Вы находитесь на странице: 1из 33

REVIEW

Zinc Metabolism in the Brain:


Relevance to Human Neurodegenerative Disorders
Math P. Cuajungco* and Gordon J. Lees*
,
*Department of Psychiatry and Behavioural Science and

Department of Pharmacology
and Clinical Pharmacology, University of Auckland School of Medicine, Private Bag 92019,
Auckland, New Zealand
Received May 12, 1997; accepted August 25, 1997
Zinc is an important trace element in biology. An important pool of zinc in the brain is the one present in
synaptic vesicles in a subgroup of glutamatergic neurons. In this form it can be released by electrical
stimulation and may serve to modulate responses at receptors for a number of different neurotransmit-
ters. These include both excitatory and inhibitory receptors, particularly the NMDA and GABA
A
receptors. This pool of zinc is the only formof zinc readily stained histochemically (the chelatable zinc
pool), but constitutes only about 8% of the total zinc content in the brain. The remainder of the zinc is
moreor lesstightlyboundtoproteinswhereit actseither asacomponent of thecatalyticsiteof enzymesor in
astructural capacity. Themetabolismof zincinthebrainisregulatedbyanumber of transport proteins, some
of which have been recently characterized by gene cloning techniques. The intracellular concentration may
be mediated both by efflux fromthe cell by the zinc transporter ZrT1 and by complexing with apothionein to
form metallothionein. Metallothionein may serve as the source of zinc for incorporation into proteins,
including a number of DNA transcription factors. However, zinc is readily released from metallothionein by
disuldes, increasing concentrations of which are formed under oxidative stress. Metallothionein is a very
good scavenger of free radicals, and zinc itself can also reduce oxidative stress by binding to thiol groups,
decreasing their oxidation. Zinc is also a very potent inhibitor of nitric oxide synthase. Increased levels of
chelatable zinc have beenshowntobe present incell cultures of immune cells undergoingapoptosis. This is
very reminiscent of the zinc staining of neuronal perikarya dying after an episode of ischemia or seizure
activity. Thus a possible role of zinc in causing neuronal death in the brain needs to be fully investigated.
Intraventricular injections of calciumEDTAhave already been shown to reduce neuronal death after a period
of ischemia. Pharmacological doses of zinc cause neuronal death, and some estimates indicate that
extracellular concentrations of zinc could reach neurotoxic levels under pathological conditions. Zinc is
released in high concentrations fromthe hippocampus during seizures. Unfortunately, there are contrasting
observations as to whether this zinc serves to potentiate or decrease seizure activity. Zinc may have an
additional role in causing death in at least some neurons damaged by seizure activity and be involved in the
sprouting phenomenon which may give rise to recurrent seizure propagation in the hippocampus. In
Alzheimers disease, zinc has been shown to aggregate -amyloid, a form which is potentially neurotoxic.
The zinc-dependent transcription factors NF-B and Sp1 bind to the promoter region of the amyloid
precursor protein (APP) gene. Zinc also inhibits enzymes which degradeAPPto nonamyloidogenic peptides
and which degrade the soluble form of -amyloid. The changes in zinc metabolism which occur during
oxidative stress may be important in neurological diseases where oxidative stress is implicated, such as
Alzheimers disease, Parkinsons disease, and amyotrophic lateral sclerosis (ALS). Zinc is a structural
component of superoxide dismutase 1, mutations in which give rise to one form of familiar ALS. After HIV
infection, zinc deciency is found which may be secondary to immune-induced cytokine synthesis. Zinc is
involved in the replication of the HIV virus at a number of sites. These observations should stimulate further
researchintotheroleof zincinneuropathology.

1997 Academic Press
Key Words: zinc metabolism; metallothionein; zinc neurotoxicity; Alzheimers disease; epilepsy;
cerebral ischemia; neurodegenerative disease.
Neurobiology of Disease 4, 137169 (1997)
Article No. NB970163
137
0969-9961/97 $25.00
Copyright

1997 by Academic Press
All rights of reproduction in any form reserved.
INTRODUCTION
Zinc is one of the important trace elements in
biological systems (Bettger & ODell, 1981; Golden,
1989; Mills, 1989; Vallee, 1983). It is redox inert and has
structural, catalytic, and regulatory roles in cellular
biology (Bettger & ODell, 1981; Vallee, 1983; Williams,
1989). Zinc is a small ion (0.65 ) and binds mostly to
nitrogen (N) and sulfur (S) donors, but easily ex-
changes ligands due to its lowligand eld stabilization
energy (Williams, 1989). Thus, zinc is exchangeable
between cell compartments. In addition to enzymatic
and structural roles for zinc, several motifs of deoxyri-
bonucleic acid (DNA)-binding proteins have been
identied and characterized as zinc ngers, zinc clus-
ters, and zinc twist (Vallee et al., 1991). Hence, cellular
zinc is also crucial for genetic expression. The biochem-
istry of zinc has been extensively reviewed (Vallee &
Falchuk, 1993).
There is circumstantial evidence suggesting that zinc
might be involved in several neurological dysfunc-
tions. Postmortem studies have revealed signicant
variations of zinc content in the pathologic brain
(Andrasi et al., 1995; Constantinidis, 1990, 1991; Corri-
gan et al., 1993; Deibel et al., 1996; Deng et al., 1994;
Dexter et al., 1989a,b, 1991; Jenner et al., 1992; Wenstrup
et al., 1990; Yasui et al., 1993). Further, chelatable zinc is
mobilized and accumulates in degenerating neurons
after transient cerebral ischemia (Koh et al., 1996;
Tonder et al., 1990) and seizure activity (Frederickson et
al., 1989). Some calcium-binding proteins (CBPs) with
helixloophelix motifs (EF-hand motif) contain not
only calcium, but also high-affinity zinc-binding sites
(Baudier et al., 1983; Calissano et al., 1974; Filipek et al.,
1990). Of these proteins, S-100 has been functionally
correlated with various diseases of the brain (Griffin et
al., 1989, 1995; Sheng et al., 1996; Van Eldik & Griffin,
1994). Zinc interacts with the proteins, -amyloid and
its precursor protein, which are believed to be in-
volved in the causation of degenerative processes in
the brain, particularly in Alzheimers disease (Bush et
al., 1993, 1994a, 1994b, 1995, 1996; Clements et al., 1996;
Esler et al., 1996; Mantyh et al., 1993). Zinc has inhibi-
tory effects on several enzymes critical for normal cell
functioning (Brewer et al., 1979; Donaldson et al., 1971;
Link &von Jagow, 1995; Mustafa et al., 1971; Persechini
et al., 1995; Zhang et al., 1990). This review will
concentrate on both vesicular and ionic zinc contents
in cells that could possibly be perturbed during parox-
ysmal activity and degenerative events in the brain.
We will present and evaluate current understanding
with regard to the role that zinc might play in cellular
degeneration and/or proliferation. We will explore
possible connections between zinc and its effects on
proteinprotein interactions, genetic expression, neuro-
trophic activity, oxidative stress, and programmed cell
death.
1. OVERVIEW OF ZINC METABOLISM IN
THE BRAIN
Localization of Chelatable Zinc in the Central
Nervous System
The content of transition metals in the mammalian
central nervous system (CNS), especially zinc, has
been studied using techniques such as histochemistry
(Danscher, 1981; Danscher et al., 1985; Haug, 1967;
Kozma et al., 1981; Slomianka et al., 1990), histouores-
cence (Frederickson et al., 1987a), stable-isotope dilu-
tion (Frederickson et al., 1982, 1983), and retrograde/
anterograde techniques (Casanovas-Aguilar et al., 1995;
Christensen et al., 1992; Haug et al., 1971; Howell &
Frederickson, 1989; Kesslak et al., 1987; Long et al.,
1995). There are three main sources of zinc in the brain,
namely: (a) a vesicular pool localized in the synaptic
vesicles of nerve terminals; (b) a membrane-bound,
metalloprotein, or proteinmetal complex pool in-
volved in both metabolic reactions and nonmetabolic
reactions such as structural support for biomembranes
and protein folds; and (c) an ionic pool of free or
loosely bound ions in the cytoplasm (Frederickson,
1989). There are undoubtedly both vesicular and metal-
locomplex pools of zinc in the CNS. However, the
existence of free zinc (Zn
2
) in the cellular milieu
seems unlikely (Frederickson, 1989) since a number of
zinc-binding proteins are present in brain tissue (Ebadi,
1991; Ebadi & Hama, 1986; Itoh et al., 1983). Neverthe-
less, there is an indication that very lowconcentrations
of an intracellular ionic pool of zinc exist (Reyes, 1996;
Simons, 1991; Williams, 1989). Chelatable or labile
refers to the ability of complexing agents to abstract
Zn
2
from zinc-bound cellular proteins within a given
parameter (e.g., binding affinity of the metalligand
complex). Chelatable zinc is easily detected using the
various histological techniques. The source of the
chelatable zinc pool is believed to be vesicular compart-
ments, loosely bound zinc protein, and ionic zinc.
Among these, the vesicular zinc pool is quantitatively
the most signicant (Danscher et al., 1985; Frederick-
son et al., 1983; Haug, 1967; Perez-Clausell &Danscher,
1985). Neurons that have a chelatable zinc pool in their
terminal boutons and release zinc during synaptic
138 Cuajungco and Lees
Copyright

1997 by Academic Press
All rights of reproduction in any form reserved.
activity are termed zinc-enriched neurons (ZEN)
(Danscher, 1996; Frederickson, 1989). The identica-
tion of ZEN in the mammalian brain has been previ-
ously reviewed (Frederickson &Moncrieff, 1994; Slomi-
anka, 1992). The metabolic or structural pool of zinc
may not be chelatable due to its strong complexation
with proteins and/or location of zinc deep within the
protein structure (Frederickson, 1989). In the CNS,
labile zinc is mainly detected in the telencephalon,
with particularly high concentrations in the vesicular
compartments of the hippocampal mossy ber termi-
nals (Danscher et al., 1985; Frederickson et al., 1983;
Haug, 1967; Perez-Clausell & Danscher, 1985; Slomi-
anka et al., 1990). The neo-Timms sulde/silver tech-
nique is a histochemical staining method used to
detect heavy metals, has a high selectivity for zinc, and
is the most common method used to detect zinc in
mammalian brain (e.g. see Danscher, 1981, 1996; Haug,
1967; Sloviter, 1982). The TimmDanscher reaction
products label only the chelatable pool of zinc (Dan-
scher, 1981, 1996; Danscher et al., 1985; Perez-Clausell
& Danscher, 1985).
Cellular Zinc Levels
The tissue zinc content in the gray matter is between
150 and 200 M (Ebadi, 1991; Sato et al., 1984), while a
220300 M concentration of terminal bouton zinc is
estimated to be present in the vesicles of the mossy
bers (Frederickson et al., 1983).
Zn
2
is crucial to over 200 proteins and enzymes
(Prasad, 1996; Vallee & Falchuk, 1993). Zn
2
concentra-
tions range from 10
9
M within the cytoplasm in
most cells to 10
3
M in some vesicles (Williams,
1989). In transgenic baby hamster kidney (BHK) cells
that express zinc inux transporter proteins, it was
estimated that vesicular zinc concentrations reach up
to 14 mMwhen these cells are exposed to high levels of
exogenous zinc (Palmiter et al., 1996a). Whether these
vesicular zinc concentrations can be attained by the
synaptic vesicles of ZENs under pathological condi-
tions in vivo is currently unknown.
Cell survival in vitro is compromised when cells are
exposed to extracellular zinc concentrations ([Zn
2
]
e
)
between 225 and 1000 M in neuronal cells (Choi et al.,
1988; Koh & Choi, 1994; Yokoyama et al., 1986) and
between 7.5 and 200 M in nonneuronal cells (Provin-
cialli et al., 1995; Telford & Fraker, 1995). [Zn
2
]
e
may
reach 300 M during seizure activity (Assaf & Chung,
1984). Hence, it is essential that cells regulate their
intracellular zinc concentrations ([Zn
2
]
i
) through ma-
nipulation of ion inux and efflux. In addition, physi-
ological chelation of Zn
2
to apothionein or other
zinc-sequestering proteins, and amino acids ensures
homeostatic control (Ebadi, 1991; Palmiter & Findley,
1995; Palmiter et al., 1992, 1996a,b).
Zinc Release and Uptake
Zn
2
is released in a calcium-dependent manner
from the hippocampal mossy ber terminals during
spontaneous activity (Charton et al., 1985), after stimu-
lation evoked electrically (Aniksztejn et al., 1987; How-
ell et al., 1984), after administration of potassium ions
(Aniksztejn et al., 1987; Assaf & Chung, 1984; Charton
et al., 1985), or after administration of kainic acid (KA)
(Assaf & Chung, 1984; Charton et al., 1985). Whether
the liberated zinc is in an ionic form or in a low-affinity
zinc complex remains to be determined. The amount of
zinc that may be released from synaptic vesicles of the
mossy ber terminals is approximately 8% of the total
hippocampal zinc (Frederickson et al., 1983). It was
suggested that the concentration of zinc released may
reach 300 M at the peak of convulsive activity (Assaf
& Chung, 1984). This concentration was determined
based on the accumulated Zn
2
in the perfusing
medium of hippocampal tissue slices. The actual con-
centration within the synaptic cleft may be even
greater as the estimated value from the perfusate is an
average concentration of Zn
2
which has been diluted
by the surrounding extracellular uid.
Neuronal presynaptic terminals sequester Zn
2
from
the extracellular space (Howell et al., 1984; Perez-
Clausell & Danscher, 1986; Wensink et al., 1988). High-
and low-affinity uptake mechanisms for zinc were
found by Howell and colleagues (1984) with K
D
15
M and K
D
361 M, respectively; Wensink et al.
(1988) reported an even higher binding affinity con-
stant of K
D
0.25 M. Zinc uptake is saturable and
unaffected by adenosine triphosphate (ATP) levels or
the concentration gradient of sodium ions (Wensink et
al., 1988).
Very recently, three homologous mammalian zinc
transporters (ZnTs) involved in the sequestration
(ZnT-2/ZnT-3) and extrusion (ZnT-1) of zinc from
transfected BHK cells were cloned and functionally
identied (Palmiter & Findley, 1995; Palmiter et al.,
1996a,b). Interestingly, mammalian ZnTs have se-
quence homology with ZnT genes identied in yeast
(Kamizono et al., 1989; Zhao & Eide, 1996). ZnT-1
appears to be mostly localized in the plasma mem-
brane (Palmiter & Findley, 1995) and is expressed in
Zinc Metabolism in the Brain 139
Copyright

1997 by Academic Press
All rights of reproduction in any form reserved.
the CNS and other organs (Palmiter et al., 1996b).
According to Palmiter and Findley (1995), the rate of
zinc efflux increases as the extracellular concentration
is increased, which initially indicates that efflux is an
energy-dependent mechanism and that ZnT-1 is not a
channel or a facilitated transporter. However, it was
found that efflux was unaffected by metabolic inhibi-
tors of oxidative phosphorylation (sodium azide or
potassium cyanide). A similar resistance to azide or
cyanide was observed with the zinc effluxer in fungi
(Vepachedu & Mohan, 1996), indicating functional
homology with the mammalian ZnT-1 transporter.
Thus, ZnT-1 was postulated to be some type of second-
ary active transporter (Palmiter &Findley, 1995). Mean-
while, ZnT-2 was suggested to be a component protein
of vesicular (acidic) intracellular compartments which
facilitates zinc accumulation, but has a low affinity for
zinc and is hardly detectable in the brain (Palmiter et
al., 1996a,b). On the other hand, ZnT-3, another zinc
transporter that may be involved in zinc uptake into
synaptic vesicles of ZENs was very recently cloned by
Palmiters laboratory (Palmiter et al., 1996b). ZnT-3 is
expressed exclusively in murine brain and testis. ZnT-3
messenger ribonucleic acid (mRNA) is detected most
strongly in neurons of the hippocampal dentate gran-
ule cells and pyramidal cells of the CA1 and CA3
regions, as well as in the cerebral cortex. Immunolabel-
ing of ZnT-3 produced staining patterns similar to that
shown by neo-Timms sulde/silver method in detect-
ing vesicular zinc in the CNS.
Zinc uptake from the blood to the brain is under
homeostatic control (Blair-West et al., 1990; Kasarskis,
1984; Pullen et al., 1995). However, homeostatic control
of intracellular zinc by vesicular sequestration in the
CNS has never been evaluated.
Zinc and Metallothionein
The importance of zinc in cellular functions makes
metallothionein (MT) of critical importance in the
control of its bioavailability (Vallee & Falchuk, 1993).
Metallothionein proteins are a group of low-molecular-
weight (M
r
67 kDa), single-polypeptide chains of
60 or more amino acid residues. Currently, there are
only four identied functional mammalian isoforms
(MT-I, -II, -III, and -IV), although an additional six
human genes coding for MT have been characterized
(Stennard et al., 1994). MT has relatively high cysteine
(2530%) and serine/lysine (1218%) contents with a
complete absence of heterocyclic and aromatic amino
acids. It is able to bind up to a total of 7 g atom
zinc/molecule of thionein protein. The reader is re-
ferred to a recent review on the functions of MT
(Vallee, 1995).
MT is ubiquitous in most organs and can be induced
by different endogenous or exogenous stimuli includ-
ing heavy metals and glucocorticoids (Kagi & Schaffer,
1988; Vallee, 1995). Studies have found MT to modu-
late the transfer and homeostasis of metal ions (Cano-
Gauci & Sarkar, 1996; Ebadi, 1991; Ebadi et al., 1995,
1996; Masters et al., 1994; Zeng et al., 1991a,b). MT
prevents zinc deciency and toxicity in vivo (Kelly et
al., 1996). MT has also been proposed to function as a
detoxifying agent of reactive metals and free radicals
(Sato & Bremner, 1993; Thornalley & Vasak, 1985).
Furthermore, it also regulates gene expression in vitro
via interactions with transcription factors (Cano-Gauci
&Sarkar, 1996; Zeng et al., 1991a,b). In cultured Ehrlich
cells, the cellular concentration of zinc-bound MT was
determined to be approximately 13 M(Krezoski et al.,
1988). MT, however, is not a long-term storage for zinc
as it has a short biological half-life in these cells
(Krezoski et al., 1988). Hence, both vesicular storage
and the control of ZnTs also provide important contri-
butions to zinc homeostasis (Fig. 1).
Expression of MT isoforms is tissue-specic. For
example, MT-I and MT-II are expressed in a zinc-
dependent manner (Atar et al., 1995; Durnam & Pal-
miter, 1981) and are detectable in most mammalian
organs including the brain (Ebadi, 1991; Ebadi et al.,
1995, 1996; Palmiter et al., 1992). Investigations on the
mechanism of convulsive seizures and the consequent
observations of changes in zinc staining in the hippo-
campus led to the discovery of MT in the brain (Ebadi,
1991; Ebadi et al., 1995; Itoh et al., 1983). Two recently
discovered isoforms, MT-III and MT-IV, were observed
to be expressed mainly in brain (Palmiter et al., 1992;
Pountney et al., 1994; Tsuji et al., 1992; Uchida et al.,
1991), and in squamous epithelia, respectively (Quaife
et al., 1994). MT-III has 7 additional amino acids, giving
a total of 68 residues (Tsuji et al., 1992; Uchida et al.,
1991). MT-III is localized among a subset of astrocytes
within the vicinity of neuronal perikarya and den-
drites (Tsuji et al., 1992; Uchida et al., 1991) and is
expressed independently of the common potent induc-
ers of MT-I and MT-II (Palmiter et al., 1992). Recently,
MT-III was detected in ZENs of the hippocampal
mossy ber in vivo (Erickson et al., 1997; Masters et al.,
1994). It was observed that cultured brain cells express-
ing MT-III sequester exogenously added zinc as evi-
denced by an increased intensity of Timms stain.
MT-III is unique among the other known isoforms in
that it has an inhibiting effect on neuronal growth;
140 Cuajungco and Lees
Copyright

1997 by Academic Press
All rights of reproduction in any form reserved.
FIG. 1. Zinc metabolism cycle. The diagram illustrates cell regulation of intracellular zinc following uptake or permeation through ionotropic
glutamate receptors. Homeostatic control of zinc levels is accomplished via vesicular-mediated and membrane protein-transporter-mediated
sequestration of ions (ZnTs). Zinc buffering is also accomplished by metallothionein protein (MT). MT has and clusters that can bind up to a
total of seven zinc atoms. MT synthesis is induced by zinc (and other agents) by liberating the transcription factor MTF-1 from MT transcription
inhibitor (MTI). MTF1 is believed to bind to the metal responsive element (MRE) of the MT gene promoter, thus initiating transcription.
However, the mechanism that controls the expression of the MTF-1 gene or the ZnT gene is not known. TF, transcription factor.
Zinc Metabolism in the Brain 141
Copyright

1997 by Academic Press
All rights of reproduction in any form reserved.
hence, it is also known as growth inhibitory factor
(Erickson et al., 1994; Pountney et al., 1994; Tsuji et al.,
1992; Uchida et al., 1991).
Zinc, Metallothionein, and Gene Expression
In addition to regulation of MT synthesis, zinc also
plays a major role in the control of the activity of DNA
and RNA(Martinez-Balbas et al., 1995; Rubin & Koide,
1973; Vallee, 1983; Vallee & Falchuk, 1993). Activator
proteins (AP) regulating transcription and replication
contain zinc (Vallee & Falchuk, 1993). The genetic
advantage and relative importance of incorporating
zinc for folding protein domains have been reviewed
(OHalloran, 1993; Rhodes & Klug, 1993; Schwabe &
Klug, 1996). Zinc is essential for binding of eukaryotic
Xenopus laevis transcription factor IIIA (TFIIIA) to the
DNA site of the 5S-RNA gene (Martinez-Balbas et al.,
1995; Vallee & Falchuk, 1993). Zinc nger proteins
have high affinities for zinc (K
D
10
8
10
12
M; Shang
et al., 1989). The removal of zinc by various complexing
agents, including MT, antagonizes binding to DNA
genes by TFIIIA (Shang et al., 1989; Zeng et al., 1991a)
and specicity protein type 1 (Sp1) (Zeng et al., 1991b).
Metal abstraction, however, is reversed by zinc supple-
mentation (Zeng et al., 1991a,b). Sp1 is a zinc-
dependent member of the activator protein family of
the transcription apparatus and the rst human tran-
scription factor identied to recognize the specic
regulatory sequence of the guanine-cytosine box (Tjian,
1995).
Nuclear factor-B (NF-B) is a zinc-dependent tran-
scription factor that translocates between cell cytoplasm
andnucleus (Baeuerle, 1991). NF-Bcontrols the expres-
sion of a variety of genes, including that of immunore-
ceptors, cytokines, and several viruses (Baeuerle, 1991).
NF-B requires zinc for its normal DNA-binding activ-
ity (Zabel et al., 1991). Gold-mediated oxidation of the
zinc-bound cysteines, as well as chelation of zinc using
1,10-orthophenanthroline (1 mM), can disable NF-kB
binding activity to DNA (Yang et al., 1995). Other
zinc-dependent transcription factors include estrogen
receptor (EsR; Cano-Gauci & Sarkar, 1996), erythroid
receptor (Eryf-1 also known as GATA-1, NF-E1, or
GF-1; Omichinski et al., 1993), Zif268 (also known as
Krox-24, NGFI-A, or Egr-1; Pavletich &Pabo, 1991), and
glucocorticoid receptor (Luisi et al., 1991), all of which
contain the zinc nger DNA-binding motif.
Zinc Involvement in Cytokine Function
Cytokines are various polypeptides that mediate
cellular responses in immunity and inammation in
both the peripheral nervous system and the CNS
(Hopkins & Rothwell, 1995; Rothwell & Hopkins,
1995). Inammatory and immune responses are known
to deplete serum zinc bioavailability (Bremner & May,
1989; Goldblum et al., 1987; Taubeneck et al., 1995). A
possible mechanism by which zinc may be depleted is
through enhanced activities of zinc-requiring enzymes
(involved in DNA/nucleic acid synthesis) of mitogen-
or cytokine-responsive cells (Odeh, 1992).
Many cytokines including tumor necrosis factor-
(TNF-) and interleukin-1 (IL-1) have been implicated
in neurological dysfunctions related to neurodegenera-
tive conditions (Griffin et al., 1989, 1995; McGeer &
McGeer, 1995; Stanley et al., 1994). In vivo, TNF-
causes alterations in zinc metabolism leading to deple-
tion of zinc known as hypozincemia (Taubeneck et al.,
1995), while IL-1 induces not only hypozincemia, but
also hypoferremia (iron deciency) (Goldblum et al.,
1987).
The presence of zinc has been histochemically iden-
tied in the 7S-NGF complex (a pro-protein of which
NGF- is a constituent subunit) present in nonneural
granule cells (Frederickson et al., 1987b). 7S-NGF has a
high affinity for zinc (Pattison & Dunn, 1976b), while
murine NGF-, a functional dimer, also contains zinc-
binding sites (Holland et al., 1994). Zinc is a potent and
specic inhibitor of the native 7S-NGF -subunit con-
taining esteropeptidase (Pattison & Dunn, 1976b). It
has been observed in vitro that chelation of zinc from
7S-NGF using 2,2,2-terpyridine or 8-hydroxyquino-
line-5-sulfonic acid activates the -esteropeptidase en-
zyme and liberates the , , and subunits from the
oligomer (Pattison & Dunn, 1976a,b).
2. GENERAL MECHANISMS WHICH LINK
ZINC TO NEURODEGENERATIVE
DISEASES
Oxidative Stress
It has been proposed that oxidative stress due to
diminished antioxidant defenses and/or increased
production of reactive oxygen species (ROS) and free
radicals could initiate a cascade of events leading to
142 Cuajungco and Lees
Copyright

1997 by Academic Press
All rights of reproduction in any form reserved.
neurodegeneration (Olanow, 1993). However, the pos-
sibility that free radical production is a consequence of
underlying degenerative processes, rather than a ma-
jor cause of degeneration, needs to be borne in mind
(for reviews see Coyle &Puttfarcken, 1993; Lees, 1993).
This section will provide an overview of oxidative
stress and other processes which link zinc to disease
progression. Recent reports suggest a possible involve-
ment of zinc in the regulation of oxidative stress and,
in particular, some form of association with the gluta-
thione (GSH) antioxidative system.
Energy release involves aerobic and anaerobic
mechanisms. In the aerobic process, the reduction of
molecular oxygen (O
2
) to water (H
2
O) results in the
formation of ROS (Halliwell & Gutteridge, 1984; Ola-
now, 1993). ROS and free radical species such as
superoxide radical (O
2

), hydrogen peroxide (H
2
O
2
),
peroxynitrite (ONOO

), nitric oxide (NO

), hydroxyl
radical (OH

), and hypochlorous acid (HOCl) can


cause oxidative stress and cell death when their rate of
production exceeds the cellular capacity to metabolize
them (Halliwell & Gutteridge, 1984).
The glutathione peroxidase family is an important
cellular antioxidant defense mechanism (Jenner, 1993;
Meister, 1983). Peroxidases catalyze the reduction of
H
2
O
2
to H
2
O and O
2
. Oxidation of GSH, catalyzed by
glutathione peroxidase (GPO), results in the formation
of oxidized-glutathione (GSSG; glutathione disulde).
Glutathione reductase (GR) catalyzes the reduction of
GSSG back to GSH. In vivo depletion of GSH shows a
marked correlation to a decrease of GR activity in rat
brains, thus decreasing the capacity to reduce GSSG
and increasing the vulnerability to oxidative stress
(Barker et al., 1996). GSH may also be converted to
mercapturate products catalyzed by glutathione S-trans-
ferase (GST) or degraded back to cysteinylglycine and
-glutamyl amino acid by -glutamyl-transpeptidase
(Jenner, 1993, Meister, 1983). GSH and GSH-related
enzymes are differentially distributed in the CNS
(Huang & Philbert, 1995; Kudo et al., 1990).
Disulde products are abundant in cellular systems
and were recently found to interact with MT by
facilitating the release of zinc from its clusters (Maret,
1994, 1995). Therefore, alterations in the GSH/GSSG
ratio that result in increased concentrations of GSSG
and other biological disuldes could release and el-
evate cytoplasmic zinc levels. As a homeostatic re-
sponse, increased production of MT and/or efflux of
zinc may occur. Interestingly, hepatic GSH deciency
results in the synthesis of MT I and II isoforms (Sato et
al., 1995). It is possible that the loss of GSH antioxidant
defense triggers a compensatory response by synthesiz-
ing MT to neutralize free radicals and to prevent
cytoplasmic accumulation of zinc caused by MT
disulde interaction. Both zinc and MT have been
proposed as antioxidants against ROS and free radicals
(for reviews see Bray & Bettger, 1990 and Sato &
Bremner, 1993, respectively).
The knowledge gained involving the interactions of
zinc, MT, and the GSH antioxidant family marks an
important event that links their functions to cellular
senescence and death.
The Neurotoxicity Of Zinc
The antioxidant protective effects of zinc can be
overridden at higher levels of zinc. The importance of
increases in endogenous cellular zinc as a neuroprotec-
tive or neurotoxic mechanism (Fig. 2) is discussed in
more detail in subsequent sections. This section dis-
cusses the toxic effects of pharmacological doses of
zinc, which nevertheless correspond to levels which
may be reached in pathological states. The neurotoxic
potential of exogenous zinc in neurons has been
observed both in vivo (Chuah et al., 1995; Cuajungco &
Lees, 1996; Lees & Leong, 1994; Lees et al., 1990) and in
vitro (Choi et al., 1988; Duncan et al., 1992; Koh et al.,
1996; Yokoyama et al., 1986). It is not known, however,
if zinc-induced neurotoxicity occurs only by a necrotic
mechanism of death (cell edema and lysis). Zinc-
induced neuronal death is associated with excitotoxic
injury (see next section). Even though excitotoxic
injury typically results in necrosis, recent reports
showed that apoptosis (gene-directed cell death) may
also occur after an excitotoxic insult (Ankarcrona et al.,
1995; Bonfoco et al., 1995; Portera-Cailliau et al., 1995).
This section will concentrate on necrotic cell death,
while the association of zinc with apoptotic death will
be discussed later.
Zinc exposure at a nonphysiological dose (600
M) for 15 min resulted in neuronal death in mixed
cultures of cortical and glial cells, while exposure for a
longer duration (1824 h) at lower concentrations
(300 M) induces glial cell death (Choi et al., 1988;
Yokoyama et al., 1986). In vivo, intranasal irrigation of
zinc (17 mol) in mice produced axonal degeneration
of olfactory nerve bundles (Schwann cells) (Chuah et
al., 1995). Some olfactory Schwann cells were immu-
nopositive for S-100 protein and appeared to have
expanded perinuclear space (Chuah et al., 1995). An
association between S-100 proteins and neurodegenera-
Zinc Metabolism in the Brain 143
Copyright

1997 by Academic Press
All rights of reproduction in any form reserved.
FIG. 2. Aschematic summary representing several cellular protein and enzyme activities inuenced by both intracellular and extracellular zinc
levels. (A) The diagram shows that an increase of extracellular levels of zinc through vesicular release and ZnT-mediated efflux may affect
several extracellular processes. Several inhibitory and excitatory receptors, neurotransmitter uptake proteins, and matrix metalloproteinases
(MMPs), enzymes that degrade basement membrane (BM) of the extracellular matrix (ECM), have been found to be modulated by the presence
of zinc. (B) The diagram illustrates several mechanisms where intracellular zinc elevation may occur. Ascenario is depicted where intracellular
zinc overload may overwhelm homeostatic control mechanisms in cells, consequently producing detrimental effects on various protein and
enzyme activities necessary for normal cell functioning. It is interesting to note that zinc has also been observed to yield a positive effect on
apoptotic cell death, whereby it may inhibit endonuclease-mediated DNA cleavage or it may activate PARP DNA repair enzyme, thus,
precluding apoptosis.
144
tion is discussed elsewhere in this review. While the
above in vitro and in vivo observations clearly show
that zinc is neurotoxic, due to diffusion and dilution by
the extracellular uid (in vivo) or intracellular uid (in
vitro), the nal zinc levels reached in cells after expo-
sure to zinc is difficult to determine. In addition, the
minimum concentration required to induce neuronal
death in vivo is unknown (Lees et al., 1990).
Blocking the Toxic Effects of Zinc
Our endeavor to prevent zinc neurotoxicity in vivo
led us to look for complexing agents that would inhibit
the effects of exogenous zinc-induced neuronal death
(Cuajungco & Lees, 1996). This rationale is analogous
to a report that chelating calcium in vitro and in vivo
protects neurons fromdamage due to prolonged hippo-
campal tissue stimulation (Scharfman & Schwartz-
kroin, 1989) or focal cortical ischemia (Tymianski et al.,
1993), respectively. The protective effect of the calcium
chelator, however, was evident only in cells that do not
contain CBPs (Scharfman & Schwartzkroin, 1989).
CBPs are known to buffer Ca
2
and contribute to
intracellular calcium concentration ([Ca
2
]
i
) homeosta-
sis (Heizmann & Braun, 1995). Using metal complex-
ing agents, it was shown that membrane permeable
chelators N,N,N,N-tetrakis(2-pyridylmethyl)ethylene-
diamine (TPEN) (Cuajungco &Lees, 1996) and pyrithi-
one (Cuajungco &Lees, manuscript submitted) attenu-
ate zinc neurotoxicity in vivo, while the membrane
impermeant chelator ethylenediamine tetraacetic acid
(EDTA) reduces zinc toxicity in vitro (Koh et al., 1996).
Note that neither TPEN nor pyrithione showed inher-
ent toxicity in the absence of exogenous metal ions in
the rat hippocampus. The implications of chelating
intracellular zinc using TPEN both in vivo and in vitro
are discussed in the forthcoming section on apoptosis.
Putative Mechanisms of Zinc-Induced Neuronal
Cell Death
The presence of zinc-binding proteins (Ebadi, 1991;
Ebadi & Hama, 1986; Itoh et al., 1983) and possibly an
efflux protein transporter similar to ZnT-1 (Palmiter &
Findley, 1995) serves as homeostatic control mecha-
nisms. However, aberrations in the metabolism and
buffering of zinc resulting in an excess of [Zn
2
]
i
and
its subsequent extrusion resulting in elevated [Zn
2
]
e
could be detrimental to cells. The exact mechanism of
zinc-induced cell death is currently unknown. There
are several ways zinc may augment or directly trigger
neuronal death by interactions with various cell recep-
tors, enzymes, and/or other metal ions present in the
cellular environment (Fig. 2).
a. Receptor Interactions
The ultrastructural localization of several neuro-
transmitters present in the hippocampal mossy ber
terminals, especially glutamate (Beaulieu et al., 1992;
Bramham et al., 1990; Kesslak et al., 1987; Stengaard-
Pedersen et al., 1983; Storm-Mathisen & Ottersen,
1990), coincides with the location of ZENs (Frederick-
son et al., 1982, 1983; Haug, 1967; Haug et al., 1971;
Perez-Clausell & Danscher, 1985). Modication of the
toxicity of excitatory amino acids (EAAs) is thus a
possible mechanism by which zinc may induce neuro-
nal death. For example, zinc exacerbates the toxicity of
-amino-3-hydroxy-5-methyl-4-isoxazole propionic
acid (AMPA) (Freund & Reddig, 1994; Koh & Choi,
1988; Koh et al., 1996; Weiss et al., 1993) and KA(Nave
& Connor, 1993; Shiraishi et al., 1993; Yin & Weiss,
1995), but reduces N-methyl-D-aspartate (NMDA) tox-
icity (Koh & Choi, 1988; Peters et al., 1987; Xie et al.,
1993). The alteration of EAAneurotoxicity is due to an
interaction of zinc with both ionotropic and metabo-
tropic glutamate receptors (Bresink et al., 1996; Dreixler
& Leonard, 1994; Harrison & Gibbons, 1994; Koh &
Choi, 1994; Smart et al., 1994; Xie et al., 1993).
In addition, zinc binds to the -aminobutyric acid
(GABA) receptors and noncompetitively inhibits
GABA-mediated responses (Kumamoto & Murata,
1995; Kume et al., 1994; Mayer & Vyklicky, 1989;
Ricciardi & Malouf, 1995; Smart, 1992; Smart et al.,
1994; Turgeon &Albin, 1992; Wang et al., 1995; White &
Gurley, 1995). The neuronal death in cerebral ischemia
is due in part to the ratios of glutamate, glycine, and
GABA released extracellularly (the excitotoxic index)
(Globus et al., 1991). An interaction of zinc with these
receptors could thus modify the excitotoxic index and
be region specic, due to the localization of neurons
containing vesicular zinc. Interestingly, zinc modulates
both glutamate (Dreixler & Leonard, 1994) and GABA
(Wang et al., 1995; White & Gurley, 1995) responses of
neurons that contain only specic splice-variant recep-
tor subunits of NMDA and GABA
A
, respectively. The
dissociation constants (K
D
) for zinc antagonism of
NMDA (Mayer et al., 1989) and GABA (Mayer &
Vyklicky, 1989) responses are approximately 13 and 11
M, respectively. These are well within the concentra-
tions released during synaptic activity, which implies
an association of zinc with the regulation of neurotrans-
mission. Likewise, zinc can modulate the ATP P
2X
-
Zinc Metabolism in the Brain 145
Copyright

1997 by Academic Press
All rights of reproduction in any form reserved.
purinergic (Cloues, 1995; Cloues et al., 1993; Koizumi et
al., 1995), opiate (Connor & Chavkin, 1992), and gly-
cine (Kumamoto & Murata, 1996; Laube et al., 1995)
receptors, as well as the dopamine transporter com-
plex (Richeld, 1993). Recently, zinc was observed to
inhibit the actions of a competitive NMDA receptor
antagonist, CGP-39653, on glutamate and glycine bind-
ing on NMDAreceptors (Terse &Komiskey, 1997). This
nding has implications regarding the efficacy of
competitive NMDA receptor antagonists for use as
neuroprotectants in CNS injuries caused by cerebral
ischemia or epilepsy.
b. Interactions with the Activities of Critical
Enzymes and Transport Proteins
Zinc can inhibit the cell respiratory chain (K
i
10
7
M) by blocking the initial step of respiration of electron
transfer between ubiquinone (coenzyme Q) and cyto-
chrome b of the bc
1
complex (complex III) (Hunter &
Ford, 1955; Kleiner & von Jagow, 1972; Link & von
Jagow, 1995). At higher concentrations (10
3
M), zinc
may further inhibit at the levels of avoprotein 1 or 2
(complex I or II) and cytochrome c oxidase (complex
IV) activities (Skulachev et al., 1967).
Zinc also inhibits the ion transporter enzymes such
as Na

/K

-ATPase both in vitro and in vivo (ID


50
20
M, Donaldson et al., 1971; Hexum, 1974; Mustafa et
al., 1971) and calcium-ATPase in vitro (IC
50
50100
M; Brewer, 1980; Brewer et al., 1979; Zhang et al.,
1990). The Na

/K

-ATPase enzyme activity has been


found to be highly concentrated in the hippocampus
and hypothalamus (Donaldson et al., 1971). Inhibition
of Na

/K

-ATPase is known to cause neuronal death


(e.g., see Lees, 1991; Lees & Leong, 1994, 1995).
Zinc inhibits a number of enzymes critical for
metabolismof neurotransmitters (Dreosti, 1989; Freder-
ickson, 1989). It inhibits uptake transporters for both
glutamate and GABA (K
m
50 M) in synaptosomal
fractions of mice (Gabrielsson et al., 1986). Inhibition of
the glutamate uptake mechanism produces progres-
sive extracellular glutamate accumulation and subse-
quent death of cortical neurons in vitro (Velasco et al.,
1996). However, the ability of physiological zinc concen-
trations to depress glutamate/aspartate uptake trans-
porter activity was recently disputed (Killinger et al.,
1995).
c. Interactions with Structural Proteins
Zinc has been shown to inuence both assembly and
disassembly of tubulin (K
D
10
5
10
7
M; Eagle et al.,
1983) and a number of microtubule-associated proteins
in vitro (Banerjee et al., 1982; Gaskin & Kress, 1977;
Gaskin et al., 1978; Kress et al., 1981). An excess zinc
concentration in nervous tissues could interfere with
the microtubule structure by disrupting the normal
functions of the cytoskeleton (Kress et al., 1981). The
metal chelator, TPEN, prevented protein kinase C
(PKC)-mediated actin cytoskeletal disruption induced
by phorbol ester in cultured C6 rat glioma cells, while
the addition of zinc reversed the protective effects of
TPEN (Hedberg et al., 1994).
d. Perturbation of Intracellular Calcium and
Interactions with Calcium-Binding Proteins
It has been observed that neuronal swelling caused
by zinc is dependent on the presence of sodiumin vitro
(Choi et al., 1988). Koh and Choi (1994) proposed that
zinc inux via the NMDA-receptor channel may be a
mechanism of zinc-induced neurotoxicity. Recently,
however, Koh et al. (1996) showed that activation of
AMPA receptors but not NMDA receptors mediated
zinc toxicity in cultured neurons. Zinc inux may
depolarize the cell and consequently trigger a rise in
intracellular calcium concentration ([Ca
2
]
i
). It is also
conceivable that Zn
2
inux may perturb buffering of
[Ca
2
]
i
by interacting with CBPs, thus exacerbating the
condition. A rise in [Zn
2
]
i
could result in the activa-
tion of second-messenger systems via PKC-mediated
(Csermely et al., 1988; Murakami et al., 1987) phosphor-
ylation of receptor ion channels and/or voltage-
dependent gene expression (Atar et al., 1995; Rubin &
Koide, 1973) which could initiate cellular death path-
ways. NMDA receptor-dependent Ca
2
inux and
overload are known to result in excitotoxic damage
(Mody & MacDonald, 1995). In contrast, it has been
reported that Zn
2
attenuates NMDA receptor-acti-
vated channel currents by binding to two different
receptor sites in vitro (Mayer et al., 1989), which
suggests some protective role for zinc. Therefore, it
remains to be delineated if zinc toxicity is caused by an
excitotoxic mechanism mediated by overaccumulation
of [Ca
2
]
i
through activation of either NMDA- or
non-NMDA-gated ion channels or whether it acts via
direct inhibition of processes critical for cell survival or
both.
Anumber of EF-hand CBPs have been implicated in
neurodegenerative disorders (Heizmann & Braun,
1995). Some of these proteins contain distinct zinc-
binding sites such as S-100, S-100A1/S-100, S-100A3/
S-100E, S-100A6/calcyclin, calmodulin (CaM), and
calgranulin C (for a review, see Schafer & Heizmann,
146 Cuajungco and Lees
Copyright

1997 by Academic Press
All rights of reproduction in any form reserved.
1996). Zinc was observed to bind CaM and induces a
slight conformational change (Baudier et al., 1983).
However, calcium has a greater affinity for CaM
(K
D
10
8
10
6
M) and that zinc binding was sug-
gested to be nonspecic (Baudier et al., 1983). Zinc has
been found to inhibit the CaM-complexed Ca
2
-
ATPase of erythrocyte membranes (I-310) (Brewer,
1980; Brewer et al., 1979). Calcyclin, homologous to
S-100, is mainly present in neurons and is not found
in glial cells (Filipek et al., 1993). Its function is
currently unknown, but has a specic binding site for
zinc (Filipek et al., 1990) with a half-maximal affinity
(K
s
) of 2 mM (Pedrocchi et al., 1994). Generally, the
affinity of most CBPs for zinc is very low, such that it is
unlikely to affect them under normal physiological
zinc concentrations (typically in the nanomolar range).
Zinc Deciency and Neurological Dysfunction
The consequential effects of nutritional zinc de-
ciency in biological systems have been documented.
Lack of zinc in animals results in teratogenic and CNS
abnormalities in vivo (Dreosti, 1989; McNall et al., 1995;
Taubeneck et al., 1995). For example, chronic zinc
deciency has been observed to affect neuronal func-
tions in the hippocampal mossy bers (Hesse, 1979;
Wensink et al., 1987). Likewise, dietary zinc deciency
for 3 months in rats showed a 30% maximum reduc-
tion of mossy ber zinc content in the hippocampus
(Wensink et al., 1987). Further, Hesse (1979) found that
zinc-decient rats showed abnormal hippocampal
mossy ber synaptic responses during low-frequency
stimulation. Recently, it was found that functional
NMDA receptors in cortical synaptic membranes in
vitro were decreased due to low zinc bioavailability
(Browning & ODell, 1995). Low zinc status also
activates brain MT-III expression, which results in cell
growth retardation (Palmiter, 1995). Finally, it was
observed that maternal zinc deciency (4 days) in
rats produces excessive embryonic apoptotic cell death
especially within the neural crest (Rogers et al., 1995). It
is interesting to note that in several human neurodegen-
erative diseases, zinc is specically depleted in several
brain regions (see Part 3).
Zinc and Apoptosis
This section discusses in detail the involvement of
zinc in apoptosis occurring in cells from nonnervous
tissues (Sunderman, 1995) as it provides a basis for
comparative studies for the role of zinc in apoptotic
death in the CNS. Apoptosis (or programmed cell
death) was rst described by Kerr et al. (1972) from
their observation of distinct morphological changes
occurring in cells in vitro. Apoptosis and necrosis are
two distinct ways a cell can die. The characteristic
features of apoptosis are membrane blebbing, chroma-
tin condensation, DNAfragmentation, organellar pack-
aging, and cell shrinkage (Wyllie et al., 1973). Necrosis,
however, is characterized by cell swelling and lysis, is
accompanied by inammation, and is thus a rather
violent way for a cell to die (Wyllie et al., 1973).
Neuronal apoptosis has been postulated to be a
mechanism typically occurring in neurodegenerative
disorders (Thompson, 1995, for a review). Evidence
comes from in vitro and in vivo models of brain injury
caused by applications of excitotoxins (Ankarcrona et
al., 1995; Bonfoco et al., 1995; Portera-Cailliau et al.,
1995) or periods of hypoxiaischemia (Beilharz et al.,
1995; Nitatori et al., 1995), or paroxysmal activities
(Charriaut-Marlangue et al., 1996; Pollard et al., 1994a,b;
Simonian et al., 1996). There is also some evidence that
apoptosis occurs in the striatum and temporal lobes of
Huntingtons disease and Alzheimers disease pa-
tients, respectively (Dragunow et al., 1995; Su et al.,
1994). Recently, evidence of apoptotic cell death in the
substantia nigra of Parkinsons disease brains was also
conrmed (Mochizuki et al., 1996), although this is
contradictory to the report of Dragunow et al. (1995).
Free radicals have been proposed to play a major role
in the initiation of the apoptotic process (Bonfoco et al.,
1995; Greenlund et al., 1995; Ratan et al., 1994).
Surprisingly, the involvement of zinc in neuronal
apoptosis is just beginning to be explored. Zinc may be
very relevant in neuronal apoptosis as cells require
many zinc-dependent transcription factors in normal
functioning (OHalloran, 1993; Rhodes & Klug, 1993;
Schwabe & Klug, 1996; Vallee et al., 1991) and during
induction of apoptosis (Xu et al., 1996). There is
currently no information on the relevance of zinc in
neuronal apoptosis, and thus, we discuss correlative
evidence that zinc plays a role in neuronal apoptotic
death.
a. Accumulation of Intracellular Zinc
in Apoptotic Cells
Cultured lymphoblasts undergoing early events of
apoptosis have an increased intracellular zinc content
(Zalewski et al., 1994). This was demonstrated by the
use of a zinc-specic uorophore, Zinquin. The investi-
gators emphasized that the observed enhanced intracel-
Zinc Metabolism in the Brain 147
Copyright

1997 by Academic Press
All rights of reproduction in any form reserved.
lular zinc uorescence was not due to zinc uptake from
the mediumsince the latter was zinc-free and occurred
before cell membrane disruption (see Zalewski et al.,
1994 for discussion).
While the mechanism for [Zn
2
]
i
mobilization is not
known, enhanced zinc dissociation from metallopro-
teins in vitro can be produced by free radicals and ROS
such as HOCl, H
2
O
2
, and O
2

(Fliss & Menard, 1991,


1992; Fliss et al., 1991), EAAs such as L-glutamate,
L-aspartate, and L-cysteic acid (Baba et al., 1989), and
biological disuldes (Maret, 1994, 1995). These observa-
tions could provide a pivotal role for dissociated
Zn
2
to reduce or exacerbate pathological conditions in
certain disease states, depending on the levels of [Zn
2
]
i
.
After a period of cerebral ischemia (Koh et al., 1996;
Tonder et al., 1990) or seizure activity (Frederickson et
al., 1989), degenerating neurons stain positively for
zinc. These neurons appear to be dying by an apoptotic
process (Beilharz et al., 1995; Charriaut-Marlangue et
al., 1996; Nitatori et al., 1995; Pollard et al., 1994a,b;
Simonian et al., 1996). The increase in cellular zinc
occurs only after a delay of 324 h following an insult,
but precedes other markers for cellular damage such
as cellular staining with acid fuchsin. The marked
decrease in neuronal death brought about by EDTA
implies that cellular uptake of zinc is critical for
neuronal death caused by a period of ischemia (EDTA
is not taken up by cells) (Koh et al., 1996). Translocation
frompresynaptic ZENterminals to postsynaptic neuro-
nal perikarya appears to be the mechanism (Frederick-
son et al., 1989; Koh et al., 1996; Tonder et al., 1990).
Hence, both increased uptake and intracellular mobili-
zation are possible sources for the increased intracellu-
lar zinc following an insult.
b. Induction and Prevention of Apoptosis by
Manipulating Intracellular Zinc
i. Modulation of induction of apoptosis by the
extracellular concentration. Exogenously added zinc
can have biphasic effects on nonneuronal cells by
either initiating or preventing apoptosis (Cohen &
Duke, 1984; Provinciali et al., 1995; Sunderman, 1995;
Telford & Fraker, 1995; Treves et al., 1994). There is,
however, a variable range of concentrations where zinc
may induce or preclude apoptotic death. For example,
cultured mouse thymocytes incubated with zinc at
concentrations 15 Mundergo apoptosis (Provinciali
et al., 1995), while another study found induction of
apoptotic death when extracellular zinc concentrations
were 80200 M (Telford & Fraker, 1995). On the other
hand, zinc concentrations from 75 M and up to 5000
M have been shown to prevent DNA fragmentation
and apoptotic death in a variety of nonneural cells
induced by either serum/growth factor deprivation or
exposure to toxic/chelating agents (Cohen & Duke,
1984; Jiang et al., 1995; McCabe et al., 1993; Provincialli
et al., 1995; Shimizu et al., 1990; Telford & Fraker, 1995;
Treves et al., 1994; Zalewski et al., 1993, 1994). These
results contrast markedly with induction of c-myc-
dependent apoptosis by zinc (37.5 M). Thus, the
increased intracellular zinc in apoptotic lymphoblasts
reported by Zalewski et al. (1994) may be a protective
mechanism. In contrast, the rise of zinc in ischemic
neurons would appear to be neurotoxic as a chelating
agent reduced neuronal death (Fig. 4) (Koh et al., 1996).
ii. Activationof apoptosis by chelationof intracel-
lular zinc. Apoptosis with or without elevation of
[Ca
2
]
i
, in vitro can be triggered by chelating [Zn
2
]
i
using TPEN in thymocytes, lymphocytes, and spleno-
cytes (Jiang et al., 1995; Treves et al., 1994; Zalewski et
al., 1993). TPEN is a cell-permeant complexing agent
with a high affinity for zinc but low affinity for calcium
(Anderegg et al., 1977; Arslan et al., 1985). These results
are comparable to recent ndings in which apoptosis
was induced by exposure to lowlevels of Zn
2
(Provin-
cialli et al., 1995; Telford & Fraker, 1995). Note that zinc
chelation-mediated death is not only independent of,
but also additive to, apoptosis induced by exogenous
addition of Ca
2
(Jiang et al., 1995). These observations
clearly indicate that Zn
2
is important in at least some
apoptotic cell death pathways.
iii. Hippocampal zincchelationinvivo. We found
that intrahippocampal injections of TPEN (10 nmol)
alone were not toxic to hippocampal neurons (Cuajungco
& Lees, 1996) in spite of complexing most of, if not all,
chelatable mossy ber zinc for up to 8 h (Cuajungco &
Lees, unpublished observation). This result is in direct
contradiction with in vitro observations that nonneuro-
nal cells loaded with TPENundergo apoptosis (Jiang et
al., 1995; McCabe et al., 1993; Treves et al., 1994).
Morphological evidence of apoptosis was not apparent
4 to 14 days after TPEN administration neurons (Cu-
ajungco & Lees, 1996). However, molecular techniques
are necessary to corroborate whether in vivo exposure
of neurons to TPEN did not induce apoptotic death.
These results imply that different classes of cells may
have different intracellular zinc requirements for the
initiation of apoptosis.
c. Mechanisms by Which Zinc May Modify
Apoptotic Pathways
i. I nteractions between zinc and immediate early
genes in apoptosis. Immediate early genes (IEGs)
148 Cuajungco and Lees
Copyright

1997 by Academic Press
All rights of reproduction in any form reserved.
and their products, known as inducible transcription
factors (ITFs), are known to play a role in the cascade
of events leading to apoptosis (for a review, see
Dragunow&Preston, 1995). Several ITFs such as c-Jun,
JunB, JunD, c-Fos, Fos-B, and the zinc-dependent ITF
Krox-24 show basal, but diverse expression in the
mammalian CNS (Herdegen et al., 1995). Both neural
and nonneural cells undergoing apoptotic death have
been found to express IEGs such as c-fos, c-jun, and
hsp-70 (Dragunow et al., 1990; Estus et al., 1994).
Induction of either c-fos or c-jun in neurons is highly
correlated with apoptotic death (Dragunowet al., 1990;
Estus et al., 1994). In contrast, c-fos in nonneural murine
cell cultures was recently demonstrated to be unneces-
sary in cell suicide, as cells with a disrupted c-fos
/
gene (spleen and thymus) underwent spontaneous
and etoposide-induced apoptosis (Gajate et al., 1996).
Zinc is able to induce c-jun and c-fos expressions in
hepatocyte (H-7) cells (Xu et al., 1996). However, no
further data are available to indicate whether zinc
produces a transient or prolonged expression of c-fos
and c-jun or whether zinc-induced expression of the
two IEGs is directly associated with the induction of
cell death. A more direct involvement of zinc in IEG
regulation of apoptosis was shown when zinc (37.5
M) facilitated the c-myc-dependent apoptosis of cul-
tured H-7 cells (Xu et al., 1996). Cells exposed to either
copper or cadmium did not produce any effects,
suggesting a specic zinc-dependent step for c-myc-
induced apoptosis (Xu et al., 1996). Apoptotic death via
c-myc occurs without DNA fragmentation (Xu et al.,
1996).
ii. I nhibition of endonuclease. DNA fragmenta-
tion is one of the hallmarks of apoptosis which is
closely associated with key morphological changes
(Wyllie, 1980), but may not be essential. Glucocorticoid-
induced thymocyte apoptosis results in the activation
of the calcium/magnesium (Ca
2
/Mg
2
)-dependent
endonuclease in which DNAlinker regions are cleaved
into around 200-basepair fragments (Wyllie, 1980).
Hence, one hypothesis is that zinc halts apoptosis in
both cell cultures and cell-free isolated nuclei via
competitive inhibition of endonuclease activity (Cohen
& Duke, 1984). This proposal was challenged when
Barbieri et al. (1992) found that zinc inhibition (15
mM) of DNAfragmentation did not protect rat thymo-
cytes from spontaneous or dexamethasone-induced
apoptosis. Furthermore, it has been reported that the
main morphological features of apoptosis may occur
without DNA cleavage, thus suggesting that DNA
fragmentation is not a critical event in the death
process (Barbieri et al., 1992; Cohen et al., 1992; Xu et al.,
1996). Other recent studies have challenged the idea
that the Ca
2
/Mg
2
-dependent endonuclease activity
is controlled solely by [Ca
2
]
i
(e.g., see Jiang et al., 1995;
Treves et al., 1994). It is interesting to note that an
increase in [Ca
2
]
i
induces apoptosis in immature
thymocytes (Cohen & Duke, 1984), while mature
thymocytes could resist apoptosis even with an ele-
vated [Ca
2
]
i
(McCabe et al., 1993). This is further
evidence supporting differential induction of apopto-
sis among various cell types.
iii. Regulation of growth cycle. A rather interest-
ing observation was noted in mouse thymocytes in
which preferential inhibition or induction of apoptosis
occurs during the G
0
/G
1
phase (Provinciali et al., 1995;
Telford & Fraker, 1995). This could be explained by
previous ndings that cultured broblasts required
zinc-dependent processes (i.e., activation of thymidine
kinase and possibly DNA histone H3) (Chesters et al.,
1993) in order for them to progress from G
1
to S phase
(Chesters et al., 1989). According to Willson (1989) cells
may not divide during low intracellular zinc status at
critical sites, as zinc is thought to protect cells from
iron-catalyzed thiol oxidation during mitosis, thus
preventing intramolecular disulde formation.
iv. DNA strand breakage repair enzyme. DNA
strand breakage results in the activation of poly-
[adenosine diphosphate(ADP)-ribose] synthetase
(PARS) and poly-[ADP-ribose] polymerase (PARP),
which recognize and repair strand damage (de Murcia
& de Murcia, 1994; Szabo et al., 1996). PARP is a zinc
nger protein whose activity depends on the presence
of zinc (Zahradka & Ebisuzaki, 1984). PARP catalyzes
the attachment of poly(ADP-ribose) from the coen-
zyme nicotinamide adenine dinucleotide (NAD

) (de
Murcia & de Murcia, 1994). Activation of PARP/PARS
signicantly depletes NAD

(de Murcia & de Murcia,


1994; Szabo et al., 1996). The process of replenishing the
diminished NAD

is postulated to consume too much


energy, consequently exhausting the cellular ATPstores,
which then ultimately leads to the cells demise. While
zinc (1 mM) inhibits DNA fragmentation and cell
death, it does not prevent the N-methyl-N-nitro-N-
nitrosoguanidine-mediated increase of PARP synthesis
(Shimizu et al., 1990). Thus, the protective effects of
zinc on apoptosis do not appear to be due to a direct
inhibition of PARP activity.
v. Protein kinase regulation. PKC is an enzyme
activated by the second messengers Ca
2
and diacyl-
glycerol that then exerts its effects at certain cytoplas-
mic locations (Huang, 1989). PKC contains zinc nger
domains and has a high-affinity binding site for zinc
(Hubbard et al., 1991; Murakami et al., 1987; Schwabe &
Zinc Metabolism in the Brain 149
Copyright

1997 by Academic Press
All rights of reproduction in any form reserved.
Klug, 1996). Zinc has biphasic effects on PKC activity,
with stimulation at low concentrations (300400 M)
and inhibition at high concentrations (400 M), both
in the presence of calcium (Murakami et al., 1987).
Binding of zinc can result in the translocation of PKC
to the membrane and augment its activity on the
membrane cytoskeleton (Baba et al., 1991; Csermely et
al., 1988; Zalewski et al., 1990, 1991). Downregulation
of PKC activity is highly correlated with the induction
of apoptosis in vitro (Lin et al., 1995; Rinaudo et al.,
1995). Taken together, these results indicate that zinc
may have an intrinsic role in PKC activity in terms of
either initiating or preventing apoptotic cell death.
The above observations demonstrate an important
role for zinc in the complex processes of apoptotic cell
death and provide further evidence that diverse and
complex mechanisms of cell suicide exist among differ-
ent cell types.
3. POSSIBLE INVOLVEMENT
OF ZINC IN SELECTED
NEUROLOGICAL DISORDERS
In any discussion of causation in neurological dis-
eases, correlations can often be made between pathol-
ogy and abnormal concentrations or metabolism of
many factors. Which events are primary and which are
secondary is often unclear, and in many cases changes
may be downstream events not involved in pathology.
In the following discussion on the role of zinc in CNS
dysfunctions, aberrations in zinc metabolism are high-
lighted, with emphasis on those processes which are
likely to be involved in the causation of the disorder. In
a number of cases, the changes observed are mutually
inconsistent. However, all possibilities are listed, in the
hope of stimulating future research to determine which
changes in zinc metabolism are most important. Fu-
ture research will inevitably nd that in some, if not
most, of these diseases, abnormal zinc metabolism is a
secondary event that may be critical in the common
pathways by which neurons die (e.g., the function of
zinc in cell suicide) rather than for the initiation of the
disease process.
Hypoxia-Ischemia
The role of zinc in neuronal death caused by cerebral
ischemia was rst investigated by Tonder and col-
leagues (1990). Recent conrmation by Chois group
indicates that zinc is in fact involved in the pathology
resulting from transient global cerebral ischemia (Koh
et al., 1996). A growing amount of evidence suggests
that selective delayed neuronal death in cerebral isch-
emia occurs via apoptosis (Ankarcrona et al., 1995;
Beilharz et al., 1995; Charriaut-Marlangue et al., 1996).
Thus, it is reasonable to postulate that zinc status may
be involved in ischemia-induced apoptosis.
The apparent similarity between neurodegeneration
following ischemia and kainic acid-induced toxicity
(an experimental model of epilepsy) led Tonder and
colleagues (1990) to propose that zinc has an important
function in cell death. Transient cerebral ischemia in
rats produced a zinc-dependent increase in the uores-
cence of N-(6-methoxy-8-quinolyl)-p-toluene sulfon-
amide quinoline (TSQ) in the dentate hilus with a
reduction of TSQ uorescence in the mossy ber
terminals (see Fig. 3) (Koh et al., 1996; Tonder et al.,
1990). Neither neuronal perikarya nor glial cell bodies
normally exhibit positive Zn
2
uorescence (Koh et al.,
1996). Prior chelation of Zn
2
by dithizone abrogated
Zn
2
uorescence within degenerating neurons (Koh
et al., 1996). TSQ-positive neurons in the hilus (Koh et
al., 1996; Tonder et al., 1990) and pyramidal CA1 cells
(Koh et al., 1996) were in fact degenerating, as evi-
denced by positive staining with acid fuchsin, a marker
for damaged neurons (Fig. 3). Furthermore, other brain
areas such as the neocortical laminae (II/III/IV), tha-
lamic reticular nucleus, amygdala, and striatum, known
to selectively degenerate after transient global isch-
emia, also showed positive TSQ uorescence within
neuronal perikarya (Koh et al., 1996). Interestingly,
Chois group found that Zn
2
accumulation preceded
neuronal death and that chelation of Zn
2
via intraven-
tricular injection of calcium-EDTA (Ca-EDTA) 30 min
prior to insult decreased ischemic-related neurodegen-
eration by up to 70% (Fig. 4) (Koh et al., 1996). Hence,
zinc is one of many factors involved in ischemia-
induced neuronal death. It was suggested that zinc
cytotoxicity and other excitotoxic insults may be code-
pendent processes required to induce lethal injury
after global cerebral ischemia (Koh et al., 1996). There is
a need, however, to dene the mechanisms involved in
zinc neurotoxicity, in order to devise an efficacious
counteractive measure to minimize neuropathological
sequelae following brain trauma.
Epilepsy
a. The Role of Zinc in Modulating Seizure Activity
Ever since the discovery that zinc induces seizures in
rats, it has been used as an experimental model of
150 Cuajungco and Lees
Copyright

1997 by Academic Press
All rights of reproduction in any form reserved.
epilepsy (Pei & Koyama, 1986; Pei et al., 1983). This
nding has also resulted in the suggestion that the
release of cellular zinc plays a major role in the
generation and propagation of epileptic activity (Assaf
& Chung, 1984; Frederickson et al., 1988, 1989; Pei &
Koyama, 1986; Pei et al., 1983; Sloviter, 1985). However,
opposing lines of evidence suggest that endogenous
zinc may have an anticonvulsant effect. These contrast-
ing observations are reviewed below.
Zinc in synaptic vesicles of the hippocampal mossy
bers (Danscher, 1981; Haug, 1967; Perez-Clausell &
Danscher, 1985, 1986) is proposed to be co-localized
with glutamate (Beaulieu et al., 1992; Frederickson,
1989). Glutamate is believed to be the major EAA
neurotransmitter in excitatory hippocampal pathways
(Bramham et al., 1990; Storm-Mathisen & Ottersen,
FIG. 3. TSQ staining and neuronal degeneration in the hippocam-
pus. (A) Fluorescent photomicrograph of normal rat hippocampus
after staining with TSQ, showing dense uorescence in the hilus (H)
and the stratum lucidum (SL) of CA3. In addition, TSQ staining is
seen in the stratum radiatum (SR) and the stratum oriens (SO) of
CA1 (DG, dentate gyrus). No Zn
2
uorescence is seen in the
stratum pyramidale (SP) or alveus (alv.) of CA1 (dark bands). (B)
Twenty-four hours after a 10-min period of forebrain ischemia, TSQ
uorescence is reduced in presynaptic terminals and is newly
apparent in the cell bodies of some hilar neurons. (C) The same
hippocampal section as in (B), with subsequent acid fuchsin stain-
ing. All the TSQ-uorescent neurons in (C) exhibited ischemic
acidophilic changes (pink cytoplasm, arrows). (D) Seventy-two
hours after a 10-min period of ischemia, dense TSQ staining
appeared in degenerating CA1 pyramidal neurons. (E) The same
section as in (D), with acid fuchsin staining. All the CA1 neurons
with Zn
2
uorescence showed acidophilic changes. Bars, 800 m
(A); 200 m (B and C); 100 m (D and E). Reprinted with permission
from Koh et al. (1996) Science 272, 10131016. Copyright 1996
American Association for the Advancement of Science.
FIG. 4. CaEDTA protects hippocampal hilar and CA1 neurons
against ischemic injury. (A) Hilar neuron counts (mean SEM) in
normal rats (n 4) or in saline-injected controls (n 20), Ca-EDTA-
treated rats (n 12), and Zn-EDTA-treated rats (n 7), 72 hours
after a 10-min period of ischemia. (B) CA1 pyramidal neuron counts
for the same rats. Three rats injected with Ca-EDTA were studied
after 14 days. Asterisks denote difference from controls ( p 0.05
two-tailed t test with Bonferroni correction for multiple compari-
sons). Reprinted from Koh et al. (1996) Science 272, 10131016 with
permission.
Zinc Metabolism in the Brain 151
Copyright

1997 by Academic Press
All rights of reproduction in any form reserved.
1990). The suggestion that glutamate and zinc are
co-localized within neurons comes from observations
that: (i) ZENs in the hippocampus release the cation in
a calcium-dependent manner during synaptic activity
(Aniksztejn et al., 1987; Assaf & Chung, 1984; Charton
et al., 1985; Howell et al., 1984); and (ii) the release of
zinc is abolished following chemical destruction of the
hippocampal mossy bers (Aniksztejn et al., 1987). It is
noteworthy that not all glutamatergic neurons contain
zinc (Beaulieu et al., 1992; Frederickson, 1989).
There are conicting results regarding the role of
endogenous zinc in the propagation of seizures
(Mitchell &Barnes, 1993; Mitchell et al., 1990; Morton et
al., 1990; Theoret et al., 1988). For example, Mitchell
and Barnes (1993) observed that intraperitoneal injec-
tion (ip) of diethyldithiocarbamate (DEDTC, 50 mg/
kg) chelates endogenous zinc and lowers the threshold
for convulsions when the rat hippocampal perforant
path was stimulated, indicating a protective role for
the release of zinc. Furthermore, DEDTC and dithi-
zone were found to augment the severity of KA-
induced seizures (Mitchell et al., 1990). However,
dithizone (100 mg/kg) and DEDTC (1000 mg/kg)
administered ip 30 min prior to intrahippocampal (ih)
injection of zinc (10 nmol) did not protect hippocampal
neurons from zinc-induced neurotoxicity (Lees, Cu-
ajungco, & Leong, unpublished observation). Like-
wise, co-injection (ih) of DEDTC (100 nmol) with zinc
(10 nmol) had no signicant effect on zinc neurotoxic-
ity (Cuajungco & Lees, unpublished observation).
Thus, it is uncertain whether the proconvulsant effect
of DEDTC and dithizone reported by other groups is
due to chelation of zinc or a side effect of the chelating
drugs. Morton et al. (1990) found that subcutaneous
preadministration of zinc (100 mg/kg) signicantly
decreased the frequency of audiogenic seizures in
mice, but did not signicantly attenuate KA-induced
seizures. There is, however, a considerable doubt as to
whether this effect is due to a central action of zinc. The
amount of zinc crossing the bloodbrain barrier within
minutes of injection is likely to be insufficient to
modify seizures. This is because previous studies have
shown that peripheral zinc is only slowly taken up into
the cerebrospinal uid (Blair-West et al., 1990) and into
the brain (Kasarskis, 1984), where it is rapidly bound to
plasma proteins and nonneural tissues (Kasarskis,
1984). Although Morton et al. (1990) estimated that a
minimum of 2 M zinc (i.e., 0.2% of 100 mg/kg
dosage) would have reached the extracellular compart-
ments of the CNS via facilitated diffusion of the
microligand-bound fraction, it can be argued that the
brain has high-affinity zinc-binding proteins to buffer
any excess extracellular zinc (Ebadi, 1991; Ebadi &
Hama, 1986; Itoh et al., 1983).
A very recent study using resected hippocampal
tissue from medial temporal lobe sclerosis patients has
reported that exogenously applied zinc (300 M)
attenuated the dentate granule cell hyperexcitability,
thus limiting seizure activity in situ (Williamson &
Spencer, 1995). From this nding, the authors pro-
posed that zinc may act as inhibitory neuromodulator
in pathological situations, as zinc is only normally
released during repetitive stimulation (Williamson
& Spencer, 1995). More recently, however, Buhl and
co-workers (1996) have suggested that zinc released
from aberrantly sprouted mossy ber terminals in the
human temporal lobe epilepsy (HTLE) kindling
model could cause a direct collapse of an intensied
autoinhibition by GABA through attenuation of the
GABA response. This is evidenced by the observation
that zinc (200 M) signicantly blocked the inhibitory
postsynaptic currents in kindled dentate granule cells
of hippocampal slices, whereas similar concentrations
of zinc did not produce any effects on control neurons
(Buhl et al., 1996). Whether zinc is important for the
neuronal death resulting from sustained seizure activ-
ity is a separate question from its involvement in
seizure activity per se. Nonetheless, the association of
zinc in epilepsy needs conrmation.
b. Zinc and Neuronal Death in Epilepsy
Fromthe clear involvement of zinc in neurotransmis-
sion and neuromodulation, it was proposed that synap-
tic zinc mobilization during paroxysmal activity could
be involved in the consequent neuropathology of
epileptic activity (Assaf & Chung, 1984; Frederickson
et al., 1988, 1989; Fukahori et al., 1988; Sloviter, 1985).
Some researchers have observed an elevated level of
zinc in the brains of rats and mice following epileptic
seizures (Chung & Johnson, 1983; Mody & Miller,
1985), while another study reported otherwise (Fuka-
hori et al., 1988). These variations could be due to
differences between the types of seizure model used
and the time between seizure and sacrice. Mody and
Miller (1985) used a rat commissural-kindling model,
Chung and Johnson (1983) used the audiogenic seizure-
prone DBA/2J mice strain, while Fukahori et al. (1988)
used E1 mice, a strain that is seizure-prone to postural
stimulations or paraboloid movements. Mody and
Miller (1985) sacriced and measured hippocampal
zinc levels of the rats 24 h after the last stimulation,
whereas Fukahori et al. (1988) sacriced the E1 mice at
7 days after the last induced seizure. Chung and
152 Cuajungco and Lees
Copyright

1997 by Academic Press
All rights of reproduction in any form reserved.
Johnson (1983) did not specify the time at which
hippocampal zinc levels of the DBA/2J mice were
measured. It is possible then that the observed differ-
ences in hippocampal zinc concentration after a sei-
zure could be due to variations in the lag time before
zinc levels were measured. Moreover, the severity and
latency of paroxysmal activities may also explain the
anomalous ndings of the above studies.
A consistent observation is that zinc is translocated
from presynaptic boutons to postsynaptic neurons as
revealed by TSQ uorescence 12 h after KA-induced
seizures (Frederickson et al., 1988, 1989) or by Timms
stain 24 h following perforant path stimulation-
induced seizures (Sloviter, 1985). Experimental models
of epilepsy have provided increasing evidence that
neurons severely affected by seizures die by an apop-
totic process (Charriaut-Marlangue et al., 1996; Pollard
et al., 1994a,b; Simonian et al., 1996). While the rise in
[Ca
2
]
i
resulting from the activation of glutamate
receptors is one likely cause of cell death (Choi, 1988,
1992), zinc may also be involved in the pathology.
However, it remains to be determined whether the
toxic and convulsant effects of pharmacological doses
of zinc occur in vivo during release of endogenous zinc
from neurons.
c. Zinc, Cytokines, and Neuronal Reorganization
Following Seizure Activity
Dense sprouting and distinctive reorganization of
the hippocampal mossy bers are commonly found in
HTLE tissue and experimental brain models (Babb et
al., 1991; Cronin & Dudek, 1988; Sundstrom et al., 1993;
Sutula et al., 1989). Sundstrom et al. (1993) found an
increased staining of ZEN terminal bers of the den-
tate granule cell layer (inner molecular) after KA
lesioning of the CA3 pyramidal cells and CA4 commis-
sural association bers. The group inferred that the
presence of suprapyramidal band stain provides a
direct link between aberrant sprouting and propaga-
tion of epileptiform activity. A zinc-containing growth
factor-like protein has been suggested to be the cause
of sympathetic sprouting in deafferented septo-
hippocampal projections (Kesslak et al., 1987; Frederick-
son et al., 1983; Stewart et al., 1984; Yu & Crutcher,
1995). The zinc-containing NGF- is one such growth
factor which promotes highly specic sprouting of
dennervated septo-hippocampal (cholinergic) connec-
tions (Kesslak et al., 1987; Frederickson et al., 1983;
Stewart et al., 1984; Yu & Crutcher, 1995).
Another important nding is the signicant and
specic up-regulation of S-100 expression, but not
GFAP in HTLE resected tissues, possibly signifying a
physiological role for this protein in aberrant axonal
sprouting seen in epilepsy (Griffin et al., 1995). It was
suggested that overexpression of S-100 could be a
factor in excessive formation of dystrophic neurites in
HTLE, which are also typically found in neuritic
plaques in Alzheimers disease (Griffin et al., 1995;
Ihara, 1988; Marshak et al., 1991). S-100 has a higher
affinity for zinc than for calcium (Baudier & Gerard,
1983), induces calcium inux which elevates [Ca
2
]
i
in
glial and neuronal cells (Barger &Van Eldik, 1992), and
promotes neurite outgrowth (Kligman & Marshak,
1985), as well as axonal and glial cell proliferation
(Reeves et al., 1994; Selinfreund et al., 1991). These
factors may be substantial enough to associate interac-
tions of zinc with neurotrophins which promote aber-
rant sproutings seen in HTLE hippocampus. Studies to
show whether zinc chelation affects seizure-induced
sprouting may prove useful.
Alzheimers Disease
Senile dementia exists in many forms. Alzheimers
disease (AD) is a commonly observed type of demen-
tia which afflicts millions worldwide. The disease is
progressive, exhibiting a widespread deterioration of
memory, intellect, will, self-care, and social functions
(Drachman & Leavitt, 1974; Rubin, 1990). Cortical
atrophy in the frontal and temporal cortices (especially
in the hippocampus) is the neuropathological hall-
mark of the disease (Ball, 1988). Furthermore, selective
loss of central cholinergic neurons in the brains of AD
patients is also observed (Davies & Maloney, 1976;
Whitemore et al., 1982; Wurtman et al., 1990). In the
basal nucleus system, the cholinergic receptor activity
is of the muscarinic type, is believed to be important in
memory, and is markedly affected in AD (Drachman &
Leavitt, 1974). Other ndings in AD brains include a
signicant accumulation of senile and diffuse plaques
and neurolament plaques (paired helical lament)
(Brun et al., 1990; Mattson, 1995a,b; Selkoe et al., 1990).
Insoluble amyloid- protein (A) is the main constitu-
ent of senile plaques and vascular tissue amyloid; it
typically has 3943 amino acid residues with a molecu-
lar mass of around 4 kDa (Mattson, 1995a; Selkoe et al.,
1990). It originates from cleavage of the amyloid
precursor protein (APP) by the putative proteinase
enzymes designated , , and secretases (Checler,
1995; Mattson, 1995b). Several genetic mutations iden-
tied from familial AD have been identied and
characterized (Levy-Lahad et al., 1995a,b; St. George-
Zinc Metabolism in the Brain 153
Copyright

1997 by Academic Press
All rights of reproduction in any form reserved.
Hyslop et al., 1987; Sherrington et al., 1995; Tanzi et al.,
1987).
a. Zinc Interactions with Amyloid Precursor
Protein Gene Expression and Metabolism
The genetic mutation that links Aand the causation
of one form of familial AD has been identied in
chromosome 21 (St. George-Hyslop et al., 1987; Tanzi et
al., 1987). It is believed that possible overexpression of
the APP gene has etiologic consequences for the AD
and AD-like neuropathology seen in Downs syn-
drome (trisomy 21) (for a review, see Ashall & Goate,
1994, and Hendriks & Van Broeckhoven, 1996). The
function of zinc in regulating the synthesis of APP is
thus of critical importance.
Zinc is important in gene transcription as evidenced
by its diverse presence in many transcription factors
(OHalloran, 1993; Rhodes & Klug, 1993; Schwabe &
Klug, 1996). Of these, APP synthesis is regulated by
zinc-containing transcription factors NF-B and Sp1.
The APP promoter contains two NF-B/Re1 (desig-
nated APB/APB) binding sites in its proximal (5-
regulatory) region, which, upon activation, increases
gene transcription by several-fold (Grilli et al., 1995;
Quitschke, 1994). Furthermore, Sp1 has been found to
bind to the positive regulatory region of the mouse
APP gene promoter which is highly homologous to the
human APP gene (Izumi et al., 1992). Zinc is easily
abstracted from NF-B and Sp1, but is essential for
their activity (Yang et al., 1995; Zabel et al., 1991; Zeng et
al., 1991a,b). Whether their activity in vivo is regulated
by zinc availability is hence relevant to the synthesis of
APP.
APP has a ligand-binding site for zinc (764 nM), and
zinc (50 M) inhibits degradation of APP and proteo-
lytic cleavage of A at its secretase site (Bush et al.,
1993, 1994b). While this concentration is high, it is
nevertheless well within the range of concentrations
which may occur under pathological conditions (dis-
cussed in previous sections). Thus, zinc may be a factor
in both stimulating the synthesis of APP and inhibiting
its subsequent degradation. The regulation by zinc of
the activities of NF-B and Sp1 and their subsequent
effects on APP expression should be fruitful areas of
research.
b. Zinc Interactions with Basement Membrane and
Matrix Metalloproteinase
Basement membranes (BMs) are distinct types of
extracellular matrices (ECMs) that support cells, pro-
vide ltration, and sequester growth factors (Yur-
chenco & Schittny, 1990). BMs have been implicated in
the pathogenesis of AD as APP isoforms were found to
interfere with the normal BM proteinprotein interac-
tions in vitro (Narindrasorasak et al., 1995). Several BM
proteins have been found to contain zinc-binding sites
(Reinhardt et al., 1993), as well as a zinc nger-like
structural motif (Narindrasorasak et al., 1992). Zinc (10
M) can stimulate the binding of laminin to several
APP isoforms in vitro (Narindrasorasak et al., 1992).
Likewise, zinc (1 M) enhances binding of APP to
heparin side chains of proteoglycans in vitro (Mul-
thaup et al., 1994). Recently, zinc (15 M) was observed
to effectively augment both lamininentactin and lam-
inincollagen type IV associations (Ancsin & Kisi-
levsky, 1996). This is the rst report of a function for
zinc as a cofactor/modulator of BM assembly and the
rst identication of high-affinity proteinprotein inter-
actions involving an extracellular zinc nger motif
(Ancsin & Kisilevsky, 1996).
Matrix metalloproteinases (MMPs) degrade and re-
structure the ECM (Baramova & Foidart, 1995;
Gottschall & Yu, 1995). MMPs are a family of enzymes
(collagenases and gelatinases) whose activities are
dependent on the presence of zinc in their catalytic
sites (Baramova & Foidart, 1995; Gottschall & Yu,
1995). In vitro, zinc has been shown to completely
inhibit two types of MMPs (MMP-2 and MMP-9)
found in the CNS (Backstrom et al., 1992). However,
the potency of zinc as an inhibitor is unknown (only
one high concentration of zinc was tested). MMP-2
degrades soluble forms of A
140
/A
142
, but not in-
soluble aggregates of A
140
/A
142
(Roher et al., 1994).
MMP-9 is also capable of degrading synthetic A
140
peptides (Backstrom et al., 1996). Thus, regulation of
MMP activity by zinc could also affect the degradation
of A, a possible initiator of neuronal death.
Further investigations are required in order to deter-
mine whether zinc plays a major role in the changes
that occur between APP metabolism and its interac-
tions with BMproteins or with the metabolismof BMs,
and any association of these activities with amyloid
deposition.
c. Zinc, -Amyloid, and Oxidative Stress
The aggregated form of A is cytotoxic (Iversen et
al., 1995) and can induce apoptosis both in vivo (LaFerla
et al., 1995) and in vitro, although this action is depen-
dent on the type of cultured cells used (Gschwind &
Huber, 1995). A toxicity can trigger activation of the
stress-related transcription factor NF-B (Behl et al.,
154 Cuajungco and Lees
Copyright

1997 by Academic Press
All rights of reproduction in any form reserved.
1994). Recent studies have implicated oxidative injury
to mitochondrial DNA damage and malfunctions in
the mitochondrial respiration chains as key players in
the pathogenesis of AD (Hutchin & Cortopassi, 1995;
Mecocci et al., 1994; Mutisya et al., 1994), as well as to
the mediation of A toxicity in vitro (Behl et al., 1994).
The relevance of oxidative stress to the release of zinc
from MT has been discussed previously.
Several metal ions such as iron and aluminum
(Mantyh et al., 1993), as well as zinc (Bush et al.,
1994a,b, 1995, 1996; Clements et al., 1996; Esler et al.,
1996; Fuson et al., 1996; Mantyh et al., 1993), induce A
protein aggregation. Of these metal ions, it is claimed
that only zinc can induce A precipitates at low,
physiological concentrations 10
6
M (Bush et al.,
1994a,b, 1995, 1996; Fuson et al., 1996). However, other
recent studies do not support the initial report of Bush
and colleagues (e.g., see Clements et al., 1996; Esler et
al., 1996), and the subject is rather controversial (e.g.,
see Kaiser, 1994). It is interesting to note that zinc-
mediated A
140
precipitation per se is not sufficient for
neurodegeneration as co-addition of zinc in rat hippo-
campal primary cultures specically and in a dose-
dependent manner reduced A
140
neurotoxicity (Fu-
son et al., 1996).
Postmortem studies report that zinc is signicantly
reduced in the temporal lobes including the hippocam-
pus, as well as in the inferior parietal and primary
visual cortices of AD brain (Andrasi et al., 1995;
Constantinidis, 1990, 1991; Corrigan et al., 1993; Deng
et al., 1994; Ward & Mason, 1987; Wenstrup et al., 1990).
On the other hand, other investigations report no
change (Basun et al., 1991; Wenstrup et al., 1990) or an
increase in specic brain areas (Danscher et al., per-
sonal communication; Deibel et al., 1996; Samudralwar
et al., 1995; Thompson et al., 1988). Thus, the specicity
and location in brain regions/neurons of any changes
in zinc need further work. It is not known if zinc
depletion is simply due to a reduction in number of
zinc-containing neurons (i.e., a later-stage event) or
whether it is the cause of neuronal death found at
specic areas of AD brains. Observations indicating a
decrease in brain zinc concentration contrast with the
postulated role of zinc inAPPmetabolism, Aaggrega-
tion, and abnormal BM metabolism. This paradox
could be resolved if the observed zinc reduction in AD
brain is a secondary effect of neurodegenerative pro-
cesses (Cuajungco & Lees, 1997). There is also a
controversy regarding the observed levels of MT-III in
AD brains (e.g., see Tsuji et al., 1992, and Uchida et al.,
1991, versus Erickson et al., 1994).
d. Zinc and EF-Hand S-100 Protein
S-100 protein is an isoform localized in the brain
(Baudier & Gerard, 1983; Baudier et al., 1983; Kligman
& Marshak 1985). S-100 has a greater affinity for zinc
than calcium, and zinc increases its affinity for calcium
by one order of magnitude (Baudier & Gerard, 1983).
The physical binding of zinc or calcium produces
distinct conformational changes in the proteins struc-
ture (Baudier & Gerard, 1983; Calissano et al., 1974). It
is not known whether the conformational changes
induced by either ion are necessary for the normal
functions of the protein. S-100 is mainly observed in
glial cells and is believed to regulate their proliferation
(Reeves et al., 1994; Selinfreund et al., 1991). However,
it was also suggested to be involved in neuronal
differentiation and maturation (Azmitia et al., 1990;
Van Eldik et al., 1991; Yang et al., 1996).
Several lines of evidence have implicated S-100 in
the pathogenesis of neurological dysfunctions in AD
and Downs syndrome (DS) (Allore et al., 1988; Griffin
et al., 1989; Marshak et al., 1991; Sheng et al., 1996; Van
Eldik & Griffin, 1994). S-100, interleukin-1 (IL-1), and
glial brillary acidic protein (GFAP) are increased in
the temporal lobes of AD and DS brains where neuritic
plaques are densely located (Marshak et al., 1991;
Griffin et al., 1989). The observed elevation of S-100
content in the brain of rats (Kato et al., 1990) and
humans (Sheng et al., 1996) is closely correlated with
aging and age-dependent changes in the CNS, lending
support for a role of S-100 in AD neuropathology. On
the contrary, elevation of S-100 has no effect on APP
mRNA induction nor does it promote A deposits in
the brains of aging transgenic mice (Yao et al., 1995).
e. AD and Protein Kinases
The role of zinc in regulating the activity of PKC and
the relationship between down-regulation of PKC and
apoptosis has already been described. PKC activation
has been found to modulate APP degradation, thus
inhibiting Aprotein formation (Buxbaum et al., 1993;
Gabuzda et al., 1993; Hung et al., 1993). The regulation
of APP secretion by PKC activity was found to be
defective in cultured neurons (Wang et al., 1994) and
broblasts (Bergamaschi et al., 1995) obtained fromAD
patients. Both cytosolic and membrane-associated PKC
responses were attenuated in frontal and temporal
cortices, including the hippocampi of ADbrains (Wang
et al., 1994). Further evidence comes from a recent
report that both Ca
2
-dependent and Ca
2
-indepen-
dent PKC activities in cytosolic and membranous
fractions obtained from AD specimens were signi-
Zinc Metabolism in the Brain 155
Copyright

1997 by Academic Press
All rights of reproduction in any form reserved.
cantly decreased (Matsushima et al., 1996). The tyro-
sine kinase trk proteins are another group of protein
kinases which are signicantly reduced in AD hippo-
campi (Vener et al., 1993). Zinc (0.5 mM, or magnesium
at 10 mM) added to membrane fractions of the AD
hippocampus activates the phosphorylation of trk
proteins (Vener et al., 1993). It is conceivable then that a
decreased zinc content in the AD hippocampus (if
conrmed) could produce defective PKC and/or pro-
tein trk activities (Vener et al., 1993). A future study
would be to determine if addition of zinc to membrane
fractions taken from specic areas of AD brain would
lead to a recovery of PKC activity.
Amyotrophic Lateral Sclerosis
Amyotrophic lateral sclerosis (ALS) is a neurodegen-
erative disorder that affects motor neurons in the
spinal cord, brain stem, and cerebral (motor) cortex
(Olanow, 1993). Growing evidence suggests that ALS
pathology involves free radical toxicity (Olanow, 1993;
Przedborski et al., 1996).
Even though the etiology of sporadic ALS is still
unknown, there are approximately 510% reported
familial cases of this disease (FALS). Agenetic defect in
the superoxide dismutase type 1 (SOD1) gene of FALS
was discovered in chromosome q21 which codes for
Cu,Zn-SOD (Rosen et al., 1993). It is not known,
however, how this particular mutation may lead to
FALS symptomatology and pathology. Recent studies
suggest that FALS may not be a result of decreased
enzymic function owing to mutation in the SOD1 gene,
but due to a gain-of-function by the enzyme (Rabiza-
deh et al., 1995; Yimet al., 1996). In vitro studies showed
that the SOD1 mutation is proapoptotic in both neu-
rons (Rabizadeh et al., 1995) and immune cells (Peled-
Kamar et al., 1995).
Cu,Zn-SOD is a cytosolic antioxidant that catalyzes
the dismutation of O
2

to H
2
O
2
and O
2
, and it repre-
sents the rst line of defense against oxidative toxicity
(Olanow, 1993). There are two zinc atoms and two
copper atoms in the enzyme, and all four ions are
necessary for optimal SOD activity (Willson, 1989).
Zinc is believed to have a structural function, while
copper is involved in the catalytic sites of dismutation
(Willson, 1989). SOD is structurally very stable and
remains active even at high (8 M) urea concentrations
(Willson, 1989). SOD has a relatively low affinity for
zinc (K
D
10
5
M) (Fabris & Mocchegiani, 1995).
Nonetheless, the metal chelators 1,10-phenanthroline
(K
D
Zn
10
17
M, pH 5.5) and EDTA (K
D
Zn
10
16
M,
pH 7.4), chelate only 13 and 10% of Zn
2
in SOD,
respectively (Willson, 1989), which indicates that most
of the structural zinc is inaccessible to zinc chelators.
Very recently, the effect of the mutation on the binding
of the metal constituents of the enzyme was reported
(Lyons et al., 1996). It was found that SOD1 mutation
affects the native zinc-binding sites of the enzyme,
which may possibly result in an unstable protein
structure.
MMP-2 and MMP-9 activities were recently reported
to be signicantly different in brain and spinal cord of
ALS patients compared with controls (Lim et al., 1996).
MMP-9 activity was markedly elevated in the frontal
and occipital cortices, particularly in the motor cortex,
and in the thoracic and lumbar regions of the spinal
cord of ALS sufferers (Lim et al., 1996). On the other
hand, MMP-2 activity in the motor cortex was signi-
cantly reduced. As zinc may inhibit both MMP-2 and
MMP-9 activities in vitro (Backstrom et al., 1992),
evidence for an abnormal zinc metabolism in ALS
would not explain these observed changes.
PKC activity in ALS spinal cord is signicantly
increased, possibly attributable to abnormal Ca
2
buff-
ering (for a recent review, see Krieger et al., 1996) or
aberrant zinc levels which could inuence its activity
(Csermely et al., 1988; Murakami et al., 1987; Zalewski
et al., 1990, 1991). Note that MT expression is signi-
cantly increased in the spinal cord gray matter of
protoplasmic astrocytes (Sillevis-Smitt et al., 1992a), as
well as in liver and kidney (Sillevis-Smitt et al., 1992b)
of ALS patients. These reports could mean that either
metal ion levels are elevated or an increased oxidative
stress is present (reviewed in Sillevis-Smitt et al., 1994).
The observations presented here suggest some associa-
tion of metal cations and oxidative injury in the
pathogenesis of ALS. Hence, investigations need to be
done to determine whether abnormal zinc or copper
function in mutant SOD has any relevance to disease
progression.
GuamALS-Parkinsonism Dementia
Guamanian ALS-PD is a subtype of ALS. The pres-
ence of cycad (genus Cycas) in food or medicine has
been implicated in the pathogenesis of ALS-PD of
Guam (Duncan et al., 1992). It was found that cycad
our samples produce a signicant but dose-depen-
dent cytotoxicity to mesencephalic and cerebellar gran-
ule cell cultures (Duncan et al., 1992). Zinc was found
to be the neurotoxic constituent which contaminates
cycad our, and similar zinc levels (100 M) calculated
to be present in the our samples also produced a
156 Cuajungco and Lees
Copyright

1997 by Academic Press
All rights of reproduction in any form reserved.
concentration-dependent toxicity on cultured mesence-
phalic and cerebellar granule cells. Addition of EDTA
(200 M) in the culture medium conferred protection
against zinc neurotoxicity. Meanwhile, zinc levels in
the gray matter of the frontal cortex of Guam ALS-PD
brains were reported to be markedly reduced com-
pared with normal controls (Yasui et al., 1993). Further
investigations are necessary in order to understand the
relationship between alterations in zinc metabolism
and exacerbation of the disease.
Parkinsons Disease
As with ALS, oxidative stress may be an etiologic
factor in Parkinsons disease (PD) (Jenner et al., 1992).
According to Dexter et al. (1989a), postmortem studies
revealed that polyunsaturated fatty acid levels in the
substantia nigra (SN) of patients with PD were re-
duced in comparison with other brain regions and that
of control (normal) tissue. Moreover, malondialdehyde
levels were elevated in the SN, suggesting occurrence
of lipid peroxidation (Dexter et al., 1989a). Zinc and
both cytoplasmic and mitochondrial SOD activities in
SN of patients with PD were also markedly increased
(Dexter et al., 1989b, 1991, 1992; Jenner et al., 1992),
while GSH levels were reduced in both early- and
late-stage PD (Jenner et al., 1992; Soc et al., 1992). In
contrast, there were no signicant alterations in the
activity of H
2
O
2
-degrading enzymes (catalase and
GPO) or in the concentrations of free radical scaven-
gers (ascorbic acid and -tocopherol) (Jenner et al.,
1992), as well as GSSG (Soc et al., 1992). The SN
contains a collection of neurons that utilize dopamine
as their major transmitter and which are selectively
lost in PD. These dopaminergic neurons also contain
melanin, an autoxidation by-product of dopamine
metabolism. Nonmelanized neurons are believed to be
less susceptible to oxidative stress than melanized
cells, since they are not likely to experience iron
melanin interactions that could produce excessive free
radicals (Ben-Shachar et al., 1991; Hirsch, 1992). Metabo-
lism of catecholamines such as DA autoxidation pro-
duces ROS and free radicals, thus making nigral cells
more susceptible to oxidative stress (Ben-Shachar et al.,
1991; Hirsch, 1992). Intracerebral injection of 6-hy-
droxydopamine in rats, a drug known to produce free
radicals, resulted in a reduction of zinc and MT
concentrations in the striatum, but not in any other
brain areas tested (Shiraga et al., 1993). Shiraga et al.
(1993) postulated that MT may reduce the effects of
generating ROS and free radicals by releasing zinc to
neuronal membrane and/or receptors, suggesting an
antioxidant role for both zinc and MT (Bray & Bettger,
1990; Sato & Bremner, 1993). Further, zinc may prevent
excessive NO

production as it inhibits nNOS activity


(one Zn
2
per NOS molecule) (Persechini et al., 1995).
One possible source of zinc is its release fromMT, since
cells have been estimated to contain approximately 13
M zinc-bound MT in vitro (Krezoski et al., 1988).
HIV Neuropathy/AIDS Dementia Complex
The mechanisms of human immunodeciency virus
(HIV) neuropathy and acquired immunodeciency
syndrome (AIDS) dementia complex (ADC) are not
fully understood. ADC is progressive and is character-
ized by memory impairment, motor decits, organic
psychosis, and other behavioral dysfunctions (Navia et
al., 1986). There is a great deal of neuropathological
variability in patients with CNS HIV-1 infection and
ADC (Price et al., 1988). The major brain areas that
show abnormalities are central white matter and deep
gray matter, including the basal ganglia, thalamus,
brain stem, and spinal cord, with the exception of
relative sparing of the cerebral cortex (Navia et al.,
1986). The potential involvement of zinc-dependent
processes on HIV neuropathy and ADC is on three
fronts: (i) zinc-dependent synthesis of, or binding to,
cytokines; (ii) oxidative stress-mediatedneuronal death;
and (iii) regulation of viral replication.
An abundance of reactive microglial cells is com-
monly found in HIV-infected brain, as well as intense
IL-1 immunoreactivity (Stanley et al., 1994). Reactive
astroglial cells with up-regulated S-100 expression
are also observed (Stanley et al., 1994). Generally, glial
changes are associated with an elevated -APP expres-
sion in neurons and overgrown neurites, including
detection of tau-2 immunoreactivity among NFT-like
formations (Stanley et al., 1994). Not surprisingly, the
histopathological detection of NFTs and reactive glial
cells, as well as marked elevations of IL-1 and S-100
inADCsuggest a close association with the correspond-
ing neuropathological changes in both AD and DS
pathologies (Griffin et al., 1989). Although the above
observations suggest a direct relationship between
enhanced expression of cytokines and neurodegenera-
tion, it remains to be proven whether these entities
provide a primary involvement in the pathogenesis of
neurological dysfunctions. The physiological role of
zinc in AIDS and its importance in the immune
function have been reviewed (Odeh, 1992; Ripa &
Ripa, 1995). There is a signicant reduction in the
Zinc Metabolism in the Brain 157
Copyright

1997 by Academic Press
All rights of reproduction in any form reserved.
plasma zinc level among asymptomatic HIV patients,
which suggests that deciencies in trace metals and
antioxidants are not secondary effects of malnutrition
commonly found in the late AIDS stage (Favier et al.,
1994). Zinc deciency, on the other hand, could also be
due to an intense immune response against HIV,
because endotoxin- and viral-induced immune/inam-
matory reactions result in low serum zinc (Bremner &
May, 1989; Goldblumet al., 1987; Taubeneck et al., 1995)
and iron (Goldblum et al., 1987). Zinc deciency may
also be due to utilization by HIV during replication
(see below). In contrast, a previous study did not show
any association between HIV infection and zinc status
(Walter et al., 1990).
HIV type 1 (HIV-1) infection has been linked with
oxidative stress-mediated cell death (McCord &Flores,
1994). Evidence comes from ndings that GSH levels
(Staal et al., 1992) and other antioxidants (Favier et al.,
1994) were signicantly decreased in tissues of in-
fected patients. Such condition may increase the intra-
cellular zinc pool as disuldeMT interactions are
known to liberate zinc from MT (Maret, 1994, 1995).
In regards to the involvement of zinc in the replica-
tion of the HIV virus, zinc (10 M) was recently found
to stimulate the Mg
2
-dependent 3-processing activ-
ity of HIV-1 integrase in vitro (Lee & Han, 1996). Note
also that zinc is an essential component of retroviral
nucleocapsid protein which is present in viral RNA
genomes (Green & Berg, 1990). Zinc is also important
for the activity of gag polyprotein, which serves to
identify and formviral RNAgenome complexes (Sum-
mers et al., 1990). The nal mature viral product is
HIV-1 nucleocapsid protein type 7 (NCp7), a zinc-
dependent protein (Mely et al., 1996). The importance
of zinc in HIV replication was demonstrated when
HIV activity and infectivity in vitro were inhibited by
ejecting zinc from intact HIV-1 virions and NCp7,
using disulde benzamides and C-nitroso compounds
(Rice et al., 1993; Tummino et al., 1996). Furthermore,
zinc is an essential component of NF-B, which regu-
lates the transcription of immuno-receptors, cytokines,
and viruses (Baeuerle, 1991). The zinc-dependent,
oxidoreductive activation of NF-B protein stimulates
viral replication by binding to HIV long terminal
repeats (LTRs) (Baeuerle, 1991), although it is not
essential as other transcription promoters are involved
(Leonard et al., 1989). Deletion of two NF-B-binding
sites and the three Sp1 sites, or mutation of the
tat-responsive region, completely abolished viral repli-
cation (Leonard et al., 1989). TNF- has been found to
independently instigate HIV-1 transcription, and it
acts in synergy with the presence of mitogens in vitro,
via the enhancer element of the HIV-LTR NF-B-
binding region (Israel et al., 1989). Further, TNF- was
shown to induce zinc deciency and teratogenicity in
developing mice (Taubeneck et al., 1995).
The mechanism of damage in cerebral ischemia is
also closely associated with the possible pathogenesis
of HIV neuropathy and ADC. The glycoprotein coat of
HIV-1 (gp120) was found to increase [Ca
2
]
i
in astro-
cytes (Codazzi et al., 1996) and increase NO

levels
among cortical neurons in vitro (Dawson et al., 1993).
HIV-1 gp120 induces apoptosis in hippocampal neu-
rons and interneurons in vitro (Aggoun-Zouaoui et al.,
1996) and in cortical neurons in vivo (Bagetta et al.,
1996; Charriaut-Marlangue et al., 1996). NO

-mediated
toxicity may be one of the underlying mechanisms of
death in both HIV neuropathy and ADC neuropathol-
ogy, since inhibition of NOS abolished the cytotoxic
effects of gp120 (Dawson et al., 1993). Likewise, inhibi-
tion of NOS by L-NAME reduces the extracellular
glutamate concentration in the rat hippocampus
(Rigaud-Monnet et al., 1995). Thus, zinc depletion
could be detrimental as it is a potent inhibitor of nNOS
activity (Persechini et al., 1995).
Provided that the chelation of zinc is not too toxic for
human cells (e.g., due to inhibition of nNOS activity),
the above results indicate that chelation of zinc from
HIVproteins could be effective in reducing the infectiv-
ity of the virus.
CONCLUDING REMARKS
Zinc is critical for the growth and survival of cells.
However, an abnormal metabolism of zinc in cells can
have deleterious effects. The evidence presented here
indicates that while the importance of changes in zinc
metabolism in causing neurological dysfunctions can
be debated, zinc does seem to play a role in ischemic-
related neuropathology. Although there is a high asso-
ciation with zinc and epilepsy, some contentious issues
need to be resolved. Studies have shown that zinc can
induce cytotoxicity in nervous and nonnervous tissue.
Zinc was also shown to play a role in apoptotic death
of nonneuronal cells. Whether zinc also serves any
function in neuronal apoptosis needs to be elucidated.
The involvement of zinc in Alzheimers and other
neurodegenerative diseases is far from being demon-
strated, but is currently the subject of thorough investi-
gation. Whether zinc is directly responsible in the
etiologic processes of these disorders or is a secondary
consequence of the disease process remains to be
proven. Therefore, further investigations are needed in
158 Cuajungco and Lees
Copyright

1997 by Academic Press
All rights of reproduction in any form reserved.
order to determine if there is any role for zinc in
dysfunctions of the central nervous system. The knowl-
edge gained from investigations of the mechanism of
zinc-induced cell death may enable researchers to
develop therapeutic interventions that would reduce
or perhaps prevent further destructive processes occur-
ring in the brain.
ACKNOWLEDGMENTS
We thank the NZ Lotteries Grant Office, the University of
Auckland Staff Research Fund, and the Health Research Council of
NZ for providing nancial support to G.J.L. M.P.C. is a recipient of
the Miller scholarship fromthe NZ Neurological Foundation. We are
grateful to Professor Richard Palmiter and Professor GormDanscher
for providing results prior to publication, and we thank the Ameri-
can Association for the Advancement of Science for allowing us to
reproduce published materials.
REFERENCES
Aggoun-Zouaoui, D., Charriaut-Marlangue, C., Rivera, S., Jorquera,
I., Ben-Ari, Y., & Represa, A. (1996) The HIV-1 envelope protein
gp120 induces neuronal apoptosis in hippocampal slices. Neuro-
Report 7, 433436.
Allore, R., OHanlon, D., Price, R., Neilson, K., Willard, H. F., Cox,
D. R., Marks, A., &Dunn, J. N. (1988) Gene encoding the subunit
of S-100 protein is on chromosome 21: Implications for Down
syndrome. Science 239, 13111313.
Ancsin, J. B., & Kisilevsky, R. (1996) Laminin interactions important
for basement membrane assembly are promoted by zinc and
implicate laminin zinc nger-like sequences. J. Biol. Chem. 271,
68456851.
Anderegg, G., Hubmann, E., Podder, N. G., & Wenk, F. (1977)
Pyridine derivatives as complexing agents XI: Thermodynamics
of metal complex formation with bis-, tris-, and tetrakis[(2-
pyridyl)methyl]-amines. Helv. Chim. Acta 60, 123140.
Andrasi, E., Farkas, E., Scheibler, H., Reffy, A., & Bezur, L. (1995) Al,
Zn, Cu, Mn and Fe levels in brain in Alzheimers disease. Arch.
Gerontol. Geriatr. 21, 8997.
Aniksztejn, L., Charton, G., & Ben-Ari, Y. (1987) Selective release of
endogenous zinc from the hippocampal mossy bers in situ. Brain
Res. 404, 5864.
Ankarcrona, M., Dypbukt, J. M., Bonfoco, E., Zhivotovsky, B.,
Orrenius, S., Lipton, S. A., & Nicotera, P. (1995) Glutamate-
induced neuronal death: A succession of necrosis or apoptosis
depending on mitochondrial function. Neuron 15, 961973.
Arslan, P., Di Virgilio, F., Beltrame, M., Tsien, R., & Pozzan, T. (1985)
Cytosolic Ca
2
homeostasis in Ehrlich and Yoshida carcinomas. J.
Biol. Chem. 260, 27192727.
Ashall, F. & Goate, A. M. (1994) Role of the -amyloid precursor
protein in Alzheimers disease. Trends Biochem. Sci. 19, 4246.
Assaf, S. Y. & Chung, S. H. (1984) Release of endogenous Zn
2
from
brain tissue during activity. Nature 308, 734736.
Atar, D., Backx, P. H., Appel, M. M., Gao, W. D., & Marban, E. (1995)
Excitationtranscription coupling mediated by zinc inux through
voltage-dependent calciumchannels. J. Biol. Chem. 270, 24732477.
Azmitia, E. C., Dolan, K., & Whitaker-Azmitia, P. M. (1990) S-100
but not NGF, EGF, insulin, or calmodulin is a CNS serotonergic
growth factor. Brain Res. 516, 354356.
Baba, A., Etoh, S., & Iwata, H. (1991) Inhibition of NMDA-induced
protein kinase C translocation by a Zn
2
chelator: Implication of
intracellular Zn
2
. Brain Res. 557, 103108.
Baba, A., Kihara, T., Sawada, T., & Iwata, H. (1989) Excitatory amino
acids enhance dissociation of zinc from soluble protein in cytosol
of rat hippocampus. Brain Res. 486, 372375.
Babb, T. L., Kupfer, W. R., Pretorius, J. K., Crandall, P. H., &
Levesque, M. F. (1991) Synaptic reorganization by mossy bers in
human epileptic fascia dentata. Neuroscience 42, 351363.
Backstrom, J. R., Lim, G. P., Cullen, M. J., &Tokes, Z. A. (1996) Matrix
metalloproteinase-9 (MMP-9) is synthesized in neurons of the
human hippocampus and is capable of degrading the amyloid-
peptide (140). J. Neurosci. 16, 79107919.
Backstrom, J. R., Miller, C. A., & Tokes, Z. A. (1992) Characterization
of neutral proteinases from Alzheimer-affected and control brain
specimens: Identication of calcium-dependent metalloprotein-
ases from the hippocampus. J. Neurochem. 58, 983992.
Baeuerle, P. A. (1991) The inducible transcription activator NF-B:
Regulation by distinct protein subunits. Biochim. Biophys. Acta
1072, 6380.
Bagetta, G., Corasaniti, M. T., Aloe, L., Berliocchi, L., Costa, N.,
Finazzi-Agro, A., & Nistico, G. (1996) Intracerebral injection of
human immunodeciency virus type 1 coat protein gp120 differen-
tially affects the expression of nerve growth factor and nitric oxide
synthase in the hippocampus of rat. Proc. Natl. Acad. Sci. USA 93,
928933.
Ball, M. J. (1988) Hippocampal histopathologyA critical substrate
for dementia of the Alzheimer type. In: Interdisciplinary Topics in
Gerontology: Histology and Histopathology of the Aging Brain, (J.
Ulrich, Ed.), pp. 1637. Karger, Basel.
Banerjee, A., Roychowdhury, S., & Bhattacharyya, B. (1982) Zinc-
induced self-assembly of goat brain tubulin: Some novel aspects.
Biochem. Biophys. Res. Commun. 105, 15031510.
Baramova, E., & Foidart, J. M. (1995) Matrix metalloproteinase
family. Cell Biol. Int. 19, 239242.
Barbieri, D., Troiano, L., Grassilli, E., Agnesini, C., Cristofalo, E. A.,
Monti, D., Capri, M., Cossarizza, A., & Franceschi, C. (1992)
Inhibition of apoptosis by zinc: A reappraisal. Biochem. Biophys.
Res. Commun. 187, 12561261.
Barger, S. W., & Van Eldik, L. J. (1992) S-100 stimulates calcium
uxes in glial and neuronal cells. J. Biol. Chem. 267, 96899694.
Barker, J. E., Heales, S. J. R., Cassidy, A., Bolanos, J. P., Land, J. M., &
Clark, J. B. (1996) Depletion of brain glutathione results in a
decrease of glutathione reductase activity, an enzyme susceptible
to oxidative damage. Brain Res. 716, 118122.
Basun, H., Forssell, L. G., Wetterberg, L., &Winbald, B. (1991) Metals
and trace elements in plasma and cerebrospinal uid in normal
ageing and Alzheimers disease. J. Neural Transm. Parkinsons Dis.
Dementia Sect. 4, 231258.
Baudier, J., & Gerard, D. (1983) Ions binding to S-100 proteins:
Structural changes induced by calcium and zinc on S-100 and
S-100 proteins. Biochem. 22, 33603369.
Baudier, J., Haglid, K., Haiech, J., & Gerard, D. (1983) Zinc ion
binding to human brain calcium-binding proteins calmodulin and
S-100 protein. Biochem. Biophys. Res. Commun. 114, 11381146.
Beaulieu, C., Dyck, R., & Cynader, M. (1992) Enrichment of gluta-
mate in zinc-containing terminals of the cat visual cortex. Neuro-
Report 3, 861864.
Zinc Metabolism in the Brain 159
Copyright

1997 by Academic Press
All rights of reproduction in any form reserved.
Behl, C., Davis, D. R., Lesley, R., & Schubert, D. (1994) Hydrogen
peroxide mediates amyloid protein toxicity. Cell 77, 817827.
Beilharz, E. J., Williams, C. E., Dragunow, M., Sirimanne, E. S., &
Gluckman, P. D. (1995) Mechanisms of delayed cell death follow-
ing hypoxicischemic injury in the immature rat: Evidence for
apoptosis during selective neuronal loss. Mol. Brain Res. 29, 114.
Ben-Shachar, D., Riederer, P., & Youdim, M. B. H. (1991) Iron
melanin interaction and lipid peroxidation: Implications for Par-
kinsons disease. J. Neurochem. 57, 16091614.
Bergamaschi, S., Binetti, G., Govoni, S., Wetsel, W. C., Battaini, F.,
Trabucchi, M., Bianchetti, A., & Racchi, M. (1995) Defective
phorbol ester-stimulated secretion of -amyloid precursor protein
fromAlzheimers disease broblasts. Neurosci. Lett. 201, 14.
Bettger, W. J., & ODell, B. L. (1981) A critical physiological role of
zinc in the structure and function of biomembrane. Life Sci. 28,
14251438.
Blair-West, J. R., Denton, D. A., Gibson, A. P., &McKinely, M. J. (1990)
Opening the bloodbrain barrier to zinc. Brain Res. 507, 610.
Bonfoco, E., Krainc, D., Ankarcrona, M., Nicotera, P., & Lipton, S. A.
(1995) Apoptosis and necrosis: Two distinct events induced,
respectively, by mild and intense insults with N-methyl-D-
aspartate or nitric oxide/superoxide in cortical cell cultures. Proc.
Natl. Acad. Sci. USA 92, 71627166.
Bramham, C. R., Torp, R., Zhang, N., Storm-Mathisen, J., & Ottersen,
O. P. (1990) Distribution of glutamate-like immunoreactivity in
excitatory hippocampal pathways: A semiquantitative electron
microscopic study in rats. Neuroscience 39, 405417.
Bray, T. M., &Bettger, W. J. (1990) The physiological role of zinc as an
antioxidant. Free Radicals Biol. Med. 8, 281291.
Bremner, I., &May, P. M. (1989) Systemic interactions of zinc. In: Zinc
in Human Biology, (C. F. Mills, Ed.), pp. 95108. Springer-Verlag,
London.
Bresink, I., Ebert, B., Parsons, C. G., & Mutschler, E. (1996) Zinc
changes AMPAreceptor properties: Results of binding studies and
patch clamp recordings. Neuropharmacology 35, 503509.
Brewer, G. J. (1980) Calmodulin, zinc, and calcium in cellular and
membrane regulation: An interpretive review. Am. J. Hemat. 8,
231248.
Brewer, G. J., Aster, J. C., Knutsen, C. A., & Kruckeberg, W. C. (1979)
Zinc inhibition of calmodulin: A proposed molecular mechanism
of zinc action on cellular functions. Am. J. Hemat. 7, 5360.
Browning, J. D., & ODell, B. L. (1995) Zinc deciency decreases the
concentration of N-methyl-D-aspartate receptors in guinea pig
cortical synaptic membranes. J. Nutrition 125, 20832089.
Brun, A., Gustafson, L., & Englund, E. (1990) Subcortical pathology
of Alzheimers disease. In: Advances in Neurology (R. J. Wurtman,
S. Corkin, J. H. Growdon, & E. Ritter-Walker, Eds.), pp. 7377.
Raven Press, NewYork.
Buhl, E. H., Otis, T. S., & Mody, I. (1996) Zinc-induced collapse of
augmented inhibition by GABA in a temporal lobe epilepsy
model. Science 271, 369373.
Bush, A. I., Atwood, C. S., Huang, X., Moir, R. D., Glabe, C. G., &
Tanzi, R. E. (1996) Abnormal homeostasis of ph and zinc: The
prelude for cerebral A amyloid formation? Eur. Neuropsychophar-
macol. 6 (Supplement 3), 20(S-4-2).
Bush, A. I., Moir, R. D., Rosenkranz, K. M., & Tanzi, R. E. (1995) Zinc
and Alzheimers disease. Science 268, 19211923.
Bush, A. I., Multhaup, G., Moir, R. D., Williamson, T. G., Small, D. H.,
Rumble, B., Pollwein, P., Beyreuther, K., & Masters, C. L. (1993) A
novel zinc(II) binding site modulates the function of the A4
amyloid protein precursor of Alzheimers disease. J. Biol. Chem.
268, 1610916112.
Bush, A. I., Pettingel, W. H., Multhaup, G., Paradis, M. D., Vonsattel,
J. P., Gusella, J. F., Beyreuther, K., Masters, C. L., & Tanzi, R. E.
(1994a) Rapid induction of Alzheimer A amyloid formation by
zinc. Science 265, 14641467.
Bush, A. I., Pettingel, W. H., Paradis, M. D., & Tanzi, R. E. (1994b)
Modulation of A adhesiveness and secretase site cleavage by
zinc. J. Biol. Chem. 269, 1215212158.
Buxbaum, J. D., Koo, E. H., &Greengard, P. (1993) Protein phosphor-
ylation inhibits production of Alzheimer amyloid /A4 peptide.
Proc. Natl. Acad. Sci. USA 90, 91959198.
Calissano, P., Alema, S., & Fasella, P. (1974) Interaction of S-100
protein with cations and liposomes. Biochem. 13, 45534560.
Cano-Gauci, D. F., & Sarkar, B. (1996) Reversible zinc exchange
between metallothionein and the estrogen receptor zinc nger.
FEBS Lett. 386, 14.
Casanovas-Aguilar, C., Christensen, M. K., Reblet, C., Martinez-
Garcia, F., Perez-Clausell, J., & Bueno-Lopez, J. L. (1995) Callosal
neurones give rise to zinc-rich boutons in the rat visual cortex.
NeuroReport 6, 497500.
Charriaut-Marlangue, C., Aggoun-Zouaoui, D., Represa, A., & Ben-
Ari, Y. (1996) Apoptotic features of selective neuronal death in
ischemia, epilepsy, and gp120 toxicity. Trends Neurosci. 19, 109114.
Charton, G., Rovira, C., Ben-Ari, Y., & Leviel, V. (1985) Spontaneous
and evoked release of endogenous Zn
2
in the hippocampal
mossy ber zone of the rat in situ. Exp. Brain Res. 58, 202205.
Checler, F. (1995) Processing of the -amyloid precursor protein and
its regulation in Alzheimers disease. J. Neurochem. 65, 14311444.
Chesters, J. K., Petrie, L., & Lipson, K. E. (1993) Two zinc-dependent
steps during G1 to S phase transition. J. Cell. Physiol. 155, 445451.
Chesters, J. K., Petrie, L., & Vint, H. (1989) Specicity and timing of
the Zn
2
requirement for DNAsynthesis by 3T3 cells. Exp. Cell Res.
184, 499508.
Choi, D. W. (1988) Calcium-mediated neurotoxicity: relationship to
specic channel types and role in ischemic damage. Trends
Neurosci. 11, 465469.
Choi, D. W. (1992) Bench to bedside: The glutamate connection.
Science 258, 241243.
Choi, D. W., Yokoyama, M., & Koh, J. (1988) Zinc neurotoxicity in
cortical cell culture. Neuroscience 24, 6779.
Christensen, M. K., Frederickson, C. J., & Danscher, G. (1992)
Retrograde tracing of zinc-containing neurons by selenide ions: A
survey of seven selenium compounds. J. Histochem. Cytochem. 40,
575579.
Chuah, M. I., Tennet, R., & Jacobs, I. (1995) Response of olfactory
Schwann cells to intranasal zinc sulfate irrigation. J. Neurosci. Res.
42, 470478.
Chung, S. H., & Johnson, M. S. (1983) Divalent transition-metal ions
(Cu
2
and Zn
2
) in the brains of epileptogenic and normal mice.
Brain Res. 280, 323334.
Clements, A., Allsop, D., Walsh, D. M., & Williams, C. H. (1996)
Aggregation and metal-binding of mutant forms of the amyloid
A peptide of Alzheimers disease. J. Neurochem. 66, 740747.
Cloues, R. (1995) Properties of ATP-gated channels recorded fromrat
sympathetic neurons: Voltage dependence and regulation by Zn
2
ions. J. Neurophysiol. 73, 312319.
Cloues, R., Jones, S., & Brown, D. A. (1993) Zn
2
potentiates
ATP-activated currents in rat sympathetic neurons. Eur. J. Physiol.
424, 152158.
Codazzi, F., Racchetti, G., Grohovaz, F., & Meldolesi, J. (1996)
Transduction signals induced in rat brain cortex astrocytes by the
HIV-1 gp120 glycoprotein. FEBS Lett. 384, 135137.
160 Cuajungco and Lees
Copyright

1997 by Academic Press
All rights of reproduction in any form reserved.
Cohen, G. M., Sun, X. M., Snowden, R. T., Dinsdale, D., & Skilleter,
D. N. (1992) Key morphological features of apoptosis may occur in
the absence of internucleosomal DNA fragmentation. Biochem. J.
286, 331334.
Cohen, J. J., & Duke, R. C. (1984) Glucocorticoid activation of a
calcium-dependent endonuclease in thymocyte nuclei leads to cell
death. J. Immunol. 132, 3842.
Connor, M. A., & Chavkin, C. (1992) Ionic zinc may function as an
endogenous ligand for the haloperidol-sensitive 2 receptor in rat
brain. Mol. Pharmacol. 42, 471479.
Constantinidis, J. (1990) Alzheimers disease: The zinc theory.
LEncephale 16, 231239.
Constantinidis, J. (1991) The hypothesis of zinc deciency in the
pathogenesis of neurobrillary tangles. Med. Hypo. 35, 319323.
Corrigan, F. M., Reynolds, G. P., & Ward, N. I. (1993) Hippocampal
tin, aluminum, and zinc in Alzheimers disease. Biometals 6,
149154.
Coyle, J. T., & Puttfarcken, P. (1993) Oxidative stress, glutamate, and
neurodegenerative disorders. Science 262, 689695.
Cronin, J., & Dudek, F. E. (1988) Chronic seizures and collateral
sprouting of dentate mossy bers after kainic acid treatment in
rats. Brain Res. 474, 181184.
Csermely, P., Szamel, M., Resch, K., & Somogyi, J. (1988) Zinc can
increase the activity of protein kinase C and contributes to its
binding to plasma membranes in T lymphocytes. J. Biol. Chem. 263,
64876490.
Cuajungco, M. P., & Lees, G. J. (1996) Prevention of zinc neurotoxic-
ity in vivo by N,N,N,N-tetrakis(2-pyridylmethyl)ethylenediamine
(TPEN). NeuroReport 7, 13011304.
Cuajungco, M. P., & Lees, G. J. (1997) Zinc and Alzheimers disease:
Is there a direct link? Brain Res. Rev., 23, 219236.
Danscher, G. (1981) Histochemical demonstration of heavy metals: A
revised version of the sulphide silver method suitable for both
light and electron microscopy. Histochemistry 71, 116.
Danscher, G. (1996) The autometallographic zinc-sulphide method.
A new approach involving in vivo creation of nanometer-sized
zinc sulphide crystal lattices in zinc-enriched synaptic and secre-
tory vesicles. Histochem. J. 28, 361373.
Danscher, G., Howell, G., Perez-Clausell, J., & Hertel, N. (1985) The
dithizone, Timms sulphide silver and selenium methods demon-
strate a chelatable pool of zinc in CNS. Histochemistry 83, 419422.
Davies, P., & Maloney, A. J. F. (1976) Selective loss of central
cholinergic neurons in Alzheimers disease. Lancet 2, 1403.
Dawson, V. L., Dawson, T. M., Uhl, G. R., & Snyder, S. H. (1993)
Human immunodeciency virus type 1 coat protein neurotoxicity
mediated by nitric oxide in primary cortical cultures. Proc. Natl.
Acad. Sci. USA 90, 32563259.
Deibel, M. A., Ehmann, W. D., & Markesbery, W. R. (1996) Copper,
iron, and zinc imbalances in severely degenerated brain regions in
Alzheimers disease: Possible relation to oxidative stress. J. Neurol.
Sci. 143, 137142.
de Murcia, G., & de Murcia, J. M. (1994) Poly(ADP-ribose) polymer-
ase: Amolecular nick-sensor. Trends Biochem. Sci. 19, 172176.
Deng, Q. S., Turk, G. C., Brady, D. R., & Smith, Q. R. (1994)
Evaluation of brain element composition in Alzheimers disease
using inductively-coupled plasma mass spectrometry. Neurobiol.
Aging 15 (Suppl.), S113(464).
Dexter, D. T., Carayon, A., Javoy-Agid, F., Agid, Y., Wells, F. R.,
Daniel, S. E., Lees, A. J., Jenner, P., & Marsden, C. D. (1991)
Alterations in the levels of iron, ferritin, and other trace metals in
Parkinsons disease and other neurodegenerative diseases affect-
ing the basal ganglia. Brain 114, 19531975.
Dexter, D. T., Carter, C. J., Javoy-Agid, F., Agid, Y., Lees, A., Jenner, P.,
& Marsden, C. D. (1989a) Basal lipid peroxidation in substantia
nigra is increased in Parkinsons disease. J. Neurochem. 52, 381389.
Dexter, D. T., Jenner, P., Schapira, A. H. V., & Marsden, C. D. (1992)
Alterations in levels of iron, ferritin, and other trace metals in
neurodegenerative diseases affecting the basal ganglia. Ann.
Neurol. 32, S94S100.
Dexter, D. T., Wells, F., Lees, A. J., Agid, F., Jenner, P., & Marsden,
C. D. (1989b) Increased nigral iron content and alterations in other
metal ions occurring in brain in Parkinsons disease. J. Neurochem.
52, 18301836.
Donaldson, J., St. Pierre, T., Minnich, J., &Barbeau, A. (1971) Seizures
in rats associated with divalent cation inhibition of Na

/K

-
ATPase. Can. J. Biochem. 49, 12171224.
Drachman, D. A., & Leavitt, J. (1974) Human memory and the
cholinergic system. Arch. Neurol. 30, 113121.
Dragunow, M., Faull, R. L. M., Lawlor, P., Beilharz, E. J., Singleton,
K., Walker, E. B., & Mee, E. (1995) In situ evidence for DNA
fragmentation in Huntingtons disease temporal lobes. NeuroRe-
port 6, 10531057.
Dragunow, M., Goulding, M., Faull, R. L. M., Ralph, R., Mee, E., &
Frith, R. (1990) Induction of c-fos mRNA and protein in neurons
and glia after traumatic brain injury: Pharmacological characteriza-
tion. Exp. Neurol. 107, 236248.
Dragunow, M., & Preston, K. (1995) The role of inducible transcrip-
tion factors in apoptotic nerve cell death. Brain Res. Rev. 21, 128.
Dreixler, J. C., & Leonard, J. P. (1994) Subunit-specic enhancement
of glutamate receptor responses by zinc. Mol. Brain Res. 22,
144150.
Dreosti, I. E. (1989) Neurobiology of zinc. In: Zinc in Human Biology,
(C. F. Mills, Ed.), pp. 235248. Springer-Verlag, London.
Duncan, M. W., Marini, A. M., Watters, R., Kopin, I. J., &Markey, S. P.
(1992) Zinc, a neurotoxin to cultured neurons, contaminates cycad
our prepared by traditional Guamanian methods. J. Neurosci. 12,
15231537.
Durnam, D. M., &Palmiter, R. D. (1981) Transcriptional regulation of
the mouse metallothionein-I gene by heavy metals. J. Biol. Chem.
256, 57125716.
Eagle, G. R., Zombola, R. R., & Himes, R. H. (1983) Tubulin-zinc
interactions: Binding and polymerization studies. Biochemistry 22,
221228.
Ebadi, M. (1991) Metallothioneins and other zinc-binding proteins in
brain. Methods Enzymol. 205, 363387.
Ebadi, M., &Hama, Y. (1986) Zinc-binding proteins in the brain. Adv.
Exp. Med. Biol. 203, 557570.
Ebadi, M., Iversen, P. L., Hao, R., Cerutis, D. R., Rojas, P., Happe,
H. K., Murrin, L. C., & Pfeiffer, R. F. (1995) Expression and
regulation of brain metallothionein. Neurochem. Int. 27, 122.
Ebadi, M., Perini, F., Mountjoy, K., & Garvey, J. S. (1996) Amino acid
composition, immunoreactivity, sequence analysis, and function
of bovine hippocampal metallothionein isoforms. J. Neurochem. 66,
21212127.
Erickson, J. C., Hollopeter, G., Thomas, S. A., Froelick, G. J., &
Palmiter, R. D. (1997) Disruption of the metallothionein-III gene in
mice: Analysis of brain zinc, behavior, and neuron vulnerability to
metals, aging, and seizures. J. Neurosci. 17, 12711281.
Erickson, J. C., Sewell, A. K., Jensen, L. T., Winge, D. R., & Palmiter,
R. D. (1994) Enhanced neurotrophic activity in Alzheimers dis-
ease cortex is not associated with down-regulation of metallothio-
nein-III (GIF). Brain Res. 649, 297304.
Zinc Metabolism in the Brain 161
Copyright

1997 by Academic Press
All rights of reproduction in any form reserved.
Esler, W. P., Stimson, E. R., Jennings, J. M., Ghilardi, J. R., Mantyh,
P. W., & Maggio, J. E. (1996) Zinc-induced aggregation of human
and rat -amyloid peptides in vitro. J. Neurochem. 66, 723732.
Estus, S., Zaks, W. J., Freeman, R. S., Gruda, M., Bravo, R., &Johnson,
E. M. J. (1994) Altered gene expression in neurons during pro-
grammed cell death: Identication of c-jun as necessary for
neuronal apoptosis. J. Cell Biol. 127, 17171727.
Fabris, N., &Mocchegiani, E. (1995) Zinc, human diseases and aging.
Aging Clin. Exp. Res. 7, 7793.
Favier, A., Sappey, C., Leclerc, P., Faure, P., & Micoud, M. (1994)
Antioxidant status and lipid peroxidation in patients infected with
HIV. Chem. Biol. Inter. 91, 165180.
Filipek, A., Heizmann, C. W., & Kuznicki, J. (1990) Calcyclin is a
calcium and zinc binding protein. FEBS Lett. 264, 263266.
Filipek, A., Puzianowska, M., Cieslak, B., & Kuznicki, J. (1993)
CalcyclinCa
2
-binding protein homologous to glial S-100 is
present in neurones. NeuroReport 4, 383386.
Fliss, H., &Menard, M. (1991) Hypochlorous acid-induced mobiliza-
tion of zinc from metalloproteins. Arch. Biochem. Biophys. 287,
175179.
Fliss, H., & Menard, M. (1992) Oxidant-induced mobilization of zinc
from metallothionein. Arch. Biochem. Biophys. 293, 195199.
Fliss, H., Menard, M., & Desai, M. (1991) Hypochlorous acid
mobilizes cellular zinc. Can. J. Physiol. Pharmacol. 69, 16861691.
Frederickson, C. J. (1989) Neurobiology of zinc and zinc-containing
neurons. Int. Rev. Neurobiol. 31, 145238.
Frederickson, C. J., Hernandez, M. D., Goik, S. A., Morton, J. D., &
McGinty, J. F. (1988) Loss of zinc staining from hippocampal
mossy bers during kainic acid induced seizures: A histouores-
cence study. Brain Res. 446, 383386.
Frederickson, C. J., Hernandez, M. D., & McGinty, J. F. (1989)
Translocation of zinc may contribute to seizure-induced death of
neurons. Brain Res. 480, 317321.
Frederickson, C. J., Kasarskis, E. J., Ringo, D., & Frederickson, R. E.
(1987a) A quinoline uorescence method for visualizing and
assaying the histochemically reactive zinc (bouton zinc) in the
brain. J. Neurosci. Methods 20, 91103.
Frederickson, C. J., Klitenick, M. A., Manton, W. I., & Kirkpatrick,
J. B. (1983) Cytoarchitectonic distribution of zinc in the hippocam-
pus of man and the rat. Brain Res. 273, 335339.
Frederickson, C. J., Manton, W. I., Frederickson, M. H., Howell, G. A.,
& Mallory, M. A. (1982) Stable-isotope dilution measurement of
zinc and lead in rat hippocampus and spinal cord. Brain Res. 246,
338341.
Frederickson, C. J., & Moncrieff, D. W. (1994) Zinc-containing
neurons. Biol. Signals 3, 127139.
Frederickson, C. J., Perez-Clausell, J., & Danscher, G. (1987b) Zinc-
containing 7S-NGF complex: Evidence from zinc histochemistry
for localization in salivary secretory granules. J. Histochem. Cyto-
chem. 35, 579583.
Freund, W. D., &Reddig, S. (1994) AMPA/Zn
2
-induced neurotoxic-
ity in rat primary cortical cultures: involvement of L-type calcium
channels. Brain Res. 654, 257264.
Fukahori, M., Itoh, M., Oomagari, K., & Kawasaki, H. (1988) Zinc
content in discrete hippocampal and amygdaloid areas of the
epilepsy (E1) mouse and normal mice. Brain Res. 455, 381384.
Fuson, K. S., Boggs, L. N., & May, P. C. (1996) Zinc promotes A
aggregation but attenuates A neurotoxicity. Neurobiol. Aging 17
(Supplement), S108(431).
Gabrielsson, B., Robson, T., Norris, D., & Chung, S. H. (1986) Effects
of divalent metal ions on the uptake of glutamate and GABAfrom
synaptosomal fractions. Brain Res. 384, 218223.
Gabuzda, D., Busciglio, J., & Yankner, B. A. (1993) Inhibition of
-amyloid production by activation of protein kinase C. J. Neuro-
chem. 61, 23262329.
Gajate, C., Alonso, M. T., Schimmang, T., & Mollinedo, F. (1996) C-fos
is not essential for apoptosis. Biochem. Biophys. Res. Commun. 218,
267272.
Gaskin, F., &Kress, Y. (1977) Zinc ion-induced assembly of tubulin. J.
Biol. Chem. 252, 69186924.
Gaskin, F., Kress, Y., Brosnan, C., & Bornstein, M. (1978) Abnormal
tubulin aggregates induced by zinc sulfate in organotypic cultures
of nerve tissue. Neuroscience 3, 11171128.
Globus, M. Y. T., Ginsberg, M. D., & Busto, R. (1991) Excitotoxic
indexA biochemical marker of selective vulnerability. Neurosci.
Lett. 127, 3942.
Goldblum, S. E., Cohen, D. A., Jay, M., & McClain, C. J. (1987)
Interleukin 1-induced depression of iron and zinc: Role of granulo-
cytes and lactoferrin. Am. J. Physiol. 252, E27E32.
Golden, B. E. (1989) Zinc in cell division and tissue growth:
Physiological aspects. In: Zinc in Human Biology, (C. F. Mills, Ed.),
pp. 119128. Springer-Verlag, London.
Gottschall, P. E., & Yu, X. (1995) Cytokines regulate gelatinase Aand
B (matrix metalloproteinase 2 and 9) activity in cultured rat
astrocytes. J. Neurochem. 64, 15131520.
Green, L. M., & Berg, J. M. (1990) Retroviral nucleocapsid protein-
metal ion interactions: Folding and sequence variants. Proc. Natl.
Acad. Sci. USA 87, 64036407.
Greenlund, L. J. S., Deckwerth, T. L., & Johnson, E. M. J. (1995)
Superoxide dismutase delays neuronal apoptosis: A role for
reactive oxygen species in programmed neuronal death. Neuron
14, 303315.
Griffin, W. S. T., Stanley, L. C., Ling, C., White, L., MacLeod, V.,
Perrot, L. J., White III, C. L., &Araoz, C. (1989) Brain interleukin-1
and S-100 immunoreactivity are elevated in Down syndrome and
Alzheimers disease. Proc. Natl. Acad. Sci. USA 86, 76117615.
Griffin, W. S. T., Yeralan, O., Sheng, J. G., Boop, F. A., Mrak, R. E.,
Rovnaghi, C. R., Burnett, B. A., Feoktistova, A., & Van Eldik, L. J.
(1995) Overexpression of the neutrophic cytokine S100 in human
temporal lobe epilepsy. J. Neurochem. 65, 228233.
Grilli, M., Ribola, M., Alberici, A., Valerio, A., Memo, M., & Spano, P.
(1995) Identication and characterization of a B/Rel binding site
in the regulatory region of the amyloid precursor protein gene. J.
Biol. Chem. 270, 2677426777.
Gschwind, M., & Huber, G. (1995) Apoptotic cell death induced by
-amyloid
142
peptide is cell type dependent. J. Neurochem. 65,
292300.
Halliwell, B., & Gutteridge, J. M. C. (1984) Oxygen toxicity, oxygen
radicals, transition metals and disease. Biochem. J. 219, 114.
Harrison, N. L., & Gibbons, S. J. (1994) Zn
2
: An endogenous
modulator of ligand- and voltage-gated ion channels. Neurophar-
macol. 33, 935952.
Haug, F. M. S. (1967) Electron microscopical localization of the zinc
in hippocampal mossy bre synapses by a modied sulde silver
procedure. Histochemie 8, 355368.
Haug, F. M. S., Blackstad, T. W., Simonsen, A. H., & Zimmer, J. (1971)
Timms sulde silver reaction for zinc during experimental antero-
grade degeneration of hippocampal mossy bers. J. Comp. Neurol.
142, 2332.
Hedberg, K. K., Birrell, G. B., Mobley, P. L., & Griffith, O. H. (1994)
Transition metal chelator TPENcounteracts phorbol ester-induced
actin cytoskeletal disruption in c6 rat glioma cells without inhibit-
ing activation or translocation of protein kinase C. J. Cell. Physiol.
158, 337346.
162 Cuajungco and Lees
Copyright

1997 by Academic Press
All rights of reproduction in any form reserved.
Heizmann, C. W., & Braun, K. (1995) Calcium Regulation by Calcium-
Binding Proteins in Neurodegenerative Disorders. Springer-Verlag/
Landes, Austin, TX.
Hendriks, L., & Van Broeckhoven, C. (1996) The A4 amyloid
precursor protein gene and Alzheimers disease. Eur. J. Biochem.
237, 615.
Herdegen, T., Kovary, K., Buhl, A., Bravo, R., Zimmermann, M., &
Gass, P. (1995) Basal expression of the inducible transcription
factors c-jun, JunB, JunD, c-Fos, FosB, and krox-24 in the adult rat
brain. J. Comp. Neurol. 354, 3956.
Hesse, G. (1979) Chronic zinc deciency alters neuronal function of
hippocampal mossy bers. Science 205, 10051007.
Hexum, T. D. (1974) Studies on the reaction catalyzed by transport
(Na,K) adenosine triphosphatease-I: Effects of divalent metals.
Biochem. Pharmacol. 23, 34413447.
Hirsch, E. C. (1992) Why are nigral catecholaminergic neurons more
vulnerable than other cells in Parkinsons disease. Ann. Neurol. 32,
S88S93.
Holland, D. R., Cousens, L. S., Meng, W., & Matthews, B. W. (1994)
Nerve growth factor in different crystal forms displays structural
exibility and reveals zinc binding sites. J. Mol. Biol. 239, 385400.
Hopkins, S. J., & Rothwell, N. J. (1995) Cytokines and the nervous
system. I. Expression and recognition. Trends Neurosci. 18, 8388.
Howell, G. A., & Frederickson, C. J. (1989) A retrograde transport
method for mapping zinc-containing ber systems in the brain.
Brain Res. 515, 277286.
Howell, G. A., Welch, M. G., &Frederickson, C. J. (1984) Stimulation-
induced uptake and release of zinc in hippocampal slices. Nature
308, 736738.
Huang, K. P. (1989) The mechanism of protein kinase C activation.
Trends Neurosci. 12, 425432.
Huang, J., & Philbert, M. A. (1995) Distribution of glutathione and
glutathione-related enzyme systems in mitochondria and cytosol
of cultured cerebellar astrocytes and granule cells. Brain Res. 680,
1622.
Hubbard, S. R., Bishop, W. R., Kirschmeier, P., George, S. J., Cramer,
S. P., & Hendrickson, W. A. (1991) Identication and characteriza-
tion of zinc binding sites in protein kinase C. Science 254,
17761779.
Hung, A. Y., Haass, C., Nitsch, R. M., Qiu, W. Q., Citron, M.,
Wurtman, R. J., Growdon, J. H., & Selkoe, D. J. (1993) Activation of
protein kinase C inhibits cellular production of the amyloid-
protein. J. Biol. Chem. 268, 2295922962.
Hunter, F. E. J., & Ford, L. (1955) Inactivation of oxidative and
phosphorylative systems in mitochondria by preincubation with
phosphate and other ions. J. Biol. Chem. 216, 357369.
Hutchin, T., & Cortopassi, G. (1995) A mitochondrial DNA clone is
associated with increased risk for Alzheimer disease. Proc. Natl.
Acad. Sci. USA 92, 68926895.
Ihara, Y. (1988) Massive somatodendritic sprouting of cortical neu-
rons in Alzheimers disease. Brain Res. 459, 138144.
Israel, N., Hazan, U., Alcami, J., Munier, A., Arenzana-Seisdedos, F.,
Bachelerie, F., Israel, A., & Virelizier, J. L. (1989) Tumor necrosis
factor stimulates transcription of HIV-1 in human t lymphocytes,
independently and synergistically with mitogens. J. Immunol. 143,
39563960.
Itoh, M., Ebadi, M., & Swanson, S. (1983) The presence of zinc-
binding proteins in brain. J. Neurochem. 41, 823829.
Iversen, L. L., Mortishire-Smith, R. J., Pollack, S. J., &Shearman, M. S.
(1995) The toxicity in vitro of -amyloid protein. Biochem. J. 311,
116.
Izumi, R., Yamada, T., Yoshikai, S. I., Hattori, M., & Sakaki, Y. (1992)
Positive and negative regulatory elements for the expression of the
Alzheimers disease amyloid precursor-encoding gene in mouse.
Gene 112, 189195.
Jenner, P. (1993) Altered mitochondrial function, iron metabolism
and glutathione levels in Parkinsons disease. Acta Neurol. Scand.
Suppl. 146, 613.
Jenner, P., Dexter, D. T., Sian, J., Schapira, A. H. V., & Marsden, C. D.
(1992) Oxidative stress as a cause of nigral cell death in Parkin-
sons disease and incidental Lewy body disease. Ann. Neurol. 32,
S82S87.
Jiang, S., Chow, S. C., McCabe, M. J. J., & Orrenius, S. (1995) Lack of
Ca
2
involvement in thymocyte apoptosis induced by chelation of
intracellular Zn
2
. Lab. Invest. 73, 111117.
Kagi, J. H. R., & Schaffer, A. (1988) Biochemistry of metallothionein.
Biochem. 27, 85098515.
Kaiser, J. (1994) Alzheimers: Could there be a zinc link? Science 265,
1365.
Kamizono, A., Nishizawa, M., Teranishi, Y., Murata, K., & Kimura,
A. (1989) Identication of a gene conferring resistance to zinc and
cadmiumions in the yeast Saccharomyces cerevisiae. Mol. Gen. Genet.
219, 161167.
Kasarskis, E. J. (1984) Zinc metabolism in normal and zinc-decient
rat brain. Exp. Neurol. 85, 114127.
Kato, K., Suzuki, F., Morishita, R., Asano, T., & Sato, T. (1990)
Selective increase in S-100 protein by aging in rat cerebral cortex.
J. Neurochem. 54, 12691274.
Kelly, E. J., Quaife, C. J., Froelick, G. J., & Palmiter, R. D. (1996)
Metallothionein I and II protect against zinc deciency and zinc
toxicity in mice. J. Nutrition 126, 17821790.
Kerr, J. F. R., Wyllie, A. H., & Currie, A. R. (1972) Apoptosis: Abasic
biological phenomenon with wide-ranging implications in tissue
kinetics. Br. J. Cancer 26, 239257.
Kesslak, J. P., Frederickson, C. J., & Gage, F. H. (1987) Quantication
of hippocampal noradrenaline and zinc changes after selective cell
destruction. Exp. Brain Res. 67, 7784.
Killinger, S., Li, Y., &Balcar, V. J. (1995) Sensitivity of the binding sites
on glutamate transporters to neurotoxic agents. NeuroReport 6,
12901292.
Kleiner, D., & von Jagow, G. (1972) On the inhibition of mitochon-
drial electron transport by Zn
2
ions. FEBS Lett. 20, 229232.
Kligman, D., & Marshak, D. R. (1985) Purication and characteriza-
tion of a neurite extension factor from bovine brain. Proc. Natl.
Acad. Sci. USA 82, 71367139.
Koh, J. Y., & Choi, D. W. (1988) Zinc alters excitatory amino acid
neurotoxicity on cortical neurons. J. Neurosci. 8, 21642171.
Koh, J. Y., & Choi, D. W. (1994) Zinc toxicity on cultured cortical
neurons: involvement of N-methyl-D-aspartate receptors. Neurosci-
ence 60, 10491057.
Koh, J. Y., Suh, S. W., Gwag, B. J., He, Y. Y., Hsu, C. Y., & Choi, D. W.
(1996) The role of zinc in selective neuronal death after transient
global cerebral ischemia. Science 272, 10131016.
Koizumi, S., Ikeda, M., Inoue, K., Nakazawa, K., & Inoue, K. (1995)
Enhancement by zinc of ATP-evoked dopamine release from rat
pheochromocytoma PC12 cells. Brain Res. 673, 7582.
Kozma, M., Szerdahelyi, P., & Kasa, P. (1981) Histochemical detec-
tion of zinc and copper in various neurons of the central nervous
system. Acta Histochem. 69, 1217.
Kress, Y., Gaskin, F., Brosnan, C. F., & Levine, S. (1981) Effects of zinc
on the cytoskeletal proteins in the central nervous system of the
rat. Brain Res. 220, 139149.
Zinc Metabolism in the Brain 163
Copyright

1997 by Academic Press
All rights of reproduction in any form reserved.
Krezoski, S. K., Villalobos, J., Shaw III, C. F., & Petering, D. H. (1988)
Kinetic lability of zinc bound to metallothionein in Erhlich cells.
Biochem. J. 255, 483491.
Krieger, C., Lanius, R. A., Pelech, S. L., & Shaw, C. A. (1996)
Amyotrophic lateral sclerosis: The involvement of intracellular
Ca
2
and protein kinase C. Trends Pharmacol. Sci. 17, 114120.
Kudo, H., Kokunai, T., Kondoh, T., Tamaki, N., & Matsumoto, S.
(1990) Quantitative analysis of glutathione rat central nervous
system: Comparison of GSH in infant brain with that in adult
brain. Brain Res. 511, 326328.
Kumamoto, E., & Murata, Y. (1995) Characterization of GABA
current in rat septal cholinergic neurons in culture and its
modulation by metal cations. J. Neurophysiol. 74, 20122027.
Kumamoto, E., & Murata, Y. (1996) Glycine current in rat septal
cholinergic neuron in culture: Monophasic positive modulation by
Zn
2
. J. Neurophysiol. 76, 227241.
Kume, A., Sakurai, S. Y., & Albin, R. L. (1994) Zinc inhibition of
t-[3H]butyl bicycloorthobenzoate binding to the GABA
A
receptor
complex. J. Neurophysiol. 62, 602607.
LaFerla, F. M., Tinkle, B. T., Bieberich, C. J., Haudenschild, C. C., &
Jay, G. (1995) The Alzheimers A peptide induces neurodegenera-
tion and apoptotic cell death in transgenic mice. Nature Gen. 9,
2129.
Laube, B., Kuhse, J., Rundstrom, N., Kirsch, J., Schmieden, V., &Betz,
H. (1995) Modulation by zinc ions of native rat and recombinant
human inhibitory glycine receptors. J. Physiol. 483, 613619.
Lee, S. P., & Han, M. K. (1996) Zinc stimulates Mg
2
-dependent
3-processing activity of human immunodeciency virus type 1
integrase in vitro. Biochemistry 35, 38373844.
Lees, G. J. (1991) Inhibition of sodium-potassium ATPase: A poten-
tially ubiquitous mechanism contributing to central nervous
system neuropathology. Brain Res. Rev. 16, 221228.
Lees, G. J. (1993) Contributory mechanisms in the causation of
neurodegenerative disorders. Neuroscience 54, 287322.
Lees, G. J., Lehmann, A., Sandberg, M., & Hamberger, A. (1990) The
neurotoxicity of zinc in the rat hippocampus. Neurosci. Lett. 120,
155158.
Lees, G. J., & Leong, W. (1994) Brain lesions induced by specic and
non-specic inhibitors of sodium-potassium ATPase. Brain Res.
649, 225233.
Lees, G. J., & Leong, W. (1995) The sodiumpotassium ATPase
inhibitor ouabain is neurotoxic in the rat substantia nigra and
striatum. Neurosci. Lett. 188, 113116.
Leonard, J., Parrott, C., Buckler-White, A. J., Turner, W., Ross, E. K.,
Martin, M. A., & Rabson, A. B. (1989) The NF-B binding sites in
the human immunodeciency virus type 1 long terminal repeat
are not required for virus infectivity. J. Virol. 63, 49194924.
Levy-Lahad, E., Wasco, W., Poorkaj, P., Romano, D. M., Oshima, J.,
Pettingel, W. H., Yu, C. E., Jondro, P. D., Schmidt, S. D., Wang, K.,
Crowley, A. C., Fu, Y. H., Guenette, S. Y., Galas, D., Nemens, E.,
Wijsman, E. M., Bird, T. D., Schellenberg, G. D., & Tanzi, R. E.
(1995b) Candidate gene for the chromosome 1 familial Alzhei-
mers disease locus. Science 269, 973977.
Levy-Lahad, E., Wijsman, E. M., Nemens, E., Anderson, L., Goddard,
K. A. B., Weber, J. L., Bird, T. D., & Schellenberg, G. D. (1995a) A
familial alzheimers disease locus on chromosome 1. Science 269,
970973.
Lim, G. P., Backstrom, J. R., Cullen, M. J., Miller, C. A., Atkinson,
R. D., & Tokes, Z. A. (1996) Matrix metalloproteinase in the
neocortex and spinal cord of amyotrophic lateral sclerosis pa-
tients. J. Neurochem. 67, 251259.
Lin, Y. S., Kao, S. F., Jan, M. S., Cheng, M. L., Wing, L. Y. C., Chang,
W. C., Lei, H. Y., & Lin, M. T. (1995) Changes of protein kinase C
subspecies in staphylococcal enterotoxin-B-induced thymocyte
apoptosis. Biochem. Biophys. Res. Commun. 213, 11321139.
Link, T. A., & von Jagow, G. (1995) Zinc ions inhibit the Qp center of
bovine heart mitochondrial bcl complex by blocking a protonable
group. J. Biol. Chem. 270, 2500125006.
Long, Y., Hardwick, A. L., & Frederickson, C. J. (1995) Zinc-
containing innervation of the subicular region in the rat. Neuro-
chem. Int. 27, 95103.
Luisi, B. F., Xu, W. X., Otwinowski, Z., Freedman, L. P., Yamamoto,
K. R., & Sigler, P. B. (1991) Crystallographic analysis of the inter-
action of the glucocorticoid receptor with DNA. Nature
352, 497505.
Lyons, T. J., Liu, H., Goto, J. J., Nersissian, A., Roe, J. A., Graden, J. A.,
Cafe, C., Ellerby, L. M., Bredesen, D. E., Butler Gralla, E., &
Selverstone Valentine, J. (1996) Mutations in copper-zinc superox-
ide dismutase that cause amyotrophic lateral sclerosis alter the
zinc binding site and the redox behavior of the protein. Proc. Natl.
Acad. Sci. USA 93, 1224012244.
Mantyh, P. W., Ghilardi, J. R., Rogers, S., DeMaster, E., Allen, C. J.,
Stimson, E. R., & Maggio, J. E. (1993) Aluminum, iron, and zinc
ions promote aggregation of physiological concentrations of
-amyloid peptide. J. Neurochem. 61, 11711174.
Maret, W. (1994) Oxidative metal release from metallothionein via
zinc-thiol/disulde interchange. Proc. Natl. Acad. Sci. USA 91,
237241.
Maret, W. (1995) Metallothionein/disulde interactions, oxidative
stress, and the mobilization of cellular zinc. Neurochem. Int. 27,
111117.
Marshak, D. R., Pesce, S. A., Stanley, L. C., & Griffin, W. S. T. (1991)
Increased S-100 neurotrophic activity in Alzheimers disease
temporal lobe. Neurobiol. Aging 13, 17.
Martinez-Balbas, M. A., Jimenez-Garcia, E., & Azorin, F. (1995)
Zinc(II) ions selectively interact with DNA sequences present at
the TFIIIAbinding site of the Xenopous 5S-RNAgene. Nucleic Acids
Res. 23, 24642471.
Masters, B. A., Quaife, C. A., Erickson, J. C., Kelly, E. J., Froelick, G. J.,
Zambrowicz, B. P., Brinster, R. L., & Palmiter, R. D. (1994)
Metallothionein III is expressed in neurons that sequester zinc in
synaptic vesicles. J. Neurosci. 14, 58445857.
Matsushima, H., Shimohama, S., Chachin, M., Taniguchi, T., &
Kimura, J. (1996) Ca
2
-dependent and Ca
2
-independent protein
kinase C changes in the brains of patients with Alzheimers
disease. J. Neurochem. 67, 317323.
Mattson, M. P. (1995a) Untangling the pathophysiochemistry of
-amyloid. Nature Struc. Biol. 2, 926928.
Mattson, M. P. (1995b) Degenerative and protective signalling
mechanisms in the neurobrillary pathology of AD. Neurobiol.
Aging 16, 447463.
Mayer, M. L., & Vyklicky, L. J. (1989) The action of zinc on synaptic
transmission and neuronal excitability in cultures of mouse
hippocampus. J. Physiol. 415, 351365.
Mayer, M. L., Vyklicky, L. J., &Westbrook, G. L. (1989) Modulation of
excitatory amino acid receptors by group IIB metal cations in
cultured mouse hippocampal neurones. J. Physiol. 415, 329350.
McCabe, M. J. J., Jiang, S. A., & Orrenius, S. (1993) Chelation of
intracellular zinc triggers apoptosis in mature thymocytes. Lab.
Invest. 69, 101110.
McCord, J. M., & Flores, S. C. (1994) The human immunodeciency
virus and oxidative balance. In: Oxidative Processes and Antioxi-
dants (R. Paoletti, B. Samuelsson, A. L. Catapano, A. Poli, & M.
Rinetti, Eds.), pp. 1323. Raven Press, NewYork.
164 Cuajungco and Lees
Copyright

1997 by Academic Press
All rights of reproduction in any form reserved.
McGeer, P. L., & McGeer, E. G. (1995) The inammatory response
system of brain: Implications for therapy of Alzheimer and other
neurodegenerative diseases. Brain Res. Rev. 21, 195218.
McNall, A. D., Etherton, T. D., & Fosmire, G. J. (1995) The impaired
growth induced by zinc deciency in rats is associated with
decreased expression of the hepatic insulin-like growth factor I
and growth hormone receptor genes. J. Nutrition 125, 874879.
Mecocci, P., MacGarvey, U., &Beal, M. F. (1994) Oxidative damage to
mitochondrial DNA is increased in Alzheimers disease. Ann.
Neurol. 36, 747751.
Meister, A. (1983) Selective modication of glutathione metabolism.
Science 200, 472477.
Mely, Y., De Rocquigny, H., Morellet, N., Roques, B. P., & Gerard, D.
(1996) Zinc binding to the HIV-1 nucleocapsid protein: Athermo-
dynamic investigation by uorescence spectroscopy. Biochem. 35,
51755182.
Mills, C. F. (1989) The biological signicance of zinc for man:
Problems and prospects. In: Zinc in Human Biology (C. F. Mills,
Ed.), pp. 371381. Springer-Verlag, London.
Mitchell, C. L., & Barnes, M. I. (1993) Proconvulsant action of
diethyldithiocarbamate in stimulation of the perforant path. Neu-
rotoxicol. Teratol. 15, 165171.
Mitchell, C. L., Barnes, M. I., &Grimes, L. M. (1990) Diethyldithiocar-
bamate and dithizone augment the toxicity of kainic acid. Brain
Res. 506, 327330.
Mochizuki, H., Goto, K., Mori, H., & Mizuno, Y. (1996) Histochemi-
cal detection of apoptosis in Parkinsons disease. J. Neurol. Sci. 137,
120123.
Mody, I., & MacDonald, J. F. (1995) NMDA receptor-dependent
excitotoxicity: The role of intracellular Ca
2
release. Trends Pharma-
col. Sci. 16, 356359.
Mody, I., & Miller, J. J. (1985) Levels of hippocampal calcium and
zinc following kindling-induced epilepsy. Can. J. Physiol. Pharma-
col. 63, 159161.
Morton, J. D., Howell, G. A., & Frederickson, C. J. (1990) Effects of
subcutaneous injections of zinc chloride on seizures induced by
noise and by kainic acid. Epilepsia 31, 139144.
Multhaup, G., Bush, A. I., Pollwein, P., & Masters, C. L. (1994)
Interaction between the zinc(II) and the heparin binding site of the
Alzheimers disease a4 amyloid precursor protein. FEBS Lett.
355, 151154.
Murakami, K., Whiteley, M. K., & Routtenberg, A. (1987) Regulation
of protein kinase C activity by cooperative interaction of Zn
2
and
Ca
2
. J. Biol. Chem. 262, 1390213906.
Mustafa, M. G., Cross, C. E., Munn, R. J., &Hardie, J. A. (1971) Effects
of divalent metal ions on alveolar macrophage membrane adeno-
sine triphosphatase activity. J. Lab. Clin. Med. 77, 563571.
Mutisya, E. M., Bowling, A. C., & Beal, M. F. (1994) Cortical
cytochrome oxidase activity is reduced in Alzheimers disease. J.
Neurochem. 63, 21792184.
Narindrasorasak, S., Altman, R. A., Gonzales-DeWhitt, P., Green-
berg, B. D., & Kisilevsky, R. (1995) An interaction between
basement membrane and Alzheimer amyloid precursor proteins
suggests a role in the pathogenesis of Alzheimers disease. Lab.
Invest. 72, 272282.
Narindrasorasak, S., Lowery, D. E., Altman, R. A., Gonzales-
DeWhitt, P. A., Greenberg, B. D., & Kisilevsky, R. (1992) Character-
ization of high affinity binding between laminin and Alzheimers
disease amyloid precursor proteins. Lab. Invest. 67, 643652.
Nave, J. M., & Connor, J. D. (1993) Inuence of ZnCl
2
pretreatment
on behavioral and histological responses to kainic acid in rats.
Brain Res. 604, 298303.
Navia, B. A., Jordan, B. D., & Price, R. W. (1986) The AIDS dementia
complex. I. Clinical features. Ann. Neurol. 19, 517524.
Nitatori, T., Sato, N., Waguri, S., Karasawa, Y., Araki, H., Shibanai,
K., Kominami, E., & Uchiyama, Y. (1995) Delayed neuronal death
in the CA1 pyramidal cell layer of the gerbil hippocampus
following transient ischemia is apoptosis. J. Neurosci. 15, 1001
1011.
Odeh, M. (1992) The role of zinc in acquired immunodeciency
syndrome. J. Internal Med. 231, 463469.
OHalloran, T. V. (1993) Transition metals in control of gene expres-
sion. Science 261, 715725.
Olanow, C. W. (1993) A radical hypothesis for neurodegeneration.
Trends Neurosci. 16, 439444.
Omichinski, J. G., Trainor, C., Evans, T., Gronenborn, A. M., Clore,
G. M., &Felsenfeld, G. (1993) Asmall single-nger peptide from
the erythroid transcription factor GATA-1 binds specically to
DNA as a zinc or iron complex. Proc. Natl. Acad. Sci. USA 90,
16761680.
Palmiter, R. D. (1995) Constitutive expression of metallothionein-III
(MT-III), but not MT-I, inhibits growth when cells become zinc
decient. Toxicol. Appl. Pharmacol. 135, 139146.
Palmiter, R. D., & Findley, S. D. (1995) Cloning and functional
characterization of a mammalian zinc transporter that confers
resistance to zinc. EMBO J. 14, 639649.
Palmiter, R. D., Cole, T. B., & Findley, S. D. (1996a) ZnT-2, a
mammalian protein that confers resistance to zinc by facilitating
vesicular sequestration. EMBO J. 15, 17841791.
Palmiter, R. D., Cole, T. B., Quaife, C. J., & Findley, S. D. (1996b)
ZnT-3, a putative transporter of zinc into synaptic vesicles. Proc.
Natl. Acad. Sci. USA 93, 1493414939.
Palmiter, R. D., Findley, S. D., Whitmore, T. E., & Durnam, D. M.
(1992) MT-III, a brain-specic member of the metallothionein gene
family. Proc. Natl. Acad. Sci. USA 89, 63336337.
Pattison, S. E., & Dunn, M. F. (1976a) On the mechanism of divalent
metal ion chelator induced activation of the 7S nerve growth factor
esteropeptidase: activation by 2,2,2-terpyridine and by 8-hy-
droxyquinoline-5-sulfonic acid. Biochem. 15, 36913696.
Pattison, S. E., & Dunn, M. F. (1976b) On the mechanism of divalent
metal ion chelator induced activation of the 7S nerve growth factor
esteropeptidase: Thermodynamics and kinetics of activation. Bio-
chem. 15, 36963703.
Pavletich, N. P., & Pabo, C. O. (1991) Zinc nger-DNA recognition:
Crystal structure of a Zif268DNA complex at 2.1 angstrom.
Science 252, 809817.
Pedrocchi, M., Schafer, B. W., Durussel, I., Cox, J. A., & Heizmann,
C. W. (1994) Purication and characterization of the recombinant
human calcium-binding S-100 proteins CAPLand CACY. Biochem-
istry 33, 67326738.
Pei, Y. Q., & Koyama, I. (1986) Features of seizures and behavioral
changes induced by intrahippocampal injection of zinc sulfate in
the rabbit: A new experimental model of epilepsy. Epilepsia 27,
183188.
Pei, Y. Q., Zhao, D., Huang, J., & Cao, L. (1983) Zinc-induced
seizures: A new experimental model of epilepsy. Epilepsia 24,
169176.
Peled-Kamar, M., Lotem, J., Okon, E., Sachs, L., & Groner, Y. (1995)
Thymic abnormalities and enhanced apoptosis of thymocytes and
bone marrow cells in transgenic mice overexpressing Cu/Zn-
superoxide dismutase: Implications for Down syndrome. EMBO J.
14, 49854993.
Zinc Metabolism in the Brain 165
Copyright

1997 by Academic Press
All rights of reproduction in any form reserved.
Perez-Clausell, J., &Danscher, G. (1985) Intravesicular localization of
zinc in rat telencephalic boutons: Ahistochemical study. Brain Res.
337, 9198.
Perez-Clausell, J., & Danscher, G. (1986) Release of zinc sulphide
accumulation into synaptic clefts after in vivo injection of sodium
sulphide. Brain Res. 362, 358361.
Persechini, A., McMillan, K., & Masters, B. S. S. (1995) Inhibition of
nitric oxide synthase activity by zinc ion. Biochemistry 34, 15091
15095.
Peters, S., Koh, J. Y., & Choi, D. W. (1987) Zinc selectively blocks the
action of N-methyl-D-aspartate on cortical neurons. Science 236,
589593.
Pollard, H., Cantagrel, S., Charriaut-Marlangue, C., Moreau, J., &
Ben-Ari, Y. (1994a) Apoptosis associated DNA fragmentation in
epileptic brain damage. NeuroReport 5, 10531055.
Pollard, H., Charriaut-Marlangue, C., Cantagrel, S., Represa, A.,
Robain, O., Moreau, J., & Ben-Ari, Y. (1994b) Kainate-induced
apoptotic cell death in hippocampal neurons. Neuroscience 63,
718.
Portera-Cailliau, C., Hedreen, J. C., Price, D. L., & Koliatsos, V. E.
(1995) Evidence for apoptotic cell death in Huntingtons disease
and excitotoxic animal models. J. Neurosci. 15, 37753787.
Posewitz, M. C., & Wilcox, D. E. (1995) Properties of the Sp1 zinc
nger 3 peptide: Coordination chemistry, redox reactions, and
metal binding competition with metallothionein. Chem. Res. Toxi-
col. 8, 10201028.
Pountney, D. L., Fundel, S. M., Faller, P., Birchler, N. E., Hunziker,
P., & Vasak, M. (1994) Isolation, primary structures and metal
binding properties of neuronal growth inhibitory factor (GIF)
from bovine and equine brain. FEBS Lett. 345, 193197.
Prasad, A. S. (1996) Zinc: The biology and therapeutics of an ion.
Ann. Int. Med. 125, 142144.
Price, R. W., Brew, B., Sidtis, J., Rosenblaum, M., Scheck, A., &Cleary,
P. (1988) The brain in AIDS: Central nervous system HIV-1
infection and AIDS dementia complex. Science 239, 586592.
Provinciali, M., Di Stefano, G., & Fabris, N. (1995) Dose-dependent
opposite effect of zinc on apoptosis in mouse thymocytes. Int. J.
Immunopharmacol. 17, 735744.
Przedborski, S., Donaldson, D., Jakowec, M., Kish, S. J., Guttman, M.,
Rosoklija, G., & Hays, A. P. (1996) Brain superoxide dismutase,
catalase, and glutathione peroxidase activities in amyotrophic
lateral sclerosis. Ann. Neurol. 39, 158165.
Pullen, R. G. L., Franklin, P. A., &Hall, G. H. (1991)
65
Zn uptake from
blood into brain in the rat. J. Neurochem. 56, 485489.
Quaife, C. J., Findley, S. D., Erickson, J. C., Froelick, G. J., Kelly, E. J.,
Zambrowicz, B. P., & Palmiter, R. D. (1994) Induction of a new
metallothionein isoform (MT-IV) occurs during differentiation of
stratied squamous epithelia. Biochemistry 33, 72507259.
Quitschke, W. W. (1994) Two nuclear factor binding domains activate
expression from the human amyloid -protein precursor pro-
moter. J. Biol. Chem. 269, 2122921233.
Rabizadeh, S., Gralla, E. B., Borchelt, D. R., Gwinn, R., Valentine,
J. R., Sisodia, S., Wong, P., Lee, M., Hahn, H., & Bredesen, D. E.
(1995) Mutations associated with amyotrophic lateral sclerosis
convert superoxide dismutase from an antiapoptotic gene to a
proapoptotic gene: studies in yeast and neural cells. Proc. Natl.
Acad. Sci. USA 92, 30243028.
Ratan, R. R., Murphy, T. H., & Baraban, J. M. (1994) Oxidative stress
induces apoptosis in embryonic cortical neurons. J. Neurochem. 62,
376379.
Reeves, R. H., Yao, J., Crowley, M. R., Buck, S., Zhang, X., Yarowsky,
P., Gearhart, J. D., & Hilt, D. C. (1994) Astrocytosis and axonal
proliferation in the hippocampus of S-100 transgenic mice. Proc.
Natl. Acad. Sci. USA 91, 53595363.
Reinhardt, D., Mann, K., Nischt, R., Fox, J. W., Chu, M. L., Krieg, T., &
Timpl, R. (1996) Mapping of nidogen binding sites for collagen
type IV, heparan sulfate proteoglycan, and zinc. J. Biol. Chem. 268,
1088110887.
Reyes, J. G. (1996) Zinc transport in mammalian cells. Am. J. Physiol.
270, C401C410.
Rhodes, D., & Klug, A. (1993) Zinc ngers. Sci. Am. 268, 3239.
Ricciardi, T. N., & Malouf, A. T. (1995) Differential effects of zinc on
hyperpolarizing and depolarizing GABA
A
synaptic potentials in
hippocampal slice cultures. Brain Res. 680, 8087.
Rice, W. G., Schaeffer, C. A., Harten, B., Villinger, F., South, T. L.,
Summers, M. F., Henderson, L. E., Bess, J. W. J., Arthur, L. O.,
McDougal, J. S., Orloff, S. L., Mendeleyev, J., & Kun, E. (1993)
Inhibition of HIV-1 infectivity by zinc-ejecting aromatic C-nitroso
compounds. Nature 361, 473475.
Richeld, E. K. (1993) Zinc modulation of drug binding, cocaine
affinity states, and dopamine uptake on the dopamine uptake
complex. Mol. Pharmacol. 43, 100108.
Rigaud-Monnet, A. S., Heron, A., Seylaz, J., & Pinard, E. (1995) Effect
of inhibiting NO synthesis on hippocampal extracellular gluta-
mate concentration in seizures induced by kainic acid. Brain Res.
673, 297303.
Rinaudo, M. S., Su, K., Falk, L. A., Haldar, S., & Mufson, R. A. (1995)
Human interleukin-3 receptor modulates bcl-2 mRNAand protein
levels through protein kinase C in TF-1 cells. Blood 86, 8088.
Ripa, S., & Ripa, R. (1995) Zinc and immune function. Minerva Med.
86, 315318.
Rogers, J. M., Taubeneck, W. M., Daston, G. P., Sulik, K. K., Zucker,
R. M., Elstein, K. H., Jankowski, M. A., & Keen, C. L. (1995) Zinc
deciency causes apoptosis but not cell cycle alterations in
organogenesis-stage rat embryos: Effect of varying duration of
deciency. Teratology 52, 149159.
Roher, A. E., Kasunic, T. C., Woods, A. S., Cotter, R. J., Ball, M. J., &
Fridman, R. (1994) Proteolysis of A peptide from Alzheimers
disease brain by gelatinase A. Biochem. Biophys. Res. Commun. 205,
17551761.
Rosen, D. R., Siddique, T., Patterson, D., Figlewicz, D. A., Sapp, P.,
Hentati, A., Donaldson, D., Goto, J., ORegan, J. P., Deng, H. X.,
Rahmani, Z., Krizus, A., Mckenna-Yasek, D., Cayabyab, A., Gas-
ton, S. M., Berger, R., Tanzi, R. E., Halperin, J. J., Herzfeldt, B., Van
den Bergh, R., Hung, W. Y., Bird, T., Deng, G., Mulder, D. W.,
Smyth, C., Laing, N. G., Soriano, E., Pericak-Vance, M. A., Haines,
J., Rouleau, G. A., Gusella, J. S., Horvitz, H. R., & Brown, R. H. J.
(1993) Mutations in Cu/Zn superoxide dismutase gene are associ-
ated with familial amyotrophic lateral sclerosis. Nature 362, 5962.
Rothwell, N. J., & Hopkins, S. J. (1995) Cytokines and the nervous
system. II. Actions and mechanisms of action. Trends Neurosci. 18,
130136.
Rubin, E. H. (1990) Psychopathology of senile dementia of the
Alzheimer type. In: Advances in Neurology (R. J. Wurtman, S.
Corkin, J. H. Growdon, & E. Ritter-Walker, Eds.), pp. 5359. Raven
Press, NewYork.
Rubin, H., & Koide, T. (1973) Stimulation of DNA synthesis and
2-deoxy-D-glucose transport in chick embryo cultures by excessive
metal concentrations and by a carcinogenic hydrocarbon. J. Cell.
Physiol. 81, 387396.
St. George-Hyslop, P. H., Tanzi, R. E., Polinsky, R. J., Haines, J. L.,
Nee, L., Watkins, P. C., Myers, R. H., Feldman, R. G., Pollen, D.,
Drachman, D., Growdon, J., Bruni, A., Foncin, J. F., Salmon, D.,
Frommelt, P., Amaducci, L., Sorbi, S., Piacentini, S., Stewart, G. D.,
166 Cuajungco and Lees
Copyright

1997 by Academic Press
All rights of reproduction in any form reserved.
Hobbs, W. J., Conneally, P. M., & Gusella, J. F. (1987) The genetic
defect causing familial Alzheimers disease maps on chromosome
21. Science 235, 885890.
Samudralwar, D. L., Diprete, C. C., Ni, B. F., Ehmann, W. D., &
Markesbery, W. R. (1995) Elemental imbalances in the olfactory
pathway in Alzheimers disease. J. Neurol. Sci. 130, 139145.
Sato, M., & Bremner, I. (1993) Oxygen free radicals and metallothio-
nein. Free Radicals Biol. Med. 14, 325337.
Sato, M., Sasaki, M., Oguro, T., Kuroiwa, Y., & Yoshida, T. (1995)
Induction of metallothionein synthesis by glutathione depletion
after trans- and cis-stilbene oxide administration in rats. Chem.
Biol. Int. 98, 1525.
Sato, S. M., Frazier, J. M., & Goldberg, A. M. (1984) The distribution
and binding of zinc in the hippocampus. J. Neurosci. 4, 16621670.
Schafer, B. W., & Heizmann, C. W. (1996) The S-100 family of
EF-Hand calcium-binding proteins: Functions and pathology.
Trends Biochem. Sci. 21, 134140.
Scharfman, H. E., & Schwartzkroin, P. A. (1989) Protection of dentate
hilar cells from prolonged stimulation by intracellular calcium
chelation. Science 246, 257260.
Schwabe, J. W. R., & Klug, A. (1996) Zinc mining for protein
domains. Struct. Biol. 1, 345349.
Selinfreund, R. H., Barger, S. W., Pledger, W. J., & Van Eldik, L. J.
(1991) Neurotrophic protein S-100 stimulates glial cell prolifera-
tion. Proc. Natl. Acad. Sci. USA 88, 35543558.
Selkoe, D. J., Podlisny, M. B., Gronbeck, A., Mammen, A., & Kosik,
K. S. (1990) Molecular relation of amyloid laments and paired
helical laments in alzheimers disease. In: Advances in Neurology
(R. J. Wurtman, S. Corkin, J. H. Growdon, & E. Ritter-Walker,
Eds.), pp. 171179. Raven Press, NewYork.
Shang, Z., Liao, Y. D., Wu, F. Y. H., & Wu, C. W. (1989) Zinc release
from Xenopus transcription factor IIIAinduced by chemical modi-
cations. Biochem. 28, 97909795.
Sheng, J. G., Mrak, R. E., Rovnaghi, C. R., Kozlowska, E., Van Eldik,
L. J., & Griffin, W. S. T. (1996) Human brain S-100 and S-100
mRNAexpression increases with age: Pathogenic implications for
Alzheimers disease. Neurobiol. Aging 17, 359363.
Sherrington, R., Rogaev, E. I., Liang, Y., Rogaev, E. A., Levesque, G.,
Ikeda, M., Chi, H., Lin, C., Li, G., Holman, K., Tsuda, T., Mar, L.,
Foncin, J. F., Bruni, A. C., Montesi, M. P., Sorbi, S., Rainero, I.,
Pinessi, L., Polinsky, R. J., Wasco, W., Da Silva, H. A. R., Haines,
J. L., Pericak-Vance, M. A., Tanzi, R. E., Roses, A. D., Fraser, P. E.,
Rommens, J. M., & St. George-Hyslop, P. H. (1995) Cloning of a
gene bearing missense mutations in early-onset familial Alzhei-
mers disease. Nature 375, 755760.
Shimizu, T., Kubota, M., Tanizawa, A., Sano, H., Kasai, Y., Hashi-
moto, H., Akiyama, Y., & Mikawa, H. (1990) Inhibition of both
etoposide-induced DNA fragmentation and activation of poly-
(ADP-ribose) synthesis by zinc ion. Biochem. Biophys. Res. Commun.
169, 11721177.
Shiraga, H., Pfeiffer, R. F., & Ebadi, M. (1993) The effects of
6-hydroxydopamine and oxidative stress on the level of brain
metallothionein. Neurochem. Int. 23, 561566.
Shiraishi, K., Nakazawa, S., & Ito, H. (1993) Zinc enhances kainate
neurotoxicity in the rat brain. Neurol. Res. 15, 113116.
Sillevis-Smitt, P. A. E., Blaauwgeers, H. G. T., Troost, D., & Vianney
de Jong, J. M. B. (1992a) Metallothionein immunoreactivity is
increased in the spinal cord of patients with amyotrophic lateral
sclerosis. Neurosci. Lett. 144, 107110.
Sillevis-Smitt, P. A. E., Mulder, T. P. J., Verspaget, H. W., Blaauwgeers,
H. G. T., Troost, D., & Vianney de Jong, J. M. B. (1994) Metallothio-
nein in amyotrophic lateral sclerosis. Biol. Signals 3, 193197.
Sillevis-Smitt, P. A. E., van Beek, H., Baars, A. J., Troost, D.,
Louwerse, E. S., Krops-Hermus, A. C. M., de Wolff, F. A., &
Vianney de Jong, J. M. B. (1992b) Increased metallothionein in the
liver and kidney of patients with amyotrophic lateral sclerosis.
Arch. Neurol. 49, 721724.
Simonian, N. A., Getz, R. L., Leveque, J. C., Konrad, C., & Coyle, J. T.
(1996) Kainate induces apoptosis in neurons. Neuroscience 74,
675683.
Simons, T. J. B. (1991) Intracellular free zinc and zinc buffering in
human red blood cells. J. Membrane Biol. 123, 6371.
Skulachev, V. P., Chistyakov, V. V., Jasaitis, A. A., & Smirnova, E. G.
(1967) Inhibition of the respiratory chain by zinc ions. Biochem.
Biophys. Res. Commun. 26, 16.
Slomianka, L. (1992) Neurons of origin of zinc-containing pathways
and the distribution of zinc-containing boutons in the hippocam-
pal region of the rat. Neuroscience 48, 325352.
Slomianka, L., Danscher, G., & Frederickson, C. J. (1990) Labeling of
the neurons of origin of zinc-containing pathways by intraperito-
neal injections of sodium selenite. Neuroscience 38, 843854.
Sloviter, R. (1985) Aselective loss of hippocampal mossy ber Timm
stain accompanies granule cell seizure activity induced by perfo-
rant path stimulation. Brain Res. 330, 150153.
Sloviter, R. S. (1982) A simplied Timm stain procedure compatible
with formaldehyde xation and routine paraffin embedding of rat
brain. Brain Res. Bull. 8, 771774.
Smart, T. G. (1992) A novel modulatory binding site for zinc on the
GABA
A
receptor complex in cultured rat neurones. J. Physiol. 447,
587625.
Smart, T. G., Xie, X., & Krishek, B. J. (1994) Modulation of inhibitory
and excitatory amino acid receptor ion channels by zinc. Prog.
Neurobiol. 42, 393441.
Soc, E., Lange, K. W., Jellinger, K., & Riederer, P. (1992) Reduced
and oxidized glutathione in the substantia nigra of patients with
Parkinsons disease. Neurosci. Lett. 142, 128130.
Staal, F. J. T., Ela, S. W., Roederer, M., Anderson, M. T., Herzenberg,
L. A., Herzenberg, L. A. (1992) Glutathione deciency and human
immunodeciency virus infection. Lancet 339, 909912.
Stanley, L. C., Mrak, R. E., Woody, R. C., Perrot, L. J., Zhang, S. X.,
Marshak, D. R., Nelson, S. J., & Griffin, W. S. T. (1994) Glial
cytokines as neuropathogenic factors in HIV infection: Pathogenic
similarities to Alzheimers disease. J. Neuropathol. Exp. Neurol. 53,
231238.
Stengaard-Pedersen, K., Fredens, K., &Larsson, L. I. (1983) Compara-
tive localization of enkephalin and cholecystokinin immunoreac-
tivities and heavy metals in the hippocampus. Brain Res. 273,
8196.
Stennard, F. A., Holloway, A. F., Hamilton, J., & West, A. K. (1994)
Characterisation of six additional human metallothionein genes.
Biochim. Biophys. Acta. 1218, 357365.
Stewart, G. R., Frederickson, C. J., Howell, G. A., &Gage, F. H. (1984)
Cholinergic denervation-induced increase of chelatable zinc in
mossy-ber region of the hippocampal formation. Brain Res. 290,
4351.
Storm-Mathisen, J., & Ottersen, O. P. (1990) Immunocytochemistry
of glutamate at the synaptic level. J. Histochem. Cytochem. 38,
17331743.
Su, J. H., Anderson, A. J., Cummings, B. J., & Cotman, C. W. (1994)
Immunohistochemical evidence for apoptosis in Alzheimers dis-
ease. NeuroReport 5, 25292533.
Summers, M. F., South, T. L., Kim, B., & Hare, D. R. (1990) High
resolution structure of an HIV zinc ngerlike domain via a new
NMR-based distance geometry approach. Biochem. 29, 329340.
Zinc Metabolism in the Brain 167
Copyright

1997 by Academic Press
All rights of reproduction in any form reserved.
Sunderman, F. W. J. (1995) The inuence of zinc on apoptosis. Ann.
Clin. Lab. Sci. 25, 134142.
Sundstrom, L. E., Mitchell, J., & Wheal, H. V. (1993) Bilateral
reorganization of mossy bers in the rat hippocampus after a
unilateral intracerebroventricular kainic acid injection. Brain Res.
609, 321326.
Sutula, T., Cascino, G., Cavazos, J., Parada, I., & Ramirez, L. (1989)
Mossy ber synaptic reorganization in the epileptic human tempo-
ral lobe. Ann. Neurol. 26, 321330.
Szabo, C., Zingarelli, B., OConnor, M., & Salzman, A. (1996) DNA
strand breakage, activation of poly(ADP-ribose) synthetase, and
cellular energy depletion are involved in the cytotoxicity in
macrophages and smooth muscle cells exposed to peroxynitrite.
Proc. Natl. Acad. Sci. USA 93, 17531758.
Tanzi, R. E., Gusella, J. F., Watkins, P. C., Bruns, G. A. P., St. George-
Hyslop, P., Van Keuren, M. L., Patterson, D., Pagan, S., Kurnit,
D. M., &Neve, R. L. (1987) Amyloid protein gene: cDNA, mRNA
distribution, and genetic linkage near the Alzheimer locus. Science
235, 880884.
Taubeneck, M. W., Daston, G. P., Rogers, J. M., Gershwin, M. E.,
Ansari, A., & Keen, C. L. (1995) Tumor necrosis factor- alters
maternal and embryonic zinc metabolism and is developmentally
toxic in mice. J. Nutrition 125, 908919.
Telford, W. G., & Fraker, P. J. (1995) Preferential induction of
apoptosis in mouse CD
4
CD
8
--TCR
Io
CD3
Io
thymocytes by
zinc. J. Cell. Physiol. 164, 259270.
Terse, P. S., & Komiskey, H. L. (1997) Modulation of a competitive
N-methyl-D-aspartate receptor antagonist binding by zinc oxide.
Brain Res. 744, 347350.
Theoret, Y., Earnhardt, T. S., Bouldin, T. W., & Krigman, M. R. (1988)
The neurotoxicity of intrahippocampal kainic acid injection in rats
is not accompanied by a reduction of Timm stain. Brain Res. 449,
341346.
Thompson, C. B. (1995) Apoptosis in the pathogenesis and treatment
of disease. Science 267, 14561462.
Thompson, C. M., Markesbery, W. R., Ehmann, W. D., Mao, Y. X., &
Vance, D. E. (1988) Regional brain trace-element studies in
Alzheimers disease. Neurotoxicology 9, 17.
Thornalley, P. J., & Vasak, M. (1985) Possible role for metallothionein
in protection against radiation-induced oxidative stress: Kinetics
and mechanism of its reaction with superoxide and hydroxyl
radicals. Biochim. Biophys. Acta 827, 3644.
Tjian, R. (1995) Molecular machines that control genes. Sci. Am. 272,
3845.
Tonder, N., Johansen, F. F., Frederickson, C. J., Zimmer, J., & Diemer,
N. H. (1990) Possible role of zinc in the selective degeneration of
dentate hilar neurons after cerebral ischemia in the adult rat.
Neurosci. Lett. 109, 247252.
Treves, S., Trentini, P. L., Ascanelli, M., Bucci, G., & Di Virgilio, F.
(1994) Apoptosis is dependent on intracellular zinc and indepen-
dent of intracellular calcium in lymphocytes. Exp. Cell Res. 211,
339343.
Tsuji, S., Kobayashi, H., Uchida, Y., Ihara, Y., & Miyatake, T. (1992)
Molecular cloning of human growth inhibitory factor cDNA and
its down-regulation in Alzheimers disease. EMBO J. 11, 4843
4850.
Tummino, P. J., Scholten, J. D., Harvey, P. J., Holler, T. P., Maloney, L.,
Gogliotti, R., Domagala, J., & Hupe, D. (1996) The in vitro ejection
of zinc from human immunodeciency virus (HIV) type 1 nucleo-
capsid protein by disulde benzamides with cellular anti-HIV
activity. Proc. Natl. Acad. Sci. USA 93, 969973.
Turgeon, S. M., &Albin, R. L. (1992) Zinc modulates GABA
B
binding
in rat brain. Brain Res. 596, 3034.
Tymianski, M., Wallace, M. C., Spigelman, I., Uno, M., Carlen, P. L.,
Tator, C. H., & Charlton, M. P. (1993) Cell-permeant Ca
2
chelators
reduce early excitotoxic and ischemic neuronal injury in vitro and
in vivo. Neuron 11, 221235.
Uchida, Y., Takio, K., Titani, K., Ihara, Y., & Tomonaga, M. (1991) The
growth inhibitory factor that is decient in the Alzheimers
disease brain is a 68 amino acid metallothionein-like protein.
Neuron 7, 337347.
Vallee, B. L. (1983) Zinc in biology and biochemistry. In: Zinc
Enzymes (T. G. Spiro, Ed.), pp. 124. Wiley, NewYork.
Vallee, B. L. (1995) The function of metallothionein. Neurochem. Int.
27, 2333.
Vallee, B. L., Coleman, J. E., & Auld, D. S. (1991) Zinc ngers, zinc
clusters, and zinc twists in DNA-binding protein domains. Proc.
Natl. Acad. Sci. USA 88, 9991003.
Vallee, B. L., & Falchuk, K. H. (1993) The biochemical basis of zinc
physiology. Physiol. Rev. 73, 79117.
Van Eldik, L. J., Christie-Pope, B., Bolin, L. M, Shooter, E. M., &
Whetsell, W. O. J. (1991) Neurotrophic activity of S-100 in
cultures of dorsal root ganglia from embryonic chick and fetal rat.
Brain Res. 542, 280285.
Van Eldik, L. J., & Griffin, W. S. T. (1994) S-100 expression in
Alzheimers disease: Relation to neuropathology in brain regions.
Biochim. Biophys. Acta 1223, 398403.
Velasco, I., Tapia, R., & Massieu, L. (1996) Inhibition of glutamate
uptake induces progressive accumulation of extracellular gluta-
mate and neuronal damage in rat cortical cultures. J. Neurosci. Res.
44, 551561.
Vener, A. V., Aksenova, M. V., & Burbaeva, G. S. (1993) Drastic
reduction of the zinc- and magnesium-stimulated protein tyrosine
kinase activities in Alzheimers disease hippocampus. FEBS Lett.
328, 68.
Vepachedu, V. S. K. R., & Mohan, P. M. (1996) Efflux is a mechanism
for zinc resistance in Neurospora crassa. Biochemistry 35, 9301(76).
Walter, R. M. Jr., Oster, M. H., Lee, T. J., Flynn, N., & Keen, C. L.
(1990) Zinc status in human immunodeciency virus infection.
Life Sci. 46, 15971600.
Wang, H. Y., Pisano, M. R., & Friedman, E. (1994) Attenuated protein
kinase C activity and translocation in Alzheimers disease brain.
Neurobiol. Aging 15, 293298.
Wang, T. L., Hackam, A., Guggino, W. B., & Cutting, G. R. (1995) A
single histidine residue is essential for zinc inhibition of GABA1
receptors. J. Neurosci. 15, 76847691.
Ward, N. I., & Mason, J. A. (1987) Neutron activation analysis for
identifying elemental status in Alzheimers disease. J. Radioanal.
Nucl. Chem. 113, 515526.
Weiss, J. H., Hartley, D. M., Koh, J. Y., & Choi, D. W. (1993) AMPA
receptor activation potentiates zinc neurotoxicity. Neuron 10,
4349.
Wensink, J., Lenglet, W. J. M., Vis, R. D., & Van den Hamer, C. J. A.
(1987) The effect of dietary zinc deciency on the mossy ber zinc
content of the rat hippocampus. Histochemistry 87, 6569.
Wensink, J., Molenaar, A. J., Woroniecka, U. D., & Van den Hamer,
C. J. A. (1988) Zinc uptake into synaptosomes. J. Neurochem. 50,
782789.
Wenstrup, D., Ehmann, W. D., & Markesbery, W. R. (1990) Trace
element imbalances in isolated subcellular fractions of Alzhei-
mers disease brains. Brain Res. 533, 125131.
White, G., &Gurley, D. A. (1995) Alpha subunits inuence zinc block
of 2 containing GABA
A
receptor currents. NeuroReport 6, 461464.
168 Cuajungco and Lees
Copyright

1997 by Academic Press
All rights of reproduction in any form reserved.
Whitemore, P. J., Price, D. L., Struble, R. G., Clark, A. W., Coyle,
J. T., & DeLong, M. R. (1982) Alzheimers disease and senile
dementia: Loss of neurons in the basal forebrain. Science 215,
12371239.
Williams, R. J. P. (1989) An introduction to the biochemistry of zinc.
In: Zinc in Human Biology (C. F. Mills, Ed.), pp. 1531. Springer-
Verlag, London.
Williamson, A., & Spencer, D. (1995) Zinc reduces dentate granule
cell hyperexcitability in epileptic humans. NeuroReport 6, 1562
1564.
Willson, R. L. (1989) Zinc and iron in free radical pathology and
cellular control. In: Zinc in Human Biology (C. F. Mills, Ed.), pp.
147172. Springer-Verlag, London.
Wurtman, R. J., Blusztajn, J. K., Ulus, I. H., Coviella, I. L. G.,
Buyukuysal, R. L., Growdon, J. H., & Slack, B. E. (1990) Choline
metabolismin cholinergic neurons: Implications for the pathogen-
esis of neurodegenerative diseases. In: Advances in Neurology (R. J.
Wurtman, S. Corkin, J. H. Growdon, & E. Ritter-Walker, Eds.), pp.
117125. Raven Press, NewYork.
Wyllie, A. H. (1980) Glucocorticoid-induced thymocyte apoptosis is
associated with endogenous endonuclease activation. Nature 284,
Kerr, J. F. R., & Currie, A. R. (1973) Cell death in the normal
neonatal rat adrenal cortex. J. Pathol. 111, 255261.
Xie, X., Gerbert, U., Gahwiler, B. H., & Smart, T. G. (1993) Interaction
of zinc with ionotropic and metabotropic glutamate receptors in
rat hippocampal slices. Neurosci. Lett. 159, 4650.
Xu, J., Xu, Y., Nguyen, Q., Novikoff, P. M., & Czaja, M. J. (1996)
Induction of hepatoma cell apoptosis by c-myc requires zinc and
occurs in the absence of DNA fragmentation. Am. J. Physiol. 270,
G60G70.
Yang, J. P., Merin, J. P., Nakano, T., Kato, T., Kitade, Y., &Okamoto, T.
(1995) Inhibition of DNA-binding activity of NF-B by gold
compounds in vitro. FEBS Lett. 361, 8996.
Yang, Q., Hamberger, A., Wang, S., & Haglid, K. G. (1996) Appear-
ance of neuronal S-100 during development of the rat brain. Dev.
Brain Res. 91, 181189.
Yao, J., Kitt, C., & Reeves, R. H. (1995) Chronic elevation of S-100
protein does not alter APP mRNAexpression or promote -amyloid
deposition in the brains of aging transgenic mice. Brain Res. 702,
3236.
Yasui, M., Ota, K., & Garruto, R. M. (1993) Concentrations of zinc
and iron in the brains of Guamanian patients with amyotrophic
lateral sclerosis and parkinsonism-dementia. Neurotoxicol. 14,
445450.
Yim, M. B., Kang, J. H., Yim, H. S., Kwak, H. S., Chock, P. B., &
Stadtman, E. R. (1996) Again-of-function of an amyotrophic lateral
sclerosis-associated Cu,Zn-superoxide dismutase mutant: An en-
hancement of free radical formation due to a decrease in K
m
for
hydrogen peroxide. Proc. Natl. Acad. Sci. USA 93, 57095714.
Yin, H. Z., & Weiss, J. H. (1995) Zn
2
permeates Ca
2
permeable
AMPA/kainate channels and triggers selective neural injury.
NeuroReport 6, 25532556.
Yokoyama, M., Koh, J., & Choi, D. W. (1986) Brief exposure to zinc is
toxic to cortical neurons. Neurosci. Lett. 71, 351355.
Yu, C., & Crutcher, K. A. (1995) Nerve growth factor immunoreactiv-
ity and sympathetic sprouting in the rat hippocampal formation.
Brain Res. 672, 5567.
Yurchenco, P. D., & Schittny, J. C. (1990) Molecular architecture of
basement membranes. FASEB J. 4, 15771590.
Zabel, U., Schreck, R., & Baeuerle, P. A. (1991) DNA binding of
puried transcription factor NF-B. J. Biol. Chem. 266, 252260.
Zahradka, P., &Ebisuzaki, K. (1984) Poly(ADP-ribose) polymerase is
a zinc metalloenzyme. Eur. J. Biochem. 142, 503509.
Zalewski, P. D., Forbes, I. J., & Betts, W. H. (1993) Correlation of
apoptosis with change in intracellular labile Zn(II) using Zinquin
[(2-methyl-8-p-toluenesulphonamido-6-quinolyloxy) acetic acid],
a newspecic uorescent probe for Zn(II). Biochem. J. 296, 403408.
Zalewski, P. D., Forbes, I. J., Giannakis, C., & Betts, W. H. (1991)
Regulation of protein kinase C by Zn
2
-dependent interaction
with actin. Biochem. Int. 24, 11031110.
Zalewski, P. D., Forbes, I. J., Giannakis, C., Cowled, P. A., & Betts,
W. H. (1990) Synergy between zinc and phorbol ester in transloca-
tion of protein kinase C to cytoskeleton. FEBS Lett. 273, 131134.
Zalewski, P. D., Forbes, I. J., Seamark, R. F., Borlinghaus, R., Betts,
W. H., Lincoln, S. F., & Ward, A. D. (1994) Flux of intracellular
labile zinc during apoptosis (gene-directed cell death) revealed by
a specic chemical probe, Zinquin. Chem. Biol. 1, 153161.
Zeng, J., Heuchel, R., Schaffner, W., & Kagi, J. H. R. (1991a) Thionein
(apometallothionein) can modulate DNA binding and transcrip-
tion activation by zinc nger containing factor Sp1. FEBS Lett. 279,
310312.
Zeng, J., Vallee, B. L., & Kagi, J. H. R. (1991b) Zinc transfer from
transcription factor IIIA ngers to thionein clusters. Proc. Natl.
Acad. Sci. USA 88, 99849988.
Zhang, G. H., Yamaguchi, M., Kimura, S., Higham, S., & Kraus-
Friedmann, N. (1990) Effects of heavy metal on rat liver micro-
somal Ca
2
-ATPase and Ca
2
sequestering. J. Biol. Chem. 265,
21842189.
Zhao, H., & Eide, D. (1996) The yeast ZRT1 gene encodes the zinc
transporter protein of a high-affinity uptake system induced by
zinc limitation. Proc. Natl. Acad. Sci. USA 93, 24542458.
Zinc Metabolism in the Brain 169
Copyright

1997 by Academic Press
All rights of reproduction in any form reserved.

Вам также может понравиться