.F., 1991. Intramuscular desferrioxamine in patients with Alzhei-
Cardoso, S.M., Proenca, M.T., Santos, S., Santana, I., Oliveira, C.R., 2004. Cyto- mer’s disease. Lancet 337, 1304–1308.
chrome c oxidase is decreased in Alzheimer’s disease platelets. Neurobiol. Cross, J.B., Duca, J.S., Kaminski, J.J., Madison, V.S., 2002. The active site of a zinc-
Aging 25, 105–110. dependent metalloproteinase influences the computed pK(a) of ligands coor-
Carri, M.T., Ferri, A., Cozzolino, M., Calabrese, L., Rotilio, G., 2003. Neurodegeneration dinated to the catalytic zinc ion. J. Am. Chem. Soc. 124, 11004–11007.
in amyotrophic lateral sclerosis: the role of oxidative stress and altered ho- Crouch, P.J., Hung, L.W., Adlard, P.A., Cortes, M., Lal, V., Filiz, G., Perez, K.A., Nurjono,
meostasis of metals. Brain Res. Bull. 61, 365–374. M., Caragounis, A., Du, T., Laughton, K., Volitakis, I., Bush, A.I., Li, Q.X., Masters,
Carson, J.A., Turner, A.J., 2002. Beta-amyloid catabolism: roles for neprilysin (NEP) C.L., Cappai, R., Cherny, R.A., Donnelly, P.S., White, A.R., Barnham, K.J., 2009a.
and other metallopeptidases? J. Neurochem. 81, 1–8. Increasing Cu bioavailability inhibits Abeta oligomers and tau phosphorylation.
Castellani, R.J., Moreira, P.I., Liu, G., Dobson, J., Perry, G., Smith, M.A., Zhu, X., 2007. Proc. Natl. Acad. Sci. U.S.A. 106, 381–386.
Iron: the Redox-active center of oxidative stress in Alzheimer disease. Neuro- Crouch, P.J., Tew, D.J., Du, T., Nguyen, D.N., Caragounis, A., Filiz, G., Blake, R.E.,
chem. Res. 32, 1640–1645. Trounce, I.A., Soon, C.P., Laughton, K., Perez, K.A., Li, Q.X., Cherny, R.A., Masters,
Cater, M.A., McInnes, K.T., Li, Q.X., Volitakis, I., La Fontaine, S., Mercer, J.F., Bush, A.I., C.L., Barnham, K.J., White, A.R., 2009b. Restored degradation of the Alzheimer’s
2008. Intracellular copper deficiency increases amyloid-beta secretion by di- amyloid-beta peptide by targeting amyloid formation. J. Neurochem. 108,
verse mechanisms. Biochem. J. 412, 141–152. 1198–1207.
Caughman, S.W., Hentze, M.W., Rouault, T.A., Harford, J.B., Klausner, R.D., 1988. The Cuajungco, M.P., Faget, K.Y., Huang, X., Tanzi, R.E., Bush, A.I., 2000. Metal chelation as
iron-responsive element is the single element responsible for iron-dependent a potential therapy for Alzheimer’s disease. Ann. N. Y. Acad. Sci. 920, 292–304.
translational regulation of ferritin biosynthesis. Evidence for function as the Culotta, V.C., Yang, M., O’Halloran, T.V., 2006. Activation of superoxide dismutases:
binding site for a translational repressor. J. Biol. Chem. 263, 19048–19052. putting the metal to the pedal. Biochim. Biophys. Acta 1763, 747–758.
Chen, K., Kazachkov, M., Yu, P.H., 2007. Effect of aldehydes derived from oxidative Curtain, C.C., Ali, F., Volitakis, I., Cherny, R.A., Norton, R.S., Beyreuther, K., Barrow,
deamination and oxidative stress on beta-amyloid aggregation; pathological C.J., Masters, C.L., Bush, A.I., Barnham, K.J., 2001. Alzheimer’s disease amyloid-
implications to Alzheimer’s disease. J. Neural. Transm. 114, 835–839. beta binds copper and zinc to generate an allosterically ordered membrane-
Cheng, Y., Feng, Z., Zhang, Q.Z., Zhang, J.T., 2006. Beneficial effects of melatonin in penetrating structure containing superoxide dismutase-like subunits. J. Biol.
experimental models of Alzheimer disease. Acta Pharmacol. Sin. 27, 129–139. Chem. 276, 20466–20473.
Cherny, R.A., Atwood, C.S., Xilinas, M.E., Gray, D.N., Jones, W.D., McLean, C.A., Curtain, C.C., Ali, F.E., Smith, D.G., Bush, A.I., Masters, C.L., Barnham, K.J., 2003. Metal
Barnham, K.J., Volitakis, I., Fraser, F.W., Kim, Y., Huang, X., Goldstein, L.E., Moir, ions, pH, and cholesterol regulate the interactions of Alzheimer’s disease
R.D., Lim, J.T., Beyreuther, K., Zheng, H., Tanzi, R.E., Masters, C.L., Bush, A.I., 2001. amyloid-beta peptide with membrane lipid. J. Biol. Chem. 278, 2977–2982.
14 J.A. Duce, A.I. Bush / Progress in Neurobiology 92 (2010) 1–18

da Silva, G.F., Lykourinou, V., Angerhofer, A., Ming, L.J., 2009. Methionine does not Frausto da Silva, J.J.R., Williams, R.J.P., 2001. The Biological Chemistry of the
reduce Cu(II)-beta-amyloid!–rectification of the roles of methionine-35 and Elements. Oxford University Press, Oxford.
reducing agents in metal-centered oxidation chemistry of Cu(II)-beta-amyloid Frederickson, C.J., 1989. Neurobiology of zinc and zinc-containing neurons. Int. Rev.
Biochim. Biophys. Acta 1792, 49–55. Neurobiol. 31, 145–238.
Danielsson, J., Pierattelli, R., Banci, L., Graslund, A., 2007. High-resolution NMR Frederickson, C.J., Koh, J.Y., Bush, A.I., 2005. The neurobiology of zinc in health and
studies of the zinc-binding site of the Alzheimer’s amyloid beta-peptide. FEBS J. disease. Nat. Rev. Neurosci. 6, 449–462.
274, 46–59. Friedlich, A.L., Lee, J.Y., van Groen, T., Cherny, R.A., Volitakis, I., Cole, T.B., Palmiter,
Danzeisen, R., Achsel, T., Bederke, U., Cozzolino, M., Crosio, C., Ferri, A., Frenzel, M., R.D., Koh, J.Y., Bush, A.I., 2004. Neuronal zinc exchange with the blood vessel
Gralla, E.B., Huber, L., Ludolph, A., Nencini, M., Rotilio, G., Valentine, J.S., Carri, wall promotes cerebral amyloid angiopathy in an animal model of Alzheimer’s
M.T., 2006. Superoxide dismutase 1 modulates expression of transferrin recep- disease. J. Neurosci. 24, 3453–3459.
tor. J. Biol. Inorg. Chem. 11, 489–498. Fujibayashi, Y., Taniuchi, H., Yonekura, Y., Ohtani, H., Konishi, J., Yokoyama, A., 1997.
Davis, C.D., Milne, D.B., Nielsen, F.H., 2000. Changes in dietary zinc and copper affect Copper-62-ATSM: a new hypoxia imaging agent with high membrane perme-
zinc-status indicators of postmenopausal women, notably, extracellular super- ability and low redox potential. J. Nucl. Med. 38, 1155–1160.
oxide dismutase and amyloid precursor proteins. Am. J. Clin. Nutr. 71, 781–788. Gaggelli, E., Grzonka, Z., Kozlowski, H., Migliorini, C., Molteni, E., Valensin, D.,
De Deyn, P.P., Hiramatsu, M., Borggreve, F., Goeman, J., D’Hooge, R., Saerens, J., Mori, Valensin, G., 2008. Structural features of the Cu(II) complex with the rat
A., 1998. Superoxide dismutase activity in cerebrospinal fluid of patients with Abeta(1–28) fragment. Chem. Commun. (Camb.) 341–343.
dementia and some other neurological disorders. Alzheimer Dis. Assoc. Disord. Gandy, S., 2005. The role of cerebral amyloid beta accumulation in common forms of
12, 26–32. Alzheimer disease. J. Clin. Invest. 115, 1121–1129.
De Strooper, B., 2003. Aph-1, Pen-2, and Nicastrin with Presenilin generate an active Garzon-Rodriguez, W., Yatsimirsky, A.K., Glabe, C.G., 1999. Binding of Zn(II), Cu(II),
gamma-Secretase complex. Neuron 38, 9–12. and Fe(II) ions to Alzheimer’s A beta peptide studied by fluorescence. Bioorg.
Dearling, J.L., Lewis, J.S., Mullen, G.E., Welch, M.J., Blower, P.J., 2002. Copper Med. Chem. Lett. 9, 2243–2248.
bis(thiosemicarbazone) complexes as hypoxia imaging agents: structure-activ- Ghribi, O., Golovko, M.Y., Larsen, B., Schrag, M., Murphy, E.J., 2006. Deposition of iron
ity relationships. J. Biol. Inorg. Chem. 7, 249–259. and beta-amyloid plaques is associated with cortical cellular damage in rabbits
Defeudis, F.V., 2002. Bilobalide and neuroprotection. Pharmacol. Res. 46, 565–568. fed with long-term cholesterol-enriched diets. J. Neurochem. 99, 438–449.
Deibel, M.A., Ehmann, W.D., Markesbery, W.R., 1996. Copper, iron, and zinc imbal- Giese, A., Buchholz, M., Herms, J., Kretzschmar, H.A., 2005. Mouse brain synapto-
ances in severely degenerated brain regions in Alzheimer’s disease: possible somes accumulate copper-67 efficiently by two distinct processes independent
relation to oxidative stress. J. Neurol. Sci. 143, 137–142. of cellular prion protein. J. Mol. Neurosci. 27, 347–354.
Deraeve, C., Pitie, M., Meunier, B., 2006. Influence of chelators and iron ions on the Gomis-Ruth, F.X., 2003. Structural aspects of the metzincin clan of metalloendo-
production and degradation of H2O2 by beta-amyloid-copper complexes. J. peptidases. Mol. Biotechnol. 24, 157–202.
Inorg. Biochem. 100, 2117–2126. Gonzales, P.E., Solomon, A., Miller, A.B., Leesnitzer, M.A., Sagi, I., Milla, M.E., 2004.
Deshpande, A., Kawai, H., Metherate, R., Glabe, C.G., Busciglio, J., 2009. A role for Inhibition of the tumor necrosis factor-alpha-converting enzyme by its pro
synaptic zinc in activity-dependent Abeta oligomer formation and accumula- domain. J. Biol. Chem. 279, 31638–31645.
tion at excitatory synapses. J. Neurosci. 29, 4004–4015. Goodman, L., 1953. Alzheimer’s disease; a clinico-pathologic analysis of twenty-
Di Vaira, M., Bazzicalupi, C., Orioli, P., Messori, L., Bruni, B., Zatta, P., 2004. Clioquinol, three cases with a theory on pathogenesis. J. Nerv. Ment. Dis. 118, 97–130.
a drug for Alzheimer’s disease specifically interfering with brain metal metab- Green, M.A., Klippenstein, D.L., Tennison, J.R., 1988. Copper(II) bis(thiosemicarba-
olism: structural characterization of its zinc(II) and copper(II) complexes. Inorg. zone) complexes as potential tracers for evaluation of cerebral and myocardial
Chem. 43, 3795–3797. blood flow with PET. J. Nucl. Med. 29, 1549–1557.
Dikalov, S.I., Vitek, M.P., Mason, R.P., 2004. Cupric-amyloid beta peptide complex Grundke-Iqbal, I., Fleming, J., Tung, Y.C., Lassmann, H., Iqbal, K., Joshi, J.G., 1990.
stimulates oxidation of ascorbate and generation of hydroxyl radical. Free Ferritin is a component of the neuritic (senile) plaque in Alzheimer dementia.
Radic. Biol. Med. 36, 340–347. Acta Neuropathol. 81, 105–110.
Dong, J., Atwood, C.S., Anderson, V.E., Siedlak, S.L., Smith, M.A., Perry, G., Carey, P.R., Guengerich, F.P., 2009. Thematic minireview series: metals in biology. J. Biol. Chem.
2003. Metal binding and oxidation of amyloid-beta within isolated senile 284, 18557.
plaque cores: Raman microscopic evidence. Biochemistry 42, 2768–2773. Gupta, V.B., Anitha, S., Hegde, M.L., Zecca, L., Garruto, R.M., Ravid, R., Shankar, S.K.,
Donnelly, P.S., Caragounis, A., Du, T., Laughton, K.M., Volitakis, I., Cherny, R.A., Stein, R., Shanmugavelu, P., Jagannatha Rao, K.S., 2005. Aluminium in Alzhei-
Sharples, R.A., Hill, A.F., Li, Q.X., Masters, C.L., Barnham, K.J., White, A.R., 2008. mer’s disease: are we still at a crossroad? Cell Mol. Life Sci. 62, 143–158.
Selective intracellular release of copper and zinc ions from bis(thiosemicarba- Ha, C., Ryu, J., Park, C.B., 2007. Metal ions differentially influence the aggregation
zonato) complexes reduces levels of Alzheimer disease amyloid-beta peptide. J. and deposition of Alzheimer’s beta-amyloid on a solid template. Biochemistry
Biol. Chem. 283, 4568–4577. 46, 6118–6125.
Dorlet, P., Gambarelli, S., Faller, P., Hureau, C., 2009. Pulse EPR spectroscopy reveals Haeffner, F., Smith, D.G., Barnham, K.J., Bush, A.I., 2005. Model studies of cholesterol
the coordination sphere of copper(II) ions in the 1–16 amyloid-beta peptide: a and ascorbate oxidation by copper complexes: relevance to Alzheimer’s disease
key role of the first two N-terminus residues. Angew. Chem. Int. Ed. Engl. 48, beta-amyloid metallochemistry. J. Inorg. Biochem. 99, 2403–2422.
9273–9276. Hallgren, B., Sourander, P., 1958. The effect of age on the non-haemin iron in the
Drew, S.C., Masters, C.L., Barnham, K.J., 2009a. Alanine-2 carbonyl is an human brain. J. Neurochem. 3, 41–51.
oxygen ligand in Cu2+ coordination of Alzheimer’s disease amyloid-beta Hardy, P.A., Gash, D., Yokel, R., Andersen, A., Ai, Y., Zhang, Z., 2005. Correlation of R2
peptide–relevance to N-terminally truncated forms. J. Am. Chem. Soc. 131, with total iron concentration in the brains of rhesus monkeys. J. Magn. Reson.
8760–8761. Imaging 21, 118–127.
Drew, S.C., Noble, C.J., Masters, C.L., Hanson, G.R., Barnham, K.J., 2009b. Pleomorphic Harris, F.M., Brecht, W.J., Xu, Q., Mahley, R.W., Huang, Y., 2004. Increased tau
copper coordination by Alzheimer’s disease amyloid-beta peptide. J. Am. Chem. phosphorylation in apolipoprotein E4 transgenic mice is associated with acti-
Soc. 131, 1195–1207. vation of extracellular signal-regulated kinase: modulation by zinc. J. Biol.
Ebadi, M., Elsayed, M.A., Aly, M.H., 1994. The importance of zinc and metallothio- Chem. 279, 44795–44801.
nein in brain. Biol. Signals 3, 123–126. Hart, P.J., Nersissian, A.M., Herrmann, R.G., Nalbandyan, R.M., Valentine, J.S., Eisen-
Egana, J.T., Zambrano, C., Nunez, M.T., Gonzalez-Billault, C., Maccioni, R.B., 2003. berg, D., 1996. A missing link in cupredoxins: crystal structure of cucumber
Iron-induced oxidative stress modify tau phosphorylation patterns in hippo- stellacyanin at 1.6 A resolution. Protein Sci. 5, 2175–2183.
campal cell cultures. Biometals 16, 215–223. Hayakawa, M., Miyashita, H., Sakamoto, I., Kitagawa, M., Tanaka, H., Yasuda, H.,
Ehehalt, R., Keller, P., Haass, C., Thiele, C., Simons, K., 2003. Amyloidogenic proces- Karin, M., Kikugawa, K., 2003. Evidence that reactive oxygen species do not
sing of the Alzheimer beta-amyloid precursor protein depends on lipid rafts. J. mediate NF-kappaB activation. EMBO J. 22, 3356–3366.
Cell. Biol. 160, 113–123. Hensley, K., Maidt, M.L., Yu, Z., Sang, H., Markesbery, W.R., Floyd, R.A., 1998.
Ekmekcioglu, C., 2001. The role of trace elements for the health of elderly indivi- Electrochemical analysis of protein nitrotyrosine and dityrosine in the Alzhei-
duals. Nahrung 45, 309–316. mer brain indicates region-specific accumulation. J. Neurosci. 18, 8126–8132.
El Meskini, R., Crabtree, K.L., Cline, L.B., Mains, R.E., Eipper, B.A., Ronnett, G.V., 2007. Hesse, L., Beher, D., Masters, C.L., Multhaup, G., 1994. The beta A4 amyloid precursor
ATP7A (Menkes protein) functions in axonal targeting and synaptogenesis. Mol. protein binding to copper. FEBS Lett. 349, 109–116.
Cell. Neurosci. 34, 409–421. Himes, R.A., Park, G.Y., Siluvai, G.S., Blackburn, N.J., Karlin, K.D., 2008. Structural
Faller, P., Hureau, C., 2009. Bioinorganic chemistry of copper and zinc ions coordi- studies of copper(I) complexes of amyloid-beta peptide fragments: formation of
nated to amyloid-beta peptide. Dalton Trans. 1080–1094. two-coordinate bis(histidine) complexes. Angew. Chem. Int. Ed. Engl. 47, 9084–
Farris, W., Mansourian, S., Chang, Y., Lindsley, L., Eckman, E.A., Frosch, M.P., Eckman, 9087.
C.B., Tanzi, R.E., Selkoe, D.J., Guenette, S., 2003. Insulin-degrading enzyme Hoke, D.E., Tan, J.L., Ilaya, N.T., Culvenor, J.G., Smith, S.J., White, A.R., Masters, C.L.,
regulates the levels of insulin, amyloid beta-protein, and the beta-amyloid Evin, G.M., 2005. In vitro gamma-secretase cleavage of the Alzheimer’s amyloid
precursor protein intracellular domain in vivo. Proc. Natl. Acad. Sci. U.S.A. 100, precursor protein correlates to a subset of presenilin complexes and is inhibited
4162–4167. by zinc. FEBS J. 272, 5544–5557.
Feng, Z., Chang, Y., Cheng, Y., Zhang, B.L., Qu, Z.W., Qin, C., Zhang, J.T., 2004. Hou, L., Zagorski, M.G., 2006. NMR reveals anomalous copper(II) binding to the
Melatonin alleviates behavioral deficits associated with apoptosis and cholin- amyloid Abeta peptide of Alzheimer’s disease. J. Am. Chem. Soc. 128, 9260–
ergic system dysfunction in the APP 695 transgenic mouse model of Alzheimer’s 9261.
disease. J. Pineal. Res. 37, 129–136. Huang, X., Atwood, C.S., Hartshorn, M.A., Multhaup, G., Goldstein, L.E., Scarpa, R.C.,
Flagstad, T., 1977. Intestinal absorption of 65 Zinc in A46 (Adema disease) after Cuajungco, M.P., Gray, D.N., Lim, J., Moir, R.D., Tanzi, R.E., Bush, A.I., 1999a. The A
treatment with oxychinolines. Nord. Vet. Med. 29, 96–100. beta peptide of Alzheimer’s disease directly produces hydrogen peroxide
Franz, K.J., 2008. Copper shares a piece of the pi. Nat. Chem. Biol. 4, 85–86. through metal ion reduction. Biochemistry 38, 7609–7616.
J.A. Duce, A.I. Bush / Progress in Neurobiology 92 (2010) 1–18 15

Huang, X., Atwood, C.S., Moir, R.D., Hartshorn, M.A., Tanzi, R.E., Bush, A.I., 2004. prevents plaque formation, secondary pathology, and premature death. Neuron
Trace metal contamination initiates the apparent auto-aggregation, amyloid- 40, 1087–1093.
osis, and oligomerization of Alzheimer’s Abeta peptides. J. Biol. Inorg. Chem. 9, LeVine, S.M., 1997. Iron deposits in multiple sclerosis and Alzheimer’s disease
954–960. brains. Brain Res. 760, 298–303.
Huang, X., Atwood, C.S., Moir, R.D., Hartshorn, M.A., Vonsattel, J.P., Tanzi, R.E., Bush, Liewendahl, K., Lamberg, B.A., 1967. Metabolism of 125-iodochloroxyquinoline in
A.I., 1997. Zinc-induced Alzheimer’s Abeta1-40 aggregation is mediated by man. I. Absorption, binding and excretion. Nucl. Med. (Stuttg.) 6, 20–31.
conformational factors. J. Biol. Chem. 272, 26464–26470. Lind, S.E., McDonagh, J.R., Smith, C.J., 1993. Oxidative inactivation of plasmin and
Huang, X., Cuajungco, M.P., Atwood, C.S., Hartshorn, M.A., Tyndall, J.D., Hanson, G.R., other serine proteases by copper and ascorbate. Blood 82, 1522–1531.
Stokes, K.C., Leopold, M., Multhaup, G., Goldstein, L.E., Scarpa, R.C., Saunders, Ling, Y., Morgan, K., Kalsheker, N., 2003. Amyloid precursor protein (APP) and the
A.J., Lim, J., Moir, R.D., Glabe, C., Bowden, E.F., Masters, C.L., Fairlie, D.P., Tanzi, biology of proteolytic processing: relevance to Alzheimer’s disease. Int. J.
R.E., Bush, A.I., 1999b. Cu(II) potentiation of alzheimer abeta neurotoxicity. Biochem. Cell Biol. 35, 1505–1535.
Correlation with cell-free hydrogen peroxide production and metal reduction. J. Linkous, D.H., Adlard, P.A., Wanschura, P.B., Conko, K.M., Flinn, J.M., 2009. The
Biol. Chem. 274, 37111–37116. effects of enhanced zinc on spatial memory and plaque formation in transgenic
Hung, Y.H., Robb, E.L., Volitakis, I., Ho, M., Evin, G., Li, Q.X., Culvenor, J.G., Masters, mice. J. Alzheimers Dis..
C.L., Cherny, R.A., Bush, A.I., 2009. Paradoxical condensation of copper with Linkous, D.H., Flinn, J.M., Koh, J.Y., Lanzirotti, A., Bertsch, P.M., Jones, B.F., Giblin, L.J.,
elevated beta-amyloid in lipid rafts under cellular copper deficiency conditions: Frederickson, C.J., 2008. Evidence that the ZNT3 protein controls the total
implications for Alzheimer disease. J. Biol. Chem. 284, 21899–21907. amount of elemental zinc in synaptic vesicles. J. Histochem. Cytochem. 56,
Hureau, C., Balland, V., Coppel, Y., Solari, P.L., Fonda, E., Faller, P., 2009. Importance of 3–6.
dynamical processes in the coordination chemistry and redox conversion of Liu, S.F., Ye, X., Malik, A.B., 1999. Inhibition of NF-kappaB activation by pyrrolidine
copper amyloid-beta complexes. J. Biol. Inorg. Chem. 14, 995–1000. dithiocarbamate prevents In vivo expression of proinflammatory genes. Circu-
Hureau, C., Faller, P., 2009. Abeta-mediated ROS production by Cu ions: structural lation 100, 1330–1337.
insights, mechanisms and relevance to Alzheimer’s disease. Biochimie 91, Lloret, A., Badia, M.C., Mora, N.J., Pallardo, F.V., Alonso, M.D., Vina, J., 2009. Vitamin E
1212–1217. paradox in Alzheimer’s disease: it does not prevent loss of cognition and may
Iskra, M., Patelski, J., Majewski, W., 1993. Concentrations of calcium, magnesium, even be detrimental. J. Alzheimers Dis. 17, 143–149.
zinc and copper in relation to free fatty acids and cholesterol in serum of Loeffler, D.A., LeWitt, P.A., Juneau, P.L., Sima, A.A., Nguyen, H.U., DeMaggio, A.J.,
atherosclerotic men. J. Trace Elem. Electrolytes Health Dis. 7, 185–188. Brickman, C.M., Brewer, G.J., Dick, R.D., Troyer, M.D., Kanaley, L., 1996. Increased
Iwata, N., Tsubuki, S., Takaki, Y., Shirotani, K., Lu, B., Gerard, N.P., Gerard, C., Hama, E., regional brain concentrations of ceruloplasmin in neurodegenerative disorders.
Lee, H.J., Saido, T.C., 2001. Metabolic regulation of brain Abeta by neprilysin. Brain Res. 738, 265–274.
Science 292, 1550–1552. Lovell, M.A., Robertson, J.D., Teesdale, W.J., Campbell, J.L., Markesbery, W.R., 1998.
Iwatsubo, T., Odaka, A., Suzuki, N., Mizusawa, H., Nukina, N., Ihara, Y., 1994. Copper, iron and zinc in Alzheimer’s disease senile plaques. J. Neurol. Sci. 158,
Visualization of A beta 42(43) and A beta 40 in senile plaques with end-specific 47–52.
A beta monoclonals: evidence that an initially deposited species is A beta Luo, Y., Bolon, B., Kahn, S., Bennett, B.D., Babu-Khan, S., Denis, P., Fan, W., Kha, H.,
42(43). Neuron 13, 45–53. Zhang, J., Gong, Y., Martin, L., Louis, J.C., Yan, Q., Richards, W.G., Citron, M.,
Jefferies, W.A., Dickstein, D.L., Ujiie, M., 2001. Assessing p97 as an Alzheimer’s Vassar, R., 2001. Mice deficient in BACE1, the Alzheimer’s beta-secretase, have
disease serum biomarker. J. Alzheimers Dis. 3, 339–344. normal phenotype and abolished beta-amyloid generation. Nat. Neurosci. 4,
Jiang, D., Men, L., Wang, J., Zhang, Y., Chickenyen, S., Wang, Y., Zhou, F., 2007. Redox 231–232.
reactions of copper complexes formed with different beta-amyloid peptides Lyubartseva, G., Smith, J.L., Markesbery, W.R., Lovell, M.A., 2009. Alterations of zinc
and their neuropathological [correction of neuropathalogical] relevance. Bio- transporter proteins ZnT-1, ZnT-4 and ZnT-6 in preclinical Alzheimer’s disease
chemistry 46, 9270–9282. brain. Brain Pathol.
Kanemitsu, H., Tomiyama, T., Mori, H., 2003. Human neprilysin is capable of Ma, Q., Li, Y., Du, J., Liu, H., Kanazawa, K., Nemoto, T., Nakanishi, H., Zhao, Y., 2006.
degrading amyloid beta peptide not only in the monomeric form but also Copper binding properties of a tau peptide associated with Alzheimer’s disease
the pathological oligomeric form. Neurosci. Lett. 350, 113–116. studied by CD, NMR, and MALDI-TOF MS. Peptides 27, 841–849.
Karr, J.W., Akintoye, H., Kaupp, L.J., Szalai, V.A., 2005. N-Terminal deletions modify Ma, Q.F., Li, Y.M., Du, J.T., Kanazawa, K., Nemoto, T., Nakanishi, H., Zhao, Y.F., 2005.
the Cu2+ binding site in amyloid-beta. Biochemistry 44, 5478–5487. Binding of copper (II) ion to an Alzheimer’s tau peptide as revealed by MALDI-
Karr, J.W., Szalai, V.A., 2007. Role of aspartate-1 in Cu(II) binding to the amyloid- TOF MS, CD, and NMR. Biopolymers 79, 74–85.
beta peptide of Alzheimer’s disease. J. Am. Chem. Soc. 129, 3796–3797. Madaric, A., Ginter, E., Kadrabova, J., 1994. Serum copper, zinc and copper/zinc ratio
Karr, J.W., Szalai, V.A., 2008. Cu(II) binding to monomeric, oligomeric, and fibrillar in males: influence of aging. Physiol. Res. 43, 107–111.
forms of the Alzheimer’s disease amyloid-beta peptide. Biochemistry 47, 5006– Maeda, J., Ji, B., Irie, T., Tomiyama, T., Maruyama, M., Okauchi, T., Staufenbiel, M.,
5016. Iwata, N., Ono, M., Saido, T.C., Suzuki, K., Mori, H., Higuchi, M., Suhara, T., 2007.
Kawahara, M., 2005. Effects of aluminum on the nervous system and its possible Longitudinal, quantitative assessment of amyloid, neuroinflammation, and
link with neurodegenerative diseases. J. Alzheimers Dis. 8, 171–182 discussion anti-amyloid treatment in a living mouse model of Alzheimer’s disease enabled
209-115. by positron emission tomography. J. Neurosci. 27, 10957–10968.
Kawahara, M., Muramoto, K., Kobayashi, K., Mori, H., Kuroda, Y., 1994. Aluminum Mainous 3rd, A.G., Eschenbach, S.L., Wells, B.J., Everett, C.J., Gill, J.M., 2005. Choles-
promotes the aggregation of Alzheimer’s amyloid beta-protein in vitro. Bio- terol, transferrin saturation, and the development of dementia and Alzheimer’s
chem. Biophys. Res. Commun. 198, 531–535. disease: results from an 18-year population-based cohort. Fam. Med. 37, 36–42.
Ke, Y., Qian, Z.M., 2007. Brain iron metabolism: neurobiology and neurochemistry. Malm, T.M., Iivonen, H., Goldsteins, G., Keksa-Goldsteine, V., Ahtoniemi, T., Kanni-
Prog. Neurobiol. 83, 149–173. nen, K., Salminen, A., Auriola, S., Van Groen, T., Tanila, H., Koistinaho, J., 2007.
Keberle, H., 1964. The Biochemistry of Desferrioxamine and Its Relation to Iron Pyrrolidine dithiocarbamate activates Akt and improves spatial learning in APP/
Metabolism. Ann. N. Y. Acad. Sci. 119, 758–768. PS1 mice without affecting beta-amyloid burden. J. Neurosci. 27, 3712–3721.
Kikinis, Z., Eisenstein, R.S., Bettany, A.J., Munro, H.N., 1995. Role of RNA secondary Markesbery, W.R., Lovell, M.A., 2007. Damage to lipids, proteins, DNA, and RNA in
structure of the iron-responsive element in translational regulation of ferritin mild cognitive impairment. Arch. Neurol. 64, 954–956.
synthesis. Nucleic Acids Res. 23, 4190–4195. Martin, W.R., Ye, F.Q., Allen, P.S., 1998. Increasing striatal iron content associated
King, C.E., Adlard, P.A., Dickson, T.C., Vickers, J.C., 2000. Neuronal response to with normal aging. Mov. Disord. 13, 281–286.
physical injury and its relationship to the pathology of Alzheimer’s disease. Martinez Lista, E., Sole, J., Arola, L., Mas, A., 1993. Changes in plasma copper and zinc
Clin. Exp. Pharmacol. Physiol. 27, 548–552. during rat development. Biol. Neonate 64, 47–52.
Kotaki, H., Yamamura, Y., Tanimura, Y., Saitoh, Y., Nakagawa, F., Tamura, Z., 1983. Masters, C.L., Beyreuther, K., 2006. Alzheimer’s centennial legacy: prospects for
Determination of chinoform and its metabolites in plasma by gas chromatog- rational therapeutic intervention targeting the Abeta amyloid pathway. Brain
raphy–mass spectrometry. Chem. Pharm. Bull. (Tokyo) 31, 299–304. 129, 2823–2839.
Lammich, S., Kojro, E., Postina, R., Gilbert, S., Pfeiffer, R., Jasionowski, M., Haass, C., Matsubara, E., Bryant-Thomas, T., Pacheco Quinto, J., Henry, T.L., Poeggeler, B.,
Fahrenholz, F., 1999. Constitutive and regulated alpha-secretase cleavage of Herbert, D., Cruz-Sanchez, F., Chyan, Y.J., Smith, M.A., Perry, G., Shoji, M., Abe, K.,
Alzheimer’s amyloid precursor protein by a disintegrin metalloprotease. Proc. Leone, A., Grundke-Ikbal, I., Wilson, G.L., Ghiso, J., Williams, C., Refolo, L.M.,
Natl. Acad. Sci. U.S.A. 96, 3922–3927. Pappolla, M.A., Chain, D.G., Neria, E., 2003. Melatonin increases survival and
Lee, J.Y., Cole, T.B., Palmiter, R.D., Suh, S.W., Koh, J.Y., 2002. Contribution by synaptic inhibits oxidative and amyloid pathology in a transgenic model of Alzheimer’s
zinc to the gender-disparate plaque formation in human Swedish mutant APP disease. J. Neurochem. 85, 1101–1108.
transgenic mice. Proc. Natl. Acad. Sci. U.S.A. 99, 7705–7710. Mattson, M.P., 2004. Pathways towards and away from Alzheimer’s disease. Nature
Lee, J.Y., Friedman, J.E., Angel, I., Kozak, A., Koh, J.Y., 2004a. The lipophilic metal 430, 631–639.
chelator DP-109 reduces amyloid pathology in brains of human beta-amyloid Maurer, I., Zierz, S., Moller, H.J., 2000. A selective defect of cytochrome c oxidase is
precursor protein transgenic mice. Neurobiol. Aging 25, 1315–1321. present in brain of Alzheimer disease patients. Neurobiol. Aging 21, 455–462.
Lee, J.Y., Kim, J.H., Hong, S.H., Cherny, R.A., Bush, A.I., Palmiter, R.D., Koh, J.Y., 2004b. Maynard, C.J., Cappai, R., Volitakis, I., Cherny, R.A., Masters, C.L., Li, Q.X., Bush, A.I.,
Estrogen decreases zinc transporter 3 expression and synaptic vesicle zinc 2006. Gender and genetic background effects on brain metal levels in APP
levels in mouse brain. J. Biol. Chem. 279, 8602–8607. transgenic and normal mice: implications for Alzheimer beta-amyloid pathol-
Lehmann, D.J., Worwood, M., Ellis, R., Wimhurst, V.L., Merryweather-Clarke, A.T., ogy. J. Inorg. Biochem. 100, 952–962.
Warden, D.R., Smith, A.D., Robson, K.J., 2006. Iron genes, iron load and risk of Maynard, C.J., Cappai, R., Volitakis, I., Cherny, R.A., White, A.R., Beyreuther, K.,
Alzheimer’s disease. J. Med. Genet. 43, e52. Masters, C.L., Bush, A.I., Li, Q.X., 2002. Overexpression of Alzheimer’s disease
Leissring, M.A., Farris, W., Chang, A.Y., Walsh, D.M., Wu, X., Sun, X., Frosch, M.P., amyloid-beta opposes the age-dependent elevations of brain copper and iron. J.
Selkoe, D.J., 2003. Enhanced proteolysis of beta-amyloid in APP transgenic mice Biol. Chem. 277, 44670–44676.
16 J.A. Duce, A.I. Bush / Progress in Neurobiology 92 (2010) 1–18

McColl, G., Roberts, B.R., Gunn, A.P., Perez, K.A., Tew, D.J., Masters, C.L., Barnham, K.J., Opazo, C., Huang, X., Cherny, R.A., Moir, R.D., Roher, A.E., White, A.R., Cappai, R.,
Cherny, R.A., Bush, A.I., 2009. The Caenorhabditis elegans A beta 1–42 model of Masters, C.L., Tanzi, R.E., Inestrosa, N.C., Bush, A.I., 2002. Metalloenzyme-like
Alzheimer disease predominantly expresses A beta 3–42. J. Biol. Chem. 284, activity of Alzheimer’s disease beta-amyloid. Cu-dependent catalytic conver-
22697–22702. sion of dopamine, cholesterol, and biological reducing agents to neurotoxic
McDermott, J.R., Smith, A.I., Iqbal, K., Wisniewski, H.M., 1979. Brain aluminum in H(2)O(2). J. Biol. Chem. 277, 40302–40308.
aging and Alzheimer disease. Neurology 29, 809–814. Opazo, C., Luza, S., Villemagne, V.L., Volitakis, I., Rowe, C., Barnham, K.J., Strozyk, D.,
McMaster, D., McCrum, E., Patterson, C.C., Kerr, M.M., O’Reilly, D., Evans, A.E., Love, Masters, C.L., Cherny, R.A., Bush, A.I., 2006. Radioiodinated clioquinol as a
A.H., 1992. Serum copper and zinc in random samples of the population of biomarker for beta-amyloid: Zn complexes in Alzheimer’s disease. Aging Cell
Northern Ireland. Am. J. Clin. Nutr. 56, 440–446. 5, 69–79.
Mekmouche, Y., Coppel, Y., Hochgrafe, K., Guilloreau, L., Talmard, C., Mazarguil, H., Osterman, P.O., 1971. Myelopathy after clioquinol treatment. Lancet 2, 544.
Faller, P., 2005. Characterization of the ZnII binding to the peptide amyloid- Ostrakhovitch, E.A., Lordnejad, M.R., Schliess, F., Sies, H., Klotz, L.O., 2002. Copper
beta1-16 linked to Alzheimer’s disease. Chembiochemistry 6, 1663–1671. ions strongly activate the phosphoinositide-3-kinase/Akt pathway independent
Meloni, G., Sonois, V., Delaine, T., Guilloreau, L., Gillet, A., Teissie, J., Faller, P., Vasak, of the generation of reactive oxygen species. Arch. Biochem. Biophys. 397, 232–
M., 2008. Metal swap between Zn7-metallothionein-3 and amyloid-beta-Cu 239.
protects against amyloid-beta toxicity. Nat. Chem. Biol. 4, 366–372. Paik, S.R., Lee, J.H., Kim, D.H., Chang, C.S., Kim, J., 1997. Aluminum-induced struc-
Menditto, A., Morisi, G., Alimonti, A., Caroli, S., Petrucci, F., Spagnolo, A., Menotti, A., tural alterations of the precursor of the non-A beta component of Alzheimer’s
1993. Association of serum copper and zinc with serum electrolytes and with disease amyloid. Arch. Biochem. Biophys. 344, 325–334.
selected risk factors for cardiovascular disease in men aged 55–75 years. NFR Palmiter, R.D., Cole, T.B., Quaife, C.J., Findley, S.D., 1996. ZnT-3, a putative trans-
Study Group. J. Trace Elem. Electrolytes Health Dis. 7, 251–253. porter of zinc into synaptic vesicles. Proc. Natl. Acad. Sci. U.S.A. 93, 14934–
Mercer, J.F., 2001. The molecular basis of copper-transport diseases. Trends Mol. 14939.
Med. 7, 64–69. Park, I.H., Jung, M.W., Mori, H., Mook-Jung, I., 2001. Zinc enhances synthesis of
Metodiewa, D., 1998. Molecular mechanisms of cellular injury produced by neuro- presenilin 1 in mouse primary cortical culture. Biochem. Biophys. Res. Com-
toxic amino acids that generate reactive oxygen species. Amino Acids 14, 181– mun. 285, 680–688.
187. Perluigi, M., Joshi, G., Sultana, R., Calabrese, V., De Marco, C., Coccia, R., Butterfield,
Migliore, L., Coppede, F., 2009. Genetics, environmental factors and the emerging D.A., 2006. In vivo protection by the xanthate tricyclodecan-9-yl-xanthogenate
role of epigenetics in neurodegenerative diseases. Mutat. Res. 667, 82–97. against amyloid beta-peptide (1-42)-induced oxidative stress. Neuroscience
Milne, D.B., Johnson, P.E., 1993. Assessment of copper status: effect of age and 138, 1161–1170.
gender on reference ranges in healthy adults. Clin. Chem. 39, 883–887. Petri, S., Calingasan, N.Y., Alsaied, O.A., Wille, E., Kiaei, M., Friedman, J.E., Baranova,
Miyata, M., Smith, J.D., 1996. Apolipoprotein E allele-specific antioxidant activity O., Chavez, J.C., Beal, M.F., 2007. The lipophilic metal chelators DP-109 and DP-
and effects on cytotoxicity by oxidative insults and beta-amyloid peptides. Nat. 460 are neuroprotective in a transgenic mouse model of amyotrophic lateral
Genet. 14, 55–61. sclerosis. J. Neurochem. 102, 991–1000.
Moir, R.D., Atwood, C.S., Romano, D.M., Laurans, M.H., Huang, X., Bush, A.I., Smith, Phinney, A.L., Drisaldi, B., Schmidt, S.D., Lugowski, S., Coronado, V., Liang, Y., Horne,
J.D., Tanzi, R.E., 1999. Differential effects of apolipoprotein E isoforms on metal- P., Yang, J., Sekoulidis, J., Coomaraswamy, J., Chishti, M.A., Cox, D.W., Mathews,
induced aggregation of A beta using physiological concentrations. Biochemistry P.M., Nixon, R.A., Carlson, G.A., St George-Hyslop, P., Westaway, D., 2003. In vivo
38, 4595–4603. reduction of amyloid-beta by a mutant copper transporter. Proc. Natl. Acad. Sci.
Molina, J.A., Jimenez-Jimenez, F.J., Aguilar, M.V., Meseguer, I., Mateos-Vega, C.J., U.S.A. 100, 14193–14198.
Gonzalez-Munoz, M.J., de Bustos, F., Porta, J., Orti-Pareja, M., Zurdo, M., Barrios, Pierson, K.B., Evenson, M.A., 1988. 200 Kd neurofilament protein binds Al, Cu and
E., Martinez-Para, M.C., 1998. Cerebrospinal fluid levels of transition metals in Zn. Biochem. Biophys. Res. Commun. 152, 598–604.
patients with Alzheimer’s disease. J. Neural. Transm. 105, 479–488. Pinero, D.J., Hu, J., Connor, J.R., 2000. Alterations in the interaction between iron
Monget, A.L., Galan, P., Preziosi, P., Keller, H., Bourgeois, C., Arnaud, J., Favier, A., regulatory proteins and their iron responsive element in normal and Alzhei-
Hercberg, S., 1996. Micronutrient status in elderly people. Geriatrie/Min. Vit. mer’s diseased brains. Cell Mol. Biol. (Noisy-le-grand) 46, 761–776.
Aux Network. Int. J. Vitam. Nutr. Res. 66, 71–76. Puglielli, L., Friedlich, A.L., Setchell, K.D., Nagano, S., Opazo, C., Cherny, R.A., Barn-
Morris, C.M., Keith, A.B., Edwardson, J.A., Pullen, R.G., 1992. Uptake and distribution ham, K.J., Wade, J.D., Melov, S., Kovacs, D.M., Bush, A.I., 2005. Alzheimer disease
of iron and transferrin in the adult rat brain. J. Neurochem. 59, 300–306. beta-amyloid activity mimics cholesterol oxidase. J. Clin. Invest. 115, 2556–
Morris, C.M., Kerwin, J.M., Edwardson, J.A., 1994. Non-haem iron histochemistry of 2563.
the normal and Alzheimer’s disease hippocampus. Neurodegeneration 3, 267– Qian, Y., Tiffany-Castiglioni, E., Welsh, J., Harris, E.D., 1998. Copper efflux from
275. murine microvascular cells requires expression of the menkes disease Cu-
Morris, M.C., Evans, D.A., Tangney, C.C., Bienias, J.L., Schneider, J.A., Wilson, R.S., ATPase. J. Nutr. 128, 1276–1282.
Scherr, P.A., 2006. Dietary copper and high saturated and trans fat intakes Quinta-Ferreira, M.E., Matias, C.M., 2005. Tetanically released zinc inhibits hippo-
associated with cognitive decline. Arch. Neurol. 63, 1085–1088. campal mossy fiber calcium, zinc and synaptic responses. Brain Res. 1047, 1–9.
Munro, H.N., Suter, P.M., Russell, R.M., 1987. Nutritional requirements of the elderly. Rae, T.D., Schmidt, P.J., Pufahl, R.A., Culotta, V.C., O’Halloran, T.V., 1999. Undetect-
Annu. Rev. Nutr. 7, 23–49. able intracellular free copper: the requirement of a copper chaperone for
Murray, I.V., Liu, L., Komatsu, H., Uryu, K., Xiao, G., Lawson, J.A., Axelsen, P.H., 2007. superoxide dismutase. Science 284, 805–808.
Membrane-mediated amyloidogenesis and the promotion of oxidative lipid Raman, B., Ban, T., Yamaguchi, K., Sakai, M., Kawai, T., Naiki, H., Goto, Y., 2005. Metal
damage by amyloid beta proteins. J. Biol. Chem. 282, 9335–9345. ion-dependent effects of clioquinol on the fibril growth of an amyloid {beta}
Murray, I.V., Sindoni, M.E., Axelsen, P.H., 2005. Promotion of oxidative lipid mem- peptide. J. Biol. Chem. 280, 16157–16162.
brane damage by amyloid beta proteins. Biochemistry 44, 12606–12613. Rauk, A., 2009. The chemistry of Alzheimer’s disease. Chem. Soc. Rev. 38, 2698–
Nadal, R.C., Rigby, S.E., Viles, J.H., 2008. Amyloid beta-Cu2+ complexes in both 2715.
monomeric and fibrillar forms do not generate H2O2 catalytically but quench Ravaglia, G., Forti, P., Maioli, F., Nesi, B., Pratelli, L., Savarino, L., Cucinotta, D., Cavalli,
hydroxyl radicals. Biochemistry 47, 11653–11664. G., 2000. Blood micronutrient and thyroid hormone concentrations in the
Nagano, S., Huang, X., Moir, R.D., Payton, S.M., Tanzi, R.E., Bush, A.I., 2004. Peroxidase oldest-old. J. Clin. Endocrinol. Metab. 85, 2260–2265.
activity of cyclooxygenase-2 (COX-2) cross-links beta-amyloid (Abeta) and Regland, B., Lehmann, W., Abedini, I., Blennow, K., Jonsson, M., Karlsson, I., Sjogren,
generates Abeta-COX-2 hetero-oligomers that are increased in Alzheimer’s M., Wallin, A., Xilinas, M., Gottfries, C.G., 2001. Treatment of Alzheimer’s disease
disease. J. Biol. Chem. 279, 14673–14678. with clioquinol. Dement. Geriatr. Cogn. Disord. 12, 408–414.
Nakano, E., Williamson, M.P., Williams, N.H., Powers, H.J., 2004. Copper-mediated Reznichenko, L., Amit, T., Zheng, H., Avramovich-Tirosh, Y., Youdim, M.B., Weinreb,
LDL oxidation by homocysteine and related compounds depends largely on O., Mandel, S., 2006. Reduction of iron-regulated amyloid precursor protein and
copper ligation. Biochim. Biophys. Acta 1688, 33–42. beta-amyloid peptide by ()-epigallocatechin-3-gallate in cell cultures: impli-
Namekata, K., Imagawa, M., Terashi, A., Ohta, S., Oyama, F., Ihara, Y., 1997. Associa- cations for iron chelation in Alzheimer’s disease. J. Neurochem. 97, 527–536.
tion of transferrin C2 allele with late-onset Alzheimer’s disease. Hum. Genet. Ridge, P.G., Zhang, Y., Gladyshev, V.N., 2008. Comparative genomic analyses of
101, 126–129. copper transporters and cuproproteomes reveal evolutionary dynamics of
Neldner, K.H., Hambidge, K.M., 1975. Zinc therapy of acrodermatitis enteropathica. copper utilization and its link to oxygen. PLoS One 3, e1378.
N. Engl. J. Med. 292, 879–882. Ritchie, C.W., Bush, A.I., Mackinnon, A., Macfarlane, S., Mastwyk, M., MacGregor, L.,
Nelson, T.J., Alkon, D.L., 2005. Oxidation of cholesterol by amyloid precursor protein Kiers, L., Cherny, R., Li, Q.X., Tammer, A., Carrington, D., Mavros, C., Volitakis, I.,
and beta-amyloid peptide. J. Biol. Chem. 280, 7377–7387. Xilinas, M., Ames, D., Davis, S., Beyreuther, K., Tanzi, R.E., Masters, C.L., 2003.
Niciu, M.J., Ma, X.M., El Meskini, R., Pachter, J.S., Mains, R.E., Eipper, B.A., 2007. Metal-protein attenuation with iodochlorhydroxyquin (clioquinol) targeting
Altered ATP7A expression and other compensatory responses in a murine Abeta amyloid deposition and toxicity in Alzheimer disease: a pilot phase 2
model of Menkes disease. Neurobiol. Dis. 27, 278–291. clinical trial. Arch. Neurol. 60, 1685–1691.
Nurmi, A., Goldsteins, G., Narvainen, J., Pihlaja, R., Ahtoniemi, T., Grohn, O., Kois- Robson, K.J., Lehmann, D.J., Wimhurst, V.L., Livesey, K.J., Combrinck, M., Merry-
tinaho, J., 2006. Antioxidant pyrrolidine dithiocarbamate activates Akt-GSK weather-Clarke, A.T., Warden, D.R., Smith, A.D., 2004. Synergy between the C2
signaling and is neuroprotective in neonatal hypoxia-ischemia. Free Radic. Biol. allele of transferrin and the C282Y allele of the haemochromatosis gene (HFE) as
Med. 40, 1776–1784. risk factors for developing Alzheimer’s disease. J. Med. Genet. 41, 261–265.
Oken, B.S., Storzbach, D.M., Kaye, J.A., 1998. The efficacy of Ginkgo biloba on Rogers, J.T., Randall, J.D., Cahill, C.M., Eder, P.S., Huang, X., Gunshin, H., Leiter, L.,
cognitive function in Alzheimer disease. Arch. Neurol. 55, 1409–1415. McPhee, J., Sarang, S.S., Utsuki, T., Greig, N.H., Lahiri, D.K., Tanzi, R.E., Bush, A.I.,
Omar, R.A., Chyan, Y.J., Andorn, A.C., Poeggeler, B., Robakis, N.K., Pappolla, M.A., Giordano, T., Gullans, S.R., 2002. An iron-responsive element type II in the 50 -
1999. Increased expression but reduced activity of antioxidant enzymes in untranslated region of the Alzheimer’s amyloid precursor protein transcript. J.
Alzheimer’s disease. J. Alzheimers Dis. 1, 139–145. Biol. Chem. 277, 45518–45528.
J.A. Duce, A.I. Bush / Progress in Neurobiology 92 (2010) 1–18 17

Roskams, A.J., Connor, J.R., 1994. Iron, transferrin, and ferritin in the rat brain during Sohal, R.S., Wennberg-Kirch, E., Jaiswal, K., Kwong, L.K., Forster, M.J., 1999. Effect of
development and aging. J. Neurochem. 63, 709–716. age and caloric restriction on bleomycin-chelatable and nonheme iron in
Rozga, M., Bal, W., 2009. The Cu(II)/Abeta/human serum albumin model of control different tissues of C57BL/6 mice. Free Radic. Biol. Med. 27, 287–293.
mechanism for copper-related amyloid neurotoxicity. Chem. Res. Toxicol. 23, Sparks, D.L., Schreurs, B.G., 2003. Trace amounts of copper in water induce beta-
298–308. amyloid plaques and learning deficits in a rabbit model of Alzheimer’s disease.
Rozga, M., Protas, A.M., Jablonowska, A., Dadlez, M., Bal, W., 2009. The Cu(II) Proc. Natl. Acad. Sci. U.S.A. 100, 11065–11069.
complex of Abeta40 peptide in ammonium acetate solutions. Evidence for Squitti, R., Lupoi, D., Pasqualetti, P., Dal Forno, G., Vernieri, F., Chiovenda, P., Rossi, L.,
ternary species formation. Chem. Commun. (Camb.) 1374–1376. Cortesi, M., Cassetta, E., Rossini, P.M., 2002a. Elevation of serum copper levels in
Sacharczuk, M., Zagulski, T., Sadowski, B., Barcikowska, M., Pluta, R., 2005. Lacto- Alzheimer’s disease. Neurology 59, 1153–1161.
ferrin in the central nervous system. Neurol. Neurochir. Pol. 39, 482–489. Squitti, R., Pasqualetti, P., Dal Forno, G., Moffa, F., Cassetta, E., Lupoi, D., Vernieri, F.,
Sayre, L.M., Perry, G., Harris, P.L., Liu, Y., Schubert, K.A., Smith, M.A., 2000. In situ Rossi, L., Baldassini, M., Rossini, P.M., 2005. Excess of serum copper not related
oxidative catalysis by neurofibrillary tangles and senile plaques in Alzheimer’s to ceruloplasmin in Alzheimer disease. Neurology 64, 1040–1046.
disease: a central role for bound transition metals. J. Neurochem. 74, 270–279. Squitti, R., Rossini, P.M., Cassetta, E., Moffa, F., Pasqualetti, P., Cortesi, M., Colloca, A.,
Scheuermann, S., Hambsch, B., Hesse, L., Stumm, J., Schmidt, C., Beher, D., Bayer, T.A., Rossi, L., Finazzi-Agro, A., 2002b. d-penicillamine reduces serum oxidative
Beyreuther, K., Multhaup, G., 2001. Homodimerization of amyloid precursor stress in Alzheimer’s disease patients. Eur. J. Clin. Invest. 32, 51–59.
protein and its implication in the amyloidogenic pathway of Alzheimer’s Stoltenberg, M., Bush, A.I., Bach, G., Smidt, K., Larsen, A., Rungby, J., Lund, S., Doering,
disease. J. Biol. Chem. 276, 33923–33929. P., Danscher, G., 2007. Amyloid plaques arise from zinc-enriched cortical layers
Schlief, M.L., Craig, A.M., Gitlin, J.D., 2005. NMDA receptor activation mediates in APP/PS1 transgenic mice and are paradoxically enlarged with dietary zinc
copper homeostasis in hippocampal neurons. J. Neurosci. 25, 239–246. deficiency. Neuroscience 150, 357–369.
Schlief, M.L., West, T., Craig, A.M., Holtzman, D.M., Gitlin, J.D., 2006. Role of the Su, X.Y., Wu, W.H., Huang, Z.P., Hu, J., Lei, P., Yu, C.H., Zhao, Y.F., Li, Y.M., 2007.
Menkes copper-transporting ATPase in NMDA receptor-mediated neuronal Hydrogen peroxide can be generated by tau in the presence of Cu(II). Biochem.
toxicity. Proc. Natl. Acad. Sci. U.S.A. 103, 14919–14924. Biophys. Res. Commun. 358, 661–665.
Schneider, A., Rajendran, L., Honsho, M., Gralle, M., Donnert, G., Wouters, F., Hell, Suh, J.H., Moreau, R., Heath, S.H., Hagen, T.M., 2005. Dietary supplementation with
S.W., Simons, M., 2008. Flotillin-dependent clustering of the amyloid precursor (R)-alpha-lipoic acid reverses the age-related accumulation of iron and deple-
protein regulates its endocytosis and amyloidogenic processing in neurons. J. tion of antioxidants in the rat cerebral cortex. Redox Rep. 10, 52–60.
Neurosci. 28, 2874–2882. Suh, S.W., Jensen, K.B., Jensen, M.S., Silva, D.S., Kesslak, P.J., Danscher, G., Freder-
Schneider, L.S., DeKosky, S.T., Farlow, M.R., Tariot, P.N., Hoerr, R., Kieser, M., 2005. A ickson, C.J., 2000. Histochemically-reactive zinc in amyloid plaques, angiopathy,
randomized, double-blind, placebo-controlled trial of two doses of Ginkgo biloba and degenerating neurons of Alzheimer’s diseased brains. Brain Res. 852, 274–
extract in dementia of the Alzheimer’s type. Curr. Alzheimer Res. 2, 541–551. 278.
Schreck, R., Meier, B., Mannel, D.N., Droge, W., Baeuerle, P.A., 1992. Dithiocarba- Suh, Y.H., Checler, F., 2002. Amyloid precursor protein, presenilins, and alpha-
mates as potent inhibitors of nuclear factor kappa B activation in intact cells. J. synuclein: molecular pathogenesis and pharmacological applications in Alz-
Exp. Med. 175, 1181–1194. heimer’s disease. Pharmacol. Rev. 54, 469–525.
Scott, L.E., Orvig, C., 2009. Medicinal inorganic chemistry approaches to passivation Sullivan, P.G., Brown, M.R., 2005. Mitochondrial aging and dysfunction in
and removal of aberrant metal ions in disease. Chem. Rev. 109, 4885–4910. Alzheimer’s disease. Prog. Neuropsychopharmacol. Biol. Psychiatry 29,
Sensi, S.L., Paoletti, P., Bush, A.I., Sekler, I., 2009. Zinc in the physiology and 407–410.
pathology of the CNS. Nat. Rev. Neurosci. 10, 780–791. Syme, C.D., Nadal, R.C., Rigby, S.E., Viles, J.H., 2004. Copper binding to the amyloid-
Seshadri, S., Beiser, A., Selhub, J., Jacques, P.F., Rosenberg, I.H., D’Agostino, R.B., beta (Abeta) peptide associated with Alzheimer’s disease: folding, coordination
Wilson, P.W., Wolf, P.A., 2002. Plasma homocysteine as a risk factor for geometry, pH dependence, stoichiometry, and affinity of Abeta-(1–28): insights
dementia and Alzheimer’s disease. N. Engl. J. Med. 346, 476–483. from a range of complementary spectroscopic techniques. J. Biol. Chem. 279,
Shearer, J., Szalai, V.A., 2008. The amyloid-beta peptide of Alzheimer’s disease binds 18169–18177.
Cu(I) in a linear bis-his coordination environment: insight into a possible Syme, C.D., Viles, J.H., 2006. Solution 1H NMR investigation of Zn2+ and Cd2+
neuroprotective mechanism for the amyloid-beta peptide. J. Am. Chem. Soc. binding to amyloid-beta peptide (Abeta) of Alzheimer’s disease. Biochim.
130, 17826–17835. Biophys. Acta 1764, 246–256.
Shimada, Y., Tsuji, T., Igata, A., Steinitz, H., 1971. Halogenated oxyquinoline deri- Szallasi, Z., Bogi, K., Gohari, S., Biro, T., Acs, P., Blumberg, P.M., 1996. Non-equivalent
vatives and neurological syndromes. Lancet 2, 41–43. roles for the first and second zinc fingers of protein kinase Cdelta. Effect of their
Shin, R.W., Kruck, T.P., Murayama, H., Kitamoto, T., 2003. A novel trivalent cation mutation on phorbol ester-induced translocation in NIH 3T3 cells. J. Biol. Chem.
chelator Feralex dissociates binding of aluminum and iron associated with 271, 18299–18301.
hyperphosphorylated tau of Alzheimer’s disease. Brain Res. 961, 139–146. Tabner, B.J., Turnbull, S., El-Agnaf, O.M., Allsop, D., 2002. Formation of hydrogen
Shinall, H., Song, E.S., Hersh, L.B., 2005. Susceptibility of amyloid beta peptide peroxide and hydroxyl radicals from A(beta) and alpha-synuclein as a possible
degrading enzymes to oxidative damage: a potential Alzheimer’s disease spiral. mechanism of cell death in Alzheimer’s disease and Parkinson’s disease. Free
Biochemistry 44, 15345–15350. Radic. Biol. Med. 32, 1076–1083.
Shishodia, S., Sethi, G., Aggarwal, B.B., 2005. Curcumin: getting back to the roots. Takeda, A., 2004. Essential trace metals and brain function. Yakugaku Zasshi 124,
Ann. N. Y. Acad. Sci. 1056, 206–217. 577–585.
Siddiq, A., Ayoub, I.A., Chavez, J.C., Aminova, L., Shah, S., LaManna, J.C., Patton, S.M., Takeda, A., Minami, A., Takefuta, S., Tochigi, M., Oku, N., 2001. Zinc homeostasis in
Connor, J.R., Cherny, R.A., Volitakis, I., Bush, A.I., Langsetmo, I., Seeley, T., the brain of adult rats fed zinc-deficient diet. J. Neurosci. Res. 63, 447–452.
Gunzler, V., Ratan, R.R., 2005. Hypoxia-inducible factor prolyl 4-hydroxylase Tapiero, H., Townsend, D.M., Tew, K.D., 2003. Trace elements in human physiology
inhibition. A target for neuroprotection in the central nervous system. J. Biol. and pathology. Copper. Biomed. Pharmacother. 57, 386–398.
Chem. 280, 41732–41743. Terry, R.D., 1996. The pathogenesis of Alzheimer disease: an alternative to the
Silvestri, L., Camaschella, C., 2008. A potential pathogenetic role of iron in Alzhei- amyloid hypothesis. J. Neuropathol. Exp. Neurol. 55, 1023–1025.
mer’s disease. J. Cell. Mol. Med. 12, 1548–1550. Torsdottir, G., Kristinsson, J., Sveinbjornsdottir, S., Snaedal, J., Johannesson, T., 1999.
Simons, A., Ruppert, T., Schmidt, C., Schlicksupp, A., Pipkorn, R., Reed, J., Masters, C.L., Copper, ceruloplasmin, superoxide dismutase and iron parameters in Parkin-
White, A.R., Cappai, R., Beyreuther, K., Bayer, T.A., Multhaup, G., 2002. Evidence son’s disease. Pharmacol. Toxicol. 85, 239–243.
for a copper-binding superfamily of the amyloid precursor protein. Biochemis- Tsubaki, T., Honma, Y., Hoshi, M., 1971. Neurological syndrome associated with
try 41, 9310–9320. clioquinol. Lancet 1, 696–697.
Sinha, S., Anderson, J.P., Barbour, R., Basi, G.S., Caccavello, R., Davis, D., Doan, M., Ujiie, M., Dickstein, D.L., Jefferies, W.A., 2002. p97 as a biomarker for Alzheimer
Dovey, H.F., Frigon, N., Hong, J., Jacobson-Croak, K., Jewett, N., Keim, P., Knops, J., disease. Front Biosci. 7, e42–47.
Lieberburg, I., Power, M., Tan, H., Tatsuno, G., Tung, J., Schenk, D., Seubert, P., Valensin, D., Mancini, F.M., Luczkowski, M., Janicka, A., Wisniewska, K., Gaggelli, E.,
Suomensaari, S.M., Wang, S., Walker, D., Zhao, J., McConlogue, L., John, V., 1999. Valensin, G., Lankiewicz, L., Kozlowski, H., 2004. Identification of a novel high
Purification and cloning of amyloid precursor protein beta-secretase from affinity copper binding site in the APP(145–155) fragment of amyloid precursor
human brain. Nature 402, 537–540. protein. Dalton Trans. 16–22.
Smart, T.G., Hosie, A.M., Miller, P.S., 2004. Zn2+ ions: modulators of excitatory and Vallee, B.L., Coleman, J.E., Auld, D.S., 1991. Zinc fingers, zinc clusters, and zinc twists
inhibitory synaptic activity. Neuroscientist 10, 432–442. in DNA-binding protein domains. Proc. Natl. Acad. Sci. U.S.A. 88, 999–1003.
Smith, D.G., Cappai, R., Barnham, K.J., 2007. The redox chemistry of the Alzheimer’s Vasak, M., Hasler, D.W., 2000. Metallothioneins: new functional and structural
disease amyloid beta peptide. Biochim. Biophys. Acta 1768, 1976–1990. insights. Curr. Opin. Chem. Biol. 4, 177–183.
Smith, D.P., Smith, D.G., Curtain, C.C., Boas, J.F., Pilbrow, J.R., Ciccotosto, G.D., Lau, Vassar, R., Bennett, B.D., Babu-Khan, S., Kahn, S., Mendiaz, E.A., Denis, P., Teplow,
T.L., Tew, D.J., Perez, K., Wade, J.D., Bush, A.I., Drew, S.C., Separovic, F., Masters, D.B., Ross, S., Amarante, P., Loeloff, R., Luo, Y., Fisher, S., Fuller, J., Edenson, S., Lile,
C.L., Cappai, R., Barnham, K.J., 2006a. Copper-mediated amyloid-beta toxicity is J., Jarosinski, M.A., Biere, A.L., Curran, E., Burgess, T., Louis, J.C., Collins, F.,
associated with an intermolecular histidine bridge. J. Biol. Chem. 281, 15145– Treanor, J., Rogers, G., Citron, M., 1999. Beta-secretase cleavage of Alzheimer’s
15154. amyloid precursor protein by the transmembrane aspartic protease BACE.
Smith, J.L., Xiong, S., Lovell, M.A., 2006b. 4-Hydroxynonenal disrupts zinc export in Science 286, 735–741.
primary rat cortical cells. Neurotoxicology 27, 1–5. Vaughan, D.W., Peters, A., 1981. The structure of neuritic plaque in the cerebral
Smith, M.A., Perry, G., Richey, P.L., Sayre, L.M., Anderson, V.E., Beal, M.F., Kowall, N., cortex of aged rats. J. Neuropathol. Exp. Neurol. 40, 472–487.
1996. Oxidative damage in Alzheimer’s. Nature 382, 120–121. Venti, A., Giordano, T., Eder, P., Bush, A.I., Lahiri, D.K., Greig, N.H., Rogers, J.T., 2004.
Smith, M.A., Richey Harris, P.L., Sayre, L.M., Beckman, J.S., Perry, G., 1997. Wide- The integrated role of desferrioxamine and phenserine targeted to an iron-
spread peroxynitrite-mediated damage in Alzheimer’s disease. J. Neurosci. 17, responsive element in the APP-mRNA 50 -untranslated region. Ann. N. Y. Acad.
2653–2657. Sci. 1035, 34–48.
18 J.A. Duce, A.I. Bush / Progress in Neurobiology 92 (2010) 1–18

Viles, J.H., Cohen, F.E., Prusiner, S.B., Goodin, D.B., Wright, P.E., Dyson, H.J., 1999. reduction to iron (II) reverses the aggregation: implications in the formation of
Copper binding to the prion protein: structural implications of four neurofibrillary tangles of Alzheimer’s disease. J. Neurochem. 82, 1137–1147.
identical cooperative binding sites. Proc. Natl. Acad. Sci. U.S.A. 96, Yan, R., Bienkowski, M.J., Shuck, M.E., Miao, H., Tory, M.C., Pauley, A.M., Brashier, J.R.,
2042–2047. Stratman, N.C., Mathews, W.R., Buhl, A.E., Carter, D.B., Tomasselli, A.G., Parodi,
Vyas, S.B., Duffy, L.K., 1995. Stabilization of secondary structure of Alzheimer beta- L.A., Heinrikson, R.L., Gurney, M.E., 1999. Membrane-anchored aspartyl prote-
protein by aluminum(III) ions and D-Asp substitutions. Biochem. Biophys. Res. ase with Alzheimer’s disease beta-secretase activity. Nature 402, 533–537.
Commun. 206, 718–723. Yang, F., Lim, G.P., Begum, A.N., Ubeda, O.J., Simmons, M.R., Ambegaokar, S.S., Chen,
Wang, D.S., Iwata, N., Hama, E., Saido, T.C., Dickson, D.W., 2003. Oxidized neprilysin P.P., Kayed, R., Glabe, C.G., Frautschy, S.A., Cole, G.M., 2005. Curcumin inhibits
in aging and Alzheimer’s disease brains. Biochem. Biophys. Res. Commun. 310, formation of amyloid beta oligomers and fibrils, binds plaques, and reduces
236–241. amyloid in vivo. J. Biol. Chem. 280, 5892–5901.
Watanabe, S., Nagano, S., Duce, J., Kiaei, M., Li, Q.X., Tucker, S.M., Tiwari, A., Brown Jr., Yassin, M.S., Ekblom, J., Lofberg, C., Oreland, L., 1998. Transmethylation reactions
R.H., Beal, M.F., Hayward, L.J., Culotta, V.C., Yoshihara, S., Sakoda, S., Bush, A.I., and autoradiographic distribution of vitamin B12: effects of clioquinol treat-
2007. Increased affinity for copper mediated by cysteine 111 in forms of mutant ment in mice. Jpn. J. Pharmacol. 78, 55–61.
superoxide dismutase 1 linked to amyotrophic lateral sclerosis. Free Radic. Biol. Yassin, M.S., Ekblom, J., Xilinas, M., Gottfries, C.G., Oreland, L., 2000. Changes in
Med. 42, 1534–1542. uptake of vitamin B(12) and trace metals in brains of mice treated with
Wender, M., Szczech, J., Hoffmann, S., Hilczer, W., 1992. Electron paramagnetic clioquinol. J. Neurol. Sci. 173, 40–44.
resonance analysis of heavy metals in the aging human brain. Neuropatol. Pol. Yong, V.W., Krekoski, C.A., Forsyth, P.A., Bell, R., Edwards, D.R., 1998. Matrix
30, 65–72. metalloproteinases and diseases of the CNS. Trends Neurosci. 21, 75–80.
West, D.X., Liberta, A.E., Rajendran, K.G., Hall, I.H., 1993. The cytotoxicity of Yoshiike, Y., Tanemura, K., Murayama, O., Akagi, T., Murayama, M., Sato, S., Sun, X.,
copper(II) complexes of heterocyclic thiosemicarbazones and 2-substituted Tanaka, N., Takashima, A., 2001. New insights on how metals disrupt amyloid
pyridine N-oxides. Anticancer Drugs 4, 241–249. beta-aggregation and their effects on amyloid-beta cytotoxicity. J. Biol. Chem.
White, A.R., Du, T., Laughton, K.M., Volitakis, I., Sharples, R.A., Xilinas, M.E., Hoke, 276, 32293–32299.
D.E., Holsinger, R.M., Evin, G., Cherny, R.A., Hill, A.F., Barnham, K.J., Li, Q.X., Bush, Yoshimoto, N., Tasaki, M., Shimanouchi, T., Umakoshi, H., Kuboi, R., 2005. Oxidation
A.I., Masters, C.L., 2006. Degradation of the Alzheimer disease amyloid beta- of cholesterol catalyzed by amyloid beta-peptide (Abeta)-Cu complex on lipid
peptide by metal-dependent up-regulation of metalloprotease activity. J. Biol. membrane. J. Biosci. Bioeng. 100, 455–459.
Chem. 281, 17670–17680. Yu, W.H., Lukiw, W.J., Bergeron, C., Niznik, H.B., Fraser, P.E., 2001. Metallothionein III
White, A.R., Reyes, R., Mercer, J.F., Camakaris, J., Zheng, H., Bush, A.I., Multhaup, G., is reduced in Alzheimer’s disease. Brain Res. 894, 37–45.
Beyreuther, K., Masters, C.L., Cappai, R., 1999. Copper levels are increased in the Zandi, P.P., Anthony, J.C., Khachaturian, A.S., Stone, S.V., Gustafson, D., Tschanz, J.T.,
cerebral cortex and liver of APP and APLP2 knockout mice. Brain Res. 842, 439– Norton, M.C., Welsh-Bohmer, K.A., Breitner, J.C., 2004. Reduced risk of Alzhei-
444. mer disease in users of antioxidant vitamin supplements: the Cache County
Whitson, J.S., Selkoe, D.J., Cotman, C.W., 1989. Amyloid beta protein enhances the Study. Arch. Neurol. 61, 82–88.
survival of hippocampal neurons in vitro. Science 243, 1488–1490. Zatta, P., Tognon, G., Carampin, P., 2003. Melatonin prevents free radical formation
Wu, J., Basha, M.R., Brock, B., Cox, D.P., Cardozo-Pelaez, F., McPherson, C.A., Harry, J., due to the interaction between beta-amyloid peptides and metal ions [Al(III),
Rice, D.C., Maloney, B., Chen, D., Lahiri, D.K., Zawia, N.H., 2008. Alzheimer’s Zn(II), Cu(II), Mn(II), Fe(II)]. J. Pineal. Res. 35, 98–103.
disease (AD)-like pathology in aged monkeys after infantile exposure to envi- Zecca, L., Gallorini, M., Schunemann, V., Trautwein, A.X., Gerlach, M., Riederer, P.,
ronmental metal lead (Pb): evidence for a developmental origin and environ- Vezzoni, P., Tampellini, D., 2001. Iron, neuromelanin and ferritin content in the
mental link for AD. J. Neurosci. 28, 3–9. substantia nigra of normal subjects at different ages: consequences for iron
Wu, W., Graves, L.M., Jaspers, I., Devlin, R.B., Reed, W., Samet, J.M., 1999. Activation storage and neurodegenerative processes. J. Neurochem. 76, 1766–1773.
of the EGF receptor signaling pathway in human airway epithelial cells exposed Zhang, A.S., Enns, C.A., 2009. Iron homeostasis: recently identified proteins provide
to metals. Am. J. Physiol. 277, L924–931. insight into novel control mechanisms. J. Biol. Chem. 284, 711–715.
Xie, C.X., Mattson, M.P., Lovell, M.A., Yokel, R.A., 1996. Intraneuronal aluminum Zhang, X., Luhrs, K.J., Ryff, K.A., Malik, W.T., Driscoll, M.J., Culver, B., 2009. Suppres-
potentiates iron-induced oxidative stress in cultured rat hippocampal neurons. sion of nuclear factor kappa B ameliorates astrogliosis but not amyloid burden
Brain Res. 743, 271–277. in APPswe/PS1dE9 mice. Neuroscience 161, 53–58.
Xu, N., Majidi, V., Markesbery, W.R., Ehmann, W.D., 1992. Brain aluminum in Zhou, L.X., Du, J.T., Zeng, Z.Y., Wu, W.H., Zhao, Y.F., Kanazawa, K., Ishizuka, Y.,
Alzheimer’s disease using an improved GFAAS method. Neurotoxicology 13, Nemoto, T., Nakanishi, H., Li, Y.M., 2007. Copper (II) modulates in vitro aggre-
735–743. gation of a tau peptide. Peptides 28, 2229–2234.
Yamamoto, A., Shin, R.W., Hasegawa, K., Naiki, H., Sato, H., Yoshimasu, F., Kitamoto, Zhu, X., Su, B., Wang, X., Smith, M.A., Perry, G., 2007. Causes of oxidative stress in
T., 2002. Iron (III) induces aggregation of hyperphosphorylated tau and its Alzheimer disease. Cell Mol. Life Sci. 64, 2202–2210.

You might also like