You are on page 1of 11

View Article Online / Journal Homepage / Table of Contents for this issue

Metallomics Dynamic Article Links

Cite this: Metallomics, 2011, 3, 239–249

www.rsc.org/metallomics CRITICAL REVIEW


The potential application of iron chelators for the treatment of
neurodegenerative diseasesw
Robert C. Hider, Sourav Roy, Yong Min Ma, Xiao Le Kong and Jane Preston
Published on 22 February 2011 on http://pubs.rsc.org | doi:10.1039/C0MT00087F

Received 24th November 2010, Accepted 25th January 2011


DOI: 10.1039/c0mt00087f

Many forms of neurodegenerative disease, for instance Alzheimer’s disease, Parkinson’s disease,
Friedreich’s ataxia, Hallervorden Spatz syndrome and macular degeneration, are associated with
elevated levels of redox active metals in the brain and eye. A logical therapeutic approach
Downloaded by Duke University on 11 January 2013

therefore, is to remove the toxic levels of these metals, copper and iron in particular, by selective
chelation. The increased number of iron-selective chelators now available for clinical use has
enhanced interest in this type of therapy. This review summarises the recent developments in the
design of chelators for treatment of neurodegenerative disease, identifies some of the essential
properties for such molecules and suggests some future strategies.

1. Introduction against them through a set of antioxidants and detoxifying


enzymes that include superoxide dismutase, catalase and
The main agent of risk in most neurodegenerative disorders is glutathione. When this imbalance occurs, oxidatively modified
age and this may be directly linked to oxidative stress (lipid molecules (lipids, proteins, nucleotides) accumulate in the
peroxidation, protein oxidation, DNA and RNA oxidation), cellular compartment causing dysfunction.1 In the case of very
which increases in the brain with age and plays a central role in sensitive cells such as neurons, the lack of control of defence
the pathogenic mechanisms of neurodegeneration. systems may eventually lead to cell death. Under physiological
Oxidative stress may be defined as an imbalance between the conditions, free radicals are byproducts of cellular oxygen
production of free radicals and the ability of the cell to defend metabolism, with superoxide (O2 ), hydroxyl (OH ) and
nitric oxide (NO ) species being prevalent, while hydrogen
King’s College London, Institute of Pharmaceutical Science, peroxide (H2O2) and peroxynitrite (ONOO), not radicals
Franklin-Wilkins Building, 150 Stamford Street, London, themselves, contributing to the cellular redox state and
UK SE1 9NH
w This article is published as part of a themed issue on Metals in eventually produce radicals through various chemical
Neurodegenerative Diseases, Guest Edited by David Brown. reactions. Mitochondrial oxidative metabolism phospholipid

Professor Bob Hider is Sourav is a PhD in Pharma-


emeritus professor of medicinal ceutical Sciences (King’s
chemistry at King’s College College London) with an
London, where he has worked MSc in Drug Discovery
since 1987. Prior to this he (University of London) and a
was a lecturer in biological B-Pharm (Bachelor of
chemistry at Essex University. Pharmacy, Rajiv Gandhi
He has worked with University of Health Sciences,
siderophore-based iron uptake Bangalore, India). His PhD
processes in microorganisms research involved the evalua-
and the absorption of iron by tion of blood-brain barrier
mammalian cells. His work on permeability and neuro-
membrane structure and protective properties of novel
transport mechanisms has led iron chelators intended for the
Robert C. Hider to the development of novel Sourav Roy prophylaxis of Alzheimer’s
oral iron chelators for the disease. Through this multi-
treatment of iron overload. As a result, N-alkyl-3-hydroxypyridin- disciplinary work two lead iron chelators possessing neuro-
4-ones have been identified as possessing potential for clinical protective properties were identified.
application. Deferiprone is now used worldwide for the treatment
of iron overload.

This journal is c The Royal Society of Chemistry 2011 Metallomics, 2011, 3, 239–249 239
View Article Online

metabolism, proteolytic pathways and metal ions are potential neurons. A key pathological hallmark of AD is the presence of
sources of intracellular free radicals. senile plaques constituted by a highly dense core formed of a
The brain is at risk from oxidative damage because of the mixture of 39–43 residue polypeptides derived from Amyloid
following specific characteristics: Precursor Protein (APP) that accumulate in the cortical inter-
 high oxygen consumption (20% of the total body basal O2 stitium and cerebrovasculature in a characteristic manner. One
consumption); such peptide, amyloid b-peptide 1–42 (Ab1–42) is a minor
 critically high levels of both iron and ascorbate soluble species but possesses a fibrillogenic activity that
 relatively low levels of antioxidant protective agents renders it central to the pathogenesis and to be particularly toxic
 tendency to accumulate metals with age. to cells in the early stage of the peptide aggregation process.2,3
Oxidative damage induced by the redox activity of a target There is strong evidence of a relationship between oxidative
protein, which interacts with free radicals and metal ions, has stress and Ab cortical deposits and this characteristic is likely
been found as a typical hallmark in many neurodegenerative to derive mostly from its ability to bind metals and, as a
Published on 22 February 2011 on http://pubs.rsc.org | doi:10.1039/C0MT00087F

disorders such as Alzheimer’s disease (AD) and Parkinson’s consequence, to mediate redox reactions.4,5 In fact, Ab1–42 is
disease (PD). a metallo-binding peptide with binding sites for both Cu(II)
and Fe(III)6 (Fig. 1). Metal homeostasis is altered during AD,
and as a consequence, metals are reported to accumulate in the
1.1 Oxidative stress and Alzheimer’s disease
neuropil with concentrations that are 3–5-fold increased
Oxidative stress is believed to play a major role in the compared to agematched controls.7
Downloaded by Duke University on 11 January 2013

dysfunction and degeneration occurring in AD, as one of the The proposed metal coordinating site can in principle bind
earliest events that takes place in the cytoplasm of vulnerable either iron (Fig. 2A) or copper (Fig. 2B) and provide sites
which will readily redox cycle as judged by the relative
selectivity of histidine, tyrosine and aspartate for the ion pairs
Yong Min Ma was born in iron(III)/iron(II) and copper(II)/copper(I) (Table 1). The
1976 in Zhejiang, P.R. China. proposed binding site for iron (Fig. 2A and 3) is similar to
He studied chemistry at that of transferrin (2Y, 1D, 1H) except that a second histidine
Zhejiang University, where he
obtained his bachelors degree
in 1998 and master’s degree in
2001 under the direction of
Prof. Yong Min Zhang. He
then moved to London for
PhD study in 2002 in
Prof. Hider’s group at King’s
College London. After
completion of his PhD in
2005, he continued to work as
Yong Min Ma a postdoctoral fellow in Prof.
Hider’s group. His research Fig. 1 Sequence of the Ab1–42 peptide, which is a proteolytic
interests are centred on the design of iron chelators for the fragment of the amyloid precursor protein. The key amino acids for
treatment of iron overload diseases. iron(III) coordination are His 6, Asp 7, Tyr 10 and His 14.

Xiaole Kong was originally Dr Preston is Senior Lecturer


educated as a computer in the Institute of Pharma-
engineer. After successfully ceutical Science at King’s
assembling an automatic College London. She com-
fluorescent titration system pleted her PhD and post-
for Professor Hider’s lab, he doctoral training at St Thomas’
became more interested in Hospital Medical School and
scientific research. He Brown University, USA and
commenced a one year joined King’s in 1995. She
training course and then has been both Vice-Chairman
undertook study for a PhD— (2004) and Organizing
‘‘Spectrophotometric Deter- Chairman (2006) of the
mination of Stability Con- Gordon Research Conference
stants of Iron Chelators’’. He ‘Barriers of the CNS’ USA
Xiao Le Kong was awarded the ‘‘2009 Jane Preston and founding member of the
Tadion–Rideal Prize for International Brain Barriers
Molecular Science’’. His specialty is the determination of Society. Research interests are in drug delivery to CNS,
affinity constants. cerebrospinal fluid (CSF) dynamics, and brain-barrier function
in ageing, stroke, brain tumour challenge using in vivo and BBB
cell culture models.

240 Metallomics, 2011, 3, 239–249 This journal is c The Royal Society of Chemistry 2011
View Article Online

hypothesis that oxidative stress generates the cascade of


events, which are responsible of the preferential degeneration
of melanised dopaminergic neurons in the substantia nigra
pars compacta in PD. Parkinsonian brains possess elevated
levels of iron in microglia, astrocytes, oligodendrocytes and
dopaminergic neurones of substantia nigra pars compacta.11–15
Iron can in principle mediate the generation of hydroxyl
radicals by interaction with a-synuclein, a 140-amino-acid
presynaptic protein that accumulates intracellularly in the
form of fibrillar aggregates in neurons and sometimes also in
glia cells. The synucleins (a-, b- and g-) are highly expressed
proteins in the nervous tissue. Two missense mutations in the
Published on 22 February 2011 on http://pubs.rsc.org | doi:10.1039/C0MT00087F

a-synuclein gene (A53T and A30P) are responsible for rare


forms of familial PD. The accumulation of fibrillar forms of
Fig. 2 Schematic representation of iron(III) and copper(II) bound to
the protein is one of the decisive events occurring in the
the metal coordination site of Ab1–42. It is proposed that the
pathogenesis of PD.16–18 The toxicity in part may be generated
octahedral field of iron is completed by the coordination of two water
molecules.
by oxidative damage, induced by interactions between
a-synuclein fibrils and iron.19,20
Downloaded by Duke University on 11 January 2013

Table 1 Ligand selectivity of iron, copper and zinc cations 1.3 Therapeutic strategies

Iron(III) O (anionic) A wide range of drugs with diverse mechanisms of action has
been investigated. Although some of these approaches have
Iron(II) O (anionic), N (aliphatic), N (imidazole), S demonstrated potential, when studied in cellular and animal
Copper(II) O (anionic), N (aliphatic), N (imidazole), S
Copper(I) N (imidazole), S models of acute or chronic neurodegeneration, only a few have
Zinc(II) O (anionic), N (aliphatic), N (imidazole), S provided convincing clinical results. Among the relevant
therapeutic strategies, it is noteworthy to mention the use of
antioxidants, excitotoxicity modulators, the inhibition of the
expression of amyloidogenic protein, inhibition of the release
of amyloidogenic peptide and the inhibition of Ab-amyloid
aggregation. Oxidative stress, protein aggregation and redox
active metal ions are also considered to be promising pharma-
cological targets. The apparently critical involvement of
metals, particularly iron and copper, in both oxidative stress
and protein aggregation processes therefore renders chelation
therapy a sensible strategy.21 The key feature of a suitable
chelating agent would be the ability firstly to scavenge the free
redox active metal present in excess in the brain and to form a
nontoxic metal complex, which is then excreted.

2. Chelation-based therapy
One of the dominant properties of any therapeutic chelator is
Fig. 3 Possible conformation of iron bound to the metal coordina- metal selectivity, typically a high selectivity being required, for
tion site of Ab1–42. Light blue, carbon; red, oxygen; dark blue,
instance in the treatment of iron overload associated with
nitrogen; white, hydrogen; yellow, iron. The structure was energy
b-thalassaemia. In this situation, ligands with a high selectivity
minimised by HyperChems 8.0 Amber.
for iron over copper and zinc are essential, as chelation
therapy is maintained for life. Unfortunately, with the
replaces one of the tyrosines (Y, 1D, 2H). This substitution
proposed treatment of neurodegenerative diseases by chelation
would render the site with a higher affinity for iron(II) than
therapy, the identity of the putative toxic metal is not always
that of transferrin, thereby facilitating redox cycling. The
firmly established. With AD, for instance, iron, copper and zinc
proposed octahedral site has two sites occupied by water,
have all been associated with the progression of the disease. In
which by virtue of their lability would render the iron
contrast, with PD, iron is clearly the major target. Although
accessible to both reducing and oxidising agents.8
there are clear guidelines for the design of iron-selective
chelating agents,22 this is not the situation with copper.
1.2 Oxidative stress and Parkinson’s disease
With the necessity of ready permeation of the blood–brain
Post-mortem studies in PD brains indicate that a wide range of barrier (BBB), the size of useful chelators should probably be
molecules undergo oxidative damage, including lipids, limited to less than 300Da, thereby excluding hexadentate
proteins and DNA.9,10 In fact, significant neurochemical, ligands and further limiting the potential for the design of
physical, histochemical and biochemical evidence confirm the selective copper(II) chelators. As PD is clearly associated with

This journal is c The Royal Society of Chemistry 2011 Metallomics, 2011, 3, 239–249 241
View Article Online

Table 2 Metal affinity constants for selected ligands (Martell & Smith, 1974–1989)

Log cumulative stability constant


Ligand Fe(III) Al(III) Ga(II) Cu(II) Zn(II) Fe(II) pFeIII
DFO (1) 30.6 25.0 27.6 14.1 11.1 7.2 26
EDTA (2) 25.1 16.5 21.0 18.8 16.5 14.3 23.4
N,N-dimethyl-2,3-dihydroxy-benzamide (DMB) (3) 40.2 24.9 13.5 17.5 15
Acetohydroxamic acid (4) 28.3 21.5 7.9 9.6 8.5 13
3-hydroxypyridin-4-one (deferiprone) (5) 37.2 35.8 32.6 21.7 13.5 12.1 20.5b
8-hydroxyquinoline (6) 37.7 40.5 22.9 15.8 22.2 20.6
a
pFeIII = log[Fe3+] when [Fe3+]total = 106 M and [ligand]total = 105 M at pH 7.4. b This value agrees with that recently reported by Nurchi
et al.25 but is different to the widely quoted figure of 19.021 which was based on measurements of logb3 by Martell & Smith of 35.8. The value of
cumulative affinity constant has recently been reported to be higher, namely 36.725 and 37.1.26 Hexadentate ligands: DFO (1), EDTA (2); bidentate
Published on 22 February 2011 on http://pubs.rsc.org | doi:10.1039/C0MT00087F

ligands: N,N-dimethyl-2,3-dihydroxy-benzamide (DMB) (3), acetohydroxamic acid (4), deferiprone (5), 8-hydroxyquinoline (6).

an abnormal distribution of iron in the brain and as most


strong iron(III) chelators also possess an appreciable affinity
for copper(II) (Table 2) an initial search for an iron(III) chelator
Downloaded by Duke University on 11 January 2013

would appear to be an appropriate approach.


Iron(III)-selective chelators favour oxygen atoms as ligands,
notably hydroxamates and catecholates (Table 2). Most
tribasic cations, for instance aluminium(III) and gallium(III),
are not essential for living cells and thus iron(III) is a practical
target for ‘clinical chelator’ design. An additional advantage of
high affinity iron(III) chelators is that, under aerobic
conditions, they will chelate iron(II) and facilitate autoxidation
to iron(III).23
Ligands can be structurally classified according to the
number of donor atoms that each molecule possesses. When
a ligand contains two, three or six donor atoms, it is termed
bidentate, tridentate or hexadentate (Fig. 4). For biological
conditions, the pFe value is a more useful parameter than the
conventional stability constant for assessment of the ligand’s
affinity for the metal.22,24 For clinically useful iron scavengers,
a pFe3+ value X 20 (Table 2) is considered to be essential.
Molecular size is also a critical factor, as it influences the
penetration of both the wall of the gastrointestinal tract27 and
the BBB.28 Lipinski et al.29 suggest a guideline of 500. This
molecular-weight limit provides a considerable restriction on
the choice of chelator and may effectively exclude hexadentate
ligands from consideration. Fig. 5 General structure of iron(III) chelators. Hexadentate: desfer-
Bidentate and tridentate ligands, by virtue of their much rioxamine (DFO, 1), ethylenediaminetetraacetic acid (EDTA, 2);
bidentate: N,N-demethyl-2-3,-dihydroxybenzamide (DMB, 3),
lower molecular weights, are predicted to possess higher
acetohydroxamic acid (4), hydroxypyridinone (5, R1 = R2 = CH3,
absorption efficiencies. The fraction of the absorbed dose for
deferiprone; 5a, R1 = R2 = C2H5, CP94; 5b, R1 = nC4H9, R2 =
a range of bidentate hydroxypyridin-4-ones (5) has, for CH3, CP24), 8-hydroxyquinoline (6), Clioquinol (7), VK20 (8),
instance, been found to fall between 50 and 70%.30 M30 (9), CP241 (10), CP242 (11).

Hydroxypyridinones have also been demonstrated to


penetrate the BBB.31,32
Another important feature for chelation therapy of neuro-
degeneration is the ability not only to scavenge redox active
metal but to remove this excess metal from the brain. One of
the successes of deferiprone in the treatment of transfusion—
induced iron overloaded thalassaemia patients is its ability to
remove iron from the heart.33 It is able to achieve this by
forming a non charged water soluble iron(III) complex
Fig. 4 Schematic representation of chelate ring formation in iron- possessing a molecular weight less than 500 and therefore
ligand complexes. capable of permeating membranes by non facilitated diffusion.

242 Metallomics, 2011, 3, 239–249 This journal is c The Royal Society of Chemistry 2011
View Article Online

6-OHDA at very low doses. Moreover, this study has shown


that the mechanism of action of VK-28 is more likely to be
related to iron chelation properties than to any direct inter-
ference with 6-OHDA, since intranigral or intraventricular
6-OHDA initiates an increase in total iron in the substantia
nigra and striatum at the sites of neurodegeneration, in
monkeys, rats and mice.46

3. Chelator design
The toxicity associated with chelators originates from a
number of factors, such as inhibition of metalloenzymes, lack
Published on 22 February 2011 on http://pubs.rsc.org | doi:10.1039/C0MT00087F

of metal selectivity and redox cycling of complexes (Fig. 8).

3.1 Redox properties of iron complex


Fig. 6 Schematic representation of the penetration of deferiprone If all the coordinating ligands are charged oxygen atoms, as is
[LH]o through the BBB. The bidentate ligand scavenges loosely bound the case with 3-hydroxypyridinones, then both iron and
iron forming the 3 : 1 complex [inset] which also carries zero net charge. copper complexes will favour the most oxidised form of the
Downloaded by Duke University on 11 January 2013

Diffusion through the BBB as the iron complex leads to iron excretion. metal ion, namely iron(III) and copper(II) (Table 1). This in
turn prevents redox cycling, as the corresponding iron(II) and
copper(I) complexes are unstable. This is reflected in the
extremely low redox potential of iron(III) hydroxypyridinone
complexes, 610 mV.47 In contrast, if some of the coordinating
ligands are nitrogen atoms, as is the case with 8-hydroxy-
quinolines, then redox cycling becomes possible with the
concomitant generation of free radicals. This is reflected in
the higher affinity of 8-hydroxyquinolines for iron(II) and
copper(I) (Table 2), when compared with those of hydroxy-
pyridinones and the higher redox potential of the iron/
hydroxyquinoline complex of 150 mV.48

Fig. 7 Brain iron in the ferrocene-loaded rat. 3,5,5-trimethylhexanoyl 3.2 Inhibition of enzyme activity
ferrocene is used to load the rat brain. Brain regions: 1, cerebellum; 2,
In general, iron chelators do not directly inhibit haem iron-
cerebral cortex; 3, hippocampus; 4, brain stem; 5, striatum; 6,
substantia nigra. Structure of CP94 given in Fig. 5.
containing enzymes due to the inaccessibility of porphyrin
bound iron to chelating agents. In contrast, in many non-haem
iron-containing enzymes, such as lipoxygenase, the aromatic
This property indicates that hydroxypyridinones may also be hydroxylase family and ribonucleotide reductase are susceptible
able to remove excess iron from the brain (Fig. 6). Indeed to chelator-induced inhibition.49 Generally, hydrophobic
studies with hydroxypyridinones in the ferrocene—loaded rat chelators inhibit lipoxygenases, therefore the introduction of
demonstrate the ability of such compounds to remove excess hydrophilic characteristics into a chelator tend to minimise
iron from the brain (Fig. 7). It is this ability to remove excess such inhibitory potential,50 particularly if their introduction
labile iron from cells and mitochondria that has led to the also induces steric interference of the chelation process
success of deferiprone therapy in patients suffering from at the enzyme active site.51 By careful modification of
Friedreich’s ataxia34,35 and NBIA (Neurodegeneration with physicochemical properties, iron chelators can therefore be
brain iron accumulation).36 designed which exert minimal inhibitory influence on many
A range of 8-hydroxyquinoline analogues have also been metalloenzymes.52,53
investigated for their potential to treat neurodegeneration.
Clioquinol (7) is a small lipophilic molecule which has been
demonstrated to have beneficial effects in both AD and PD
models37–40 and in clinical studies.41,42 8-Hydroxyquinolines,
like hydroxypyridinones, form neutral 3 : 1 iron complexes.
Unfortunately, halogenated hydroxy-quinolines possess
neurotoxic side effects.43,44 These side effects may be avoided
by the use of nonhalogenated analogues, for instance the brain
permeable VK-28 (8) and M30 (9).45 A study centred on rats,
with 6-OHDA—induced striatal dopaminergic lesions, has
shown that, when injected either intraventricularly (1 mg in 5 ml)
or intraperitoneally (1 or 5 mg kg1 day1 for 10 and 7 days, Fig. 8 Redox cycling activity of iron complexes. The reducing agent
respectively), VK-28 is able to provide neuroprotection against could be vitamin C or reduced coenzymes, for instance NADPH.

This journal is c The Royal Society of Chemistry 2011 Metallomics, 2011, 3, 239–249 243
View Article Online

Table 3 Inhibition of Tyrosine Hydroxylase introduction of a single hydroxyl function dramatically


decreases the ability of the hydroxypyridinone to penetrate
the BBB (Fig. 9). There is a practical limit to increasing the log
P value in order to enhance BBB permeability due to increasingly
efficient first past extraction by the liver, which in turn leads to
decreasing oral bioavailability. In an attempt to optimise BBB
% inhibition penetration while not enhancing liver first pass extraction,
of tyrosine a range of fluorinated hydroxypyridinones have been
Hydroxypyridinone R R1 hydroxylasea synthesised56 (Table 4). Brain content was determined by
CP94 (5a) — — 43 in situ perfusion of guinea pig brains,57,58 having made
CP142 H CH3 18.5 allowance for the vascular space by use of 3H-mannitol
CP132 CH3 CH3 2.5 (Fig. 10). There was no correlation with logPoctanol values,
CP133 CH3 CH2CH(CH3)2 15.5
Published on 22 February 2011 on http://pubs.rsc.org | doi:10.1039/C0MT00087F

CP135 CH3 CH2Ph 3.0 but four were found to cross the BBB more efficiently than
CP137 CH2Ph CH2CH(CH3)2 46.5 deferiprone, the most effective being CP241 2-fluoro-3-hydroxy-
a
[chelator] = 10 mM. 1-n-propylpyridin-4-one (10).59

3.4 Ability to inhibit prooxidant insults


One series of enzymes which requires careful attention in all
Hydroxypyridinones have been demonstrated to protect
Downloaded by Duke University on 11 January 2013

studies centred on neurodegeneration are those associated


with the synthesis and metabolism of dopamine and cultured cortical neurones against prooxidant damage induced
5-hydroxytryptamine, namely tyrosine hydroxylases, trypto- by iron(III) NTA, hydrogen peroxide and Ab1–40.55,60
phan hydroxylases and catechol-O-methyl transferase. In a Deferiprone is capable of providing protection to these cells at
study with ferrocene-loaded rats the hydroxypyridinone CP94 10 mM (Fig. 11). With the amyloid peptide Ab1–40, the
(5a) was found to decrease both striatal dopamine and prooxidant damage almost certainly results from the peptide
5-hydroxytryptamine levels.54 Significantly the more hydro- scavenging iron from the medium and providing a binding site
philic analogue deferiprone (5) failed to change these levels.54 which is capable of facile redox cycling (Fig. 3). These results
A series of hydroxypyridinones have subsequently been are summarised for 20 mM deferiprone in Fig. 12 where it is
designed which possess a very low ability to inhibit tyrosine clear that deferiprone confers protection against amyloid b—
hydroxylase55 (Table 3). induced neurotoxicity. A similar finding is observed with the

3.3 Ability to cross the blood brain barrier Table 4 Structure of fluorinated hydroxypyridinones

By virtue of their small molecular size and noncharged nature,


hydroxypyridinones readily cross the blood brain barrier
(BBB),31,32 there being a clear relationship with the log Poctanol
values of N-alkyl derivatives (Fig. 9).31 However the

R1 R2 R5 R6 logP oct.
Deferiprone CH3 CH3 H H 0.77
CP227 Et F H H 0.58
CP233 Et H F H 0.89
CP241 n-Prop F H H 0.07
CP242 n-Prop H F H 0.46
CP243 iso-Prop H F H 0.92
CP228 H CF3 H H 0.08
CP246 H F CONHMe H 1.00
CP221 H COCF3 H H 0.97

Fig. 9 Relationship between blood-brain barrier permeability


(Log PS) of a range of hydroxypyridinones with Log Poctanol in adult
wistar rats. The vascular perfusion time was 1 min values are expressed
as means  SD (n = 3)31 CP20 (deferiprone) (5), R1 = R2 = CH3;
CP21, R1 = CH3CH2, R2 = CH3; CP94 (5a); CP24 (5b); CP25, R1 =
CH3(CH2)4CH2, R2 = CH3; CP29, R1 = CH3(CH2)3CH2, R2 = CH3;
CP41, R1 = (CH2)3OH, R2 = CH3; CP102, R1 = (CH2)2OH, Fig. 10 Brain content of fluorinated hydroxypyridinones. Values are
R2 = Et; CP107, R1 = (CH2)3OH, R2 = Et. mean  S.E.M where n = 6–12.59

244 Metallomics, 2011, 3, 239–249 This journal is c The Royal Society of Chemistry 2011
View Article Online
Published on 22 February 2011 on http://pubs.rsc.org | doi:10.1039/C0MT00087F
Downloaded by Duke University on 11 January 2013

Fig. 11 Neurotoxicity studies in primary cortical neurones exposed to human Ab1–40.60 Cytotoxicity was assessed by morphometric analysis of
cell viability by HOE 33324 (blue) and/or cell death by propidium iodide (red) using an IN1000 Cell Analyser. Cultures were treated with Ab1–40
(3 or 20 mM, t = 24 h) (a–c) and/or deferiprone (10, 30 or 100 mM) (d–l).

induced by Ab peptide is completely reversed with chelator


concentrations of 30 mM.

4. Targeting chelators to the brain


As indicated in the previous section, there is a clear limit to the
maximum chelator trans-BBB flux achievable by non
facilitated diffusion. To increase this flux above this limit it
will be necessary to facilitate chelator transfer. Two
approaches have been reported, the use of nanoparticles and
the conjugation of chelators to sugars, leading to facilitated
transport across the BBB.

4.1 Nanoparticle bound iron chelators


Fig. 12 Deferiprone (10, 30 or 100 mM, t = 24 h) confers protection
against amyloid b-induced neurotoxicity (#p o 0.001 vs. control).60 There are a number of nanoparticle approaches available for
[Ab] = 20mM. The results shown are the mean  SEM of three the transfer of chelators into brain.61 Polymeric nanoparticles
independent experiments in quadruplicate. possess high drug loading capacities and have been targeted to
the BBB62,63; nanogels, which consist of net works of
fluoro substituted hydroxypyridinones CP241 (10) and CP242 cross-linked polymers can also be used to deliver low
(11) where both compounds offer protection against 10 mM molecular weight drugs64; polymeric nanomicelles possess a
Ab1–40 peptide (Fig. 13).59 In both cases the neurotoxicity core-shell structure, with a hydrophobic cove within a shell of

This journal is c The Royal Society of Chemistry 2011 Metallomics, 2011, 3, 239–249 245
View Article Online
Published on 22 February 2011 on http://pubs.rsc.org | doi:10.1039/C0MT00087F
Downloaded by Duke University on 11 January 2013

Fig. 13 Neurotoxicity studies in primary cortical neurones exposed to human Ab1–40. Cytotoxicity was assessed by formazan blue production,
which monitors the viability of mitochondria. The cells were incubated for 24 h in the presence of Ab1–40 (10 mM) and various chelators (30 mM).
All values are mean  SEM from three independent experiments, in quadruplicate.

hydrophilic polymers, the core is capable of accommodating


up to 25% w/w of drugs65; and polymeric nanoliposomes,
which are vesicular structures composed of one or more
lipid bilayers surrounding an internal aqueous compartment.66
These nanoparticles can be targeted to the BBB using a
variety of address systems,67 including melanotransferrin
receptor,68 transferrin receptor69 and apolipoprotein
receptors.70
Clioquinol (7)—encapsulated in poly(butylcyanoacrylate)
nanoparticles (PBCA) has been prepared as a vector for the
invivo brain imaging of b-amyloid plaques. These preparations
cross the BBB with high efficiency.71 Further more,
125
I-labelled clioquinol incorporated into nanoparticle has
been utilised for invivo biodistribution studies in mice.72 PBCA
nanoparticles have been used to deliver a range of drugs to the
CNS,73 for instance doxorubicin.74 The drug is loaded during
the initial polymerisation process and the resulting particles Fig. 14 An illustration of a typical structure of a polymer-based
are coated with polysorbate 80. Following intravenous nanoparticle with externally attached chelators. The hydroxypyridi-
administration the surface of the particles become further nones are covalently attached via amide links to the polymer matrix.
coated with absorbed plasma proteins, most prominently
apolipoprotein E. It has been suggested that these naturally
coated particles are mistaken for low-density lipoprotein structure (Fig. 14).77 An interesting feature of these nano-
particles by the cerebral endothelium and internalised by the particles is that they are reported to have the potential of
LDL uptake system.75,76 Nanoparticles have also been recrossing the BBB and entering the blood stream. Thus if they
prepared with chelators, desferrioxamine (1) and hydroxy- bind apolipoprotein A-I in the CSF they will be susceptible to
pyridinones (5) covalently attached to the outside of the efflux from the brain.78

246 Metallomics, 2011, 3, 239–249 This journal is c The Royal Society of Chemistry 2011
View Article Online

An important limitation to nanoparticle delivery of 3-glucopyranosyloxy derivative of deferiprone (15) failed to


chelators to the brain is they would have to be administered cross the guinea pig BBB.89
by a parenteral route, oral presentation would not be possible.

4.2 Facilitated transport of chelator-sugar conjugates


In principle pendant glucose molecules can be attached to
drugs in order to enhance their ability to permeate the BBB.
Due the heavy requirement of the brain for glucose, the BBB is
endowed with a high concentration of the GLUT 1 hexose
transporter protein.79 This glycosylation strategy has been
applied to a range of low molecular weight compounds with
mixed success. Thus glycosyl dopamine derivatives offered
Published on 22 February 2011 on http://pubs.rsc.org | doi:10.1039/C0MT00087F

little advantage over L-dopa or dopamine,80 whereas an


enkephalin glycopeptides produced strong analagesic effects
when administered peripherally81 and a glucose conjugate
of 7-chlorokynurenic acid was found to excert a strong
anticonvulsant effect in rodents.82 The concept has been
extended to chelators, but most of these conjugates have not Clearly this approach has potential, as sugar conjugates
Downloaded by Duke University on 11 January 2013

been directly tested for BBB penetration. Thus glycosylated which behave as a substrate for hexose transporters, will also
tetrahydrosalens (12) have been synthesised which possess a experience facilitated transport by the duodenum. The
high affinity for both copper(II) and zinc(II),83,84 but it remains addition of the sugar lowers the log P value to the
unclear whether or not they are capable of penetrating the (3)–(4) range and thus first pass extraction by the liver
BBB. Feralex (13), a glucose conjugate of a hydroxy- should be minimal. A possible scenario for this class of
pyridinone has been studied in a range of in vitro and cell molecule is presented in Fig. 15.
culture studies85–87 but again no direct demonstration of BBB
permeation. A small series of prodrugs containing a
5. Conclusion
3-glucopyranosyloxy group have also been investigated88
and the 4 iodophenyl derivative (14) has been reported It is clear that iron- and copper chelators will play an
to cross the BBB in rat. However the corresponding appreciable role in the future treatment of various forms of

Fig. 15 Schematic representation of iron removal from CSF by a hydroxypyridinone-sugar conjugate. L Sug—Hydroxypyridinone-sugar
conjugate; L—Hydroxy-pyridinone; FeL3—3 : 1 Hydroxypyridinone / iron complex. The hydroxypyridinone-sugar conjugate (L Sug) is absorbed
from the GIT by facilitated diffusion. By virtue of its hydrophilic nature (log P o 3) it is not susceptible to efficient liver first pass extraction and
so enters the systemic circulation. By virtue of the high concentration of glucose transporters in the BBB, L Sug enters the brain by facilitated
transport, where it will be metabolised to the free hydroxypyridinone which can scavenge iron to form the hydrophobic FeL3 complex, which can
permeate the BBB by nonfacilitated diffusion. Once in the systemic circulation FeL3 will be excreted via the urine.

This journal is c The Royal Society of Chemistry 2011 Metallomics, 2011, 3, 239–249 247
View Article Online

neurodegeneration. Indeed, with the identification of bio- 12 M. E. Gotz, G. Künig, P. Riederer and M. B. H. Youdim,
markers for the various forms of neurodegenerative disease, Pharmacol. Ther., 1994, 63, 37–122.
13 C. W. Olanow and M. B. H. Youdim, New York: Academic Press,
selective chelators many find a role not only for treatment of 1996, 55–69.
the active disease, but also in prophylaxis of disease. The 14 F. Q. Ye, P. S. Allen and W. R. W. Martin, Movement Disord.,
abnormal distribution of these two metals is probably not the 1996, 11, 243–249.
primary cause of the various forms of neurodegeneration, 15 K. A. Jelliger, Drugs Aging, 1999, 14, 115–140.
16 K. Ueda, H. Fukushima, E. Masliah, Y. Xia, A. Iwai,
but is undoubtedly associated with disease progression. M. Yoshimoto, D. A. C. Otero, J. Kondo, Y. Ihara and
Deferiprone (5) is currently involved in clinical trials for the T. Saitoh, Proc. Natl. Acad. Sci. U. S. A., 1993, 90, 11282–11286.
treatment of Friedreich’s ataxia, Parkinson’s disease, macular 17 M. H. Polymeropoulos, C. Lavedan, E. Leroy, S. E. Ide,
A. Dehejia, A. Dutra, B. Pike, H. Root, J. Rubenstein, R. Boyer,
degeneration and Hallervorden-Spatz syndrome. By
E. S. Stenroos, S. Chandrasekharappa, A. Athanassiadou,
modifying the structure of this extremely simple molecule, in T. Papapetropoulos, W. G. Johnson, A. M. Lazzarini,
order to facilitate penetration of the BBB, it is likely that a R. C. Duvoisin, G. Di Iorio, L. I. Golbe and R. L. Nussbaum,
Published on 22 February 2011 on http://pubs.rsc.org | doi:10.1039/C0MT00087F

range of iron-selective chelators will be identified, some of Science, 1997, 276, 2045–2047.
18 K. A. Conway, E. W. Baxter, K. M. Felsenstein and A. B. Reitz,
which may find application in neurodegeneration therapy. Curr. Pharm. Des., 2003, 9, 427–447.
19 S. R. Paik, H. J. Shin, J. H. Lee, C. S. Chang and J. Kim, Biochem.
J., 1999, 340, 821–828.
List of abbreviations 20 S. Miranda, C. Opazo, L. F. Larrondo, F. J. Munoz, F. Ruiz,
F. Leighton and N. C. Inestrosa, Prog. Neurobiol., 2000, 62,
AD Alzheimer’s disease 633–648.
Downloaded by Duke University on 11 January 2013

PD Parkinson’s disease 21 L. E. Scott and C. Orvig, Chem. Rev., 2009, 109, 4885–4910.
Ab1–42 Amyloid b-peptide 1–42 22 Z. D. Liu and R. C. Hider, Med. Res. Rev., 2002, 22, 26–64.
23 D. C. Harris and P. Aisen, Biochim. Biophys. Acta, Gen. Subj.,
Ab1–40 Amyloid b-peptide 1–40 1973, 329, 156–158.
BBB Blood-brain barrier 24 K. N. Raymond, G. Müller and B. F. Matzanke, Top Curr. Chem.,
6-OHDA 6-Hydroxydopamine 1984, 58, 49–102.
CSF Cerebral spinal fluid 25 V. M. Nurchi, G. Crisponi, T. Pivetta, M. Donatoni and
M. Remelli, J. Inorg. Biochem., 2008, 102, 684–692.
NTA Nitrilotriacetic acid 26 X. L. Kong and R. C. Hider, unpublished observations.
PBCA poly(butylcyanoacrylate) 27 D. Holander, D. Ricketts and C. A. R. Boyd, Can. J. Gastroenterol.,
LDL Low density lipoprotein 1988, 2, 35A–38A.
28 W. H. Oldendorf, Proc. Soc. Expt. Bio. Med., 1974, 147, 813–816.
pFe3+ log[Fe3+ 6H2O] at pH 7.4 when [L]Total = 105 M 29 C. A. Lipinski, F. Lombardo, B. W. Dominy and P. J. Feeney,
and [Fe]Total = 106 M. Adv. Drug Delivery Rev., 1997, 23, 3–25.
30 R. A. Yokel, A. M. Fredenburg, K. A. Meurer and T. L. Skinner,
Drug Metab. Dispos., 1995, 23, 1178–1180.
Acknowledgements 31 M. D. Habgood, Z. D. Liu, L. S. Dehkordi, H. H. Khodr,
J. Abbott and R. C. Hider, Biochem. Pharmacol., 1999, 57,
SR’s study was financed by Research Councils UK by the 1305–1310.
provision of a Dorothy Hodgkin postgraduate award. RCH, 32 A. M. Fredenburg, R. K. Sethi, D. D. Allen and R. A. Yokel,
YM and XLK acknowledge general support from Toxicology, 1996, 108, 191–199.
33 D. J. Pennell, V. Berdoukas, M. Karagiorga, V. Ladis, A. Piga,
Apotex, Canada and Vifor, Switzerland. We wish to thank A. Aessopos, E. D. Gotsis, M. A. Tanner, G. C. Smith,
Dr. Francisco Molina-Holgado for the supervision of the M. A. Westwood, B. Wonke and R. Galanello, Blood, 2006, 107,
neurotoxicity studies. 3738–3744.
34 N. Boddaert, K. H. Le Quan Sang, A. Rötlg, A. Leroy-Willig,
S. Gallet, F. Brunelle, D. Sidi, J. C. Thalabard, A. Munnich and
References Z. I. Cabantchik, Blood, 2008, 110, 401–408.
35 O. Kakhlon, H. Manning, W. Breuer, N. Melamed-Book, C. Lu,
1 R. A. Floyd and K. Hensley, Neurobiol. Aging, 2002, 23, 795–807. G. Cortopassi, A. Munnich and Z. I. Cabantchik, Blood, 2008,
2 W. L. Klein, G. A. Krafft and C. E. Finch, Trends Neurosci., 2001, 112, 5219–5227.
24, 219–224. 36 G. L. Forni, M. Balocco, L. Cremonesi, G. Abbruzzese,
3 S. S. Wang, A. Becerra-Arteaga and T. A. Good, Biotechnol. R. C. Parodi and R. Marchese, Movement Disord., 2008, 23,
Bioeng., 2002, 80, 50–59. 904–907.
4 D. A. Butterfield and J. Kanski, Mech. Ageing Dev., 2001, 122, 37 R. A. Cherny, C. S. Atwood, M. E. Xilinas, D. N. Gray,
945–962. W. D. Jones, C. A. Mclean, K. J. Barnham, I. Volitakis,
5 D. A. Butterfield, J. Drake, C. Pocernich and A. Castegna, Trends F. W. Fraser and Y. S. Kim, Neuron, 2001, 30, 665–676.
Mol. Med., 2001, 7, 548–554. 38 A. E. Finefrock, A. I. Bush and P. M. Doraiswamy, J. Am. Geriatr.
6 X. Huang, C. S. Atwood, M. A. Hartshorn, G. Multhaup, Soc., 2003, 51, 1143–1148.
L. E. Goldstein, R. C. Scarpa, M. P. Cuajungco, D. N. Gray, 39 D. Kaur, F. Yantiri, S. Rajagopalan, J. Kumar, J. Q. Mo,
J. Lim, R. D. Moir, R. E. Tanzi and A. I. Bush, Biochemistry, R. Boonplueang, V. Viswanath, R. Jacobs, L. Yang, M. F. Beal,
1999, 38, 7609–7616. D. Di Monte, I. Volitaskis, L. Ellerby, R. A. Cherny, A. I. Bush
7 M. A. Lovell, J. D. Robertson, W. J. Teesdal, J. L. Campbell and and J. K. Anderson, Neuron, 2003, 37, 899–909.
W. R. Markesbery, J. Neurol. Sci., 1998, 158, 47–52. 40 T. Rival, R. M. Page, D. S. Chandraratna, T. J. Sendall, E. Ryder,
8 L. D. Devanur, H. Neubert and R. C. Hider, J. Pharm. Sci., 2008, B. Liu, H. Lewis, T. Rosahl, R. C. Hider, L. M. Camargo,
97, 1454–1467. M. S. Shearman, D. C. Crowther and D. A. Lomas, Eur. J.
9 D. T. Dexter, F. R. Wells and A. J. Lees, J. Neurochem., 1989, 52, Neurosci., 2009, 29, 1335–1347.
381–389. 41 B. Regland, W. Lehmann, I. Abedini, K. Blennow, M. Jonsson,
10 J. Sanchez-Ramos, E. Overvick and B. N. Ames, Neurodegeneration, I. Karlsson, M. Sjögren, A. Wallin, M. Xilinas and C. G. Gottfries,
1994, 3, 197–204. Dementia Geriatr. Cognit. Disord., 2001, 12, 408–414.
11 P. Riederer, E. Sofic, W. D. Rausch, K. Jellinger and M. B. H. 42 A. I. Bush, Trends Neurosci., 2003, 26, 207–214.
Youdim, J. Neurochem., 1989, 52, 515–520. 43 T. Tsubaki, Y. Honma and M. Hosh, Lancet, 1971, 297, 696–697.

248 Metallomics, 2011, 3, 239–249 This journal is c The Royal Society of Chemistry 2011
View Article Online

44 G. P. Oakley, JAMA, J. Am. Med. Assoc., 1973, 225, 395–397. 69 S. Skarlatos, T. Yoshikawa and W. M. Pardridge, Brain Res., 1995,
45 T. Amit, Y. Avramovich-Tirosh, M. B. H. Youdim and S. Mandel, 683, 164–171.
FASEB J., 2008, 22, 1296–1305. 70 B. J. Spencer and I. M. Verma, Proc. Natl. Acad. Sci. U. S. A.,
46 D. Ben-Shachar, N. Kahana, V. Kampel, A. Warshawsky and 2007, 104, 7594–7599.
M. B. H. Youdim, Neuropharmacology, 2004, 46, 254–263. 71 R. A. Cherney, C. S. Atwood, M. E. Xilinas, D. N. Gray,
47 M. Merkofer, R. Kissner, R. C. Hider and W. H. Koppenol, Helv. W. D. Jones, C. A. McLean, K. J. Barnham, I. Volitakis,
Chim. Acta, 2004, 87, 3021–3034. F. W. Fraser, Y. S. Kim, X. Huang, L. E. Goldstein,
48 F. P. Dwyer and D. P. Mellor, Chelating agents and Metal R. D. Moir, J. T. Lim, K. Beyreuther, H. Zheng, R. E. Tanzi,
Chelates, Academic Press, New York, 1964. C. L. Masters and A. I. Bush, Neuron, 2001, 30, 665–676.
49 R. C. Hider, Toxicol. Lett., 1995, 82–83, 961–967. 72 J. Kreuter, J. Nanosci. Nanotechnol., 2004, 4, 484–488.
50 R. D. Abeysinghe, P. J. Roberts, C. E. Cooper, K. H. Maclean, 73 J. Kreuter, Adv. Drug Delivery Rev., 2001, 47, 65–81.
R. C. Hider, J. B. Porter and J. Biol, Chem., 1996, 271, 7965–7972. 74 S. C. J. Steiniger1, J. Kreuter1, A. S. Khalansky, I. N. Skidan,
51 Z. D. Liu, R. Kayyali, R. C. Hider, J. B. Porter and A. I. Bobruskin, Z. S. Smirnova, S. E. Severin, R. Uhl, M. Kock,
A. E. Theobald, J. Med. Chem., 2002, 45, 631–639. K. D. Geiger and S. E. Gelperina, Int. J. Cancer, 2004, 109,
52 C. E. Cooper, G. R. Lynagh, K. P. Hoyes, R. C. Hider, 759–767.
Published on 22 February 2011 on http://pubs.rsc.org | doi:10.1039/C0MT00087F

R. Cammack and J. B. Porter, J. Biol. Chem., 1996, 271, 75 J. Kreuter, P. Ramge, V. Petrov, S. Hamm, S. E. Gelperina,
20291–20299. B. Engelhardt, R. Alyautdin, H. von Briesen and D. J. Begley,
53 Z. D. Liu, M. Lockwood, S. Rose, A. E. Theobald and Pharm. Res., 2003, 20, 409–416.
R. C. Hider, Biochem. Pharmacol., 2001, 61, 285–290. 76 J. Kreuter, D. Shamenkov, V. Petrov, P. Ramge, K. Cychutek,
54 R. J. Ward, D. Dexter, A. Florence, F. Aouad, R. C. Hider. C. Koch-Brandt and R. Alyautdin, J. Drug Targeting, 2002, 10,
P. Jenner and R. R. Crichton, Biochem. Pharmacol., 1995, 49, 317–325.
1812–1826. 77 G. Liu, P. Men, P. L. R. Harris, R. K. Rolston, G. Perry and
55 A. Gaeta, F. Molina-Holgado, X. L. Kong, S. Salvage, S. Fakih, M. A. Smith, Neurosci. Lett., 2006, 406, 189–193.
Downloaded by Duke University on 11 January 2013

P. T. Francis, R. J. Williams and R. C. Hider, Bioorg. Med. Chem., 78 H. Davson and M. B. Segal, Physiology of the CSF and Blood-
2011 inpress. Brain Barriers, CRC press, 1996.
56 Y. M. Ma and R. C. Hider, Tetrahedron Lett., 2010, 51, 79 W. M. Pardridge, R. J. Boado and C. R. Farrell, J. Biological
5230–5233. Chem., 1990, 265, 18035–18040.
57 J. E. Preston, H. al-Sarraf and M. B. Segal, Dev. Brain Res., 1995, 80 C. Fernandez, O. Nieto, E. Rivas, G. Montenegro, J. A. Fontenla
87, 69–76. and A. Fernandez-Mayoralas, Carbohydr. Res., 2000, 327,
58 J. E. Gibbs and S. A. Thomas, J. Neurochem., 2002, 80, 392–404. 353–365.
59 S. Roy, Iron chelator design: evaluation of blood-brain barrier 81 E. J. Bilsky, R. D. Egleton, S. A. Mitchell, M. M. Palian, P. Davis,
permeability and neuroprotective properties, PhD Thesis, London J. D. Huber, H. Jones, H. I. Yamamura, J. Janders, T. P. Davis,
University, 2009. F. Porreca, V. J. Hruby and R. Polt, J. Med. Chem., 2000, 43,
60 F. Molina-Holgado, A. Gaeta, P. T. Francis, R. J. P. Williams and 2586–2590.
R. C. Hider, J. Neurochem., 2008, 105, 2466–2476. 82 G. Battaglia, M. La Russa, V. Bruno, L. Arenare, R. Ippolito,
61 G. Modi, V. Pillay and Y. E. Choonara, Ann. N. Y. Acad. Sci., A. Copani, f. Bonina and F. Nicoletti, Brain Res., 2000, 860,
2010, 1184, 154–172. 149–156.
62 P. Calvo, B. Gouritin, H. Chacun, D. Desmaële, J. D 0 Angelo, 83 T. Storr, M. Merkel, G. X. Song-Zhao, L. E. Scott, D. E. Green,
J. P. Noel, D. Georgin, E. Fattal, J. P. Andreux and P. Couvreur, M. L. Bowen, K. H. Thompson, B. O. Patrick, H. J. Schugar and
Pharm. Res., 2001, 18, 1157–1166. C. Orvig, J. Am. Chem. Soc., 2007, 129, 7453–7463.
63 R. N. Alyaudtin, A. Reichel, R. Löbenberg, P. Ramge, J. Kreuter 84 T. Storr, L. E. Scott, M. L. Bowen, D. E. Green, K. H. Thompson,
and D. J. Begley, J. Drug Targeting, 2001, 9, 409–416. H. J. Schugar and C. Orvig, Dalton Trans., 2009, 3034–3043.
64 R. H. Muller and C. M. Keck, J. Nanosci. Nanotechnol., 2004, 4, 85 T. P. A. Kruck and T. E. Burrow, J. Inorg. Biochem., 2002, 88,
471–483. 19–24.
65 M. Oishi, H. Hayashi, M. Iijima and Y. Nagasaki, J. Mater. 86 R. W. Shin, T. P. A. Kruck, H. Murayama and T. Kitamoto,
Chem., 2007, 17, 3720–3725. Brain Res., 2003, 961, 139–146.
66 V. P. Chekhonin, Y. A. Zhirkov, O. I. Gurina, I. A. Ryabukhin, 87 T. P. Kruck, J. G. Cui, M. E. Percy and W. J. Lukiw, Cell. Mol.
S. V. Lebedev, I. A. Kashparov and T. B. Dmitriyeva, Drug Neurobiol., 2004, 24, 443–459.
Delivery, 2005, 12, 1–6. 88 H. Schugar, D. E. Green, M. L. Bowen, L. E. Scott, T. Storr,
67 J. Lichota, T. Skjørringe, L. B. Thomsen and T. Moos, K. Böhmerle, F. Thomas, D. D. Allen, P. R. Lockman, M. Merkel,
J. Neurochem., 2010, 113, 1–13. K. H. Thompson and C. Orvig, Angew. Chem., Int. Ed., 2007, 46,
68 D. Karkan, C. Pfeifer, T. Z. Vitalis, G. Arthur, M. Ujiie, Q. Chen, 1716–1718.
S. Tsai, G. Koliatis, R. Gabathuler and W. A. Jefferies, PLoS 89 S. Roy, J. E. Preston, R. C. Hider and Y. M. Ma, J. Med. Chem.,
ONE, 2008, 25(e2), 469. 2010, 53, 5886–5889.

This journal is c The Royal Society of Chemistry 2011 Metallomics, 2011, 3, 239–249 249

You might also like