Вы находитесь на странице: 1из 13

Current Drug Targets - CNS & Neurological Disorders, 2005, 4, 541-552 541

Protein Kinase C Isozymes: Memory Therapeutic Potential

Miao-Kun Sun*,1 and Daniel L. Alkon1,2

1
Blanchette Rockefeller Neurosciences Institute, West Virginia University and 2Johns Hopkins University, Rockville, MD.
USA

Abstract: PKC plays an important role in many types of learning and memory. Evidence has been provided that PKC
activation and translocation are induced in associative learning tasks. PKC inhibition, on the other hand, impairs
learning and memory, consistent with the observations that transgenic animal models with a particular PKC isoform
deficit exhibit impaired capacity in cognition. The dramatic impact of PKC pharmacology on learning and memory is
further emphasized by a regulatory role of PKC isozymes in amyloid production and accumulation. Recent study reveals
that PKC activation greatly reduces neurotoxic amyloid production and accumulation. PKC activators, therefore, may
have important therapeutic values in the treatment of dementia, especially when fine-tuning of selective isoform
activity can be effectively achieved pharmacologically, with further development of isozymes-specific agents. The
success of antidementia therapy with agents that act on PKC signaling cascades depends on whether such agents at their
effective doses would significantly disrupt or interfere with other vital functions that rely on a narrow range of PKC
activities.

Protein kinase C (PKC), ubiquitous in the central nervous associated with the development of Alzheimer’s disease (AD)
system and activated by Ca2+, phospholipids and and other neurodegenerative disorders. Changes in synaptic
diacylglycerol, or phorbol-ester [1], has been found to mediate transmission and signal processing in this structure are bound
various extracellular signals into the cell and intracellular to have dramatic consequences on cognition and behavior.
signal events [2]. The PKC signaling pathway plays an Expression of a PKC inhibitor in Purkinje cells using the
important and regulatory role in a wide range of vital promoter of the Purkinjie cell-specific gene pcp-2(L7) in mice
biological functions and processes [3-5], including has also been reported to block cerebellar long-term depression
proliferation, altered gene expression, synaptic plasticity, (LTD) and adaptation of the vestibulo-ocular reflex [19]. These
neuronal injury, differentiation, cell growth and apoptosis, and and other evidence (see below) point to an important role of
oncogenesis, depending on cell identity and states. various isoforms of PKC in the formation of memory traces
and thus potential opportunities in the development of
The importance of a normal functional control through
cognition therapeutic agents that modify PKC activity and
PKC signaling cascades for brain health and functions is well
signal cascade(s). On the other hand, PKC is sensitive to a
established. The brain has, in fact, the highest concentration of
variety of neurotoxic/injurious events and may mediate some
PKC than any organ in the body [6], especially in vital neural
pathological responses of the nervous system to injury. In this
structures that are involved in cognition and mood regulation.
short review article, we intend to summarize the roles of PKC
In the nervous system, activation of PKC leads to changes in
in learning and memory and neural injury, and to indicate
such vital responses as enhancing calcium action potentials,
potential of developing cognition therapeutic agents that act on
increasing neurotransmitter release, and decreasing voltage-
+ + the PKC system.
gated Na currents [7-9] and voltage-dependent K currents [10,
11] as well as calcium-activated potassium current in
hippocampus, all relevant to information processing in PROTEIN KINASE C ISOFORMS AND MOLECULAR
cognition. Particularly, PKC activation in the hippocampus TARGETS
and related neural structures produces potentiation of synaptic PKC activity, with the active site usually located at the C-
responses, mimicking those seen in long-term potentiation terminus to phosphorylate serine and threonine residues, is
(LTP) of the glutamatergic post-synaptic responses [12] and mediated by a family of at least twelve PKC isoforms [20-22]
enhanced summation of excitatory post-synaptic potentials (Fig. 1). Each isoform of the twelve identified PKC isozymes
with rabbit eye blink conditioning [13]. The hippocampus in is encoded by a separate gene, with the exception of the I and
mammals is involved in many types of cognitive performances II isoforms, which are splice variants.
as well as mood regulation [14-17] and plays a critical role in
The structural features of PKC include two main regulatory
episodic, declarative, and spatial learning and memory [18]. It 2+
domains (the activator-binding C1 and the Ca -binding C2
is a highly plastic and vulnerable brain structure that may be
domains) at the NH2-terminal that mediate membrane
damaged in major depression and by a variety of injuries that
association and activation, a C-terminal active site that
change PKC signaling cascades. Dysfunction of PKC is
functions as a serine/threonine kinase, and a pseudosubstrate
sequence adjacent to the C1 domain (Fig. 1). The C2 domain
*Address correspondence to the author at the Blanchette Rockefeller also contains binding sites for lipids and proteins. The
Neurosciences Institute, 9601 Medical Center Drive, Academic and catalytic domain of PKC isozymes has the C3 and C4
Research Bldg., Room 319, Rockville, MD. USA; E-mail: mksun@brni- domains. The former possesses the binding site for ATP, the
jhu.org phosphate donor for phosphotransferase activity. The latter has

1568-007X/05 $50.00+.00 © 2005 Bentham Science Publishers Ltd.


542 Current Drug Targets - CNS & Neurological Disorders, 2005, Vol. 4, No. 5 Sun and Alkon

Fig. (1). Structures of the PKC isozymes. PS: pseudosubstrate.

the binding site for substrates. The C1 domains represent zinc in most cases which particular PKC isozyme(s) may mediate
2+
finger structures that bind diacylglycerol and other non-Ca the observed effects. Obviously, signal cascades that are not
PKC activators. For instance, the regulatory domains of novel triggered by calcium waves are better relayed by the calcium-
PKC (nPKC) contain two C1 sub-domains (C1A and C1B), independent pathways. aPKC appears to have higher structural
each of which interact with sn-1,2-diacylglycerol and phorbol restriction for activators than the other subgroups, although the
esters [23-25]. The C1A and C1B domains are, however, not biological meaning of such difference remains to be studied.
equivalent, with their relative affinities depending on the Nevertheless, evidence has been provided that atypical PKC
isozymes. The C1A domain of PKC, for example, has high whose expression levels are correlated with nucleotide excision
affinity for diacylglycerol and is critical for the diacylglycerol- repair activity and protein levels, probably regulates the
induced membrane binding and PKC activation, while the expression of nucleotide repair proteins [33]. Unfortunately, the
C1B domain exhibits high affinity for phorbol esters [26]. PKC pharmacology has been limited by the availability of
These differences indicate potential for developing isozyme- water soluble, specific inhibitors of PKC isozymes for an in
selective PKC activators. The activator-binding sites within the vivo administration. The pharmacological antagonism and
C1 domain of PKC are also able to bind alcohols and isozyme involvement therefore remain to be studied in detail
anesthetics [27]. In the inactive conformation, a in the future.
pseudosubstrate sequence interacts with the substrate-binding PKC is activated by synaptic inputs and intracellular
region, keeping the enzyme in the inactive state. signals that are involved in information processing in
The twelve PKC isoforms can be further divided into three cognition, including glutamatergic inputs [34], cholinergic
subgroups, cPKC, nPKC, and aPKC, based on their molecular inputs [9], serotonergic inputs [7, 8], dopaminergic inputs
structures and thus co-factor requirements. The cPKC, or [35], intracellular calcium and diacylglycerol elevations, and
classical PKCs, contains the activator-binding C1 domain and other hormones [36]. It is gradually recognized that structural
2+
the C2 region corresponding to the Ca binding site [28]. The plasticity may also play an important role in signal processing
2+
cPKC requires Ca as a co-factor for activation and consists of in cognition and involves PKC activity. Thus, in addition to
,  I, II, and  isoforms. The nPKC, including ,
,
’, , , an essential involvement of PKC in synaptic plasticity and
and  isoforms, on the other hand, is calcium-independent transmission, PKC activity signals growth. Growth and other
2+
[28]. These isoforms do not have the Ca binding C2 region stimuli cause a transient increase in levels of cellular
[28-30] and do not require Ca2+ as a co-factor for activation. diacylglycerol, which activates PKC as well as produces some
They contain a C2-like structure, which may be involved in other effects that appear to be PKC-independent [37]. PKC
interaction with proteins and phospholipids. A common feature signaling triggered by local cell-cell contact appears to be a
of these two subgroup PKCs is that they contain both C1A mechanism through which astrocytes regulate neuronal
and C1B sub-domains and can all be activated by development. The mechanism involved can be briefly described
diacylglycerol, phorbol esters, and bryostatins. The third as that local contact with astrocytes via integrin receptors
group, aPKC, atypical PKC isozymes, includes  and / elicits PKC activation and excitatory synaptogenesis in
isoforms. The aPKC contains a single C1 domain and does not neurons, through the arachidonic acid cascade [38]. The cascade
bind diacylglycerol, phorbol esters, or bryostatins. It does not represents an underlying mechanism for global neuronal
2+
contain the Ca binding C2 region and one of the repeated maturation following local astrocyte adhesion, probably
cysteine-rich zinc finger binding motifs within the C1 domain including neurogenesis in adults such as that matured
[31, 32]. It is not clear how aPKC activity is regulated in vivo. hippocampal astrocytes regulate neurogenesis by instructing
Many of the isoforms are expressed in the same cells. The stem cells to adopt a neuronal fate [39, 40]. It is not clear,
however, whether the integrin-PKC-cascade plays an essential
biological significance of the heterogeneity of the PKC family
role in the formation of memory traces in the brain in adults.
has not been really established. For instance, we do not know
Protein Kinase C Isozymes: Memory Therapeutic Potential Current Drug Targets - CNS & Neurological Disorders, 2005, Vol. 4, No. 5 543

PKCs interact with a large diverse family of PKC- in the dissociation of the kinase. Such dissociation, however,
interacting proteins [4, 41, 42], including the receptors for is effectively prevented in the presence of high-frequency
activated C-kinase (RACKs). These phosphoproteins are the calcium spikes and persistent calcium increases. Only high-
PKC substrates, through which PKC regulates cellular frequency calcium spikes and persistent calcium increases can
functions. One of the phosphoproteins that are heavily lead to an integration of cPKC activity, inducing a much
phosphorylated by PKC is localized to the presynaptic higher amplitude of and lasting kinase activity.
terminal. It was first identified by Gispen and his colleagues in While cPKC is not activated when calcium spikes are
the Netherlands and was named as B50 by them for the brain absent, other PKC isoforms are well positioned to respond to
protein with an approximate molecular weight of 50 kDa (for signal events without associated calcium spikes. Protein kinase
review, see [43]. It was then named as F1 [44] and GAP-43 for C regulates the nuclear localization of diacylglycerol kinase-
a growth-associated protein with an approximate molecular [58], which terminates signaling from diacylglycerol by
weight of 43 kDa [45]. GAP-43 is found exclusively in the converting it to phosphatidic acid. The nuclear-localization
brain [43], is localized to the membranes of axonal processes, signal of diacylglycerol kinase- is localized in a region that is
and is not found in membranes of dendritic processes ([46-48], homologous to the phosphorylation-site domain of the
except in the olfactory system [49]). Its expression is up- MARCKS protein, a major target for PKC. Reducing nuclear
regulated during spatial learning and memory [50]. Transgenic diacylglycerol levels by diacylglycerol kinase- attenuates cell
mice overexpressing GAP-43 have been found to exhibit an growth.
enhanced memory in a maze task [44], while heterozygous
GAP-43 knockout mice have impaired hippocampus-dependent PROTEIN KINASE C AND MEMBRANE
memory [51]. RECEPTOR/CHANNEL ACTIVITY
The separate localization of PKC and its membrane-
Receptor/channel states and activity have dramatic
associated substrates, such as GAP-43, in the cells dictates an
consequence on signal transfer through the neural network and
essential step in the PKC activation and translocation. PKC
synaptic plasticity that underlie learning and memory. These
translocation to membrane component thus becomes a marker
include the potassium channels, glutamatergic cationic
indicating PKC activation and is commonly studied. Early
channels, voltage-gated Na+ channels, and -aminobutyric acid
studies defined an important involvement of PKC activation in
(GABA)A receptor channels, all sensitive to changes in PKC
associative learning [10]. Such activation is indeed associated
activity. The calcium- and phospholipid-dependent PKCs
with a characterized PKC translocation from cytosol to
control membrane electrical properties [59, 60] and modulate
membrane component [52, 53], an essential step leading to
neurotransmitter receptor function [61-63]. For instance,
substrate activation. The PKC translocation from the
activation of PKC by phorbol esters and synthetic
cytoplasm to the membrane component of the cells is
diacylglyceroles has been shown to attenuate voltage-gated
associated with phosphorylation of GAP-43 and with the
potassium currents in a number of cell types, including the
induction of LTP [54]. The phenomenon of PKC translocation
CA1 pyramidal cells, the “place” cells, in the hippocampus
occurs presynaptically, as in the case of phosphorylation of
[10, 11, 19, 64-67]. PKC phosphorylates the voltage-gated
GAP-43 [54] and LTP [55], as well as postsynaptically, as in +
Na channels in hippocampal neurons and is responsible for
the case of associative learning [52]. +
PKC-mediated reduction of Na currents in the neurons [9]. A
reduction in the voltage-gated Na+ currents is expected to
PROTEIN KINASE C AND CELLULAR SIGNALING
increase spike threshold and reduce spike train duration, as
In memory formation, it is generally believed that the shown in prefrontal cortex neurons [7, 8]. Reduction in
frequency and intensity of calcium spikes/waves represent potassium channel conductance, on the other hand, represents
important signal of the experienced events. an important change that is associated with potentiated
synaptic transmission and many types of learning and memory.
PKC can ‘detect’ the calcium spike frequency and intensity,
thus the pattern and frequency of synaptic inputs, in a unique PKC modulates the glutamate system, including the
pattern and cycle of calcium-induced translocation and ionotropic receptors and metabotropic receptors. PKC
diacylglycerol-mediated kinase activation [56]. For instance, activation, for example, with 12- O-tetradecanoylphorbol
receptor stimuli that trigger repetitive calcium spikes induce modifies both -amino-3-hydroxy-5-methyl-4-
PKC in a way in which PKC responds to persistent isoxazoleproprionate receptor (AMPAR) and N-methyl-D-
diacylglycerol increases combined with high- but not low- aspartate receptors (NMDAR) mobility, increasing their
frequency calcium spikes [57]. This is because calcium acts extrasynaptic and synaptic diffusion [68]. In addition, PKC
rapidly. The calcium spike can translocate PKC and minimally activation has been found to result in phosphorylation of
activate cPKC. Diacylglycerol binding to PKC is initially glutamate receptor (GluR) 2/3 (at serine 880) in the Purkinje
prevented by a pseudosubstrate clamp, keeping the cells. PKC activation in the hippocampal CA1 pyramidal cells
diacylglycerol-binding site inaccessible and delaying calcium- has also been reported to produce an inhibition of the slow
and diacylglycerol-mediated kinase activation. After after-hyperpolarization, thus, resulting in an enhanced
termination of calcium signals, bound diacylglycerol prolongs excitability. Evidence has also been provided that metabotropic
kinase activity. Once diacylglycerol concentration is high, low- GluR (mGluR)6-containing kainate receptors are probably the
frequency calcium spikes lead to a small and oscillating cPKC trigger for PKC-mediated inhibition of slow after-
activity with an activity that is reversed after each spike. hyperpolarization in the hippocampal CA1 pyramidal neurons
Diacylglycerol alone is not sufficient to retain PKC at the [69]. Glutamate also desensitizes mGluR5a and mGluR5b,
plasma membrane. A drop in calcium concentration will result
544 Current Drug Targets - CNS & Neurological Disorders, 2005, Vol. 4, No. 5 Sun and Alkon

through PKC-mediated phosphorylation of mGluR5 at neurons, for example, activation of PKC induces rapid
multiple sites [70]. morphological plasticity in dendrites and spines [82].
In the brain at large and in the hippocampus particularly, The involvement of PKC activity in synaptic plasticity has
the GABA system plays an important role in controlling been investigated intensively. Early studies have revealed that
neuronal activity and information transfer through the neural by using inhibitors for both kinases (PKC and CaMKII), LTP
network related to cognition. PKC also appears to mediate of the glutamatergic synapses is sustained by persistent protein
brain-derived neurotrophic factor (BDNF)-induced modulation kinase activity but is not due to a substrate that remains stably
of GABAA receptor phosphorylation and activity [71]. The phosphorylated [83]. LTP is not due to a long-lived kinase
impact of BDNF-PKC-GABA transmission cascade in the activator, since the persistently active kinase exists in the
hippocampus and related brain structures on cognition and absence of an activator. However, the same study suggests that
mood remains to be evaluated in the future. the process of autophosphorylation, while potentially
contributing to the persistent kinase activity, does not
PROTEIN KINASE C AND NEUROTRANSMISSION completely account for the sustained kinase activation. When
more selective antagonists became available, the antagonists to
It is well-established that synaptic transmission through either PKC or CaMKII were postsynaptically injected into
neurotransmitter release is fundamental to signal processing hippocampal pyramidal cells, resulting in a blockade of
between neurons and through neural networks. PKC appears to induction of LTP of the glutamatergic synapses [84, 85]. Thus,
play an important role in the regulation of synaptic both PKC and calcium-calmodulin-dependent protein kinase II
transmission in the nervous system. are involved in the induction of LTP. These investigators also
First, the synthesis of many neurotransmitters is under found that once LTP was established, intracellular injection of
PKC regulation. For instance, cholinergic neurons play an the non-selective kinase inhibitor H-7 had no effect.
important role in learning and memory and their damaged Antagonists that block the autophosphorylation of CaMKII
function in fact characterizes the early brain injury and memory were also shown to block the induction of LTP (for review, see
deficits in AD patients. Choline acetyltransferase synthesizes [86]. These are exclusively postsynaptic. However, Malinow et
acetylcholine in the cholinergic neurons. The enzyme is al. [83] showed that potentially there is presynaptic
differentially phosphorylated by PKC isoforms on four serines involvement. When H-7 was applied extracellularly instead of
(Ser-440, -346, -347, -476) and one threonine (Thr-255), injecting it intracellularly, LTP remains sensitive to
regulating its catalytic activity [72]. PKC activity plays an extracellular kinase inhibition. Thus, while intracellular
important regulatory role in GABAergic synaptic transmission postsynaptic events are insensitive to kinase inhibition, once
[73]. In the glutamate system, phorbol esters have been shown LTP is established, events occurring outside of the intracellular
to increase the size of the readily releasable pool and the rate at postsynaptic realm remain sensitive to kinase inhibition.
which the pool refills at glutamatergic hippocampal synapses. Location exposed to extracellularly applied H-7 includes the
This effect is produced by a PKC-dependent mechanism [74]. presynaptic terminal as well as pre- and postsynaptic events
The PKC pathway may therefore regulate synaptic strength by occurring within the domain of the synaptic cleft. Of course,
modulating the readily releasable pool of the glutamatergic the contribution and maintenance of LTP differ depending on
vesicles. The calcium- and phospholipid-dependent PKCs also the fiber tract and cell group that is being examined [87].
regulate neurotransmitter release [74-76] and regulate gene However, the PKC effects on LTP of the glutamatergic
expression in mature neurons [77]. However, others report that synapses may not necessarily mean corresponding effects on
phorbol ester- and diacylglycerol-induced augmentation of cognition. Dissociation of memory performance from LTP has
transmitter release is mediated by Munc13s, a presynaptic been observed and reported frequently. For instance, mice
protein that contains diacylglycerol/phorbol ester binding C1 deficit in PKC have been found to show normal brain
domain, but not by PKCs [78]. anatomy and normal hippocampal synaptic transmission,
normal paired facilitation, and normal LTP of the
PROTEIN KINASE C AND SYNAPTIC PLASTICITY glutamatergic synaptic responses, but a loss of learning and
memory in both cued and contextual fear conditioning [88].
It is generally believed that learned experience depends on
the expression of modified synaptic transmission in the The involvement of PKC activity in LTD of glutamatergic
relevant synapses. The involvement of PKC isozymes in synapses has also been suggested. LTD in the hippocampal
synaptic plasticity and learning and memory has been the focus CA1 field is associated with a transient decrease in pre- and/or
of many investigations. Activation of PKC leads to the postsynaptic PKC substrate phosphorylation [89, 90]. But, the
phosphorylation of numerous proteins, including glutamate exact impact of such an involvement on learning and memory
receptor 2/3 (at serine 880) in Purkinje cells. The glutamate remains to be studied.
receptor 2/3 phosphorylation appears to be the critical step for Not only does synaptic facilitation and synaptogenesis
parallel fiber LTD [79]. In the neural networks, PKC activation involve PKC activation, synaptic activity-induced synapse
generally facilitates synaptic plasticity, including such elimination may also involve PKC. One possibility is that
responses as an enhancement of calcium influx, an increase in different isoforms may be involved. PKC in both presynaptic
neurotransmitter release, a decrease in a calcium-activated and postsynaptic elements has been proposed to play an
current in the hippocampus, actions that produce a potentiation important role in activity-dependentsynaptic elimination at the
of synaptic responses [10-13, 56, 80, 81]. These functional neuromuscular synapse [91]. PKC-deficient neurons that
changes are likely associated with morphological changes that innervate normal muscle exhibited a marked deficit in activity-
are also induced by PKC activation. In cultured hippocampal related synapse reduction.
Protein Kinase C Isozymes: Memory Therapeutic Potential Current Drug Targets - CNS & Neurological Disorders, 2005, Vol. 4, No. 5 545

PROTEIN KINASE C AND MEMORY basolateral nucleus of the amygdala in mice [88]. Mice
deficient in PKC have been found to exhibit a loss of learning
The central question in memory pharmaceutics based on in both cued and contextual fear conditioning [88] although
PKC activity modulation is the importance of PKC activity in their brain anatomy, hippocampal synaptic transmission,
learning and memory, and in memory disorders. The evidence responses to paired facilitation, and LTP of the hippocampal
available indicates an essential role of PKC in learning and glutamatergic synapses all appear normal.
memory. PKCs have been implicated in many types of
cognitive tasks, including learning and memory of eye blink One important anti dementia action of PKC activation
conditioning [80, 81, 92-94], spatial learning and memory [52, could be to facilitate PKC translocation and has been observed
95-99], olfactory discrimination learning [100], conditioned to occur in a variety of associative memory tasks [105]. A
taste aversion [101], contextual fear memory [102, 103], and second potentially important action of PKC in anti-dementia
conditioned avoidance [104]. is its role in regulation and modulation of amyloid production.
Activation of PKC and  is known to enhance sAPP
The involvement of PKC in brain functions and disorders, production [117, 118] and to reduce A production [119].
including cognition, mood regulation, and neurodegeneration, Evidence has been provided that the administration of
is implicated by several factors. Brain, especially the bryostatin-1 reduced A40 in the brains of AD transgenic mice
hippocampus and related structures that play critical roles in and both brain A40 and A42 in AD double-transgenic mice
cognition, express the highest levels of PKC in the body. In [120]. The action has an obvious therapeutic value for
addition, neural events and activated inputs that occur in bryostatin-1 to be an anti-dementic agent, as long as neurotoxic
learning and memory activate PKC. Associative learning amyloid defines much of the pathogenesis of AD. Some of this
produces translocation of PKC activity from the cytosolic to apparent therapeutic benefit of bryostatin-1 may be produced
the membrane compartment of the CA1 region of the through its persistent activation of PKC in the absence of
hippocampus [52]. Glutamate activates PKC in the down-regulation of the isoform [121].
hippocampus [34]. PKC may also mediate, at least partly,
arginine vasopressin-induced facilitation of memory retention Furthermore, some of the AD neuropathology and
in a step-through passive avoidance task in mice [36]. functional impairment in cognition may be mediated by an A-
induced PKC inhibition. As mentioned above, all of the PKC
PKC is activated in many cognitive tasks with enhanced isozymes, except PKC (human) and its murine homolog,
phosphorylation of its substrate proteins. GAP-43, for PKD, contain a pseudosubstrate motif in the regulatory region.
example, has been found to be phosphorylated in association The motif is an autoinhibitory domain of PKCs, maintaining
with LTP (for review, see [43]) and the phosphorylation of this their inactive state. It turns out that As contain a putative
presynaptic protein lasts for several days. Bryostatin-1, a PKC PKC pseudosubstrate domain (A 28-30). A is found in the
activator, enhances spatial memory in rats [105]. Furthermore, cerebrospinal fluid and blood of normal human subjects.
reduced PKC activity is associated with AD dementia [106, Evidence has been presented that A (1 M) directly inhibits
107] and suicide victims [108], suggesting an involvement in PKC and  activation by phorbol-12,13-dibutyrate [122]. The
cognition and mood regulation. Age-related spatial memory action may play a physiological role and/or a
impairment is associated with altered subcellular concentrations pathophysiological one. Through its binding to PKC, A can
of PKC and may be indicative of deficient signal transduction induce PKC degradation [107], reduce PKC-mediated
and neuronal plasticity in the hippocampal formation. The phosphorylation [123, 124], and decrease PKC membrane
connection between PKC and memory and dementia is translocation [125]. In cultured brain endothelial cells, A1-40
illustrated by a decreased PKC and  levels and activities in induced translocation of PKC from membrane fraction to
the temporal cortex [109, 110] without changes in PKC cytosol [125]. The interaction between PKC and A suggests
mRNA [111] and in crude extracts from the hippocampus, that restoration of the inhibited PKC activity may have direct
temporal and frontal cortex [112]. Activity of PKC enzyme therapeutic effects to cognition. Furthermore, recent evidence
extract purified from endogenous modulators is not modified indicates that PKC activation inhibits glycogen synthase 3
in all brain areas of AD patients [113]. These observations kinase [126, 127] and thereby reduces tau phosphorylation
suggest either changed isoform degradation or increased [128]. This last event is considered important for the
endogenous inhibitors of PKC activity in AD brain. Another production of another pathological hallmark of AD:
important change is RACKs, the substrate for activated PKCs. neurofibrillary tangles.
RACKs are important regulators of PKC function [114].
Deficit in RACK1 levels has been reported in both soluble and PKC ACTIVATORS
particulate fractions from AD brain frontal cortex [115].
The main focus of agents that act on PKC isoforms and
PKC inhibition has been consistently found to significantly
cascades as potential memory therapeutics is PKC activators,
impair performance of cognitive tasks. Intracerebroventricular
since PKC activators facilitate synaptic plasticity, enhance
injection of PKC inhibitors, for example, has been found to
learning and memory, reduce neurotoxic amyloid production
cause marked memory impairment in passive avoidance task
and accumulation, and inhibit tau phosphorylation.
and water maze task [116]. Relationships between spatial
learning (water maze) in rats and hippocampal concentrations PKC activators, such as diacylglycerol, arachidonic acid
of calcium-dependent PKC are isoform-specific and mainly of (Fig. 2), phorbol esters, bryostatins, aplysiatoxins, and
PKC isoform [98]. Studies with transgenic animals further teleocidins, bind to a hydrophilic cleft in a largely hydrophobic
support the critical role of PKC in learning and memory. surface of the C1 domains, resulting in an enhanced
PKC, for instance, was found to predominantlyexpress in the hydrophobicity of the surface and promoting the interaction
neocortex, in area CA1 of the hippocampus, and in the between the C1 domain and the phospholipid bilayer of the
546 Current Drug Targets - CNS & Neurological Disorders, 2005, Vol. 4, No. 5 Sun and Alkon

O
H
R O COOH
O R

O
OH
Arachidonic acid
(S)-1,2-Diacyl-sn-glycerol
O O
O O
O O
O O

H
H H
H H
H
OH
OH
OH
OH OH
OH O
O
Phorbol 12,13-dibutyrate Phorbol 12-myristate 13-acetate

OH
O OAc
B A
O O O

3 O
OH H OH
O O
C
26
O OH
O
O
O

Bryostatin-1

Fig. (2). Chemical structures of some PKC activators. R and R’ are either saturated or unsaturated hydrocarbon chains.

cell membranes and driving removal of the pseudosubstrate a partial agonist, can selectively modulate PKC isozymes, such
region from the catalytic site of the enzyme. as PKC, PKC and PKC [132-135], at very low
concentrations (at nM or sub-nM concentrations), but prevents
Diacylglycerol is an endogenous PKC activator, which
tumor cell growth, probably through a selective down-
binds to the C1 domain of cPKC and nPKC, with binding
3 regulation of PKC in several cancer cell lines [136, 137] and
affinity at M levels (for displacing bound [20- H]phorbol
preventing certain PKC from undergoing down-regulation
12,13-dibutyrate from; [129]. The hydrocarbon chains are to
facilitate partitioning into the lipid-rich membrane [138]. These profiles make it an especially attractive lead for
environment. the design of new PKC probes. The binding of bryostatin-1 to
PKC results in PKC activation, autophosphorylation, and
Phorbol esters, such as phorbol 12.13-dibutyrate and translocation to the cell membrane. Bryostatin-1-bound PKC
phorbol 12-myristate 13-acetate (Fig. 2), also bind to the C1 is then down-regulated by ubiquitination and degradation in
domain of cPKC and nPKC, with binding affinities over 2 proteasomes. Short exposures to bryostatin-1 activate PKC,
orders of magnitude greater than those of diacylglycerol. Their while prolonged exposures are followed by down-regulation of
higher potencies are derived from a conformationally rigid PKC. Down-regulation is most significant when PKC is
orientation of hydrophilic pharmacophores. There are reports, exposed to high and/or prolonged high concentrations of
however, that some observed effects that are produced with bryostatin-1 [132, 133, 139, 140]. Studies of structure-activity
phorbol esters and viewed as PKC-mediated may be mediated relations of brayostatin-1 for PKC isoforms [141, 142] indicate
by other signaling molecules, and not by PKCs [78]. that the intact 20-membered macrolactone ring is needed for
Bryostatin-1 [130, 131], a macrocyclic lactone (Fig. 2), is good PKC binding activity, while the A-ring and the B-ring
an antineoplastic agent that potently activates PKC with exocyclic olefin can be deleted from the 20-membered structure
chemical structure unrelated to phorbol esters. It is isolated without inducing much change in PKC-binding affinity. The
from the marine Bugula neritina [131]. Bryostatin-1, acting as C(26) free hydroxyl is essential for a good interaction with
Protein Kinase C Isozymes: Memory Therapeutic Potential Current Drug Targets - CNS & Neurological Disorders, 2005, Vol. 4, No. 5 547

PKC isozymes, while C(3)-hydroxyl with (R)-stereochemistry genetic changes. At high levels, however, V1 expression
is important for a high affinity. depresses cardiac contractile function and disrupts cytoskeletal
structures to form aggregates, in addition to increased
OTHER CONCERNS expression of fetal cardiac genes [156]. PKC and  isoforms
may also mediate isoflurane-induced improvement of
An involvement of PKC activity in the synaptic plasticity functional and metabolic recovery in isolated ischemic rat
and cognition indicates a possibility that agents acting on the hearts [157]. Some of the cardiac effects may involve PKC-
PKC signal cascade(s) may have therapeutic values in mediated desensitization of 1- and 2-adrenergic receptors
cognition and dementia. As briefly mentioned above, PKC [158]. In cases of PKC involvement in ischemic
regulates and modulates a variety of other biological functions, preconditioning, a small degree of PKC activation would be
raising valid concerns of specificities and other side and expected to provide a period of protection of the ischemia-
adverse effects that may be produced by agents that change sensitive neurons from stroke/ischemic injury.
PKC activity and signal cascade(s). In addition to an
involvement of PKCs in the development of morphine 2. Sensation of Thermal and Inflammatory Pain
tolerance [143], the main concerns for the development of PKC
agents in the treatment of dementia and cognition enhancers are PKC is involved in mediating the sensation of thermal and
their potential adverse effects. inflammatory pain, including bradykinin- and anandamide-
induced nociception [159]. In the nociceptive dorsal root
1. Cardiovascular System ganglion neurons, PKC enhances the activity of the slowly
+
inactivating Na channels [160]. There is therefore a possibility
One major concern is the impact of modulation of PKC that PKC activators may facilitate thermal and inflammatory
activity on cardiovascular functions. PKC is involved in nociception, especially in those sensitive patients.
regulating cardiovascular functions [144, 145]. The effects of
PKC activity on the health of cardiovascular system are 3. Hypoxic and Ischemic Injury
complicated by the fact that PKC activity may be both
essential and damaging to a normal cardiovascular function, The second major concern in developing agents acting on
emphasizing the importance of defining isoform specificity in PKC cascade(s) as cognition therapeutics is potential impact of
functional involvements and controlling the levels of PKC such agents on neural responses to stroke and ischemic injury.
isoform-specific activation. Hypoxia/ischemia triggers glutamate release and accumulation
in the extracellular space and activates several PKC isoforms in
PKC activators are known to stimulate cardiomyocyte
neurons. It is well-established that hypoxia/ischemia produces
hypertrophy [146], with an elevated risk for the development 2+
NMDA-, Ca -, and PKC-dependent LTP [161] and increases
of heart failure. The particular isoforms involved include
calcium influx through ionotropic glutamate receptors of the
PKC [147], PKC [148], PKC, and PKC [149].
cortical neurons, via PKC-mediated phosphorylationof NMDA
Activation of PKC and mitochondrial translocation occur at
receptors [162]. In cultured rat hippocampal neurons,
the onset of heart reperfusion after cardiac ischemia and
application of 0.5 mM glutamate for 5 min (neurotoxicity)
mediates apoptosis by facilitating the accumulation and
increases PKC phosphotransferase activity by 3.7-fold and
dephosphorylation of the pro-apoptotic BAD (Bcl-2 associated
protein levels by 5.5-fold in the membrane fraction, without
death promoter), dephosphorylation of Akt, cytochrome c
changing PKC activity in the cytosol [34]. Such PKC
release, PARP cleavage, and DNA laddering [150]. Evidence
activation would be expected to mediate or enhance neuronal
has also been provided that inhibition of PKC is able to
stroke/ischemic injury, rather than to provide preconditioning-
provide the protection of the heart from ischemia and
like neural protection.
reperfusion-induced damage [151, 152]. In addition, enhanced
activation of PKC is correlated with the vascular complications Hypoxia/ischemia is known to induce changes in gene
of diabetes mellitus. LY333531, a specific PKC inhibitor, expression, some of which are actually mediated by PKC.
has been shown to antagonize retinal, cardiac and renal vascular Regulation of EPO mRNA by cellular O2 tension is known to
dysfunction in diabetic rats [153] and patients [154]. These represent a homeostatic mechanism for maintaining tissue
cardiac responses to PKC activation represent adverse effects oxygenation [163]. In RIF (radiation-induced fibrosarcoma)
that have to be addressed in cognition therapy especially when tumor cells, PKC, , and  isoforms are expressed, but only
the same PKC isoforms are the pharmacological targets in the PKC is specifically translocated to the membrane and
cognition therapies. involved in the transcriptional activation of heat shock
However, PKC and  isoforms are not only involved in transcription factor (HSF) and hypoxia-inducible factor- (HIF
heart failure and myocardial hypertrophy and ischemic injury [164]) 1 by hypoxia [165]. A bioflavonoid quercetin (QCT;
to the cardiomyocytes, but also in ischemic preconditioning. 100 M), a well-known inhibitor of hsp gene expression, was
PKC may represent one of the pathways converged to inhibit found to significantly inhibit the transcriptional activation of
glycogen synthase kinase-3 and the mitochondrial HSF and HIF-1, by abrogating completely the PKC
permeability transition pore [155]. PKC is also involved in translocation [165]. QCT, however, is known to inhibit several
maintaining normal cardiomyocyte cytoskeletal integrity. protein kinases, including tyrosine kinase, casein kinase II, and
Cardiac ischemia in mice triggers PKC translocation, which phosphorylase kinase. EPO rRNA accumulation was also
was inhibited by transgenic expression of PKC-specific observed as early as 1 h after hypoxia and lasted for 8 h, and
translocation inhibitor protein fragment (V1) [156]. Thus, the accumulation was blocked by QCT [165]. In addition,
levels of V1 expression confer striking resistance to inhibiting PKC activation, either pharmacologically with
postischemic dysfunction without inducing structural and phorbol 12-myristate 13-acetate (400 nM) or with
548 Current Drug Targets - CNS & Neurological Disorders, 2005, Vol. 4, No. 5 Sun and Alkon

bisindolymaleimide II (50 nM) or genetically by transient tyrosine-phosphorylate PKCs with a modification in catalytic
transfection of a dominant negative PKC, significantly activity, including an activation of PKC [184]. PKC is also
inhibited the transcriptional activation of HSF and HIF-1 by sensitive to oxidative stress, with its activation leading to
hypoxia [165]. The membrane translocation of PKC depends neuronal death [185]. The effects of oxidants on PKC can be
on the activation of phosphoinositol 3-kinase (PI3K). attenuated or eliminated by lipid-soluble antioxidants, such as
Treatment with PI3K inhibitor, wortmannin (50 nM) or vitamin E, which also exhibits a PKC-dependent and
LY294002 (5 M), abrogated not only PKC translocation but antioxidant-independent inhibition of platelet aggregation
also the subsequent transcriptional activation of HSF and HIF- [186]. In rabbit hippocampus, immunocytochemical analysis
1 by hypoxia [165]. reveals a significant aging-related increase in PKC1, 2, and
 immunoreactivity in the cell bodies and dendrites of the
PI3K activation generates D3-phosphorylated
principal cells and interneurons in the hilar region and a
phosphoinositide (D3PPI), which has been shown to act on the
decreased number of PKC and -positive interneurons in the
phosphoinositides-dependent kinase-1 (PDK-1) [166, 167] and
stratum oriens [187], although it has not been evaluated
atypical PKCs [167]. In other systems, Akt was suggested to
whether such increase is related to aging-related oxidant
be a likely target of the D3PPI rather than PDK-1 [168, 169].
accumulation. The enhanced PKC immunoreactivity may be
Hippocampal CA1 pyramidal cells and network are related to reduced protein-protein interactions of PKC with its
particularly sensitive to hypoxic/ischemic damage [170-174]. anchoring protein RACK1.
In the hippocampus, anoxic LTP, a pathophysiological
response to acute hypoxia and ischemia in CA1 pyramidal cells 6. Amyloid and Dementia
was reported to be blocked, not by the specific cAMP-
dependent protein kinase (PKA) inhibitor Rp-cyclic adenosine Although the majority of studies report that PKC activation
3’,5’-monophospnate (25 M), but by PKC inhibitors NPC- results in an increased sAPP production [117, 118] and a
15437 (M), H-7 (20 M), or intracellular application of PKC decreased neurotoxic A production [119, 120], there is a
inhibitor (PKCI) 19-36 (50 M) into the recorded neurons concern that PKC may mediate some A neurotoxicity. In
[175], indicating that induction of anoxic LTP requires primary neuronal cultures, treatment with A1-40 for 48 h
2+
activation of postsynaptic PKC. The involvement of caused a significant increase in the content and activity of Ca -
2+
intracellular release of Ca is indicated by the effectiveness of dependent (, II) and -independent (, ) PKC isoforms,
2+
intracellular Ca chelator BAPTA into the postsynaptic CA1 followed by down-regulation of 72 h, probably resulting from
neurons in preventing the anoxic LTP [175]. over-activation [188]. The A1-40-induced biphasic PKC
activation and down-regulation was exacerbated by hypoxia
PKCs are indicated in mediating ischemic and reperfusion [188], indicating increased vulnerability in AD patients with
damage. A selective PKC peptide inhibitor, for example, has vascular complications.
been found to reduce cellular injury in a rat hippocampal slice
model of cerebral ischemia, when present both during the Some of the pathophysiological actions of PKCs appear to
ischemic episode and for the first 3 hours of reperfusion. The be involved in dementia. Riluzole, the only drug used to
inhibitor decreased infarct size in vivo in rats with transient control amyotrophic lateral sclerosis, directly inhibits PKC
middle cerebral artery occlusion when administered at the [189], in addition to an antiglutamatergic action. PKC may
2+
onset, at 1 hr, or at 6 hr of reperfusion [176]. Global ischemia also mediate Pb -induced inhibition of nicotinic cholinergic
increases PKC mRNA and protein levels in the cortex and modulation of synaptic transmission in the hippocampus
hippocampus and these levels are also increased in the [190]. A recent study reveals that GAP-43 is consistently
compromised peri-infarct region after local cerebral ischemia elevated in a subfield of the hippocampus in AD brain,
[177, 178]. The increased PKC expression in the penumbral associated with elevated growth protein and neural sprouting,
area may be responsible for the delayed neuronal damage [179]. and that increased expression GAP-43 may lead to a mis-
wiring of circuits critical for memory function [191].
4. Stress
7. Neuronal Survival
Stress, on the other hand, is also known to induce high
levels of PKC activity in the prefrontal cortex. The enhanced PKC is both pro-growth and apoptotic. Different PKC
PKC over-activity can disrupt prefrontal cortical regulation of isoforms may exhibit differential effects on growth and
behavior and thought, and impair working memory [180]. apoptosis, depending on cell types and states. Some PKC
PKC mediates modulation of the serotonergic regulation of activators may act as tumor promoters and others, as anti-
GABA transmission by the stress-related neuropeptide tumor agents. nPKCs are involved in tumor promotion [192-
corticotropin-releasing factor [181]. PKC activators may 194]. Only careful evaluation will reveal whether they possess
potentiate stress-mediated psychiatric disorders. Such an tumor-promoting capabilities, since we have very little
interaction between corticotropin-releasingfactor and 5-HT may knowledge of isoform involvement and underlying
play an important role in psychiatric disorders. mechanisms. PKC, for instance, has been found to activate
caspases and be a proapoptotic intermediate in several types of
5. Oxidants cells [195-200], including neurons [201, 2002]. Experimental
brain injury is associated with PKC activation, such as the 
Part of the signal factors in neurodegenerativedisorders and and  isoforms, in regions undergoing neuronal degeneration
aging is oxidants. All the PKC isoforms are sensitive to [203].
oxidants [182, 183]. Hydrogen peroxide, for instance, can
Protein Kinase C Isozymes: Memory Therapeutic Potential Current Drug Targets - CNS & Neurological Disorders, 2005, Vol. 4, No. 5 549

CONCLUDING REMARKS mediated by PKC-independent pathways [32]. All these can


only be answered by further intensive studies. It appears that
There is no doubt that PKCs play an important role in the ligand-binding sites of the PKC isozymes possess some
learning and memory and that agents acting on the PKC signal different profiles so that highly potent isozyme-selective PKC
cascade(s) have promising therapeutic values in anti-dementia activators may be developed in the near future [215, 216] and
therapy. Activation of PKC represents a potential therapeutic that PKC activators have powerful anti-dementia/memory-
strategy in improving memory and mood. PKC may be enhancing and antidepressive activities in animal models (105,
involved in regulation of 5-HT receptor activity [204]. The 217-219). With a full understanding of the intracellular targets
phorbol esters have been reported to be potent PKC of individual PKC isozymes in different cell types (220), the
stimulators. However, their potential as therapeutic drugs is use of pharmacological or molecular therapeutic approaches
limited by their action as tumor promoters [205]. Bryostatin-1 based on modulation of PKC activity/cascades might be
is distinct from the phorbol esters in several properties. For justified clinically in the future.
instance, unlike the phorbol esters, it lacks tumor-promoting
capabilities and actually counteracts tumor promotion induced
by phorbol esters [206]. In clinical trials as an anti-tumor LIST OF ABBREVIATIONS
agent, bryostatin-1 is reasonably well-tolerated, having a
-2 -1
maximum tolerated dose of 25 g m week [207]. Its main A = Amyloid- peptide
dose-limiting toxicity is myalgia. AD = Alzheimer’s disease
Several issues, however, need to be addressed before AMPA = -Amino-3-hydroxy-5-methyl-isoxazole-propio-
bryostatin-1 and its analogues might be developed as nic acid
therapeutic drugs. PKC is involved in a variety of functions
and neurological disorders. It remains to be demonstrated AMPAR = AMPA receptor
whether a relatively non-selective activator of this pathway aPKC = Atypical PKC
would be clinically useful in the treatment of the psychiatric
illness. Defining the underlying isozyme(s) and developing BDNF = Brain-derived neurotropic factor
isozyme-specific agents are thus essential. For instance, the cPKC = Classical PKC
PKC and PKC are implicated in heart failure and myocardial
hypertrophy. Inhibition of PKC and activation of PKC D3PPI = D3-phosphorylated phsophoinositide
appear to protect against myocardial ischemia, but a low level GABA = -Aminobutyric acid
of PKC activation is essential to maintain normal
GAP = Growth-associated protein
cardiomyocyte cytoskeletal integrity [156]. PKC isozyme
activation plays an important pathological role in anoxic LTP GluR = Glutamate receptor
of the glutamatergic synaptic responses in the CA1 pyramidal HIF = Hypoxia-inducible factor
cells of the hippocampus [208], glutamate excitotoxicity in
cultured rat hippocampal neurons [34], tau phosphorylation HSF = Heat shock transcription factor
[209], caspases-mediated apoptosis [210], and amyloid LTD = Long-term depression
neurotoxicity [188]. On the other hand, bryostatin-1 activates
-secretase and thus reduces the production and accumulation LTP = Long-term potentiation
of neurotoxic amyloid [119, 120]. PKC is also capable of mGluR = Metabotropic glutamate receptor
activating the extracellular signal-regulated kinase signaling
pathway through Ras-dependent and –independent cascades, NMDA = N-methyl-D-aspartate
regulating cell proliferation and enhancing cell survival [32], NMDAR = NMDA receptor
opposite to its pro-apoptotic action. Potential toxic side effects
of PKC activation in the periphery might be addressed with co- nPKC = Novel PKC
administration of CNS-impermeant PKC inhibitors/ PKC PDK-1 = Phosphoinositides-dependent kinase-1
cascade inhibitors and/or dosing schedules that only
PKC = Protein kinase C
periodically activate PKC in the brain.
PKCI = PKCI
These concerns, including impact on cardiovascular
function, stroke/ischemia, may be addressed when more studies RACK = Receptor for activated C kinase
are performed to define the PKC isoform involvement in
particular actions and when more selective PKC isoform REFERENCES
activators and inhibitors are developed. Currently, the [1] Inoue, M.; Kishimoto, A.; Takai, Y.; Nishizuka, Y. J. Biol. Chem.
pharmacological advantages possessed by the bryostatins are 1977, 252, 7610.
exciting. However, their selectivity for PKC isoforms is still [2] Tanaka, C.; Nishizuka, Y. Ann. Rev. Neurosci. 1994, 17, 551.
rather limited. In addition, the C1 domain structures [32] also [3] Choi, B.-H.; Chae, H.-D.; Park, T.-J.; Oh, J.; Lim, J.; Kang, S.-S.; Ha,
exist in other molecules, such as PKD kinases [211], chimaerin H.; Kim, K.-T. J. Neurochem. 2004, 90, 442.
[4] Poole, A.W.; Pula, G.; Hers, I.; Crosby, D.; Jones, M.L. Trends
Rac GTPase-activating proteins [212], Ras guanyl nucleotide- Pharmacol. Sci. 2004, 25, 528.
releasing protein [213], Ras and Rap1 exchange factors, [5] Young, K.W.; Garro, M.A.; Challiss, R.A.J.; Nahorski, S.R. J.
MUNC13 scaffolding proteins [78] and diacylglycerol kinases Neurochem. 2004, 89, 1537.
 and  [214]. Some of the effects that have been attributed to [6] Saito, N.; Kikkawa, U.; Nishizuka, Y.; Tanaka, C. J. Neurosci. 1988,
PKC isozymes in response to phorbol esters may therefore be 8, 369.
550 Current Drug Targets - CNS & Neurological Disorders, 2005, Vol. 4, No. 5 Sun and Alkon

[7] Carr, D.B.; Cooper, D.C.; Ulrich, S.L.; Spruston, N.; Surmeier, D.J. [48] Goslin, K.; Schreyer, D.J.; Pate Skene, J.H.; Banker, G. Nature 1988,
J. Neurosci. 2002, 22, 6846. 336, 672.
[8] Carr, D.B.; Day, M.; Cantrell, A.R.; Held, J.; Scheuer, T.; Catterall, [49] Verhaagen, J.; Oestreicher, A.B.; Gispen, W.H.; Margolis, F.L. J.
W.A.; Surmeier, D.J. Neuron 2003, 39, 793. Neurosci. 1989, 9, 683.
[9] Chen, Y.; Cantrell, A.R.; Messing, R.O.; Scheuer, T.; Catterall, W.A. [50] Pascale, A.; Gusev, P.A.; Amadio, M.; Govoni, S.; Alkon, D.L.;
J. Neurosci. 2005, 25, 507. Quattrone, A. Proc. Natl. Acad. Sci. USA 2004, 101, 1217.
[10] Alkon, D.L.; Neary, J.T.; Naito, S.; Coulter, D.; Kubota, M.; [51] Rekart, J.L.; Meiri, K.; Routtenberg, A. Hippocampus 2004, in press.
Rasmussen, H. Biochem. Biophys. Res. Commun. 1986, 134, 1253. [52] Olds, J.L.; Anderson, M.L.; McPhie, D.L.; Alkon, D.L. Science 1989,
[11] Farley, J.; Auerbach, S. Nature 1986, 319, 220. 245, 866.
[12] Kaczmarek, L.K. Trends Neurosci. 1987, 10, 30. [53] McPhie, D.L.; Matzel, L.D.; Olds, J.L.; Lester, D.S.; Kuzirian, A.M.;
[13] LoTurco, J.L.; Coulter, D.A.; Alkon, D.L. Proc. Nat. Acad. Sci. USA Alkon, D.L. J. Neurochem. 1993, 60, 646.
1988, 85, 1672. [54] Akers, R.F.; Routtenberg, A. J. Neurosci. 1987, 7, 3976.
[14] Sheline, Y.I.; Wang, P.W.; Gado, M.H.; Csernansky, J.G.; Vannier, [55] Akers, R.F.; Lovinger, D.M.; Colley, P.A.; Linden, D.J.; Routtenberg,
M.W. Proc. Nat. Acad. Sci. USA 1996, 93, 1908. A. Science 1986, 231, 587.
[15] Bremner, J.D.; Narayan, M.; Anderson, E.R.; Staib, L.H.; Miller, [56] Alkon, D.L.; Rasmussen, H. Science 1988, 239, 998.
H.L.; Charney, D.S. Am. J. Psychiatry 2000, 157, 115-118. [57] Oancea, E.; Meyer, T. Cell 1998, 95, 307.
[16] Cryan, J.F.; Markou, A.; Lucki, I. Trends Pharmacol. Sci. 2002, 23, [58] Topham, M.K.; Bunting, M.; Zimmerman, G.A.; McIntyre, T.M.;
238. Blackshera, P.J.; Prescott, S.M. Nature 1998, 394, 697.
[17] Maviel, T.; Durkin, T.P.; Menzaghi, F.; Bontempi, B. Science 2004, [59] Hoffman, D.A.; Johnston, D. J. Neurosci. 1998, 18, 3528.
305, 96. [60] Manseau, F.; Sossin, W.S.; Castellucci, V.F. J. Neurophsyiol. 1998,
[18] Riedel, G.; Micheau, J.; Lam, A.G.M.; Roloff, E.V.L.; Martin, S.J.; 79, 1210.
Bridge, H.; de Hoz, L.; Poeschel, B.; McCulloch, J.; Morris, R.G.M. [61] Roche, K.W.; O’Brien, R.J.; Mammen, A.L.; Bernhardt, J.; Huganir,
Nature Neurosci. 1999, 2, 898. R.L. Neuron 1996, 16, 1179.
[19] De Zeeuw, C.I.; Hansel, C.; Bian, F.; Koekkoek, S.K.E.; van Alphen, [62] Macek, T.A.; Schaffhauser, H.; Conn, P.J. J. Neurosci. 1998, 18,
A.M.; Linden, D.J.; Oberdick, J. Neuron 1998, 20, 495. 6138.
[20] Toker, A. Front Biosci. 1998, 3, D1134. [63] Suen, P.C.; Wu, K.; Xu, J.L.; Lin, S.Y.; Levine, E.S.; Black, I.B. Brain
[21] Rabah, D.; Grant, S.; Ma C.; Conrad, D.H. J. Immunol. 2001, 167, Res. Mol. Brain Res. 1998, 59, 215.
4910. [64] Grega, D.S.; Werz, M.A.; MacDonald, R.L. Science 1987, 235, 345.
[22] Oster, H.; Eichele, G.; Leitges, M. Mol. Brain Res. 2004, 127, 79. [65] Doerner, D.; Pitler, T.A.; Alger, B.E. J. Neurosci. 1988, 8, 4069.
[23] Hurley, J.H.; Newton, A.C.; Parker, P.J.; Blumberg, P.M.; Nishizuka, [66] Linden, D.J.; Smeyne, M.; Sun, S.C.; Connor, J.A. J. Neurosci. 12,
Y. Protein Sci. 1997, 6, 477. 3601.Louat, T.; Canitrot, Y.; Jousseaume, S.; Baudouin, C.; Canal, P.;
[24] Irie, K.; Oie, K.; Nakahara, A.; Yanai, Y.; Ohigashi, H.; Wender, Laurent, G.; Lautier, D. FEBS Lett. 2004, 574, 121.
P.A.; Fukuda, H.; Konishi, H.; Kikkawa, U.; J. Am. Chem. Soc. 1998, [67] Mao, J.; Wang, X.; Wang R.; Rojas, A.; Shi, Y.; Piao, H.; Jinag, C.
120, 9159. Proc. Natl. Acad. Sci. USA 2004, 101, 1087.
[25] Irie, K.; Nakahara, A.; Nakagawa, Y.; Ohigashi, H.; Shindo, M.; [68] Groc, L.; Heine, M.; Cognet, L.; Brickley, K.; Stepenson, F.A.;
Fukuda, H.; Konishi, H.; Kikkawa, K.; Saito, N. Pharmacol. Ther. Lounis, B.; Choquet, D. Nature Neurosci. 2004, 7, 695.
2002, 93, 271. [69] Melyan, Z.; Wheal, H.V.; Lancaster, B. Neuron 2002, 34, 107.
[26] Stahelin, R.V.; Digman, M.A.; Medkova, M.; Anathannarayanan, B.; [70] Gereau IV, R.W.; Heinemann, S.F. Neuron 1998, 20, 143.
Rafter, J.D.; Melowic, H.R.; Cho, W. J. Biol. Chem. 2004, 279, [71] Jovanovic, J.N.; Thomas, P.; Kittler, J.T.; Smart, T.G.; Moss, S.J. J.
29501. Neurosci. 2004, 24, 522.
[27] Slater. S.J.; Malinowski, S.A.; Stubbs, C.D. Biochemistry 2004, 43, [72] Dobransky, T.; Doherty-Kirby, A.; Kim, A.; Brewer, D.; Lajoie, G.;
7601. Rylett, R.J. J. Biol. Chem. 2004, 279, 52059.
[28] Nishizuka, Y. Nature 1988, 334, 661. [73] Okada, M.; Zhu, G.; Yoshida, S.; Hirose, S.; Kaneko, S.
[29] Gszchwendt, M.; Leipbersperger, H.; Kittstein, W.; Marks, F. FEBS Neuropharmacology 2004, 47, 485.
Lett. 1992, 307, 151. [74] Stevens, C.F.; Sullivan, J.M. Neuron 1998, 21, 885.
[30] Ways, D.K.; Cook, P.P.; Webster, C.; Parker, P.J. J. Biol. Chem. [75] Malenka, R.C.; Madison, D.V.; Nicoll, R.A. Nature 1986, 321, 175.
1992, 267, 4799. [76] Nicolls, D.G. Prog. Brain Res. 1998, 116, 15.
[31] Nishizuka, Y. Science 1992, 258, 607. [77] Roberson, E.D.; English, J.D.; Adams, J.P.; Selcher, J.C.; Kondratick,
[32] Yang, C.-F.; Kazanietz, M.G. Trends Pharmacol. Sci. 2003, 24, 602. C.; Sweatt, J.D. J. Neurosci. 1999, 19, 4337.
[33] Louat, T.; Canitrot, Y.; Jousseaurne, S.; Baudouin, C.; Canal, P.; [78] Rhee, J.-S.; Betz, A.; Pyott, S.; Reim, K.; Varoqueaux, F.; Augustin,
Laurent, G.; Lautier, D. FEBS Lett. 2004, 574, 121. I.; Hesse, D.; Südhof, T.C.; Takahashi, M.; Rosenmund, C.; Brose, N.
[34] Hasham, M.I.; Pelech, S.L.; Krieger, C. Neurosci. Lett. 1997, 228, Cell 2002, 108, 121.
115. [79] Chung, H.J.; Steinberg, J.P.; Huganir, R.L.; Linden, D.J. Science
[35] Maurice, N.; Tkatch, T.; Meisler, M.; Sprunger, L.K.; Surmeier, D.J. 2003, 300, 1751.
J. Neurosci. 2001, 21, 2268. [80] Bank, B.; DeWeer, A.; Kuzirian, A.M.; Rasmussen, H.; Alkon, D.L.
[36] Sato, T.; Tanaka, K.; Teramoto, T.; Ohnishi, Y.; Hirate, K.; Irifune, Proc. Nat. Acad. Sci. USA 1988, 85, 1988.
M.; Nishikawa, T. Peptides 2004, 25, 1139. [81] Alkon, D.L.; Nelson, T.J.; Zhao, W.Q.; Cavallaro, S. Trends
[37] Andoh, T.; Itoh, H.; Higashi, T.; Saito, Y.; Ishiwa, D.; Kamiya, Y.; Neurosci. 1998, 21, 529.
Yamada, Y. Brain Res. 2004, 1013, 125. [82] Pilpel, Y.; Segal, M. Eur. J. Neurosci. 2004, 19, 3151.
[38] Hama, H.; Hara, C.; Yamaguchi, K.; Miyawaki, A. Neuron 2004, 41, [83] Malinow, R.; Madison, D.V.; Tsien, R.W. Nature 1988, 335, 820.
405. [84] Malenka, R.C.; Kauer, J.A.; Perkel, D.J.; Mauk, M.D.; Kelly, P.T.;
[39] Song, H.; Stevens, C.F.; Gage, F.H. Nature 2002, 417, 39. Nicoll, R.A.; Waxham, M.N. Nature 1989, 340, 554.
[40] Song, H.; Stevens, C.F.; Gage, F.H. Nature Neurosci. 2002b, 5, 438. [85] Malinow, R.; Schulman, H.; Tsien, R.W. Science 1989, 245, 862.
[41] Ron, D.; Kazanietz, M.G. FASEB. J. 1999, 13, 1658. [86] Malenka, R.C.; Kauer, J.A.; Perkel, D.J.; Nicoll, R.A. Trends
[42] Newton, A.C. Chem. Rev. 2001, 101, 2353. Neurosci. 1989, 12, 444.
[43] Melchers, B.P.C.; DeGraan, P.N.E.; Schrama, L.H.; Wadman, W.J.; [87] Zalutsky, R.A.; Nicoll, R.A. Science 1990, 248, 1619.
Lopez da Silva, F.H.; Gispen, W.-H. In A Long-Term Potentiation: [88] Weeber, E.J.; Atkins, C.M.; Selcher, J.C.; Varga, A.W.; Mirnikjoo,
From Biophysics to Behavior, Landfield, P.W.; Deadwyler, S.A. B.; Paylor, R.; Leitges, M.; Sweatt, J.D. J. Neurosci. 2000, 20, 5906.
Eds.; Alan R. Liss: New York, 1988; pp. 307-327. [89] Ramakers, G.M.; Heinen, K.; Gispen, W.H.; de Graan, P.N. J. Biol.
[44] Routtenberg, A.; Cantallops, I.; Zaffuto, S.; Serrano, P.; Namgung, U. Chem. 2000, 275, 28682.
Proc. Natl. Acad. Sci. USA 2000, 97, 7657. [90] van Dam, E.J.; Ruiter, B.; Kamal, A.; Ramakers, G.M.; Gispen,
[45] Skene, P.J.H.; Jacobson, R.D.; Snipes, G.J.; McGuire, C.B.; Norden, W.H.; de Graan, P.N. Neurosci. Lett. 2002, 320, 129.
J.J.; Freeman, J.A. Science 1986, 233, 783. [91] Li, M.-X.; Jia, M.; Yang, L.-X.; Jiang, H.; Lanuza, M.A.; Gonzalez,
[46] Jacobson, R.D.; Virag, I.; Pate Skene, J.H. J. Neurosci. 1986, 6, 1843. C.M.; Nelson, P.G. J. Neurosci. 2004, 24, 3762.
[47] Benowitz, L.I.; Apostolides, P.J.; Perronne-Bizzozero, N.; [92] Schreurs, B.G.; Oh, M.M.; Alkon, D.L. J. Neurophysiol. 1996, 75,
Finklestein, S.P.; Zwiers, H. J. Neurosci. 1988, 8, 339. 1051.
Protein Kinase C Isozymes: Memory Therapeutic Potential Current Drug Targets - CNS & Neurological Disorders, 2005, Vol. 4, No. 5 551

[93] Schreurs, B.G.; Tomsic, D.; Gusev, P.A.; Alkon, D.L. J. [132] Isakov, N.; Galron, D.; Mustelin, T.; Pettit, G.R.; Altman. A. J.
Neurophysiol. 1997, 77, 86. Immunol. 1993, 150, 1195.
[94] Van der Zee, E.A.; Kronforst-Collins, M.A.; Maizels, E.T.; [133] Galron, D.; Tamir, A.; Altman, A.; Isakov, N. Cell. Immunol. 1994,
Hunzicker,-Dunn, M.; Disterhoft, J.F. Hippocampus 1997, 7, 271. 158, 195.
[95] Olds, J.L.; Golski, S.; McPhie, D.L.; Olton, D.; Mishkin, M.; Alkon, [134] Zhang, X.H.; Zhang, R.W.; Zhao, H.; Cai, H.Y.; Gush, K.A.; Kerr,
D.L. J. Neurosci. 1990, 10, 3707. R.G.; Pettit, G.R.; Kraft, A.S. Cancer Res. 1996, 56, 802.
[96] Paylor, R.; Rudy, J.W.; Wehner, J.M. Behav. Brain Res. 1991, 45, [135] Yoo, J.; Nichols, A.; Song, J.C.; Mun, E.C.; Matthews, J.B. Surgery
189. 2001, 130, 310.
[97] Paylor, R.; Morrison, S.K.; Rudy, J.W.; Waltrip, L.T.; Wehner, J.M. [136] Jalava, A.; Lintunen, M.; Heikkila, J. Biochem. Biophys. Res.
Behav. Brain Res. 1992, 52, 49. Commun. 1993, 191, 472.
[98] Colombo, P.J.; Wetsel, W.C.; Gallagher, M. Proc. Natl. Acad. Sci [137] Stanwell, C.; Gescher, A.; Bradshaw, T.D.; Pettit, G.R. Int. J. Cancer
USA 1997, 94, 14195. 1994, 56, 585.
[99] Vázquez, A.; de Ortiz, P. Toxicol. Appl. Phamacol. 2004, 200, 27. [138] Szallasi, Z.; Denning, M.F.; Smith, C.B.; Diugosz, A.A.; Yuspa, S.H.;
[100] Olds, J.L.; Bhalla, U.S.; McPhie, D.L.; Lester, D.S.; Bower, J.M.; Pettit, G.R.; Blumberg, P.M. Mol. Pharmacol. 1994, 46, 840.
Alkon, D.L. Behav. Brain Res. 1994, 61, 37. [139] Lee, H.W.; Smith, L.; Pettit, G.R.; Bingham, S.J. Am. J. Physiol. 1996,
[101] Yasoshima, Y.; Yamamoto, T. NeuroReport 1997, 8, 1363. 271, C304.
[102] Ahi, J.; Radulovic, J.; Spiess, J. Behav. Brain Res. 2004, 149, 17. [140] Song, J.C.; Hanson, C.M.; Tsai, V.; Farokhad, O.C.; Lotz, M.;
[103] Levenson, J.M.; O’Riordan, K.J.; Brown, K.D.; Trinh, M.A.; Matthews, J.B. Am. J. Physiol. Cell Physiol. 2001, 281, C649.
Molfese, D.L.; Sweatt, J.D. J. Biol. Chem. 2004, 279, 40545. [141] Wender, P.A.; Hinkle, K.W.; Koehler, M.F.T.; Lippa, B. Med. Res.
[104] Jerusalinsky, D.; Quillfeldt, J.A.; Walz, R.; Da Silva, R.C.; Medina, Rev. 1999, 19, 388.
J.H.; Izquierdo, I. Behav. Neural Biol. 1994, 61, 107. [142] Wender, P.A.; Baryza, J.L.; Bennett, C.E.; Bi, F.C.; Brenner, S.E.;
[105] Sun, M.-K.; Alkon, D.L. Eur. J. Pharmacol. 2005, submitted. Clarke, M.O.; Horan, J.C.; Kan, C.; Lacote, E.; Lippa, B.; Nell, P.G.;
[106] Cole, G.; Dobkins, K.R.; Hansen, L.A.; Terry, R.D.; Saitoh, T. Brain Turner, T.M. J. Am. Chem. Soc. 2002, 124, 13648.
Res. 1988, 452, 165. [143] Bailey, C.P.; Kelly, E.; Henderson, G. Mol. Pharmacol. 2004, 66,
[107] Favit, A.; Grimaldi, M.; Nelson, T.J.; Alkon, D.L. Proc. Natl. Acad. 1592.
Sci. USA 1998, 95, 5562. [144] Domanska-Janik, K. Acta Neurobiol. Exp (Warsz) 1996, 56, 579.
[108] Pandey, G.N.; Dwivedi, Y.; Ren, X.; Rizavi, H.S.; Roberts, R.C.; [145] Krupinski, J.; Slevin, M.A.; Kumar, P.; Gaffney, J.; Kaluza, J. Acta
Conley, R.R.; Tamminga, C. J. Psychiatr. Res. 2003, 37, 421. Neurobiol. Exp. (Warsz) 1998, 58, 13.
[109] Matsushima, H.; Shimohama, S.; Chachin, M.; Taniguchi, T.; Kimura, [146] Vega, R.B.; Harrison, B.C.; Meadows, E.; Roberts, C.R.; Papst, P.J.;
J. J. Neurochem. 1996, 67, 317. Olson, E.N.; McKinsey, T.A. Mol. Cell. Biol. 2004, 24, 8374.
[110] Kurimatani, T.; Fastbom, J.; Bonkale, W.L.; Bogdanovic, N.; [147] Braz, J.C.; Bueno, O.F.; De Windt, J.; Molkentin, J.D. J. Cell Biol.
Wingblad, B.; Ohm, T.G.; Cowburn, R.F. Brain Res. 1998, 796, 209. 2002, 156, 905.
[111] Chachin, M.; Shimohama, S.; Kunugi, Y.U.; Taniguchi, T. Jpn. J. [148] Bowman, J.C.; Steinberg, S.F.; Jiang, T.; Geenen, D.L.; Fishman, G.I.;
Pharmcaol. 1996, 71, 175. Buttrick, P.M. J. Clin. Investig. 1997, 100, 2189.
[112] Wang, H.-Y.; Pisano, M.R.; Friedman, E. Neurobiol. Aging 1994, 15, [149] Mochly-Rosen, D.; Wu, G.; Hahn, H.; Osinska, H.; Liron, T.; Lorenz,
293. J.N.; Yatani, A.; Robbins, J.; Dorn, Jr. G.W. Circ. Res. 2000, 86,
[113] Lanius, R.A.; Wagey, R.; Sahl, B.; Beattie, B.L.; Feldman, H.; Pelech, 1173.
S.L.; Krieger, C. Brain Res. 1997, 764, 75. [150] Murriel, C.L.; Churchill, E.; Inagaki, K.; Szweda, L.I.; Mochly-
[114] Battaini, F.; Pascale, A.; Paoletti, R.; Govoni, S. Trends Neurosci. Rosen, D. J. Biol. Chem. 2004, 279, 47985.
nd
1997, 20, 410. [151] Inagaki, K.; Hahn, H.S.; Dorn, G.W., 2 ; Mochly-Rosen, D.
[115] Battaini, F.; Pascale, A.; Lucchi, L.; Pasinetti, G.M.; Govoni, S. Exp. Circulation 2003, 108, 869.
Neurol. 1999, 159, 559. [152] Inagaki, K.; Chen, L.; Ikeno, F.; Lee, F.H.; Imahashi, K.; Bouley,
[116] Takashima, A.; Yokota, T.; Maeda, Y.; Itoh, S. Peptides 1991, 12, D.M.; Rezaee, M.; Yock, P.G.; Murphy, E.; Mochly-Rosen, D.
699. Circulation 2003b, 108, 2304.
[117] Kinouchi, T.; Sorimachi, H.; Maruyama, K.; Mizuno, K.; Ohno, S.; [153] Ishii, H.; Jirousek, M.R.; Koya, D.; Takagi, C.; Xia, P.; Clermont, A.;
Ishiura, S.; Suzuki, K. FEBS Lett. 1995, 364, 203. Bursell, S.-E.; Kern, T.S.; Ballas, L.M.; Heath, W.F.; Stramm, L.E.;
[118] Yeon, S.W.; Jung, M.W.; Ha, M.J.; Kim, S.U.; Huh, K.; Savage, M.J.; Peener, E.P.; King, G.L. Science 1996, 272, 728.
Masliah, E.; Mook-Jung, I. Biochem. Biophys. Res. Commun. 2001, [154] Aiello, L.P.; Bursell, S.; Devries, T.; Alatorre, C.; King, G.; Ways, K.
280, 782. Diabetes 1999, 48, A19.
[119] Zhu, G.; Wang, D.; Lin, Y.H.; McMahon, T.; Koo, E.H.; Messing, [155] Juhaszova, M.; Zorov, D.B.; Kim, S.-H.; Pepe, S.; Fu, Q.; Fishbein,
R.O. Biochem. Biophys. Res. Commun. 2001, 285, 997. K.W.; Ziman, B.D.; Wang, S.; Ytrehus, K.; Antos, C.L.; Olson, E.N.;
[120] Etcheberrigaray, R.; Tan, M.; Dewachter, I.; Kuipéri, C.; Van der Sollott, S.J. J. Clin. Invest. 2004, 113, 1535.
Auwera, I.; Wera, S.; Qiao, L.; Bank, B.; Nelson, T.J.; Kozikowski, [156] Hahn, H.S.; Yussman, M.G.; Toyokawa, T.; Mareez, Y.; Barrett,
A.P.; Van Leuven, F.; Alkon, D.L. Proc. Natl. Acad. Sci. USA 2004, T.J.; Hilty, K.C.; Osinska, H.; Robbins, J.; Dorn II, G.W. Circ. Res.
101, 11141. 2002, 91, 741.
[121] Lorenzo, P.S.; Bögi, K.; Ács, P.; Pettit, G.R.; Blumberg, P.M. J. Biol. [157] Xu, P.; Wang, J.; Kodavatiganti, R.; Zeng, Y.; Kass, I.S. Anesth.
Chem. 1997, 272, 33338. Analg. 2004, 99, 993.
[122] Lee, W.; Boo, J.H.; Jung, M.W.; Park, S.D.; Kim, Y.H.; Kim, S.U.; [158] Guimond, J.; Mamarbachi, A.M.; Allen, B.G.; Rindt, H.; Hébert, T.E.
Mook-Jung, I. Mol. Cell. Neurosci. 2004, 26, 222. Cell. Signal. 2005, 17, 49.
[123] Chauhan, A.; Chauhan, V.P.; Brockerhoff, H.; Wisniewski, H.M. [159] Premkumar, L.S.; Ahern, G.P. Nature 2000, 408, 985.
Life Sci. 1991, 49, 1555. [160] Gold, M.S.; Levine, J.D.; Correa, A.M. J. Neurosci. 1998, 18, 10345.
[124] Govoni, S.; Berganaschi, S.; Racchi, M.; Battaini, F.; Binetti, G.; [161] Hammond, C.; Crepel, V.; Gozlan, H.; Ben-Ari, Y. Trends Neuroci.
Bianchetti, A.; Trabucchi, M. Neurology 1993, 43, 2581. 1994, 17, 497.
[125] Pakaski, M.; Balaspiri, L.; Checler, F.; Kasa, P. Neurochem. Int. [162] Bickler, P.E.; Fahlman, C.S.; Ferriero, D.M. J. Neurochem. 2004, 88,
2002, 41, 409. 878.
[126] Lavoiee, L.; Band, C.J.; Kong, M.; Bergeron, J.J.M.; Posner, B.I. J. [163] Wang, G.L.; Semenza, G.L. Blood 1993, 82, 3610.
Biol. Chem. 1999, 274, 28279. [164] Wang, G.L.; Semenza, G.L. J. Biol. Chem. 1995, 270 1230.
[127] Fang, X.; Yu, S.; Tanyi, J.L.; Lu, Y.; Woodgett, J.R.; Mills, G.B. Mol. [165] Baek, S.-H.; Lee, U.-Y.; Park, E.-M.; Han, M.-Y.; Lee, Y.-S.; Park,
Cell. Biol. 2002, 22, 2099. Y.-M. J. Cell. Physiol. 2001, 188, 223.
[128] Cho, J.H.; Johnson, G.V. J. Biol. Chem. 2004, 279, 54716. [166] Stephens, L.; Anderson, K.; Stokoe, D.; Erdjument-Bromage, H.;
[129] Kang, J.-H.; Kim, S.Y.; Lee, J.; Marquez, V.E.; Lewin, N.E.; Pearce, Painter, G.F.; Molmes, A.B.; Gaffney, P.R.; Reese, C.B.; McCormick,
L.V.; Blumberg, P.M. J. Med. Chem. 2004, 47, 4000. F.; Tempst, P.; Coadwell, J.; Hawkins, P.T. Science 1998, 279, 710.
[130] Trenn, G.; Pettit, G.R.; Takayama, N.; Hu-Li, J.; Sitkovsky, M.V. J. [167] Chan, T.O.; Rittenhouse, S.E.; Tsichlis, P.N. Annu. Rev. Biochem.
Immunol. 1988, 140, 433. 1999, 68, 965.
[131] Pettit, G. R. Fortschr. Chem. Org. Naturst. 1991, 57, 153. [168] Marshall, C.J. Cell 1995, 80, 179.
552 Current Drug Targets - CNS & Neurological Disorders, 2005, Vol. 4, No. 5 Sun and Alkon

[169] Alessi, D.R.; Deak, M.; Cassamayor,A.; Caudwell, F.B.; Morrice, N.; [195] Li, L.; Lorenzo, P.S.; Bogi, K.; Blumberg, P.M.; Yuspa, S.H. Mol. Cell
Norman, D.G.; Gaffney, P.; Reese, C.B.; MacDougall, C.N.; Biol. 1999, 19, 8547.
Haribison, D.; Ashworth, A.; Bownes, M. Curr. Biol. 1997, 7, 776. [196] Cross, T.; Griffiths, G.; Deacon, E.; Sallis, R.; Gough, M.; Watters,
[170] Sun, M.-K. IUBMB Life 1999, 48, 373. D.; Lord, J.M. Oncogene 2000, 19, 2331.
[171] Sun, M.-K.; Alkon, D.L. LifeXY 2001, 1, 1075. [197] Dempsey, E.C.; Newton, A.C.; Mochly-Rosen, D.; Fields, A.P.;
[172] Sun, M.-K.; Xu, H.; Alkon, D.L. J. Pharmacol. Exp. Ther. 2002, 300, Reyland, M.E.; Insel, P.A.; Messing, R.O. Am. J. Phsyiol. 2000, 279,
408. L429.
[173] Sun, M.-K.; Alkon, D.L. In Focus on Alzheimer’s Disease Research, [198] Basu, A.; Woolard, M.D.; Johnson, C.L. Cell Death Differ. 2001, 8,
Welsh, E.M. Ed.; Nova Biomedical Books: New York, 2003; pp. 1- 899.
49. [199] Matassa, A.A.; Carpenter, L.; Biden, T.J.; Humphries, M.J.; Reyland,
[174] Sun, M.-K.; Alkon, D.L. J. Alzheim. Dis. 2004, 6, 355. M.E. J. Biol. Chem. 2001, 276, 29719.
[175] Hsu, K.-S.; Huang, C.-C. Br. J. Pharmacol. 1997, 122, 671. [200] Durrant, D.; Liu, J.; Yang, H.-S.; Lee, R.M. Biochem. Biophys. Res.
[176] Bright, R.; Raval, A.P.; Dembner, J.M.; Pérez-Pinzón, M.A.; Commun. 2004, 321, 905.
Steinberg, G.K.; Yenari, M.A.; Mochly-Rosen, D. J. Neurosci. 2004, [201] Pongracz, J.; Webb, P.; Wang, K.; Deacon, E.; Lunn, O.J.; Lord, J.M.
24, 6880. J. Biol. Chem. 1999, 274, 37329.
[177] Miettinen, S.; Roivainen, R.; Keinanen, R.; Hokfelt, T.; Koistinaho, [202] Naher, P. J. Neurosci. 2001, 21, 2929.
J. J. Neurosci. 1996, 16, 6236. [203] Padmaperuma, B.; Mark, R.; Dhillon, H.S.; Mattson, M.P.; Prasad,
[178] Koponen, S.; Goldsteins, G.; Keinanen, R.; Koistinaho, J. J. Cereb. M.R. Brain Res. 1996, 714, 19.
Blood Flow Metab. 2000, 20, 93. [204] Rahimian, R.; Hrdina, P.D. Can. J. Physiol. Pharmacol. 1995, 73,
[179] Phan, T.G.; Wright, P.M.; Markus, R.; Howells, D.W.; Davis, S.M.; 1686.
Donnan, G.A. Clin. Exp. Pharmacol. Phsyiol. 2002, 29, 1. [205] Blumberg, P.M. Cancer Res. 1988, 48, 1.
[180] Birnbaum, S.G.; Yuan, P.X.; Wang, M.; Vijayraghavan, S.; Bloom, [206] Hennings, H.; Blumberg, P.M.; Pettit, G.R.; Herald, C.L.; Shores, R.;
A.K.; Davis, D.J.; Gobeske, K.T.; Sweatt, J.D.; Manji, H.K.; Arnsten, Yuspa, S.H. Carcinogenesis 1987, 8, 1343.
A.F.T. Science 2004, 306, 882. [207] Clamp, A.R.; Blackhall, F.H.; Vasey, P.; Soukop, M.; Coleman, R.;
[181] Tan, H.; Zhong, P.; Yan Z. J. Neurosci. 2004, 24, 5000. Halbert, G.; Robson, L.; Jayson, G.C. Br. J. Cancer 2003, 89, 1152.
[182] Armstead, W.A. Stroke 1999, 30, 153. [208] Hsu, K.-S.; Huang, C.-C. NeuroReport 1998, 9, 3525.
[183] Battaini, F. Pharmacol. Res. 2001, 44, 353. [209] Ekinci, F.J.; Shea, T.B. Int. J. Dev. Neurosci. 1997, 15, 867.
[184] Konish, H.; Yamauchi, E.; Taniguchi, H.; Yamamoto, T.; Matsusaki, [210] Yashida, K.; Wang, H.G.; Miki, Y.; Kufe, D. EMBO J. 2003, 22,
H.; Takemura, Y.; Ohmae, K.; Kikkawa, U.; Nishizuka, Y. Proc. 1431.
Natl. Acad. Sci. USA 2001, 98, 6587. [211] Van Lint, J.; Rykx, A.; Maeda, Y.; Vantus, T.; Sturany, S.; Malhotra,
[185] Jung, Y.-S.; Ryu, B.R.; Lee, B.K.; Mook-Jung, I.; Kim, S.U.; Lee, V.; Vandenheede, J.R.; Seufferlein, T. Trends Cell. Biol. 2002, 12,
S.H.; Baik, E.J.; Moon, C.-H. Biochem. Biophys. Res. Commun. 2004, 193.
320, 789. [212] Caloca, M.J.; Wang, H.; Kazanietz, M.G. Biochem. J. 2003, 375,
[186] Freedman, J.E.; Keaney, Jr. J.F. J. Nutrit. 2001, 131, 374S. 313.
[187] Van der Zee, E.A.; Palm, I.F.; O’Connor, M.; Maizels, E.T.; [213] Ebinu, J.O.; Bottorff, D.A.; Chan, E.Y.; Stang, S.L.; Dunn, R.J.; Stone,
Hunzicker-Dunn, M.; Disterholft, J.F. Hippocampus 2004, 14, 849. J.C. Science 1998, 280, 1082.
[188] Kuperstein, F.; Reiss, N.; Koudinova, N.; Yavin, E. J. Neurochem. [214] Shindo, M.; Irie, K.; Ohigashi, H.; Kuriyama, M.; Saito, N. Biochem.
2001, 76, 758. Biophysi. Res. Commun. 2001, 289, 451.
[189] Noh, K.M.; Hwang, J.Y.; Shin, H.C.; Koh, J.Y. Neurobiol. Dis. 2000, [215] Nakagawa, Y.; Irie, K.; Yamanaka, N.; Ohigashi, H.; Tsuda, K.-I.
7, 375. Bioorg. Med. Chem. Lett. 2003, 13, 3015.
[190] Braga, M.F.M.; Pereira, E.F.R.; Mike, A.; Albuquerque, E.X. J. [216] Sridhar, J.; Wei, Z.-L.; Nowak, I.; Lewin, N.E.; Ayres, J.A.; Pearce,
Pharmacol. Exp.Ther. 2004, 311, 700. L.V.; Blumberg, P.M.; Kozikowski, A.P. J. Med. Chem. 2003, 46,
[191] Rekart, J.L.; Quinn, B.; Mesulam, M.M.; Routtenberg, A. 4196.
Neuroscience 2004, 126, 579. [217] Sun, M.-K.; Alkon, D.L. Curr. Drug Targets: CNS Neurol. Disord.
[192] Lu, Z.; Hornia, A.; Jiang, Y.-W.; Zang, Q.; Ohno, S.; Foster, D.A. 2002, 1, 575.
Mol. Cell. Biol. 1997, 17, 3418. [218] Sun, M.-K.; Alkon, D.L. J. Neurosci. Meth. 2003, 126, 35-409.
[193] Reddig, P.J.; Dreckschimdt, N.E.; Ahrens, H.; Simsiman, R.; Tseng, [219] Sun, M.-K.; Alkon, D.L. Neuroscience 2004, 129, 129.
C.-P.; Zou, J.; Oberley, T.D.; Verma, A.K. Cancer Res. 1999, 59, [220] Hopf, F.W.; Mailliard, W.S.; Gonzalez, G.F.; Diamond, I.; Bonci, A.
5710. J. Neurosci. 2005, 25, 985.
[194] Reddig, P.J.; Dreckschmidt, N.E.; Bourguignon, S.E.; Oberley, T.D.;
Verma, A.K. Cancer Res. 2000,60, 595.

Вам также может понравиться