You are on page 1of 13

Coordination Chemistry Reviews 256 (2012) 2129–2141

Contents lists available at SciVerse ScienceDirect

Coordination Chemistry Reviews


journal homepage: www.elsevier.com/locate/ccr

Review

Copper, zinc and iron in neurodegenerative diseases


(Alzheimer’s, Parkinson’s and prion diseases)
Henryk Kozlowski a,∗ , Marek Luczkowski a , Maurizio Remelli b , Daniela Valensin c
a
Faculty of Chemistry, University of Wroclaw, 50-383 Wroclaw, Poland
b
Dipartimento di Chimica, Università di Ferrara, Ferrara, Italy
c
Dipartimento di Chimica, Università di Siena, Siena, Italy

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2129
2. Brain metal homeostasis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2130
2.1. Copper . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2130
2.2. Iron . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2131
2.3. Zinc . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2132
3. Neurodegenerative diseases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2132
3.1. Alzheimer’s disease . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2133
3.2. Parkinson’s disease . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2133
3.3. Prion diseases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2133
4. Metal binding to neurodegenerative proteins . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2134
4.1. Alzheimer’s disease . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2134
4.2. Parkinson’s disease . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2135
4.3. Prion diseases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2135
4.4. Metal–protein affinity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2136
5. Protein misfolding . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2136
6. Oxidative stress . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2138
7. Metal based therapeutic strategies for neurogenerative disease . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2138
Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2139
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2139

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

Article history: The basic role of metal ions including copper, zinc and iron in neurological pathologies is generally
Received 27 December 2011 accepted. The relationship between the development of disease and particular metal ions is very compli-
Received in revised form 20 February 2012 cated and complex. Thus, comprehension of metal homeostasis, details of transport and interactions with
Accepted 9 March 2012
biomolecules is essential for understanding the normal and pathological processes occurring in the living
Available online 30 March 2012
system. Homeostasis of metal ions usually involves a huge set of proteins which regulate the proper metal
biology. Disorder in metal homeostasis may result in serious pathologies including neurodegenerative
Keywords:
diseases. Metal ions, especially copper, zinc and iron play very important roles in neurodegeneration
Brain metal homeostasis
Neurodegenerative diseases
having impact on both protein structure (misfolding) and oxidative stress. Metal ion binding to proteins
Metal binding to neurodegenerative involved in neurodegeneration is therefore an important factor for whole brain damage processes. All
proteins these aspects are discussed in the review.
Protein misfolding © 2012 Elsevier B.V. All rights reserved.
Oxidative stress

1. Introduction

The fundamentals of homeostasis and the basic roles of copper,


zinc and iron in biological functioning and neurological disorders
∗ Corresponding author. Tel.: +48 713757251; fax: +48 713757251.
in the brain will be covered in this review. All three metals play
E-mail addresses: henryk.kozlowski@chem.uni.wroc.pl, henrykoz@wchuwr.pl
(H. Kozlowski), marek.luczkowski@chem.uni.wroc.pl (M. Luczkowski), critical roles in the brain biology being present in some regions
rmm@unife.it (M. Remelli), daniela.valensin@unisi.it (D. Valensin). of the brain in millimolar concentrations. Iron and copper are

0010-8545/$ – see front matter © 2012 Elsevier B.V. All rights reserved.
doi:10.1016/j.ccr.2012.03.013
2130 H. Kozlowski et al. / Coordination Chemistry Reviews 256 (2012) 2129–2141

also effective catalysts for production of oxygen radicals so they breakdown of the metal homeostatic mechanisms. It is, therefore,
are major cause of the oxidative stress. Furthermore, the metal crucial to understand the structural and functional interactions
ions mentioned above may have a basic impact on the protein between metals and the various intracellular and extracellular
misfolding and progression of the neurodegenerative processes. components of the CNS, both in healthy and in pathological
All these aspects are discussed below. conditions.
Here, we will focus on the three major biological transition met-
als involved in neurodegeneration, copper, zinc and iron. For all the
2. Brain metal homeostasis metal ions similar mechanisms of damage were found including:
(a) free radical production (for redox active metals), (b) protein
Metal ions are required to sustain a huge number of cellular pro- aggregation and (c) metal transport alteration [3,6,14–17].
cesses, including neuronal ones. Transition metal ions like Fe, Zn,
and Cu play an important role in regulating neuronal activity within 2.1. Copper
the synapses and they are essential for the biological functions of
several metallo-proteins such as Cu/Zn superoxide dismutase and Copper is a redox-active transition metal occurring in two main
cytochrome c oxidase. In the central nervous system (CNS) sev- oxidation states, +2 and +1. It coordinates to different biological
eral transition metal ions like iron, copper and zinc are required for ligands such as imidazole nitrogen, amide and amino nitrogen,
many enzymatic activities, mitochondrial function, neurotransmis- carboxylate oxygen, cysteine and methionine sulfur donor atoms
sion, learning and memory. For example Zn2+ stored in presynaptic [18–23]. Although the copper redox activity is fundamental for the
vesicles is involved in neuron signaling process, it is released in biological activity of a huge number of enzymes, it can become
brief pulses into the synaptic cleft, and acts on recognition sites in highly dangerous if not accurately regulated. For these reasons,
the postsynaptic membrane [1]. specific proteins guarantee that no copper ions can freely circulate.
Even if these transition metal ions are vital to ensure many cel- The full comprehension of copper transport within the CNS is
lular processes, their concentrations within the cell must be finely not yet reached, although the proteins involved in copper homeo-
tuned since any aberrations may lead to cell death and severe stasis in the brain and their importance in the CNS have been
diseases. Metal crossing through the blood–brain barrier (BBB) is identified [24]. The proteins which play a key roles in regulating
strictly regulated; many transporters, enzymes and chaperones copper distribution and transport in peripheral tissues are also
tightly regulate the metal ion content within the brain controlling present in the brain as described by the schematic representation
metal uptake and delivery to specific domains and preventing a pas- shown in Fig. 1. The major copper binding proteins (cuproenzymes
sive flux of metals from the circulation to the brain [2]. In a healthy and/or carrier proteins) in extracellular fluids are ceruloplasmin,
state the concentration of free metal ions is maintained at a very albumin, and transcuprein, which transport the absorbed copper
low level, each physiological metal ion is selectively delivered to to all tissues. Ctr1 and ATP7A or ATP7B transport copper across cel-
its own site of action in order to tightly control metal reactivity (i.e. lular membrane transferring copper into (import) or out of (export)
redox activity). the cytosol, respectively. Once inside cells, cytosolic copper chap-
In order to maintain metal-ion homeostasis cells have devel- erones deliver the metal to intracellular protein targets (e.g. copper
oped a well refined machinery whose failure might lead to several chaperone of superoxide dismutase (CCS), delivers copper to super-
disease state, among which are neurodegenerative disorders [3–8]. oxide dismutase 1, SOD1) [25].
It is well accepted that transition metal ions are involved in Copper distribution within CNS varies according to species. Each
Alzheimer’s (AD), Parkinson’s (PD) and prion diseases (TSEs). High region of the brain is characterized by specific copper distribu-
content of copper and zinc is present in amyloid plaques from AD tion which increases with aging. The regions showing the highest
brains (25.0 ± 7.8 and 69.0 ± 18.4 ␮g/g respectively) [9] and in the copper concentrations might be associated with the regions with
cerebrospinal fluid (CSF) of patients affected by both AD and PD the highest copper metabolic demand. In rat, the areas showing
[10]. In particular Cu and Zn concentration in CSF are almost two high copper enrichment are in the (i) medial geniculate nucleus
or three fold higher than control levels (10.2 and 5.3 ng/g for cop- (the region responsible of the visual information processing), (ii)
per and zinc respectively). Accumulation of iron is also frequently
observed in brain areas linked to both Parkinson’s and Alzheimer’s
diseases [11].
Proteins associated with neurodegenerative diseases (amyloid-
␤, A␤; ␣-synuclein, ␣S; prion protein, PrP) also bind several
transition metal ions and metal binding affects the aggregation
process of A␤ and ␣S (vide infra Sections 4 and 5). In both PD and AD
copper and iron are associated with reactive oxygen species (ROS)
generation. The Fenton reactions (Cu+ + H2 O2 → Cu2+ + OH• + OH−
or Fe2+ + H2 O2 → Fe3+ + OH• + OH− ) lead to the formation of the
highly reactive hydroxyl radical (OH• ) which induces oxidative
stress in substantia nigra pars compacta dopamine neurons by
lipid membrane peroxidation, DNA damage and protein oxidation
or misfolding (vide infra Section 6). As for PrP, strong evidence
indicates that it has a role in the regulation of brain copper
metabolism. It has been suggested that PrP–copper interactions
might modulate protection against Cu2+ mediated oxidative stress,
copper transport and copper dependent cellular signaling [12].
The possible beneficial or detrimental role of copper in PrP disease
is still a matter of debate [13].
The involvement of iron, copper and zinc in many neurode- Fig. 1. Schematic representation of copper distribution and transport in a cell of
generative diseases is not simply correlated with increased and CNS.
toxicological exposure to these metals, but rather it depends on a Source: Figure reproduced from Ref. [24] with permission.
H. Kozlowski et al. / Coordination Chemistry Reviews 256 (2012) 2129–2141 2131

superior colliculus (a component of the midbrain associated with functions seem to have more iron than non-motor related regions
motor functions), (iii) periaqueductal gray (the gray matter modu- explaining why movement disorders are often associated with iron
lating pain and defensive behavior), (iv) the lateral amygdala (area loading [35]. As for copper, iron concentration in the brain seems
involved in memory and emotional reactions) and in the (v) dor- to increase with aging [35].
somedial aspect of the diencephalon (the part of the forebrain Different proteins have been identified as iron binding pro-
containing thalamus, hypothalamus and the posterior portion of teins in the brain: (i) H- and L-chain ferritin in oligodendrocytes,
the pituitary gland). In humans the highest copper concentration L-ferritin in microglia and neuromelanin in neurons. Since H- and L-
occurs in the hippocampus [26]. ferritin have different functions their specific locations might point
As already mentioned, patients with neurological symptoms to distinctive biological roles. In fact, L-ferritin is mainly involved
related to dementia show increased copper concentrations [10,27]. with iron storage, while, on the contrary H-ferritin is associated
In Alzheimer’s-like dementia, copper content in the CNS, is 2-fold with stress responses catalysing Fe2+ /Fe3+ oxidation via the fer-
higher than in age matched controls [10,28,29]. roxidase center. Iron transport within the brain is modulated by
Copper, released at the synaptic cleft by ATP7A acts as a modula- pathways including both transferrin and non-transferrin bound
tor of neurotransmission. Recent studies indicate that copper yields iron. In the brain, iron uptake seems to be controlled by transfer-
biphasic changes in neurotransmission in rat hippocampal neu- rin receptors, expressed exclusively in gray matter, and by ferritin
rons by both blocking and enhancing neurotransmission [30]. It has receptors present only in white matter.
also been proposed that copper release protects against excitotoxic Iron levels in AD and PD are higher than those expected by nor-
cell death by regulating the activities of extra-synaptic N-methyl mal aging but the mechanisms associated with such accumulation
d-aspartate (NMDA) receptors [31,32]. are still elusive [36]. One hypothesis is that Fe accumulation is
correlated with neuronal death through glutamate receptors [37].
2.2. Iron Iron damage can also depend on altered metal transport mech-
anisms [17]. The divalent metal transporter (DMT1) is responsible
Like copper, iron is a redox active metal occurring in two main for iron transport across the BBB. Iron accumulation, induced
oxidation states, +2 and +3. During the last ten years the scien- by neurotoxins such as kainate, 6-hydroxydopamine (6-OHDA)
tific community has made enormous progress to understand iron and MPTP (N-methyl-4-phenyl-1,2,3,6-tetrahydropyridine), might
homeostasis [33], however it is still obscure how the brain regulates depend either on increased iron uptake or release from DMT1 and
fluxes and storage of iron into neurons, oligodentrites, astrocytes ferritin, respectively. Such improper handling of iron is thought
and glial cells. In the healthy state the iron level strictly depends to form a labile iron pool, also called ionic iron, which involves
on specific areas of the brain. Higher iron concentrations are usu- iron in the Fenton reaction with hydrogen peroxide, generated
ally found in the substantia nigra and the globus pallidus [34]. The by dopamine metabolism by monoamine oxidase (MAO), auto-
dentate gyrus, interpeduncular nucleus, thalamus, ventral pallidus, oxidation and other oxidative processes. Ionic iron might also
nucleus basilis and red nucleus also contain relatively high iron determine aggregation of ␣S and A␤, which in turn, can initiate
concentrations. Generally the brain regions responsible for motor OH• generation, causing oxidative stress (Fig. 2).

Fig. 2. Representation of labile iron pool formation and related generation of oxidative stress mechanisms in Parkinson’s disease.
Source: Figure adapted from Ref. [17] with permission.
2132 H. Kozlowski et al. / Coordination Chemistry Reviews 256 (2012) 2129–2141

Fig. 3. Representation of the disruption of synaptic activity by A␤–Zn2+ interaction at the synaptic cleft.
Source: Figure adapted from Ref. [65] with permission.

2.3. Zinc to ZnT expression has been found. Both increased or decreased
zinc levels are measured in serum [54–56], hippocampal [38] and
Zinc is the most abundant trace metal in the brain where it CSF [55] of AD patients compared to age-matched controls. Sim-
plays various functions. It interacts with several proteins confer- ilarly, ZnT expression during AD progression, is both increased
ring on them catalytic or structural properties. As with copper, zinc or decreased when compared to control [57–59]. High concen-
is released into the synaptic cleft, where it modulates NMDA and trations of copper, iron, and zinc are found in amyloid plaques
␣-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid (AMPA) [9,60,61]. Unlike copper and iron, whose toxicities might depend on
receptors regulating the activity of glutamatergic synapses, sug- oxidative-stress-type mechanisms [61,62], zinc plays a crucial role
gesting its possible role as a neurotransmitter in its own right in A␤ aggregation. It can control the rate of self-assembly of A␤ and
[38–43]. regulates the amyloid morphology via specific coordination sites
Several different factors, such as balance between the cellular [63]. Furthermore, zinc can coordinate A␤ via intra-and/or inter-
redox state, the concentrations of other biological chelating agents molecular binding modes, promoting A␤ aggregation [64] and that
and the energy status of the cell are determinant for zinc distribu- synaptic zinc induces the formation of A␤ oligomer with their con-
tion within the cell [44]. sequent accumulation at excitatory synapses (Fig. 3) [65].
Zn homeostasis is controlled by the joint activity of several
Zn transporter families, particularly those involved in zinc ingress 3. Neurodegenerative diseases
and egress from cells and organelles. In particular Zn2+ homeo-
stasis within neuronal cells results from the balanced control The term neurodegenerative diseases refers to the group of
among (i) influx, (ii) buffering and (iii) extrusion processes. Two disorders that share common features and molecular patterns sug-
different families of zinc transporters are known: the ZnT fam- gesting that their pathology may be directly comparable. They are
ily, which decreases intracellular zinc concentrations and the ZIP characterized by abnormal accumulation of proteins and selective
family, which imports the metal ions from the extracellular space neuronal degeneration. Depositions of insoluble, fibrous amyloid
or organellar lumen into the cytoplasm. The zinc ingress into proteins are found either extra- or intracellularly in the brain. The
neurons takes place through voltage-sensitive Ca channels and fibrils consist of various amino acid sequences but all share a ␤-
Ca-permeable ionotropic glutamate receptors as well [45]; on the pleated-sheet secondary structure. This sets down the hallmark of
other hand sequestration and buffering require metallothioneins the conformational disorders like Alzheimer’s disease, Parkinson’s
(MTs), mitochondria, zincosomes and lysosomes activity [46–50]. disease and transmissible spongiform encephalopathies where
The three MTs isoforms expressed in the brain are MT-1, MT-2 and aggregation of specific proteins has been observed. These proteins
MT-3 present in astrocytes and in neurons [51,52]. are usually natively partly unfolded but, after misfolding, aggregate
There is a strong indication that Zn2+ is involved in AD pro- and accumulate as deposits in the brain. A large body of evidence
gression only (for review see [53]). Many investigations have been indicates that transition metal ions like Cu, Zn and Fe have strong
addressed to monitor changes of zinc levels associated with AD impact on the development of neurodegenerative disorders. As a
onset. No consensus on what happens to zinc concentrations and matter of fact, metal binding may result in severe conformational
H. Kozlowski et al. / Coordination Chemistry Reviews 256 (2012) 2129–2141 2133

changes of proteins involved in neurodegeneration. The fact that 3.2. Parkinson’s disease
Cu, Zn and Fe interact with the major protein components is rather
because of the deregulation of the mechanisms that maintain metal Parkinson’s disease is a severe, progressive motor disorder
homeostasis [66]. caused by changes in the central nervous system, tightly linked to
degeneration of dopaminergic neurons of basal ganglia known as
3.1. Alzheimer’s disease substantia nigra. Familiar or inherited forms of disease are linked
to the mutation of genes coding ␣-synuclein, the key PD protein
Alzheimer’s disease is the most common form of dementia that [84]. The histopathological hallmark of disease is the presence of
is normally preceded by up to eight years of mild cognitive impair- the Lewy bodies that contain the fibrous ␣-synuclein deposits.
ment, the transition stage between the disease and the normal Although various biochemical studies suggested that the protein
aging [67]. AD is characterized by the extracellular deposition of may have an impact on numerous processes in the cell, like the
amyloid plaques formed by the aggregated amyloid-␤ peptide that reassembly of the SNARE complex required neurotransmitter
is cleaved from its integral membrane precursor protein (APP) release [85], so far none has shown that it is absolutely essential
and the intracellular deposition of paired helical filaments of the for anything. Hence, the role of protein remains elusive, over
hyperphosphorylated tau protein in neurofibrillary tangles [68]. and above the mechanism leading to its conversion to species
The function of the APP is unknown. However, recent evidence tending to aggregate. The likely factors that could be involved in
suggests that it plays a role in maintaining copper homeostasis, per- the latter process are the transition metal ions, in particular Cu
haps employing its ferroxidase activity [69–71]. The original loci of that interacts with both the soluble and membrane-bound helical
the amyloid deposition within the brain have been determined in form of ␣-synuclein.
the synaptic clefts of glutamatergic synapses of the hippocampus Actually, altered metal homeostasis has been reported in PD.
and cortex. These synapses are involved in long-term potentia- Zinc and iron are augmented, while copper is decreased in substan-
tion, the fundamental process contributing to memory formation. tia nigra. The elevated iron level may contribute to oxidative stress
The latest studies elucidated them as sites of Cu and Zn release and deposition of intracellular inclusion bodies, thus facilitating the
[72–74]. The pool of vesicular zinc formed as an outcome of ZnT3 progress of pathology. Under normal physiological conditions the
protein activity is most likely excreted with glutamate during neu- excess iron is sequestered by ferritin and neuromelanin [86,87].
rotransmission to deteriorate the response of postsynaptic NDMA The reduced expression of the former one and the loss of neurome-
receptors. The Menkes ATPase dependent release of copper from lanin cells in substantia nigra coupled with structure alterations of
postsynaptic neurites results in a series of events employing the the neuromelanin itself may compromise the ability of neurons to
metal related S-nitrosylation of NDMA receptors that modulate sequester the iron and result in elevated level of this redox active
effects on the latter one [74]. The increasing number of recently metal. Iron promotes the aggregation of the key PD protein prob-
published papers suggests that glutamatergic synapse might have ably by distraction of ubiquitin dependent proteasomal system of
unique properties as it may be the site of convergence of chem- protein clearing [88]. The protective role of ferritin and iron chelat-
ically exchangeable pools of Zn and Cu [66]. When these metal ing agents in animal models supports these hypotheses.
ions meet the amyloid-␤, the peptide with its ability to precipi-
tate or crosslink, the resulting series of events engaging interaction 3.3. Prion diseases
of these metal ions with the peptide might either lead to cop-
per related free radical toxicity accompanied by aggregation or to Human prion diseases are a group of fatal neurodegenerative
inhibition of destructive effects leading to oxidative stress by pep- disorders that share a common pathogenic agent, the prion [89].
tide precipitation induced by Zn binding. This hypothesis might There are five known human prion diseases: Creutzfeld–Jakob dis-
be supported by the metal swap experiments applying Zn loaded ease (CJD), Gerstmann Straussler Scheinker disease, kuru, new
metallothionein 3 as a factor that can reduce Cu induced toxicity variant Creutzfeld–Jakob disease and fatal familial insomnia. These
toward cell cultures [75]. Metallothionein 3 is indeed released by diseases are unusual because they can be transferred between
neighboring astrocytes at the glutametergic synaptic cleft [76]. individuals or even between the species. Regardless of etiology,
Zn, Cu and Fe are found in amyloid plaques in relatively high, prion diseases are identified by neuronal vacuolation (spongiosis)
even millimolar or submillimolar concentrations (1 mM, 400 ␮M and deposition of amyloid-like plaques. The diverse prion diseases
and 1 mM, respectively) [9]. Zinc and copper collocate within A␤ share various isoforms of common pathogenic unit known as PrPSc ,
deposits where they seem to be bound, whereas iron does not the abnormal form of host-encoded cellular prion protein (PrPC )
co-purify with the A␤ extracted from plaques and is found pre- [90]. The only differences between them seem to be the monomer
dominantly as a ferritin bound form in neuritic processes associated conformation and the resulting aggregation properties. It has been
with the plaques [77,78]. Interestingly, iron levels are elevated in proposed that prions are self-propagating amyloid forms of cellular
the cortex, whereas copper levels are decreased [17,79]. The former counterparts that replicate by fiber fragmentation that can act as
one might be the consequence of atypical distribution of proteins templates for monomer recruitment [91]. The exact mechanism by
that regulate iron homeostasis including ceruloplasmin, the cop- which PrP can cause neurodegeneration and cell death is elusive.
per dependent protein. For that reason, altered iron homeostasis is The cellular form of prion protein is also widely known as a cop-
rather a secondary than a primary effect [17,80]. per binding protein. The binding of copper is possibly necessary
The interplay of copper and zinc is not limited only to amyloid- for its normal cellular function [92]. One of the primary manifes-
␤ but spreads through all the key Alzheimer’s disease proteins. To tations of this metal ion binding is internalization of the protein
begin with, Cu+ binds to N-terminal extracellular domain of APP from cellular surface. Another consequence of copper binding to
thus promoting its export outside the neuron [81]. In addition, cel- PrP is the plausible neuroprotective role of the holoprotein fulfilled
lular Cu promotes the expression of APP mRNA [82,83]. C-terminal by indirect rather than direct involvement in antioxidant protec-
domain of ␤-secretase binds Cu+ that allows the protein to interact tion of cells. These activities are lost when PrPC is converted to
with CCS1, the copper chaperone that delivers Cu+ to SOD1. Zn2+ PrPSc during disease [93,94]. Studies on other metal ion binding
ions, even at low concentration, negatively modulate the activity to human PrP suggested possible binding of Zn, Ni and Mn [95].
of ␥-secretase, another metalloproteinase directly involved in APP However, low affinity of the protein toward Zn makes these sugges-
processing. However, physiological implications of the majority of tions questionable and biologically irrelevant. The question of Mn
these interactions are unclear. interaction with the protein is still open to discussion. Manganese
2134 H. Kozlowski et al. / Coordination Chemistry Reviews 256 (2012) 2129–2141

Fig. 4. Species distribution diagram for the Cu/A␤16 system at [Cu2+ ]tot = [L]tot Fig. 5. Proposed structure of the mixed metal Cu2+ /Zn2+ complex of A␤ peptide.
= 0.001 M. Eight mono-nuclear, variously protonated complexes are formed in the Only amino acid residues involved in metal complexation are shown for clarity.
pH range 3–11. The concentration of each species is reported as percentage of copper Source: Figure reproduced from Ref. [120] with permission.
involved, as a function of pH.
Source: Diagram drawn using data reported in Ref. [101].

bound protease resistant forms similar to PrPSc are isolated from the capability of A␤ to bind up to four Cu2+ ion, at alkaline pH,
the brains of patients with CJD [96]. This binding might be the con- was also reported [102]. The few available investigations on Cu2+
sequence rather than cause of disease. The potential interaction of complexation with amyloid fibrils suggest a coordination behavior
prion protein with Fe might be neglected, since the protein’s affin- similar to that reported in solution [110], with the proved involve-
ity toward this metal would favor the metal ion sequestration by ment of ␧-nitrogen atoms of His13 and His14, as well as carboxylate
other proteins present in the milieu [92]. side-chains of Glu3, Glu11 and Glu22 [111], in a high variety of
coordination modes based on both intra- and inter-molecular inter-
4. Metal binding to neurodegenerative proteins actions, due to the high concentration of binding sites at the solid
state. The involvement of the C-terminus was also suggested, while
4.1. Alzheimer’s disease the structure of hydrophobic regions of A␤40 does not look to be
affected by copper coordination [111]. Ni2+ , often used as a probe
The amyloid-␤ peptide (A␤), whose aggregation results in the to investigate Cu2+ coordination behavior, forms complex species
formation of plaques typical of AD [97], binds metal ions influencing with A␤ of a structure very similar to that found for Cu2+ [112].
their metabolism. It is widely accepted that the hydrophilic frag- The coordination chemistry of Cu+ with A␤ is of much inter-
ment containing the first 16 residues of the A␤ peptide (A␤16) is est, since Cu2+ can be reduced in the brain to produce ROS species
highly representative of the coordination behavior of the wild pep- through the Fenton reaction. Cu+ coordination is rather different
tide toward metal ions, as also reported in some recent reviews from Cu2+ , as expected: Cu+ ion is coordinated to A␤ by means of
[97–100]. two imidazole rings of His residues in a linear fashion. All three his-
As far as Cu2+ is concerned, solution-equilibria studies on model tidines are involved in binding, with a dynamic exchange between
peptides [101,102] agree with spectroscopic investigations about different ligand combinations [113–115]. The existence of a dra-
the presence of different species at physiological pH (Fig. 4). The matic difference between the coordination mode of Cu+ to A␤
A␤16 fragment contains several potential donor sites: Asp1 (NH2 ), monomers or oligomers has been reported [116]: in the latter case
Asp1 (CO), Asp1 (COO− ), Ala2 (CO), Glu3 (COO− ), His6 (Im), Asp7 tetrahedral geometry is preferred.
(COO− ), Tyr10 (O− ), Glu11 (COO− ), His13 (Im), His14 (Im) and There is a general consensus that the Zn2+ complex species
amidyl from the peptide backbone; also water molecules can be formed at neutral pH have 1:1 stoichiometry with all imidazoles
taken into consideration. Some of these have been definitively involved in coordination [99]. As further binding groups, termi-
excluded from coordination, like e.g. Tyr10 [102,103] even though nal amine, Asp1 or Glu11 carboxylates or water molecules have
this residue plays a role in amyloid aggregation. The main complex been suggested [117], while amide nitrogen atoms, Tyr10 and Arg5
structures around physiological pH have been classified as Form side chains have been excluded. The minimal binding site was rec-
I (prevailing at pH 6.5) and Form II (prevailing at pH higher than ognized as A␤6–14 (HDSGYEVHH), with the region 11–14 playing
8); both of them are distorted square planar, with at least three the main recognition role [118]. Also in the Zn2+ case, solid-state
nitrogen donor-atoms in equatorial position, giving 3N1O or 4N binding-geometries in fibrils are very similar to those described in
coordination modes [104]. Both the terminal amino group and side solution. The possibility of forming mixed-metal Cu2+ /Zn2+ com-
chain imidazoles of His6, His13 and His14 are involved in com- plexes, in the presence of Zn2+ excess, has been recently reported
plexation [105], in many different combinations and in dynamic [119]. In this case, Cu2+ complex geometry would be modified,
equilibrium [106,107]. The involvement of the amidyl nitrogen of being the His13 and His14 sites engaged with Zn2+ , thus reducing
the Asp1 Ala2 peptide bond is likely for Form II [108], possibly in the redox activity of the copper–A␤ complex (Fig. 5).
place of His13 and/or His14, with a consequent change in complex Iron coordination to A␤ has been very recently investigated,
geometry. The identity of the oxygen equatorial donor eventually using NMR spectroscopy [120,121]. The key role in complex-
participating in complex-formation is still debated: the best candi- formation played by the A␤16 fragment has been confirmed for
dates are carboxylates of Asp1 or Asp7, while, in the case of Form the Fe2+ ion. Experimental data suggest that Fe2+ is hexacoordi-
II, the involvement of Ala2 carbonyl has also been suggested [109]. nated, most likely involving the following residues: Asp1 (NH2 , CO
The presence of a second, independent coordination site for Cu2+ and COO− ), Glu3 (COO− ), His6 (Im and CO), His13 (Im), His14(Im),
have also been put forward, although its estimated affinity con- but not Tyr10 nor Met35. The binding-site geometry is quite sim-
stant rules out any possible biological relevance for this site [99]; ilar to Form I of copper (see above), but, different from Cu2+ , no
H. Kozlowski et al. / Coordination Chemistry Reviews 256 (2012) 2129–2141 2135

4.3. Prion diseases

TSEs are spongiform encephalopathies with the peculiarity to be


transmissible: the infectious agent is a mutated form of prion pro-
tein (PrPSc ) which, compared to the normal protein (PrPC ), is highly
insoluble in physiological medium and resistant to proteases, thus
accumulating in neurons as plaques. The exact mechanism of PrPC
post-translational modification to PrPSc is not completely under-
stood. On the other hand it is widely demonstrated that PrPC can
bind metal ions, especially Cu2+ and Zn2+ , with high affinities: the
complex-formation imposes structural constraints on the protein
Fig. 6. Coordination geometry of the primary Cu2+ binding site in ␣S. [132], suggesting that its metal-binding capability is the key for
Source: Figure reproduced from Ref. [122] with permission.
PrPSc formation, if not for the disease development.
PrPC is a 254 amino acid protein: under physiological condi-
tions, its N-terminal domain is unstructured, while the C-terminal
pH dependence was observed around physiological pH. As far as
domain contains three ␣-helices and two ␤-sheets. It is widely
Fe3+ and Al3+ are concerned, no complex species was detected at
accepted that the main Cu2+ binding sites are located in the
physiological pH [121].
N-terminal domain, containing six histidines (61, 69, 77, 85, 96,
111) which can act as metal anchoring sites [133]. A PrPC peculiar-
4.2. Parkinson’s disease ity, highly conserved among mammals and found with some differ-
ences also in birds and fish, is the presence of four repetitions of the
␣-Synuclein (␣S) is the copper-binding protein involved in Lewy same sequence of eight residues (PHGGGWGQ), called “octarepeat”
bodies formation which are distinctive of PD. It consists of 140 and located in the fragment 60–91. Each octarepeat contains one
amino acids and its native monomeric form is unstructured. ␣S His and can bind copper or zinc. The binding mode highly depends
is composed of three distinct domains: (i) the N-terminal region on both pH and/or metal/ligand ratio. In fact, at acidic pH or in the
(1–60) containing two strong anchoring sites for the Cu2+ ion at presence of (sub-)stoichiometric levels of Cu2+ ions, the so-called
Met1 (NH2 ) and His50 (Im); (ii) the so-called non-amyloidogenic “inter-repeat” binding mode is adopted: up to four histidines of
component (NAC), spanning the residues 61–95, and (iii) the neg- this region bind the metal at the same time, with the formation of
atively charged C-terminal domain (96–140), rich in Asp and Glu a 3N or 4N species where imidazole nitrogen atoms act as donor
residues, potential additional metal-binding sites. atoms (Fig. 7a). Increasing pH, the copper ion can displace one or
The binding capability of ␣S toward Cu2+ has recently been more protons from the amide nitrogen atoms of peptidic backbone,
reviewed [100]. It is widely accepted that ␣S can bind at least three thus changing the complex geometry to the “intra-repeat” mode
Cu2+ ions, two of them with micromolar or lower affinity and the (Fig. 7b). This coordination mode is identical to that found both in
third one with millimolar affinity [123]. The two stronger bind- solution and at the solid state with peptide models like Ac-HGGGW-
ing sites are located in the N-terminal domain and correspond to NH2 [134] or Ac-PHGGGWGQ-NH2 [135]; the presence of a Pro
the Met1 (NH2 ) and His50 (Im) anchoring sites [22]; there is no residue before the His anchoring site forces the amide coordination
agreement in the literature on the fact that they are independent toward the C-terminus, forming a rather unusual seven-membered
[124,125] or that the Cu2+ ion can form a macrochelate, binding at chelate ring. Another interesting feature is the stabilizing role of the
the same time both amine and imidazole groups of Met1 and His50 Trp indole ring, interacting with the metal ion through an axially
residues, respectively (Fig. 6) [122,123,126]. The third Cu2+ bind- bound water-molecule. At a sufficiently high copper/PrP ratio, up
ing site is located at the C-terminal; it is weak and unspecific [100]. to four Cu2+ ions can be bound to the PrP octarepeat domain [23].
The presence of different species, characterized by different donor One or even two additional Cu2+ ions can be tethered to PrP
atom sets, has been ascertained at neutral pH, by means of both at the so-called “fifth binding-site”, located in the amyloidogenic
spectroscopic [127] and thermodynamic investigations [127–129]. PrP region of (90–126), where two His residues (96 and 111)
The comparison between ␣S and ␤S behavior confirmed most of are present. Most of the literature agrees that these histidines
the above observations [130]. behave independently [137] and that the preferred anchoring site,
Cu+ coordination to ␣S has recently been investigated [131]: in humans, is His111 [133]. The most accredited coordination
the reduced copper ion can bind ␣S with micromolar affinity, the geometry at physiological pH is [NIm , 2N− , O] for both the sites;
donor-set mainly involving Met1 and Met5 residues. the presence of an equilibrium between different species where

Fig. 7. Inter-repeat (a) and intra-repeat (b) structures of Cu2+ –PrP complexes with the octarepeat domain.
Source: Figure (a) reproduced from Ref. [136] with permission. Figure (b) drawn using crystallographic data of the Cu2+ /Ac-HGGGW-NH2 complex [134], deposited at the
Cambridge Structural Database.
2136 H. Kozlowski et al. / Coordination Chemistry Reviews 256 (2012) 2129–2141

the amide coordination is directed toward either the N- or the C- 5. Protein misfolding
terminus has been suggested [138]. The involvement of Met109
and/or Met112 residues in complexation at the site centered at On the basis of the growing number of pathological processes
His111 is still a matter of debate [139,140]. in which folding abnormalities are involved, it has been proposed
The “inter-repeat” binding mode is uniquely available for Zn2+ that at least half of human diseases are somehow associated with
(with an affinity much lower than copper), since this metal ion protein misfolding [145]. These folding abnormalities are either
is unable to displace amide protons [100]. However, due to the associated with diseases caused by decreased concentrations of a
high zinc concentration that can be present in the brain, Zn2+ can specific protein that never achieves its folded functional structure
influence the Cu2+ distribution in PrP complexes [141,142] in the (lack of function mechanism) or might be due to intra- or extracel-
presence of low copper levels. lular accumulation of insoluble, abnormally aggregated proteins
(gain of function mechanism) [146]. Diseases based on the first
mechanism are distributed over each phase of life and are mostly
4.4. Metal–protein affinity
associated with genetic diseases, while those induced by protein
aggregation are predominantly associated with aging. Neurode-
The classical way that biochemists use to estimate the affinity
generative disorders constitute one of the most common groups
of a ligand (e.g. a metal ion, M) for a substrate (e.g. a protein, P)
of the latter type of protein misfolding diseases, with exception
is the evaluation of the dissociation constant, Kd , referring to the
of the prion diseases that might be elucidated both by “loss of
equilibrium:
function” and “gain of function” mechanisms (vide supra). In most
[M][P] cases large structural rearrangements between the frequently ␣-
MP  M + P Kd = (1) helical or unordered structure of the monomeric native protein and
[MP]
the aggregated material rich in ␤-sheet conformation is observed
where charges are omitted for the sake of simplicity. Kd is a [147]. The initiating event of the mechanism of formation of amy-
conditional constant, usually obtained by experimentally deter- loid fibrils is protein misfolding, which results in the generation of
mining the free metal concentration, [M], when [P] = [MP], at known aggregation-prone structures that grow by an autocatalytic mech-
total concentration of P and under physiological conditions [133]. anism. The critical event in that process, known as nucleation
Whatever experimental technique (most often spectroscopic) is dependent model, generates protein oligomers that act as a seed of
employed to measure [M], this procedure deals with the chem- successive propagation of protein misfolding [148–150]. Aggrega-
ical system, where many complex-formation equilibria are often tion starts when protein exceeds critical concentration. In addition
active at the same time (e.g. see Fig. 4), as the main (if not the to mature fibrils, number of other structures has been described
unique) species forming was the 1:1 complex. Taking into account as a meta-stable intermediate of the protein aggregation process,
this simplification and the differences in the experimental meth- including soluble oligomers, pores, annular structures, spherical
ods employed, in the matrix (usually containing a buffer and/or a micelles and protofibrils. These diverse species with progressive
competitor), in temperature and ionic strength, the high variability degrees of aggregation are present simultaneously in dynamic
of Kd values reported in the literature for a specific system is not equilibrium between each other [151–153].
surprising. In particular, a problem commonly not correctly faced The discovery of A␤ in the senile plaques set the principle for
is the account in which competitor’s action has to be taken; more- the amyloid cascade hypothesis where the imbalance of A␤ pep-
over, if ternary species, containing the competitor, are formed, the tide metabolism followed by aggregation and deposition leads to
Kd value is unreliable [143]. In addition, in some cases it not possi- neuronal death and in consequence to dementia in AD. This idea
ble to classify the “binding sites” of a substrate but it is much better has been further implemented to other neurodegenerative disor-
to distinguish among different “binding modes” [133,144]. A big ders. However, the hypothesis that deposited aggregates are toxic
advantage of this method is the possibility of working with very has been challenged by the results of studies indicating that these
low amounts of high-molecular-weight substrates, like recombi- plaques were found in people without any clinical symptoms of
nant proteins. AD. Moreover, in some animal models of AD and other neurode-
An alternative way to investigate the metal–protein affinity is generative disorders the symptoms have been identified before the
the rigorous determination of the speciation model (stoichiometry occurrence of protein aggregates has been detected. As a conse-
and thermodynamic stability constant of every species present in quence, today the most accepted hypothesis affirms that protein
solution), through the study of simplified model systems, mimick- misfolding and early stages of oligomerization are actual culprits
ing the binding site(s). In this case, the experimental conditions and of neurodegeneration. Perhaps formation of amyloid-like fibrils
procedures are strictly defined and highly reproducible, no buffer or is a protective mechanism of sequestration and isolation of toxic
competitor is normally necessary; in addition, the knowledge of the misfolded intermediates [154], unless aggregates and associated
speciation model allows the evaluation of Kd and the calculation of proteins are in dynamic equilibrium with soluble versions of the
distribution and competition diagrams. Details of calculation pro- protein, thus acting as reservoir of toxic oligomeric species (Fig. 8)
cedure requested to estimate Kd from a given speciation model are [153,155,156].
described elsewhere [133,144]. A drawback here can derive from All of the proteins involved in the most common neurodegener-
the different behavior between the full-length protein and its low ative disorders, amyloid-␤ peptide and hyperphosphorylated tau
molecular-weight models, possibly due to the protein secondary or protein for AD, ␣-synuclein for PD and prion protein for TSEs are
tertiary structure or the action of residues not included in the pep- affected by metals when it comes to their aggregation.
tide model. The former problem can be neglected if the binding site Focusing on AD, protein misfolding associated with amyloid-␤
is located in an unstructured domain of the protein. In summary, aggregation is greatly affected by various biophysical and chemi-
the two methods are complementary and, together, they can give cal factors including metal ions as Cu, Zn, and Fe which have been
a complete picture of the system under study. found in senile plaques in post-mortem AD brains. However, the
The affinities of above proteins, involved in the main neurode- exact role of these metal ions, especially Fe, in ␤-sheet formation
generative disorders, with metal ions, and the resulting biological is rather unclear. Among them Zn2+ is very efficient inductor of
consequences, have been recently reviewed [99,122,128,133,144] fast precipitation of the amyloid peptide leading to formation of
and they will be extensively examined in another chapter of this protease-resistant aggregates, whereas Cu and Fe demonstrate lim-
volume; for these reasons they will not be discussed here. ited ability to induce aggregation in the physiological pH range that
H. Kozlowski et al. / Coordination Chemistry Reviews 256 (2012) 2129–2141 2137

Fig. 8. Two pathways of amyloidogenic protein aggregation.


Source: Figure reproduced from Ref. [157] with permission.

drastically increases in slightly acidic pH [158–162]. The proamy- membrane bound form is believed to have higher propensity for
loidogenic activity of Zn is neuroprotective rather than neurotoxic the formation of toxic oligomeric species, that form pores in the
effect, since Zn attenuates A␤ toxicity in cortical cultures. The membrane of the vesicle [176]. Coordination of Cu2+ could lock
possible mechanism of action might be interrelated with compe- the protein in ␣-helical conformation thus leading to increased
tition with Cu for the peptide binding sites thus reducing the Cu toxicity [177]. Another factor that may stabilize the toxic oligomers
related Fenton type redox chemistry of amyloid-␤ species. In gen- are dopamine-quinones that are the components of the modified
eral, the role of Cu depends on the conditions and type of aggregate form of neuromelanin [178].
[163,164], since some reports suggest the inhibitory effect of Cu on The basic concept of so-called “prion hypothesis” is that mis-
the amyloidogenic processes [158,165,166]. folded prion protein interacts specifically with host PrPC catalyzing
A number of hypothetical mechanisms of A␤-induced neu- its conversion to the pathogenic form of the protein. The true neu-
rotoxicity have been proposed. All of them require the peptide rotoxic species, like in AD and PD is almost certainly represented by
to aggregate, all of them were toxic, despite what the actual the oligomeric intermediate. Available data suggest that neurotoxic
aggregation state was. Although generally speaking, most of the infectious species are represented by the oligomeric intermedi-
aggregation states have been poorly characterized, there is a grow- ates that act as a seeds to bind PrPC and catalyze its conversion
ing interest in the scientific community in oligomers that are into the misfolded form by incorporation into the growing polymer
particularly toxic. Oxidative stress induced by Cu interaction with [179,180].
hallmark peptide or the pool of labile Fe may indeed contribute to At the same time long PrPSc polymers break into smaller pieces
the formation of oligomers through the formation of non-native that cause the increase in the number of effective nuclei to pro-
covalent bonds, e.g. dityrosine [167,168]. mote further conversion of PrPC to its pathogenic form [181]. Metal
The aggregation properties of ␣-synuclein, the key PD protein, ions may influence the equilibrium between different forms of PrP
rely on the N-terminal domain of the protein. The key step is the aggregates. Copper inhibits the potential of the protein to aggregate
transformation of the native protein to the partially folded inter- [182], although in some cases non-specific interactions accelerate
mediate that gains the ␤-sheet content. The residues 8–16 were aggregation [183]. The former hypothesis is supported by the obser-
claimed to be critical to the formation of ␤-sheet structure [169]. vations that critical metal ion increases prion infectivity in animal
However, these observations are in contrast to another study that models [184], since it promotes oligomer formation in other sys-
suggests that residues 64–100 regulate protein aggregation [170]. tems [167,168]. Manganese binding to PrP may generate the form
Current research has suggested that presence of metal ions like Cu of the protein that may act as a seed catalyzing the aggregation,
may promote self-oligomerization of the protein [171]. The actual with no effect of the metal on the aggregation itself [185].
mechanism leading to self-association relies on induction of con- The nucleation dependent mechanism of aggregation provides
formational changes rather than induction of aggregation. Beside a reasonable explanation for the infectious nature of prions. More-
copper, aluminum and iron were assigned the most effective pro- over, it comprises most reliable model elucidating the mechanism
moters of polymerization [172]. Metal ion concentrations required of other neurodegenerative disorders [149,186]. These observa-
to cause the aggregation are well above physiological values sug- tions suggest that protein misfolding has an intrinsic ability to
gesting that the metals themselves are not necessary for the process be transmissible. This raises the question as to whether or not
and promote the aggregation through oxidative stress induction neurodegenerative diseases are transmitted by infection through
[173]. On the contrary other reports indicate that even at physio- a prion-like phenomenon [181].
logical concentration Cu2+ may accelerate the rate of ␣-synuclein Although it is theoretically possible, biological and pharmacoki-
aggregation [123]. netic barriers would prevent some amyloid aggregates, e.g. toxic
The biological activity of ␣-synuclein is feasibly related to the oligomeric species, from acting like prions [150]. It is likely to be
control of vesicular transport of neurotransmitters within the the problem especially for some of the intracellular aggregates like
dopaminergic neuron. This requires the protein to interact with Lewy bodies in PD. The unusual resistance of PrPSc to proteases
the vesicular membranes that influences its structure and pro- may be the key in the efficiency of prions as infectious agents [187].
motes aggregation, where the membrane surface serves as “seed” The other reason for non transmissibility of neurodegenerative dis-
for the polymerization [174,175]. The predominantly ␣-helical eases other than prion diseases is the formation of hyperstable
2138 H. Kozlowski et al. / Coordination Chemistry Reviews 256 (2012) 2129–2141

aggregates that are poor at propagating the misfolding [188]. In the occurrence of preorganization mechanism characterized by
fact fragmentation of aggregates is essential for effective propa- small positive reduction potential (0.3 eV) [205].
gation of prion diseases [189,190]. Despite that, although the key As copper, Fe2+ can generate reactive hydroxyl radicals from
proteins of amyloid-like diseases are less hardy than prion, they the Fenton reaction (Fe2+ + H2 O2 → Fe3+ + OH• + OH− ) [200]. The
seem to spread in an infection-like manner within regions of the hydroxyl radical is highly reactive, its reaction with DNA bases
brain most vulnerable to the disorder [157]. or deoxyribose backbone might produce damaged bases or strand
breaks yielding to deleterious effects [206].
As a matter of fact, iron can stimulate (i) free radical forma-
6. Oxidative stress tion, (ii) increased protein and DNA oxidation in the Alzheimer’s
and Parkinson’s brain, (iii) enhanced lipid peroxidation, and (iv)
All the mentioned neurodegenerative diseases, although decreased level of cytochrome c oxidase and advanced glycation
exhibiting different symptoms and affecting different parts of the end products. In addition the iron-dependent HIF-1 prolyl-4-
brain, share many common features such as impaired mitochon- hydroxylase might be activated by excess of iron in brain such
drial function [66] and increased oxidative damage [191]. to modulate the proteasomal-mediated degradation of HIF. Iron-
Mitochondria play key role in regulating cellular survival and chelating drugs stabilize HIF-1, which, in turn, would transactivate
death [192] and they are an important source of reactive oxygen the expression of established protective genes, including vascular
species (ROS) within most mammalian cells [193–195] such that endothelial growth factor (VEGF), erythropoietin, aldolase and p21
mitochondria damage highly contributes to neurodegeneration. [207].
Free radicals are normally produced in the organism and they Contrary to copper and iron, zinc is a redox inert metal and
can interact with cellular constituents. Such reaction is strongly does not take part in oxidation-reduction reactions. However zinc
dependent on aging yielding activation of cytosolic stress signaling deficiency has been associated with increased levels of oxidative
pathways. The term oxidative stress is usually referred to the level damage including increased lipid, protein and DNA oxidation [208].
of oxidative damage in a cell caused by ROS. They can cause serious Zinc can act as an antioxidant by protecting sulfhydryl groups of
injuries affecting proteins, nucleic acid and lipid membranes. In the proteins against free radical attack and by behaving as antagonists
healthy state ROS toxicity is normally handled (ROS are converted of redox-active transition metals, such as iron and copper [209].
into metabolically nondestructive molecules or are scavenged after In the latter case the redox active metal is pushed off its binding
their formation) and any ROS damage is readily repaired. Once site and it is replaced by an isostructurally similar redox-inactive
triggered, ROS induced damage is often irreversible and creates a metal (e.g. copper replacement by zinc) [210]. Such substitution
positive feed-forward process, such that, when mitochondrial func- allows the redox metal to leave the cells reducing the formation of
tions are impaired by ROS further ROS generation is favored. This hydroxyl radicals.
vicious relationship finally yields oxidative stress and accelerated The beneficial effects of zinc have been also ascribed to the
aging. induction of other species that serves as definitive antioxidants.
The generation of free radicals is strongly related to the involve- To this end, metallothioneins (MTs) seem the most effective pro-
ment of redox-active trace metals. In particular Fe2+ /Fe3+ and teins [211]. Recent studies have shown that the MTs represent a
Cu+ /Cu2+ redox couples modulate the redox state of the cell which connection between cellular zinc and the redox state of the cell
should be strictly maintained to physiological limits. [212]. Under conditions of high oxidative stress, changes in the cel-
Copper induced oxidative stress can occur by means of two dif- lular redox state result in release of zinc from metallothionein as a
ferent mechanisms. The first one is related to ROS generation via result of sulfide/disulfide exchange. Finally it has been shown that
a Fenton-like reaction [196,197]. Cu2+ , in the presence of superox- metallothioneins efficiently prevents deleterious redox activity
ide anion radical or biological reductants, can be reduced to Cu+ associated with copper binding to both ␣S and A␤ [213,214,75,215].
(Cu2+ + O2 − → Cu+ + O2 ) which in turn, catalyzes the formation of
reactive hydroxyl radicals through the decomposition of hydrogen 7. Metal based therapeutic strategies for neurogenerative
peroxide via the Fenton reaction (Cu+ + H2 O2 → Cu2+ + OH• + OH− ) disease
[198–200]. The resulting hydroxyl radical, extremely reactive, can
cause extensive injury to biological molecules resulting in lipid As largely discussed Cu, Fe and Zn play key roles in many
peroxidation and DNA damage. neurodegenerative diseases. Many reports support the use of
Copper modulates oxidative stress by also reducing the levels compounds that modulate metal binding to neurodegenerative
of glutathione [201]. Normally glutathione (GSH) concentration proteins as promising therapeutic strategy for neurodegeneration,
in the cells is in the millimolar range. GSH acts as potent cel- especially for AD. The use of metal chelators results in regulation
lular antioxidant and as substrate for several enzymes removing of metal-induced A␤ aggregation and neurotoxicity both in vitro
ROS. Glutathione is involved in intracellular copper metabolism and in vivo [7,216–218]. These compounds, differentiate from
and it can modulate copper toxicity by directly coordinating the the high affinity metal depleting chelators traditionally used for
metal with high affinity [202]. Moreover the stabilization of copper metal overload disease (e.g. hemochromatosis (Fe) and Wilson’s
reduced state, makes it unavailable for redox cycling. disease) and they are based on selectively targeting the copper
Alteration of copper homeostasis, with the formation of an and zinc bound to A␤. They are usually referred to metal–protein
extensive pool of copper, might result in reduced glutathione lev- attenuating compounds (MPACs) [217]. Among these compounds,
els [203], which in turn may enhance the cytotoxic effect of ROS clioquinol and 8-hydroxiquinoline improved cognition in phase
by favoring the copper catalytic activity and as a consequence ROS II clinical trials [219,220]. Recently, copper and iron chelators
production. have resulted effective in both inhibiting A␤ fibrils aggregation
Cu+ may catalyze free radical oxidation of A␤ and ␣S via the and potentially capable to cross the blood–brain barrier (BBB)
generation of free radicals by the Fenton reaction. The neurotoxi- [221–226]. Similarly a novel prochelator, HLA20A has been devel-
city of Cu2+ –A␤ interaction is correlated with the metal reduction oped for AD targeting. The effective molecule HLA20, released
to Cu+ and subsequent generation of hydrogen peroxide that after HLA20A binding to acetylcholinesterase (AChE), modulates
can be catalytically decomposed to form hydroxyl radical [204]. APP regulation and A␤ reduction and equilibrate metal ions con-
Investigations on electron transfer reactivity of Cu(I)/Cu(II)–A␤ centration in the brain [227]. Finally, metal based drugs (platinum
complexes reports conflicting results, recent observations support and/or ruthenium complexes), inhibiting A␤/metal interactions
H. Kozlowski et al. / Coordination Chemistry Reviews 256 (2012) 2129–2141 2139

by selectively occupying the copper and/or zinc metal binding site [46] J.J. Hwang, S.J. Lee, T.Y. Kim, J.H. Cho, J.Y. Koh, J. Neurosci. 28 (2008) 3114.
show promising properties in affecting metal-induced aggregation [47] L.M. Malaiyandi, O. Vergun, K.E. Dineley, I.J. Reynolds, J. Neurochem. 93 (2005)
1242.
and ROS activity associated with copper binding [228–231]. [48] W. Maret, J. Nutr. 130 (2000) 1455S.
[49] N.E. Saris, K. Niva, FEBS Lett. 356 (1994) 195.
Acknowledgments [50] S.L. Sensi, H.Z. Yin, J.H. Weiss, Eur. J. Neurosci. 12 (2000) 3813.
[51] M. Aschner, FASEB J. 10 (1996) 1129.
[52] Y. Manso, P.A. Adlard, J. Carrasco, M. Vasak, J. Hidalgo, J. Biol. Inorg. Chem. 16
We thank National Science Center (NCN 2011/01/B/ST5/03936) (2011) 1103.
and MIUR (PRIN 2008) for financial support. [53] N.T. Watt, I.J. Whitehouse, N.M. Hooper, Int. J. Alzheimers Dis. 2011 (2010)
971021.
[54] L. Baum, I.H. Chan, S.K. Cheung, W.B. Goggins, V. Mok, L. Lam, V. Leung, E. Hui,
References C. Ng, J. Woo, H.F. Chiu, B.C. Zee, W. Cheng, M.H. Chan, S. Szeto, V. Lui, J. Tsoh,
A.I. Bush, C.W. Lam, T. Kwok, Biometals 23 (2010) 173.
[1] C.J. Frederickson, S.W. Suh, D. Silva, R.B. Thompson, J. Nutr. 130 (2000) 1471S. [55] F.J. Jimenez-Jimenez, J.A. Molina, M.V. Aguilar, I. Meseguer, C.J. Mateos-Vega,
[2] C.E. Outten, T.V. O’Halloran, Science 292 (2001) 2488. M.J. Gonzalez-Munoz, F. de Bustos, A. Martinez-Salio, M. Orti-Pareja, M. Zurdo,
[3] S. Bolognin, L. Messori, P. Zatta, Neuromol. Med. 11 (2009) 223. M.C. Martinez-Para, J. Neural Transm. 105 (1998) 497.
[4] A. Budimir, Acta Pharm. 61 (2011) 1. [56] L.L. Rulon, J.D. Robertson, M.A. Lovell, M.A. Deibel, W.D. Ehmann, W.R. Markes-
[5] R.R. Crichton, D.T. Dexter, R.J. Ward, J. Neural Transm. 118 (2011) 301. ber, Biol. Trace Elem. Res. 75 (2000) 79.
[6] J.A. Duce, A.I. Bush, Prog. Neurobiol. 92 (2010) 1. [57] M.A. Lovell, J. Alzheimers Dis. 16 (2009) 471.
[7] V.B. Kenche, K.J. Barnham, Br. J. Pharmacol. 163 (2011) 211. [58] M.A. Lovell, J.L. Smith, S. Xiong, W.R. Markesbery, Neurotox. Res. 7 (2005) 265.
[8] A. Santner, V.N. Uversky, Metallomics 2 (2010) 378. [59] J.L. Smith, S. Xiong, W.R. Markesbery, M.A. Lovell, Neuroscience 140 (2006)
[9] M.A. Lovell, J.D. Robertson, W.J. Teesdale, J.L. Campbell, W.R. Markesbery, J. 879.
Neurol. Sci. 158 (1998) 47. [60] J. Dong, C.S. Atwood, V.E. Anderson, S.L. Siedlak, M.A. Smith, G. Perry, P.R.
[10] I. Hozumi, T. Hasegawa, A. Honda, K. Ozawa, Y. Hayashi, K. Hashimoto, M. Carey, Biochemistry 42 (2003) 2768.
Yamada, A. Koumura, T. Sakurai, A. Kimura, Y. Tanaka, M. Satoh, T. Inuzuka, J. [61] M.A. Smith, P.L. Harris, L.M. Sayre, G. Perry, Proc. Natl. Acad. Sci. U.S.A. 94
Neurol. Sci. 303 (2011) 95. (1997) 9866.
[11] J. Stankiewicz, S.S. Panter, M. Neema, A. Arora, C.E. Batt, R. Bakshi, Neurother- [62] D.A. Butterfield, T. Reed, S.F. Newman, R. Sultana, Free Radic. Biol. Med. 43
apeutics 4 (2007) 371. (2007) 658.
[12] E.D. Walter, D.J. Stevens, A.R. Spevacek, M.P. Visconte, A. Dei Rossi, G.L. Mill- [63] J. Dong, J.E. Shokes, R.A. Scott, D.G. Lynn, J. Am. Chem. Soc. 128 (2006) 3540.
hauser, Curr. Protein Pept. Sci. 10 (2009) 529. [64] Y. Miller, B. Ma, R. Nussinov, Proc. Natl. Acad. Sci. U.S.A. 107 (2010) 9490.
[13] L. Varela-Nallar, A. Gonzalez, N.C. Inestrosa, Curr. Pharm. Des. 12 (2006) 2587. [65] A. Deshpande, H. Kawai, R. Metherate, C.G. Glabe, J. Busciglio, J. Neurosci. 29
[14] V. Desai, S.G. Kaler, Am. J. Clin. Nutr. 88 (2008) 855S. (2009) 4004.
[15] S. Rivera-Mancia, I. Perez-Neri, C. Rios, L. Tristan-Lopez, L. Rivera-Espinosa, S. [66] K.J. Barnham, A.I. Bush, Curr. Opin. Chem. Biol. 12 (2008) 222.
Montes, Chem. Biol. Interact. 186 (2010) 184. [67] M.T. Fodero-Tavoletti, V.L. Villemagne, C.C. Rowe, C.L. Masters, K.J. Barnham,
[16] A.R. White, P. Faller, C.S. Atwood, P. Zatta, Int. J. Alzheimers Dis. 2011 (2011) R. Cappai, Int. J. Biochem. Cell Biol. 43 (2011) 1247.
659424. [68] M.P. Mattson, Nature 430 (2004) 631.
[17] L. Zecca, M.B. Youdim, P. Riederer, J.R. Connor, R.R. Crichton, Nat. Rev. Neu- [69] P.A. Adlard, A.I. Bush, J. Alzheimers Dis. 10 (2006) 145.
rosci. 5 (2004) 863. [70] J.A. Duce, A. Tsatsanis, M.A. Cater, S.A. James, E. Robb, K. Wikhe, S.L. Leong,
[18] I. Bertini, Biological Inorganic Chemistry: Structure and Reactivity, University K. Perez, T. Johanssen, M.A. Greenough, H.H. Cho, D. Galatis, R.D. Moir, C.L.
Science Books, Sausalito, CA, 2007. Masters, C. McLean, R.E. Tanzi, R. Cappai, K.J. Barnham, G.D. Ciccotosto, J.T.
[19] E. Gaggelli, F. Bernardi, E. Molteni, R. Pogni, D. Valensin, G. Valensin, M. Rogers, A.I. Bush, Cell 142 (2010) 857.
Remelli, M. Luczkowski, H. Kozlowski, J. Am. Chem. Soc. 127 (2005) 996. [71] G.K.W. Kong, J.J. Adams, H.H. Harris, J.F. Boas, C.C. Curtain, D. Galatis, C.L. Mas-
[20] E. Gaggelli, Z. Grzonka, H. Kozlowski, C. Migliorini, E. Molteni, D. Valensin, G. ters, K.J. Barnham, W.J. McKinstry, R. Cappai, M.W. Parker, J. Mol. Biol. 367
Valensin, Chem. Commun. (Camb.) (2008) 341. (2007) 148.
[21] E. Gaggelli, E. Jankowska, H. Kozlowski, A. Marcinkowska, C. Migliorini, P. [72] G. Danscher, M. Stoltenberg, J. Histochem. Cytochem. 53 (2005) 141.
Stanczak, D. Valensin, G. Valensin, J. Phys. Chem. B 112 (2008) 15140. [73] C.J. Frederickson, L.J. Giblin, R.V. Balaji, R. Masalha, Y.P. Zeng, E.V. Lopez, J.Y.
[22] D. Valensin, F. Camponeschi, M. Luczkowski, M.C. Baratto, M. Remelli, G. Koh, U. Chorin, L. Besser, M. Hershfinkel, Y. Li, R.B. Thompson, A. Krezel, J.
Valensin, H. Kozlowski, Metallomics 3 (2011) 292. Neurosci. Methods 158 (2006) 169.
[23] D. Valensin, M. Luczkowski, F.M. Mancini, A. Legowska, E. Gaggelli, G. Valensin, [74] M.L. Schlief, A.M. Craig, J.D. Gitlin, J. Neurosci. 25 (2005) 239.
K. Rolka, H. Kozlowski, Dalton Trans. (2004) 1284. [75] G. Meloni, V. Sonois, T. Delaine, L. Guilloreau, A. Gillet, J. Teissie, P. Faller, M.
[24] S. Lutsenko, A. Bhattacharjee, A.L. Hubbard, Metallomics 2 (2010) 596. Vasak, Nat. Chem. Biol. 4 (2008) 366.
[25] H. Tapiero, D.M. Townsend, K.D. Tew, Biomed. Pharmacother. 57 (2003) 386. [76] Y. Uchida, F. Gomi, T. Masumizu, Y. Miura, J. Biol. Chem. 277 (2002) 32353.
[26] J. Dobrowolska, M. Dehnhardt, A. Matusch, M. Zoriy, N. Palomero-Gallagher, [77] I. Grundke-Iqbal, J. Fleming, Y.C. Tung, H. Lassmann, K. Iqbal, J.G. Joshi, Acta
P. Koscielniak, K. Zilles, J.S. Becker, Talanta 74 (2008) 717. Neuropathol. 81 (1990) 105.
[27] V. Nischwitz, A. Berthele, B. Michalke, Anal. Chim. Acta 627 (2008) 258. [78] C. Quintana, S. Bellefqih, J.Y. Laval, J.L. Guerquin-Kern, T.D. Wu, J. Avila, I.
[28] H. Basun, L.G. Forssell, L. Wetterberg, B. Winblad, J. Neural. Transm. Park. Dis. Ferrer, R. Arranz, C. Patino, J. Struct. Biol. 153 (2006) 42.
Dement. Sect. 3 (1991) 231. [79] D. Religa, D. Strozyk, R.A. Cherny, I. Volitakis, V. Haroutunian, B. Winblad, J.
[29] C.O. Hershey, L.A. Hershey, A. Varnes, S.D. Vibhakar, P. Lavin, W.H. Strain, Naslund, A.I. Bush, Neurology 67 (2006) 69.
Neurology 33 (1983) 1350. [80] D.G. Smith, R. Cappai, K.J. Barnham, Biochim. Biophys. Acta: Biomembr. 1768
[30] C. Peters, B. Munoz, F.J. Sepulveda, J. Urrutia, M. Quiroz, S. Luza, G.V. De Ferrari, (2007) 1976.
L.G. Aguayo, C. Opazo, J. Neurochem. 119 (2011) 78. [81] S.A. Bellingham, G.D. Ciccotosto, B.E. Needham, L.R. Fodero, A.R. White, C.L.
[31] M.L. Schlief, J.D. Gitlin, Mol. Neurobiol. 33 (2006) 81. Masters, R. Cappai, J. Camakaris, J. Neurochem. 91 (2004) 423.
[32] M.L. Schlief, T. West, A.M. Craig, D.M. Holtzman, J.D. Gitlin, Proc. Natl. Acad. [82] A.D. Armendariz, M. Gonzalez, A.V. Loguinov, C.D. Vulpe, Physiol. Genomics
Sci. U.S.A. 103 (2006) 14919. 20 (2004) 45.
[33] R.R. Crichton, J.R. Boelaert, Inorganic Biochemistry of Iron Metabolism: From [83] S.A. Bellingham, D.K. Lahiri, B. Maloney, S. La Fontaine, G. Multhaup, J.
Molecular Mechanisms to Clinical Consequences, 2nd ed., Wiley, Chichester, Camakaris, J. Biol. Chem. 279 (2004) 20378.
NY, 2001. [84] D.R. Brown, FEBS J. 274 (2007) 3766.
[34] M.E. Gotz, K. Double, M. Gerlach, M.B. Youdim, P. Riederer, Ann. N.Y. Acad. Sci. [85] J. Burre, M. Sharma, T. Tsetsenis, V. Buchman, M.R. Etherton, T.C. Sudhof,
1012 (2004) 193. Science 329 (2010) 1663.
[35] W.A. Cass, R. Grondin, A.H. Andersen, Z. Zhang, P.A. Hardy, L.K. Hussey- [86] J. Sian-Hulsmann, S. Mandel, M.B.H. Youdim, P. Riederer, J. Neurochem. 118
Andersen, W.S. Rayens, G.A. Gerhardt, D.M. Gash, Neurobiol. Aging 28 (2007) (2011) 939.
258. [87] F. Tribl, E. Asan, T. Arzberger, T. Tatschner, E. Langenfeld, H.E. Meyer, G. Bring-
[36] D. Berg, M.B. Youdim, Top. Magn. Reson. Imaging 17 (2006) 5. mann, P. Riederer, M. Gerlach, K. Marcus, Mol. Cell. Proteomics 8 (2009) 1832.
[37] J.H. Cheah, S.F. Kim, L.D. Hester, K.W. Clancy, S.E. Patterson 3rd, V. Papadopou- [88] K.S. McNaught, R. Belizaire, O. Isacson, P. Jenner, C.W. Olanow, Exp. Neurol.
los, S.H. Snyder, Neuron 51 (2006) 431. 179 (2003) 38.
[38] G. Danscher, K.B. Jensen, C.J. Frederickson, K. Kemp, A. Andreasen, S. Juhl, M. [89] S.B. Prusiner, Proc. Natl. Acad. Sci. U.S.A. 95 (1998) 13363.
Stoltenberg, R. Ravid, J. Neurosci. Methods 76 (1997) 53. [90] S.B. Prusiner, Science 216 (1982) 136.
[39] C.J. Frederickson, A.I. Bush, Biometals 14 (2001) 353. [91] J. Collinge, A.R. Clarke, Science 318 (2007) 930.
[40] C.J. Frederickson, G. Danscher, Prog. Brain Res. 83 (1990) 71. [92] D.R. Brown, Metallomics 3 (2011) 229.
[41] J.Y. Koh, J. Clin. Neurol. 1 (2005) 121. [93] D.R. Brown, Neurobiol. Dis. 15 (2004) 534.
[42] L. Slomianka, G. Danscher, C.J. Frederickson, Neuroscience 38 (1990) 843. [94] S.J. Collins, V.A. Lawson, C.L. Masters, Lancet 363 (2004) 51.
[43] Y. Tian, Z. Yang, T. Zhang, Neurosci. Res. 68 (2010) 167. [95] D.R. Brown, F. Hafiz, L.L. Glasssmith, B.S. Wong, I.M. Jones, C. Clive, S.J. Haswell,
[44] C. Jacob, W. Maret, B.L. Vallee, Proc. Natl. Acad. Sci. U.S.A. 95 (1998) 3489. EMBO J. 19 (2000) 1180.
[45] S.L. Sensi, H.Z. Yin, S.G. Carriedo, S.S. Rao, J.H. Weiss, Proc. Natl. Acad. Sci. U.S.A. [96] B.S. Wong, S.G. Chen, M. Colucci, Z.L. Xie, T. Pan, T. Liu, R.L. Li, P. Gambetti,
96 (1999) 2414. M.S. Sy, D.R. Brown, J. Neurochem. 78 (2001) 1400.
2140 H. Kozlowski et al. / Coordination Chemistry Reviews 256 (2012) 2129–2141

[97] P. Faller, ChemBioChem 10 (2009) 2837. [144] G. Arena, G. Pappalardo, I. Sovago, E. Rizzarelli, Coord. Chem. Rev. 256 (2012)
[98] S.C. Drew, K.J. Barnham, Acc. Chem. Res. 44 (2011) 1146. 3.
[99] P. Faller, C. Hureau, Dalton Trans. (2009) 1080. [145] J. Bradbury, Lancet 361 (2003) 1194.
[100] C. Migliorini, E. Porciatti, M. Luczkowski, D. Valensin, Coord. Chem. Rev. 256 [146] F. Chiti, C.M. Dobson, Annu. Rev. Biochem. 75 (2006) 333.
(2012) 352. [147] O.S. Makin, L.C. Serpell, Biochem. Soc. Trans. 30 (2002) 521.
[101] T. Kowalik-Jankowska, M. Ruta, K. Wisniewska, L. Lankiewicz, J. Inorg. [148] D.C. Gajdusek, in: J. Bjornsson, R.I. Carp, A. Love, H.M. Wisniewski (Eds.), Slow
Biochem. 95 (2003) 270. Infections of the Central Nervous System: The Legacy of Dr Bjorn Sigurdsson,
[102] C.A. Damante, K. Osz, Z. Nagy, G. Pappalardo, G. Grasso, G. Impellizzeri, E. 1994, p. 173.
Rizzarelli, I. Sovago, Inorg. Chem. 47 (2008) 9669. [149] J.D. Harper, P.T. Lansbury, Annu. Rev. Biochem. 66 (1997) 385.
[103] S.C. Drew, C.J. Noble, C.L. Masters, G.R. Hanson, K.J. Barnham, J. Am. Chem. Soc. [150] C. Soto, L. Estrada, J. Castilla, Trends Biochem. Sci. 31 (2006) 150.
131 (2009) 1195. [151] B. Caughey, P.T. Lansbury, Annu. Rev. Neurosci. 26 (2003) 267.
[104] P. Dorlet, S. Gambarelli, P. Faller, C. Hureau, Angew. Chem. Int. Ed. 48 (2009) [152] C.G. Glabe, R. Kayed, Neurology 66 (2006) S74.
9273. [153] D.B. Teplow, N.D. Lazo, G. Bitan, S. Bernstein, T. Wyttenbach, M.T. Bowers,
[105] B.-k. Shin, S. Saxena, J. Phys. Chem. A 115 (2011) 9590. A. Baumketner, J.E. Shea, B. Urbanc, L. Cruz, J. Borreguero, H.E. Stanley, Acc.
[106] J. Ali-Torres, J.-D. Marechal, L. Rodriguez-Santiago, M. Sodupe, J. Am. Chem. Chem. Res. 39 (2006) 635.
Soc. 133 (2011) 15008. [154] P.T. Lansbury, H.A. Lashuel, Nature 443 (2006) 774.
[107] L.A. Hong, T.M. Carducci, W.D. Bush, C.G. Dudzik, G.L. Millhauser, J.D. Simon, [155] C. Haass, D.J. Selkoe, Nat. Rev. Mol. Cell Biol. 8 (2007) 101.
J. Phys. Chem. B 114 (2010) 11261. [156] S. Kim, E.A.A. Nollen, K. Kitagawa, V.P. Bindokas, R.I. Morimoto, Nat. Cell Biol.
[108] C. Hureau, Y. Coppel, P. Dorlet, P.L. Solari, S. Sayen, E. Guillon, L. Sabater, P. 4 (2002) 826.
Faller, Angew. Chem. Int. Ed. 48 (2009) 9522. [157] J. Schnabel, Nature 475 (2011) S12.
[109] S.C. Drew, C.L. Masters, K.J. Barnham, J. Am. Chem. Soc. 131 (2009) 8760. [158] C.S. Atwood, R.D. Moir, X.D. Huang, R.C. Scarpa, N.M.E. Bacarra, D.M. Romano,
[110] C.J. Sarell, C.D. Syme, S.E.J. Rigby, J.H. Viles, Biochemistry 48 (2009) 4388. M.K. Hartshorn, R.E. Tanzi, A.I. Bush, J. Biol. Chem. 273 (1998) 12817.
[111] S. Parthasarathy, F. Long, Y. Miller, Y. Xiao, D. McElheny, K. Thurber, B. Ma, R. [159] C.S. Atwood, R.C. Scarpa, X.D. Huang, R.D. Moir, W.D. Jones, D.P. Fairlie, R.E.
Nussinov, Y. Ishii, J. Am. Chem. Soc. 133 (2011) 3390. Tanzi, A.I. Bush, J. Neurochem. 75 (2000) 1219.
[112] E. Jozsa, K. Osz, C. Kallay, P. de Bona, C.A. Damante, G. Pappalardo, E. Rizzarelli, [160] A.I. Bush, Trends Neurosci. 26 (2003) 207.
I. Sovago, Dalton Trans. 39 (2010) 7046. [161] X.D. Huang, M.P. Cuajungco, C.S. Atwood, M.A. Hartshorn, J.D.A. Tyndall, G.R.
[113] C. Hureau, V. Balland, Y. Coppel, P.L. Solari, E. Fonda, P. Faller, J. Biol. Inorg. Hanson, K.C. Stokes, M. Leopold, G. Multhaup, L.E. Goldstein, R.C. Scarpa, A.J.
Chem. 14 (2009) 995. Saunders, J. Lim, R.D. Moir, C. Glabe, E.F. Bowden, C.L. Masters, D.P. Fairlie, R.E.
[114] S. Furlan, C. Hureau, P. Faller, G. La Penna, J. Phys. Chem. B 114 (2010) 15119. Tanzi, A.I. Bush, J. Biol. Chem. 274 (1999) 37111.
[115] Y. Lu, M. Prudent, L.A. Qiao, M.A. Mendez, H.H. Girault, Metallomics 2 (2010) [162] P.W. Mantyh, J.R. Ghilardi, S. Rogers, E. Demaster, C.J. Allen, E.R. Stimson, J.E.
474. Maggio, J. Neurochem. 61 (1993) 1171.
[116] J. Shearer, P.E. Callan, T. Tran, V.A. Szalai, Chem. Commun. 46 (2010) [163] J. Dong, J.M. Canfield, A.K. Mehta, J.E. Shokes, B. Tian, W.S. Childers, J.A. Sim-
9137. mons, Z. Mao, R.A. Scott, K. Warncke, D.G. Lynn, Proc. Natl. Acad. Sci. U.S.A.
[117] E. Gaggelli, A. Janicka-Klos, E. Jankowska, H. Kozlowski, C. Migliorini, E. 104 (2007) 13313.
Molteni, D. Valensin, G. Valensin, E. Wieczerzak, J. Phys. Chem. B 112 (2008) [164] J.T. Pedersen, J. Ostergaard, N. Rozlosnik, B. Gammelgaard, N.H.H. Heegaard,
100. J. Biol. Chem. 286 (2011) 26952.
[118] P.O. Tsvetkov, A.A. Kulikova, A.V. Golovin, Y.V. Tkachev, A.I. Archakov, S.A. [165] B. Raman, T. Ban, K. Yamaguchi, M. Sakai, T. Kawai, H. Naiki, Y. Goto, J. Biol.
Kozin, A.A. Makarov, Biophys. J. 99 (2010) L84. Chem. 280 (2005) 16157.
[119] C.A. Damante, K. Osz, Z. Nagy, G. Grasso, G. Pappalardo, E. Rizzarelli, I. Sovago, [166] J. Zou, K. Kajita, N. Sugimoto, Angew. Chem. Int. Ed. 40 (2001) 2274.
Inorg. Chem. 50 (2011) 5342. [167] C.S. Atwood, G. Perry, H. Zeng, Y. Kato, W.D. Jones, K.Q. Ling, X.D. Huang, R.D.
[120] F. Bousejra-ElGarah, C. Bijani, Y. Coppel, P. Faller, C. Hureau, Inorg. Chem. 50 Moir, D.D. Wang, L.M. Sayre, M.A. Smith, S.G. Chen, A.I. Bush, Biochemistry 43
(2011) 9024. (2004) 560.
[121] D. Valensin, C. Migliorini, G. Valensin, E. Gaggelli, G. La Penna, H. Kozlowski, [168] K.J. Barnham, F. Haeffner, G.D. Ciccotosto, C.C. Curtain, D. Tew, C. Mavros, K.
C. Gabbiani, L. Messori, Inorg. Chem. 50 (2011) 6865. Beyreuther, D. Carrington, C.L. Masters, R.A. Cherny, R. Cappai, A.I. Bush, FASEB
[122] C.G. Dudzik, E.D. Walter, G.L. Millhauser, Biochemistry 50 (2011) 1771. J. 18 (2004) 1427.
[123] R.M. Rasia, C.W. Bertoncini, D. Marsh, W. Hoyer, D. Cherny, M. Zweckstetter, C. [169] A.M. Bodles, D.J.S. Guthrie, P. Harriott, P. Campbell, G.B. Irvine, Eur. J. Biochem.
Griesinger, T.M. Jovin, C.O. Fernandez, Proc. Natl. Acad. Sci. U.S.A. 102 (2005) 267 (2000) 2186.
4294. [170] O.M.A. El-Agnaf, K.E. Paleologou, B. Greer, A.M. Abogrein, J.E. King, S.A. Salem,
[124] J.C. Lee, H.B. Gray, J.R. Winkler, J. Am. Chem. Soc. 130 (2008) 6898. N.J. Fullwood, F.E. Benson, R. Hewitt, K.J. Ford, F.L. Martin, P. Harriot, M.R.
[125] Y.H. Sung, C. Rospigliosi, D. Eliezer, Biochim. Biophys. Acta: Proteins Pro- Cookson, D. Allsop, FASEB J. 18 (2004) 1315.
teomics 1764 (2006) 5. [171] S.R. Paik, H.J. Shin, J.H. Lee, C.S. Chang, J. Kim, Biochem. J. 340 (1999) 821.
[126] M. Bortolus, M. Bisaglia, A. Zoleo, M. Fittipaldi, M. Benfatto, L. Bubacco, A.L. [172] V.N. Uversky, J. Li, A.L. Fink, J. Biol. Chem. 276 (2001) 10737.
Maniero, J. Am. Chem. Soc. 132 (2010) 18057. [173] G. Yamin, C.B. Glaser, V.N. Uversky, A.L. Fink, J. Biol. Chem. 278 (2003) 27630.
[127] S.C. Drew, S.L. Leong, C.L.L. Pham, D.J. Tew, C.L. Masters, L.A. Miles, R. Cappai, [174] H.J. Lee, C. Choi, S.J. Lee, J. Biol. Chem. 277 (2002) 671.
K.J. Barnham, J. Am. Chem. Soc. 130 (2008) 7766. [175] M. Zhu, J. Li, A.L. Fink, J. Biol. Chem. 278 (2003) 40186.
[128] L.A. Hong, J.D. Simon, Metallomics 3 (2011) 262. [176] M.J. Volles, P.T. Lansbury, Biochemistry 41 (2002) 4595.
[129] T. Kowalik-Jankowska, A. Rajewska, E. Jankowska, Z. Grzonka, Dalton Trans. [177] H.R. Lucas, J.C. Lee, Metallomics 3 (2011) 280.
(2006) 5068. [178] M.J. Volles, S.J. Lee, J.C. Rochet, M.D. Shtilerman, T.T. Ding, J.C. Kessler, P.T.
[130] A. Binolfi, G.R. Lamberto, R. Duran, L. Quintanar, C.W. Bertoncini, J.M. Souza, Lansbury, Biochemistry 40 (2001) 7812.
C. Cervenansky, M. Zweckstetter, C. Griesinger, C.O. Fernandez, J. Am. Chem. [179] B. Caughey, Br. Med. Bull. 66 (2003) 109.
Soc. 130 (2008) 11801. [180] C. Soto, G.P. Saborio, L. Anderes, Trends Neurosci. 25 (2002) 390.
[131] A. Binolfi, A.A. Valiente-Gabioud, R. Duran, M. Zweckstetter, C. Griesinger, C.O. [181] C. Soto, L.D. Estrada, Arch. Neurol. 65 (2008) 184.
Fernandez, J. Am. Chem. Soc. 133 (2011) 194. [182] O.V. Bocharova, L. Breydo, V.V. Salnikov, I.V. Baskakov, Biochemistry 44 (2005)
[132] N.D. Younan, M. Klewpatinond, P. Davies, A.V. Ruban, D.R. Brown, J.H. Viles, J. 6776.
Mol. Biol. 410 (2011) 369. [183] E. Quaglio, R. Chiesa, D.A. Harris, J. Biol. Chem. 276 (2001) 11432.
[133] H. Kozlowski, M. Luczkowski, M. Remelli, Dalton Trans. 39 (2010) 6371. [184] D. McKenzie, J. Bartz, J. Mirwald, D. Olander, R. Marsh, J. Aiken, J. Biol. Chem.
[134] C.S. Burns, E. Aronoff-Spencer, C.M. Dunham, P. Lario, N.I. Avdievich, W.E. 273 (1998) 25545.
Antholine, M.M. Olmstead, A. Vrielink, G.J. Gerfen, J. Peisach, W.G. Scott, G.L. [185] M.W. Brazier, P. Davies, E. Player, F. Marken, J.H. Viles, D.R. Brown, J. Biol.
Millhauser, Biochemistry 41 (2002) 3991. Chem. 283 (2008) 12831.
[135] G.L. Millhauser, Annu. Rev. Phys. Chem. 58 (2007) 299. [186] M.R.H. Krebs, L.A. Morozova-Roche, K. Daniel, C.V. Robinson, C.M. Dobson,
[136] H. Kozlowski, A. Janicka-Klos, P. Stanczak, D. Valensin, G. Valensin, K. Kulon, Protein Sci. 13 (2004) 1933.
Coord. Chem. Rev. 252 (2008) 1069. [187] C. Soto, G.P. Saborio, Trends Mol. Med. 7 (2001) 109.
[137] M. Remelli, D. Valensin, D. Bacco, E. Gralka, R. Guerrini, C. Migliorini, H. [188] J.R. Silveira, G.J. Raymond, A.G. Hughson, R.E. Race, V.L. Sim, S.F. Hayes, B.
Kozlowski, New J. Chem. 33 (2009) 2300. Caughey, Nature 437 (2005) 257.
[138] E. Gralka, D. Valensin, E. Porciatti, C. Gajda, E. Gaggelli, G. Valensin, W. Kamysz, [189] D.S. Kryndushkin, I.M. Alexandrov, M.D. Ter-Avanesyan, V.V. Kushnirov, J. Biol.
R. Nadolny, R. Guerrini, D. Bacco, M. Remelli, H. Kozlowski, Dalton Trans. Chem. 278 (2003) 49636.
(2008) 5207. [190] G.P. Saborio, B. Permanne, C. Soto, Nature 411 (2001) 810.
[139] M. Remelli, D. Valensin, L. Toso, E. Gralka, R. Guerrini, E. Marzola, H. Kozłowski, [191] B. Halliwell, Drugs Aging 18 (2001) 685.
Metallomics, submitted. [192] N.N. Danial, S.J. Korsmeyer, Cell 116 (2004) 205.
[140] L. Rivillas-Acevedo, R. Grande-Aztatzi, I. Lomeli, J.E. Garcia, E. Barrios, S. [193] V. Adam-Vizi, C. Chinopoulos, Trends Pharmacol. Sci. 27 (2006) 639.
Teloxa, A. Vela, L. Quintanar, Inorg. Chem. 50 (2011) 1956. [194] R.S. Balaban, S. Nemoto, T. Finkel, Cell 120 (2005) 483.
[141] F. Stellato, A. Spevacek, O. Proux, V. Minicozzi, G. Millhauser, S. Morante, Eur. [195] E. Cadenas, K.J. Davies, Free Radic. Biol. Med. 29 (2000) 222.
Biophys. J. Biophys. Lett. 40 (2011) 92. [196] S.I. Liochev, I. Fridovich, Free Radic. Biol. Med. 16 (1994) 29.
[142] E.D. Walter, D.J. Stevens, M.P. Visconte, G.L. Millhauser, J. Am. Chem. Soc. 129 [197] J. Prousek, Pure Appl. Chem. 79 (2007) 2325.
(2007) 15440. [198] O.I. Aruoma, B. Halliwell, E. Gajewski, M. Dizdaroglu, Biochem. J. 273 (Pt 3)
[143] V. Tougu, A. Tiiman, P. Palumaa, Metallomics 3 (2011) 250. (1991) 601.
H. Kozlowski et al. / Coordination Chemistry Reviews 256 (2012) 2129–2141 2141

[199] K. Barbusinski, Ecol. Chem. Eng. S 16 (2009) 347. [219] P.A. Adlard, R.A. Cherny, D.I. Finkelstein, E. Gautier, E. Robb, M. Cortes, I.
[200] J. Prousek, Chem. Listy 89 (1995) 11. Volitakis, X. Liu, J.P. Smith, K. Perez, K. Laughton, Q.X. Li, S.A. Charman, J.A.
[201] H. Speisky, M. Gomez, F. Burgos-Bravo, C. Lopez-Alarcon, C. Jullian, C. Olea- Nicolazzo, S. Wilkins, K. Deleva, T. Lynch, G. Kok, C.W. Ritchie, R.E. Tanzi, R.
Azar, M.E. Aliaga, Bioorg. Med. Chem. 17 (2009) 1803. Cappai, C.L. Masters, K.J. Barnham, A.I. Bush, Neuron 59 (2008) 43.
[202] M.D. Mattie, J.H. Freedman, Am. J. Physiol. Cell Physiol. 286 (2004) C293. [220] R.A. Cherny, C.S. Atwood, M.E. Xilinas, D.N. Gray, W.D. Jones, C.A. McLean,
[203] M.C. Linder, C.A. Goode, Biochemistry of Copper, Plenum Press, New York, K.J. Barnham, I. Volitakis, F.W. Fraser, Y.S. Kim, X.D. Huang, L.E. Goldstein,
1991. R.D. Moir, J.T. Lim, K. Beyreuther, H. Zheng, R.E. Tanzi, C.L. Masters, A.I. Bush,
[204] M.P. Cuajungco, L.E. Goldstein, A. Nunomura, M.A. Smith, J.T. Lim, C.S. Atwood, Neuron 30 (2001) 665.
X. Huang, Y.W. Farrag, G. Perry, A.I. Bush, J. Biol. Chem. 275 (2000) 19439. [221] A.S. DeToma, S. Salamekh, A. Ramamoorthy, M.H. Lim, Chem. Soc. Rev. 41
[205] V. Balland, C. Hureau, J.M. Saveant, Proc. Natl. Acad. Sci. U.S.A. 107 (2010) (2012) 608.
17113. [222] R. Jakob-Roetne, H. Jacobsen, Angew. Chem. Int. Ed. 48 (2009) 3030.
[206] O. Kakhlon, Z.I. Cabantchik, Free Radic. Biol. Med. 33 (2002) 1037. [223] L.R. Perez, K.J. Franz, Dalton Trans. 39 (2010) 2177.
[207] T. Amit, Y. Avramovich-Tirosh, M.B. Youdim, S. Mandel, FASEB J. 22 (2008) [224] A. Rauk, Chem. Soc. Rev. 38 (2009) 2698.
1296. [225] L.E. Scott, C. Orvig, Chem. Rev. 109 (2009) 4885.
[208] A.S. Prasad, Curr. Opin. Clin. Nutr. Metab. Care 12 (2009) 646. [226] P. Zatta, D. Drago, S. Bolognin, S.L. Sensi, Trends Pharmacol. Sci. 30 (2009) 346.
[209] T.M. Bray, W.J. Bettger, Free Radic. Biol. Med. 8 (1990) 281. [227] H.L. Zheng, M.B.H. Youdim, M. Fridkin, J. Med. Chem. 52 (2009) 4095.
[210] E.R. Stadtman, Free Radic. Biol. Med. 9 (1990) 315. [228] K.J. Barnham, V.B. Kenche, G.D. Ciccotosto, D.P. Smith, D.J. Tew, X. Liu, K. Perez,
[211] S.R. Powell, J. Nutr. 130 (2000) 1447S. G.A. Cranston, T.J. Johanssen, I. Volitakis, A.I. Bush, C.L. Masters, A.R. White,
[212] W. Maret, Exp. Gerontol. 43 (2008) 363. J.P. Smith, R.A. Cherny, R. Cappai, Proc. Natl. Acad. Sci. U.S.A. 105 (2008) 6813.
[213] J. Durand, G. Meloni, C. Talmard, M. Vasak, P. Faller, Metallomics 2 (2010) [229] A. Kumar, L. Moody, J.F. Olaivar, N.A. Lewis, R.L. Khade, A.A. Holder, Y. Zhang,
741. V. Rangachari, ACS Chem. Neurosci. 1 (2010) 691.
[214] G. Meloni, P. Faller, M. Vasak, J. Biol. Chem. 282 (2007) 16068. [230] G.L. Ma, F. Huang, X.W. Pu, L.Y. Jia, T. Jiang, L.Z. Li, Y.Z. Liu, Chem. Eur. J. 17
[215] G. Meloni, M. Vasak, Free Radic. Biol. Med. 50 (2011) 1471. (2011) 11657.
[216] A.I. Bush, R.E. Tanzi, Neurotherapeutics 5 (2008) 421. [231] D. Valensin, P. Anzini, E. Gaggelli, N. Gaggelli, G. Tamasi, R. Cini, C. Gabbiani,
[217] C.W. Ritchie, A.I. Bush, C.L. Masters, Expert Opin. Inv. Drugs 13 (2004) 1585. E. Michelucci, L. Messori, H. Kozlowski, G. Valensin, Inorg. Chem. 49 (2010)
[218] R. Squitti, C. Salustri, Curr. Alzheimer Res. 6 (2009) 476. 4720.

You might also like