You are on page 1of 25

Progress in Neurobiology 116 (2014) 33–57

Contents lists available at ScienceDirect

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

Metabolism and functions of copper in brain


Ivo F. Scheiber a, Julian F.B. Mercer b, Ralf Dringen c,d,*
a
Department of Parasitology, Faculty of Science, Charles University, Prague, Czech Republic
b
Centre for Cellular and Molecular Biology, School of Life and Environmental Sciences, Deakin University, Burwood Campus, 221 Burwood Highway,
Burwood, Victoria 3125, Australia
c
Center for Biomolecular Interactions Bremen, University of Bremen, P.O. Box 330440, D-28334 Bremen, Germany
d
Center for Environmental Research and Sustainable Technology, Leobener Strasse, D-28359 Bremen, Germany

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

Article history: Copper is an important trace element that is required for essential enzymes. However, due to its redox
Received 11 July 2013 activity, copper can also lead to the generation of toxic reactive oxygen species. Therefore, cellular
Received in revised form 8 January 2014 uptake, storage as well as export of copper have to be tightly regulated in order to guarantee sufficient
Accepted 8 January 2014
copper supply for the synthesis of copper-containing enzymes but also to prevent copper-induced
Available online 17 January 2014
oxidative stress. In brain, copper is of importance for normal development. In addition, both copper
deficiency as well as excess of copper can seriously affect brain functions. Therefore, this organ possesses
Keywords:
ample mechanisms to regulate its copper metabolism. In brain, astrocytes are considered as important
Astrocytes
Brain
regulators of copper homeostasis. Impairments of homeostatic mechanisms in brain copper metabolism
Copper have been associated with neurodegeneration in human disorders such as Menkes disease, Wilson’s
Neurodegeneration disease and Alzheimer’s disease. This review article will summarize the biological functions of copper in
Transport the brain and will describe the current knowledge on the mechanisms involved in copper transport,
storage and export of brain cells. The role of copper in diseases that have been connected with
disturbances in brain copper homeostasis will also be discussed.
ß 2014 Elsevier Ltd. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
2. Importance of copper for brain function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
2.1. Energy metabolism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
2.2. Antioxidative defense . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
2.3. Iron metabolism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
2.4. Neurotransmitter and neuropeptide synthesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
2.5. Neuromodulatory functions of copper . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
3. Copper content, transport and storage in the brain . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
3.1. Copper uptake into the brain . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37

Abbreviations: Ab, amyloid b; AD, Alzheimer’s disease; AMPA, a-amino-3-hydroxy-5-methyl-4-isoxazole propionic acid; APP, amyloid precursor protein; Atox 1, human
antioxidant protein 1; Atx1, antioxidant protein 1; BBB, blood–brain barrier; BCB, blood–cerebrospinal fluid barrier; CCS, copper chaperone for superoxide dismutase; Cp,
ceruloplasmin; CNS, central nervous system; CSF, cerebrospinal fluid; Ctr, copper transporter; DbM, dopamine-b-monoxygenase; DMT1, divalent metal transporter 1; DOPA,
3,4-dihydroxyphenylalanine; EPR, electron paramagnetic resonance; GABA, g-aminobutyric acid; GPI, glycosylphosphatidylinositol; GSH, glutathione; hCtr1, human copper
transporter 1; HD, Huntington’s disease; Hif-1a, hypoxia-inducible factor 1a; Hspa5, heat shock 70 kDa protein 5; IMS, intermembrane space; LA-ICP-MS, laser ablation
inductively coupled plasma mass spectroscopy; LOX, lysyl oxidase; LTP, long-term potentiation; MAPK, mitogen-activated protein kinase; MBD, metal binding domain; MT,
metallothionein; NGF, nerve growth factor; NMDA, N-methyl-D-aspartate; PAM, peptidylglycine a-amidating monoxygenase; PD, Parkinson’s disease; PINA, pineal gland
night-specific ATPase; PLC-PKC, phospholipase C-protein kinase C; PrP, prion protein; PrPSc, pathogenic form of the prion protein; ROS, reactive oxygen species; SOD,
superoxide dismutase; TGN, trans-Golgi network; ZIP, ZRT-/IRT-like protein.
* Corresponding author at: Center for Biomolecular Interactions Bremen, University of Bremen, P.O. Box 330440, D-28334 Bremen, Germany. Tel.: +49 421 218 63230; fax:
+49 421 218 63244.
E-mail address: ralf.dringen@uni-bremen.de (R. Dringen).

0301-0082/$ – see front matter ß 2014 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.pneurobio.2014.01.002
34 I.F. Scheiber et al. / Progress in Neurobiology 116 (2014) 33–57

3.2. Copper content and spatial distribution in brain . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37


3.3. Cellular copper uptake . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
3.4. Copper storage and copper trafficking . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
3.4.1. Glutathione . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
3.4.2. Metallothioneins . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
3.4.3. Copper chaperones . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
3.5. Copper export . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
4. Astrocyte-neuron coupling in copper metabolism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
5. Copper and neurodegenerative diseases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
5.1. Menkes disease . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
5.2. Wilson’s disease . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
5.3. Alzheimer’s disease. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
5.4. Prion diseases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
5.5. Parkinson’s disease . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
5.6. Huntington’s disease. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
6. Conclusions and outlook . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47

1. Introduction metabolism (cytochrome c oxidase), antioxidative defense


(Zn,Cu-containing superoxide dismutases), iron metabolism (ce-
Copper is an indispensable element for all organisms that have ruloplasmin), neurotransmitter synthesis (dopamine-b-monoxy-
an oxidative metabolism. Copper is after iron and zinc the third genase) and neuropeptide synthesis (peptidylglycine-a-amidating
most abundant essential transition metal in human liver enzyme).
(Lewinska-Preis et al., 2011). Although some compounds exist
with copper in the oxidation states Cu3+ and Cu4+, copper 2.1. Energy metabolism
biochemistry is largely dominated by Cu+ and Cu2+ compounds,
as these ions form numerous complexes with both organic and The brain is one of the most energy-demanding tissues of the
inorganic ligands. The soft Cu+ ion prefers ligands that have large body (Rossi et al., 2004). Most of this energy is required for active
polarizable electron clouds, such as sulfur ligands or unsaturated ion transport processes (Vergun et al., 2007). Since 95% of total ATP
nitrogen donors. Such ligands usually exert coordination numbers in the brain is estimated to be generated in mitochondria (Vergun
from two to four with linear, trigonal or tetrahedral coordination, et al., 2007), mitochondrial efficiency is essential for brain function.
while the hard Cu2+ ion prefers sp3 hybridized nitrogen and oxygen The final step of the electron transfer in the mitochondrial
ligands (Crichton and Pierre, 2001; Kaim and Rall, 1996; Rubino respiratory chain, the oxidation of reduced cytochrome c by
and Franz, 2012; Tisato et al., 2010; Wadas et al., 2007). The dioxygen, is catalyzed by cytochrome c oxidase, also known as
reduction potential of the Cu2+/Cu+ redox pair varies dramatically complex IV of the respiratory chain (Diaz, 2010; Ferguson-Miller
depending on the ligand environment and the pH. For example, the and Babcock, 1996; Hatefi, 1985; Tsukihara et al., 1996). This
one electron oxidation of various Cu+-complexes toward dioxygen protein is a member of the super-family of heme-copper contain-
has been reported to vary between 1.5 and +1.3 V (Tisato et al., ing oxidases (Ferguson-Miller and Babcock, 1996; Popovic et al.,
2010), while in copper proteins the reduction potential for Cu2+/ 2010; Stiburek et al., 2009). Mammalian cytochrome c oxidase is a
Cu+ ranges from +0.32 V to +0.78 V (Rubino and Franz, 2012). multimeric protein complex consisting of 13 subunits, encoded by
The brain concentrates heavy metals including copper for both the mitochondrial and nuclear genome (Hatefi, 1985; Leary
metabolic use (Bush, 2000). Copper is of great importance for the et al., 2009b; Stiburek and Zeman, 2010; Tsukihara et al., 1995,
normal development and function of the brain. As a cofactor of 1996). The mitochondria encoded subunits of cytochrome c
several enzymes and/or as structural component, copper is oxidase, Cox1, Cox2 and Cox3, constitute the catalytic core at
involved in many physiological pathways in the brain. This review which the dioxygen reduction and proton translocation take place
will summarize the functions of copper in the brain for various (Diaz, 2010; Ferguson-Miller and Babcock, 1996; Hamza and
biochemical pathways and will describe the current knowledge on Gitlin, 2002; Hatefi, 1985). Cox1 contains two heme moieties
the copper homeostasis by addressing copper transport, storage (hemes a and a3) and one copper ion (CuB; Hatefi, 1985; Tsukihara
and export in brain cells. Finally, disturbances in the copper et al., 1995, 1996). Cox2 contains a binuclear copper center (CuA)
homeostasis that have been connected with neurodegenerative which serves as the initial electron acceptor from cytochrome c
disorders will be discussed. (Tsukihara et al., 1995, 1996). During dioxygen reduction electrons
derived from cytochrome c are transferred from the CuA center first
2. Importance of copper for brain function to heme a and then to the site of dioxygen binding and reduction
which is a binuclear center consisting of heme a3 and CuB
Copper is utilized in the brain for general metabolic as well as (Ferguson-Miller and Babcock, 1996; Tsukihara et al., 1995, 1996).
for more brain specific functions (Lutsenko et al., 2010). Copper is Cytochrome c oxidase deficiency is one of the most common
an essential cofactor and/or a structural component of a number of causes of respiratory chain defects in humans (Borisov, 2002; Diaz,
important enzymes (Scheiber and Dringen, 2013) which are 2010; Hamza and Gitlin, 2002). Human cytochrome c oxidase
involved in redox reactions (Kaim and Rall, 1996; Rubino and deficiency manifests as wide variety of disorders with distinct
Franz, 2012). The relatively high reduction potential of the Cu2+/ clinical phenotypes resulting from a number of unique genetic
Cu+ system enables many of the copper enzymes to directly oxidize abnormalities (Borisov, 2002; Diaz, 2010; Hamza and Gitlin, 2002).
their substrates, for example superoxide by superoxide dismutase Pathological features range from metabolic acidosis, weakness and
and catechols by tyrosinase (Tisato et al., 2010). Copper-dependent cardiomyopathy to neurodegeneration (Borisov, 2002; Diaz, 2010;
enzymes participate in biological processes such as energy Hamza and Gitlin, 2002). Cytochrome c oxidase deficiency rarely
I.F. Scheiber et al. / Progress in Neurobiology 116 (2014) 33–57 35

arises from mutations located in mitochondrial or nuclear genes plasma, lymph and cerebrospinal fluid, where it protects cell
encoding the cytochrome c oxidase subunits, but is rather membranes against oxidative stress (Antonyuk et al., 2009;
secondary to loss-of-function mutations in genes encoding for Petersen and Enghild, 2005).
proteins required for the assembly of the functional holoprotein In brain of Menkes disease patients SOD1 immunoreactivity is
(Diaz, 2010; Shoubridge, 2001). Thus, mutations in genes encoding strongly reduced, whereas the expression level of SOD2 is elevated
Sco1 and Sco2, both critical for the metallation of the CuA site, (Shibata et al., 1995), a well-characterized response to oxidative
result in impaired cytochrome c oxidase function (Leary et al., stress (Uriu-Adams and Keen, 2005). However, while SOD1 activity
2004). Reduced insertion of copper might also be the reason for the under copper deficient conditions in rodents is strongly reduced in
cytochrome c oxidase impairment observed in the brain of Menkes liver (Lai et al., 1994; Prohaska, 1991), it is maintained at almost
disease patients (Maehara et al., 1983) and in animal models of this normal levels in brain (Gybina and Prohaska, 2006; Lai et al., 1994;
disease (Yoshimura et al., 1993) as well as in copper deficient Prohaska, 1991; Prohaska et al., 2003).
rodents (Gybina and Prohaska, 2006, 2008a,b; Kunz et al., 1999; Besides its function in the detoxification of superoxide, SOD1
Prohaska, 1991; Prohaska et al., 1995; Rusinko and Prohaska, has been connected with intracellular signaling (Juarez et al., 2008;
1985). Mondola et al., 2004). The activity of several growth factors, e.g.
epidermal growth factor, platelet-derived growth factor and
2.2. Antioxidative defense vascular endothelial growth factor, is redox regulated (Valko
et al., 2007) and SOD1 has been demonstrated to play an essential
The brain is sensitive to oxidative stress, as it has a high rate of role in mitogen-activated protein kinase (MAPK) signaling by
oxidative metabolism, contains compared to other organs only low mediating a hydrogen peroxide-dependent oxidation and inacti-
specific activities of antioxidative enzymes and is rich in redox- vation of phosphatases in several tumor cell lines (Juarez et al.,
active metals and polyunsaturated fatty acids (Dringen, 2000; 2006, 2008). Furthermore, secreted SOD1 has been shown to bind
Halliwell, 2006; Rossi et al., 2004). Copper deficit may render the to SK-N-BE neuroblastoma cells, leading to an increase in
brain even more susceptible to oxidative stress, since defective intracellular calcium concentrations through a phospholipase C-
cytochrome c oxidase activity may result in increased superoxide protein kinase C (PLC-PKC)-dependent pathway and subsequently
production by the respiratory chain and/or impaired activity of the to an activation of the MAPK extracellular signal-regulated kinases
copper-dependent superoxide dismutases (SOD) 1 and 3 may 1 and 2 (Mondola et al., 2004). Interestingly, the activation of the
weaken the antioxidative defense. PLC-PKC pathway by SOD1 was independent of its superoxide
SOD1 and SOD3 are members of the ubiquitous family of dismutase activity (Mondola et al., 2004). Since SOD1 is also
enzymes that convert superoxide to dioxygen and hydrogen present in the neuronal microenvironment, a neuromodulatory
peroxide (Miao and St Clair, 2009; Perry et al., 2010) for further role of SOD1 has been suggested (Mondola et al., 2004).
disposal by catalase, glutathione peroxidase and peroxiredoxins
(Bell and Hardingham, 2011; Hirrlinger and Dringen, 2010). 2.3. Iron metabolism
Superoxide is produced as byproduct during the electron flow in
the respiratory chain and as product of the autoxidation of Like copper, iron is also a redox-active metal and is an essential
catecholamines and their metabolites (Dringen and Hirrlinger, cofactor for many enzymatic processes. The brain has a substan-
2010; Halliwell and Gutteridge, 2007; Hirrlinger and Dringen, tial requirement for this essential metal as iron deficiency or
2010). Superoxide is also product of some enzyme reactions. For excess can seriously affect brain function (Crichton et al., 2011;
example NADPH oxidases in macrophages and microglial cells Dringen et al., 2007). The copper-containing ferroxidase cerulo-
produce superoxide after activation of these cells during the plasmin (Cp) plays an important role in iron-homeostasis (Healy
immune response (Dringen, 2005; Halliwell and Gutteridge, 2007; and Tipton, 2007; Hellman and Gitlin, 2002; Kosman, 2010b) and
Morel et al., 1991). Since excess of superoxide can contribute to thus links copper and iron metabolism. Cp belongs to the family of
oxidative stress that will lead to oxidation of cellular constituents multicopper oxidases that couple the one-electron oxidation of
and initiate lipid peroxidation (Halliwell and Gutteridge, 2007), four substrate molecules with the four-electron reduction of
SODs represent a first line of defense against the toxicity of dioxygen to water (Healy and Tipton, 2007; Hellman and Gitlin,
superoxide. 2002; Kosman, 2010a; Quintanar et al., 2007). The reducing
In eukaryotic cells three distinct isoforms of SOD, SOD1–3, that substrates utilized in this reaction vary among the enzyme family
are encoded by three different genes have been identified (Miao members (Hellman and Gitlin, 2002). The known substrates are
and St Clair, 2009; Perry et al., 2010). In contrast to SOD1 and SOD3 diverse and include compounds such as low valent transition
which contain catalytic copper and structural zinc ions in their metal ions (Fe2+, Mn2+, Cu+), bilirubin, ascorbate, phenols and
active sites (Antonyuk et al., 2009; Carrico and Deutsch, 1970; nitrate (Hellman and Gitlin, 2002; Kosman, 2010a; Quintanar
Miao and St Clair, 2009; Tibell et al., 1987; Weisiger and Fridovich, et al., 2007).
1973b), SOD2 contains manganese as metal cofactor (Miao and St The importance of Cp in the maintenance of iron homeostasis is
Clair, 2009; Weisiger and Fridovich, 1973a, 1973b). SOD1 is a demonstrated by aceruloplasminemia (Kono, 2013), an autosomal
homodimeric protein located largely in the cytosol with minor recessive disorder resulting from a loss-of-function mutation in
fractions being present in intracellular compartments including the Cp gene (Harris et al., 1995; Takahashi et al., 1996; Yoshida
the nucleus, the intermembrane space of mitochondria, lysosomes et al., 1995). This disease is characterized by marked iron
and peroxisomes (Crapo et al., 1992; Okado-Matsumoto and accumulation in the brain and other tissues (Gonzalez-Cuyar
Fridovich, 2001; Thomas et al., 1974; Weisiger and Fridovich, et al., 2008; Harris et al., 1995; Miyajima et al., 1996; Yoshida et al.,
1973a). However, some cell types also secrete SOD1 (Cimini et al., 1995). Symptoms include neurological abnormalities such as
2002; Mondola et al., 1996, 1998). The homotetrameric SOD2 is a dysarthria, dystonia and mild dementia (Gonzalez-Cuyar et al.,
mitochondrial enzyme that resides within the matrix and is 2008; Harris et al., 1995; McNeill et al., 2008; Miyajima et al., 1996;
associated with the inner membrane of mitochondria (Miao and St Takahashi et al., 1996; Yoshida et al., 1995). These neurological
Clair, 2009; Okado-Matsumoto and Fridovich, 2001; Perry et al., symptoms mirror the site of iron deposition in the brain (Miyajima,
2010; Weisiger and Fridovich, 1973a). The homotetrameric 2003) and iron-mediated oxidative stress is likely to contribute to
glycoprotein SOD3 is secreted by fibroblasts and glial cells and the pathogenesis of aceruloplasminemia (Gonzalez-Cuyar et al.,
has been found in the extracellular matrix of tissues as well as in 2008; Kaneko et al., 2002a, 2002b; Kono and Miyajima, 2006;
36 I.F. Scheiber et al. / Progress in Neurobiology 116 (2014) 33–57

Miyajima et al., 2002). The basal ganglia and the cortical gray 2.5. Neuromodulatory functions of copper
matter of patients with Wilson’s disease (a genetic copper toxicity
condition) contain more iron than the normal control (Cumings, Synaptosomes and primary hippocampal neurons have been
1948). This increase is likely to be a consequence of impaired shown to release copper following depolarization, resulting in
transfer of copper to apo-Cp, which has been reported at least for copper concentrations in the synaptic cleft that have been reported
the liver of Wilson’s disease patients (Meng et al., 2004; Terada to range from 15 mM to 250 mM (Hartter and Barnea, 1988; Hopt
et al., 1998). et al., 2003; Kardos et al., 1989; Schlief et al., 2005). This suggests a
Besides its critical role in iron-homeostasis, Cp has been potential neuromodulatory function of copper as also demonstrat-
suggested to be an important contributor to the plasma ed for zinc (Frederickson and Bush, 2001). Indeed, exogenous
antioxidative capacity, since it displays ferroxidase, cuprous application of already low micromolar concentrations of copper
oxidase and glutathione peroxidase activities and has the ability produces an antagonistic effect on N-methyl-D-aspartate (NMDA)-,
to scavenge reactive oxygen species (ROS) (Atanasiu et al., 1998; kainate-, a-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid
Healy and Tipton, 2007; Stoj and Kosman, 2003). Cp also possesses (AMPA)- and g-aminobutyric acid (GABA)-receptors as well as on
amine oxidase, catechol oxidase and ascorbate oxidase activities, the P2Y1, P2X4 and P2X7 purinoreceptors (Coddou et al., 2003;
although the physiological importance of these enzymatic activi- Doreulee et al., 1997; Horning and Trombley, 2001; Huidobro-Toro
ties remains to be elucidated (Bielli and Calabrese, 2002; Healy and et al., 2008; Kardos et al., 1989; Li et al., 1996; Peters et al., 2011;
Tipton, 2007). Even though copper bound to Cp accounts for up to Sharonova et al., 2000; Trombley and Shepherd, 1996; Vlachova
95% of plasma copper, there is no evidence for a direct role of this et al., 1996; Weiser and Wienrich, 1996; Xiong et al., 1999; Zhu
plasma protein in copper transport to tissues (Bielli and Calabrese, et al., 2002). Copper has further been shown to block glycine-
2002; Choi and Zheng, 2009; Healy and Tipton, 2007; Meyer et al., mediated currents, when glycine receptors are activated by low,
2001). non-desensitizing concentrations of glycine (Trombley and Shep-
herd, 1996; Wang et al., 2002) and to potentiate P2X2 receptor
2.4. Neurotransmitter and neuropeptide synthesis activity (Huidobro-Toro et al., 2008; Xiong et al., 1999). In contrast
to its acute inhibitory effect on AMPA- and GABA-mediated
Dopamine-b-monoxygenase (DbM) and peptidylglycine a- currents, treatment of rat hippocampal neurons with copper for 3 h
amidating monoxygenase (PAM) belong to a small class of copper enhances synaptic activity (Peters et al., 2011). In addition to its
proteins found exclusively in mammals (Klinman, 2006). The effect on ligand-gated receptors, copper has been shown to inhibit
latter enzyme is composed of two enzymatic domains, pepti- voltage-gated Ca2+- and K+-channels (Horning and Trombley,
dylglycine a-hydroxylating monoxygenase and peptidyl-a- 2001; Niu et al., 2005, 2006).
hydroxyglycine a-amidating lyase (Bousquet-Moore et al., AMPA/kainite and NMDA receptors are noncompetitively
2010a; Klinman, 1996; Prigge et al., 1997, 2000). Both DbM blocked by copper in cultured rat cortical neurons (Vlachova
and PAM catalyze the dioxygen- and ascorbate-dependent et al., 1996; Weiser and Wienrich, 1996), whereas GABAA receptors
hydroxylation of specific C–H-bonds in their target substrates and glycine receptors are blocked in a competitive manner
(Chen and Solomon, 2004; Klinman, 1996, 2006; Solomon et al., (Sharonova et al., 2000; Wang et al., 2002; Zhu et al., 2002). The
2011). inhibitory effect of copper on GABA-, NMDA- and glycine receptor-
DbM catalyzes the final step in noradrenaline synthesis, the mediated currents is voltage-independent, in contrast to the effect of
oxidative hydroxylation of dopamine to noradrenaline, and thus Mg2+, suggesting that copper acts at a specific neuromodulatory site
plays an important role in the catecholamine metabolism (Kim rather than as a channel blocker (Trombley and Shepherd, 1996;
et al., 2002; Klinman, 1996, 2006; Stewart and Klinman, 1988; Vlachova et al., 1996; Wang et al., 2002; Weiser and Wienrich, 1996).
Timmers et al., 2004). Noradrenaline is the principal sympathetic Experimental evidence suggests that copper catalyzes an S-
neurotransmitter and an important modulator of mood, attention, nitrosylation of the NMDA receptor, which is a known regulatory
arousal and cardiovascular function (Berridge, 2008; Goddard mechanism of NMDA receptors (Schlief et al., 2006). An oxidation
et al., 2010; Goldstein et al., 2003; Kim et al., 2002). However, process has also been proposed to be involved in the inhibition of
despite the fundamental role of noradrenaline in CNS functions, AMPA/kainate receptors by copper (Weiser and Wienrich, 1996). In
patients with congenital DbM deficiency exhibit only subtle signs contrast, an increase in AMPAergic neurotransmission has been
of neural dysfunction (Kim et al., 2002; Robertson et al., 1991). observed in response to prolonged exposure to copper, which has
DbM is located in granulated vesicles of adrenal medulla been ascribed to a copper-induced enrichment of the AMPA receptor
chromaffin cells, sympathetic nerve terminals, and noradrenergic at the plasma membrane (Peters et al., 2011).
and adrenergic neurons of the brain, where it exists as both a The exact role of copper in synaptic physiology remains to be
soluble and a membrane-bound protein (Geffen et al., 1969; Kim elucidated (for review see: Gaier et al., 2013). However,
et al., 2002; Klinman, 1996, 2006; Stewart and Klinman, 1988; synaptically released endogenous copper as well as exogenously
Timmers et al., 2004). The stimulus-dependent secretion of the applied copper protect primary hippocampal neurons against
soluble enzyme accounts for the presence of DbM in blood and CSF NMDA-mediated excitotoxic cell death, most likely by lowering
(Kim et al., 2002; Stewart and Klinman, 1988). the NMDA receptor-mediated intracellular Ca2+-elevation follow-
Amidated neuropeptides are synthesized in almost every ing depolarization/activation (Schlief et al., 2006). Copper also
neuron and are involved in a variety of functions in the brain, appears to be involved in the modulation of synaptic plasticity by
including neuronal proliferation, energy metabolism and neuro- inhibition of NMDA receptor-mediated induction and the mainte-
modulation (Bousquet-Moore et al., 2010b; Hansel et al., 2001; nance of long-term potentiation (LTP) (Voglis and Tavernarakis,
Magistretti et al., 1998). PAM is the only enzyme known to catalyze 2006), as hippocampus slices that had been exposed to exogenous
the a-amidation of peptide precursors (Bousquet-Moore et al., copper (Doreulee et al., 1997; Salazar-Weber and Smith, 2011) as
2010a; Klinman, 1996; Prigge et al., 1997, 2000). Due to its well as hippocampal slices of rats that had been fed a high-copper
physiological importance, lack of functional PAM in mice is diet did not show any LTP (Goldschmith et al., 2005; Leiva et al.,
embryonic lethal (Bousquet-Moore et al., 2010a). Expression of 2009). Although LTP has been associated with learning and
PAM in adults is highest in the secretory vesicles of atrial myocytes, memory in many organisms (Voglis and Tavernarakis, 2006),
endocrine cells of the pituitary gland and in many neurons, but not despite blocking LTP in the hippocampus, copper did not alter
limited to theses cell types (Prigge et al., 2000; Rhodes et al., 1990). learning and memory in rats (Leiva et al., 2009).
I.F. Scheiber et al. / Progress in Neurobiology 116 (2014) 33–57 37

Several of the neuromodulatory functions of copper appear to


be directly linked to interactions of copper with receptors (Gaier
et al., 2013). However, copper may exert its neuromodulatory
functions also by other means. At least for peripheral tissues
copper has been reported to induce mitogen-activated protein
kinase (Samet et al., 1998; Wu et al., 1999) as well as
phosphoinositide-3-kinase/Akt signaling (Barthel et al., 2007;
Walter et al., 2006). Copper may activate these signaling cascades
via the inhibition of protein tyrosine phosphatases (Barthel et al.,
2007; Samet et al., 1998; Wu et al., 1999) which has been reported
on a cellular level (Wu et al., 1999) as well as in cell-free systems
(Kim et al., 2000). Copper may also affect intracellular signaling
through activation of the metal-responsive transcription factor 1
Fig. 1. Uptake and efflux of copper at brain barriers. Copper is transported into the
(Balamurugan and Schaffner, 2006; Gunther et al., 2012) and/or brain via brain endothelial cells which comprise the blood–brain barrier (BBB).
the copper chaperone Atox1 that also functions as a copper- These cells take up copper from the blood by the copper transporter 1 (Ctr1), located
dependent transcription factor (Itoh et al., 2008, 2009; Muller and at the apical side and release copper into the brain parenchyma by ATP7A. Excess
Klomp, 2009). In addition, copper may modulate the action of copper is released from brain cells into the cerebrospinal fluid (CSF) where it is
taken up by polarized epithelial cells of the choroid plexus which form the Blood-CSF
neurotrophic factors. For example, the presence of copper
barrier (BCB). Copper taken up by these cells is either stored for potential transport
decreases the proliferative action of brain-derived neurotrophic into CSF by ATP7B or released into the blood by ATP7A.
factor on neuroblastoma cells (Travaglia et al., 2012), while it
increased cell proliferation in the presence of nerve growth factor
(NGF) (Travaglia et al., 2011). On the other hand, NGF promotes 3.2. Copper content and spatial distribution in brain
copper uptake that is required for protein methylation reactions
(Birkaya and Aletta, 2005). Consistent with an involvement of The estimated copper content of the human brain ranges from
protein methylation in NGF-stimulated neurite outgrowth (Cimato
3.1 to 5.1 mg/g wet weight (Davies et al., 2013, 2014; Lech and
et al., 1997), copper depletion has been demonstrated to interfere Sadlik, 2007; Rahil-Khazen et al., 2002). Compared to human
with this process (Birkaya and Aletta, 2005).
brains, mice appear to have a similar brain copper content (5.5 mg
copper/g wet weight; Waggoner et al., 2000), whereas rat brains
contain less copper (1–2.5 mg/g wet weight; Olusola et al., 2004;
3. Copper content, transport and storage in the brain

3.1. Copper uptake into the brain

Brain copper is derived from peripheral copper that is


transported across the blood–brain barrier (BBB) and/or the
blood–cerebrospinal fluid barrier (BCB), which separate the
brain interstitial space from blood and cerebrospinal fluid (CSF),
respectively (Zheng and Monnot, 2012). At both barriers copper
is transported primarily as free ion (Choi and Zheng, 2009).
Although the copper uptake into cerebral capillaries is much
slower than into the choroid plexus, the copper acquired by
cerebral capillaries appears to be more readily transported into
the brain parenchyma than copper from the choroid plexus to the
CSF (Choi and Zheng, 2009; Monnot et al., 2011). In fact, recent
evidence indicates that the role of the BCB in brain copper
homeostasis is rather to export copper from the CSF to the blood
than to import copper (Monnot et al., 2011). Based on these
findings, a model (Fig. 1) has been postulated for the copper
transport between the blood and the brain (Monnot et al., 2011;
Zheng and Monnot, 2012). The BBB represents the major route
for the transport of copper from the blood circulation into the
brain parenchyma, where copper is utilized and subsequently
released into the CSF via the brain interstitial fluid. The copper in
the CSF can be taken up by choroid epithelial cells, from where it Fig. 2. Copper metabolism of brain cells. Copper is taken up into brain cells by the
copper transporter Ctr1. Also the divalent metal transporter 1 (DMT1) may
may be stored or exported to the blood. Thus, while the BBB contribute to copper uptake. Since DMT1 and Ctr1 have been reported to prefer Cu+
determines the influx of copper into the brain, the BCB as substrate, an ecto-cuprireductase will provide the reduced copper species for
contributes to the maintenance of the copper homeostasis in uptake. Accumulated copper is sequestered by GSH or stored in metallothioneins
the brain extracellular fluids. However, the situation might be (MT). In addition, copper is shuttled to its specific cellular targets by the copper
chaperones CCS to superoxide dismutase 1 (SOD1), by Cox17 to Sco1/2 and Cox11
different in the developing brain for which the BCB has been
for subsequent incorporation into cytochrome c oxidase and by antioxidant protein
suggested to be the primary route of copper entry (Donsante 1 (Atox1) to ATP7A and ATP7B. Both, ATP7A and ATP7B transport copper into the
et al., 2010). trans-Golgi network (TGN) for subsequent incorporation into copper-dependent
The knowledge of the copper handling machinery in different enzymes such as tyrosinase (T), ceruloplasmin (Cp), dopamine-b-monoxygenase
cell types of the central nervous system is rudimentary (Lutsenko (DbM), peptidylglycine a-amidating monoxygenase (PAM) or lysyl oxidase (LOX).
When the cellular copper level rises above a certain threshold, ATP7A and ATP7B
et al., 2010). Nevertheless, it is clear that most of the key players translocate reversibly via vesicles toward the plasma membrane. ATP7A imports
(Fig. 2) of peripheral copper homeostasis are also present in brain copper into vesicles for release by fusion with the plasma membrane and/or exports
cells (Lutsenko et al., 2010; Tiffany-Castiglioni et al., 2011). copper directly.
38 I.F. Scheiber et al. / Progress in Neurobiology 116 (2014) 33–57

Sugawara et al., 1992). The cerebrospinal fluid contains copper in a 0.4–1.7 mg/g wet weight (Nooijen et al., 1981; Willemse et al.,
concentration of 0.3–0.5 mM (Forte et al., 2004; Nischwitz et al., 1982).
2008; Strozyk et al., 2009; Stuerenburg, 2000). However, the
extracellular copper concentration in brain tissue may be higher. 3.3. Cellular copper uptake
At least for the synaptic cleft copper concentrations of up to
250 mM have been reported (Hopt et al., 2003; Kardos et al., 1989). As in peripheral cells (Eisses et al., 2005; Kim et al., 2008; Lee
Several studies have demonstrated that copper is unevenly et al., 2001; Maryon et al., 2007), the copper transporter Ctr1 is
distributed in the brain (Becker et al., 2005; Bonilla et al., 1984; likely to be the major pathway for copper entry into brain cells.
Davies et al., 2013, 2014; Dobrowolska et al., 2008; Faa et al., 2001; Ctr1 mediates copper uptake with an apparent KM-value of 1–
Popescu et al., 2009; Serpa et al., 2008; Smeyers-Verbeke et al., 8 mM (Eisses and Kaplan, 2002, 2005; Lee et al., 2002a,b). The
1974; Wang et al., 2010; Warren et al., 1960). In general, copper transport of copper by human Ctr1 (hCtr1) does not depend on ATP
contents are higher in the gray matter (1.6–6.5 mg/g wet weight) hydrolysis or an ion gradient (Lee et al., 2002a), suggesting that
than in the white matter (0.9–2.5 mg/g wet weight; Becker et al., hCtr1 is neither an ion pump nor a secondary active transporter.
2005; Bonilla et al., 1984; Dobrowolska et al., 2008; Smeyers- Since Ctr1-mediated copper uptake is effectively inhibited by Ag+,
Verbeke et al., 1974; Warren et al., 1960). Quantification of copper which is isoelectric to Cu+, and strongly stimulated by ascorbate,
in human brain sections demonstrated that copper is enriched in monovalent copper is thought to be the copper species transported
the locus coeruleus and the substantia nigra, which both are by Ctr1 (Bertinato et al., 2010; Lee et al., 2002a,b). hCtr1 is
pigmented tissues and contain catecholaminergic cells (Davies composed of 190 amino acids, has a molecular mass of 23 kDa and
et al., 2013; Goldberg and Allen, 1981; Popescu et al., 2009; Warren shares more than 90% sequence identity with the mouse and rat
et al., 1960). Two-dimensional copper maps of human brain slices, Ctr1 homologues (Lee et al., 2000; Zhou and Gitschier, 1997).
obtained by laser ablation inductively coupled plasma mass Mammalian Ctr1 have an overall architecture consisting of an
spectroscopy (LA-ICP-MS), revealed that areas strongly enriched extracellular N-terminus, an intracellular loop connecting the first
in copper are also present within the hippocampus (Dobrowolska and second transmembrane domain, three membrane-spanning a-
et al., 2008). Compared to other gray matter regions the thalamus helices and an intracellular C-terminus. In contrast to most other
of human brain contains lower copper levels (Smeyers-Verbeke transporters that are predicted to have 6–12 membrane-spanning
et al., 1974; Warren et al., 1960), whereas the copper content of the domains, Ctr1 contains only three membrane-spanning a-helices
thalamus of rats is higher (Jackson et al., 2006), indicating that not (Eisses and Kaplan, 2002; Klomp et al., 2003). Since three a-helices
only the copper content but also the copper distribution within the are insufficient to form a translocation pore, Ctr1 has to
brain varies between species. oligomerize to provide a pore that allows copper to pass through
Histochemical and more recently LA-ICP-MS investigations the plasma membrane (De Feo et al., 2007; Eisses and Kaplan,
showed that copper contents in brain are higher in glial cells than 2005). Indeed, biochemical, genetic and electron microscopy
in neurons, under both normal and pathological conditions (Becker studies have demonstrated that Ctr1 forms homotrimers (Aller
and Salber, 2010; Kodama et al., 1991; Szerdahelyi and Kasa, 1986). and Unger, 2006; Eisses and Kaplan, 2002; Lee et al., 2002a) and
In addition, some glial cells that are localized close to the brain that oligomerization is essential for Ctr1 functionality (Aller et al.,
ventricles contain very high copper levels (Szerdahelyi and Kasa, 2004).
1986). Very recently, astrocytes in the subventricular zone of In mammalian brain, Ctr1 is ubiquitously distributed with its
rodent brain have been reported to contain very high copper expression being most abundant in the choroid plexus and high in
contents and to even increase their copper content with age the endothelial cells of the capillaries (Davies et al., 2013; Gybina
(Pushkar et al., 2013). These finding indicate that glial cells store and Prohaska, 2006; Kuo et al., 2006; Nishihara et al., 1998; Zheng
copper and presumably these cells play an important role in the and Monnot, 2012). Ctr1 is enriched on the apical membrane of the
copper metabolism of the brain. In the locus coeruleus and the choroid plexus where it may extract copper from the CSF, consistent
substantia nigra, the regions with the highest copper content in with the proposed function of the choroid plexus in the mainte-
human brain, copper was predominately observed in glial cells and nance of the copper homeostasis in the brain extracellular fluids
not in neurons (Szerdahelyi and Kasa, 1986), although another (Kuo et al., 2006; Zheng and Monnot, 2012). In brain capillary
study reported that most of the copper in the locus coeruleus was endothelial cells, Ctr1 has been proposed to be located at the
located on presynaptic membranes of a part of the afferent luminal side where it may regulate the uptake of copper from the
terminals contacting dendrites or somatic spines of neurons (Sato blood (Zheng and Monnot, 2012). Overexpression of Ctr1 has
et al., 1994). The presence of substantial amounts of copper in recently been reported for brain capillary endothelial cells and
nerve terminals was also shown by subcellular fractionation of the choroid epithelial cells in the brains of North Ronaldsay sheep
rat cerebral cortex demonstrating that 23% of the total copper (Haywood and Vaillant, in press), supporting the view that Ctr1 is
content is contained in synaptosomes (Matsuba and Takahashi, involved in the transport of copper from the periphery into the
1970). The copper concentration in synaptosomes was estimated brain. In rat dorsal root ganglion tissue intense Ctr1-immunoreac-
to be about 15 mM, whereas the copper concentration in synaptic tivity has been observed in large-sized neurons, which are thought
vesicles is about 290 mM, suggesting that an active transport is to have a strong need for Ctr1 to meet their high demands for
involved in the uptake of copper into synaptic vesicles (Hopt et al., copper delivery to cytochrome c oxidase (Ip et al., 2010; Liu et al.,
2003). 2009). In human visual cortex, anterior cingulate cortex, caudate
Both brain copper content and distribution change during and putamen, expression of Ctr1 has been observed primarily in
development, with age and in neurodegenerative diseases (Serpa neurons, while in the cerebellum Ctr1 expression appeared to be
et al., 2008; Tarohda et al., 2004; Wang et al., 2010). The overall restricted to Bergmann glia (Davies et al., 2013).
brain copper content has been reported to increase with aging in Ctr1 plays an essential role in embryonic development as
mouse (Wang et al., 2010) and bovine brain (Zatta et al., 2008). deletion of Ctr1 is embryonic lethal, most likely due to an
Compared to a normal human brain copper content of 13–60 mg/g insufficient supply of the developing embryo with copper (Kuo
dry weight, the copper content in brains of Wilson’s disease et al., 2001b; Lee et al., 2001). In brain and spleen of Ctr1-
patients was increased to 125 mg/g dry weight (Faa et al., 2001), heterozygous knock-out mice copper levels are lowered to about
while the brain copper contents of Menkes disease patients have 50% of that of wild type animals, but were normal in other organs,
been reported to be strongly reduced from 2.3 mg/g wet weight to suggesting that Ctr1 is especially important for copper uptake in
I.F. Scheiber et al. / Progress in Neurobiology 116 (2014) 33–57 39

brain and spleen, while an alternative Ctr1-independent copper Finally, Ctr1-independent copper uptake may not require a
transport systems is able to supply other organs with copper (Lee specific transporter, but may occur by endocytosis (Ferruzza et al.,
et al., 2001). The identities of these alternative copper transport 2000; Zimnicka et al., 2007). In support of such a mechanism,
pathways and their potential involvement in copper metabolism apical copper uptake in CaCo2 cells has been shown to be not
within the brain remain to be elucidated. saturable (Ferruzza et al., 2000; Zimnicka et al., 2007). However,
In humans a second member of the Ctr family, hCtr2, has been this lack of saturation could also be explained by the contribution
identified (Zhou and Gitschier, 1997), which also mediates the of an unidentified transporter with a high KM-value for copper
uptake of Cu+ (Bertinato et al., 2008; van den Berghe et al., 2007). ions (Zimnicka et al., 2007). Endocytotic uptake is most likely also
However, since hCtr2 transports copper with lower affinity involved in the uptake of copper oxide nanoparticles which has
compared to hCtr1 (Bertinato et al., 2008), does not complement recently been reported for cultured astrocytes (Bulcke et al.,
the yeast copper uptake deficiency (Zhou and Gitschier, 1997) and 2014).
is predominately located in intracellular compartments (Bertinato
et al., 2008; van den Berghe et al., 2007), its exact role in copper 3.4. Copper storage and copper trafficking
transport in mammalian cells remains unclear (Gupta and
Lutsenko, 2009). Under physiological conditions cells contain less than one free
Experimental evidence was provided that the divalent metal copper ion (Rae et al., 1999). This extremely low concentration of
transporter (DMT1), which is also expressed in brain cells (Burdo free cellular copper is maintained by binding of copper to
et al., 2001; Pelizzoni et al., 2012), is involved in apical copper metallothioneins (MTs) and other ligands of low molecular mass
uptake by intestinal cells (Arredondo et al., 2003). DMT1 is such as glutathione (GSH). In addition, mitochondria have been
constituted of 561 amino acids, contains 12 membrane-spanning suggested to be involved in cellular copper buffering (Cobine et al.,
domains and is widely expressed in several tissues such as 2004; Leary et al., 2009b; Maxfield et al., 2004). A group of proteins
intestine, kidney and brain (Gunshin et al., 1997). It mediates termed copper chaperones shuttle copper to its specific cellular
ferrous iron transport across the plasma membrane and from targets, thereby protecting it from being scavenged by MTs or GSH
intracellular compartments into the cytosol by a process that is (Robinson and Winge, 2010).
coupled to a proton co-transport (Gunshin et al., 1997). In addition Also copper that has entered brain cells is likely to be
to iron, other transition metal ions including manganese, sequestered by GSH and either stored as MT-copper complex or
cadmium and copper provoked large inward currents in DMT1- shuttled by copper chaperones to their specific cellular target sites.
expressing oocytes, suggesting that these metal ions might also be Elevated brain copper levels are associated with an increase in MT
transported by DMT1 (Gunshin et al., 1997). Indeed, a decreased levels, most likely reflecting a compensatory response against
apical uptake of copper by CaCo2 cells was observed following copper-induced toxicity (Haywood et al., 2008; Haywood and
partial knock-down of DMT1 and competition experiments Vaillant, in press; Ono et al., 1997; Zatta et al., 2008). Detailed
revealed that copper inhibits apical iron uptake and vice versa studies of copper chaperones in the brain are scarce. However, the
(Arredondo et al., 2003). However, another study found no copper chaperones Atox1, copper chaperone for superoxide
evidence for a role of DMT1 in apical copper transport of CaCo2 dismutase (CCS) and the copper chaperones for cytochrome c
cells (Zimnicka et al., 2007) and recently was reported that at least oxidase, Cox17, Sco1 and Sco2, have been shown to be expressed in
one DMT1 isoform does not mediate copper transport (Illing et al., brain (Gybina and Prohaska, 2006; Hamza et al., 2000; Klomp et al.,
2012). 1997; Naeve et al., 1999; Nishihara et al., 1998; Papadopoulou
Embryonic Ctr1-homozygous knock-out cells have been et al., 1999; Punter et al., 2000; Rothstein et al., 1999; Takahashi
demonstrated to possess a residual copper transport activity of et al., 2002).
approximately 30% of that of wild type cells (Lee et al., 2002b). In
contrast to Ctr1-mediated copper uptake, this Ctr1-independent 3.4.1. Glutathione
copper transport activity was of low affinity (KM-value: 10 mM) The tripeptide GSH is present in millimolar concentrations in
and most likely represents the transport of divalent copper (Lee cells and is important for the detoxification of peroxides and ROS as
et al., 2002b). Even though the KM-value of this transport pathway well as for intracellular redox regulation and signaling (Hirrlinger
is close to that reported for hCtr2 (Bertinato et al., 2008) the lack of and Dringen, 2010; Schmidt and Dringen, 2012). In addition, GSH
inhibition by Ag+ makes Ctr2 unlikely to contribute to the copper has been connected with the detoxification of xenobiotics, toxins
uptake observed in Ctr1-deficient cells. Also, iron, manganese and and metals (Ballatori et al., 2009; Jomova et al., 2010).
cadmium, high-affinity substrates for DMT1 (Garrick et al., 2003), GSH forms stable complexes with Cu+ in cell-free systems
only slightly inhibited the residual copper uptake (Lee et al., even in the presence of oxygen (Ascone et al., 1993; Banci et al.,
2002b), while zinc strongly competed for Ctr1-independent copper 2010; Ciriolo et al., 1990; Kachur et al., 1998). Consistent with
transport. This observation led to the assumption that the GSH functioning as an intracellular copper chelator, an elevation
alternative copper transport could be mediated by members of of the cellular GSH content has been shown to lower the
the ZIP (ZRT-/IRT-like protein) family of metal transporters (Lee bioavailable intracellular copper pool (Chen et al., 2008).
et al., 2002b). In mammals, 14 ZIP transporters have been Copper–GSH complexes are actually considered as major
described (Kambe et al., 2004; Lichten and Cousins, 2009). These contributor to the exchangeable cytosolic copper pool (Banci
proteins have broad substrate specificity and were shown to et al., 2010; Freedman et al., 1989; Miras et al., 2008). In
transport zinc, manganese, iron and cadmium ions (Eide, 2006; accordance with this function, Cu(I)-GSH is capable in vitro of
Kambe et al., 2004). A contribution of ZIP members in copper transferring copper to MTs (Ferreira et al., 1993) and SOD1
transport has not been reported, but copper has been shown to (Ascone et al., 1993; Carroll et al., 2004; Ciriolo et al., 1990). Also
compete for zinc and cadmium uptake by Zip1, Zip2 and Zip14 for the in vivo situation, Cu(I)–GSH has been demonstrated to be
(Gaither and Eide, 2000, 2001; Girijashanker et al., 2008). However, required for the incorporation of copper into MTs and for the
no direct evidence for a contribution of Zip1 and Zip2 to Ctr1- CCS-independent activation of SOD1 (Carroll et al., 2004;
independent copper uptake activity has been found (Lee et al., Freedman et al., 1989; Huang et al., 2011; Jensen and Culotta,
2002b). Whether other members of this protein family may 2005; Leitch et al., 2009a). Cu(I)–GSH has further been
contribute to Ctr1-independent copper accumulation, remains to suggested to represent the major source of copper for Atox1
be elucidated. (Miras et al., 2008; Poger et al., 2008).
40 I.F. Scheiber et al. / Progress in Neurobiology 116 (2014) 33–57

The susceptibility of cells to copper toxicity correlates strongly et al., 1992; Vasak and Meloni, 2011). MTs are primarily
with their cellular GSH content (Chen et al., 2008; Du et al., 2008; cytoplasmic proteins, however, they have been observed in the
Freedman et al., 1989; Steinebach and Wolterbeek, 1994; nucleus under several physiological and pathological conditions
Vidyashankar and Patki, 2010; White et al., 1999c). Sequestration (Kiningham et al., 1995; Nagano et al., 2000; Nishimura et al.,
of copper by GSH protects cells from the toxic effects of free copper 1989). Despite the absence of a classical secretion sequence, MT-1,
that arise from its ability to bind indiscriminately to essential MT-2 and MT-3 are secreted by some cells including brain
proteins and/or to catalyze ROS generation (Burkitt, 2001; Hanna astrocytes (Chung et al., 2008; Miyazaki et al., 2011; Moltedo et al.,
and Mason, 1992; Jimenez and Speisky, 2000; Milne et al., 1993). 2000; Trayhurn et al., 2000; Uchida et al., 2002). Mammalian MTs
However, Cu(I)–GSH is redox-active and continuously reacts with consist of a single polypeptide chain of 61–68 amino acids, 20 of
molecular oxygen to produce superoxide (Aliaga et al., 2010, 2011; which are strictly conserved cysteines that are arranged in a
Speisky et al., 2008, 2009). This feature accounts for the ability of distinct pattern of CxC, CC and CxxC sequences (Kissling and Kagi,
Cu(I)–GSH to reduce Fe3+ and to stimulate the release of Fe2+ from 1977; Kojima et al., 1976; Palmiter et al., 1992; Vasak and Meloni,
ferritin (Aliaga et al., 2011), thus fostering iron-catalyzed produc- 2011). MTs contain two metal binding domains, the N-terminal b-
tion of hydroxyl radicals by the Fenton reaction in the presence of domain and the C-terminal a-domain. These domains differ in
hydrogen peroxide. This pro-oxidative behavior of the Cu(I)–GSH their abilities to bind metal ions and their folding is induced by
complex has been postulated to underlie the oxidative stress and metal coordination (Boulanger et al., 1982; Duncan et al., 2006;
toxicity associated with the over-exposure of cells to copper Hunt et al., 1984; Salgado and Stillman, 2004; Vasak and Meloni,
(Aliaga et al., 2011). 2011; Winge and Miklossy, 1982). Within cells MTs are present in
GSH depletion has been found to lower the rate of copper their apo-forms as well as in their metal bound forms (Petering
uptake into liver cells and trophoblasts (Freedman et al., 1989; et al., 2006; Yang et al., 2001). One MT can reversibly bind up to
Tong and McArdle, 1995), suggesting that the formation of the seven divalent metal ions in a three-metal-thiolate cluster (b-
Cu(I)-GSH complex is involved in copper uptake. Alternatively, this domain) and a four-metal-thiolate cluster (a-domain) or up to 12
lowered copper uptake might reflect the adaption of the cells to monovalent ions in two six-metal-thiolate clusters (Boulanger
their intracellular bioavailable pool of copper. In accord with this et al., 1982; Duncan et al., 2006; Hunt et al., 1984; Roschitzki and
view, down-regulation of hCtr1 lowered copper uptake in GSH- Vasak, 2002). Although all mammalian MT isoforms can bind a
depleted liver cells, whereas hCtr1 abundance and copper uptake variety of heavy metals, copper and zinc are the primary metals
rates are increased in cells with elevated GSH levels (Chen et al., that bind to MTs under normal physiological conditions (Bogumil
2008). Despite having slower copper uptake, GSH depleted cells et al., 1996; Orlowski and Piotrowski, 1998; Quaife et al., 1994;
accumulate more copper than control cells as copper export is Stillman, 1995; Uchida et al., 1991; Vasak and Meloni, 2011).
impaired in these cells (Singleton et al., 2010). Multiple lines of evidence suggest that MTs are involved in the
In addition to its copper chelating and copper trafficking maintenance of cellular copper homeostasis. MTs have been
properties, GSH can participate in cellular copper homeostasis by postulated to function as cytosolic copper store in normal copper
regulating the activities of copper transport proteins via glutathio- metabolism and to serve as metal reservoir in the event of copper
nylation/deglutathionylation (Lim et al., 2006; Singleton et al., deficit (Ogra et al., 2006; Suzuki et al., 2002; Tapia et al., 2004).
2010; Voronova et al., 2007a,b). The activities of the copper Consistent with such functions, cells lacking MT-1 and MT-2 have
transporting ATPases ATP7A and ATP7B have recently been lower basal copper contents and are more sensitive to copper
demonstrated to be regulated by this posttranslational mechanism depletion than wild-type cells (Miyayama et al., 2009; Ogra et al.,
(Singleton et al., 2010). Glutathionylation might also be involved in 2006). However, MT-1 and MT-2 appear not to play an obligatory
the regulation of the copper chaperone function of Cox17 as it has role in normal copper homeostasis, since MT-1/2 knock-out mice
been shown to form mixed disulfide with GSH in vitro (Voronova do not show a copper-related phenotype (Camakaris et al., 1999;
et al., 2007a,b). Whether also in brain cells the copper metabolism Michalska and Choo, 1993). The expression of MTs is induced by an
is regulated by glutathionylation remains to be elucidated. excess of copper (Hidalgo et al., 1994; Jiang et al., 2002; Kawai
et al., 2000; Kuo et al., 2001a; Wake and Mercer, 1985). Elevated
3.4.2. Metallothioneins MT contents have been demonstrated in liver and kidney of
Metallothioneins (MTs) constitute a family of low-molecular Wilson’s disease patients and in some animal models (Klein et al.,
weight, cysteine-rich proteins (Bremner and Davies, 1975; Buhler 1990; Mercer et al., 1992; Mulder et al., 1992; Sakurai et al., 1992;
and Kagi, 1974; Kagi et al., 1974; Palacios et al., 2011; Pulido et al., Suzuki-Kurasaki et al., 1997). Since MTs are capable of binding
1966; Vasak and Meloni, 2011). In addition to their suggested role excess cellular copper, an increase in the cellular MT content
in the homeostasis of copper and zinc, functions of MTs include the confers resistance against copper-induced toxicity (Freedman
detoxification of non-essential metal ions, neuroprotection, et al., 1986; Jiang et al., 2002; Kawai et al., 2000). Hence, the rise in
maintenance of the intracellular thiol redox potential and MT levels reflects an adaption of cells to copper overload
regulation of cell proliferation (Cherian and Kang, 2006; Colvin conditions. Despite the function of MTs in the cellular detoxifica-
et al., 2010; Hidalgo et al., 2001; Maret, 2011; Thirumoorthy et al., tion of copper, cells lacking MT-1 and MT-2 are only slightly more
2011; Vasak and Meloni, 2011). sensitive to excess copper than wild type cells (Jiang et al., 2002;
In mammals there are four MT isoforms, MT-1 to MT-4 (Kagi Kelly and Palmiter, 1996; Park et al., 2001; Tapia et al., 2004). This
and Valee, 1960; Palmiter et al., 1992; Quaife et al., 1994; counterintuitive finding can be attributed to a lower copper uptake
Raudenska et al., 2013; Thirumoorthy et al., 2011; Vasak and by these cells (Armendariz et al., 2006; Tapia et al., 2004). Thus, in
Meloni, 2011). The major isoforms MT-1 and MT-2 are ubiqui- addition to their role in cytosolic copper storage and copper
tously expressed in almost all organs and tissues and inducible by detoxification, MTs appear to be involved in the regulation of
various stressors including heavy metals, oxidative stress and pro- copper uptake by controlling the intracellular concentration of free
inflammatory cytokines (Andrews, 2000; Haq et al., 2003; Kawai copper.
et al., 2000; Leone et al., 1985; Santon et al., 2008; Thirumoorthy There are three MT isoforms in the brain. MT1, MT2 and MT3
et al., 2011; Vasak and Meloni, 2011). MT-3 and MT-4 are are expressed in the BBB and BCB as well as in astrocytes and
constitutively expressed mainly in brain and stratified epithelium, neurons, whereas microglial cells and oligodendrocytes have been
respectively (Garrett et al., 1999a,b; Hozumi et al., 2008; Palmiter reported to be essentially devoid of MTs (Anezaki et al., 1995;
et al., 1992; Quaife et al., 1994; Thirumoorthy et al., 2011; Tsuji Hidalgo et al., 1994, 2001; Uchida et al., 1991). Elevated copper
I.F. Scheiber et al. / Progress in Neurobiology 116 (2014) 33–57 41

levels in brain are accompanied by an increase in MT levels, most Biogenesis of cytochrome c oxidase is a complicated process
likely reflecting a compensatory response against copper-induced that requires several specific proteins, so-called assembly factors,
toxicity (Haywood et al., 2008; Haywood and Vaillant, in press; and even a greater number of proteins with broader substrate
Ono et al., 1997; Zatta et al., 2008). specificities, such as mitochondrial ATP-dependent proteases
(Shoubridge, 2001; Hamza and Gitlin, 2002; Leary et al., 2004;
3.4.3. Copper chaperones Diaz, 2010; Stiburek and Zeman, 2010). Several proteins have been
Detailed studies of copper chaperones in the brain are scarce. identified so far to be involved in the insertion of copper ions into
However, the copper chaperones Atox1, CCS and the copper the CuB and CuA sites in Cox1 and Cox2 of mammalian cytochrome
chaperones for cytochrome c oxidase Cox17, Sco1 and Sco2 are c oxidase (Banci et al., 2008c; Barros et al., 2004; Heaton et al.,
expressed in the brain (Davies et al., 2013; Gybina and Prohaska, 2000; Nobrega et al., 2002; Rigby et al., 2007). Cox17 mediates the
2006; Hamza et al., 2000; Klomp et al., 1997; Naeve et al., 1999; transfer of copper to the Sco-proteins Sco1 and Sco2 and Cox11
Nishihara et al., 1998; Papadopoulou et al., 1999; Punter et al., (Banci et al., 2008b; Banting and Glerum, 2006; Beers et al., 1997;
2000; Rothstein et al., 1999; Takahashi et al., 2002). Horng et al., 2004; Leary et al., 2009b; Robinson and Winge, 2010),
Atox1 is the human homologue of the yeast antioxidant protein which is followed by the incorporation of copper into the CuA and
1 (Atx1). It is a 68 amino acid protein that shares 47% amino acid CuB sites (Banting and Glerum, 2006; Beers et al., 1997; Leary et al.,
identity and 58% similarity with Atx1 (Klomp et al., 1997). Atox1 is 2009b; Robinson and Winge, 2010).
abundant in all human tissues and distributed throughout the Cox17 is a small cysteine-rich, hydrophilic protein essential for
cytosol and nucleus of cells (Hamza et al., 1999; Klomp et al., the metallation of eukaryotic cytochrome c oxidases (Glerum et al.,
1997). Both expression level and intracellular localization of Atox1 1996; Heaton et al., 2000; Oswald et al., 2009; Takahashi et al.,
do not depend on the intracellular copper content (Hamza et al., 2002). Its dual localization in the intermembrane space (IMS) and
1999; Klomp et al., 1997). Atox1 has been demonstrated to bind in the cytosol (Beers et al., 1997; Glerum et al., 1996; Maxfield
and transfer Cu+ to the N-terminal metal-binding domains (MBDs) et al., 2004; Oswald et al., 2009) suggests a function of Cox17 in the
of the copper transporting P-type ATPases ATP7A and ATP7B (Banci transfer of copper from cytosol into IMS. However, the transfer of
et al., 2007, 2008a, 2009; Benitez et al., 2011; Hamza et al., 1999; copper to Sco1, Sco2 and Cox11 is considered as primary function
Hung et al., 1998; Hussain et al., 2009; Ralle et al., 2003; Rodriguez- of Cox17 (Horng et al., 2004; Maxfield et al., 2004). Sco proteins are
Granillo et al., 2010; Tanchou et al., 2004; Wernimont et al., 2004). involved in the formation of the CuA site of cytochrome c oxidase
In addition, Atox1 is involved in the regulation of the copper (Horng et al., 2005; Leary et al., 2004; Mattatall et al., 2000;
mediated intracellular trafficking and catalytic activity of both Stiburek et al., 2009). In humans, the homologs hSco1 and hSco2
ATP7A and ATP7B (Banci et al., 2007; Tanchou et al., 2004). In brain, transfer copper to the binuclear copper center and act as thiol-
Atox1 is expressed ubiquitously (Davies et al., 2013; Nishihara disulphide oxidoreductases (Horng et al., 2005; Leary et al., 2004,
et al., 1998). While expression levels in the brain were found to be 2009a). Consistent with such functions, mutations in Sco1 and
highest in the choroid plexus (Nishihara et al., 1998), in brain Sco2 cause severe deficits of cytochrome c oxidase (Diaz, 2010;
parenchyma strongest expression was reported in the substantia Horng et al., 2005; Leary et al., 2007; Stiburek et al., 2009). In
nigra (Davies et al., 2013). addition, hSco1 and hSco2 have been implicated in the mainte-
The copper chaperone for superoxide dismutase (CCS) is found nance of cellular copper homeostasis (Leary et al., 2007, 2009b;
widely distributed throughout eukaryotes (Casareno et al., 1998; Stiburek et al., 2009; Stiburek and Zeman, 2010), especially in the
Leitch et al., 2009b; Rae et al., 2001). In human brain, CCS is regulation of copper efflux (Leary et al., 2007). The CuB site of
expressed at a similar level in various brain regions, with the cytochrome c oxidase receives its copper ion from Cox11 (Carr
exception of white matter from where it is largely absent et al., 2002; Hiser et al., 2000). Accordingly, cells lacking functional
(Rothstein et al., 1999). On a cellular level CCS was found to be Cox11 have impaired cytochrome c oxidase activity (Banting and
more abundant in neurons than in astrocytes (Rothstein et al., Glerum, 2006; Carr et al., 2002; Hiser et al., 2000; Thompson et al.,
1999). CCS is involved in the maturation of SOD1 by inserting 2010).
copper into this enzyme (Culotta et al., 1997; Furukawa et al., Since the insertion of copper ions into cytochrome c oxidase
2004; Leitch et al., 2009b; Rae et al., 2001). Although SOD1 from takes place within the mitochondrial IMS it requires the delivery
most species can be activated independently of CCS, maximal of copper into this mitochondrial compartment. Mitochondria
SOD1 activity relies on the presence of CCS (Leitch et al., 2009a, have been shown to contain a significant matrix pool of non-
2009b; Rae et al., 2001). Primarily localized to the cytosol and protein bound copper which has been suggested to supply the IMS
nucleus of cells, CCS possesses a similar cellular distribution as its with copper (Cobine et al., 2004; Maxfield et al., 2004). Recently,
target protein SOD1 (Casareno et al., 1998; Culotta et al., 1997). The the mitochondrial carrier family protein Pic2 has been identified
cellular expression level of CCS depends on the cellular copper in yeast as the first protein involved in the import of copper into
content (Bertinato and L’Abbe, 2003; Caruano-Yzermans et al., the mitochondrial matrix (Vest et al., 2013). Other components of
2006; Prohaska et al., 2003) and copper deficiency has been the machinery by which copper is shuttled into mitochondria
demonstrated to result in an increase in CCS protein abundance remain unknown.
due to a lowered rate of proteosomal CCS degradation (Bertinato
and L’Abbe, 2003; Caruano-Yzermans et al., 2006). The activation 3.5. Copper export
of SOD1 by CCS is not a simple copper transfer reaction, but an
oxygen-dependent redox process in which copper incorporation is Copper export from mammalian cells relies on the function of
accompanied by the generation of the critical intramolecular two proteins, ATP7A and ATP7B. These proteins belong to the protein
disulfide between Cys57 and Cys146 (Brown et al., 2004; Furukawa family of P1B-type ATPases and have key functions in metal
et al., 2004; Stasser et al., 2007). In the heterodimeric CCS-SOD1 homeostasis in organisms of all kind of phyla as they use the
complex, an intermolecular disulfide bond forms between Cys57 of energy of ATP hydrolysis to transport heavy metals across cellular
SOD1 and a cysteine of the CxC motif in domain 3 of CCS is believed membranes (Arguello et al., 2007; Lutsenko et al., 2007; Vulpe et al.,
to facilitate copper insertion by opening the SOD1 active site and is 1993). Both ATP7A and ATP7B mediate Cu+ translocation with
subsequently converted to an intramolecular disulfide in the active apparent KM-values in the low micromolar range (Voskoboinik et al.,
SOD1 (Barry and Blackburn, 2008; Furukawa et al., 2004; Lamb 1998, 1999, 2001a, 2001b) and shuttle copper to the secretory
et al., 2001; Robinson and Winge, 2010). pathway for subsequent incorporation into copper-dependent
42 I.F. Scheiber et al. / Progress in Neurobiology 116 (2014) 33–57

enzymes (El Meskini et al., 2003; Hardman et al., 2007; Kosonen 2007; Schlief et al., 2005). In cerebellum of adult mice, ATP7A is
et al., 1997; Meng et al., 2004; Niciu et al., 2007; Petris et al., 2000; expressed in Bergmann glial cells but not in Purkinje cells (Barnes
Ray et al., 2007; Setty et al., 2008; Steveson et al., 2003; Tchaparian et al., 2005), while in human cerebellum ATP7A has been reported
et al., 2000; Terada et al., 1998; Wang and Hebert, 2006). The to be expressed in Purkinje neurons but not in Bergmann glia
importance of these proteins in the maintenance of copper (Davies et al., 2013). ATP7A has further been shown to be strongly
homeostasis is dramatically illustrated by the human genetic expressed in small-sized neurons of the rat dorsal root ganglion
disorders Menkes and Wilson’s disease that are caused by mutations that are the primary site of substance P synthesis in this tissue (Ip
in ATP7A and ATP7B, respectively (Huster, 2010; Kaler, 2011; et al., 2010). Since the biosynthesis of substance P requires PAM the
Kodama et al., 2011; Pfeiffer, 2007; Scheinberg and Sternlieb, 1996; strong expression of ATP7A in these neurons is thought to meet
Tumer and Moller, 2010). their high demand for copper delivery to PAM (Ip et al., 2010). In
Human ATP7A and ATP7B are homologous multispanning addition to its well-known role in the export and delivery of copper
membrane proteins (Bull et al., 1993; Hung et al., 1997; Payne and to cuproenzymes in peripheral tissues, ATP7A is also required in
Gitlin, 1998; Tanzi et al., 1993; Yamaguchi et al., 1996) that contain the brain to release copper from hippocampal neurons upon
a cytosolic N-terminus, eight transmembrane helices, an ATP- NMDA-activation (Schlief et al., 2005). NMDA-receptor mediated
binding domain, an actuator domain, and a cytosolic C-terminus Ca2+-entry induces the identical trafficking of ATP7A to a
(Barry et al., 2010; Boal and Rosenzweig, 2009; Lutsenko et al., cytoplasmic compartment adjacent to the plasma membrane as
2007). When copper levels are within the physiological range, both observed with copper (Schlief et al., 2005, 2006) where ATP7A is
ATP7A and ATP7B display steady state localization at the trans- thought to accumulate copper into a membrane bound compart-
Golgi network (TGN) where they transport copper from the cytosol ment, forming and/or replenishing a pool of potentially releasable
to the TGN lumen for incorporation into copper-dependent copper (Schlief et al., 2005; Schlief and Gitlin, 2006). Moreover, an
enzymes (Hung et al., 1997; La Fontaine et al., 1998; Petris and ATP7A-mediated copper efflux has been suggested to play a role in
Mercer, 1999; Petris et al., 1996; Yamaguchi et al., 1996). Using an neuronal process outgrowth and/or synaptogenesis of maturing
extracellular myc epitope tag ATP7A has been shown to continu- olfactory receptor neurons (El Meskini et al., 2005). However, a
ously recycle between the TGN and the plasma membrane and/or a secretion of copper into the extracellular space has not been
cytosolic vesicular compartment (Petris and Mercer, 1999; Petris observed during the formation of neuritic processes (Finney et al.,
et al., 1996). Rising cytosolic copper levels stimulate the 2007, 2009).
translocation of both ATP7A and ATP7B from the TGN to the Compared to ATP7A, the function of ATP7B in the brain is less
plasma membrane and/or to a distinct cytosolic vesicular clear. ATP7B is expressed in many brain regions (Bull et al., 1993;
compartment in close proximity to the plasma membrane (Forbes Niciu et al., 2007; Tanzi et al., 1993; Yamaguchi et al., 1996) and has
and Cox, 2000; Hung et al., 1997; La Fontaine et al., 1998; Nyasae been observed in brain capillary endothelial cells, at the apical
et al., 2007; Pase et al., 2004; Petris and Mercer, 1999; Petris et al., surface of choroidal epithelial cells, in ependymal cells as well as in
1996; Voskoboinik et al., 1998, 1999). When cellular copper levels Purkinje neurons of adult mice (Barnes et al., 2005; Davies et al.,
return to normal, ATP7A and ATP7B relocate back to the TGN 2013; Niciu et al., 2006; Qian et al., 1998; Zheng and Monnot,
(Nyasae et al., 2007; Pase et al., 2004; Petris et al., 1996). Copper- 2012). The expression of ATP7B in Purkinje neurons is consistent
induced redistribution in their cellular localization is linked to the with the expression of Cp in these cells (Barnes et al., 2005). The
ability of ATP7A and ATP7B to export copper (La Fontaine and pineal gland and adult retina express an alternative-splice form of
Mercer, 2007). Therefore, mutations that impair the copper- ATP7B, the pineal gland night-specific ATPase (PINA), which lacks
dependent trafficking of these proteins have been associated with the N-terminal MBDs and the first four membrane segments
Menkes and Wilson’s disease (Forbes and Cox, 2000; Kim et al., (Borjigin et al., 1999). Although representing only the C-terminal
2003). half of ATP7B, PINA has been shown to possess some copper
ATP7A is crucial for the supply of the brain with copper as transport activity (Borjigin et al., 1999). The pineal gland is a
demonstrated by the low brain copper levels observed in Menkes functional compartment of the circadian timing system and the
disease patients (Nooijen et al., 1981; Willemse et al., 1982) and in expression of PINA is 100-fold higher at night than at daytime
animal models of this disease (Hunt, 1974; Kumode et al., 1993). (Borjigin et al., 1999), suggesting a function of rhythmic copper
ATP7A was shown to be widely expressed in brain (Chelly et al., metabolism in circadian rhythm. Interestingly, mice subjected to
1993; Davies et al., 2013; Niciu et al., 2006, 2007; Nishihara et al., total darkness for up to 60 days showed a marked elevation in their
1998; Paynter et al., 1994; Vulpe et al., 1993). In human brain, brain copper content (Beltramini et al., 2004).
highest levels of ATP7A were found in the cerebellum and on the Further information on the roles of ATP7A and ATP7B in the
basolateral surface of the polarized choroid plexus cells (Davies copper homeostasis of the brain can be found in a very recent
et al., 2014). In contrast to this observation, Kaler has proposed that review (Telianidis et al., 2013).
ATP7A should be apically located in the choroid plexus cells based
on functional considerations (Kaler, 2011). In mouse brain high 4. Astrocyte-neuron coupling in copper metabolism
levels of ATP7A were found in the choroid plexus and Purkinje
neurons (Niciu et al., 2006). Irrespective of the exact location of Astrocytes, which constitute the main class of neuroglia, are the
ATP7A in the choroid plexus the high expression of this copper most abundant cells in the brain (Markiewicz and Lukomska, 2006;
ATPase is consistent with an important role of choroid plexus cells Sofroniew and Vinters, 2010). These cells are distributed throughout
in the copper homeostasis of the brain. the entire brain and fulfill a range of important functions essential
A dysfunction of ATP7A has been shown to result in copper for brain physiology (Nedergaard et al., 2003; Parpura et al., 2012;
accumulation in brain capillaries of macular and brindled mutant Sofroniew and Vinters, 2010). Among other functions, astrocytes
mice (Kodama, 1993; Yoshimura, 1994), indicating that ATP7A have been proposed to be involved in the control of brain
plays a role in copper transport across the BBB. The importance of extracellular ion homeostasis, providing metabolic support of
ATP7A in copper export from brain capillary endothelial cells has neurons, the maintenance of the BBB as well as in the modulation
also been demonstrated in a cell culture model for these cells (Qian of synaptic transmission and synaptic plasticity (Barker and Ullian,
et al., 1998). In brain parenchyma, ATP7A has been shown to be 2010; Barros and Deitmer, 2010; Deitmer and Rose, 2010; Dienel,
expressed in both neurons and non-neuronal cells (Barnes et al., 2012; Hirrlinger and Dringen, 2010; Nedergaard et al., 2003; Pellerin
2005; Davies et al., 2013; El Meskini et al., 2005; Niciu et al., 2006, et al., 2007; Perea and Araque, 2010). Moreover astrocytes have
I.F. Scheiber et al. / Progress in Neurobiology 116 (2014) 33–57 43

by uptake into astrocytes this protection of neurons against copper


toxicity could involve the release of antioxidants or copper
chelating substances by astrocytes that prevent the copper-
mediated inactivation/degradation of extracellular GSH (Pope
et al., 2008). Since trafficking of GSH from astrocytes to neurons is
essential to maintain neuronal GSH levels (Dringen et al., 1999;
Hirrlinger and Dringen, 2010), this stabilization of extracellular
GSH may also have neuroprotective functions (Pope et al., 2008).
Alterations of brain GSH (Adams et al., 1991; Sian et al., 1994;
Fig. 3. Proposed model of astrocyte-neuron coupling in copper metabolism. In the Aoyama and Nakaki, 2013) and energy metabolism (Fukuyama
brain, copper is efficiently taken up into astrocytes via Ctr1 and other, yet unknown, et al., 1994; Mosconi et al., 2008) have been linked to
transporters. Excess copper is stored by astrocytes in metallothioneins (MT) or as neurodegenerative diseases such as Alzheimer’s and Parkinson’s
GSH-complex. Copper is released from astrocytes by ATP7A to supply neurons with disease. In this context, disturbances of astrocytic functions have to
copper.
be considered, since astrocytes provide neighboring neurons with
energy substrates (Barros and Deitmer, 2010; Dienel, 2012;
been considered to play a key role in the metal metabolism of the Pellerin et al., 2007) and with GSH precursors (Hirrlinger and
brain (Dringen et al., 2007, 2013; Scheiber and Dringen, 2013; Dringen, 2010; Schmidt and Dringen, 2012; Aoyama and Nakaki,
Tiffany-Castiglion and Qian, 2001; Tiffany-Castiglioni et al., 2011). 2013). Since both glyolytic flux (Scheiber and Dringen, 2011b) and
Astrocytes possess several features that allow them to function GSH export (Scheiber and Dringen, 2011a) are accelerated at least
as regulators for the uptake and distribution of essential metals to in copper-treated cultured astrocytes, the presence of an excess of
other types of brain cells and to serve as metal depots (Dringen copper in the brain may also cause an increase in the extracellular
et al., 2007, 2013; Scheiber and Dringen, 2013; Tiffany-Castiglioni concentrations of astrocyte-derived lactate and GSH. Whether
et al., 2001, 2011). These cells have a strategically important such alterations in the metabolism of astrocytes are neuroprotec-
location in the brain, being in close contact to neuronal cell bodies tive or may contribute to the progression of neurodegenerative
and to capillary endothelial cells via their cytoplasmic processes diseases that have been connected with an elevated copper content
that terminate in so-called end-feet (Demeuse et al., 2002; in brain remains to be elucidated.
Nedergaard et al., 2003). Astrocytic end-feet cover most of the
brain capillaries (Demeuse et al., 2002; Mathiisen et al., 2010; 5. Copper and neurodegenerative diseases
Nedergaard et al., 2003) and have been shown to express metal
transport proteins such as DMT1 (Burdo et al., 2001) or ferroportin Menkes and Wilson’s disease (Kodama et al., 2011; Lorincz,
(Wu et al., 2004). Thus, astrocytes are the first brain parenchyma 2010) are the best examples for pathological conditions that are
cells to encounter metals that cross the BBB. In addition, astrocytes connected with an impaired copper homeostasis that leads to
can express ferritin in an iron-dependent manner (Hoepken et al., neurodegeneration. While many of the clinical symptoms associ-
2004) and contain high cellular contents of MTs (Aschner, 1997; ated with the severe copper deficiency in Menkes disease can be
Aschner et al., 1997; Dineley et al., 2000; Hidalgo et al., 2001) and attributed to a decrease in the activities of copper-dependent
GSH (Dringen and Hamprecht, 1998; Hirrlinger and Dringen, enzymes, copper toxicity in Wilson’s disease is most likely a
2010), endowing them with a high capacity to store and to prevent consequence of the redox activity of copper (Britton, 1996; Burkitt,
the toxicity of metals and of metal-induced oxidative stress. 2001; Gunther et al., 1995; Halliwell and Gutteridge, 2007; Uriu-
It is likely that astrocytes play an important role in the copper Adams and Keen, 2005) and direct inhibition of protein functions
homeostasis of the brain (Fig. 3). Within the brain copper has been (Boulard et al., 1972; Letelier et al., 2005, 2006; Pamp et al., 2005;
shown histochemically to concentrate in astrocytes and it has been Schwerdtle et al., 2007). Alterations of copper homeostasis have
suggested that astrocytes may regulate the copper supply to also been associated with neurodegenerative diseases such as
neurons (Kodama, 1993; Kodama et al., 1991; Szerdahelyi and prion diseases, Alzheimer’s disease, Parkinson’s disease and
Kasa, 1986). Consistent with such function, astrocytes efficiently Huntington’s disease (Gaggelli et al., 2006; Greenough et al.,
accumulate copper in vitro (Brown, 2004; Kodama et al., 1991; 2013; Liddell et al., 2013; Rivera-Mancia et al., 2010). However, in
Scheiber et al., 2010a, 2010b) and in vivo (Haywood et al., 2008; contrast to Menkes and Wilson’s disease the role of copper in these
Haywood and Vaillant, in press) and have the ability to export diseases is not fully understood.
copper (Scheiber et al., 2012). Copper accumulation by astrocytes
most likely involves both, Ctr1-mediated and Ctr1-independent 5.1. Menkes disease
transport processes (Scheiber et al., 2010a), while ATP7A is likely to
contribute to the copper export from astrocytes (Kodama, 1993; Menkes disease is a rare X-linked inherited disease (Tumer and
Kodama et al., 1991; Scheiber et al., 2012). Moller, 2010; Tümer, 2013). It results from a deficit in ATP7A
In cell culture, astrocytes have been reported to take up copper activity that leads to a failure of delivery of copper from intestine
more efficiently than cultured neurons and to protect neurons epithelia into the blood stream and consequently is characterized
from copper toxicity (Brown, 2004). In addition to their capacity to by a general copper deficit (Kaler, 2011; Kodama et al., 2011;
efficiently accumulate extracellular copper and to their high Tumer and Moller, 2010). As a consequence of the reduced delivery
antioxidative potential, astrocytes are able to upregulate their of copper to the brain, Menkes disease patients exhibit severe
storage capacity for copper upon copper exposure, which is likely mental and developmental impairment that are in part due to a
to contribute to the protection of other brain cells against copper- decrease in the activity of copper-dependent enzymes (Kaler,
induced toxicity. For example, the elevated brain copper content in 2011; Kodama et al., 2011; Tumer and Moller, 2010). Besides
the North Ronaldsay sheep was accompanied by copper accumu- severe progressive neurological degeneration, clinical features of
lation in astrocytes and by a strong astrocytic immunoreactivity for Menkes disease include connective tissue abnormalities, muscular
MTs (Haywood et al., 2008; Haywood and Vaillant, in press). hypotonia and hypopigmentation of skin and hair (Kaler, 2011;
Similarly, cultured astrocytes that have been exposed to copper Kodama et al., 2011; Tumer and Moller, 2010).
upregulate their cellular contents of MTs and GSH (Dringen et al., Impaired mitochondrial function has been discussed as major
2013; Scheiber and Dringen, 2013). Besides the removal of copper factor in the devastating brain damage associated with Menkes
44 I.F. Scheiber et al. / Progress in Neurobiology 116 (2014) 33–57

disease (Kaler, 1998). Deficient cytochrome c oxidase activity has accumulation by cultured astrocytes and its toxic consequences
been observed in the brain of Menkes disease patients (Maehara are lowered by zinc (Scheiber et al., 2010b).
et al., 1983) and animal models of this disease (Yoshimura et al.,
1993). Also elevated brain lactate levels and decreased content of 5.3. Alzheimer’s disease
N-acetylaspartate in brain, the synthesis of which is coupled to
mitochondrial energy production in neurons (Moffett et al., 2007), Alzheimer’s disease (AD) is a progressive neurodegenerative
are further indicators for a failure of oxidative phosphorylation disease that causes memory loss and psychiatric disturbances
(Rusinko and Prohaska, 1985). Also decreased PAM activity and (Castellani et al., 2010; Salawu et al., 2011). It represents the most
low levels of several a-amidated peptides have been observed in common neurodegenerative disease in humans, with currently an
the brains of animal models of Menkes disease (Bousquet-Moore estimate of 18–25 million people worldwide suffering from this
et al., 2010b; Hansel et al., 2001; Steveson et al., 2003). The lack of disease (Ballard et al., 2011; Castellani et al., 2010; Kenche and
amidated peptides due to decreased PAM activity is though to Barnham, 2011). The major risk factor in AD is advanced age. Thus,
contribute to the neurodevelopmental delay and increased seizure as a direct consequence of increased life span expectancies, the
frequency associated with Menkes disease (Bousquet-Moore et al., number of people suffering from AD is projected to increase
2010b; Kaler, 2011; Steveson et al., 2003). An elevated dopamine to dramatically within the next decades (Ballard et al., 2011;
noradrenaline ratio has also been observed in the plasma of Castellani et al., 2010; Kenche and Barnham, 2011; Salawu
Menkes disease patients as well as in the brains of animal models et al., 2011). Aside from age, other risk factors include genetic
of Menkes disease (Goldstein et al., 2009; Kaler et al., 2008; factors, gender and environmental factors (Ballard et al., 2011;
Prohaska and Smith, 1982), indicating a partial deficiency of DbM Castellani et al., 2010). Extracellular senile plaques in brain and
under severe copper deficiency conditions. However, despite the intracellular neurofibrillary tangles are considered as pathological
fundamental role of noradrenaline in CNS function, the fact that hallmarks of AD (Ballard et al., 2011; Castellani et al., 2010;
patients with congenital DbM deficiency exhibit only subtle signs Nalivaeva and Turner, 2013; Pooler et al., 2013). Senile plaques
of CNS dysfunction (Kim et al., 2002; Robertson et al., 1991) consist predominately of amyloid-b (Ab) polypeptides of lengths
suggest that DbM deficiency is unlikely to be a major contributor of 40 and 42 amino acid residues. These peptides are generated by
to the neurodegeneration in Menkes disease patients. consecutive proteolytic cleavage of the integral membrane
Parental administrations of copper compounds are used for amyloid precursor protein (APP) by b- and g-secretases (Ballard
treatment of Menkes disease and may improve clinical outcome, at et al., 2011; Borchardt et al., 1999; Budimir, 2011; Castellani et al.,
least if initiated shortly after birth (Kaler, 2011; Kodama et al., 2010; Tougu et al., 2011).
2011; Tumer and Moller, 2010). Newborns with Menkes disease APP is able to bind copper with nanomolar affinity (Hesse et al.,
with residual ATP7A activity are more likely to respond to this 1994; Kong et al., 2008) and to catalytically reduce Cu2+ (Multhaup
treatment than those with complete loss-of-function mutations et al., 1996; Ruiz et al., 1999). Additionally, both expression
(Kaler, 2011). Recently, brain-directed ATP7A gene addition in a (Bellingham et al., 2004b; Borchardt et al., 1999) and localization
mouse model of Menkes disease has shown promising results (Acevedo et al., 2011) of APP are regulated by intracellular copper
(Donsante et al., 2011). In these experiments the AAV-5 vector levels. Thus, beside numerous other physiological functions that
directed the expression of ATP7A to the choroid plexus, and this was have been ascribed to APP (Nalivaeva and Turner, 2013), APP has
sufficient to increase the survival of the mutant mice, consistent also been implicated in the regulation of copper homeostasis.
with the choroid plexus playing an important role in brain copper Indeed, APP levels have been shown to modulate brain copper
homeostasis (Donsante et al., 2011). Thus, gene therapy that contents. While APP knock-out mice contain elevated copper
restores at least low levels of functional ATP7A may become an contents in the cerebral cortex (White et al., 1999b), the brains of
alternative treatment option in the future (Donsante et al., 2011; mice overexpressing APP contain lower copper levels than control
Kaler, 2011). mice (Maynard et al., 2002). However, on the cellular level the role
of APP in copper homeostasis is yet not clear. Consistent with a
5.2. Wilson’s disease proposed role of APP in cellular copper efflux (Bayer and Multhaup,
2005; Treiber et al., 2004), the knock-out of APP increased copper
Wilson’s disease is an inherited disorder connected with accumulation and overexpression of APP decreased copper
hepatic, ophthalmologic, neurological and/or psychiatric symp- accumulation in cultured primary cortical neurons (Bellingham
toms that originates from a genetic defect in the copper et al., 2004a). On the other hand, overexpression of APP in HEK293
transporting ATPase ATP7B (Huster, 2010; Pfeiffer, 2007; Purchase, cells increased copper accumulation which has been attributed to
2013; Scheinberg and Sternlieb, 1996). Due to impaired ATP7B the APP-mediated increase in their capacity to reduce extracellular
function, copper cannot be secreted from hepatocytes into the bile Cu2+ (Suazo et al., 2009). Also there is conflicting evidence for the
and accumulates in the liver and other tissues, including the brain, role of APP in copper-induced toxicity. While APP levels in mouse
leading to chronic copper toxicity (Huster, 2010; Pfeiffer, 2007; primary cortical neurons correlated with an increased sensitivity
Purchase, 2013; Scheinberg and Sternlieb, 1996). The occurrence of to copper (White et al., 1999a, 2002), APP has been reported to
abnormal astrocytes, i.e. Alzheimer type I and II cells as well as protect neurons in rat hippocampus in vivo (Cerpa et al., 2004) and
Opalski cells is a typical neuropathological feature of Wilson’s HEK293 cells (Suazo et al., 2009) from copper toxicity.
disease (Bertrand et al., 2001; Mossakowski et al., 1970). These Also Ab peptides have been shown to bind copper with high
morphological changes of normal astrocytes have been discussed affinity (Atwood et al., 2000) and senile plaques are strongly
as consequence of a strong copper accumulation (Mossakowski enriched in copper (Lovell et al., 1998). In addition, copper ions
et al., 1970). induce the precipitation of Ab peptides in vitro (Atwood et al.,
Wilson’s disease is commonly treated by systemic chelation 2000; Faller et al., 2013). These observations suggest that copper
therapy with D-penicillamine or triethylenetetramine, which tends triggers the formation of plaques in brain (Atwood et al., 2000;
to restore and maintain copper homeostasis in the body (Huster, Kenche and Barnham, 2011; Roberts et al., 2012; Tougu et al.,
2010; Pfeiffer, 2007). For the treatment of pre-symptomatic 2011). Indeed, Ab deposition begins at glutamatergic synapses
patients zinc is applied which aims to limit copper absorption (Cater et al., 2008), where both copper (Hopt et al., 2003; Kardos
in the intestine (Huster, 2010; Pfeiffer, 2007). Zinc may also help to et al., 1989) and the Ab peptide (Lesne et al., 2005; Tougu et al.,
limit copper accumulation into brain cells. At least the copper 2011) are released during synaptic transmission. A soluble
I.F. Scheiber et al. / Progress in Neurobiology 116 (2014) 33–57 45

oligomeric Ab species is now considered as the most toxic form of et al., 2007), the physiological roles of the normal PrP are still under
Ab peptides (Karran et al., 2011; Roberts et al., 2012). Since Ab can discussion. Among others it has been implicated in the protection
reduce Cu2+ to Cu+, copper-induced ROS formation may be against apoptosis and oxidative stress, in transmembrane signaling
involved in the promotion of the toxicity of Ab oligomers (Eskici as well as in the formation and function of synapses (Brown and
and Axelsen, 2012; Huang et al., 1999; Kenche and Barnham, 2011; Sassoon, 2002; Hu et al., 2007; Linden et al., 2008; Rachidi et al.,
Multhaup et al., 2002; Robertson et al., 1991). In addition, copper 2003; Westergard et al., 2007).
has been reported to increase the inhibition of cytochrome c Several lines of evidence suggest a role for PrP in the protection
oxidase by Ab (Crouch et al., 2005) and copper-bound Ab-species against copper-induced toxicity and in normal cellular copper
have been demonstrated to disrupt the K+-homeostasis in cultured homeostasis. Cells expressing PrP are much more resistant to
cortical neurons and to induce changes in axonal morphology copper treatment than PrP-deficient cells (Brown et al., 1998;
(Howells et al., 2012). Haigh and Brown, 2006; Rachidi et al., 2003; Canello et al., 2012;
Although the copper-induced enhancement of Ab-toxicity in Wang et al., 2011). Since PrP is known to bind copper with low
vitro suggest a detrimental role of copper in AD, the lowered brain nanomolar to micromolar affinity (Brown et al., 1997; Westergard
copper contents in AD patients (Deibel et al., 1996; Loeffler et al., et al., 2007; Nadal et al., 2009) it may protect against copper-
1996) and mouse models for AD (Bayer et al., 2003; Maynard et al., induced toxicity by capturing extracellular copper and thereby will
2002) suggest that also a local copper deficit may contribute to the lower copper-mediated ROS production (Brown et al., 1998; Haigh
neurodegeneration. Very recently was reported that the lowered and Brown, 2006; Millhauser, 2007; Rachidi et al., 2003).
copper level in AD brain is due to a lower content of soluble, Alternatively, PrP may protect against copper-mediated ROS
extractable copper (Rembach et al., 2013). The view that copper production by its suggested SOD-like function (Brown et al.,
deprivation may be connected with AD is supported by the 1999; Treiber et al., 2007).
improved survival after copper supplementation of mouse models Based on the finding that endocytosis of PrP is stimulated by
of AD (Bayer et al., 2003) and by the enhanced cognitive functions copper, PrP has been suggested to serve as a receptor for cellular
after application of Cu(gtsm) as copper source (Crouch et al., 2009). uptake of copper (Brown and Harris, 2003; Pauly and Harris, 1998).
However, a phase 2 clinical trial with copper supplementation did However, while some studies reported that the rate of cellular
not show any effect on cognition in patients with mild AD, copper uptake increases with the expression level of PrP (Brown,
although a positive effect on a relevant AD biomarker was observed 2004; Urso et al., 2010, 2012) and the copper contents of mouse
(Kessler et al., 2008a,b). brain fractions have been observed to increase with increasing
Mechanistically, copper deficiency may accelerate the amyloi- abundance of PrP (Brown et al., 1997) other studies did not find any
dogenic pathway by influencing APP-processing (Cater et al., alteration in the copper content of brain fractions of mice
2008). Indeed, copper treatment of cell lines has been shown to expressing different levels of PrP (Giese et al., 2005; Waggoner
lower Ab synthesis (Borchardt et al., 1999) and to up-regulate Ab et al., 2000) or failed to show a contribution of PrP to cellular
degrading metalloproteinases (White et al., 2006). In addition, copper uptake (Rachidi et al., 2003).
copper deficiency in AD may also compromise energy metabolism As PrP has been shown to reduce Cu2+ to Cu+ with maximal
and antioxidative defence, as lower activities of cytochrome c activity at pH 6.5 (Miura et al., 2005), it has been concluded that the
oxidase (Maurer et al., 2000; Mutisya et al., 1994; Parker et al., PrP does not only transport extracellular copper into endosomes,
1994) and SOD1 (Chen et al., 1994; Marcus et al., 1998) have been but also reduces Cu2+ prior to its release into the cytosol (Miura
reported for the AD brain. et al., 2005). In addition, PrP has been suggested to play a role in
In view of the high copper contents in senile plaques (Lovell regulating copper release at the synapse, as the copper content of
et al., 1998) and the overall low copper content in AD brain, it is synaptosomes prepared from PrP-deficient mice was lower than
now widely acknowledged that copper is abnormally redistributed that of synaptosomes from wild type mice (Brown et al., 1997).
to senile plaques in AD, leaving the tissue and cells deficient in Finally, the expression levels of PrP have been shown to influence
copper (Kaden et al., 2011; Macreadie, 2008; Roberts et al., 2012). the expression of several proteins involved in cellular copper
Therapeutic strategies aiming to restore the normal copper homeostasis such as DMT-1, Ctr2, ATP7A and Atox1 (Kralovicova
distribution in AD brain are currently under investigation (Kenche et al., 2009), further supporting an important role for PrP in cellular
and Barnham, 2011; Zatta et al., 2009). The copper ionophore PBT- copper homeostasis.
2, which has been shown to reduce senile plaques and to enhance An alteration of copper homeostasis may also play a role in the
brain copper levels (Adlard et al., 2008), has recently been shown development and progression of prion diseases (Leach et al., 2006;
to improve AD biomarkers as well as cognition in AD patients in a Millhauser, 2007; Rana et al., 2009). Brain copper contents of
phase 2 clinical trial (Lannfelt et al., 2008). scrapie infected mice were found to be decreased by around 60%,
pointing to a severe copper deficit in prion diseases (Mitteregger
5.4. Prion diseases et al., 2009; Thackray et al., 2002; Wong et al., 2001). Consistently,
copper supplementation of these mice slowed the progression of
Prion diseases constitute a group of disorders that have been scrapie (Mitteregger et al., 2009). In contrast to this, the onset of
associated with the refolding of the prion protein (PrP) into an prion diseases was reported to be delayed by copper chelation
abnormal pathogenic conformation (PrPSc). Prion diseases include (Sigurdsson et al., 2003) and in mice with severe brain copper
rare but fatal transmissible prion diseases or spongiform deficit due to a mutation in ATP7A (Siggs et al., 2012). However,
encephalopathies, such as Creutzfeld–Jacob disease, Kuru or fatal while copper contents were strongly decreased, manganese
familial insomnia in humans (Acquatella-Tran Van Ba et al., 2013; contents were shown to be elevated in prion-infected brain tissue
Aguzzi et al., 2008; Cobb and Surewicz, 2009; Holman et al., 2010; (Mitteregger et al., 2009; Thackray et al., 2002; Wong et al., 2001;
Moore et al., 2009). Creutzfeld–Jacob disease has an annual Johnson et al., 2013) and manganese supplementation accelerated
incidence rate of about 1 case per 1,000,000 and is characterized by disease progression in scrapie infected mice (Mitteregger et al.,
a rapidly progressive dementia and neurological features including 2009). Since manganese bound PrP is able to induce polymeriza-
ataxia, tremor, bradykinesia and rigidity, with death occurring tion of PrP (Brazier et al., 2008), copper deficiency was hypothe-
within 1–2 year of disease onset (Aguzzi et al., 2008; Brown and sized to foster the development of prion diseases by favoring the
Mastrianni, 2010; Cobb and Surewicz, 2009). Although ubiqui- incorporation of manganese into PrP instead of copper (Brown
tously expressed (Hu et al., 2007; Linden et al., 2008; Westergard et al., 2000). Indeed it was reported recently that prion plaques in
46 I.F. Scheiber et al. / Progress in Neurobiology 116 (2014) 33–57

hamsters infected with transmissible spongiform encephalopa- potential to generate radicals, defects of copper homeostasis might
thies were depleted of copper while they were enriched in yet prove to be of fundamental importance in this disease. In this
manganese (Johnson et al., 2013). However, whether the observed context it is important to mention that a-synuclein has been
perturbation of copper and manganese in the brain are a cause or a reported to have cellular ferrireductase activity which generates
consequence of prion diseases remains to be evaluated (Rana et al., ferrous iron and that this activity depends on the presence of
2009). bound copper (Davies et al., 2011b).

5.5. Parkinson’s disease 5.6. Huntington’s disease

With an overall prevalence ranging between 0.1% and 0.3%, Huntington’s disease (HD) is a rare autosomal-dominant,
Parkinson’s disease (PD) is the second most common neurodegen- progressive neurodegenerative disease that results in motor,
erative disease in humans (Wirdefeldt et al., 2011). PD is cognitive and psychiatric symptoms (Anderson, 2011; Shannon,
characterized by a complex motor disorder known as Parkinson- 2011). Molecular base of HD is an expansion of the cytosine-
ism that manifests with resting tremor, bradykinesia, rigidity and adenine-guanine (CAG) repeat in exon 1 of the huntingtin gene
postural instability, but clinical symptoms also include cognitive which will generate a huntingtin protein that contains a N-
and psychiatric problems (Breen and Barker, 2010; Ferrer et al., terminus with an expanded polyglutamine domain (McFarland
2011; Thomas and Beal, 2007). Pathological hallmarks of PD and Cha, 2011; Bordelon, 2013). The length of this extension
include a loss of neuromelanin containing dopaminergic neurons determines the onset as well as the rate of the clinical progression
in the substantia nigra pars compacta and the presence of Lewy of HD (Anderson, 2011; McFarland and Cha, 2011; Shannon, 2011).
bodies which are an aggregated form of the protein a-synuclein Among others, aggregation of the mutant huntingtin protein,
(Bendor et al., 2013; Ferrer et al., 2011; Jomova et al., 2010; Sian- oxidative stress, impaired energy metabolism, loss of neurotrophic
Hulsmann et al., 2011; Thomas and Beal, 2007). The majority of support and transcriptional dysregulation have been discussed to
cases are idiopathic with less than 10% of PD having a strict familial contribute to the development and progression of HD (Chen, 2011;
etiology (Jomova et al., 2010; Sian-Hulsmann et al., 2011; Thomas Gauthier et al., 2004; McFarland and Cha, 2011).
and Beal, 2007; Wirdefeldt et al., 2011). Mitochondrial dysfunc- The brain copper levels of HD patients (Dexter et al., 1991) and
tion, oxidative stress and inflammation have been suggested in the of a mouse model for HD (Fox et al., 2007) have been found to be
etiology of idiopathic PD (Jomova et al., 2010; Thomas and Beal, elevated compared to controls. Huntingtin binds to copper and
2007), but the underlying mechanisms of idiopathic PD remain to the complex formed is redox active (Fox et al., 2007). Of
be elucidated. significance for the pathogenesis of HD is that copper was found
There is strong and growing evidence of abnormalities in to promote the aggregation of huntingtin and the aggregated form
copper homeostasis in PD. Parkinsonism is a frequent symptom in with the expanded polyglutamine repeat is thought to be
neurological Wilson’s disease (Lorincz, 2010). Copper has been neurotoxic (Fox et al., 2007). The role of copper in HD has
demonstrated to bind to both soluble and membrane-bound a- received strong support from a very recent study on a Drosophila
synuclein with high affinity (Dudzik et al., 2013) and to accelerate model which demonstrated that altered expression of copper
aggregation of soluble a-synuclein (Davies et al., 2011a; Uversky homeostasis genes could modulate the HD phenotype (Xiao et al.,
et al., 2001). A copper-binding oligomer of a-synuclein has been 2013). Also dietary copper reduction markedly reduced the level
discussed as neurotoxic form of a-synuclein (Brown, 2010). of oligomerized huntingtin, while removal of the putative copper
Several lines of evidence argue for a regional copper deficit binding sites (Met8 and His82) completely ablated the copper
contributing to the pathogenesis of PD. Thus, while the total copper related toxicity of the polyQ huntingtin (Xiao et al., 2013). A
content in brains of PD patients is similar to that of healthy potential role of copper in HD progression is further supported by
controls, copper levels in the substantia nigra of PD patients are the observations that treatment with the copper chelators
substantially lower (Dexter et al., 1989, 1991; Loeffler et al., 1996; tetrathiomolybdate (Tallaksen-Greene et al., 2009) or clioquinol
Popescu et al., 2009). Most recently Davies and coworkers used (Nguyen et al., 2005) delayed the decline in motor function in
synchrotron-based X-ray fluorescence microscopy to show that mouse models for HD. Currently, the copper ionophore PBT-2 is
there was a significant reduction in copper contents in the under clinical investigation for the treatment of HD (Clinical-
substantia nigra and locus coeruleus in PD patients. Moreover, they trials.gov identifier: NCT01590888).
found a significant decrease in the amount of the copper
transporter Ctr1 in neurons of these regions, suggesting some 6. Conclusions and outlook
compromise of copper homeostasis in the vulnerable regions of PD
brains (Davies et al., 2014). More systemic changes in copper Within the last decade the knowledge on the metabolism and
homeostasis have been found in PD patients, e.g. ceruloplasmin functions of copper in brain has dramatically increased. Most of the
levels were found to be low in the CSF of affected individuals key players that are known to be involved in uptake, distribution,
(Olivieri et al., 2011). Using a sensitive stable isotope method, storage and export of copper in peripheral cell types and contribute
Larner and colleagues have found that PD patients showed highly to peripheral copper homeostasis are also present in brain cells.
variable copper measurements, well outside the normal range, However, substantial parts of information on the function and the
again suggesting some underlying defect in copper transport regulation of the various proteins that contribute to the copper
mechanisms in this disorder (Larner et al., 2013). metabolism of the different types of brain cells are still missing.
Copper supplementation has been reported to prevent the Future studies should elucidate in more detail the mechanisms
increase in lipid peroxidation, the striatal dopamine depletion and involved in copper uptake, storage and export in the different types
the inactivation of tyrosine hydroxylase in an animal model for PD of brain cells as well as the interactions of different types of brain
(Alcaraz-Zubeldia et al., 2001, 2009), while copper chelation was cells in brain copper homeostasis. In this context astrocytes are of
not protective in PD animal models (Youdim et al., 2007). Thus, particular interest, as these cells have been proposed to be involved
although the underlying pathological mechanisms of idiopathic PD in the pathology of neurological diseases such as Alzheimer’s
remains unclear, the reported abnormalities in copper metabolism disease and Parkinson’s disease (Halliday and Stevens, 2011;
could in turn lead to defects of iron homeostasis which has been Parpura et al., 2012) and as astrocytes are likely to play a key role in
connected with PD (Double, 2012). As both metals have the the copper metabolism of the brain due to their ability to efficiently
I.F. Scheiber et al. / Progress in Neurobiology 116 (2014) 33–57 47

take up, store and export copper and to even upregulate their Aoyama, J., Nakaki, T., 2013. Impaired glutathione synthesis in neurodegeneration.
Int. J. Mol. Sci. 14, 21044–21921.
copper storage capacity upon exposure to copper (Scheiber and Arguello, J.M., Eren, E., Gonzalez-Guerrero, M., 2007. The structure and function of
Dringen, 2013). heavy metal transport P1B-ATPases. Biometals 20, 233–248.
Disturbances in copper homeostasis have clearly been linked to Armendariz, A.D., Olivares, F., Pulgar, R., Loguinov, A., Cambiazo, V., Vulpe, C.D.,
Gonzalez, M., 2006. Gene expression profiling in wild-type and metallothionein
several neurodegenerative disorders. Both brain copper deficiency, mutant fibroblast cell lines. Biol. Res. 39, 125–142.
for example in Menkes disease and AD, as well as excess of copper, Arredondo, M., Munoz, P., Mura, C.V., Nunez, M.T., 2003. DMT1, a physiologically
for example in Wilson’s disease, have been connected with relevant apical Cu1+ transporter of intestinal cells. Am. J. Physiol. Cell Physiol.
284, C1525–C1530.
neurodegeneration. A better knowledge of the contributions and Aschner, M., 1997. Astrocyte metallothioneins (MTs) and their neuroprotective role.
the functions of the different types of brain cells in the copper Ann. N. Y. Acad. Sci. 825, 334–347.
homeostasis in the brain is greatly needed in order to develop new Aschner, M., Cherian, M.G., Klaassen, C.D., Palmiter, R.D., Erickson, J.C., Bush, A.I.,
1997. Metallothioneins in brain – the role in physiology and pathology. Toxicol.
strategies to reestablish physiologically normal copper levels in
Appl. Pharmacol. 142, 229–242.
copper-related diseases. The next decade is likely to further Ascone, I., Longo, A., Dexpert, H., Ciriolo, M.R., Rotilio, G., Desideri, A., 1993. An X-ray
consolidate the substantial contribution of disturbances of brain absorption study of the reconstitution process of bovine Cu,Zn superoxide
copper metabolism to neurodegenerative disorders and will dismutase by Cu(I)-glutathione complex. FEBS Lett. 322, 165–167.
Atanasiu, R.L., Stea, D., Mateescu, M.A., Vergely, C., Dalloz, F., Briot, F., Maupoil, V.,
hopefully lead to the development of therapeutically effective Nadeau, R., Rochette, L., 1998. Direct evidence of caeruloplasmin antioxidant
drugs which will successfully remove toxic copper from the brain properties. Mol. Cell. Biochem. 189, 127–135.
in copper overload conditions and will supply copper to brain cells Atwood, C.S., Scarpa, R.C., Huang, X., Moir, R.D., Jones, W.D., Fairlie, D.P., Tanzi, R.E.,
Bush, A.I., 2000. Characterization of copper interactions with Alzheimer amy-
in copper deficiency conditions. loid beta peptides: identification of an attomolar-affinity copper binding site on
amyloid b 1-42. J. Neurochem. 75, 1219–1233.
Balamurugan, K., Schaffner, W., 2006. Copper homeostasis in eukaryotes: teetering
Acknowledgments
on a tightrope. Biochim. Biophys. Acta 1763, 737–746.
Ballard, C., Gauthier, S., Corbett, A., Brayne, C., Aarsland, D., Jones, E., 2011. Alzhei-
Ivo F. Scheiber would like to thank the European Social Fund mer’s disease. Lancet 377, 1019–1031.
and the State Budget of the Czech Republic for financing (Project Ballatori, N., Krance, S.M., Marchan, R., Hammond, C.L., 2009. Plasma membrane
glutathione transporters and their roles in cell physiology and pathophysiology.
no. CZ.1.07/2.3.00/30.0061). Ralf Dringen and Julian Mercer would Mol. Aspects Med. 30, 13–28.
like to thank the Deakin University for financial support by the Banci, L., Bertini, I., Cantini, F., Della-Malva, N., Migliardi, M., Rosato, A., 2007. The
Thinkers in Residence Program. different intermolecular interactions of the soluble copper-binding domains of
the menkes protein, ATP7A. J. Biol. Chem. 282, 23140–23146.
Banci, L., Bertini, I., Cantini, F., Massagni, C., Migliardi, M., Rosato, A., 2009. An NMR
study of the interaction of the N-terminal cytoplasmic tail of the Wilson disease
References protein with copper(I)-HAH1. J. Biol. Chem. 284, 9354–9360.
Banci, L., Bertini, I., Cantini, F., Rosenzweig, A.C., Yatsunyk, L.A., 2008a. Metal
Acevedo, K.M., Hung, Y.H., Dalziel, A.H., Li, Q.X., Laughton, K., Wikhe, K., Rembach, A., binding domains 3 and 4 of the Wilson disease protein: solution structure and
Roberts, B., Masters, C.L., Bush, A.I., Camakaris, J., 2011. Copper promotes the interaction with the copper(I) chaperone HAH1. Biochemistry 47, 7423–
trafficking of the amyloid precursor protein. J. Biol. Chem. 286, 8252–8262. 7429.
Acquatella-Tran Van Ba, I., Imberdis, T., Perrier, V., 2013. From prion diseases to Banci, L., Bertini, I., Ciofi-Baffoni, S., Hadjiloi, T., Martinelli, M., Palumaa, P., 2008b.
prion-like propagation mechanisms of neurodegenerative diseases. Int. J. Cell Mitochondrial copper(I) transfer from Cox17 to Sco1 is coupled to electron
Biol. 2013, 975832. transfer. Proc. Natl. Acad. Sci. U.S.A. 105, 6803–6808.
Adams Jr., J.D., Klaidman, L.K., Odunze, I.N., Shen, H.C., Miller, C.A., 1991. Alzheimer’s Banci, L., Bertini, I., Ciofi-Baffoni, S., Janicka, A., Martinelli, M., Kozlowski, H.,
and Parkinson’s disease. Brain levels of glutathione, glutathione disulfide, and Palumaa, P., 2008c. A structural–dynamical characterization of human
vitamin E. Mol. Chem. Neuropathol. 14, 213–226. Cox17. J. Biol. Chem. 283, 7912–7920.
Adlard, P.A., Cherny, R.A., Finkelstein, D.I., Gautier, E., Robb, E., Cortes, M., Volitakis, Banci, L., Bertini, I., Ciofi-Baffoni, S., Kozyreva, T., Zovo, K., Palumaa, P., 2010. Affinity
I., Liu, X., Smith, J.P., Perez, K., Laughton, K., Li, Q.X., Charman, S.A., Nicolazzo, J.A., gradients drive copper to cellular destinations. Nature 465, 645–648.
Wilkins, S., Deleva, K., Lynch, T., Kok, G., Ritchie, C.W., Tanzi, R.E., Cappai, R., Banting, G.S., Glerum, D.M., 2006. Mutational analysis of the Saccharomyces
Masters, C.L., Barnham, K.J., Bush, A.I., 2008. Rapid restoration of cognition in cerevisiae cytochrome c oxidase assembly protein Cox11p. Eukaryot. Cell 5,
Alzheimer’s transgenic mice with 8-hydroxy quinoline analogs is associated 568–578.
with decreased interstitial Abeta. Neuron 59, 43–55. Barker, A.J., Ullian, E.M., 2010. Astrocytes and synaptic plasticity. Neuroscientist 16,
Aguzzi, A., Baumann, F., Bremer, J., 2008. The prion’s elusive reason for being. Annu. 40–50.
Rev. Neurosci. 31, 439–477. Barnes, N., Tsivkovskii, R., Tsivkovskaia, N., Lutsenko, S., 2005. The copper-trans-
Alcaraz-Zubeldia, M., Boll-Woehrlen, M.C., Montes-Lopez, S., Perez-Severiano, F., porting ATPases, menkes and wilson disease proteins, have distinct roles in
Martinez-Lazcano, J.C., Diaz-Ruiz, A., Rios, C., 2009. Copper sulfate prevents adult and developing cerebellum. J. Biol. Chem. 280, 9640–9645.
tyrosine hydroxylase reduced activity and motor deficits in a Parkinson’s Barros, L.F., Deitmer, J.W., 2010. Glucose and lactate supply to the synapse. Brain
disease model in mice. Rev. Invest. Clin. 61, 405–411. Res. Rev. 63, 149–159.
Alcaraz-Zubeldia, M., Rojas, P., Boll, C., Rios, C., 2001. Neuroprotective effect of acute Barros, M.H., Johnson, A., Tzagoloff, A., 2004. COX23, a homologue of COX17,
and chronic administration of copper (II) sulfate against MPP+ neurotoxicity in is required for cytochrome oxidase assembly. J. Biol. Chem. 279, 31943–
mice. Neurochem. Res. 26, 59–64. 31947.
Aliaga, M.E., Carrasco-Pozo, C., Lopez-Alarcon, C., Olea-Azar, C., Speisky, H., 2011. Barry, A.N., Blackburn, N.J., 2008. A selenocysteine variant of the human copper
Superoxide-dependent reduction of free Fe3+ and release of Fe2+ from ferritin by chaperone for superoxide dismutase A Se-XAS probe of cluster composition at
the physiologically-occurring Cu(I)-glutathione complex. Bioorg. Med. Chem. the domain 3-domain 3 dimer interface. Biochemistry 47, 4916–4928.
19, 534–541. Barry, A.N., Shinde, U., Lutsenko, S., 2010. Structural organization of human Cu-
Aliaga, M.E., Lopez-Alarcon, C., Barriga, G., Olea-Azar, C., Speisky, H., 2010. Redox- transporting ATPases: learning from building blocks. J. Biol. Inorg. Chem. 15,
active complexes formed during the interaction between glutathione and 47–59.
mercury and/or copper ions. J. Inorg. Biochem. 104, 1084–1090. Barthel, A., Ostrakhovitch, E.A., Walter, P.L., Kampkotter, A., Klotz, L.O., 2007.
Aller, S.G., Eng, E.T., De Feo, C.J., Unger, V.M., 2004. Eukaryotic CTR copper uptake Stimulation of phosphoinositide 3-kinase/Akt signaling by copper and zinc
transporters require two faces of the third transmembrane domain for helix ions: mechanisms and consequences. Arch. Biochem. Biophys. 463, 175–182.
packing, oligomerization, and function. J. Biol. Chem. 279, 53435–53441. Bayer, T.A., Multhaup, G., 2005. Involvement of amyloid beta precursor protein
Aller, S.G., Unger, V.M., 2006. Projection structure of the human copper transporter (AbPP) modulated copper homeostasis in Alzheimer’s disease. J. Alzheimers
CTR1 at 6-A resolution reveals a compact trimer with a novel channel-like Dis. 8, 201–206.
architecture. Proc. Natl. Acad. Sci. U.S.A. 103, 3627–3632. Bayer, T.A., Schafer, S., Simons, A., Kemmling, A., Kamer, T., Tepest, R., Eckert, A.,
Anderson, K.E., 2011. Huntington’s disease. Handb. Clin. Neurol. 100, 15–24. Schussel, K., Eikenberg, O., Sturchler-Pierrat, C., Abramowski, D., Staufenbiel, M.,
Andrews, G.K., 2000. Regulation of metallothionein gene expression by oxidative Multhaup, G., 2003. Dietary Cu stabilizes brain superoxide dismutase 1 activity
stress and metal ions. Biochem. Pharmacol. 59, 95–104. and reduces amyloid Abeta production in APP23 transgenic mice. Proc. Natl.
Anezaki, T., Ishiguro, H., Hozumi, I., Inuzuka, T., Hiraiwa, M., Kobayashi, H., Yuguchi, Acad. Sci. U.S.A. 100, 14187–14192.
T., Wanaka, A., Uda, Y., Miyatake, T., Yamada, K., Tohyama, M., Tsuji, S., 1995. Becker, J.S., Salber, D., 2010. New mass spectrometric tools in brain research. TrAC
Expression of growth inhibitory factor (GIF) in normal and injured rat brains. Trends Anal. Chem. 29, 966–979.
Neurochem. Int. 27, 89–94. Becker, J.S., Zoriy, M.V., Pickhardt, C., Palomero-Gallagher, N., Zilles, K., 2005.
Antonyuk, S.V., Strange, R.W., Marklund, S.L., Hasnain, S.S., 2009. The structure of Imaging of copper, zinc, and other elements in thin section of human brain
human extracellular copper–zinc superoxide dismutase at 1.7 A resolution: samples (hippocampus) by laser ablation inductively coupled plasma mass
insights into heparin and collagen binding. J. Mol. Biol. 388, 310–326. spectrometry. Anal. Chem. 77, 3208–3216.
48 I.F. Scheiber et al. / Progress in Neurobiology 116 (2014) 33–57

Beers, J., Glerum, D.M., Tzagoloff, A., 1997. Purification, characterization, and Brown, D.R., Wong, B.S., Hafiz, F., Clive, C., Haswell, S.J., Jones, I.M., 1999. Normal
localization of yeast Cox17p, a mitochondrial copper shuttle. J. Biol. Chem. prion protein has an activity like that of superoxide dismutase. Biochem. J. 344
272, 33191–33196. (Pt 1) 1–5.
Bell, K.F., Hardingham, G.E., 2011. CNS peroxiredoxins and their regulation in health Brown, D.R., Hafiz, F., Glasssmith, L.L., Wong, B.S., Jones, I.M., Clive, C., Haswell, S.J.,
and disease. Antioxid. Redox Signal. 14, 1467–1477. 2000. Consequences of manganese replacement of copper for prion protein
Bellingham, S.A., Ciccotosto, G.D., Needham, B.E., Fodero, L.R., White, A.R., Masters, function and proteinase resistance. EMBO J. 19, 1180–1186.
C.L., Cappai, R., Camakaris, J., 2004a. Gene knockout of amyloid precursor Brown, K., Mastrianni, J.A., 2010. The prion diseases. J. Geriatr. Psychiatry Neurol. 23,
protein and amyloid precursor-like protein-2 increases cellular copper levels 277–298.
in primary mouse cortical neurons and embryonic fibroblasts. J. Neurochem. 91, Brown, L.R., Harris, D.A., 2003. Copper and zinc cause delivery of the prion protein
423–428. from the plasma membrane to a subset of early endosomes and the Golgi. J.
Bellingham, S.A., Lahiri, D.K., Maloney, B., La Fontaine, S., Multhaup, G., Camakaris, J., Neurochem. 87, 353–363.
2004b. Copper depletion down-regulates expression of the Alzheimer’s disease Brown, N.M., Torres, A.S., Doan, P.E., O’Halloran, T.V., 2004. Oxygen and the copper
amyloid-beta precursor protein gene. J. Biol. Chem. 279, 20378–20386. chaperone CCS regulate posttranslational activation of Cu,Zn superoxide dis-
Beltramini, M., Di Pisa, C., Zambenedetti, P., Wittkowski, W., Mocchegiani, E., mutase. Proc. Natl. Acad. Sci. U.S.A. 101, 5518–5523.
Musicco, M., Zatta, P., 2004. Zn and Cu alteration in connection with astrocyte Budimir, A., 2011. Metal ions, Alzheimer’s disease and chelation therapy. Acta
metallothionein I/II overexpression in the mouse brain upon physical stress. Pharm. 61, 1–14.
Glia 47, 30–34. Buhler, R.H., Kagi, J.H., 1974. Human hepatic metallothioneins. FEBS Lett. 39, 229–
Bendor, J.T., Logan, T.P., Edwards, R.H., 2013. The function of a-synuclein. Neuron 234.
79, 1044–1066. Bulcke, F., Thiel, K., Dringen, R., 2014. Uptake and toxicity of copper oxide nano-
Benitez, J.J., Keller, A.M., Huffman, D.L., Yatsunyk, L.A., Rosenzweig, A.C., Chen, P., particles in cultured primary brain astrocytes. Nanotoxicology 8, 775–785.
2011. Relating dynamic protein interactions of metallochaperones with metal Bull, P.C., Thomas, G.R., Rommens, J.M., Forbes, J.R., Cox, D.W., 1993. The Wilson
transfer at the single-molecule level. Faraday Discuss. 148, 71–82. disease gene is a putative copper transporting P-type ATPase similar to the
Berridge, C.W., 2008. Noradrenergic modulation of arousal. Brain Res. Rev. 58, 1–17. Menkes gene. Nat. Genet. 5, 327–337.
Bertinato, J., Cheung, L., Hoque, R., Plouffe, L.J., 2010. Ctr1 transports silver into Burdo, J.R., Menzies, S.L., Simpson, I.A., Garrick, L.M., Garrick, M.D., Dolan, K.G., Haile,
mammalian cells. J. Trace Elem. Med. Biol. 24, 178–184. D.J., Beard, J.L., Connor, J.R., 2001. Distribution of divalent metal transporter 1
Bertinato, J., L’Abbe, M.R., 2003. Copper modulates the degradation of copper and metal transport protein 1 in the normal and Belgrade rat. J. Neurosci. Res.
chaperone for Cu,Zn superoxide dismutase by the 26 S proteosome. J. Biol. 66, 1198–1207.
Chem. 278, 35071–35078. Burkitt, M.J., 2001. A critical overview of the chemistry of copper-dependent low
Bertinato, J., Swist, E., Plouffe, L.J., Brooks, S.P., L’Abbe, M.R., 2008. Ctr2 is partially density lipoprotein oxidation: roles of lipid hydroperoxides, alpha-tocopherol,
localized to the plasma membrane and stimulates copper uptake in COS-7 cells. thiols, and ceruloplasmin. Arch. Biochem. Biophys. 394, 117–135.
Biochem. J. 409, 731–740. Bush, A.I., 2000. Metals and neuroscience. Curr. Opin. Chem. Biol. 4, 184–191.
Bertrand, E., Lewandowska, E., Szpak, G.M., Hoogenraad, T., Blaauwgers, H.G., Camakaris, J., Voskoboinik, I., Mercer, J.F., 1999. Molecular mechanisms of copper
Czlonkowska, A., Dymecki, J., 2001. Neuropathological analysis of pathological homeostasis. Biochem. Biophys. Res. Commun. 261, 225–232.
forms of astroglia in Wilson’s disease. Folia Neuropathol. 39, 73–79. Canello, T., Friedman-Levi, Y., Mizrahi, M., Binyamin, O., Cohen, E., Frid, K., Gabizon,
Bielli, P., Calabrese, L., 2002. Structure to function relationships in ceruloplasmin: a R., 2012. Copper is toxic to PrP-ablated mice and exacerbates disease in a mouse
‘moonlighting’ protein. Cell. Mol. Life Sci. 59, 1413–1427. model of E200K genetic prion disease. Neurobiol. Dis. 45, 1010–1017.
Birkaya, B., Aletta, J.M., 2005. NGF promotes copper accumulation required for Carr, H.S., George, G.N., Winge, D.R., 2002. Yeast Cox11, a protein essential for
optimum neurite outgrowth and protein methylation. J. Neurobiol. 63, 49–61. cytochrome c oxidase assembly, is a Cu(I)-binding protein. J. Biol. Chem. 277,
Boal, A.K., Rosenzweig, A.C., 2009. Structural biology of copper trafficking. Chem. 31237–31242.
Rev. 109, 4760–4779. Carrico, R.J., Deutsch, H.F., 1970. The presence of zinc in human cytocuprein and
Bogumil, R., Faller, P., Pountney, D.L., Vasak, M., 1996. Evidence for Cu(I) clusters and some properties of the apoprotein. J. Biol. Chem. 245, 723–727.
Zn(II) clusters in neuronal growth-inhibitory factor isolated from bovine brain. Carroll, M.C., Girouard, J.B., Ulloa, J.L., Subramaniam, J.R., Wong, P.C., Valentine, J.S.,
Eur. J. Biochem. 238, 698–705. Culotta, V.C., 2004. Mechanisms for activating Cu- and Zn-containing superox-
Bonilla, E., Salazar, E., Villasmil, J.J., Villalobos, R., Gonzalez, M., Davila, J.O., 1984. ide dismutase in the absence of the CCS Cu chaperone. Proc. Natl. Acad. Sci.
Copper distribution in the normal human brain. Neurochem. Res. 9, 1543–1548. U.S.A. 101, 5964–5969.
Borchardt, T., Camakaris, J., Cappai, R., Masters, C.L., Beyreuther, K., Multhaup, G., Caruano-Yzermans, A.L., Bartnikas, T.B., Gitlin, J.D., 2006. Mechanisms of the
1999. Copper inhibits beta-amyloid production and stimulates the non- copper-dependent turnover of the copper chaperone for superoxide dismutase.
amyloidogenic pathway of amyloid-precursor-protein secretion. Biochem. J. J. Biol. Chem. 281, 13581–13587.
344 (Pt 2) 461–467. Casareno, R.L., Waggoner, D., Gitlin, J.D., 1998. The copper chaperone CCS directly
Bordelon, Y.M., 2013. Clinical neurogenetics: Huntington disease. Neurol. Clin. 31, interacts with copper/zinc superoxide dismutase. J. Biol. Chem. 273, 23625–
1085–1094. 23628.
Borisov, V.B., 2002. Defects in mitochondrial respiratory complexes III and IV, and Castellani, R.J., Rolston, R.K., Smith, M.A., 2010. Alzheimer disease. Dis. Mon. 56,
human pathologies. Mol. Aspects Med. 23, 385–412. 484–546.
Borjigin, J., Payne, A.S., Deng, J., Li, X., Wang, M.M., Ovodenko, B., Gitlin, J.D., Snyder, Cater, M.A., McInnes, K.T., Li, Q.X., Volitakis, I., La Fontaine, S., Mercer, J.F., Bush, A.I.,
S.H., 1999. A novel pineal night-specific ATPase encoded by the Wilson disease 2008. Intracellular copper deficiency increases amyloid-beta secretion by di-
gene. J. Neurosci. 19, 1018–1026. verse mechanisms. Biochem. J. 412, 141–152.
Boulanger, Y., Armitage, I.M., Miklossy, K.A., Winge, D.R., 1982. 113Cd NMR study of a Cerpa, W.F., Barria, M.I., Chacon, M.A., Suazo, M., Gonzalez, M., Opazo, C., Bush, A.I.,
metallothionein fragment. Evidence for a two-domain structure. J. Biol. Chem. Inestrosa, N.C., 2004. The N-terminal copper-binding domain of the amyloid
257, 13717–13719. precursor protein protects against Cu2+ neurotoxicity in vivo. FASEB J. 18, 1701–
Boulard, M., Blume, K.G., Beutler, E., 1972. The effect of copper on red cell enzyme 1703.
activities. J. Clin. Invest. 51, 459–461. Chelly, J., Tumer, Z., Tonnesen, T., Petterson, A., Ishikawa-Brush, Y., Tommerup, N.,
Bousquet-Moore, D., Mains, R.E., Eipper, B.A., 2010a. Peptidylgycine alpha-amidat- Horn, N., Monaco, A.P., 1993. Isolation of a candidate gene for Menkes disease
ing monooxygenase and copper: a gene-nutrient interaction critical to nervous that encodes a potential heavy metal binding protein. Nat. Genet. 3, 14–19.
system function. J. Neurosci. Res. 88, 2535–2545. Chen, C.M., 2011. Mitochondrial dysfunction, metabolic deficits, and increased
Bousquet-Moore, D., Prohaska, J.R., Nillni, E.A., Czyzyk, T., Wetsel, W.C., Mains, R.E., oxidative stress in Huntington’s disease. Chang Gung Med. J. 34, 135–152.
Eipper, B.A., 2010b. Interactions of peptide amidation and copper: novel bio- Chen, H.H., Song, I.S., Hossain, A., Choi, M.K., Yamane, Y., Liang, Z.D., Lu, J., Wu, L.Y.,
markers and mechanisms of neural dysfunction. Neurobiol. Dis. 37, 130–140. Siddik, Z.H., Klomp, L.W., Savaraj, N., Kuo, M.T., 2008. Elevated glutathione
Brazier, M.W., Davies, P., Player, E., Marken, F., Viles, J.H., Brown, D.R., 2008. levels confer cellular sensitization to cisplatin toxicity by up-regulation of
Manganese binding to the prion protein. J. Biol. Chem. 283, 12831–12839. copper transporter hCtr1. Mol. Pharmacol. 74, 697–704.
Breen, D.P., Barker, R.A., 2010. Parkinson’s disease and 2009: recent advances. J. Chen, L., Richardson, J.S., Caldwell, J.E., Ang, L.C., 1994. Regional brain activity of free
Neurol. 257, 1224–1228. radical defense enzymes in autopsy samples from patients with Alzheimer’s
Bremner, I., Davies, N.T., 1975. The induction of metallothionein in rat liver by zinc disease and from nondemented controls. Int. J. Neurosci. 75, 83–90.
injection and restriction of food intake. Biochem. J. 149, 733–738. Chen, P., Solomon, E.I., 2004. Oxygen activation by the noncoupled binuclear copper
Britton, R.S., 1996. Metal-induced hepatotoxicity. Semin. Liver Dis. 16, 3–12. site in peptidylglycine alpha-hydroxylating monooxygenase. Reaction mecha-
Brown, D.R., 2004. Role of the prion protein in copper turnover in astrocytes. nism and role of the noncoupled nature of the active site. J. Am. Chem. Soc. 126,
Neurobiol. Dis. 15, 534–543. 4991–5000.
Brown, D.R., 2010. Oligomeric a-synuclein and its role in neuronal death. IUBMB Cherian, M.G., Kang, Y.J., 2006. Metallothionein and liver cell regeneration. Exp. Biol.
Life 62, 334–339. Med. (Maywood) 231, 138–144.
Brown, D.R., Sassoon, J., 2002. Copper-dependent functions for the prion protein. Choi, B.S., Zheng, W., 2009. Copper transport to the brain by the blood–brain barrier
Mol. Biotechnol. 22, 165–178. and blood–CSF barrier. Brain Res. 1248, 14–21.
Brown, D.R., Qin, K., Herms, J.W., Madlung, A., Manson, J., Strome, R., Fraser, P.E., Chung, R.S., Penkowa, M., Dittmann, J., King, C.E., Bartlett, C., Asmussen, J.W.,
Kruck, T., von Bohlen, A., Schulz-Schaeffer, W., Giese, A., Westaway, D., Hidalgo, J., Carrasco, J., Leung, Y.K., Walker, A.K., Fung, S.J., Dunlop, S.A.,
Kretzschmar, H., 1997. The cellular prion protein binds copper in vivo. Nature Fitzgerald, M., Beazley, L.D., Chuah, M.I., Vickers, J.C., West, A.K., 2008. Redefin-
390, 684–687. ing the role of metallothionein within the injured brain: extracellular metal-
Brown, D.R., Schmidt, B., Kretzschmar, H.A., 1998. Effects of copper on survival of lothioneins play an important role in the astrocyte-neuron response to injury. J.
prion protein knockout neurons and glia. J. Neurochem. 70, 1686–1693. Biol. Chem. 283, 15349–15358.
I.F. Scheiber et al. / Progress in Neurobiology 116 (2014) 33–57 49

Cimato, T.R., Ettinger, M.J., Zhou, X., Aletta, J.M., 1997. Nerve growth factor-specific plexus results in long-term rescue of the lethal copper transport defect in a
regulation of protein methylation during neuronal differentiation of PC12 cells. Menkes disease mouse model. Mol. Ther. 19, 2114–2123.
J. Cell Biol. 138, 1089–1103. Doreulee, N., Yanovsky, Y., Haas, H.L., 1997. Suppression of long-term potentiation
Cimini, V., Ruggiero, G., Buonomo, T., Seru, R., Sciorio, S., Zanzi, C., Santangelo, F., in hippocampal slices by copper. Hippocampus 7, 666–669.
Mondola, P., 2002. CuZn-superoxide dismutase in human thymus: immunocy- Double, K.L., 2012. Neuronal vulnerability in Parkinson’s disease. Parkinsonism
tochemical localisation and secretion in thymus-derived epithelial and fibro- Relat. Disord. 18 (Suppl. 1) S52–S54.
blast cell lines. Histochem. Cell Biol. 118, 163–169. Dringen, R., 2000. Metabolism and functions of glutathione in brain. Prog. Neuro-
Ciriolo, M.R., Desideri, A., Paci, M., Rotilio, G., 1990. Reconstitution of Cu,Zn- biol. 62, 649–671.
superoxide dismutase by the Cu(I) glutathione complex. J. Biol. Chem. 265, Dringen, R., 2005. Oxidative and antioxidative potential of brain microglial cells.
11030–11034. Antioxid. Redox Signal. 7, 1223–1233.
Cobb, N.J., Surewicz, W.K., 2009. Prion diseases and their biochemical mechanisms. Dringen, R., Bishop, G.M., Koeppe, M., Dang, T.N., Robinson, S.R., 2007. The pivotal
Biochemistry 48, 2574–2585. role of astrocytes in the metabolism of iron in the brain. Neurochem. Res. 32,
Cobine, P.A., Ojeda, L.D., Rigby, K.M., Winge, D.R., 2004. Yeast contain a non- 1884–1890.
proteinaceous pool of copper in the mitochondrial matrix. J. Biol. Chem. 279, Dringen, R., Hamprecht, B., 1998. Glutathione restoration as indicator for cellular
14447–14455. metabolism of astroglial cells. Dev. Neurosci. 20, 401–407.
Coddou, C., Morales, B., Huidobro-Toro, J.P., 2003. Neuromodulator role of zinc and Dringen, R., Hirrlinger, J., 2010. Antioxidative defense of brain microglial cells. In:
copper during prolonged ATP applications to P2X4 purinoceptors. Eur. J. Phar- Packer, L., Sies, H., Eggersdorf, M., Cadenas, E. (Eds.), Micronutrients and Brain
macol. 472, 49–56. Health. Taylor & Francis, Boca Raton, FL, USA, pp. 393–401.
Colvin, R.A., Holmes, W.R., Fontaine, C.P., Maret, W., 2010. Cytosolic zinc buffering Dringen, R., Pfeiffer, B., Hamprecht, B., 1999. Synthesis of the antioxidant glutathi-
and muffling: their role in intracellular zinc homeostasis. Metallomics 2, 306– one in neurons: supply by astrocytes of CysGly as precursor for neuronal
317. glutathione. J. Neurosci. 19, 562–569.
Crapo, J.D., Oury, T., Rabouille, C., Slot, J.W., Chang, L.Y., 1992. Copper,zinc superox- Dringen, R., Scheiber, I.F., Mercer, J.F., 2013. Copper metabolism of astrocytes. Front.
ide dismutase is primarily a cytosolic protein in human cells. Proc. Natl. Acad. Aging Neurosci. 5, 9.
Sci. U.S.A. 89, 10405–10409. Du, T., Ciccotosto, G.D., Cranston, G.A., Kocak, G., Masters, C.L., Crouch, P.J., Cappai,
Crichton, R.R., Dexter, D.T., Ward, R.J., 2011. Brain iron metabolism and its pertur- R., White, A.R., 2008. Neurotoxicity from glutathione depletion is mediated by
bation in neurological diseases. J. Neural Transm. 118, 301–314. Cu-dependent p53 activation. Free Radic. Biol. Med. 44, 44–55.
Crichton, R.R., Pierre, J.L., 2001. Old iron, young copper: from Mars to Venus. Dudzik, C.G., Wakter, E.D., Abrams, B.S., Jurica, M.S., Millhauser, G.L., 2013. Coordi-
Biometals 14, 99–112. nation of copper to the membrane-bound form of a-synuclein. Biochemistry 52,
Crouch, P.J., Blake, R., Duce, J.A., Ciccotosto, G.D., Li, Q.X., Barnham, K.J., Curtain, C.C., 53–60.
Cherny, R.A., Cappai, R., Dyrks, T., Masters, C.L., Trounce, I.A., 2005. Copper- Duncan, K.E., Ngu, T.T., Chan, J., Salgado, M.T., Merrifield, M.E., Stillman, M.J., 2006.
dependent inhibition of human cytochrome c oxidase by a dimeric conformer of Peptide folding, metal-binding mechanisms, and binding site structures in
amyloid-beta1-42. J. Neurosci. 25, 672–679. metallothioneins. Exp. Biol. Med. (Maywood) 231, 1488–1499.
Crouch, P.J., Hung, L.W., Adlard, P.A., Cortes, M., Lal, V., Filiz, G., Perez, K.A., Nurjono, Eide, D.J., 2006. Zinc transporters and the cellular trafficking of zinc. Biochim.
M., Caragounis, A., Du, T., Laughton, K., Volitakis, I., Bush, A.I., Li, Q.X., Masters, Biophys. Acta 1763, 711–722.
C.L., Cappai, R., Cherny, R.A., Donnelly, P.S., White, A.R., Barnham, K.J., 2009. Eisses, J.F., Chi, Y., Kaplan, J.H., 2005. Stable plasma membrane levels of hCTR1
Increasing Cu bioavailability inhibits Ab oligomers and tau phosphorylation. mediate cellular copper uptake. J. Biol. Chem. 280, 9635–9639.
Proc. Natl. Acad. Sci. U.S.A. 106, 381–386. Eisses, J.F., Kaplan, J.H., 2002. Molecular characterization of hCTR1, the human
Culotta, V.C., Klomp, L.W., Strain, J., Casareno, R.L., Krems, B., Gitlin, J.D., 1997. The copper uptake protein. J. Biol. Chem. 277, 29162–29171.
copper chaperone for superoxide dismutase. J. Biol. Chem. 272, 23469–23472. Eisses, J.F., Kaplan, J.H., 2005. The mechanism of copper uptake mediated by human
Cumings, J.N., 1948. The copper and iron content of brain and liver in the normal and CTR1: a mutational analysis. J. Biol. Chem. 280, 37159–37168.
in hepato-lenticular degeneration. Brain 71, 410–415. El Meskini, R., Cline, L.B., Eipper, B.A., Ronnett, G.V., 2005. The developmentally
Davies, K.M., Hare, D.J., Cottam, V., Chen, N., Hilgers, L., Halliday, G., Mercer, J.F., regulated expression of Menkes protein ATP7A suggests a role in axon exten-
Double, K.L., 2013. Localization of copper and copper transporters in the human sion and synaptogenesis. Dev. Neurosci. 27, 333–348.
brain. Metallomics 5, 43–51. El Meskini, R., Culotta, V.C., Mains, R.E., Eipper, B.A., 2003. Supplying copper to the
Davies, K.M., Bohic, S., Carmona, A., Ortega, R., Cottam, V., Hare, D.J., Finberg, J.P., cuproenzyme peptidylglycine alpha-amidating monooxygenase. J. Biol. Chem.
Reyes, S., Halliday, G.M., Mercer, J.F., Double, K.L., 2014. Copper pathology in 278, 12278–12284.
vulnerable brain regions in Parkinson’ disease. Neurobiol. Aging 35, 858–866. Eskici, G., Axelsen, P.H., 2012. Copper and oxidative stress in the pathogenesis of
Davies, P., Wang, X., Sarell, C.J., Drewett, A., Marken, F., Viles, J.H., Brown, D.R., 2011a. Alzheimer’s disease. Biochemistry 51, 6289–6311.
The synucleins are a family of redox-active copper binding proteins. Biochem- Faa, G., Lisci, M., Caria, M.P., Ambu, R., Sciot, R., Nurchi, V.M., Silvagni, R., Diaz, A.,
istry 50, 37–47. Crisponi, G., 2001. Brain copper, iron, magnesium, zinc, calcium, sulfur and
Davies, P., Moualla, D., Brown, D.R., 2011b. a-Synuclein is a cellular ferrireductase. phosphorus storage in Wilson’s disease. J. Trace Elem. Med. Biol. 15, 155–160.
PLoS ONE 6, e15814. Faller, P., Hureau, C., Berthoumieu, O., 2013. Role of metal ions in the self-assembly
De Feo, C.J., Aller, S.G., Unger, V.M., 2007. A structural perspective on copper uptake of the Alzheimer’s amyloid-b peptide. Inorg. Chem. 52, 12193–12206.
in eukaryotes. Biometals 20, 705–716. Ferguson-Miller, S., Babcock, G.T., 1996. Heme/copper terminal oxidases. Chem.
Deibel, M.A., Ehmann, W.D., Markesbery, W.R., 1996. Copper, iron, and zinc imbal- Rev. 96, 2889–2908.
ances in severely degenerated brain regions in Alzheimer’s disease: possible Ferreira, A.M., Ciriolo, M.R., Marcocci, L., Rotilio, G., 1993. Copper(I) transfer into
relation to oxidative stress. J. Neurol. Sci. 143, 137–142. metallothionein mediated by glutathione. Biochem. J. 292 (Pt 3) 673–676.
Deitmer, J.W., Rose, C.R., 2010. Ion changes and signalling in perisynaptic glia. Brain Ferrer, I., Martinez, A., Blanco, R., Dalfo, E., Carmona, M., 2011. Neuropathology of
Res. Rev. 63, 113–129. sporadic Parkinson disease before the appearance of parkinsonism: preclinical
Demeuse, P., Kerkhofs, A., Struys-Ponsar, C., Knoops, B., Remacle, C., van den Bosch Parkinson disease. J. Neural Transm. 118, 821–839.
de Aguilar, P., 2002. Compartmentalized coculture of rat brain endothelial cells Ferruzza, S., Sambuy, Y., Ciriolo, M.R., De Martino, A., Santaroni, P., Rotilio, G.,
and astrocytes: a syngenic model to study the blood–brain barrier. J. Neurosci. Scarino, M.L., 2000. Copper uptake and intracellular distribution in the human
Methods 121, 21–31. intestinal Caco-2 cell line. Biometals 13, 179–185.
Dexter, D.T., Carayon, A., Javoy-Agid, F., Agid, Y., Wells, F.R., Daniel, S.E., Lees, A.J., Finney, L., Mandava, S., Ursos, L., Zhang, W., Rodi, D., Vogt, S., Legnini, D., Maser, J.,
Jenner, P., Marsden, C.D., 1991. Alterations in the levels of iron, ferritin and other Ikpatt, F., Olopade, O.I., Glesne, D., 2007. X-ray fluorescence microscopy reveals
trace metals in Parkinson’s disease and other neurodegenerative diseases large-scale relocalization and extracellular translocation of cellular copper
affecting the basal ganglia. Brain 114 (Pt 4) 1953–1975. during angiogenesis. Proc. Natl. Acad. Sci. U.S.A. 104, 2247–2252.
Dexter, D.T., Wells, F.R., Lees, A.J., Agid, F., Agid, Y., Jenner, P., Marsden, C.D., 1989. Finney, L., Vogt, S., Fukai, T., Glesne, D., 2009. Copper and angiogenesis: unravelling
Increased nigral iron content and alterations in other metal ions occurring in a relationship key to cancer progression. Clin. Exp. Pharmacol. Physiol. 36, 88–
brain in Parkinson’s disease. J. Neurochem. 52, 1830–1836. 94.
Diaz, F., 2010. Cytochrome c oxidase deficiency: patients and animal models. Forbes, J.R., Cox, D.W., 2000. Copper-dependent trafficking of Wilson disease
Biochim. Biophys. Acta 1802, 100–110. mutant ATP7B proteins. Hum. Mol. Genet. 9, 1927–1935.
Dienel, G.A., 2012. Brain lactate metabolism: the discoveries and the controversies. Forte, G., Bocca, B., Senofonte, O., Petrucci, F., Brusa, L., Stanzione, P., Zannino, S.,
J. Cereb. Blood Flow Metab. 32, 1107–1138. Violante, N., Alimonti, A., Sancesario, G., 2004. Trace and major elements in
Dineley, K.E., Scanlon, J.M., Kress, G.J., Stout, A.K., Reynolds, I.J., 2000. Astrocytes are whole blood, serum, cerebrospinal fluid and urine of patients with Parkinson’s
more resistant than neurons to the cytotoxic effects of increased [Zn2+](i). disease. J. Neural Transm. 111, 1031–1040.
Neurobiol. Dis. 7, 310–320. Fox, J.H., Kama, J.A., Lieberman, G., Chopra, R., Dorsey, K., Chopra, V., Volitakis, I.,
Dobrowolska, J., Dehnhardt, M., Matusch, A., Zoriy, M., Palomero-Gallagher, N., Cherny, R.A., Bush, A.I., Hersch, S., 2007. Mechanisms of copper ion mediated
Koscielniak, P., Zilles, K., Becker, J.S., 2008. Quantitative imaging of zinc, copper Huntington’s disease progression. PLoS ONE 2, e334.
and lead in three distinct regions of the human brain by laser ablation induc- Frederickson, C.J., Bush, A.I., 2001. Synaptically released zinc: physiological func-
tively coupled plasma mass spectrometry. Talanta 74, 717–723. tions and pathological effects. Biometals 14, 353–366.
Donsante, A., Johnson, P., Jansen, L.A., Kaler, S.G., 2010. Somatic mosaicism in Freedman, J.H., Ciriolo, M.R., Peisach, J., 1989. The role of glutathione in copper
Menkes disease suggests choroid plexus-mediated copper transport to the metabolism and toxicity. J. Biol. Chem. 264, 5598–5605.
developing brain. Am. J. Med. Genet. A 152A, 2529–2534. Freedman, J.H., Weiner, R.J., Peisach, J., 1986. Resistance to copper toxicity of
Donsante, A., Yi, L., Zerfas, P.M., Brinster, L.R., Sullivan, P., Goldstein, D.S., Prohaska, cultured hepatoma cells Characterization of resistant cell lines. J. Biol. Chem.
J., Centeno, J.A., Rushing, E., Kaler, S.G., 2011. ATP7A gene addition to the choroid 261, 11840–11848.
50 I.F. Scheiber et al. / Progress in Neurobiology 116 (2014) 33–57

Fukuyama, H., Ogawa, M., Yamauchi, H., Yamaguchi, S., Kimura, J., Yonekura, Y., Hamza, I., Gitlin, J.D., 2002. Copper chaperones for cytochrome c oxidase and human
Konishi, J., 1994. Altered cerebral energy metabolism in Alzheimer’s disease: a disease. J. Bioenerg. Biomembr. 34, 381–388.
PET study. J. Nucl. Med. 35, 1–6. Hamza, I., Klomp, L.W., Gaedigk, R., White, R.A., Gitlin, J.D., 2000. Structure,
Furukawa, Y., Torres, A.S., O’Halloran, T.V., 2004. Oxygen-induced maturation of expression, and chromosomal localization of the mouse Atox1 gene. Genomics
SOD1: a key role for disulfide formation by the copper chaperone CCS. EMBO J. 63, 294–297.
23, 2872–2881. Hamza, I., Schaefer, M., Klomp, L.W., Gitlin, J.D., 1999. Interaction of the copper
Gaggelli, E., Kozlowski, H., Valensin, D., Valensin, G., 2006. Copper homeostasis and chaperone HAH1 with the Wilson disease protein is essential for copper
neurodegenerative disorders (Alzheimer’s, prion, and Parkinson’s diseases and homeostasis. Proc. Natl. Acad. Sci. U.S.A. 96, 13363–13368.
amyotrophic lateral sclerosis). Chem. Rev. 106, 1995–2044. Hanna, P.M., Mason, R.P., 1992. Direct evidence for inhibition of free radical
Gaier, E.D., Eipper, B.A., Mains, R.E., 2013. Copper signaling in the mammalian formation from Cu(I) and hydrogen peroxide by glutathione and other potential
nervous system: synaptic effects. J. Neurosci. Res. 91, 2–19. ligands using the EPR spin-trapping technique. Arch. Biochem. Biophys. 295,
Gaither, L.A., Eide, D.J., 2000. Functional expression of the human hZIP2 zinc 205–213.
transporter. J. Biol. Chem. 275, 5560–5564. Hansel, D.E., May, V., Eipper, B.A., Ronnett, G.V., 2001. Pituitary adenylyl cyclase-
Gaither, L.A., Eide, D.J., 2001. The human ZIP1 transporter mediates zinc uptake in activating peptides and alpha-amidation in olfactory neurogenesis and neuro-
human K562 erythroleukemia cells. J. Biol. Chem. 276, 22258–22264. nal survival in vitro. J. Neurosci. 21, 4625–4636.
Garrett, S.H., Sens, M.A., Shukla, D., Nestor, S., Somji, S., Todd, J.H., Sens, D.A., 1999a. Haq, F., Mahoney, M., Koropatnick, J., 2003. Signaling events for metallothionein
Metallothionein isoform 3 expression in the human prostate and cancer- induction. Mutat. Res. 533, 211–226.
derived cell lines. Prostate 41, 196–202. Hardman, B., Michalczyk, A., Greenough, M., Camakaris, J., Mercer, J., Ackland, L.,
Garrett, S.H., Sens, M.A., Todd, J.H., Somji, S., Sens, D.A., 1999b. Expression of MT-3 2007. Distinct functional roles for the Menkes and Wilson copper translocating
protein in the human kidney. Toxicol. Lett. 105, 207–214. P-type ATPases in human placental cells. Cell. Physiol. Biochem. 20, 1073–1084.
Garrick, M.D., Dolan, K.G., Horbinski, C., Ghio, A.J., Higgins, D., Porubcin, M., Moore, Harris, Z.L., Takahashi, Y., Miyajima, H., Serizawa, M., MacGillivray, R.T., Gitlin, J.D.,
E.G., Hainsworth, L.N., Umbreit, J.N., Conrad, M.E., Feng, L., Lis, A., Roth, J.A., 1995. Aceruloplasminemia: molecular characterization of this disorder of iron
Singleton, S., Garrick, L.M., 2003. DMT1: a mammalian transporter for multiple metabolism. Proc. Natl. Acad. Sci. U.S.A. 92, 2539–2543.
metals. Biometals 16, 41–54. Hartter, D.E., Barnea, A., 1988. Evidence for release of copper in the brain: depolari-
Gauthier, L.R., Charrin, B.C., Borrell-Pages, M., Dompierre, J.P., Rangone, H., Corde- zation-induced release of newly taken-up 67copper. Synapse 2, 412–415.
lieres, F.P., De Mey, J., MacDonald, M.E., Lessmann, V., Humbert, S., Saudou, F., Hatefi, Y., 1985. The mitochondrial electron transport and oxidative phosphoryla-
2004. Huntingtin controls neurotrophic support and survival of neurons by tion system. Annu. Rev. Biochem. 54, 1015–1069.
enhancing BDNF vesicular transport along microtubules. Cell 118, 127–138. Haywood, S., Vaillant, C., in press. Overexpression of copper transporter CTR1 in the
Geffen, L.B., Livett, B.G., Rush, R.A., 1969. Immunohistochemical localizatio of brain barrier of North Ronaldsay sheep: implications for the study of neuro-
protein components of catecholamine storage vesicles. J. Physiol. 204, 593–605. degnerative diseases. J. Comp. Pathol. (in press).
Giese, A., Buchholz, M., Herms, J., Kretzschmar, H.A., 2005. Mouse brain synapto- Haywood, S., Paris, J., Ryvar, R., Botteron, C., 2008. Brain copper elevation and
somes accumulate copper-67 efficiently by two distinct processes independent neurological changes in North Ronaldsay sheep: a model for neurodegenerative
of cellular prion protein. J. Mol. Neurosci. 27, 347–354. disease? J. Comp. Pathol. 139, 252–255.
Girijashanker, K., He, L., Soleimani, M., Reed, J.M., Li, H., Liu, Z., Wang, B., Dalton, T.P., Healy, J., Tipton, K., 2007. Ceruloplasmin and what it might do. J. Neural Transm.
Nebert, D.W., 2008. Slc39a14 gene encodes ZIP14, a metal/bicarbonate sym- 114, 777–781.
porter: similarities to the ZIP8 transporter. Mol. Pharmacol. 73, 1413–1423. Heaton, D., Nittis, T., Srinivasan, C., Winge, D.R., 2000. Mutational analysis of the
Glerum, D.M., Shtanko, A., Tzagoloff, A., 1996. Characterization of COX17, a yeast mitochondrial copper metallochaperone Cox17. J. Biol. Chem. 275, 37582–
gene involved in copper metabolism and assembly of cytochrome oxidase. J. 37587.
Biol. Chem. 271, 14504–14509. Hellman, N.E., Gitlin, J.D., 2002. Ceruloplasmin metabolism and function. Annu. Rev.
Goddard, A.W., Ball, S.G., Martinez, J., Robinson, M.J., Yang, C.R., Russell, J.M., Nutr. 22, 439–458.
Shekhar, A., 2010. Current perspectives of the roles of the central norepineph- Hesse, L., Beher, D., Masters, C.L., Multhaup, G., 1994. The bA4 amyloid precursor
rine system in anxiety and depression. Depress. Anxiety 27, 339–350. protein binding to copper. FEBS Lett. 349, 109–116.
Goldberg, W.J., Allen, N., 1981. Determination of Cu, Mn, Fe, and Ca in six regions of Hidalgo, J., Aschner, M., Zatta, P., Vasak, M., 2001. Roles of the metallothionein
normal human brain, by atomic absorption spectroscopy. Clin. Chem. 27, 562– family of proteins in the central nervous system. Brain Res. Bull. 55, 133–145.
564. Hidalgo, J., Garcia, A., Oliva, A.M., Giralt, M., Gasull, T., Gonzalez, B., Milnerowicz, H.,
Goldschmith, A., Infante, C., Leiva, J., Motles, E., Palestini, M., 2005. Interference of Wood, A., Bremner, I., 1994. Effect of zinc, copper and glucocorticoids on
chronically ingested copper in long-term potentiation (LTP) of rat hippocam- metallothionein levels of cultured neurons and astrocytes from rat brain. Chem.
pus. Brain Res. 1056, 176–182. Biol. Interact. 93, 197–219.
Goldstein, D.S., Eisenhofer, G., Kopin, I.J., 2003. Sources and significance of plasma Hirrlinger, J., Dringen, R., 2010. The cytosolic redox state of astrocytes: mainte-
levels of catechols and their metabolites in humans. J. Pharmacol. Exp. Ther. nance, regulation and functional implications for metabolite trafficking. Brain
305, 800–811. Res. Rev. 63, 177–188.
Goldstein, D.S., Holmes, C.S., Kaler, S.G., 2009. Relative efficiencies of plasma Hiser, L., Di Valentin, M., Hamer, A.G., Hosler, J.P., 2000. Cox11p is required for stable
catechol levels and ratios for neonatal diagnosis of menkes disease. Neurochem. formation of the Cu(B) and magnesium centers of cytochrome c oxidase. J. Biol.
Res. 34, 1464–1468. Chem. 275, 619–623.
Gonzalez-Cuyar, L.F., Perry, G., Miyajima, H., Atwood, C.S., Riveros-Angel, M., Lyons, Hoepken, H.H., Korten, T., Robinson, S.R., Dringen, R., 2004. Iron accumulation, iron-
P.F., Siedlak, S.L., Smith, M.A., Castellani, R.J., 2008. Redox active iron accumu- mediated toxicity and altered levels of ferritin and transferrin receptor
lation in aceruloplasminemia. Neuropathology 28, 466–471. in cultured astrocytes during incubation with ferric ammonium citrate. J.
Greenough, M.A., Camakaris, J., Bush, A.I., 2013. Metal dyshomeostasis and oxida- Neurochem. 88, 1194–1202.
tive stress in Alzheimer’s disease. Neurochem. Int. 62, 540–555. Holman, R.C., Belay, E.D., Christensen, K.Y., Maddox, R.A., Minino, A.M., Folkema,
Gunshin, H., Mackenzie, B., Berger, U.V., Gunshin, Y., Romero, M.F., Boron, W.F., A.M., Haberling, D.L., Hammett, T.A., Kochanek, K.D., Sejvar, J.J., Schonberger,
Nussberger, S., Gollan, J.L., Hediger, M.A., 1997. Cloning and characterization of L.B., 2010. Human prion diseases in the United States. PLoS ONE 5, e8521.
a mammalian proton-coupled metal-ion transporter. Nature 388, 482–488. Hopt, A., Korte, S., Fink, H., Panne, U., Niessner, R., Jahn, R., Kretzschmar, H., Herms, J.,
Gunther, M.R., Hanna, P.M., Mason, R.P., Cohen, M.S., 1995. Hydroxyl radical 2003. Methods for studying synaptosomal copper release. J. Neurosci. Methods
formation from cuprous ion and hydrogen peroxide: a spin-trapping study. 128, 159–172.
Arch. Biochem. Biophys. 316, 515–522. Horng, Y.C., Cobine, P.A., Maxfield, A.B., Carr, H.S., Winge, D.R., 2004. Specific copper
Gunther, V., Lindert, U., Schaffner, W., 2012. The taste of heavy metals: gene transfer from the Cox17 metallochaperone to both Sco1 and Cox11 in the
regulation by MTF-1. Biochim. Biophys. Acta 1823, 1416–1425. assembly of yeast cytochrome C oxidase. J. Biol. Chem. 279, 35334–35340.
Gupta, A., Lutsenko, S., 2009. Human copper transporters: mechanism, role in Horng, Y.C., Leary, S.C., Cobine, P.A., Young, F.B., George, G.N., Shoubridge, E.A.,
human diseases and therapeutic potential. Future Med. Chem. 1, 1125–1142. Winge, D.R., 2005. Human Sco1 and Sco2 function as copper-binding proteins. J.
Gybina, A.A., Prohaska, J.R., 2006. Variable response of selected cuproproteins in rat Biol. Chem. 280, 34113–34122.
choroid plexus and cerebellum following perinatal copper deficiency. Genes Horning, M.S., Trombley, P.Q., 2001. Zinc and copper influence excitability of rat
Nutr. 1, 51–59. olfactory bulb neurons by multiple mechanisms. J. Neurophysiol. 86, 1652–
Gybina, A.A., Prohaska, J.R., 2008a. Copper deficiency results in AMP-activated 1660.
protein kinase activation and acetylCoA carboxylase phosphorylation in rat Howells, C., Saar, K., Eaton, E., Ray, S., Palumaa, P., Shabala, L., Adlard, P.A., Bennett,
cerebellum. Brain Res. 1204, 69–76. W., West, A.K., Guillemin, G.J., Chung, R.S., 2012. Redox-active Cu(II)-Abeta
Gybina, A.A., Prohaska, J.R., 2008b. Fructose-2,6-bisphosphate is lower in copper causes substantial changes in axonal integrity in cultured cortical neurons in an
deficient rat cerebellum despite higher content of phosphorylated AMP- oxidative-stress dependent manner. Exp. Neurol. 237, 499–506.
activated protein kinase. Exp. Biol. Med. (Maywood) 233, 1262–1270. Hozumi, I., Suzuki, J.S., Kanazawa, H., Hara, A., Saio, M., Inuzuka, T., Miyairi, S.,
Haigh, C.L., Brown, D.R., 2006. Prion protein reduces both oxidative and non- Naganuma, A., Tohyama, C., 2008. Metallothionein-3 is expressed in the brain
oxidative copper toxicity. J. Neurochem. 98, 677–689. and various peripheral organs of the rat. Neurosci. Lett. 438, 54–58.
Halliday, G.M., Stevens, C.H., 2011. Glia: initiators and progressors of pathology in Hu, W., Rosenberg, R.N., Stuve, O., 2007. Prion proteins: a biological role beyond
Parkinson’s disease. Mov. Disord. 26, 6–17. prion diseases. Acta Neurol. Scand. 116, 75–82.
Halliwell, B., 2006. Oxidative stress and neurodegeneration: where are we now? J. Huang, C.H., Kuo, W.Y., Weiss, C., Jinn, T.L., 2011. Copper chaperone-dependent and
Neurochem. 97, 1634–1658. -independent activation of three copper–zinc superoxide dismutase homologs
Halliwell, B., Gutteridge, J.M., 2007. Free Radicals in Biology and Medicine. Oxford localized in different cellular compartments in arabidopsis. Plant Physiol. 158,
University Press Inc., New York. 737–746.
I.F. Scheiber et al. / Progress in Neurobiology 116 (2014) 33–57 51

Huang, X., Atwood, C.S., Hartshorn, M.A., Multhaup, G., Goldstein, L.E., Scarpa, R.C., Kaneko, K., Yoshida, K., Arima, K., Ohara, S., Miyajima, H., Kato, T., Ohta, M., Ikeda,
Cuajungco, M.P., Gray, D.N., Lim, J., Moir, R.D., Tanzi, R.E., Bush, A.I., 1999. The A S.I., 2002b. Astrocytic deformity and globular structures are characteristic of the
beta peptide of Alzheimer’s disease directly produces hydrogen peroxide brains of patients with aceruloplasminemia. J. Neuropathol. Exp. Neurol. 61,
through metal ion reduction. Biochemistry 38, 7609–7616. 1069–1077.
Huidobro-Toro, J.P., Lorca, R.A., Coddou, C., 2008. Trace metals in the brain: Kardos, J., Kovacs, I., Hajos, F., Kalman, M., Simonyi, M., 1989. Nerve endings from rat
allosteric modulators of ligand-gated receptor channels, the case of ATP-gated brain tissue release copper upon depolarization A possible role in regulating
P2X receptors. Eur. Biophys. J. 37, 301–314. neuronal excitability. Neurosci. Lett. 103, 139–144.
Hung, I.H., Casareno, R.L., Labesse, G., Mathews, F.S., Gitlin, J.D., 1998. HAH1 is a Karran, E., Mercken, M., De Strooper, B., 2011. The amyloid cascade hypothesis for
copper-binding protein with distinct amino acid residues mediating copper Alzheimer’s disease: an appraisal for the development of therapeutics. Nat. Rev.
homeostasis and antioxidant defense. J. Biol. Chem. 273, 1749–1754. Drug Discov. 10, 698–712.
Hung, I.H., Suzuki, M., Yamaguchi, Y., Yuan, D.S., Klausner, R.D., Gitlin, J.D., 1997. Kawai, K., Liu, S.X., Tyurin, V.A., Tyurina, Y.Y., Borisenko, G.G., Jiang, J.F., St Croix,
Biochemical characterization of the Wilson disease protein and functional C.M., Fabisiak, J.P., Pitt, B.R., Kagan, V.E., 2000. Antioxidant and antiapoptotic
expression in the yeast Saccharomyces cerevisiae. J. Biol. Chem. 272, 21461– function of metallothioneins in HL-60 cells challenged with copper nitrilotria-
21466. cetate. Chem. Res. Toxicol. 13, 1275–1286.
Hunt, C.T., Boulanger, Y., Fesik, S.W., Armitage, I.M., 1984. NMR analysis of the Kelly, E.J., Palmiter, R.D., 1996. A murine model of Menkes disease reveals a
structure and metal sequestering properties of metallothioneins. Environ. physiological function of metallothionein. Nat. Genet. 13, 219–222.
Health Perspect. 54, 135–145. Kenche, V.B., Barnham, K.J., 2011. Alzheimer’s disease & metals: therapeutic oppor-
Hunt, D.M., 1974. Primary defect in copper transport underlies mottled mutants in tunities. Br. J. Pharmacol. 163, 211–219.
the mouse. Nature 249, 852–854. Kessler, H., Bayer, T.A., Bach, D., Schneider-Axmann, T., Supprian, T., Herrmann, W.,
Hussain, F., Rodriguez-Granillo, A., Wittung-Stafshede, P., 2009. Lysine-60 in copper Haber, M., Multhaup, G., Falkai, P., Pajonk, F.G., 2008a. Intake of copper has no
chaperone Atox1 plays an essential role in adduct formation with a target effect on cognition in patients with mild Alzheimer’s disease: a pilot phase 2
Wilson disease domain. J. Am. Chem. Soc. 131, 16371–16373. clinical trial. J. Neural Transm. 115, 1181–1187.
Huster, D., 2010. Wilson disease. Best Pract. Res. Clin. Gastroenterol. 24, 531– Kessler, H., Pajonk, F.G., Bach, D., Schneider-Axmann, T., Falkai, P., Herrmann, W.,
539. Multhaup, G., Wiltfang, J., Schafer, S., Wirths, O., Bayer, T.A., 2008b. Effect of
Illing, A.C., Shawki, A., Cunningham, C.L., Mackenzie, B., 2012. Substrate profile and copper intake on CSF parameters in patients with mild Alzheimer’s disease: a
metal-ion selectivity of human divalent metal-ion transporter-1. J. Biol. Chem. pilot phase 2 clinical trial. J. Neural Transm. 115, 1651–1659.
287, 30485–30496. Kim, B.E., Nevitt, T., Thiele, D.J., 2008. Mechanisms for copper acquisition, distribu-
Ip, V., Liu, J.J., Mercer, J.F., McKeage, M.J., 2010. Differential expression of ATP7A, tion and regulation. Nat. Chem. Biol. 4, 176–185.
ATP7B and CTR1 in adult rat dorsal root ganglion tissue. Mol. Pain. 6, 53. Kim, B.E., Smith, K., Petris, M.J., 2003. A copper treatable Menkes disease mutation
Itoh, S., Kim, H.W., Nakagawa, O., Ozumi, K., Lessner, S.M., Aoki, H., Akram, K., associated with defective trafficking of a functional Menkes copper ATPase. J.
McKinney, R.D., Ushio-Fukai, M., Fukai, T., 2008. Novel role of antioxidant-1 Med. Genet. 40, 290–295.
(Atox1) as a copper-dependent transcription factor involved in cell prolifera- Kim, C.H., Zabetian, C.P., Cubells, J.F., Cho, S., Biaggioni, I., Cohen, B.M., Robertson, D.,
tion. J. Biol. Chem. 283, 9157–9167. Kim, K.S., 2002. Mutations in the dopamine beta-hydroxylase gene are associ-
Itoh, S., Ozumi, K., Kim, H.W., Nakagawa, O., McKinney, R.D., Folz, R.J., Zelko, I.N., ated with human norepinephrine deficiency. Am. J. Med. Genet. 108, 140–147.
Ushio-Fukai, M., Fukai, T., 2009. Novel mechanism for regulation of extracellular Kim, J.H., Cho, H., Ryu, S.E., Choi, M.U., 2000. Effects of metal ions on the activity of
SOD transcription and activity by copper: role of antioxidant-1. Free Radic. Biol. protein tyrosine phosphatase VHR: highly potent and reversible oxidative
Med. 46, 95–104. inactivation by Cu2+ ion. Arch. Biochem. Biophys. 382, 72–80.
Jackson, B., Harper, S., Smith, L., Flinn, J., 2006. Elemental mapping and quantitative Kiningham, K., Bi, X., Kasarskis, E.J., 1995. Neuronal localization of metallothioneins
analysis of Cu, Zn, and Fe in rat brain sections by laser ablation ICP-MS. Anal. in rat and human spinal cord. Neurochem. Int. 27, 105–109.
Bioanal. Chem. 384, 951–957. Kissling, M.M., Kagi, H.R., 1977. Primary structure of human hepatic metallothio-
Jensen, L.T., Culotta, V.C., 2005. Activation of CuZn superoxide dismutases from nein. FEBS Lett. 82, 247–250.
Caenorhabditis elegans does not require the copper chaperone CCS. J. Biol. Klein, D., Bartsch, R., Summer, K.H., 1990. Quantitation of Cu-containing metal-
Chem. 280, 41373–41379. lothionein by a Cd-saturation method. Anal. Biochem. 189, 35–39.
Jiang, J.F., St Croix, C.M., Sussman, N., Zhao, Q., Pitt, B.R., Kagan, V.E., 2002. Klinman, J.P., 1996. Mechanisms whereby mononuclear copper proteins functio-
Contribution of glutathione and metallothioneins to protection against copper nalize organic substrates. Chem. Rev. 96, 2541–2562.
toxicity and redox cycling: quantitative analysis using MT+/+ and MT / Klinman, J.P., 2006. The copper-enzyme family of dopamine beta-monooxygenase
mouse lung fibroblast cells. Chem. Res. Toxicol. 15, 1080–1087. and peptidylglycine alpha-hydroxylating monooxygenase: resolving the chem-
Jimenez, I., Speisky, H., 2000. Effects of copper ions on the free radical-scavenging ical pathway for substrate hydroxylation. J. Biol. Chem. 281, 3013–3016.
properties of reduced gluthathione: implications of a complex formation. J. Klomp, A.E., Juijn, J.A., van der Gun, L.T., van den Berg, I.E., Berger, R., Klomp, L.W.,
Trace Elem. Med. Biol. 14, 161–167. 2003. The N-terminus of the human copper transporter 1 (hCTR1) is localized
Johnson, C.J., Gilbert, P.U., Abrecht, M., Baldwin, K.L., Russell, R.E., Pedersen, J.A., extracellularly, and interacts with itself. Biochem. J. 370, 881–889.
Aiken, J.M., McKenzie, D., 2013. Low copper and high manganese levels in prion Klomp, L.W., Lin, S.J., Yuan, D.S., Klausner, R.D., Culotta, V.C., Gitlin, J.D., 1997.
protein plaques. Viruses 5, 654–662. Identification and functional expression of HAH1, a novel human gene involved
Jomova, K., Vondrakova, D., Lawson, M., Valko, M., 2010. Metals, oxidative stress and in copper homeostasis. J. Biol. Chem. 272, 9221–9226.
neurodegenerative disorders. Mol. Cell. Biochem. 345, 91–104. Kodama, H., 1993. Recent developments in Menkes disease. J. Inherit. Metab. Dis.
Juarez, J.C., Betancourt Jr., O., Pirie-Shepherd, S.R., Guan, X., Price, M.L., Shaw, D.E., 16, 791–799.
Mazar, A.P., Donate, F., 2006. Copper binding by tetrathiomolybdate attenuates Kodama, H., Fujisawa, C., Bhadhprasit, W., 2011. Pathology, clinical features and
angiogenesis and tumor cell proliferation through the inhibition of superoxide treatments of congenital copper metabolic disorders–focus on neurologic
dismutase 1. Clin. Cancer Res. 12, 4974–4982. aspects. Brain Dev. 33, 243–251.
Juarez, J.C., Manuia, M., Burnett, M.E., Betancourt, O., Boivin, B., Shaw, D.E., Tonks, Kodama, H., Meguro, Y., Abe, T., Rayner, M.H., Suzuki, K.T., Kobayashi, S., Nishimura,
N.K., Mazar, A.P., Donate, F., 2008. Superoxide dismutase 1 (SOD1) is essential M., 1991. Genetic expression of Menkes disease in cultured astrocytes of the
for H2O2-mediated oxidation and inactivation of phosphatases in growth factor macular mouse. J. Inherit. Metab. Dis. 14, 896–901.
signaling. Proc. Natl. Acad. Sci. U.S.A. 105, 7147–7152. Kojima, Y., Berger, C., Vallee, B.L., Kagi, J.H., 1976. Amino-acid sequence of equine
Kachur, A.V., Koch, C.J., Biaglow, J.E., 1998. Mechanism of copper-catalyzed oxida- renal metallothionein-1B. Proc. Natl. Acad. Sci. U.S.A. 73, 3413–3417.
tion of glutathione. Free Radic. Res. 28, 259–269. Kong, G.K., Miles, L.A., Crespi, G.A., Morton, C.J., Ng, H.L., Barnham, K.J., McKinstry,
Kaden, D., Bush, A.I., Danzeisen, R., Bayer, T.A., Multhaup, G., 2011. Disturbed copper W.J., Cappai, R., Parker, M.W., 2008. Copper binding to the Alzheimer’s disease
bioavailability in Alzheimer’s disease. Int. J. Alzheimers Dis. 2011, 345614. amyloid precursor protein. Eur. Biophys. J. 37, 269–279.
Kagi, J.H., Himmelhoch, S.R., Whanger, P.D., Bethune, J.L., Vallee, B.L., 1974. Equine Kono, S., 2013. Aceruloplasminemia; an update. Int. Rev. Neurobiol. 110, 125–151.
hepatic and renal metallothioneins Purification, molecular weight, amino acid Kono, S., Miyajima, H., 2006. Molecular and pathological basis of aceruloplasmi-
composition, and metal content. J. Biol. Chem. 249, 3537–3542. nemia. Biol. Res. 39, 15–23.
Kagi, J.H., Valee, B.L., 1960. Metallothionein: a cadmium- and zinc-containing Kosman, D.J., 2010a. Multicopper oxidases: a workshop on copper coordination
protein from equine renal cortex. J. Biol. Chem. 235, 3460–3465. chemistry, electron transfer, and metallophysiology. J. Biol. Inorg. Chem. 15,
Kaim, W., Rall, J., 1996. Copper – a modern bioelement. Angew. Chem. Int. Ed. 35, 15–28.
43–60. Kosman, D.J., 2010b. Redox cycling in iron uptake, efflux, and trafficking. J. Biol.
Kaler, S.G., 1998. Metabolic and molecular bases of Menkes disease and occipital Chem. 285, 26729–26735.
horn syndrome. Pediatr. Dev. Pathol. 1, 85–98. Kosonen, T., Uriu-Hare, J.Y., Clegg, M.S., Keen, C.L., Rucker, R.B., 1997. Incorporation
Kaler, S.G., 2011. ATP7A-related copper transport diseases-emerging concepts and of copper into lysyl oxidase. Biochem. J. 327 (Pt 1) 283–289.
future trends. Nat. Rev. Neurol. 7, 15–29. Kralovicova, S., Fontaine, S.N., Alderton, A., Alderman, J., Ragnarsdottir, K.V., Collins,
Kaler, S.G., Holmes, C.S., Goldstein, D.S., Tang, J., Godwin, S.C., Donsante, A., Liew, C.J., S.J., Brown, D.R., 2009. The effects of prion protein expression on metal metab-
Sato, S., Patronas, N., 2008. Neonatal diagnosis and treatment of Menkes olism. Mol. Cell. Neurosci. 41, 135–147.
disease. N. Engl. J. Med. 358, 605–614. Kumode, M., Yamano, T., Shimada, M., 1993. Neuropathological study on
Kambe, T., Yamaguchi-Iwai, Y., Sasaki, R., Nagao, M., 2004. Overview of mammalian cerebellum of macular mutant mouse heterozygote. Acta Neuropathol. 86,
zinc transporters. Cell. Mol. Life Sci. 61, 49–68. 411–417.
Kaneko, K., Nakamura, A., Yoshida, K., Kametani, F., Higuchi, K., Ikeda, S., 2002a. Glial Kunz, W.S., Kuznetsov, A.V., Clark, J.F., Tracey, I., Elger, C.E., 1999. Metabolic
fibrillary acidic protein is greatly modified by oxidative stress in aceruloplas- consequences of the cytochrome c oxidase deficiency in brain of copper-
minemia brain. Free Radic. Res. 36, 303–306. deficient Mo(vbr) mice. J. Neurochem. 72, 1580–1585.
52 I.F. Scheiber et al. / Progress in Neurobiology 116 (2014) 33–57

Kuo, S.M., Huang, C.T., Blum, P., Chang, C., 2001a. Quercetin cumulatively enhances Liddell, J.R., Bush, A.I., White, A.R., 2013. Copper in brain and neurodegeneration. In:
copper induction of metallothionein in intestinal cells. Biol. Trace Elem. Res. 84, Culotta, V., Scott, R.A. (Eds.), Encyclopedia of Inorganic and Bioinorganic Chem-
1–10. istry. Wiley, Hoboken, NJ.
Kuo, Y.M., Gybina, A.A., Pyatskowit, J.W., Gitschier, J., Prohaska, J.R., 2006. Copper Lim, C.M., Cater, M.A., Mercer, J.F., La Fontaine, S., 2006. Copper-dependent inter-
transport protein (Ctr1) levels in mice are tissue specific and dependent on action of glutaredoxin with the N termini of the copper-ATPases (ATP7A and
copper status. J. Nutr. 136, 21–26. ATP7B) defective in Menkes and Wilson diseases. Biochem. Biophys. Res.
Kuo, Y.M., Zhou, B., Cosco, D., Gitschier, J., 2001b. The copper transporter CTR1 Commun. 348, 428–436.
provides an essential function in mammalian embryonic development. Proc. Linden, R., Martins, V.R., Prado, M.A., Cammarota, M., Izquierdo, I., Brentani, R.R.,
Natl. Acad. Sci. U.S.A. 98, 6836–6841. 2008. Physiology of the prion protein. Physiol. Rev. 88, 673–728.
La Fontaine, S., Firth, S.D., Lockhart, P.J., Brooks, H., Parton, R.G., Camakaris, J., Liu, J.J., Jamieson, S.M., Subramaniam, J., Ip, V., Jong, N.N., Mercer, J.F., McKeage, M.J.,
Mercer, J.F., 1998. Functional analysis and intracellular localization of the 2009. Neuronal expression of copper transporter 1 in rat dorsal root ganglia:
human menkes protein (MNK) stably expressed from a cDNA construct in association with platinum neurotoxicity. Cancer Chemother. Pharmacol. 64,
Chinese hamster ovary cells (CHO-K1). Hum. Mol. Genet. 7, 1293–1300. 847–856.
La Fontaine, S., Mercer, J.F., 2007. Trafficking of the copper-ATPases, ATP7A and Loeffler, D.A., LeWitt, P.A., Juneau, P.L., Sima, A.A., Nguyen, H.U., DeMaggio, A.J.,
ATP7B: role in copper homeostasis. Arch. Biochem. Biophys. 463, 149–167. Brickman, C.M., Brewer, G.J., Dick, R.D., Troyer, M.D., Kanaley, L., 1996. Increased
Lai, C.C., Huang, W.H., Askari, A., Wang, Y., Sarvazyan, N., Klevay, L.M., Chiu, T.H., regional brain concentrations of ceruloplasmin in neurodegenerative disorders.
1994. Differential regulation of superoxide dismutase in copper-deficient rat Brain Res. 738, 265–274.
organs. Free Radic. Biol. Med. 16, 613–620. Lorincz, M.T., 2010. Neurologic Wilson’s disease. Ann. N. Y. Acad. Sci. 1184,
Lamb, A.L., Torres, A.S., O’Halloran, T.V., Rosenzweig, A.C., 2001. Heterodimeric 173–187.
structure of superoxide dismutase in complex with its metallochaperone. Nat. Lovell, M.A., Robertson, J.D., Teesdale, W.J., Campbell, J.L., Markesbery, W.R., 1998.
Struct. Biol. 8, 751–755. Copper, iron and zinc in Alzheimer’s disease senile plaques. J. Neurol. Sci. 158,
Lannfelt, L., Blennow, K., Zetterberg, H., Batsman, S., Ames, D., Harrison, J., Masters, 47–52.
C.L., Targum, S., Bush, A.I., Murdoch, R., Wilson, J., Ritchie, C.W., 2008. Safety, Lutsenko, S., Barnes, N.L., Bartee, M.Y., Dmitriev, O.Y., 2007. Function and regulation
efficacy, and biomarker findings of PBT2 in targeting Abeta as a modifying of human copper-transporting ATPases. Physiol. Rev. 87, 1011–1046.
therapy for Alzheimer’s disease: a phase IIa, double-blind, randomised, place- Lutsenko, S., Bhattacharjee, A., Hubbard, A.L., 2010. Copper handling machinery of
bo-controlled trial. Lancet Neurol. 7, 779–786. the brain. Metallomics 2, 596–608.
Larner, F., Sampson, B., Rehkämper, M., Weiss, D.J., Dainty, J.R., ÓRiordan, S., Panetta, Macreadie, I.G., 2008. Copper transport and Alzheimer’s disease. Eur. Biophys. J. 37,
T., Bain, P.G., 2013. High precision isotope measurements reveal poor control of 295–300.
copper metabolism in Parkinsonism. Metallomics 5, 125–132. Maehara, M., Ogasawara, N., Mizutani, N., Watanabe, K., Suzuki, S., 1983. Cyto-
Leach, S.P., Salman, M.D., Hamar, D., 2006. Trace elements and prion diseases: a chrome c oxidase deficiency in Menkes kinky hair disease. Brain Dev. 5, 533–
review of the interactions of copper, manganese and zinc with the prion protein. 540.
Anim. Health Res. Rev. 7, 97–105. Magistretti, P.J., Cardinaux, J.R., Martin, J.L., 1998. VIP and PACAP in the CNS:
Leary, S.C., Cobine, P.A., Kaufman, B.A., Guercin, G.H., Mattman, A., Palaty, J., Lock- regulators of glial energy metabolism and modulators of glutamatergic signal-
itch, G., Winge, D.R., Rustin, P., Horvath, R., Shoubridge, E.A., 2007. The human ing. Ann. N. Y. Acad. Sci. 865, 213–225.
cytochrome c oxidase assembly factors SCO1 and SCO2 have regulatory roles in Marcus, D.L., Thomas, C., Rodriguez, C., Simberkoff, K., Tsai, J.S., Strafaci, J.A.,
the maintenance of cellular copper homeostasis. Cell Metab. 5, 9–20. Freedman, M.L., 1998. Increased peroxidation and reduced antioxidant enzyme
Leary, S.C., Kaufman, B.A., Pellecchia, G., Guercin, G.H., Mattman, A., Jaksch, M., activity in Alzheimer’s disease. Exp. Neurol. 150, 40–44.
Shoubridge, E.A., 2004. Human SCO1 and SCO2 have independent, cooperative Maret, W., 2011. Redox biochemistry of mammalian metallothioneins. J. Biol. Inorg.
functions in copper delivery to cytochrome c oxidase. Hum. Mol. Genet. 13, Chem. 16, 1079–1086.
1839–1848. Markiewicz, I., Lukomska, B., 2006. The role of astrocytes in the physiology and
Leary, S.C., Sasarman, F., Nishimura, T., Shoubridge, E.A., 2009a. Human SCO2 is pathology of the central nervous system. Acta Neurobiol. Exp. (Wars) 66,
required for the synthesis of CO II and as a thiol-disulphide oxidoreductase for 343–358.
SCO1. Hum. Mol. Genet. 18, 2230–2240. Maryon, E.B., Molloy, S.A., Zimnicka, A.M., Kaplan, J.H., 2007. Copper entry into
Leary, S.C., Winge, D.R., Cobine, P.A., 2009b. Pulling the plug on cellular copper: the human cells: progress and unanswered questions. Biometals 20, 355–364.
role of mitochondria in copper export. Biochim. Biophys. Acta 1793, 146–153. Mathiisen, T.M., Lehre, K.P., Danbolt, N.C., Ottersen, O.P., 2010. The perivascular
Lech, T., Sadlik, J.K., 2007. Copper concentration in body tissues and fluids in normal astroglial sheath provides a complete covering of the brain microvessels: an
subjects of southern Poland. Biol. Trace Elem. Res. 118, 10–15. electron microscopic 3D reconstruction. Glia 58, 1094–1103.
Lee, J., Pena, M.M., Nose, Y., Thiele, D.J., 2002a. Biochemical characterization of the Matsuba, Y., Takahashi, Y., 1970. Spectrophotometric determination of copper with
human copper transporter Ctr1. J. Biol. Chem. 277, 4380–4387. N,N,N’,N’-tetraethylthiuram disulfide and an application of this method for
Lee, J., Petris, M.J., Thiele, D.J., 2002b. Characterization of mouse embryonic cells studies on subcellular distribution of copper in rat brain. Anal. Biochem. 36,
deficient in the ctr1 high affinity copper transporter Identification of a Ctr1- 182–191.
independent copper transport system. J. Biol. Chem. 277, 40253–40259. Mattatall, N.R., Jazairi, J., Hill, B.C., 2000. Characterization of YpmQ, an accessory
Lee, J., Prohaska, J.R., Dagenais, S.L., Glover, T.W., Thiele, D.J., 2000. Isolation of a protein required for the expression of cytochrome c oxidase in Bacillus subtilis.
murine copper transporter gene, tissue specific expression and functional J. Biol. Chem. 275, 28802–28809.
complementation of a yeast copper transport mutant. Gene 254, 87–96. Maurer, I., Zierz, S., Moller, H.J., 2000. A selective defect of cytochrome c oxidase is
Lee, J., Prohaska, J.R., Thiele, D.J., 2001. Essential role for mammalian copper present in brain of Alzheimer disease patients. Neurobiol. Aging 21, 455–462.
transporter Ctr1 in copper homeostasis and embryonic development. Proc. Maxfield, A.B., Heaton, D.N., Winge, D.R., 2004. Cox17 is functional when tethered to
Natl. Acad. Sci. U.S.A. 98, 6842–6847. the mitochondrial inner membrane. J. Biol. Chem. 279, 5072–5080.
Leitch, J.M., Jensen, L.T., Bouldin, S.D., Outten, C.E., Hart, P.J., Culotta, V.C., 2009a. Maynard, C.J., Cappai, R., Volitakis, I., Cherny, R.A., White, A.R., Beyreuther, K.,
Activation of Cu,Zn-superoxide dismutase in the absence of oxygen and the Masters, C.L., Bush, A.I., Li, Q.X., 2002. Overexpression of Alzheimer’s disease
copper chaperone CCS. J. Biol. Chem. 284, 21863–21871. amyloid-beta opposes the age-dependent elevations of brain copper and iron. J.
Leitch, J.M., Yick, P.J., Culotta, V.C., 2009b. The right to choose: multiple pathways Biol. Chem. 277, 44670–44676.
for activating copper,zinc superoxide dismutase. J. Biol. Chem. 284, 24679– McFarland, K.N., Cha, J.H., 2011. Molecular biology of Huntington’s disease. Handb.
24683. Clin. Neurol. 100, 25–81.
Leiva, J., Palestini, M., Infante, C., Goldschmidt, A., Motles, E., 2009. Copper sup- McNeill, A., Pandolfo, M., Kuhn, J., Shang, H., Miyajima, H., 2008. The neurological
presses hippocampus LTP in the rat, but does not alter learning or memory in presentation of ceruloplasmin gene mutations. Eur. Neurol. 60, 200–205.
the morris water maze. Brain Res. 1256, 69–75. Meng, Y., Miyoshi, I., Hirabayashi, M., Su, M., Mototani, Y., Okamura, T., Terada, K.,
Leone, A., Pavlakis, G.N., Hamer, D.H., 1985. Menkes’ disease: abnormal metal- Ueda, M., Enomoto, K., Sugiyama, T., Kasai, N., 2004. Restoration of copper
lothionein gene regulation in response to copper. Cell 40, 301–309. metabolism and rescue of hepatic abnormalities in LEC rats, an animal model of
Lesne, S., Ali, C., Gabriel, C., Croci, N., MacKenzie, E.T., Glabe, C.G., Plotkine, M., Wilson disease, by expression of human ATP7B gene. Biochim. Biophys. Acta
Marchand-Verrecchia, C., Vivien, D., Buisson, A., 2005. NMDA receptor activa- 1690, 208–219.
tion inhibits alpha-secretase and promotes neuronal amyloid-beta production. Mercer, J.F., Grimes, A., Rauch, H., 1992. Hepatic metallothionein gene expression in
J. Neurosci. 25, 9367–9377. toxic milk mice. J. Nutr. 122, 1254–1259.
Letelier, M.E., Lepe, A.M., Faundez, M., Salazar, J., Marin, R., Aracena, P., Speisky, H., Meyer, L.A., Durley, A.P., Prohaska, J.R., Harris, Z.L., 2001. Copper transport and
2005. Possible mechanisms underlying copper-induced damage in biological metabolism are normal in aceruloplasminemic mice. J. Biol. Chem. 276, 36857–
membranes leading to cellular toxicity. Chem. Biol. Interact. 151, 71–82. 36861.
Letelier, M.E., Martinez, M., Gonzalez-Lira, V., Faundez, M., Aracena-Parks, P., 2006. Miao, L., St Clair, D.K., 2009. Regulation of superoxide dismutase genes: implications
Inhibition of cytosolic glutathione S-transferase activity from rat liver by in disease. Free Radic. Biol. Med. 47, 344–356.
copper. Chem. Biol. Interact. 164, 39–48. Michalska, A.E., Choo, K.H., 1993. Targeting and germ-line transmission of a null
Lewinska-Preis, L., Jablonska, M., Fabianska, M.J., Kita, A., 2011. Bioelements and mutation at the metallothionein I and II loci in mouse. Proc. Natl. Acad. Sci.
mineral matter in human livers from the highly industrialized region of the U.S.A. 90, 8088–8092.
Upper Silesia Coal Basin (Poland). Environ. Geochem. Health 33, 595–611. Millhauser, G.L., 2007. Copper and the prion protein: methods, structures, function,
Li, C., Peoples, R.W., Weight, F.F., 1996. Cu2+ potently enhances ATP-activated and disease. Annu. Rev. Phys. Chem. 58, 299–320.
current in rat nodose ganglion neurons. Neurosci. Lett. 219, 45–48. Milne, L., Nicotera, P., Orrenius, S., Burkitt, M.J., 1993. Effects of glutathione and
Lichten, L.A., Cousins, R.J., 2009. Mammalian zinc transporters: nutritional and chelating agents on copper-mediated DNA oxidation: pro-oxidant and antioxi-
physiologic regulation. Annu. Rev. Nutr. 29, 153–176. dant properties of glutathione. Arch. Biochem. Biophys. 304, 102–109.
I.F. Scheiber et al. / Progress in Neurobiology 116 (2014) 33–57 53

Miras, R., Morin, I., Jacquin, O., Cuillel, M., Guillain, F., Mintz, E., 2008. Interplay Niciu, M.J., Ma, X.M., El Meskini, R., Pachter, J.S., Mains, R.E., Eipper, B.A., 2007.
between glutathione, Atx1 and copper 1. Copper(I) glutathionate induced Altered ATP7A expression and other compensatory responses in a murine
dimerization of Atx1. J. Biol. Inorg. Chem. 13, 195–205. model of Menkes disease. Neurobiol. Dis. 27, 278–291.
Mitteregger, G., Korte, S., Shakarami, M., Herms, J., Kretzschmar, H.A., 2009. Role of Niciu, M.J., Ma, X.M., El Meskini, R., Ronnett, G.V., Mains, R.E., Eipper, B.A., 2006.
copper and manganese in prion disease progression. Brain Res. 1292, 155–164. Developmental changes in the expression of ATP7A during a critical period in
Miura, T., Sasaki, S., Toyama, A., Takeuchi, H., 2005. Copper reduction by the postnatal neurodevelopment. Neuroscience 139, 947–964.
octapeptide repeat region of prion protein: pH dependence and implications Nischwitz, V., Berthele, A., Michalke, B., 2008. Speciation analysis of selected metals
in cellular copper uptake. Biochemistry 44, 8712–8720. and determination of their total contents in paired serum and cerebrospinal
Miyajima, H., 2003. Aceruloplasminemia, an iron metabolic disorder. Neuropathol- fluid samples: an approach to investigate the permeability of the human blood–
ogy 23, 345–350. cerebrospinal fluid-barrier. Anal. Chim. Acta 627, 258–269.
Miyajima, H., Kono, S., Takahashi, Y., Sugimoto, M., 2002. Increased lipid peroxida- Nishihara, E., Furuyama, T., Yamashita, S., Mori, N., 1998. Expression of copper
tion and mitochondrial dysfunction in aceruloplasminemia brains. Blood Cells. trafficking genes in the mouse brain. Neuroreport 9, 3259–3263.
Mol. Dis. 29, 433–438. Nishimura, H., Nishimura, N., Tohyama, C., 1989. Immunohistochemical localiza-
Miyajima, H., Takahashi, Y., Shimizu, H., Sakai, N., Kamata, T., Kaneko, E., 1996. Late tion of metallothionein in developing rat tissues. J. Histochem. Cytochem. 37,
onset diabetes mellitus in patients with hereditary aceruloplasminemia. Intern. 715–722.
Med. 35, 641–645. Niu, Z.D., Chen, J.T., Wang, S., Wang, M., Li, X.M., Ruan, D.Y., 2006. Inhibition of
Miyayama, T., Suzuki, K.T., Ogra, Y., 2009. Copper accumulation and compartmen- delayed rectifier K+ currents by copper in acutely dissociated rat hippocampal
talization in mouse fibroblast lacking metallothionein and copper chaperone, CA1 neurons. Toxicol. Lett. 165, 289–296.
Atox1. Toxicol. Appl. Pharmacol. 237, 205–213. Niu, Z.D., Yu, K., Gu, Y., Wang, M., She, J.Q., Chen, W.H., Ruan, D.Y., 2005. Effects of
Miyazaki, I., Asanuma, M., Kikkawa, Y., Takeshima, M., Murakami, S., Miyoshi, K., copper on A-type potassium currents in acutely dissociated rat hippocampal
Sogawa, N., Kita, T., 2011. Astrocyte-derived metallothionein protects dopami- CA1 neurons. Neuroreport 16, 1585–1589.
nergic neurons from dopamine quinone toxicity. Glia 59, 435–451. Nobrega, M.P., Bandeira, S.C., Beers, J., Tzagoloff, A., 2002. Characterization of
Moffett, J.R., Ross, B., Arun, P., Madhavarao, C.N., Namboodiri, A.M., 2007. COX19, a widely distributed gene required for expression of mitochondrial
N-Acetylaspartate in the CNS: from neurodiagnostics to neurobiology. Prog. cytochrome oxidase. J. Biol. Chem. 277, 40206–40211.
Neurobiol. 81, 89–131. Nooijen, J.L., De Groot, C.J., Van den Hamer, C.J., Monnens, L.A., Willemse, J.,
Moltedo, O., Verde, C., Capasso, A., Parisi, E., Remondelli, P., Bonatti, S., Alvarez- Niermeijer, M.F., 1981. Trace element studies in three patients and a fetus
Hernandez, X., Glass, J., Alvino, C.G., Leone, A., 2000. Zinc transport and metal- with Menkes’ disease. Effect of copper therapy. Pediatr. Res. 15, 284–289.
lothionein secretion in the intestinal human cell line Caco-2. J. Biol. Chem. 275, Nyasae, L., Bustos, R., Braiterman, L., Eipper, B., Hubbard, A., 2007. Dynamics of
31819–31825. endogenous ATP7A (Menkes protein) in intestinal epithelial cells: copper-
Mondola, P., Annella, T., Santillo, M., Santangelo, F., 1996. Evidence for secretion of dependent redistribution between two intracellular sites. Am. J. Physiol. Gas-
cytosolic CuZn superoxide dismutase by Hep G2 cells and human fibroblasts. trointest. Liver Physiol. 292, G1181–G1194.
Int. J. Biochem. Cell Biol. 28, 677–681. Ogra, Y., Aoyama, M., Suzuki, K.T., 2006. Protective role of metallothionein against
Mondola, P., Annella, T., Seru, R., Santangelo, F., Iossa, S., Gioielli, A., Santillo, M., copper depletion. Arch. Biochem. Biophys. 451, 112–118.
1998. Secretion and increase of intracellular CuZn superoxide dismutase con- Okado-Matsumoto, A., Fridovich, I., 2001. Subcellular distribution of superoxide
tent in human neuroblastoma SK-N-BE cells subjected to oxidative stress. Brain dismutases (SOD) in rat liver: Cu,Zn-SOD in mitochondria. J. Biol. Chem. 276,
Res. Bull. 45, 517–520. 38388–38393.
Mondola, P., Santillo, M., Seru, R., Damiano, S., Alvino, C., Ruggiero, G., Formisano, P., Olivieri, S., Conti, A., Iannaccone, S., Cannistraci, C.V., Campanella, A., Barbariga, M.,
Terrazzano, G., Secondo, A., Annunziato, L., 2004. Cu,Zn superoxide dismutase Codazzi, F., Pelizzoni, I., Magnani, G., Pesca, M., Franciotta, D., Cappa, S.F.,
increases intracellular calcium levels via a phospholipase C-protein kinase C Alessio, M., 2011. Ceruloplasmin oxidation, a feature of Parkinson’ disease
pathway in SK-N-BE neuroblastoma cells. Biochem. Biophys. Res. Commun. CSF, inhibits ferroxidase activity and promotes cellular iron retention. J. Neu-
324, 887–892. rosci. 31, 18568–18577.
Monnot, A.D., Behl, M., Ho, S., Zheng, W., 2011. Regulation of brain copper homeo- Olusola, A.O., Obodozie, O.O., Nssien, M., Adaramoye, A., Adesanoye, O., Odama, L.E.,
stasis by the brain barrier systems: effects of Fe-overload and Fe-deficiency. Emerole, G.O., 2004. Concentrations of copper, iron, and zinc in the major
Toxicol. Appl. Pharmacol. 256, 249–257. organs of the wistar albino and wild black rats: a comparative study. Biol. Trace
Moore, R.A., Taubner, L.M., Priola, S.A., 2009. Prion protein misfolding and disease. Elem. Res. 98, 265–274.
Curr. Opin. Struct. Biol. 19, 14–22. Ono, S., Koropatnick, D.J., Cherian, M.G., 1997. Regional brain distribution of
Morel, F., Doussiere, J., Vignais, P.V., 1991. The superoxide-generating oxidase of metallothionein, zinc and copper in toxic milk mutant and transgenic mice.
phagocytic cells Physiological, molecular and pathological aspects. Eur. J. Toxicology 124, 1–10.
Biochem. 201, 523–546. Orlowski, C., Piotrowski, J.K., 1998. Metal composition of human hepatic and renal
Mosconi, L., Pupi, A., De Leon, M.J., 2008. Brain glucose hypometabolism and metallothionein. Biol. Trace Elem. Res. 65, 133–141.
oxidative stress in preclinical Alzheimer’s disease. Ann. N. Y. Acad. Sci. 1147, Oswald, C., Krause-Buchholz, U., Rodel, G., 2009. Knockdown of human COX17
180–195. affects assembly and supramolecular organization of cytochrome c oxidase. J.
Mossakowski, M.J., Renkawek, K., Krasnicka, Z., Smialek, M., Pronaszko, A., 1970. Mol. Biol. 389, 470–479.
Morphology and histochemistry of Wilsonian and hepatogenic gliopathy in Palacios, O., Atrian, S., Capdevila, M., 2011. Zn- and Cu-thioneins: a functional
tissue culture. Acta Neuropathol. 16, 1–16. classification for metallothioneins? J. Biol. Inorg. Chem. 16, 991–1009.
Mulder, T.P.J., Janssens, A.R., Verspaget, H.W., Vanhattum, J., Lamers, C.B.H.W., 1992. Palmiter, R.D., Findley, S.D., Whitmore, T.E., Durnam, D.M., 1992. MT-III, a brain-
Metallothionein concentration in the liver of patients with Wilsons disease, specific member of the metallothionein gene family. Proc. Natl. Acad. Sci. U.S.A.
primary biliary-cirrhosis, and liver metastasis of colorectal cancer. J. Hepatol. 89, 6333–6337.
16, 346–350. Pamp, K., Bramey, T., Kirsch, M., De Groot, H., Petrat, F., 2005. NAD(H) enhances the
Muller, P.A., Klomp, L.W., 2009. ATOX1: a novel copper-responsive transcription Cu(II)-mediated inactivation of lactate dehydrogenase by increasing the acces-
factor in mammals? Int. J. Biochem. Cell Biol. 41, 1233–1236. sibility of sulfhydryl groups. Free Radic. Res. 39, 31–40.
Multhaup, G., Schlicksupp, A., Hesse, L., Beher, D., Ruppert, T., Masters, C.L., Beyr- Papadopoulou, L.C., Sue, C.M., Davidson, M.M., Tanji, K., Nishino, I., Sadlock, J.E.,
euther, K., 1996. The amyloid precursor protein of Alzheimer’s disease in the Krishna, S., Walker, W., Selby, J., Glerum, D.M., Coster, R.V., Lyon, G., Scalais, E.,
reduction of copper(II) to copper(I). Science 271, 1406–1409. Lebel, R., Kaplan, P., Shanske, S., De Vivo, D.C., Bonilla, E., Hirano, M., DiMauro, S.,
Multhaup, G., Scheuermann, S., Schlicksupp, A., Simons, A., Strauss, M., Kemmling, Schon, E.A., 1999. Fatal infantile cardioencephalomyopathy with COX deficien-
A., Oehler, C., Cappai, R., Pipkorn, R., Bayer, T.A., 2002. Possible mechanisms of cy and mutations in SCO2, a COX assembly gene. Nat. Genet. 23, 333–337.
APP-mediated oxidative stress in Alzheimer’s disease. Free Radic. Biol. Med. 33, Park, J.D., Liu, Y.P., Klaassen, C.D., 2001. Protective effect of metallothionein against
45–51. the toxicity of cadmium and other metals. Toxicology 163, 93–100.
Mutisya, E.M., Bowling, A.C., Beal, M.F., 1994. Cortical cytochrome oxidase activity is Parker Jr., W.D., Parks, J., Filley, C.M., Kleinschmidt-DeMasters, B.K., 1994. Electron
reduced in Alzheimer’s disease. J. Neurochem. 63, 2179–2184. transport chain defects in Alzheimer’s disease brain. Neurology 44, 1090–1096.
Nadal, R.C., Davies, P., Brown, D.R., Viles, J.H., 2009. Evaluation of Cu2+ affinities for Parpura, V., Heneka, M.T., Montana, V., Oliet, S.H., Schousboe, A., Haydon, P.G.,
the prion protein. Biochemistry 48, 8929–8931. Stout Jr., R.F., Spray, D.C., Reichenbach, A., Pannicke, T., Pekny, M., Pekna, M.,
Naeve, G.S., Vana, A.M., Eggold, J.R., Kelner, G.S., Maki, R., Desouza, E.B., Foster, A.C., Zorec, R., Verkhratsky, A., 2012. Glial cells in (patho)physiology. J. Neuro-
1999. Expression profile of the copper homeostasis gene, rAtox1, in the rat chem. 121, 4–27.
brain. Neuroscience 93, 1179–1187. Pase, L., Voskoboinik, I., Greenough, M., Camakaris, J., 2004. Copper stimulates
Nagano, T., Itoh, N., Ebisutani, C., Takatani, T., Miyoshi, T., Nakanishi, T., Tanaka, K., trafficking of a distinct pool of the Menkes copper ATPase (ATP7A) to the plasma
2000. The transport mechanism of metallothionein is different from that of membrane and diverts it into a rapid recycling pool. Biochem. J. 378, 1031–
classical NLS-bearing protein. J. Cell. Physiol. 185, 440–446. 1037.
Nalivaeva, N.N., Turner, A.J., 2013. The amyloid precursor protein: a biochemical Pauly, P.C., Harris, D.A., 1998. Copper stimulates endocytosis of the prion protein. J.
enigma in brain development, function and disease. FEBS Lett. 587, 2046– Biol. Chem. 273, 33107–33110.
2054. Payne, A.S., Gitlin, J.D., 1998. Functional expression of the menkes disease protein
Nedergaard, M., Ransom, B., Goldman, S.A., 2003. New roles for astrocytes: redefin- reveals common biochemical mechanisms among the copper-transporting P-
ing the functional architecture of the brain. Trends Neurosci. 26, 523–530. type ATPases. J. Biol. Chem. 273, 3765–3770.
Nguyen, T., Hamby, A., Massa, S.M., 2005. Clioquinol down-regulates mutant Paynter, J.A., Grimes, A., Lockhart, P., Mercer, J.F., 1994. Expression of the Menkes
huntingtin expression in vitro and mitigates pathology in a Huntington’s gene homologue in mouse tissues lack of effect of copper on the mRNA levels.
disease mouse model. Proc. Natl. Acad. Sci. U.S.A. 102, 11840–11845. FEBS Lett. 351, 186–190.
54 I.F. Scheiber et al. / Progress in Neurobiology 116 (2014) 33–57

Pelizzoni, I., Zacchetti, D., Smith, C.P., Grohovaz, F., Codazzi, F., 2012. Expression of and antioxidant enzyme activities but not copper delivery. J. Biol. Chem. 278,
divalent metal transporter 1 in primary hippocampal neurons: reconsidering its 9064–9072.
role in non-transferrin-bound iron influx. J. Neurochem. 120, 269–278. Rae, T.D., Schmidt, P.J., Pufahl, R.A., Culotta, V.C., O’Halloran, T.V., 1999. Undetect-
Pellerin, L., Bouzier-Sore, A.K., Aubert, A., Serres, S., Merle, M., Costalat, R., able intracellular free copper: the requirement of a copper chaperone for
Magistretti, P.J., 2007. Activity-dependent regulation of energy metabolism superoxide dismutase. Science 284, 805–808.
by astrocytes: an update. Glia 55, 1251–1262. Rae, T.D., Torres, A.S., Pufahl, R.A., O’Halloran, T.V., 2001. Mechanism of
Perea, G., Araque, A., 2010. GLIA modulates synaptic transmission. Brain Res. Rev. Cu,Zn-superoxide dismutase activation by the human metallochaperone hCCS.
63, 93–102. J. Biol. Chem. 276, 5166–5176.
Perry, J.J., Shin, D.S., Getzoff, E.D., Tainer, J.A., 2010. The structural biochemistry of Rahil-Khazen, R., Bolann, B.J., Myking, A., Ulvik, R.J., 2002. Multi-element analysis of
the superoxide dismutases. Biochim. Biophys. Acta 1804, 245–262. trace element levels in human autopsy tissues by using inductively coupled
Petering, D.H., Zhu, J., Krezoski, S., Meeusen, J., Kiekenbush, C., Krull, S., Specher, T., atomic emission spectrometry technique (ICP-AES). J. Trace Elem. Med. Biol. 16,
Dughish, M., 2006. Apo-metallothionein emerging as a major player in the 15–25.
cellular activities of metallothionein. Exp. Biol. Med. (Maywood) 231, 1528– Ralle, M., Lutsenko, S., Blackburn, N.J., 2003. X-ray absorption spectroscopy of the
1534. copper chaperone HAH1 reveals a linear two-coordinate Cu(I) center capable of
Peters, C., Munoz, B., Sepulveda, F.J., Urrutia, J., Quiroz, M., Luza, S., De Ferrari, G.V., adduct formation with exogenous thiols and phosphines. J. Biol. Chem. 278,
Aguayo, L.G., Opazo, C., 2011. Biphasic effects of copper on neurotransmission in 23163–23170.
rat hippocampal neurons. J. Neurochem. 119, 78–88. Rana, A., Gnaneswari, D., Bansal, S., Kundu, B., 2009. Prion metal interaction: is prion
Petersen, S.V., Enghild, J.J., 2005. Extracellular superoxide dismutase: structural and pathogenesis a cause or a consequence of metal imbalance? Chem. Biol.
functional considerations of a protein shaped by two different disulfide bridge Interact. 181, 282–291.
patterns. Biomed. Pharmacother. 59, 175–182. Raudenska, M., Gumulec, J., Podlaha, O., Sztalmachova, M., Babula, P., Eckschlager,
Petris, M.J., Mercer, J.F., 1999. The Menkes protein (ATP7A; MNK) cycles via the T., Adam, V., Kizek, R., Masarik, M., 2013. Metallothionein polymorphisms in
plasma membrane both in basal and elevated extracellular copper using a pathological processes. Metallomics 6, 55–68.
C-terminal di-leucine endocytic signal. Hum. Mol. Genet. 8, 2107–2115. Ray, K., Chaki, M., Sengupta, M., 2007. Tyrosinase and ocular diseases: some novel
Petris, M.J., Mercer, J.F., Culvenor, J.G., Lockhart, P., Gleeson, P.A., Camakaris, J., 1996. thoughts on the molecular basis of oculocutaneous albinism type 1. Prog. Retin.
Ligand-regulated transport of the Menkes copper P-type ATPase efflux pump Eye Res. 26, 323–358.
from the Golgi apparatus to the plasma membrane: a novel mechanism of Rembach, A., Hare, D.J., Lind, M., Fowler, C.J., Cherny, R.A., McLean, C., Bush, A.I.,
regulated trafficking. EMBO J. 15, 6084–6095. Masters, C.L., Roberts, B.R., 2013. Decreased copper in Alzheimer’s disease brain
Petris, M.J., Strausak, D., Mercer, J.F., 2000. The Menkes copper transporter is is predominantly in the soluble extractable fraction. Int. J. Alzheimers Dis. 2013,
required for the activation of tyrosinase. Hum. Mol. Genet. 9, 2845–2851. 623241.
Pfeiffer, R.F., 2007. Wilson’s Disease. Semin. Neurol. 27, 123–132. Rhodes, C.H., Xu, R.Y., Angeletti, R.H., 1990. Peptidylglycine alpha-amidating mono-
Poger, D., Fillaux, C., Miras, R., Crouzy, S., Delangle, P., Mintz, E., Den Auwer, C., oxygenase (PAM) in Schwann cells and glia as well as neurons. J. Histochem.
Ferrand, M., 2008. Interplay between glutathione, Atx1 and copper: X-ray Cytochem. 38, 1301–1311.
absorption spectroscopy determination of Cu(I) environment in an Atx1 dimer. Rigby, K., Zhang, L., Cobine, P.A., George, G.N., Winge, D.R., 2007. Characterization of
J. Biol. Inorg. Chem. 13, 1239–1248. the cytochrome c oxidase assembly factor Cox19 of Saccharomyces cerevisiae. J.
Pooler, A.M., Polydoro, M., Wegmann, S., Nicholls, S.B., Spires-Jones, T.L., Hyman, Biol. Chem. 282, 10233–10242.
B.T., 2013. Propagation of tau pathology in Alzheimer’s disease: identification of Rivera-Mancia, S., Perez-Neri, I., Rios, C., Tristan-Lopez, L., Rivera-Espinosa, L.,
novel therapeutic targets. Alzheimers Res. Ther. 5, 49. Montes, S., 2010. The transition metals copper and iron in neurodegenerative
Pope, S.A., Milton, R., Heales, S.J., 2008. Astrocytes protect against copper-catalysed diseases. Chem. Biol. Interact. 186, 184–199.
loss of extracellular glutathione. Neurochem. Res. 33, 1410–1418. Roberts, B.R., Ryan, T.M., Bush, A.I., Masters, C.L., Duce, J.A., 2012. The role of
Popescu, B.F., George, M.J., Bergmann, U., Garachtchenko, A.V., Kelly, M.E., McCrea, metallobiology and amyloid-beta peptides in Alzheimer’s disease. J. Neuro-
R.P., Luning, K., Devon, R.M., George, G.N., Hanson, A.D., Harder, S.M., Chapman, chem. 120 (Suppl 1) 149–166.
L.D., Pickering, I.J., Nichol, H., 2009. Mapping metals in Parkinson’s and normal Robertson, D., Haile, V., Perry, S.E., Robertson, R.M., Phillips 3rd, J.A., Biaggioni, I.,
brain using rapid-scanning x-ray fluorescence. Phys. Med. Biol. 54, 651–663. 1991. Dopamine beta-hydroxylase deficiency. A genetic disorder of cardiovas-
Popovic, D.M., Leontyev, I.V., Beech, D.G., Stuchebrukhov, A.A., 2010. Similarity of cular regulation. Hypertension 18, 1–8.
cytochrome c oxidases in different organisms. Proteins 78, 2691–2698. Robinson, N.J., Winge, D.R., 2010. Copper metallochaperones. Annu. Rev. Biochem.
Prigge, S.T., Kolhekar, A.S., Eipper, B.A., Mains, R.E., Amzel, L.M., 1997. Amidation of 79, 537–562.
bioactive peptides: the structure of peptidylglycine alpha-hydroxylating mono- Rodriguez-Granillo, A., Crespo, A., Estrin, D.A., Wittung-Stafshede, P., 2010. Copper-
oxygenase. Science 278, 1300–1305. transfer mechanism from the human chaperone Atox1 to a metal-binding
Prigge, S.T., Mains, R.E., Eipper, B.A., Amzel, L.M., 2000. New insights into copper domain of Wilson disease protein. J. Phys. Chem. B 114, 3698–3706.
monooxygenases and peptide amidation: structure, mechanism and function. Roschitzki, B., Vasak, M., 2002. A distinct Cu(4)-thiolate cluster of human
Cell. Mol. Life Sci. 57, 1236–1259. metallothionein-3 is located in the N-terminal domain. J. Biol. Inorg. Chem.
Prohaska, J.R., 1991. Changes in Cu,Zn-superoxide dismutase, cytochrome c oxidase, 7, 611–616.
glutathione peroxidase and glutathione transferase activities in copper- Rossi, L., Lombardo, M.F., Ciriolo, M.R., Rotilio, G., 2004. Mitochondrial dysfunction
deficient mice and rats. J. Nutr. 121, 355–363. in neurodegenerative diseases associated with copper imbalance. Neurochem.
Prohaska, J.R., Bailey, W.R., Lear, P.M., 1995. Copper deficiency alters rat pepti- Res. 29, 493–504.
dylglycine alpha-amidating monooxygenase activity. J. Nutr. 125, 1447– Rothstein, J.D., Dykes-Hoberg, M., Corson, L.B., Becker, M., Cleveland, D.W., Price,
1454. D.L., Culotta, V.C., Wong, P.C., 1999. The copper chaperone CCS is abundant
Prohaska, J.R., Broderius, M., Brokate, B., 2003. Metallochaperone for Cu,Zn-super- in neurons and astrocytes in human and rodent brain. J. Neurochem. 72,
oxide dismutase (CCS) protein but not mRNA is higher in organs from copper- 422–429.
deficient mice and rats. Arch. Biochem. Biophys. 417, 227–234. Rubino, J.T., Franz, K.J., 2012. Coordination chemistry of copper proteins: how
Prohaska, J.R., Smith, T.L., 1982. Effect of dietary or genetic copper deficiency on nature handles a toxic cargo for essential function. J. Inorg. Biochem. 107,
brain catecholamines, trace metals and enzymes in mice and rats. J. Nutr. 112, 129–143.
1706–1717. Ruiz, F.H., Gonzalez, M., Bodini, M., Opazo, C., Inestrosa, N.C., 1999. Cysteine 144 is a
Pulido, P., Kagi, J.H., Vallee, B.L., 1966. Isolation and some properties of human key residue in the copper reduction by the b-amyloid precursor protein. J.
metallothionein. Biochemistry 5, 1768–1777. Neurochem. 73, 1288–1292.
Punter, F.A., Adams, D.L., Glerum, D.M., 2000. Characterization and localization of Rusinko, N., Prohaska, J.R., 1985. Adenine nucleotide and lactate levels in organs
human COX17, a gene involved in mitochondrial copper transport. Hum. Genet. from copper-deficient mice and brindled mice. J. Nutr. 115, 936–943.
107, 69–74. Sakurai, H., Kamada, H., Fukudome, A., Kito, M., Takeshima, S., Kimura, M., Otaki, N.,
Purchase, R., 2013. The treatment of Wilson’s disease, a rare genetic disorder of Nakajima, K., Kawano, K., Hagino, T., 1992. Copper-metallothionein induction in
copper metabolism. Sci. Prog. 96, 19–32. the liver of LEC rats. Biochem. Biophys. Res. Commun. 185, 548–552.
Pushkar, Y., Robison, G., Sullivan, B., Fu, S.X., Kohne, M., Jiang, W., Rohr, S., Lai, B., Salawu, F.K., Umar, J.T., Olokoba, A.B., 2011. Alzheimer’s disease: a review of recent
Marcus, M.A., Zakharova, T., Zheng, W., 2013. Aging results in copper accumul- developments. Ann. Afr. Med. 10, 73–79.
tion in glial fibrillary acidic protein-positive cells in the subventricular zone. Salazar-Weber, N.L., Smith, J.P., 2011. Copper inhibits NMDA receptor-independent
Aging Cell 12, 823–832. LTP and modulates the paired-pulse ratio after LTP in mouse hippocampal
Qian, Y., Tiffany-Castiglioni, E., Welsh, J., Harris, E.D., 1998. Copper efflux from slices. Int. J. Alzheimers Dis. 2011, 864753.
murine microvascular cells requires expression of the menkes disease Salgado, M.T., Stillman, M.J., 2004. Cu+ distribution in metallothionein fragments.
Cu-ATPase. J. Nutr. 128, 1276–1282. Biochem. Biophys. Res. Commun. 318, 73–80.
Quaife, C.J., Findley, S.D., Erickson, J.C., Froelick, G.J., Kelly, E.J., Zambrowicz, B.P., Samet, J.M., Graves, L.M., Quay, J., Dailey, L.A., Devlin, R.B., Ghio, A.J., Wu, W.,
Palmiter, R.D., 1994. Induction of a new metallothionein isoform (MT-IV) occurs Bromberg, P.A., Reed, W., 1998. Activation of MAPKs in human bronchial
during differentiation of stratified squamous epithelia. Biochemistry 33, 7250– epithelial cells exposed to metals. Am. J. Physiol. 275, L551–L558.
7259. Santon, A., Formigari, A., Irato, P., 2008. The influence of metallothionein on
Quintanar, L., Stoj, C., Taylor, A.B., Hart, P.J., Kosman, D.J., Solomon, E.I., 2007. Shall exposure to metals: an in vitro study on cellular models. Toxicol. In Vitro
we dance? How a multicopper oxidase chooses its electron transfer partner. 22, 980–987.
Acc. Chem. Res. 40, 445–452. Sato, M., Ohtomo, K., Daimon, T., Sugiyama, T., Iijima, K., 1994. Localization of
Rachidi, W., Vilette, D., Guiraud, P., Arlotto, M., Riondel, J., Laude, H., Lehmann, S., copper to afferent terminals in rat locus ceruleus, in contrast to mitochondrial
Favier, A., 2003. Expression of prion protein increases cellular copper binding copper in cerebellum. J. Histochem. Cytochem. 42, 1585–1591.
I.F. Scheiber et al. / Progress in Neurobiology 116 (2014) 33–57 55

Scheiber, I.F., Dringen, R., 2011a. Copper-treatment increases the cellular GSH Stewart, L.C., Klinman, J.P., 1988. Dopamine beta-hydroxylase of adrenal chromaffin
content and accelerates GSH export from cultured rat astrocytes. Neurosci. granules: structure and function. Annu. Rev. Biochem. 57, 551–592.
Lett. 498, 42–46. Stiburek, L., Vesela, K., Hansikova, H., Hulkova, H., Zeman, J., 2009. Loss of function of
Scheiber, I.F., Dringen, R., 2011b. Copper accelerates glycolytic flux in cultured Sco1 and its interaction with cytochrome c oxidase. Am. J. Physiol. Cell Physiol.
astrocytes. Neurochem. Res. 36, 894–903. 296, C1218–C1226.
Scheiber, I.F., Dringen, R., 2013. Astrocyte functions in the copper homeostasis of the Stiburek, L., Zeman, J., 2010. Assembly factors and ATP-dependent proteases in
brain. Neurochem. Int. 62, 556–565. cytochrome c oxidase biogenesis. Biochim. Biophys. Acta 1797, 1149–1158.
Scheiber, I.F., Mercer, J.F., Dringen, R., 2010a. Copper accumulation by cultured Stillman, M.J., 1995. Metallothioneins. Coord. Chem. Rev. 144, 461–511.
astrocytes. Neurochem. Int. 56, 451–460. Stoj, C., Kosman, D.J., 2003. Cuprous oxidase activity of yeast Fet3p and human
Scheiber, I.F., Schmidt, M.M., Dringen, R., 2010b. Zinc prevents the copper-induced ceruloplasmin: implication for function. FEBS Lett. 554, 422–426.
damage of cultured astrocytes. Neurochem. Int. 57, 314–322. Strozyk, D., Launer, L.J., Adlard, P.A., Cherny, R.A., Tsatsanis, A., Volitakis, I., Blennow,
Scheiber, I.F., Schmidt, M.M., Dringen, R., 2012. Copper export from cultured K., Petrovitch, H., White, L.R., Bush, A.I., 2009. Zinc and copper modulate
astrocytes. Neurochem. Int. 60, 292–300. Alzheimer Abeta levels in human cerebrospinal fluid. Neurobiol. Aging 30,
Scheinberg, I.H., Sternlieb, I., 1996. Wilson disease and idiopathic copper toxicosis. 1069–1077.
Am. J. Clin. Nutr. 63, 842S–845S. Stuerenburg, H.J., 2000. CSF copper concentrations, blood–brain barrier function,
Schlief, M.L., Craig, A.M., Gitlin, J.D., 2005. NMDA receptor activation mediates and coeruloplasmin synthesis during the treatment of Wilson’s disease. J.
copper homeostasis in hippocampal neurons. J. Neurosci. 25, 239–246. Neural Transm. 107, 321–329.
Schlief, M.L., Gitlin, J.D., 2006. Copper homeostasis in the CNS: a novel link between Suazo, M., Hodar, C., Morgan, C., Cerpa, W., Cambiazo, V., Inestrosa, N.C., Gonzalez,
the NMDA receptor and copper homeostasis in the hippocampus. Mol. Neuro- M., 2009. Overexpression of amyloid precursor protein increases copper con-
biol. 33, 81–90. tent in HEK293 cells. Biochem. Biophys. Res. Commun. 382, 740–744.
Schlief, M.L., West, T., Craig, A.M., Holtzman, D.M., Gitlin, J.D., 2006. Role of the Sugawara, N., Ikeda, T., Sugawara, C., Kohgo, Y., Kato, J., Takeichi, N., 1992. Regional
Menkes copper-transporting ATPase in NMDA receptor-mediated neuronal distribution of copper, zinc and iron in the brain in Long-Evans Cinnamon (LEC)
toxicity. Proc. Natl. Acad. Sci. U.S.A. 103, 14919–14924. rats with a new mutation causing hereditary hepatitis. Brain Res. 588, 287–290.
Schmidt, M.M., Dringen, R., 2012. GSH synthesis and metabolism. In: Gruetter, R., Suzuki-Kurasaki, M., Okabe, M., Kurasaki, M., 1997. Copper-metallothionein in the
Choi, I.Y. (Eds.), Advances in Neurobiology. Vol. 4: Neural Metabolism in Vivo. kidney of macular mice: a model for Menkes disease. J. Histochem. Cytochem.
Springer, New York, USA, pp. 1029–1050. 45, 1493–1501.
Schwerdtle, T., Hamann, I., Jahnke, G., Walter, I., Richter, C., Parsons, J.L., Dianov, G.L., Suzuki, K.T., Someya, A., Komada, Y., Ogra, Y., 2002. Roles of metallothionein in
Hartwig, A., 2007. Impact of copper on the induction and repair of oxidative copper homeostasis: responses to Cu-deficient diets in mice. J. Inorg. Biochem.
DNA damage, poly(ADP-ribosyl)ation and PARP-1 activity. Mol. Nutr. Food Res. 88, 173–182.
51, 201–210. Szerdahelyi, P., Kasa, P., 1986. Histochemical demonstration of copper in normal rat
Serpa, R.F., de Jesus, E.F., Anjos, M.J., de Oliveira, L.F., Marins, L.A., do Carmo, M.G., brain and spinal cord Evidence of localization in glial cells. Histochemistry 85,
Correa Junior, J.D., Rocha, M.S., Lopes, R.T., Martinez, A.M., 2008. Topographic 341–347.
trace-elemental analysis in the brain of Wistar rats by X-ray microfluorescence Takahashi, Y., Kako, K., Kashiwabara, S., Takehara, A., Inada, Y., Arai, H., Nakada, K.,
with synchrotron radiation. Anal. Sci. 24, 839–842. Kodama, H., Hayashi, J., Baba, T., Munekata, E., 2002. Mammalian copper
Setty, S.R., Tenza, D., Sviderskaya, E.V., Bennett, D.C., Raposo, G., Marks, M.S., 2008. chaperone Cox17p has an essential role in activation of cytochrome C oxidase
Cell-specific ATP7A transport sustains copper-dependent tyrosinase activity in and embryonic development. Mol. Cell. Biol. 22, 7614–7621.
melanosomes. Nature 454, 1142–1146. Takahashi, Y., Miyajima, H., Shirabe, S., Nagataki, S., Suenaga, A., Gitlin, J.D., 1996.
Shannon, K.M., 2011. Huntington’s disease – clinical signs, symptoms, presymp- Characterization of a nonsense mutation in the ceruloplasmin gene resulting in
tomatic diagnosis, and diagnosis. Handb. Clin. Neurol. 100, 3–13. diabetes and neurodegenerative disease. Hum. Mol. Genet. 5, 81–84.
Sharonova, I.N., Vorobjev, V.S., Haas, H.L., 2000. Interaction between copper and Tallaksen-Greene, S.J., Janiszewska, A., Benton, K., Hou, G., Dick, R., Brewer, G.J.,
zinc at GABA(A) receptors in acutely isolated cerebellar Purkinje cells of the rat. Albin, R.L., 2009. Evaluation of tetrathiomolybdate in the R6/2 model of
Br. J. Pharmacol. 130, 851–856. Huntington disease. Neurosci. Lett. 452, 60–62.
Shibata, N., Hirano, A., Kobayashi, M., Umahara, T., Kawanami, T., Asayama, K., 1995. Tanchou, V., Gas, F., Urvoas, A., Cougouluegne, F., Ruat, S., Averseng, O., Quemeneur,
Cerebellar superoxide dismutase expression in Menkes’ kinky hair disease: an E., 2004. Copper-mediated homo-dimerisation for the HAH1 metallochaperone.
immunohistochemical investigation. Acta Neuropathol. 90, 198–202. Biochem. Biophys. Res. Commun. 325, 388–394.
Shoubridge, E.A., 2001. Cytochrome c oxidase deficiency. Am. J. Med. Genet. 106, Tanzi, R.E., Petrukhin, K., Chernov, I., Pellequer, J.L., Wasco, W., Ross, B., Romano,
46–52. D.M., Parano, E., Pavone, L., Brzustowicz, L.M., Devoto, M., Peppercorn, J., Bush,
Sian-Hulsmann, J., Mandel, S., Youdim, M.B., Riederer, P., 2011. The relevance of iron A.I., Sternlieb, I., Pirastu, M., Gusella, J.F., Evgrafov, O., Penchaszadeh, G.K., Honig,
in the pathogenesis of Parkinson’s disease. J. Neurochem. 118, 939–957. B., Edelman, I.S., Soares, M.B., Scheinberg, I.H., Gilliam, T.C., 1993. The Wilson
Sian, J., Dexter, D.T., Lees, A.J., Daniel, S., Agid, Y., Javoy-Agid, F., Jenner, P., Marsden, disease gene is a copper transporting ATPase with homology to the Menkes
C.D., 1994. Alterations in glutathione levels in Parkinson’s disease and other disease gene. Nat. Genet. 5, 344–350.
neurodegenerative disorders affecting basal ganglia. Ann. Neurol. 36, 348–355. Tapia, L., Gonzalez-Aguero, M., Cisternas, M.F., Suazo, M., Cambiazo, V., Uauy, R.,
Siggs, O.M., Cruite, J.T., Du, X., Rutschmann, S., Masliah, E., Beutler, B., Oldstone, M.B., Gonzalez, M., 2004. Metallothionein is crucial for safe intracellular copper
2012. Disruption of copper homeostasis due to a mutation of Atp7a delays the storage and cell survival at normal and supra-physiological exposure levels.
onset of prion disease. Proc. Natl. Acad. Sci. U.S.A. 109, 13733–13738. Biochem. J. 378, 617–624.
Sigurdsson, E.M., Brown, D.R., Alim, M.A., Scholtzova, H., Carp, R., Meeker, H.C., Tarohda, T., Yamamoto, M., Amamo, R., 2004. Regional distribution of manganese,
Prelli, F., Frangione, B., Wisniewski, T., 2003. Copper chelation delays the onset iron, copper, and zinc in the rat brain during development. Anal. Bioanal. Chem.
of prion disease. J. Biol. Chem. 278, 46199–46202. 380, 240–246.
Singleton, W.C., McInnes, K.T., Cater, M.A., Winnall, W.R., McKirdy, R., Yu, Y., Taylor, Tchaparian, E.H., Uriu-Adams, J.Y., Keen, C.L., Mitchell, A.E., Rucker, R.B., 2000. Lysyl
P.E., Ke, B.X., Richardson, D.R., Mercer, J.F., La Fontaine, S., 2010. Role of glutar- oxidase and P-ATPase-7A expression during embryonic development in the rat.
edoxin1 and glutathione in regulating the activity of the copper-transporting Arch. Biochem. Biophys. 379, 71–77.
P-type ATPases, ATP7A and ATP7B. J. Biol. Chem. 285, 27111–27121. Telianidis, J., Hung, Y.H., Materia, S., la Fontaine, S., 2013. Role of the P-type
Smeyers-Verbeke, J., Defrise-Gussenhoven, E., Ebinger, G., Lowenthal, A., Massart, ATPases, ATP7A and ATP7B in brain copper homeostais. Front. Aging
D.L., 1974. Distribution of Cu and Zn in human brain tissue. Clin. Chim. Acta 51, Neurosci. 5, 44.
309–314. Terada, K., Nakako, T., Yang, X.L., Iida, M., Aiba, N., Minamiya, Y., Nakai, M., Sakaki, T.,
Sofroniew, M.V., Vinters, H.V., 2010. Astrocytes: biology and pathology. Acta Miura, N., Sugiyama, T., 1998. Restoration of holoceruloplasmin synthesis in LEC
Neuropathol. 119, 7–35. rat after infusion of recombinant adenovirus bearing WND cDNA. J. Biol. Chem.
Solomon, E.I., Ginsbach, J.W., Heppner, D.E., Kieber-Emmons, M.T., Kjaergaard, C.H., 273, 1815–1820.
Smeets, P.J., Tian, L., Woertink, J.S., 2011. Copper dioxygen (bio)inorganic Thackray, A.M., Knight, R., Haswell, S.J., Bujdoso, R., Brown, D.R., 2002. Metal
chemistry. Faraday Discuss. 148, 11–39, discussion 97–108. imbalance and compromised antioxidant function are early changes in prion
Speisky, H., Gomez, M., Burgos-Bravo, F., Lopez-Alarcon, C., Jullian, C., Olea-Azar, C., disease. Biochem. J. 362, 253–258.
Aliaga, M.E., 2009. Generation of superoxide radicals by copper-glutathione Thirumoorthy, N., Shyam Sunder, A., Manisenthil Kumar, K., Senthil Kumar, M.,
complexes: redox-consequences associated with their interaction with reduced Ganesh, G., Chatterjee, M., 2011. A review of metallothionein isoforms and their
glutathione. Bioorg. Med. Chem. 17, 1803–1810. role in pathophysiology. World J. Surg. Oncol. 9, 54.
Speisky, H., Gomez, M., Carrasco-Pozo, C., Pastene, E., Lopez-Alarcon, C., Olea-Azar, Thomas, B., Beal, M.F., 2007. Parkinson’s disease. Hum. Mol. Genet. 16 (Spec. No. 2)
C., 2008. Cu(I)-glutathione complex: a potential source of superoxide radicals R183–R194.
generation. Bioorg. Med. Chem. 16, 6568–6574. Thomas, K.A., Rubin, B.H., Bier, C.J., Richardson, J.S., Richardson, D.C., 1974. The
Stasser, J.P., Siluvai, G.S., Barry, A.N., Blackburn, N.J., 2007. A multinuclear copper(I) crystal structure of bovine Cu2+,Zn2+ superoxide dismutase at 5.5-A resolution.
cluster forms the dimerization interface in copper-loaded human copper chap- J. Biol. Chem. 249, 5677–5683.
erone for superoxide dismutase. Biochemistry 46, 11845–11856. Thompson, A.K., Smith, D., Gray, J., Carr, H.S., Liu, A., Winge, D.R., Hosler, J.P., 2010.
Steinebach, O.M., Wolterbeek, H.T., 1994. Role of cytosolic copper, metallothionein Mutagenic analysis of Cox11 of Rhodobacter sphaeroides: insights into the
and glutathione in copper toxicity in rat hepatoma tissue culture cells. Toxi- assembly of Cu(B) of cytochrome c oxidase. Biochemistry 49, 5651–5661.
cology 92, 75–90. Tibell, L., Hjalmarsson, K., Edlund, T., Skogman, G., Engstrom, A., Marklund, S.L.,
Steveson, T.C., Ciccotosto, G.D., Ma, X.M., Mueller, G.P., Mains, R.E., Eipper, B.A., 1987. Expression of human extracellular superoxide dismutase in Chinese
2003. Menkes protein contributes to the function of peptidylglycine alpha- hamster ovary cells and characterization of the product. Proc. Natl. Acad. Sci.
amidating monooxygenase. Endocrinology 144, 188–200. U.S.A. 84, 6634–6638.
56 I.F. Scheiber et al. / Progress in Neurobiology 116 (2014) 33–57

Tiffany-Castiglion, E., Qian, Y., 2001. Astroglia as metal depots: molecular mecha- Vidyashankar, S., Patki, P.S., 2010. Liv.52 attenuate copper induced toxicity by
nisms for metal accumulation, storage and release. Neurotoxicology 22, inhibiting glutathione depletion and increased antioxidant enzyme activity in
577–592. HepG2 cells. Food Chem. Toxicol. 48, 1863–1868.
Tiffany-Castiglioni, E., Guerri, C., Aschner, M., Matsushima, G.K., O’Callaghan, J.P., Vlachova, V., Zemkova, H., Vyklicky Jr., L., 1996. Copper modulation of NMDA
Streit, W.J., 2001. Roles of glia in developmental neurotoxicity: session VI responses in mouse and rat cultured hippocampal neurons. Eur. J. Neurosci.
summary and research needs. Neurotoxicology 22, 567–573. 8, 2257–2264.
Tiffany-Castiglioni, E., Hong, S., Qian, Y., 2011. Copper handling by astrocytes: Voglis, G., Tavernarakis, N., 2006. The role of synaptic ion channels in synaptic
insights into neurodegenerative diseases. Int. J. Dev. Neurosci. 29, 811–818. plasticity. EMBO Rep. 7, 1104–1110.
Timmers, H.J., Deinum, J., Wevers, R.A., Lenders, J.W., 2004. Congenital dopamine- Voronova, A., Kazantseva, J., Tuuling, M., Sokolova, N., Sillard, R., Palumaa, P., 2007a.
beta-hydroxylase deficiency in humans. Ann. N. Y. Acad. Sci. 1018, 520–523. Cox17, a copper chaperone for cytochrome c oxidase: expression, purification,
Tisato, F., Marzano, C., Porchia, M., Pellei, M., Santini, C., 2010. Copper in diseases and formation of mixed disulphide adducts with thiol reagents. Protein Expr.
and treatments, and copper-based anticancer strategies. Med. Res. Rev. 30, 708– Purif. 53, 138–144.
749. Voronova, A., Meyer-Klaucke, W., Meyer, T., Rompel, A., Krebs, B., Kazantseva, J.,
Tong, K.K., McArdle, H.J., 1995. Copper uptake by cultured trophoblast cells isolated Sillard, R., Palumaa, P., 2007b. Oxidative switches in functioning of mammalian
from human term placenta. Biochim. Biophys. Acta 1269, 233–236. copper chaperone Cox17. Biochem. J. 408, 139–148.
Tougu, V., Tiiman, A., Palumaa, P., 2011. Interactions of Zn(II) and Cu(II) ions with Voskoboinik, I., Brooks, H., Smith, S., Shen, P., Camakaris, J., 1998. ATP-dependent
Alzheimer’s amyloid-beta peptide Metal ion binding, contribution to fibrilliza- copper transport by the Menkes protein in membrane vesicles isolated from
tion and toxicity. Metallomics 3, 250–261. cultured Chinese hamster ovary cells. FEBS Lett. 435, 178–182.
Travaglia, A., Arena, G., Fattorusso, R., Isernia, C., La Mendola, D., Malgieri, G., Voskoboinik, I., Greenough, M., La Fontaine, S., Mercer, J.F., Camakaris, J., 2001a.
Nicoletti, V.G., Rizzarelli, E., 2011. The inorganic perspective of nerve growth Functional studies on the Wilson copper P-type ATPase and toxic milk mouse
factor: interactions of Cu2+ and Zn2+ with the N-terminus fragment of nerve mutant. Biochem. Biophys. Res. Commun. 281, 966–970.
growth factor encompassing the recognition domain of the TrkA receptor. Voskoboinik, I., Mar, J., Strausak, D., Camakaris, J., 2001b. The regulation of catalytic
Chemistry 17, 3726–3738. activity of the menkes copper-translocating P-type ATPase Role of high affinity
Travaglia, A., La Mendola, D., Magri, A., Nicoletti, V.G., Pietropaolo, A., Rizzarelli, E., copper-binding sites. J. Biol. Chem. 276, 28620–28627.
2012. Copper, BDNF and Its N-terminal domain: inorganic features and biolog- Voskoboinik, I., Strausak, D., Greenough, M., Brooks, H., Petris, M., Smith, S., Mercer,
ical perspectives. Chemistry 18, 15618–15631. J.F., Camakaris, J., 1999. Functional analysis of the N-terminal CXXC metal-
Trayhurn, P., Duncan, J.S., Wood, A.M., Beattie, J.H., 2000. Metallothionein gene binding motifs in the human Menkes copper-transporting P-type ATPase
expression and secretion in white adipose tissue. Am. J. Physiol. Regul. Integr. expressed in cultured mammalian cells. J. Biol. Chem. 274, 22008–22012.
Comp. Physiol. 279, R2329–R2335. Vulpe, C., Levinson, B., Whitney, S., Packman, S., Gitschier, J., 1993. Isolation of a
Treiber, C., Simons, A., Strauss, M., Hafner, M., Cappai, R., Bayer, T.A., Multhaup, G., candidate gene for Menkes disease and evidence that it encodes a copper-
2004. Clioquinol mediates copper uptake and counteracts copper efflux activi- transporting ATPase. Nat. Genet. 3, 7–13.
ties of the amyloid precursor protein of Alzheimer’s disease. J. Biol. Chem. 279, Wadas, T.J., Wong, E.H., Weisman, G.R., Anderson, C.J., 2007. Copper chelation
51958–51964. chemistry and its role in copper radiopharmaceuticals. Curr. Pharm. Des. 13,
Treiber, C., Pipkorn, R., Weise, C., Holland, G., Multhaup, G., 2007. Copper is required 3–16.
for prion protein-associated superoxide dismutase-I activity in Pichia pastoris. Waggoner, D.J., Drisaldi, B., Bartnikas, T.B., Casareno, R.L., Prohaska, J.R., Gitlin, J.D.,
FEBS J. 274, 1304–1311. Harris, D.A., 2000. Brain copper content and cuproenzyme activity do not vary
Trombley, P.Q., Shepherd, G.M., 1996. Differential modulation by zinc and copper of with prion protein expression level. J. Biol. Chem. 275, 7455–7458.
amino acid receptors from rat olfactory bulb neurons. J. Neurophysiol. 76, Wake, S.A., Mercer, J.F., 1985. Induction of metallothionein mRNA in rat liver and
2536–2546. kidney after copper chloride injection. Biochem. J. 228, 425–432.
Tsuji, S., Kobayashi, H., Uchida, Y., Ihara, Y., Miyatake, T., 1992. Molecular cloning of Walter, P.L., Kampkotter, A., Eckers, A., Barthel, A., Schmoll, D., Sies, H., Klotz, L.O.,
human growth inhibitory factor cDNA and its down-regulation in Alzheimer’s 2006. Modulation of FoxO signaling in human hepatoma cells by exposure to
disease. EMBO J. 11, 4843–4850. copper or zinc ions. Arch. Biochem. Biophys. 454, 107–113.
Tsukihara, T., Aoyama, H., Yamashita, E., Tomizaki, T., Yamaguchi, H., Shinzawa-Itoh, Wang, D.S., Zhu, H.L., Hong, Z., Li, J.S., 2002. Cu2+ inhibition of glycine-activated
K., Nakashima, R., Yaono, R., Yoshikawa, S., 1995. Structures of metal sites of currents in rat sacral dorsal commissural neurons. Neurosci. Lett. 328, 117–120.
oxidized bovine heart cytochrome c oxidase at 2.8 A. Science 269, 1069–1074. Wang, L.M., Becker, J.S., Wu, Q., Oliveira, M.F., Bozza, F.A., Schwager, A.L., Hoffman,
Tsukihara, T., Aoyama, H., Yamashita, E., Tomizaki, T., Yamaguchi, H., Shinzawa-Itoh, J.M., Morton, K.A., 2010. Bioimaging of copper alterations in the aging mouse
K., Nakashima, R., Yaono, R., Yoshikawa, S., 1996. The whole structure of the brain by autoradiography, laser ablation inductively coupled plasma mass
13-subunit oxidized cytochrome c oxidase at 2.8 A. Science 272, 1136–1144. spectrometry and immunohistochemistry. Metallomics 2, 348–353.
Tumer, Z., Moller, L.B., 2010. Menkes disease. Eur. J. Hum. Genet. 18, 511–518. Wang, N., Hebert, D.N., 2006. Tyrosinase maturation through the mammalian
Tümer, Z., 2013. An overview and update of ATP7A mutations leading to Menkes secretory pathway: bringing color to life. Pigment Cell Res. 19, 3–18.
disease and occipital horn syndrome. Hum. Mutat. 34, 417–429. Wang, Z.Y., Tian, C., Jing, Y.Y., Gong, H.S., Guo, Y., Shi, Q., Chen, C., Zhu, S.Y., Dong,
Uchida, Y., Gomi, F., Masumizu, T., Miura, Y., 2002. Growth inhibitory factor X.P., 2011. Knockdown of prion protein (PrP) by RNA interference weakens the
prevents neurite extension and the death of cortical neurons caused by high protective activity of wild-type PrP against copper ion and antagonizes the
oxygen exposure through hydroxyl radical scavenging. J. Biol. Chem. 277, cytotoxicity of fCJD-associated PrP mutants in cultured cells. Int. J. Mol. Med. 28,
32353–32359. 413–421.
Uchida, Y., Takio, K., Titani, K., Ihara, Y., Tomonaga, M., 1991. The growth inhibitory Warren, P.J., Earl, C.J., Thompson, R.H., 1960. The distribution of copper in human
factor that is deficient in the Alzheimer’s disease brain is a 68 amino acid brain. Brain 83, 709–717.
metallothionein-like protein. Neuron 7, 337–347. Weiser, T., Wienrich, M., 1996. The effects of copper ions on glutamate receptors in
Uriu-Adams, J.Y., Keen, C.L., 2005. Copper, oxidative stress, and human health. Mol. cultured rat cortical neurons. Brain Res. 742, 211–218.
Aspects Med. 26, 268–298. Weisiger, R.A., Fridovich, I., 1973a. Mitochondrial superoxide simutase Site of
Urso, E., Rizzello, A., Acierno, R., Lionetto, M.G., Salvato, B., Storelli, C., Maffia, M., synthesis and intramitochondrial localization. J. Biol. Chem. 248, 4793–4796.
2010. Fluorimetric analysis of copper transport mechanisms in the b104 Weisiger, R.A., Fridovich, I., 1973b. Superoxide dismutase. Organelle specificity. J.
neuroblastoma cell model: a contribution from cellular prion protein to copper Biol. Chem. 248, 3582–3592.
supplying. J. Membr. Biol. 233, 13–21. Wernimont, A.K., Yatsunyk, L.A., Rosenzweig, A.C., 2004. Binding of copper(I) by the
Urso, E., Manno, D., Serra, A., Buccolieri, A., Rizzello, A., Danieli, A., Acierno, R., Wilson disease protein and its copper chaperone. J. Biol. Chem. 279, 12269–
Salvato, B., Maffia, M., 2012. Role of the cellular prion protein in the neuron 12276.
adaptation strategy to copper deficiency. Cell. Mol. Neurobiol. 32, 989–1001. Westergard, L., Christensen, H.M., Harris, D.A., 2007. The cellular prion protein
Uversky, V.N., Li, J., Fink, A.L., 2001. Metal-triggered structural transformations, (PrP(C)): its physiological function and role in disease. Biochim. Biophys. Acta
aggregation, and fibrillation of human alpha-synuclein A possible molecular NK 1772, 629–644.
between Parkinson’s disease and heavy metal exposure. J. Biol. Chem. 276, White, A.R., Multhaup, G., Maher, F., Bellingham, S., Camakaris, J., Zheng, H., Bush,
44284–44296. A.I., Beyreuther, K., Masters, C.L., Cappai, R., 1999a. The Alzheimer’s disease
Valko, M., Leibfritz, D., Moncol, J., Cronin, M.T., Mazur, M., Telser, J., 2007. Free amyloid precursor protein modulates copper-induced toxicity and oxidative
radicals and antioxidants in normal physiological functions and human disease. stress in primary neuronal cultures. J. Neurosci. 19, 9170–9179.
Int. J. Biochem. Cell Biol. 39, 44–84. White, A.R., Reyes, R., Mercer, J.F., Camakaris, J., Zheng, H., Bush, A.I., Multhaup, G.,
van den Berghe, P.V., Folmer, D.E., Malingre, H.E., van Beurden, E., Klomp, A.E., van Beyreuther, K., Masters, C.L., Cappai, R., 1999b. Copper levels are increased in
de Sluis, B., Merkx, M., Berger, R., Klomp, L.W., 2007. Human copper transporter the cerebral cortex and liver of APP and APLP2 knockout mice. Brain Res. 842,
2 is localized in late endosomes and lysosomes and facilitates cellular copper 439–444.
uptake. Biochem. J. 407, 49–59. White, A.R., Bush, A.I., Beyreuther, K., Masters, C.L., Cappai, R., 1999c. Exacerbation
Vasak, M., Meloni, G., 2011. Chemistry and biology of mammalian metallothioneins. of copper toxicity in primary neuronal cultures depleted of cellular glutathione.
J. Biol. Inorg. Chem. 16, 1067–1078. J. Neurochem. 72, 2092–2098.
Vergun, O., Dineley, K.E., Reynolds, I.J., 2007. Ion transport and energy metabolism. White, A.R., Multhaup, G., Galatis, D., McKinstry, W.J., Parker, M.W., Pipkorn, R.,
In: Lajtha, A. (Ed.), Handbook of Neurochemistry Molecular Neurobiology. Beyreuther, K., Masters, C.L., Cappai, R., 2002. Contrasting, species-dependent
Springer Science+Buisness Media, New York, USA, pp. 429–463I. modulation of copper-mediated neurotoxicity by the Alzheimer’s disease
Vest, K.E., Leary, S.C., Winge, D.R., Cobine, P.A., 2013. Copper import into the amyloid precursor protein. J. Neurosci. 22, 365–376.
mitochondrial matrix in Saccharomyces cerevisiae is mediated by Pic2, a White, A.R., Du, T., Laughton, K.M., Volitakis, I., Sharples, R.A., Xilinas, M.E., Hoke,
mitochondrial carrier family protein. J. Biol. Chem. 288, 23884–23892. D.E., Holsinger, R.M., Evin, G., Cherny, R.A., Hill, A.F., Barnham, K.J., Li, Q.X., Bush,
I.F. Scheiber et al. / Progress in Neurobiology 116 (2014) 33–57 57

A.I., Masters, C.L., 2006. Degradation of the Alzheimer disease amyloid beta- Yang, Y., Maret, W., Vallee, B.L., 2001. Differential fluorescence labeling of cysteinyl
peptide by metal-dependent up-regulation of metalloprotease activity. J. Biol. clusters uncovers high tissue levels of thionein. Proc. Natl. Acad. Sci. U.S.A. 98,
Chem. 281, 17670–17680. 5556–5559.
Willemse, J., Van den Hamer, C.J., Prins, H.W., Jonker, P.L., 1982. Menkes’ kinky hair Yoshida, K., Furihata, K., Takeda, S., Nakamura, A., Yamamoto, K., Morita, H.,
disease I. Comparison of classical and unusual clinical and biochemical features Hiyamuta, S., Ikeda, S., Shimizu, N., Yanagisawa, N., 1995. A mutation in the
in two patients. Brain Dev. 4, 105–114. ceruloplasmin gene is associated with systemic hemosiderosis in humans. Nat.
Winge, D.R., Miklossy, K.A., 1982. Domain nature of metallothionein. J. Biol. Chem. Genet. 9, 267–272.
257, 3471–3476. Yoshimura, N., 1994. Histochemical localization of copper in various organs of
Wirdefeldt, K., Adami, H.O., Cole, P., Trichopoulos, D., Mandel, J., 2011. Epidemiology brindled mice. Pathol. Int. 44, 14–19.
and etiology of Parkinson’s disease: a review of the evidence. Eur. J. Epidemiol. Yoshimura, N., Hatayama, I., Sato, K., Nishimura, M., 1993. Complete recovery of
26 (Suppl. 1) S1–S58. cytochrome oxidase and superoxide dismutase activities in the brain of brin-
Wong, B.S., Brown, D.R., Pan, T., Whiteman, M., Liu, T., Bu, X., Li, R., Gambetti, P., dled mice receiving copper therapy. J. Intellect. Disabil. Res. 37 (Pt 6) 561–567.
Olesik, J., Rubenstein, R., Sy, M.S., 2001. Oxidative impairment in scrapie- Youdim, M.B., Grunblatt, E., Mandel, S., 2007. The copper chelator, D-penicillamine,
infected mice is associated with brain metals perturbations and altered anti- does not attenuate MPTP induced dopamine depletion in mice. J. Neural
oxidant activities. J. Neurochem. 79, 689–698. Transm. 114, 205–209.
Wu, W., Graves, L.M., Jaspers, I., Devlin, R.B., Reed, W., Samet, J.M., 1999. Activation Zatta, P., Drago, D., Bolognin, S., Sensi, S.L., 2009. Alzheimer’s disease, metal ions and
of the EGF receptor signaling pathway in human airway epithelial cells exposed metal homeostatic therapy. Trends Pharmacol. Sci. 30, 346–355.
to metals. Am. J. Physiol. 277, L924–L931. Zatta, P., Drago, D., Zambenedetti, P., Bolognin, S., Nogara, E., Peruffo, A., Cozzi, B.,
Wu, L.J., Leenders, A.G., Cooperman, S., Meyron-Holtz, E., Smith, S., Land, W., Tsai, 2008. Accumulation of copper and other metal ions, and metallothionein I/II
R.Y., Berger, U.V., Sheng, Z.H., Rouault, T.A., 2004. Expression of the iron expression in the bovine brain as a function of aging. J. Chem. Neuroanat. 36,
transporter ferroportin in synaptic vesicles and the blood–brain barrier. Brain 1–5.
Res. 1001, 108–117. Zheng, W., Monnot, A.D., 2012. Regulation of brain iron and copper homeostasis by
Xiao, G., Fan, Q., Wang, X., Zhou, B., 2013. Huntington disease arises from a brain barrier systems: implication in neurodegenerative diseases. Pharmacol.
combinatory toxicity of polyglutamine and copper binding. Proc. Natl. Acad. Ther. 133, 177–188.
Sci. U.S.A. 110, 14995–15000. Zhou, B., Gitschier, J., 1997. hCTR1: a human gene for copper uptake identified by
Xiong, K., Peoples, R.W., Montgomery, J.P., Chiang, Y., Stewart, R.R., Weight, F.F., Li, complementation in yeast. Proc. Natl. Acad. Sci. U.S.A. 94, 7481–7486.
C., 1999. Differential modulation by copper and zinc of P2X2 and P2X4 receptor Zhu, H.L., Wang, D.S., Li, J.S., 2002. Cu2+ suppresses GABA(A) receptor-mediated
function. J. Neurophysiol. 81, 2088–2094. responses in rat sacral dorsal commissural neurons. Neurosignals 11, 322–328.
Yamaguchi, Y., Heiny, M.E., Suzuki, M., Gitlin, J.D., 1996. Biochemical characteriza- Zimnicka, A.M., Maryon, E.B., Kaplan, J.H., 2007. Human copper transporter hCTR1
tion and intracellular localization of the Menkes disease protein. Proc. Natl. mediates basolateral uptake of copper into enterocytes: implications for copper
Acad. Sci. U.S.A. 93, 14030–14035. homeostasis. J. Biol. Chem. 282, 26471–26480.

You might also like