g model article in press...g model article in press jtemb-25537; no.of pages11 p. dusek et al. /...

11
ARTICLE IN PRESS G Model JTEMB-25537; No. of Pages 11 Journal of Trace Elements in Medicine and Biology xxx (2014) xxx–xxx Contents lists available at ScienceDirect Journal of Trace Elements in Medicine and Biology jou rn al homepage: www.elsevier.de/jtemb 10th NTES Symposium Review The neurotoxicity of iron, copper and manganese in Parkinson’s and Wilson’s diseases Petr Dusek a,b,, Per M. Roos c,d , Tomasz Litwin e , Susanne A. Schneider f , Trond Peder Flaten g , Jan Aaseth h a Department of Neurology and Center of Clinical Neuroscience, Charles University in Prague, 1st Faculty of Medicine and General University Hospital in Prague, Czech Republic b Institute of Neuroradiology, University Medicine Göttingen, Göttingen, Germany c Department of Neurology, Division of Clinical Neurophysiology, Oslo University Hospital, Oslo, Norway d Institute of Environmental Medicine, Karolinska Institutet, Stockholm, Sweden e 2nd Department of Neurology, Institute of Psychiatry and Neurology, Warsaw, Poland f Neurology Department, University of Kiel, Kiel, Germany g Department of Chemistry, Norwegian University of Science and Technology, Trondheim, Norway h Department of Medicine, Innlandet Hospital Trust, Kongsvinger Hospital Division, Kongsvinger, Norway a r t i c l e i n f o Article history: Received 31 January 2014 Accepted 22 May 2014 Keywords: Iron Copper Manganese Parkinson’s disease Wilson’s disease Chelating agents a b s t r a c t Impaired cellular homeostasis of metals, particularly of Cu, Fe and Mn may trigger neurodegeneration through various mechanisms, notably induction of oxidative stress, promotion of -synuclein aggrega- tion and fibril formation, activation of microglial cells leading to inflammation and impaired production of metalloproteins. In this article we review available studies concerning Fe, Cu and Mn in Parkinson’s disease and Wilson’s disease. In Parkinson’s disease local dysregulation of iron metabolism in the subs- tantia nigra (SN) seems to be related to neurodegeneration with an increase in SN iron concentration, accompanied by decreased SN Cu and ceruloplasmin concentrations and increased free Cu concentrations and decreased ferroxidase activity in the cerebrospinal fluid. Available data in Wilson’s disease suggest that substantial increases in CNS Cu concentrations persist for a long time during chelating treatment and that local accumulation of Fe in certain brain nuclei may occur during the course of the disease. Consequences for chelating treatment strategies are discussed. © 2014 Elsevier GmbH. All rights reserved. Contents Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00 Parkinson’s disease (PD) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00 Environmental exposure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00 Animal models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00 Dysregulation of brain metal homeostasis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00 Wilson’s disease (WD) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00 Local dysregulation of metal homeostasis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00 Animal models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00 Concluding remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00 Conflict of interest . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00 Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00 Corresponding author at: Department of Neurology and Center of Clinical Neuroscience, Charles University in Prague, 1st Faculty of Medicine and General University Hospital in Prague, Katerinska 30, 120 00 Praha 2, Czech Republic. Tel.: +420 224 965 528. E-mail addresses: [email protected], [email protected] (P. Dusek). http://dx.doi.org/10.1016/j.jtemb.2014.05.007 0946-672X/© 2014 Elsevier GmbH. All rights reserved.

Upload: others

Post on 22-May-2020

5 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: G Model ARTICLE IN PRESS...G Model ARTICLE IN PRESS JTEMB-25537; No.of Pages11 P. Dusek et al. / Journal of Trace Elements in Medicine and Biology xxx (2014) xxx–xxx 3 Animal models

J

1R

TW

PTa

1b

c

d

e

f

g

h

a

ARA

KICMPWC

C

H

h0

ARTICLE IN PRESSG ModelTEMB-25537; No. of Pages 11

Journal of Trace Elements in Medicine and Biology xxx (2014) xxx–xxx

Contents lists available at ScienceDirect

Journal of Trace Elements in Medicine and Biology

jou rn al homepage: www.elsev ier .de / j temb

0th NTES Symposiumeview

he neurotoxicity of iron, copper and manganese in Parkinson’s andilson’s diseases

etr Duseka,b,∗, Per M. Roosc,d, Tomasz Litwine, Susanne A. Schneider f,rond Peder Flateng, Jan Aasethh

Department of Neurology and Center of Clinical Neuroscience, Charles University in Prague,st Faculty of Medicine and General University Hospital in Prague, Czech RepublicInstitute of Neuroradiology, University Medicine Göttingen, Göttingen, GermanyDepartment of Neurology, Division of Clinical Neurophysiology, Oslo University Hospital, Oslo, NorwayInstitute of Environmental Medicine, Karolinska Institutet, Stockholm, Sweden2nd Department of Neurology, Institute of Psychiatry and Neurology, Warsaw, PolandNeurology Department, University of Kiel, Kiel, GermanyDepartment of Chemistry, Norwegian University of Science and Technology, Trondheim, NorwayDepartment of Medicine, Innlandet Hospital Trust, Kongsvinger Hospital Division, Kongsvinger, Norway

r t i c l e i n f o

rticle history:eceived 31 January 2014ccepted 22 May 2014

eywords:ronopper

a b s t r a c t

Impaired cellular homeostasis of metals, particularly of Cu, Fe and Mn may trigger neurodegenerationthrough various mechanisms, notably induction of oxidative stress, promotion of �-synuclein aggrega-tion and fibril formation, activation of microglial cells leading to inflammation and impaired productionof metalloproteins. In this article we review available studies concerning Fe, Cu and Mn in Parkinson’sdisease and Wilson’s disease. In Parkinson’s disease local dysregulation of iron metabolism in the subs-tantia nigra (SN) seems to be related to neurodegeneration with an increase in SN iron concentration,

anganesearkinson’s diseaseilson’s disease

helating agents

accompanied by decreased SN Cu and ceruloplasmin concentrations and increased free Cu concentrationsand decreased ferroxidase activity in the cerebrospinal fluid. Available data in Wilson’s disease suggestthat substantial increases in CNS Cu concentrations persist for a long time during chelating treatmentand that local accumulation of Fe in certain brain nuclei may occur during the course of the disease.Consequences for chelating treatment strategies are discussed.

© 2014 Elsevier GmbH. All rights reserved.

ontents

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00Parkinson’s disease (PD) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00

Environmental exposure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00Animal models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00Dysregulation of brain metal homeostasis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00

Wilson’s disease (WD) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00Local dysregulation of metal homeostasis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00Animal models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00

Concluding remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00

Conflict of interest . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

∗ Corresponding author at: Department of Neurology and Center of Clinical Neuroscieospital in Prague, Katerinska 30, 120 00 Praha 2, Czech Republic. Tel.: +420 224 965 528

E-mail addresses: [email protected], [email protected] (P. Dusek).

ttp://dx.doi.org/10.1016/j.jtemb.2014.05.007946-672X/© 2014 Elsevier GmbH. All rights reserved.

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00

nce, Charles University in Prague, 1st Faculty of Medicine and General University.

Page 2: G Model ARTICLE IN PRESS...G Model ARTICLE IN PRESS JTEMB-25537; No.of Pages11 P. Dusek et al. / Journal of Trace Elements in Medicine and Biology xxx (2014) xxx–xxx 3 Animal models

ING ModelJ

2 s in M

I

giOtaaimimt�tp

iiatlbpo[rbtmttrttsc[Thphlltwit[bmFr

mrlnossippl

ARTICLETEMB-25537; No. of Pages 11

P. Dusek et al. / Journal of Trace Element

ntroduction

Impaired cellular homeostasis of metals may initiate neurode-eneration through various mechanisms. Of these, oxidative stressnduced by the formation of free radicals is the best established [1].ther possible mechanisms are impaired production of metallopro-

eins [2,3], activation of microglial cells leading to inflammation [4]nd promotion of aggregation and fibril formation of �-synuclein,

highly conserved protein with unknown function. This proteins the major constituent of Lewy bodies found as intracytoplas-

atic inclusions in Parkinson’s disease (PD) neurons [5]. Recentn vitro studies have shown that mutant �-synuclein interacts with

etals and that iron (Fe) and copper (Cu) seem to aggravate theoxicity caused by the �-synuclein [6,7]. It has been suggested that-synuclein is a Cu binding protein acting as a cellular ferrireduc-

ase [8,9]. Lewy bodies contain reactive iron along with aggregatedroteins and oxidized lipid material [10].

Our knowledge regarding cellular metal homeostasis hasmproved substantially over the last years. Several proteinsnvolved in cellular import, export, trafficking and storage of Fe, Cund manganese (Mn) have been identified [11–13]. Among those,wo novel Mn carriers involved in intracellular Mn transport intoysosomes and Golgi, ATP13A2 and SLC30A10 respectively, haveeen found. Other recent findings include roles of the copper trans-orter protein 1 (CTR1) in Cu uptake, of Cu chaperones in synthesisf specific cuproproteins, and of ceruloplasmin in cellular Fe export14]. Mechanisms of metal transport across the blood–brain bar-ier (BBB) and the blood–cerebrospinal fluid barrier (BCB) haveeen thoroughly investigated [11,15,16]. While the brain seemso be protected against high plasma Fe concentrations as docu-

ented in hereditary hemochromatosis [17], this is apparently notrue for Mn and Cu. Thus, increased plasma Cu and Mn concen-rations may lead to brain deposits and CNS damage [18]. It wasecently suggested that Mn enters the CNS predominantly throughhe BCB and that high Mn concentration impairs the integrity ofhis barrier [19]. There is ongoing research on the roles of Mnpecies and their transporter molecules in brain Mn influx; Mn-itrate was suggested as the species with the highest influx rate20], and is presumably also the major Mn species in the CSF [21,22].his is in contrast to Fe and Cu which are present in CSF mainly asigh molecular weight species [18]. Mn and Fe share the trans-ort pathway using the transferrin receptor (TfR) [23,24] and itas been suggested that Fe deficit may increase brain accumu-

ation of Mn [25,26]. Some authors argue that Fe deficiency mayead to divalent metal transporter 1 (DMT1) upregulation leadingo increased Mn uptake [27,28]. There is however controversy tohat extent the DMT1 is involved in Mn transport [16,29]. Interest-

ngly, DMT1 is supposedly implicated in Mn uptake by the olfactoryract which is another possible route for brain Mn accumulation30]. In addition to Fe and Mn interdependencies, interactionsetween Fe and Cu [31] as well as between Fe and calcium [32]etabolism have been reported. Comprehensive discussions of Mn,

e and Cu transport across BBB and BCB can be found in recenteviews [11,15,16,33].

Despite these advances, it is still unclear to what extent impairedetal metabolism affects the pathophysiology of particular neu-

odegenerative disorders. The aim of this article is to review currentiterature covering association between metal dyshomeostasis andeurodegeneration, particularly studies measuring concentrationsf Fe, Cu and Mn in the cerebrospinal fluid (CSF) or brain tis-ues of patients with PD and Wilson’s disease (WD) as well astudies examining the pathophysiology of metal abnormalities. PD

s a common neurodegenerative disorder with largely unknownathophysiology, while WD is a rare disorder with well describedathophysiology involving Cu toxicosis where established che-

ating therapy exists.

PRESSedicine and Biology xxx (2014) xxx–xxx

Parkinson’s disease (PD)

Accumulation of several metals, including Fe, Cu and Mn in subs-tantia nigra (SN) has been suggested to play a causative role in thepathophysiology of PD [34]. Metal related PD hypotheses are basedon:

(1) Epidemiological studies assessing the relationship betweenenvironmental exposure to metals and the risk of PD.

(2) Clinical observations that disorders with brain metal accu-mulation such as WD, manganism or the syndromes ofneurodegeneration with brain iron accumulation (NBIA) maymanifest with parkinsonism.

(3) Animal studies examining exposure to high levels of metals andthe effects of chelation treatment.

(4) Measurements of metal concentrations and metal regulatoryprotein expression in post-mortem brain tissue samples.

(5) Measurements of metal concentrations and metal regulatoryproteins in the CSF.

Several of these empirical observations are discussed in the fol-lowing sections.

Environmental exposure

Effects of Fe and Cu exposure have been sparsely studied anda large environmental survey lent little support for an associationbetween environmental exposure to Fe or Cu and PD [35]. How-ever, combined exposure to Cu and Fe [36] or Mn and Fe [37]lasting for two to three decades have been reported to significantlyincrease the risk of PD. Also, toxic effects of Mn have been exten-sively studied [38,39] and excessive Mn intake leads to a specificcondition with motor, cognitive and psychiatric symptoms knownas manganism, with many similarities to PD. Manganism has beendescribed largely in workers in mining, welding and smeltingindustries and in battery factories, but has also been reported incommunities using drinking water with high Mn concentrations.Manganism may also arise from intravenous Mn intake, as in long-term parenteral nutrition or in methcathinon (ephedron) abuserswho use potassium permanganate for the synthesis of the drug [12].Similar clinical symptoms as in manganism have been describedin end-stage liver diseases and due to mutations in the Mn trans-porter gene SLC30A10 associated with insufficient biliary excretionof Mn [40]. Clinical symptoms, neuroimaging findings, pathologicalexamination [41] and lack of response to dopaminergic drugs [42]suggest that manganism is not caused by dysfunction of dopami-nergic neurons in the SN and that it is pathophysiologically differentfrom idiopathic PD. It has been suggested that chronic Mn expo-sure in doses insufficient to induce manganism may be toxic tothe SN rather than the globus pallidus (GP) and can increase therisk of developing idiopathic PD [43]. Several epidemiological stud-ies have found an association between Mn exposure and risk ofidiopathic PD [36,37,44–49], however other studies do not sup-port such an association [35,50,51]. Some reviews [52,53] and onemeta-analysis [54] concluded that available data argue against arelationship between Mn exposure and PD. The latter analysis usedstrict inclusion criteria leading to exclusion of all studies favor-ing the role of Mn in PD pathophysiology and the topic is stillsubject to debate. Efforts to establish a firm conclusion are partlyhampered by difficulties in differentiating between idiopathic PDand manganism-related atypical parkinsonism in large scale epi-demiological studies [55]. Several factors may increase individual

susceptibility to PD-related Mn toxicity, e.g. age, ATP13A2 muta-tions, liver dysfunction or iron deficiency [12]. Epidemiologicalstudies of the relationship between Mn exposure and PD shouldtake these factors into account [43].
Page 3: G Model ARTICLE IN PRESS...G Model ARTICLE IN PRESS JTEMB-25537; No.of Pages11 P. Dusek et al. / Journal of Trace Elements in Medicine and Biology xxx (2014) xxx–xxx 3 Animal models

ING ModelJ

s in M

A

Ftf1tcdiomifaa

MDMPanp[

droMb

teo

D

ocrocabsMFncaIftt

vplcitpd

ARTICLETEMB-25537; No. of Pages 11

P. Dusek et al. / Journal of Trace Element

nimal models

A selective dopaminergic loss has been produced by intranigrale infusion in healthy mice [56,57], and also by repeated intraperi-oneal injection of 10 mg iron dextran three times per week duringour weeks in young Wistar rats [58]. In the 1-methyl-4-phenyl-,2,3,6-tetrahydropyridine (MPTP) PD mouse model it was shownhat unilateral MPTP treatment can induce 100% increase in Fe con-entration in the ipsilateral SN after 3–6 months [59,60] along withecreased ceruloplasmin ferroxidase activity [61]. While exam-

ning the timeline of consequences of MPTP treatment, it wasbserved that Fe deposits in SN appeared 4.5 months after treat-ent. Simultaneously a nigral cell loss of 40% had occurred [62]

ndicating that Fe increase in the SN does not precede but mayollow the onset of neurodegenerative changes. Nigral cell dam-ge in the 6-hydroxydopamine (6-OHDA) animal model of PD islso accompanied by increased Fe levels in the SN [63].

Chelator pretreatment reduces degeneration of SN neurons inPTP, 6-OHDA and lactacystin-induced PD animal models [63–69].MT1 upregulation has been documented in animal PD models.ice with dysfunctional DMT1 were partially protected against

D-inducing toxins, stressing the role of this protein in Fe medi-ted degeneration [70]. Ceruloplasmin null mice develop increasedigral Fe levels accompanied by nigral cell loss, both of which wererevented by administration of the Fe binding agent deferiprone61].

Intrastriatal injection of Mn has been shown to cause loss ofopaminergic cells [71]. Long-term exposure to Mn in developingats leads to impaired motor coordination and increased signs ofxidative stress in the striatum [72]. Contrary to results in humans,n intoxication in rodents leads to a 20–75% reduction in the num-

er of SN dopaminergic cells [73,74].Taken together, published data support the notion of Fe and Mn

oxicity directed toward nigral dopaminergic cells and protectiveffects of Fe chelating agents such as deferiprone in animal modelsf PD.

ysregulation of brain metal homeostasis

The majority of autopsy studies have reported 30–100% increasef Fe content in the SN of PD patients compared to age-matchedontrols but several studies did not confirm these findings. Possibleeasons include different analytical methods [75] and pathophysi-logical heterogeneity in PD [76]. One study indicated that the Feontent in the SN was not elevated in mildly affected but only indvanced PD patients [77], suggesting that Fe increase may note involved in the early disease development. The majority oftudies using inductively coupled plasma mass spectrometry (ICP-S) or atomic absorption spectroscopy (AAS) showed increased

e concentration in PD [78–88] (Table 1). It has been argued thategative results yielded by Mössbauer spectroscopy [89,90] are aonsequence of its relatively poor sensitivity compared to othernalytical techniques [91]. Coulometry performed using FerrochemI, which was employed by Loeffler et al. [92], is known for inter-erence with various substances [93] and it cannot be excludedhat interference with some compound in PD SN contributed tohe negative results.

Measurement of metal concentrations in tissues is prone toarious biases such as instability of metal species during sam-le preparation and analysis, sampling contamination, and metal

eaching during tissue storage. In post-mortem studies tissue metalomposition may differ from the living condition. Agonal changes

ncluding ischemia, BBB disruption and cessation of cell respira-ion might alter not only the redox state and labile metal poolroportion but also total tissue metal concentration as has beenocumented for Fe during transient brain ischemia in rats [94].

PRESSedicine and Biology xxx (2014) xxx–xxx 3

The interval between death and tissue freezing or fixation is alsoan important factor influencing the extent of lytic changes thatmight also alter tissue metal composition. Since procedures usedto avoid contamination [95,96] are either not reported or not con-sidered in the majority of studies, it cannot be excluded that errorsduring sample processing contribute to disparate results [91]. Itshould be also noted that routinely used metal analytical methodssuch as AAS or ICP-MS alone cannot distinguish between differentmetal species or oxidation states. Also, differences in total con-centration of a particular metal in a tissue may reflect differencesin the content of metallo-proteins and enzymes (e.g. ceruloplas-min in the case of serum Cu [97] or glutamine synthetase in thecase of brain Mn [98]). Thus, finding of abnormal total tissue metalconcentrations does not automatically implicate abnormal metalstores or abnormal labile (free) metal concentrations. Hyphen-ated techniques, notably high performance liquid chromatography(HPLC) and capillary zone electrophoresis (CZE) online with ICP-MS, have been used for studying Mn species in body compartmentsof healthy subjects [99,100]. These studies documented that mostof the Mn-containing compounds found in liver were Mn-enzymes(for review see Ref. [101]). In order to better understand the role ofaltered metal content in neurodegenerative diseases, similar spe-ciation studies need to be performed in brain tissue and CSF ofpatients.

Using microscopic in situ imaging techniques with resolution atthe micro- or nanometer scale it was shown that Fe deposits in theSN of PD patients are predominantly associated with neuromelaninin dopaminergic neurons [86,102] and there is evidence that Fe inthese deposits is not strongly bound and thus potentially redoxactive. Suboptimal concentrations of the ferritin protein have beenreported in the SN of PD patients and also in cases of incidentalLewy body disease [78,85,103–105]. Normal ferritin upregulationassociated with increased Fe content during aging was absent inPD patients due to sustained activity of iron regulatory protein 1(IRP1) [106]. Ferritin is the main Fe storage compound in glial cellsand its decreased levels does not necessarily affect Fe metabolism inneurons [107]. Interestingly, ferritin was shown to be a componentof neuromelanin granules suggesting its possible iron-regulatoryfunction also in neurons [108]. Moreover, neuromelanin granulesoverloaded with Fe [10,81,83] exhibit higher level of redox activityin PD [109]. Taken together, these studies suggest that disrup-tion of the ferritin complex along with other disturbances of Feregulation may eventually overwhelm the storage capacity of theneuromelanin system, which may lead to an increased labile Fepool promoting the neurodegenerative process [110]. IncreasedFe3+/Fe2+ ratio in PD was also noted in two studies from the samegroup, using spectrophotometric measurements on samples fromfresh frozen brains pretreated with hydrochloric acid and pepsinwhere ascorbic acid was used to reduce Fe. They concluded that anincreased Fe3+/Fe2+ ratio is a result of ongoing oxidative processes[77,87]. This interpretation is speculative since conditions knownto increase oxidative stress such as ischemic insult or Mn overloadcause the opposite effect, that is an increased Fe2+/Fe3+ ratio [111].Moreover, this finding was disputed by other groups who did notfind any detectable Fe2+ iron in SN using Mössbauer spectroscopy[89] and micro-XANES technique [112]. The abnormally high con-centration of Fe2+ might have been an artifact of the measurementtechnique used by Sofic et al. [91,113]. This is in line with the his-tochemical distribution of Fe3+ and Fe2+ [114] and the finding thatthe major Fe storage molecules in SN, ferritin and neuromelanin,store Fe in the ferric (Fe3+) form [115].

Iron accumulation in PD is confined to the SN, while its lev-

els in the striatum and cerebral cortex are comparable to those ofcontrols or even decreased [116]. Results in the GP are inconclu-sive since both normal [77,87], increased [82,92] and decreased[78,117] concentrations have all been reported. Also CSF studies
Page 4: G Model ARTICLE IN PRESS...G Model ARTICLE IN PRESS JTEMB-25537; No.of Pages11 P. Dusek et al. / Journal of Trace Elements in Medicine and Biology xxx (2014) xxx–xxx 3 Animal models

ARTICLE IN PRESSG ModelJTEMB-25537; No. of Pages 11

4 P. Dusek et al. / Journal of Trace Elements in Medicine and Biology xxx (2014) xxx–xxx

Table 1Concentration of metals in post-mortem basal ganglia and cortical tissue in PD patients.

Reference Method N Main findings (Fe, Cu, Mn content) Comments

Ayton [61] AAS 10 ↑Fe in SN +42%↓Cu in SN −51%=Fe, Cu in frontal GM

↓Ferroxidase activity in SN

Davies [145] SRXFMPIXE 5 ↓Cu in SN (neuromelanin) -50%, LC(neuromelanin) −57%↑Fe in SN (neuromelanin) +37%=Fe, Cu in occipital GM

=Zn in SN, LC, occipital GM↓Ctr1 in SN neurons↑SOD1 in SN

Visanji [105] AAS 3 ↑Fe in SN +100%*

=Fe in Put, pons↑Ferroportin expression in SN↑Ferritin in Put↓TfR in Put

Szczerbowska-Boruchowska (2012,2004) [208], [209]

SRXFM 4 ↑Fe in SN (neurons) +60%=Cu in SN (neurons)=Fe, Cu in SN (non-neuronal tissue)

↑Ca, Zn in SN (neurons)↑Rb, Zn in SN (non-neuronal tissue)

Barapatre (2010) [210] PIXE 6 =Fe in SN (neuromelanin/non-neuronal tissue) ↓Ca in SN (neuromelanin/non-neuronal tissue)Wypijewska (2010)

[90]MS AAS 17 =Fe, Cu in SN ↑labile Fe in SN

=Labile Cu in SNOnly Fe3+ detected in SN

Popescu [117] XFM 1 ↑Fe in SN +40%↓Fe in GP, Put, CN, cGM↓Cu in SN

↓Zn in SN, Gp, Put, CN, GM

Oakley [86] WDXMA 16 ↑Fe in SN (neurons) +100%↑Fe in SN (neuropil) +70%

Reinert (2007) [211] PIXE 1 =Fe in SN (neuromelanin)*

↓Cu in SN (neuromelanin)* −190%↑Ca in SN (neuromelanin)↓Ni in SN (neuromelanin)

Morawski (2005) [212] PIXE 6 ↑Fe in SN (neuromelanin) +100%=Fe (non-significant increase) in SN(cytoplasm, glia)

Griffiths [82,88] AAS EXAFS 6 ↑Fe in SN +100%, GP +43%= Fe in Put, CN, cGM

↑Ferritin loading in SN

Galazka-Friedman [89] MS 6 =Fe in SN Only Fe3+ detected in SNKienzl (1995) [213]

Jellinger [84]EDXMA 3 ↑Fe in SN (neuromelanin)*

=Fe in SN (Lewy bodies, cytoplasm)*Only Fe3+ detected in SN

Loeffler [92] COL 14 =Fe in SN, Put, CN, frontal GM↑Fe in GP*

=Tf in SN, GP, Put, CN, frontal GM

Mann [85] ICP-MS 18 ↑Fe in SN +56% =Zn in SN=ferritin in SN

Good [81] LAMMA 3 ↑Fe in SN (neuromelanin)* +45%=Fe in SN (cytoplasm/neuropil)*

↑Al in SN (neuromelanin)

Dexter (1989)[214,78,79]

ICP-MS 27 ↑Fe in SN +35%↓Fe in GP −29%↓Cu in SN −34%=Mn in SN, Put, CN, cGM=Cu in GP, Put, CN, cGM=Fe in Put, CN, cGM

↑Zn in SN, Put, CN=Pb in SN, cGM

Hirsch [83] WDXMA 5 ↑Fe in SN (non-neuromelanin) +240%↑Fe in SN (Lewy bodies) +400%↓Fe in SN (neuromelanin) -35%

=Zn in SN↑Al in SN (Lewy bodies)

Riederer [77] AAS 13 ↑Fe in SN (only in severe degeneration) +35%↑Cu in RN=Fe in GP, Put, CN, RN, cGM=Cu in SN, GP, Put, CN, cGM

↑Zn in RN=Ca, Mg, Zn in SN, GP, Put, CN, RN, cGM↑Ferritin in SN↑Fe3+/Fe2+ ratio

Uitti [146] AAS, AES 9 ↓Cu in SN=Fe in SN, CN=Mn in SN, CN

↓Mg in CN= Ag, Al, As, B, Be, Ca, Cd, Co, Cr, K, Pb, Mn, Mo,Na, Ni, P, Se, Ti, V, W, Zn in SN, CD

Sofic [87] SPF 8 ↑Fe in SN +76%=Fe in Put, GP, cGM

↑Fe3+/Fe2+ ratio

Earle [80] XFS 11 ↑Fe in SN, BG +100%=Mn in SN, BG

=Zn in SN, BG

AAS – atomic absorption spectroscopy, AES – atomic emission spectroscopy, ICP-AES – inductively coupled plasma-atomic emission spectrometry, ICP-MS – inductivelycoupled plasma-mass spectrometry, SPF -spectrophotometry, MS – Mössbauer spectroscopy, (SR)XFM – (synchrotron radiation) X-ray fluorescence microscopy, PIXE –particle induced X-ray emission, WDXMA – wavelength dispersive X-ray microanalysis, XFS – X-ray fluorescence spectroscopy, LAMMA – Laser Microprobe Mass Analysis,E y micrn transN

hc(nhFdP

XAFS – Extended X-ray absorption fine structure, EDXMA – energy dispersive X-raucleus, RN – red nucleus, BG – basal ganglia, (c)GM – (cortical) gray matter, Tf(R) –

= number of patients.* Not statistically significant or statistical analysis not performed.

ave reported both normal [118–121] and decreased [122–124] Feoncentration in PD suggesting no overall Fe increase in the CNSTable 2). Here it is important to note that speciation analysis wasot performed in these CSF studies and neither abnormal cellular Fe

andling, Fe2+/Fe3+ dysbalance nor increased proportion of labilee can be excluded. Alteration in Fe balance can be suspected sinceecreased CSF ferroxidase activity (see below) has been found inD [125].

oanalysis, SN – substantia nigra, LC – locus coeruleus, Put – putamen, CN – caudateferrin (receptor).

The cause of metal regulatory protein dysfunction in PD SN isunknown. Several genes involved in the causation of PD are relatedto the metabolism of metals [126] and mutations in DMT1, cerulo-plasmin and transferrin were found to be weakly associated with

the risk of developing PD [127–129]. Several studies have docu-mented abnormalities in ceruloplasmin function in PD. Specifically,decreased ferroxidase activity was found in plasma [130,131],CSF [125,132] and post-mortem SN tissue [61]. Dysfunction of
Page 5: G Model ARTICLE IN PRESS...G Model ARTICLE IN PRESS JTEMB-25537; No.of Pages11 P. Dusek et al. / Journal of Trace Elements in Medicine and Biology xxx (2014) xxx–xxx 3 Animal models

ARTICLE IN PRESSG ModelJTEMB-25537; No. of Pages 11

P. Dusek et al. / Journal of Trace Elements in Medicine and Biology xxx (2014) xxx–xxx 5

Table 2Concentration of metals in the CSF of PD patients.

Reference Method N Main findings Comments

Hozumi [119] ICP-MS 20 ↑Cu +80%, Mn +70%=Fe

↑Zn +174%=Mg

Boll [132] AAS 22 ↑Cu +230% (free) Free Cu correlated with motor impairmentAlimonti [122] ICP-AESSF-ICP-MS 42 ↓Fe −20%

=Cu, Mn↓Co, Cr, Pb, Si, Sn=Al, Ba, Be, Bi, Ca, Cd, Hg, Li, Mg, Mo, Ni, Sr, Tl, V, W, Zn, Zr

Bocca [123] ICP-AESSF-ICP-MS 91 ↓Fe −37%, =Cu, Mn ↓SiForte [124] ICP-AESSF-ICP-MS 26 ↓Fe* −55%

=Cu, Mn↓Si=Al, Ca, Mg, Zn

Boll [125] AAS 49 =Cu (total) Ferroxidase activity/total Cu inversely related to PD stageJimenez-Jimenez [121] AAS 37 =Fe, Mn, Cu ↓ZnGazzaniga [118] AAS 11 =Fe, Mn, CuPall [120] AAS 31 =Fe, Mn

↑Cu↑Cu (free) confirmed byphenantroline-copper assay

AAS – atomic absorption spectroscopy, AES – atomic emission spectroscopy, ICP-AES – inductively coupled plasma-atomic emission spectrometry, ICP-MS – inductivelycN

cttmOpeawFctmccslriLa[dwhth

biutblocPcI[isleiib

oupled plasma-mass spectrometry. = number of patients.* Not statistically significant or statistical analysis not performed.

eruloplasmin may be a consequence of oxidative changes inhis enzyme [133]. Decreased ferroxidase activity was foundo be associated with earlier disease onset [134,135] and with

arkers of iron deposition in the SN of PD patients [129,136,137].ther studies have reported abnormal expression of Fe regulatoryroteins. Nigral cells behave as if they were Fe deficient in PD;xpression of proteins involved in iron uptake as DMT1, TfR2nd lactotransferrin receptor (LTFR) is upregulated [138,139],hereas expression of proteins involved in iron efflux such as

PN1 is downregulated [140]. SN areas with the most severeell degeneration showed an increased LTFR expression [139]. Athe cellular level, regulation of Fe metabolism is dependent on

itochondria where Fe is incorporated into iron-sulfur (Fe S)lusters by the action of a complex of multiple enzymes. Fe Slusters are prosthetic groups necessary for proper function ofeveral proteins such as Complexes I–III of oxidative phosphory-ation, ubiquinone and aconitase. The latter is also known as ironesponsive element binding protein 1 (IREB1) since it is involvedn the regulation of the transcription of Fe transport proteins.ow level of Fe S cluster production leads to low cytoplasmicconitase activity and increased transcription of TfR and DMT1141]. Upregulation of Fe uptake may be related to mitochondrialysfunction associated with decreased Fe S cluster synthesis,hich was documented in nigral cells in PD [142]. Alternatively, itas been show that iron regulatory protein 2 (IRP2) is degraded byhe ubiquitin–proteasome system and its dysfunction may lead toigh IRP2 levels which in turn stimulate production of DMT1 [143].

Concerning Cu, greater variation in Cu metabolic profiles haseen shown in PD compared to healthy controls using 65Cu

sotope tracer. Most importantly, higher peak plasma Cu val-es were documented suggesting more efficient absorption inhe gut [144]. Increased concentration of redox available Cu haseen reported in PD CSF [120] and its concentration was corre-

ated with motor impairment [132]. The SN contains high levelsf Cu in healthy subjects [145]. The decrease in total Cu con-entrations reported in the SN in the majority of studies inD patients [61,78,102,146] may be a reflection of loss of theuproproteins superoxide dismutase 1 (SOD1) or ceruloplasmin.n post-mortem PD SN tissue normal SOD1 activity was found102]. Moreover, decreased Cu levels are not accompanied by sim-lar decreases in zinc (Zn) levels in the majority of post-mortemtudies (Table 1) which would be expected in the case of SOD1oss. Currently, it seems more likely that decreased SN Cu lev-

ls is a consequence of reduced ceruloplasmin content whichs in line with the low ceruloplasmin ferroxidase activity foundn SN [61]. Increased labile Cu in the CSF of PD patients maye caused by its impaired incorporation into cuproproteins or

dysfunction of the blood–CSF barrier similar to the situation inWD [147].

Abnormal concentrations of Mn have not been reported inPD brains [78–80,146], but all published studies are rather old.The majority of studies also found normal Mn levels in the CSF[118,120–124].

Wilson’s disease (WD)

WD is a genetic disorder caused by mutation of the ATP7Bprotein which is mostly expressed in the liver [148]. Dysfunc-tion of ATP7B leads to disturbed Cu metabolism with a gradualaccumulation of Cu in the body and reduced synthesis of variouscuproproteins, in particular ceruloplasmin. The traditional view onthe pathophysiology of neurologic symptoms in WD holds thatoverwhelming of the liver storage capacity for Cu leads to itsspillage into plasma and CSF, leading to cellular uptake and oxida-tive damage in the CNS [149], although alternative theories havebeen proposed [150]. However, ATP7B is expressed also in the brain,namely in the choroid plexus, basal ganglia, cerebellum and cor-tex [145] and its dysfunction may lead to neurological symptomsthrough other mechanisms than just overflow of hepatic Cu. ATP7Bdysfunction in astrocytes may hypothetically lead to impairedproduction of ceruloplasmin with the possible consequence ofimpaired Fe efflux, similar to the situation in aceruloplasminemia[151]. It has been shown that ATP7B is expressed in the choroidplexus and its dysfunction may hamper the elimination of excessCu across the blood–CSF barrier [152].

The combined evidence of animal, neuroimaging, CSF and post-mortem metal studies (see below) suggests that metals other thanCu may accumulate in WD.

Local dysregulation of metal homeostasis

In patients with the neuropsychiatric form of WD, measure-ments of Cu in CSF and post-mortem brain tissue clearly show thatCu concentrations are diffusely increased in the CNS up to 850% ofnormal levels [153] (Table 3). With chelating treatment it takesup to 4 years for CSF Cu concentrations to normalize [154,155]and high levels are a marker of non-compliance [156]. ProfoundCu accumulation is still persistent in all brain regions after 10months of chelating treatment as shown by post-mortem studies

[157–159]. Another post-portem study showed abnormal brain Cudeposits even after several years of chelating treatment, and alsothat the severity of neuropathological findings was correlated withcerebral copper content [160].
Page 6: G Model ARTICLE IN PRESS...G Model ARTICLE IN PRESS JTEMB-25537; No.of Pages11 P. Dusek et al. / Journal of Trace Elements in Medicine and Biology xxx (2014) xxx–xxx 3 Animal models

ARTICLE IN PRESSG ModelJTEMB-25537; No. of Pages 11

6 P. Dusek et al. / Journal of Trace Elements in Medicine and Biology xxx (2014) xxx–xxx

Table 3Concentration of metals in the CFS and brain tissue in WD.

Reference Method N Type of tissue Main findings Comments

Kodama [153] AAS 4 (2 neu, 2 hep) CSF ↑Cu in neuro WD +280%=Cu in hepato WD

=Zn↓Cp concentrationCu decreases with therapy

Stuerenburg [155] AAS 4 (neu) CSF ↑Cu WN Cu decreases with therapyWeisner [154] AAS 5 (neu) CSF ↑Cu +330% ↓Cp concentration

Cu decreases with therapyCumings (1948) [158] COL 3 Brain ↑Cu* in CN(+160%), Put(+600%), thal(+380%), GP(+60%),

WM(290%), GM(+320%)↑Fe* in CN(+100%), Put(+20%), thal(+160%), WM(+38%),GM(+140%)=Fe in GP

Faa [159] ICP-AES 1 Brain ↑Cu** in all brain regions↓Fe** in all brain regions

↓Mg, Zn, Ca in all brainregions

Litwin [157] ICP-MS 12 (10 neu, 2 hep) Brain ↑Cu in Put(+540%), DN(+590%), pons(+540%), frontalGM(+850%)↑Fe in ND(+74%)=Fe in Put, pons, frontal GM=Mn in Put, DN, pons, frontal GM

=Zn in Put, DN, pons, frontalGM

AAS – atomic absorption spectroscopy, COL – colorimetry, ICP-AES – inductively coupled plasma-atomic emission spectrometry, ICP-MS – inductively coupled plasma-massspectrometry.SN – substantia nigra, DN – dentate nucleus, Put – putamen, CN – caudate nucleus, GM – gray matter, WM – white matter.Neu – neuropsychiatric form, hep – hepatic form, N = number of patients.

m[bilitg[GwtadhiitPiv

ptFTmlWpdtammctSo

* No statistical analysis performed.** No control group, compared to published reference values.

Increased Fe content has been reported in the striatum, grayatter, white matter [158] and dentate nucleus of the cerebellum

157] in post-mortem studies. The latter study reported normalrain Mn concentrations, but Mn was not measured in the GP which

s the structure with the greatest predisposition toward accumu-ation of Mn and other divalent metals [26]. In fact, MRI studiesndicate that elements such as Fe and Mn may also accumulate inhe brain over the course of WD. T1 hyperintensities in GP are sug-estive of Mn deposits in patients with the hepatic form of WD161,162]. Several MRI studies documented T2 hypointensities inP and other brains regions in WD patients suggestive of depositsith paramagnetic properties [163–167]. It has been suggested that

he GP changes might reflect Cu deposits [168], since Cu2+ speciesre paramagnetic. However, little is known of which form Cu iseposited in the WD CNS, and another possibility is that these T2ypointensities correspond to the paramagnetic effect of ferritin-

ron. It has been documented that the area of T2 hypointensitiesncreases during chelating treatment but it seems unlikely thathis is due to increased brain Cu concentrations [165]. Indeed, aET study with radioactively labeled Fe in human subjects showedncreased brain Fe uptake in WD patients compared to healthyolunteers [169].

Iron accumulation in WD may be a consequence of low cerulo-lasmin levels. However, it has been reported that as little as 10% ofhe normal plasma level of ceruloplasmin was sufficient for normale metabolism in a pig model [170] and in liver preparations [171].his finding was confirmed in WD patients in whom defective Feetabolism was observed only in those who had ceruloplasmin

evels less than 5% of the normal range [172]. The fact that mostD patients have more than 10% of the normal level of cerulo-

lasmin also argues against gross changes in Fe metabolism in thisisease. Although some authors reported elevated liver Fe concen-ration in WD patients [151], results of the largest published studyrgue against a clinically relevant elevation of Fe in the liver in theajority of patients [173]. This, however, pertains to peripheral Feetabolism and CNS may be more susceptible to oxidative damage

aused by decreased ceruloplasmin function. Another factor con-ributing to the risk of Fe accumulation may be mode of treatment.tandard penicillamine chelation does not remove Fe from therganism; conversely it may lead to higher tissue Fe accumulation

because it decreases the availability of Cu for ceruloplasminproduction. Zinc treatment can have opposite effects [174]. Nodifference in brain Fe concentration was however found betweenpatients treated with Zn and penicillamine [157]. Ceruloplasminalso interacts with Mn and its decreased activity can alter the neu-rotoxicity of Mn [175].

Altogether, these data suggest that profoundly increased Cu con-centrations in the CNS of WD patients persist for a long time duringchelating treatment and that local accumulation of Fe in certainbrain nuclei may occur during the course of WD.

Animal models

Established animal models for WD are the Long Evans Cinnamon(LEC) rat [176], the toxic milk mouse [177] and the Atp7b(−/−)mouse [178]. In LEC rats markedly increased concentrations ofloosely bound Cu were found in liver and plasma when total Cu con-centration exceeded the capacity of liver metallothionein synthesis[179–181]. Increased brain Fe levels have been demonstrated inLEC rats [182] and low Fe diet as well as Fe removal by phlebotomywas beneficial in terms of decreased oxidative stress markers inthe liver [183,184]. Treatment with Zn not only reduced the liverdamage through induction of metallothionein expression but alsoled to decreased liver Cu and Fe concentrations in LEC rats [185].In the toxic milk mouse, increased Cu stores were found in all tis-sues, but elevated Fe was present only in the kidneys, possibly as aresult of hemolysis [186]. The brain Fe content was even decreasedin the hippocampus [187]. Penicillamine treatment in these miceincreased labile Cu in plasma and brain tissue as a result of itsmobilization from the large tissue deposits and enhanced oxida-tive stress, which may explain the initial worsening seen in humanpatients [188].

In the Atp7b(−/−) mouse imaging of Cu metabolism usingingested 64Cu showed Cu accumulation in liver but not in brain[189]. In the same WD model decreased ferroxidase activity ofceruloplasmin was related to decreased levels of serum Fe and

transferrin saturation [190]. In contrast, Cu intoxication itself doesnot cause Fe accumulation in liver or brain of rats [191]. This find-ing suggests indirectly that decreased ceruloplasmin, rather thanCu toxicity, affects impaired Fe homeostasis in WD.
Page 7: G Model ARTICLE IN PRESS...G Model ARTICLE IN PRESS JTEMB-25537; No.of Pages11 P. Dusek et al. / Journal of Trace Elements in Medicine and Biology xxx (2014) xxx–xxx 3 Animal models

ING ModelJ

s in M

C

baognuavtaue

pmtiTttr[tdmoofbNesdmaocuva

matesrCiso(g[wibvtb[Mi

ARTICLETEMB-25537; No. of Pages 11

P. Dusek et al. / Journal of Trace Element

oncluding remarks

Current data from studies measuring metal concentrations inrain tissue and CSF do not support the idea of a general metalccumulation in the entire CNS of PD patients. Local dysregulationf Fe metabolism seems to occur in the SN, at least in a sub-roup of PD patients. The pathophysiological relation of iron toeurodegeneration and why it selectively affects the SN remainsnanswered. It has been suggested that dopamine producing cellsre particularly vulnerable to metallotoxic effects [192]. One of theulnerable enzymes in the dopamine synthesis pathway may beyrosine hydroxylase, which requires appropriate amounts of ironnd molecular oxygen as cofactors. Animal data as well as follow-p of patients with manganism gives some support to the idea ofxposure to Mn or Cu as aggravating factors in PD [193,194].

Abundant evidence suggests that decreased activity of cerulo-lasmin in CSF and SN is contributing to Fe accumulation in PD anday be relevant to its pathophysiology. It should not be forgot-

en that in contrast to Fe deposition, neurodegenerative changesn PD are not restricted to dopaminergic neurons in the SN [195].his argues against the hypothesis that Fe accumulation alone ishe trigger of neurodegeneration in PD. Nevertheless, based onhe likely presence of labile Fe in the SN, labile Cu in the CSF andesults from animal studies, we suggest along with other authors196] that a therapeutic trial with metal chelating agents is jus-ified in PD. Several points, however, need to be considered. Theose of the chelating drug should be rather low, aiming at targetedetal detoxification and redistribution of iron to tissues capable

f reutilizing it [197]. Due to the risk of increased concentrationsf labile Cu, a drug with Cu chelating capabilities may be pre-erred in future studies. The Fe chelator deferiprone is currentlyeing tested in early stage PD (Clinicaltrials.gov NCT01539837 andCT00943748). Results from the latter study document that a mod-rate dose of deferiprone (30 mg/kg/day) given for 12 months isafe and efficient in Fe removal from the SN. Moreover, using theelayed-start paradigm this study indicated a potential disease-odifying effect of deferiprone in PD [69]. Since oxidative stress

nd glutathione depletion apparently accompany the progressionf PD treatments with antioxidants have been examined. Althoughontroversial results of such treatments have been reported, e.g.sing vitamin E or vitamin C [198–200], the glutathione preser-ing agent lipoic acid appears to possess a therapeutic potentialccording to an animal study [201].

Regarding WD, little is known about accumulation of otheretals beside Cu, but dysfunction of ceruloplasmin may lead to Fe

ccumulation while liver dysfunction may lead to Mn accumula-ion in the CNS. Both metals may contribute to neurodegeneration,specially while acting synergistically with Cu. The rationale oftudying synergistic metal toxicity is further strengthened byecent studies of multiple metals with neurotoxic properties in theSF in other neurodegenerative disorders [202,203]. More stud-

es of toxic interactions among metals are needed, because suchtudies may produce invaluable information for the optimal choicef therapeutic metal chelating agents. Disodium calcium edetateCaNa2–EDTA) has been successfully used to decrease tissue man-anese levels and improve neurological symptoms in manganism204,205]. EDTA chelation has often been administered togetherith high doses of vitamin C in order to decrease oxidative stress

n several disorders. This approach, however, does not seem toe substantiated by recent clinical observations, and paradoxicallyitamin C may enhance oxidative stress [206]. Unlike penicillamine,rientine not only chelates Cu but also Fe and Mn, although the sta-

ility constants of these complexes are lower than that with Cu207]. Thus, trientine may be useful in patients with signs of Fe and

n accumulation in combination with copper. For this reason, theres a need for further studies examining free metal concentrations in

PRESSedicine and Biology xxx (2014) xxx–xxx 7

larger groups of WD patients and studies validating the use of brainMRI in determining the regional distribution of specific metals.

Conflict of interest

The authors have no conflict of interest.

Acknowledgments

This work was supported by the Czech Ministry of Education,research project PRVOUKP26/LF1/4 and by the Czech-NorwegianResearch Programme CZ09.

References

[1] Halliwell B, Gutteridge JM. Oxygen toxicity, oxygen radicals, transition metalsand disease. Biochem J 1984;219(1):1–14.

[2] Morello M, Zatta P, Zambenedetti P, Martorana A, D‘Angelo V, MelchiorriG, et al. Manganese intoxication decreases the expression of manganopro-teins in the rat basal ganglia: an immunohistochemical study. Brain Res Bull2007;74(6):406–15.

[3] Lill R, Hoffmann B, Molik S, Pierik AJ, Rietzschel N, Stehling O, et al. The role ofmitochondria in cellular iron-sulfur protein biogenesis and iron metabolism.Biochim Biophys Acta 2012;1823(9):1491–508.

[4] Zhang W, Phillips K, Wielgus AR, Liu J, Albertini A, Zucca FA, et al. Neu-romelanin activates microglia and induces degeneration of dopaminergicneurons: implications for progression of Parkinson’s disease. Neurotox Res2011;19(1):63–72.

[5] Uversky VN, Li J, Fink AL. Metal-triggered structural transformations, aggre-gation, and fibrillation of human alpha-synuclein. A possible molecularNK between Parkinson’s disease and heavy metal exposure. J Biol Chem2001;276(47):44284–96.

[6] Chew KC, Ang ET, Tai YK, Tsang F, Lo SQ, Ong E, et al. Enhanced autophagy fromchronic toxicity of iron and mutant A53T alpha-synuclein: implications forneuronal cell death in Parkinson disease. J Biol Chem 2011;286(38):33380–9.

[7] Wang X, Moualla D, Wright JA, Brown DR. Copper binding regulates intra-cellular alpha-synuclein localisation, aggregation and toxicity. J Neurochem2010;113(3):704–14.

[8] Davies P, Moualla D, Brown DR. Alpha-synuclein is a cellular ferrireductase.PLoS ONE 2011;6(1):e15814.

[9] Brown DR. alpha-Synuclein as a ferrireductase. Biochem Soc Trans2013;41(6):1513–7.

[10] Castellani RJ, Siedlak SL, Perry G, Smith MA. Sequestration of iron by Lewybodies in Parkinson’s disease. Acta Neuropathol 2000;100(2):111–4.

[11] Zheng W, Monnot AD. Regulation of brain iron and copper homeostasis bybrain barrier systems: implication in neurodegenerative diseases. PharmacolTher 2012;133(2):177–88.

[12] Tuschl K, Mills PB, Clayton PT. Manganese and the brain. Int Rev Neurobiol2013;110:277–312.

[13] Madsen E, Gitlin JD. Copper and iron disorders of the brain. Annu Rev Neurosci2007;30:317–37.

[14] Kristinsson J, Snaedal J, Torsdottir G, Johannesson T. Ceruloplasmin and ironin Alzheimer’s disease and Parkinson’s disease: a synopsis of recent studies.Neuropsychiatr Dis Treat 2012;8:515–21.

[15] Zheng W, Aschner M, Ghersi-Egea JF. Brain barrier systems: a newfrontier in metal neurotoxicological research. Toxicol Appl Pharmacol2003;192(1):1–11.

[16] Yokel RA. Manganese flux across the blood–brain barrier. Neuromol Med2009;11(4):297–310.

[17] Russo N, Edwards M, Andrews T, O‘Brien M, Bhatia KP. Hereditary haemochro-matosis is unlikely to cause movement disorders – a critical review. J Neurol2004;251(7):849–52.

[18] Nischwitz V, Berthele A, Michalke B. Speciation analysis of selected metalsand determination of their total contents in paired serum and cerebrospinalfluid samples: an approach to investigate the permeability of the humanblood–cerebrospinal fluid-barrier. Anal Chim Acta 2008;627(2):258–69.

[19] Bornhorst J, Wehe CA, Huwel S, Karst U, Galla HJ, Schwerdtle T. Impact ofmanganese on and transfer across blood–brain and blood–cerebrospinal fluidbarrier in vitro. J Biol Chem 2012;287(21):17140–51.

[20] Yokel RA, Crossgrove JS. Manganese toxicokinetics at the blood–brain barrier.Res Rep Health Eff Inst 2004;119:7–58 [discussion 59–73].

[21] Michalke B, Berthele A, Mistriotis P, Ochsenkuhn-Petropoulou M, Halbach S.Manganese species from human serum, cerebrospinal fluid analyzed by sizeexclusion chromatography-, capillary electrophoresis coupled to inductivelycoupled plasma mass spectrometry. J Trace Elem Med Biol 2007;21(Suppl.1):4–9.

[22] Michalke B, Lucio M, Berthele A, Kanawati B. Manganese speciation in paired

serum and CSF samples using SEC-DRC-ICP-MS and CE-ICP-DRC-MS. AnalBioanal Chem 2013;405(7):2301–9.

[23] Aschner M, Gannon M. Manganese (Mn) transport across the rat blood–brainbarrier: saturable and transferrin-dependent transport mechanisms. BrainRes Bull 1994;33(3):345–9.

Page 8: G Model ARTICLE IN PRESS...G Model ARTICLE IN PRESS JTEMB-25537; No.of Pages11 P. Dusek et al. / Journal of Trace Elements in Medicine and Biology xxx (2014) xxx–xxx 3 Animal models

ING ModelJ

8 s in M

ARTICLETEMB-25537; No. of Pages 11

P. Dusek et al. / Journal of Trace Element

[24] Fitsanakis VA, Zhang N, Garcia S, Aschner M. Manganese (Mn) andiron (Fe): interdependency of transport and regulation. Neurotox Res2010;18(2):124–31.

[25] Mena I, Horiuchi K, Burke K, Cotzias GC. Chronic manganese poisoning. Indi-vidual susceptibility and absorption of iron. Neurology 1969;19(10):1000–6.

[26] Erikson KM, Syversen T, Steinnes E, Aschner M. Globus pallidus: a targetbrain region for divalent metal accumulation associated with dietary irondeficiency. J Nutr Biochem 2004;15(6):335–41.

[27] Erikson KM, Syversen T, Aschner JL, Aschner M. Interactions between exces-sive manganese exposures and dietary iron-deficiency in neurodegeneration.Environ Toxicol Pharmacol 2005;19(3):415–21.

[28] Wang X, Li GJ, Zheng W. Upregulation of DMT1 expression in choroidal epithe-lia of the blood–CSF barrier following manganese exposure in vitro. Brain Res2006;1097(1):1–10.

[29] Crossgrove JS, Yokel RA. Manganese distribution across the blood–brainbarrier III. The divalent metal transporter-1 is not the major mechanismmediating brain manganese uptake. Neurotoxicology 2004;25(3):451–60.

[30] Thompson K, Molina RM, Donaghey T, Schwob JE, Brain JD, Wessling-ResnickM. Olfactory uptake of manganese requires DMT1 and is enhanced by anemia.FASEB J 2007;21(1):223–30.

[31] Monnot AD, Zheng G, Zheng W. Mechanism of copper transport at theblood–cerebrospinal fluid barrier: influence of iron deficiency in an in vitromodel. Exp Biol Med (Maywood) 2012;237(3):327–33.

[32] Gaasch JA, Geldenhuys WJ, Lockman PR, Allen DD, Van der Schyf CJ. Voltage-gated calcium channels provide an alternate route for iron uptake in neuronalcell cultures. Neurochem Res 2007;32(10):1686–93.

[33] Michalke B, Fernsebner K. New insights into manganese toxicity and specia-tion. J Trace Elem Med Biol 2014;28(2):106–16.

[34] Graham DG. Catecholamine toxicity: a proposal for the molecular pathogen-esis of manganese neurotoxicity and Parkinson’s disease. Neurotoxicology1984;5(1):83–95.

[35] Dick FD, De Palma G, Ahmadi A, Scott NW, Prescott GJ, Bennett J, et al.Environmental risk factors for Parkinson’s disease and parkinsonism: theGeoparkinson study. Occup Environ Med 2007;64(10):666–72.

[36] Gorell JM, Johnson CC, Rybicki BA, Peterson EL, Kortsha GX, Brown GG, et al.Occupational exposures to metals as risk factors for Parkinson’s disease. Neu-rology 1997;48(3):650–8.

[37] Zayed J, Ducic S, Campanella G, Panisset JC, Andre P, Masson H, et al. Envi-ronmental factors in the etiology of Parkinson’s disease. Can J Neurol Sci1990;17(3):286–91.

[38] Racette BA, Aschner M, Guilarte TR, Dydak U, Criswell SR, Zheng W.Pathophysiology of manganese-associated neurotoxicity. Neurotoxicology2012;33(4):881–6.

[39] Aschner M, Erikson KM, Herrero Hernandez E, Tjalkens R. Manganese and itsrole in Parkinson’s disease: from transport to neuropathology. Neuromol Med2009;11(4):252–66.

[40] Maeda H, Sato M, Yoshikawa A, Kimura M, Sonomura T, Terada M, et al.Brain MR imaging in patients with hepatic cirrhosis: relationship betweenhigh intensity signal in basal ganglia on T1-weighted images and elementalconcentrations in brain. Neuroradiology 1997;39(8):546–50.

[41] Perl DP, Olanow CW. The neuropathology of manganese-induced Parkinso-nism. J Neuropathol Exp Neurol 2007;66(8):675–82.

[42] Koller WC, Lyons KE, Truly W. Effect of levodopa treatment for parkinsonismin welders: a double-blind study. Neurology 2004;62(5):730–3.

[43] Lucchini RG, Martin CJ, Doney BC. From manganism to manganese-inducedparkinsonism: a conceptual model based on the evolution of exposure. Neu-romol Med 2009;11(4):311–21.

[44] Finkelstein MM, Jerrett M. A study of the relationships between Parkinson’sdisease and markers of traffic-derived and environmental manganese air pol-lution in two Canadian cities. Environ Res 2007;104(3):420–32.

[45] Willis AW, Evanoff BA, Lian M, Galarza A, Wegrzyn A, Schootman M, et al.Metal emissions and urban incident Parkinson disease: a community healthstudy of Medicare beneficiaries by using geographic information systems. AmJ Epidemiol 2010;172(12):1357–63.

[46] Gorell JM, Rybicki BA, Cole Johnson C, Peterson EL. Occupationalmetal exposures and the risk of Parkinson’s disease. Neuroepidemiology1999;18(6):303–8.

[47] Hudnell HK. Effects from environmental Mn exposures: a review ofthe evidence from non-occupational exposure studies. Neurotoxicology1999;20(2–3):379–97.

[48] Lucchini RG, Albini E, Benedetti L, Borghesi S, Coccaglio R, Malara EC, et al.High prevalence of Parkinsonian disorders associated to manganese expo-sure in the vicinities of ferroalloy industries. Am J Ind Med 2007;50(11):788–800.

[49] Racette BA, McGee-Minnich L, Moerlein SM, Mink JW, Videen TO, PerlmutterJS. Welding-related parkinsonism: clinical features, treatment, and patho-physiology. Neurology 2001;56(1):8–13.

[50] Fored CM, Fryzek JP, Brandt L, Nise G, Sjogren B, McLaughlin JK, et al.Parkinson’s disease and other basal ganglia or movement disorders ina large nationwide cohort of Swedish welders. Occup Environ Med2006;63(2):135–40.

[51] Kenborg L, Lassen CF, Hansen J, Olsen JH. Parkinson’s disease and other neu-rodegenerative disorders among welders: a Danish cohort study. Mov Disord2012;27(10):1283–9.

[52] Guilarte TR. Manganese and Parkinson’s disease: a critical review and newfindings. Environ Health Perspect 2010;118(8):1071–80.

PRESSedicine and Biology xxx (2014) xxx–xxx

[53] Jankovic J. Searching for a relationship between manganese and welding andParkinson’s disease. Neurology 2005;64(12):2021–8.

[54] Mortimer JA, Borenstein AR, Nelson LM. Associations of welding andmanganese exposure with Parkinson disease: review and meta-analysis. Neu-rology 2012;79(11):1174–80.

[55] Park RM. Neurobehavioral deficits and parkinsonism in occupations withmanganese exposure: a review of methodological issues in the epidemio-logical literature. Saf Health Work 2013;4(3):123–35.

[56] Wesemann W, Blaschke S, Solbach M, Grote C, Clement HW, Riederer P.Intranigral injected iron progressively reduces striatal dopamine metabolism.J Neural Transm Park Dis Dement Sect 1994;8(3):209–14.

[57] Sengstock GJ, Olanow CW, Menzies RA, Dunn AJ, Arendash GW. Infusion ofiron into the rat substantia nigra: nigral pathology and dose-dependent lossof striatal dopaminergic markers. J Neurosci Res 1993;35(1):67–82.

[58] Jiang H, Song N, Wang J, Ren LY, Xie JX. Peripheral iron dextran induceddegeneration of dopaminergic neurons in rat substantia nigra. NeurochemInt 2007;51(1):32–6.

[59] Mochizuki H, Imai H, Endo K, Yokomizo K, Murata Y, Hattori N, et al.Iron accumulation in the substantia nigra of 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP)-induced hemiparkinsonian monkeys. NeurosciLett 1994;168(1–2):251–3.

[60] Temlett JA, Landsberg JP, Watt F, Grime GW. Increased iron in the substan-tia nigra compacta of the MPTP-lesioned hemiparkinsonian African greenmonkey: evidence from proton microprobe elemental microanalysis. J Neu-rochem 1994;62(1):134–46.

[61] Ayton S, Lei P, Duce JA, Wong BX, Sedjahtera A, Adlard PA, et al. Ceruloplas-min dysfunction and therapeutic potential for Parkinson disease. Ann Neurol2013;73(4):554–9.

[62] He Y, Thong PS, Lee T, Leong SK, Mao BY, Dong F, et al. Dopaminergic celldeath precedes iron elevation in MPTP-injected monkeys. Free Radic Biol Med2003;35(5):540–7.

[63] Shachar DB, Kahana N, Kampel V, Warshawsky A, Youdim MB. Neu-roprotection by a novel brain permeable iron chelator, VK-28, against6-hydroxydopamine lession in rats. Neuropharmacology 2004;46(2):254–63.

[64] Berg D, Hochstrasser H. Iron metabolism in Parkinsonian syndromes. MovDisord 2006;21(9):1299–310.

[65] Ben-Shachar D, Eshel G, Finberg JP, Youdim MB. The iron chelatordesferrioxamine (Desferal) retards 6-hydroxydopamine-induced degen-eration of nigrostriatal dopamine neurons. J Neurochem 1991;56(4):1441–4.

[66] Youdim MB, Stephenson G, Ben Shachar D. Ironing iron out in Parkinson’sdisease and other neurodegenerative diseases with iron chelators: a lessonfrom 6-hydroxydopamine and iron chelators, desferal and VK-28. Ann N YAcad Sci 2004;1012:306–25.

[67] Zhu W, Xie W, Pan T, Xu P, Fridkin M, Zheng H, et al. Prevention and restorationof lactacystin-induced nigrostriatal dopamine neuron degeneration by novelbrain-permeable iron chelators. FASEB J 2007;21(14):3835–44.

[68] Kaur D, Yantiri F, Rajagopalan S, Kumar J, Mo JQ, Boonplueang R, et al.Genetic or pharmacological iron chelation prevents MPTP-induced neuro-toxicity in vivo: a novel therapy for Parkinson’s disease. Neuron 2003;37(6):899–909.

[69] Devos D, Moreau C, Devedjian JC, Kluza J, Petrault M, Laloux C, et al. Target-ing chelatable iron as a therapeutic modality in Parkinson’s disease. AntioxidRedox Signal 2014, http://dx.doi.org/10.1089/ars.2013.5593, ahead of print.

[70] Salazar J, Mena N, Hunot S, Prigent A, Alvarez-Fischer D, Arredondo M,et al. Divalent metal transporter 1 (DMT1) contributes to neurodegener-ation in animal models of Parkinson’s disease. Proc Natl Acad Sci U S A2008;105(47):18578–83.

[71] Brouillet EP, Shinobu L, McGarvey U, Hochberg F, Beal MF. Manganese injec-tion into the rat striatum produces excitotoxic lesions by impairing energymetabolism. Exp Neurol 1993;120(1):89–94.

[72] Cordova FM, Aguiar Jr AS, Peres TV, Lopes MW, Goncalves FM, Pedro DZ, et al.Manganese-exposed developing rats display motor deficits and striatal oxida-tive stress that are reversed by Trolox. Arch Toxicol 2013;87(7):1231–44.

[73] Stanwood GD, Leitch DB, Savchenko V, Wu J, Fitsanakis VA, Anderson DJ, et al.Manganese exposure is cytotoxic and alters dopaminergic and GABAergicneurons within the basal ganglia. J Neurochem 2009;110(1):378–89.

[74] Sanchez-Betancourt J, Anaya-Martinez V, Gutierrez-Valdez AL, Ordonez-Librado JL, Montiel-Flores E, Espinosa-Villanueva J, et al. Manganese mixtureinhalation is a reliable Parkinson disease model in rats. Neurotoxicology2012;33(5):1346–55.

[75] Friedman A, Galazka-Friedman J, Koziorowski D. Iron as a cause of Parkinsondisease – a myth or a well established hypothesis? Parkinsonism Relat Disord2009;15(Suppl. 3):S212–4.

[76] Sian-Hulsmann J, Mandel S, Youdim MB, Riederer P. The relevance of iron inthe pathogenesis of Parkinson’s disease. J Neurochem 2011;118(6):939–57.

[77] Riederer P, Sofic E, Rausch WD, Schmidt B, Reynolds GP, Jellinger K, et al. Tran-sition metals, ferritin, glutathione, and ascorbic acid in parkinsonian brains. JNeurochem 1989;52(2):515–20.

[78] Dexter DT, Carayon A, Javoy-Agid F, Agid Y, Wells FR, Daniel SE, et al. Alter-ations in the levels of iron, ferritin and other trace metals in Parkinson’sdisease and other neurodegenerative diseases affecting the basal ganglia.

Brain 1991;114(Pt 4):1953–75.

[79] Dexter DT, Wells FR, Agid F, Agid Y, Lees AJ, Jenner P, et al. Increasednigral iron content in postmortem parkinsonian brain. Lancet 1987;2(8569):1219–20.

Page 9: G Model ARTICLE IN PRESS...G Model ARTICLE IN PRESS JTEMB-25537; No.of Pages11 P. Dusek et al. / Journal of Trace Elements in Medicine and Biology xxx (2014) xxx–xxx 3 Animal models

ING ModelJ

s in M

[

[

[

[

[

[

[

[

[

ARTICLETEMB-25537; No. of Pages 11

P. Dusek et al. / Journal of Trace Element

[80] Earle KM. Studies on Parkinson’s disease including X-ray fluorescentspectroscopy of formalin fixed brain tissue. J Neuropathol Exp Neurol1968;27(1):1–14.

[81] Good PF, Olanow CW, Perl DP. Neuromelanin-containing neurons of the subs-tantia nigra accumulate iron and aluminum in Parkinson’s disease: a LAMMAstudy. Brain Res 1992;593(2):343–6.

[82] Griffiths PD, Crossman AR. Distribution of iron in the basal ganglia and neo-cortex in postmortem tissue in Parkinson’s disease and Alzheimer’s disease.Dementia 1993;4(2):61–5.

[83] Hirsch EC, Brandel JP, Galle P, Javoy-Agid F, Agid Y. Iron and aluminumincrease in the substantia nigra of patients with Parkinson’s disease: an X-raymicroanalysis. J Neurochem 1991;56(2):446–51.

[84] Jellinger K, Kienzl E, Rumpelmair G, Riederer P, Stachelberger H, Ben-ShacharD, et al. Iron-melanin complex in substantia nigra of parkinsonian brains: anX-ray microanalysis. J Neurochem 1992;59(3):1168–71.

[85] Mann VM, Cooper JM, Daniel SE, Srai K, Jenner P, Marsden CD, et al. Com-plex I, iron, and ferritin in Parkinson’s disease substantia nigra. Ann Neurol1994;36(6):876–81.

[86] Oakley AE, Collingwood JF, Dobson J, Love G, Perrott HR, Edwardson JA, et al.Individual dopaminergic neurons show raised iron levels in Parkinson dis-ease. Neurology 2007;68(21):1820–5.

[87] Sofic E, Riederer P, Heinsen H, Beckmann H, Reynolds GP, Hebenstreit G, et al.Increased iron (III) and total iron content in post mortem substantia nigra ofparkinsonian brain. J Neural Transm 1988;74(3):199–205.

[88] Griffiths PD, Dobson BR, Jones GR, Clarke DT. Iron in the basal gangliain Parkinson’s disease. An in vitro study using extended X-ray absorp-tion fine structure and cryo-electron microscopy. Brain 1999;122(Pt 4):667–73.

[89] Galazka-Friedman J, Bauminger ER, Friedman A, Barcikowska M, Hechel D,Nowik I. Iron in parkinsonian and control substantia nigra – a Mossbauerspectroscopy study. Mov Disord 1996;11(1):8–16.

[90] Wypijewska A, Galazka-Friedman J, Bauminger ER, Wszolek ZK, SchweitzerKJ, Dickson DW, et al. Iron and reactive oxygen species activity in parkinsoniansubstantia nigra. Parkinsonism Relat Disord 2010;16(5):329–33.

[91] Hare DJ, Gerlach M, Riederer P. Considerations for measuring iron inpost-mortem tissue of Parkinson’s disease patients. J Neural Transm2012;119(12):1515–21.

[92] Loeffler DA, Connor JR, Juneau PL, Snyder BS, Kanaley L, DeMaggio AJ, et al.Transferrin and iron in normal, Alzheimer’s disease, and Parkinson’s diseasebrain regions. J Neurochem 1995;65(2):710–6.

[93] Palmer JH, Sapienza TJ, Zink EW. Pyrazinamide interference in the FerrochemII determination of iron. Clin Chem 1988;34(7):1510–1.

[94] Danielisova V, Gottlieb M, Burda J. Iron deposition after transient forebrainischemia in rat brain. Neurochem Res 2002;27(3):237–42.

[95] Versieck J, Barbier F, Cornelis R, Hoste J. Sample contamination as a source oferror in trace-element analysis of biological samples. Talanta 1982;29(11 Pt2):973–84.

[96] Versieck J, Vanballenberghe L, De Kesel A, Van Renterghem D. Accuracy ofbiological trace-element determinations. Biol Trace Elem Res 1987;12(1):45–54.

[97] Twomey PJ, Viljoen A, House IM, Reynolds TM, Wierzbicki AS. Relationshipbetween serum copper, ceruloplasmin, and non-ceruloplasmin-bound cop-per in routine clinical practice. Clin Chem 2005;51(8):1558–9.

[98] Wedler FC, Denman RB. Glutamine synthetase: the major Mn(II) enzyme inmammalian brain. Curr Top Cell Regul 1984;24:153–69.

[99] Michalke B. Manganese speciation using capillary electrophoresis-ICP-massspectrometry. J Chromatogr A 2004;1050(1):69–76.

100] Quintana M, Klouda AD, Gondikas A, Ochsenkuhn-Petropoulou M, MichalkeB. Analysis of size characterized manganese species from liver extractsusing capillary zone electrophoresis coupled to inductively coupledplasma mass spectrometry (CZE-ICP-MS). Anal Chim Acta 2006;573–574:172–80.

101] Michalke B, Halbach S, Nischwitz V. Speciation and toxicological relevance ofmanganese in humans. J Environ Monit 2007;9(7):650–6.

102] Davies KM, Bohic S, Carmona A, Ortega R, Cottam V, Hare DJ, et al. Copperpathology in vulnerable brain regions in Parkinson’s disease. Neurobiol Aging2014;35(4):858–66.

103] Friedman A, Arosio P, Finazzi D, Koziorowski D, Galazka-Friedman J. Fer-ritin as an important player in neurodegeneration. Parkinsonism Relat Disord2011;17(6):423–30.

104] Dexter DT, Carayon A, Vidailhet M, Ruberg M, Agid F, Agid Y, et al. Decreasedferritin levels in brain in Parkinson’s disease. J Neurochem 1990;55(1):16–20.

105] Visanji NP, Collingwood JF, Finnegan ME, Tandon A, House E, Hazrati LN. Irondeficiency in parkinsonism: region-specific iron dysregulation in Parkinson’sdisease and multiple system atrophy. J Parkinsons Dis 2013;3(4):523–37.

106] Faucheux BA, Martin ME, Beaumont C, Hunot S, Hauw JJ, Agid Y, et al. Lack ofup-regulation of ferritin is associated with sustained iron regulatory protein-1 binding activity in the substantia nigra of patients with Parkinson’s disease.J Neurochem 2002;83(2):320–30.

107] Zecca L, Gallorini M, Schunemann V, Trautwein AX, Gerlach M, Riederer P, et al.Iron, neuromelanin and ferritin content in the substantia nigra of normal sub-

jects at different ages: consequences for iron storage and neurodegenerativeprocesses. J Neurochem 2001;76(6):1766–73.

108] Tribl F, Asan E, Arzberger T, Tatschner T, Langenfeld E, Meyer HE, et al. Identi-fication of L-ferritin in neuromelanin granules of the human substantia nigra:a targeted proteomics approach. Mol Cell Proteomics 2009;8(8):1832–8.

PRESSedicine and Biology xxx (2014) xxx–xxx 9

[109] Faucheux BA, Martin ME, Beaumont C, Hauw JJ, Agid Y, Hirsch EC. Neu-romelanin associated redox-active iron is increased in the substantia nigraof patients with Parkinson’s disease. J Neurochem 2003;86(5):1142–8.

[110] Zecca L, Wilms H, Geick S, Claasen JH, Brandenburg LO, Holzknecht C, et al.Human neuromelanin induces neuroinflammation and neurodegenerationin the rat substantia nigra: implications for Parkinson’s disease. Acta Neu-ropathol 2008;116(1):47–55.

[111] Fernsebner K, Zorn J, Kanawati B, Walker A, Michalke B. Manganese leads toan increase in markers of oxidative stress as well as to a shift in the ratio ofFe(II)/(III) in rat brain tissue. Metallomics 2014;6(4):921–31.

[112] Chwiej J, Adamek D, Szczerbowska-Boruchowska M, Krygowska-Wajs A,Wojcik S, Falkenberg G, et al. Investigations of differences in iron oxida-tion state inside single neurons from substantia nigra of Parkinson’s diseaseand control patients using the micro-XANES technique. J Biol Inorg Chem2007;12(2):204–11.

[113] Friedman A, Galazka-Friedman J. The history of the research of iron in parkin-sonian substantia nigra. J Neural Transm 2012;119(12):1507–10.

[114] Morris CM, Candy JM, Oakley AE, Bloxham CA, Edwardson JA. Histochem-ical distribution of non-haem iron in the human brain. Acta Anat (Basel)1992;144(3):235–57.

[115] Zecca L, Youdim MB, Riederer P, Connor JR, Crichton RR. Iron, brain ageingand neurodegenerative disorders. Nat Rev Neurosci 2004;5(11):863–73.

[116] Yu X, Du T, Song N, He Q, Shen Y, Jiang H, et al. Decreased iron levels inthe temporal cortex in postmortem human brains with Parkinson disease.Neurology 2013;80(5):492–5.

[117] Popescu BF, George MJ, Bergmann U, Garachtchenko AV, Kelly ME, McCrea RP,et al. Mapping metals in Parkinson’s and normal brain using rapid-scanningX-ray fluorescence. Phys Med Biol 2009;54(3):651–63.

[118] Gazzaniga GC, Ferraro B, Camerlingo M, Casto L, Viscardi M, Mamoli A. A casecontrol study of CSF copper, iron and manganese in Parkinson disease. Ital JNeurol Sci 1992;13(3):239–43.

[119] Hozumi I, Hasegawa T, Honda A, Ozawa K, Hayashi Y, Hashimoto K, et al.Patterns of levels of biological metals in CSF differ among neurodegenerativediseases. J Neurol Sci 2011;303(1–2):95–9.

[120] Pall HS, Williams AC, Blake DR, Lunec J, Gutteridge JM, Hall M, et al. Raisedcerebrospinal-fluid copper concentration in Parkinson’s disease. Lancet1987;2(8553):238–41.

[121] Jimenez-Jimenez FJ, Molina JA, Aguilar MV, Meseguer I, Mateos-Vega CJ,Gonzalez-Munoz MJ, et al. Cerebrospinal fluid levels of transition metals inpatients with Parkinson’s disease. J Neural Transm 1998;105(4–5):497–505.

[122] Alimonti A, Bocca B, Pino A, Ruggieri F, Forte G, Sancesario G. Elemental profileof cerebrospinal fluid in patients with Parkinson’s disease. J Trace Elem MedBiol 2007;21(4):234–41.

[123] Bocca B, Alimonti A, Senofonte O, Pino A, Violante N, Petrucci F, et al. Metalchanges in CSF and peripheral compartments of parkinsonian patients. J Neu-rol Sci 2006;248(1–2):23–30.

[124] Forte G, Bocca B, Senofonte O, Petrucci F, Brusa L, Stanzione P, et al. Traceand major elements in whole blood, serum, cerebrospinal fluid and urine ofpatients with Parkinson’s disease. J Neural Transm 2004;111(8):1031–40.

[125] Boll MC, Sotelo J, Otero E, Alcaraz-Zubeldia M, Rios C. Reduced ferroxidaseactivity in the cerebrospinal fluid from patients with Parkinson’s disease.Neurosci Lett 1999;265(3):155–8.

[126] Funke C, Schneider SA, Berg D, Kell DB. Genetics and iron in the systemsbiology of Parkinson’s disease and some related disorders. Neurochem Int2013;62(5):637–52.

[127] He Q, Du T, Yu X, Xie A, Song N, Kang Q, et al. DMT1 polymorphism and riskof Parkinson’s disease. Neurosci Lett 2011;501(3):128–31.

[128] Borie C, Gasparini F, Verpillat P, Bonnet AM, Agid Y, Hetet G, et al. FrenchParkinson’s disease genetic study g. Association study between iron-relatedgenes polymorphisms and Parkinson’s disease. J Neurol 2002;249(7):801–4.

[129] Hochstrasser H, Bauer P, Walter U, Behnke S, Spiegel J, Csoti I, et al. Cerulo-plasmin gene variations and substantia nigra hyperechogenicity in Parkinsondisease. Neurology 2004;63(10):1912–7.

[130] Torsdottir G, Sveinbjornsdottir S, Kristinsson J, Snaedal J, Johannesson T.Ceruloplasmin and superoxide dismutase (SOD1) in Parkinson’s disease: afollow-up study. J Neurol Sci 2006;241(1–2):53–8.

[131] Torsdottir G, Kristinsson J, Sveinbjornsdottir S, Snaedal J, Johannesson T. Cop-per, ceruloplasmin, superoxide dismutase and iron parameters in Parkinson’sdisease. Pharmacol Toxicol 1999;85(5):239–43.

[132] Boll MC, Alcaraz-Zubeldia M, Montes S, Rios C. Free copper, ferroxidaseand SOD1 activities, lipid peroxidation and NO(x) content in the CSF. A dif-ferent marker profile in four neurodegenerative diseases. Neurochem Res2008;33(9):1717–23.

[133] Olivieri S, Conti A, Iannaccone S, Cannistraci CV, Campanella A, BarbarigaM, et al. Ceruloplasmin oxidation, a feature of Parkinson’s disease CSF,inhibits ferroxidase activity and promotes cellular iron retention. J Neurosci2011;31(50):18568–77.

[134] Bharucha KJ, Friedman JK, Vincent AS, Ross ED. Lower serum ceruloplasminlevels correlate with younger age of onset in Parkinson’s disease. J Neurol2008;255(12):1957–62.

[135] Jin L, Wang J, Jin H, Fei G, Zhang Y, Chen W, et al. Nigral iron deposi-

tion occurs across motor phenotypes of Parkinson’s disease. Eur J Neurol2012;19(7):969–76.

[136] Jin L, Wang J, Zhao L, Jin H, Fei G, Zhang Y, et al. Decreased serum ceruloplas-min levels characteristically aggravate nigral iron deposition in Parkinson’sdisease. Brain 2011;134(Pt 1):50–8.

Page 10: G Model ARTICLE IN PRESS...G Model ARTICLE IN PRESS JTEMB-25537; No.of Pages11 P. Dusek et al. / Journal of Trace Elements in Medicine and Biology xxx (2014) xxx–xxx 3 Animal models

ING ModelJ

1 s in M

ARTICLETEMB-25537; No. of Pages 11

0 P. Dusek et al. / Journal of Trace Element

[137] Martinez-Hernandez R, Montes S, Higuera-Calleja J, Yescas P, Boll MC, Diaz-Ruiz A, et al. Plasma ceruloplasmin ferroxidase activity correlates with thenigral sonographic area in Parkinson’s disease patients: a pilot study. Neu-rochem Res 2011;36(11):2111–5.

[138] Ke Y, Qian ZM. Brain iron metabolism: neurobiology and neurochemistry.Prog Neurobiol 2007;83(3):149–73.

[139] Faucheux BA, Nillesse N, Damier P, Spik G, Mouatt-Prigent A, Pierce A, et al.Expression of lactoferrin receptors is increased in the mesencephalon ofpatients with Parkinson disease. Proc Natl Acad Sci U S A 1995;92(21):9603–7.

[140] Wang J, Jiang H, Xie JX. Ferroportin1 and hephaestin are involved inthe nigral iron accumulation of 6-OHDA-lesioned rats. Eur J Neurosci2007;25(9):2766–72.

[141] Crichton RR, Dexter DT, Ward RJ. Brain iron metabolism and its perturbationin neurological diseases. J Neural Transm 2011;118(3):301–14.

[142] Gille G, Reichmann H. Iron-dependent functions of mitochondria-relation toneurodegeneration. J Neural Transm 2011;118(3):349–59.

[143] Li XP, Xie WJ, Zhang Z, Kansara S, Jankovic J, Le WD. A mechanistic studyof proteasome inhibition-induced iron misregulation in dopamine neurondegeneration. Neurosignals 2012;20(4):223–36.

[144] Larner F, Sampson B, Rehkamper M, Weiss DJ, Dainty JR, O‘Riordan S, et al. Highprecision isotope measurements reveal poor control of copper metabolism inparkinsonism. Metallomics 2013;5(2):125–32.

[145] Davies KM, Hare DJ, Cottam V, Chen N, Hilgers L, Halliday G, et al. Local-ization of copper and copper transporters in the human brain. Metallomics2013;5(1):43–51.

[146] Uitti RJ, Rajput AH, Rozdilsky B, Bickis M, Wollin T, Yuen WK. Regional metalconcentrations in Parkinson’s disease, other chronic neurological diseases,and control brains. Can J Neurol Sci 1989;16(3):310–4.

[147] Choi BS, Zheng W. Copper transport to the brain by the blood–brain barrierand blood–CSF barrier. Brain Res 2009;1248:14–21.

[148] Tanzi RE, Petrukhin K, Chernov I, Pellequer JL, Wasco W, Ross B, et al. TheWilson disease gene is a copper transporting ATPase with homology to theMenkes disease gene. Nat Genet 1993;5(4):344–50.

[149] Telianidis J, Hung YH, Materia S, Fontaine SL. Role of the P-Type ATPases,ATP7A and ATP7B in brain copper homeostasis. Front Aging Neurosci2013;5:44.

[150] Zischka H, Lichtmannegger J, Schmitt S, Jagemann N, Schulz S, Wartini D, et al.Liver mitochondrial membrane crosslinking and destruction in a rat model ofWilson disease. J Clin Invest 2011;121(4):1508–18.

[151] Hayashi H, Yano M, Fujita Y, Wakusawa S. Compound overload of copperand iron in patients with Wilson’s disease. Med Mol Morphol 2006;39(3):121–6.

[152] Barnes N, Tsivkovskii R, Tsivkovskaia N, Lutsenko S. The copper-transportingATPases, Menkes and Wilson disease proteins, have distinct roles in adult anddeveloping cerebellum. J Biol Chem 2005;280(10):9640–5.

[153] Kodama H, Okabe I, Yanagisawa M, Nomiyama H, Nomiyama K, Nose O, et al.Does CSF copper level in Wilson disease reflect copper accumulation in thebrain? Pediatr Neurol 1988;4(1):35–7.

[154] Weisner B, Hartard C, Dieu C. CSF copper concentration: a new parameter fordiagnosis and monitoring therapy of Wilson’s disease with cerebral manifes-tation. J Neurol Sci 1987;79(1–2):229–37.

[155] Stuerenburg HJ, copper concentrations CSF. blood–brain barrier function, andcoeruloplasmin synthesis during the treatment of Wilson’s disease. J NeuralTransm 2000;107(3):321–9.

[156] Stuerenburg HJ, Eggers C. Early detection of non-compliance in Wilson’s dis-ease by consecutive copper determination in cerebrospinal fluid. J NeurolNeurosurg Psychiatry 2000;69(5):701–2.

[157] Litwin T, Gromadzka G, Szpak GM, Jablonka-Salach K, Bulska E,Czlonkowska A. Brain metal accumulation in Wilson’s disease. J Neurol Sci2013;329(1–2):55–8.

[158] Cumings JN. The copper and iron content of brain and liver in the normal andin hepato-lenticular degeneration. Brain 1948;71(Pt. 4):410–5.

[159] Faa G, Lisci M, Caria MP, Ambu R, Sciot R, Nurchi VM, et al. Brain copper, iron,magnesium, zinc, calcium, sulfur and phosphorus storage in Wilson’s disease.J Trace Elem Med Biol 2001;15(2–3):155–60.

[160] Horoupian DS, Sternlieb I, Scheinberg IH. Neuropathological findings inpenicillamine-treated patients with Wilson’s disease. Clin Neuropathol1988;7(2):62–7.

[161] Sinha S, Taly AB, Ravishankar S, Prashanth LK, Venugopal KS, Arunodaya GR,et al. Wilson’s disease: cranial MRI observations and clinical correlation. Neu-roradiology 2006;48(9):613–21.

[162] Sinha S, Taly AB, Prashanth LK, Ravishankar S, Arunodaya GR, Vasudev MK.Sequential MRI changes in Wilson’s disease with de-coppering therapy: astudy of 50 patients. Br J Radiol 2007;80(957):744–9.

[163] Skowronska M, Litwin T, Dziezyc K, Wierzchowska A, Czlonkowska A. Doesbrain degeneration in Wilson disease involve not only copper but also ironaccumulation? Neurol Neurochir Pol 2013;47(6):542–6.

[164] Lee JH, Yang TI, Cho M, Yoon KT, Baik SK, Han YH. Widespread cerebral cor-tical mineralization in Wilson’s disease detected by susceptibility-weightedimaging. J Neurol Sci 2012;313(1–2):54–6.

[165] Engelbrecht V, Schlaug G, Hefter H, Kahn T, Modder U. MRI of the brain

in Wilson disease: T2 signal loss under therapy. J Comput Assist Tomogr1995;19(4):635–8.

[166] Brugieres P, Combes C, Ricolfi F, Degos JD, Poirier J, Gaston A. Atypical MR pre-sentation of Wilson disease: a possible consequence of paramagnetic effectof copper? Neuroradiology 1992;34(3):222–4.

PRESSedicine and Biology xxx (2014) xxx–xxx

[167] Mironov A. Decreased signal intensity of the putamen and the caudate nucleusin Wilson disease of the brain. Neuroradiology 1993;35(2):166.

[168] Fritzsch D, Reiss-Zimmermann M, Trampel R, Turner R, Hoffmann KT, SchaferA. Seven-Tesla magnetic resonance imaging in Wilson disease using quantita-tive susceptibility mapping for measurement of copper accumulation. InvestRadiol 2014;49(5):299–306.

[169] Bruehlmeier M, Leenders KL, Vontobel P, Calonder C, Antonini A, Weindl A.Increased cerebral iron uptake in Wilson’s disease: a 52Fe-citrate PET study. JNucl Med 2000;41(5):781–7.

[170] Ragan HA, Nacht S, Lee GR, Bishop CR, Cartwright GE. Effect of ceruloplasminon plasma iron in copper-deficient swine. Am J Physiol 1969;217(5):1320–3.

[171] Osaki S, Johnson DA, Frieden E. The mobilization of iron from the perfusedmammalian liver by a serum copper enzyme, ferroxidase I. J Biol Chem1971;246(9):3018–23.

[172] Roeser HP, Lee GR, Nacht S, Cartwright GE. The role of ceruloplasmin in ironmetabolism. J Clin Invest 1970;49(12):2408–17.

[173] Pfeiffenberger J, Gotthardt DN, Herrmann T, Seessle J, Merle U, SchirmacherP, et al. Iron metabolism and the role of HFE gene polymorphisms in Wilsondisease. Liver Int 2012;32(1):165–70.

[174] Medici V, Di Leo V, Lamboglia F, Bowlus CL, Tseng SC, D‘Inca R, et al. Effectof penicillamine and zinc on iron metabolism in Wilson’s disease. Scand JGastroenterol 2007;42(12):1495–500.

[175] Jursa T, Smith DR. Ceruloplasmin alters the tissue disposition and neu-rotoxicity of manganese, but not its loading onto transferrin. Toxicol Sci2009;107(1):182–93.

[176] Wu J, Forbes JR, Chen HS, Cox DW. The LEC rat has a deletion in the coppertransporting ATPase gene homologous to the Wilson disease gene. Nat Genet1994;7(4):541–5.

[177] Theophilos MB, Cox DW, Mercer JF. The toxic milk mouse is a murine modelof Wilson disease. Hum Mol Genet 1996;5(10):1619–24.

[178] Buiakova OI, Xu J, Lutsenko S, Zeitlin S, Das K, Das S, et al. Null mutationof the murine ATP7B (Wilson disease) gene results in intracellular copperaccumulation and late-onset hepatic nodular transformation. Hum Mol Genet1999;8(9):1665–71.

[179] Suzuki KT. Disordered copper metabolism in LEC rats, an animal model ofWilson disease: roles of metallothionein. Res Commun Mol Pathol Pharmacol1995;89(2):221–40.

[180] Koizumi M, Fujii J, Suzuki K, Inoue T, Inoue T, Gutteridge JM, et al. A markedincrease in free copper levels in the plasma and liver of LEC rats: an ani-mal model for Wilson disease and liver cancer. Free Radic Res 1998;28(5):441–50.

[181] Suzuki KT, Kanno S, Misawa S, Sumi Y. Changes in hepatic copper distribu-tion leading to hepatitis in LEC rats. Res Commun Chem Pathol Pharmacol1993;82(2):217–24.

[182] Kim JM, Ko SB, Kwon SJ, Kim HJ, Han MK, Kim DW, et al. Ferrous and ferriciron accumulates in the brain of aged Long-Evans Cinnamon rats, an animalmodel of Wilson’s disease. Neurosci Lett 2005;382(1–2):143–7.

[183] Sugawara N, Ohta T, Lai YR, Sugawara C, Yuasa M, Nakamura M, et al. Irondepletion prevents adenine nucleotide decomposition and an increase of xan-thine oxidase activity in the liver of the Long Evans Cinnamon (LEC) rat, ananimal model of Wilson’s disease. Life Sci 1999;65(13):1423–31.

[184] Hayashi H, Wakusawa S, Yano M. Iron removal by phlebotomy for the pro-phylaxis of fulminant hepatitis in a Wilson disease model of Long-EvansCinnamon Rats. Hepatol Res 2006;35(4):276–80.

[185] Santon A, Irato P, Medici V, D‘Inca R, Albergoni V, Sturniolo GC. Effect andpossible role of Zn treatment in LEC rats, an animal model of Wilson’s disease.Biochim Biophys Acta 2003;1637(1):91–7.

[186] Allen KJ, Buck NE, Cheah DM, Gazeas S, Bhathal P, Mercer JF. Chronologicalchanges in tissue copper, zinc and iron in the toxic milk mouse and effects ofcopper loading. Biometals 2006;19(5):555–64.

[187] Przybylkowski A, Gromadzka G, Wawer A, Bulska E, Jablonka-Salach K,Grygorowicz T, et al. Neurochemical and behavioral characteristics oftoxic milk mice: an animal model of Wilson’s disease. Neurochem Res2013;38(10):2037–45.

[188] Chen DB, Feng L, Lin XP, Zhang W, Li FR, Liang XL, et al. Penicillamine increasesfree copper and enhances oxidative stress in the brain of toxic milk mice. PLoSONE 2012;7(5):e37709.

[189] Peng F, Lutsenko S, Sun X, Muzik O. Imaging copper metabolism imbalancein Atp7b(−/−) knockout mouse model of Wilson’s disease with PET-CT andorally administered 64CuCl2. Mol Imaging Biol 2012;14(5):600–7.

[190] Merle U, Tuma S, Herrmann T, Muntean V, Volkmann M, Gehrke SG,et al. Evidence for a critical role of ceruloplasmin oxidase activity in ironmetabolism of Wilson disease gene knockout mice. J Gastroenterol Hepatol2010;25(6):1144–50.

[191] Pal A, Vasishta R, Prasad R. Hepatic and hippocampus iron status is not alteredin response to increased serum ceruloplasmin and serum “free” copper inWistar rat model for non-Wilsonian brain copper toxicosis. Biol Trace ElemRes 2013;154(3):403–11.

[192] Chakraborty S, Aschner M. Altered manganese homeostasis: implications forBLI-3-dependent dopaminergic neurodegeneration and SKN-1 protection inC. elegans. J Trace Elem Med Biol 2012;26(2–3):183–7.

[193] Kim Y. High signal intensities on T1-weighted MRI as a biomarker of exposureto manganese. Ind Health 2004;42(2):111–5.

[194] Sikk K, Haldre S, Aquilonius SM, Asser A, Paris M, Roose A, et al. Manganese-induced parkinsonism in methcathinone abusers: bio-markers of exposureand follow-up. Eur J Neurol 2013;20(6):915–20.

Page 11: G Model ARTICLE IN PRESS...G Model ARTICLE IN PRESS JTEMB-25537; No.of Pages11 P. Dusek et al. / Journal of Trace Elements in Medicine and Biology xxx (2014) xxx–xxx 3 Animal models

ING ModelJ

s in M

[

[

[

[

[

[

[

[

[

[

[

ARTICLETEMB-25537; No. of Pages 11

P. Dusek et al. / Journal of Trace Element

195] Jellinger KA. Neuropathology of sporadic Parkinson’s disease: evaluation andchanges of concepts. Mov Disord 2012;27(1):8–30.

196] Ayton S, Lei P. Nigral iron elevation is an invariable feature of Parkin-son’s disease and is a sufficient cause of neurodegeneration. Biomed Res Int2014;2014:581256.

197] Cabantchik ZI, Munnich A, Youdim MB, Devos D. Regional siderosis: a newchallenge for iron chelation therapy. Front Pharmacol 2013;4:167.

198] Beal MF, Oakes D, Shoulson I, Henchcliffe C, Galpern WR, Haas R, et al. Arandomized clinical trial of high-dosage coenzyme Q10 in early Parkinsondisease: no evidence of benefit. JAMA Neurol 2014;71(5):543–52.

199] Persson T, Popescu BO, Cedazo-Minguez A. Oxidative stress in Alzheimer’sdisease: why did antioxidant therapy fail? Oxid Med Cell Longev2014;2014:427318.

200] Shoulson I. DATATOP: a decade of neuroprotective inquiry, Parkinson StudyGroup. Deprenyl and tocopherol antioxidative therapy of parkinsonism. AnnNeurol 1998;44(3 Suppl. 1):S160–6.

201] Zaitone SA, Abo-Elmatty DM, Shaalan AA. Acetyl-l-carnitine and alpha-lipoicacid affect rotenone-induced damage in nigral dopaminergic neurons of ratbrain, implication for Parkinson’s disease therapy. Pharmacol Biochem Behav2012;100(3):347–60.

202] Squitti R. Metals in Alzheimer’s disease: a systemic perspective. Front Biosci(Landmark Ed) 2012;17:451–72.

203] Roos PM, Vesterberg O, Syversen T, Flaten TP, Nordberg M. Metal concentra-tions in cerebrospinal fluid and blood plasma from patients with amyotrophiclateral sclerosis. Biol Trace Elem Res 2013;151(2):159–70.

204] Stamelou M, Tuschl K, Chong WK, Burroughs AK, Mills PB, BhatiaKP, et al. Dystonia with brain manganese accumulation resulting from

SLC30A10 mutations: a new treatable disorder. Mov Disord 2012;27(10):1317–22.

205] Herrero Hernandez E, Discalzi G, Valentini C, Venturi F, Chio A, CarmellinoC, et al. Follow-up of patients affected by manganese-induced Parkinsonismafter treatment with CaNa2EDTA. Neurotoxicology 2006;27(3):333–9.

PRESSedicine and Biology xxx (2014) xxx–xxx 11

[206] Hininger I, Waters R, Osman M, Garrel C, Fernholz K, Roussel AM, et al. Acuteprooxidant effects of vitamin C in EDTA chelation therapy and long-termantioxidant benefits of therapy. Free Radic Biol Med 2005;38(12):1565–70.

[207] Crisponi G, Nurchi VM, Fanni D, Gerosa C, Nemolato S, Faa G. Copper-related diseases: from chemistry to molecular pathology. Coord Chem Rev2010;254(7–8):876–89.

[208] Szczerbowska-Boruchowska M, Krygowska-Wajs A, Adamek D. Elemen-tal micro-imaging and quantification of human substantia nigra usingsynchrotron radiation based x-ray fluorescence–in relation to Parkinson’sdisease. J Phys Condens Matter 2012;24(24):244104.

[209] Szczerbowska-Boruchowska M, Lankosz M, Ostachowicz J, Adamek D,Krygowska-Wajs A, Tomik B, Szczudlik A, Simionovici A, Bohic S. Topographicand quantitative microanalysis of human central nervous system tissue usingsynchrotron radiation. X-Ray Spectrometry 2004;33(1):3–11.

[210] Barapatre N, Morawski M, Butz T, Reinert T. Trace element mapping in Parkin-sonian brain by quantitative ion beam microscopy. Nuclear Instruments &Methods in Physics Research Section B-Beam Interactions with Materials andAtoms 2010;268(11–12): 2156–2159.

[211] Reinert T, Fiedler A, Morawski M, Arendt T. High resolution quantitative ele-ment mapping of neuromelanin-containing neurons. Nuclear Instruments &Methods in Physics Research Section B-Beam Interactions with Materials andAtoms 2007;260(1):227–30.

[212] Morawski M, Meinecke C, Reinert T, Dorffel AC, Riederer P, Arendt T, ButzT. Determination of trace elements in the human substantia nigra. NuclearInstruments & Methods in Physics Research Section B-Beam Interactions withMaterials and Atoms 2005;231 (1–4): 224–228.

[213] Kienzl E, Puchinger L, Jellinger K, Linert W, Stachelberger H, Jameson RF. 1995.

The role of transition metals in the pathogenesis of Parkinson’s disease. JNeurol Sci 1995: 134 (Suppl) 69–78.

[214] Dexter DT, Wells FR, Lees AJ, Agid F, Agid Y, Jenner P, Marsden CD. Increasednigral iron content and alterations in other metal ions occurring in brain inParkinson’s disease. J Neurochem 1989;52(6):1830–6.