Skip to main content
Advertisement

Main menu

  • Home
  • Articles
    • Current Issue
    • Fast Forward
    • Latest Articles
    • Archive
  • Information
    • Instructions to Authors
    • Submit a Manuscript
    • FAQs
    • For Subscribers
    • Terms & Conditions of Use
    • Permissions
  • Editorial Board
  • Alerts
    • Alerts
    • RSS Feeds
  • Virtual Issues
  • Feedback
  • Submit
  • Other Publications
    • Drug Metabolism and Disposition
    • Journal of Pharmacology and Experimental Therapeutics
    • Molecular Pharmacology
    • Pharmacological Reviews
    • Pharmacology Research & Perspectives
    • ASPET

User menu

  • My alerts
  • Log in
  • Log out
  • My Cart

Search

  • Advanced search
Pharmacological Reviews
  • Other Publications
    • Drug Metabolism and Disposition
    • Journal of Pharmacology and Experimental Therapeutics
    • Molecular Pharmacology
    • Pharmacological Reviews
    • Pharmacology Research & Perspectives
    • ASPET
  • My alerts
  • Log in
  • Log out
  • My Cart
Pharmacological Reviews

Advanced Search

  • Home
  • Articles
    • Current Issue
    • Fast Forward
    • Latest Articles
    • Archive
  • Information
    • Instructions to Authors
    • Submit a Manuscript
    • FAQs
    • For Subscribers
    • Terms & Conditions of Use
    • Permissions
  • Editorial Board
  • Alerts
    • Alerts
    • RSS Feeds
  • Virtual Issues
  • Feedback
  • Submit
  • Visit Pharm Rev on Facebook
  • Follow Pharm Rev on Twitter
  • Follow ASPET on LinkedIn
OtherReview Article

Inhibitors of Brain Phospholipase A2 Activity: Their Neuropharmacological Effects and Therapeutic Importance for the Treatment of Neurologic Disorders

Akhlaq A. Farooqui, Wei-Yi Ong and Lloyd A. Horrocks
Pharmacological Reviews September 2006, 58 (3) 591-620; DOI: https://doi.org/10.1124/pr.58.3.7
Akhlaq A. Farooqui
Department of Molecular and Cellular Biochemistry, The Ohio State University, Columbus, Ohio (A.A.F., L.A.H.); and Department of Anatomy, National University of Singapore, Singapore, Singapore (W.-Y.O.)
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
Wei-Yi Ong
Department of Molecular and Cellular Biochemistry, The Ohio State University, Columbus, Ohio (A.A.F., L.A.H.); and Department of Anatomy, National University of Singapore, Singapore, Singapore (W.-Y.O.)
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
Lloyd A. Horrocks
Department of Molecular and Cellular Biochemistry, The Ohio State University, Columbus, Ohio (A.A.F., L.A.H.); and Department of Anatomy, National University of Singapore, Singapore, Singapore (W.-Y.O.)
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
  • Article
  • Figures & Data
  • Info & Metrics
  • eLetters
  • PDF
Loading

Abstract

The phospholipase A2 family includes secretory phospholipase A2, cytosolic phospholipase A2, plasmalogen-selective phospholipase A2, and calcium-independent phospholipase A2. It is generally thought that the release of arachidonic acid by cytosolic phospholipase A2 is the rate-limiting step in the generation of eicosanoids and platelet activating factor. These lipid mediators play critical roles in the initiation and modulation of inflammation and oxidative stress. Neurological disorders, such as ischemia, spinal cord injury, Alzheimer's disease, multiple sclerosis, prion diseases, and epilepsy are characterized by inflammatory reactions, oxidative stress, altered phospholipid metabolism, accumulation of lipid peroxides, and increased phospholipase A2 activity. Increased activities of phospholipases A2 and generation of lipid mediators may be involved in oxidative stress and neuroinflammation associated with the above neurological disorders. Several phospholipase A2 inhibitors have been recently discovered and used for the treatment of ischemia and other neurological diseases in cell culture and animal models. At this time very little is known about in vivo neurochemical effects, mechanism of action, or toxicity of phospholipase A2 inhibitors in human or animal models of neurological disorders. In kainic acid-mediated neurotoxicity, the activities of phospholipase A2 isoforms and their immunoreactivities are markedly increased and phospholipase A2 inhibitors, quinacrine and chloroquine, arachidonyl trifluoromethyl ketone, bromoenol lactone, cytidine 5-diphosphoamines, and vitamin E, not only inhibit phospholipase A2 activity and immunoreactivity but also prevent neurodegeneration, suggesting that phospholipase A2 is involved in the neurodegenerative process. This also suggests that phospholipase A2 inhibitors can be used as neuroprotectants and anti-inflammatory agents against neurodegenerative processes in neurodegenerative diseases.

I. Introduction

Phospholipases A2 (PLA21; EC 3.1.1.4) form an expanding superfamily of esterases that specifically cleave the acyl ester bond at the sn-2 position of membrane phospholipids to produce a free fatty acid and lysophospholipid (Farooqui et al., 2000b). Because a large proportion of cellular arachidonic acid is found esterified at the sn-2 position of membrane phospholipids, arachidonic acid and lysophospholipid are the major products of the PLA2-catalyzed reaction. Under normal conditions, some arachidonic acid is converted to inflammatory mediators, prostaglandins, leukotrienes, and thromboxanes, whereas a majority of arachidonic acid is reincorporated into the brain phospholipids (Rapoport, 1999; Leslie, 2004). Arachidonic acid not only acts via conversion to inflammatory metabolites, but can also directly modulate neuronal function by various mechanisms, such as altering membrane fluidity and polarization state, activating protein kinase C, and regulating gene transcription (Katsuki and Okuda, 1995; Farooqui et al., 1997b). Another product of PLA2 catalyzed reactions, 1-alkyl-2-lysophospholipid, is the immediate precursor of platelet-activating factor (PAF), another potent inflammatory mediator (Farooqui and Horrocks, 2004b).

Lysophospholipids may also change membrane fluidity and permeability. These metabolites are also involved in phospholipid remodeling and membrane perturbation. Accumulation of lysophospholipids is controlled by either reacylation to native phospholipids (Farooqui et al., 2000b) or by metabolism to water-soluble glycerophosphodiesters such as glycerophosphocholine by lysophospholipases (Farooqui et al., 1985). Thus, tight regulation of PLA2 activity is necessary for maintaining basal levels of arachidonic acid, lysophospholipid, and PAF for performing normal brain function.

Increased PLA2 activity and excessive production of proinflammatory mediators, eicosanoids, and platelet-activating factor, may potentially lead to disease states and neuronal injury. Collective evidence from many recent studies suggests that increased PLA2 activity and PLA2-generated mediators play a central role not only in acute inflammatory responses in brain but also in oxidative stress associated with neurological disorders such as ischemia, Alzheimer's disease (AD), Parkinson's disease (PD), and multiple sclerosis (MS) (Kalyvas and David, 2004; Phillis and O'Regan, 2004; Sun et al., 2004). PLA2 contributes to the pathogenesis of the above disorders by attacking neural membrane phospholipids and releasing proinflammatory lipid mediators such as prostaglandins, leukotrienes, and thromboxanes, and PAF, and also by generating 4-hydroxynonenal (4-HNE). Thus, inhibition of cPLA2 activity provides an attractive approach for designing novel drugs for the treatment of inflammation and oxidative stress associated with acute neural trauma such as ischemia, spinal cord injury, and head injury and some neurodegenerative disorders such as AD, PD, and MS. A definitive proof of whether enhanced PLA2 activity represents a part of the cellular defense mechanism or whether increased PLA2 activities contribute to the pathology of the neurological disorder awaits the development of a potent, specific, and clinically useful PLA2 inhibitor for the treatment of inflammation and oxidative stress associated with neurological disorders. The purpose of this review is to describe the current knowledge of PLA2 involvement in neurological disorders and the usefulness of PLA2 inhibitors in preventing neurotoxin-mediated damage to neurons, glial cells, and the myelin sheath in brain tissue. This discussion should initiate more studies on not only the involvement of PLA2-mediated inflammation and oxidative stress in neurological disorders but also on the development of potent, specific, and nontoxic inhibitors of PLA2 activity that can cross the blood-brain barrier without harm and can be used for the treatment of neurological disorders.

II. Multiplicity of Phospholipases A2 in Brain

Recent advances in molecular and cellular biology of PLA2 have led to the identification of more than 20 isoforms with PLA2 activity. PLA2 enzymes are subdivided into several groups depending upon their structure, enzymic properties, subcellular localization, and cellular function (Table 1) (Farooqui et al., 1997c; Chakraborti, 2003; Phillis and O'Regan, 2004; Sun et al., 2004). These groups include secretory phospholipase A2 (sPLA2), cytosolic phospholipase A2 (cPLA2), plasmalogen-selective phospholipase A2 (PlsEtn-PLA2), and calcium-independent phospholipase A2 (iPLA2). Each class of PLA2 is further subdivided into isozymes of which there are 14 for sPLA2, at least 4 for cPLA2, and 2 for iPLA2. Genes coding for sPLA2, cPLA2, and iPLA2 have been shown to occur in different regions of brain, in neurons, microglia, and astrocytes (Molloy et al., 1998; Zanassi et al., 1998; Balboa et al., 2002). PLA2 isoforms have been partially purified and characterized from brain tissue (Hirashima et al., 1992; Ross et al., 1995; Yang et al., 1999), but none have been cloned and fully characterized. The following section briefly describes properties of brain PLA2 isoforms, their roles, and importance in signal transduction processes in brain metabolism (Fig. 1).

View this table:
  • View inline
  • View popup
TABLE 1

Properties of various isoforms of brain PLA2 Data summarized from Hirashima et al. (1992); Matsuzawa et al. (1996); Yang et al. (1999); and Farooqui et al. (2000b).

A. Secretory Phospholipase A2

sPLA2 is synthesized intracellularly; then it is secreted and acts extracellularly. sPLA2 has a molecular mass of 14 kDa and is mainly associated with synaptosomes and synaptic vesicle fractions (Kim et al., 1995; Matsuzawa et al., 1996). PLA2 binds to two types of cell surface receptors, namely the N type, identified in neurons, and the M type, identified in skeletal muscles, of sPLA2 receptors (Hanasaki and Arita, 2002). Brain sPLA2 contains a secretion peptide and requires millimolar concentrations of Ca2+ for enzymic activity. It shows no selectivity for particular fatty acyl chains in the phospholipids. This enzyme is present in all regions of mammalian brain. The highest activities of sPLA2 are found in medulla oblongata, pons, and hippocampus, moderate activities in the hypothalamus, thalamus, and cerebral cortex, and low activities in the cerebellum and olfactory bulb (Thwin et al., 2003). At the cellular level, the sPLA2 transcript is found in astrocytes (Mosior et al., 1998b; Zanassi et al., 1998). sPLA2 is present in differentiated PC12 cells and in rat brain synaptic vesicles, indicating that neurons also express sPLA2 activity (Matsuzawa et al., 1996).

Rat brain synaptosomes or differentiated PC12 cells release sPLA2 upon stimulation via acetylcholine and glutamate receptors or via voltage-dependent calcium channels through depolarization. Thus, sPLA2 may play an important role in neuronal metabolism (Kim et al., 1995; Matsuzawa et al., 1996). Based on pharmacological studies, the sPLA2 released from neuronal cells may modulate the degranulation process leading to the release of neurotransmitters. Inhibitors of sPLA2 activity block this release. For the expression of neurotoxicity, the released sPLA2 binds to the presynaptic membrane, enters the lumen of the synaptic vesicle during retrieval of the vesicle from the plasma membrane, and hydrolyzes phospholipids of the inner leaflet of synaptic vesicles, changing the phospholipid composition and thus impairing its endocytosis. The stimulation of sPLA2 in synaptic vesicles correlates with the induction of vesicle-vesicle aggregation. This process plays a central role in presynaptic neurotransmission (Moskowitz et al., 1983; Matsuzawa et al., 1996; Wei et al., 2003). In brain, astrocytes express sPLA2, which can be induced in response to proinflammatory cytokines such as tumor necrosis factor-α and interleukin-1β (Lin et al., 2004).

Fig. 1.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 1.

A hypothetical diagram showing interplay among lipid mediators generated by sPLA2, cPLA2, PlsEtn-PLA2, and iPLA2 in brain tissue. PM, plasma membrane; A1, agonist; R1, receptor; A2, agonist; R2, receptor; A3, agonist; R3, receptor; sPLA2-R, sPLA2 receptor; PtdCho, phosphatidylcholine; PlsEtn, ethanolamine plasmalogen; LysoPtdCho, lysophosphatidylcholine; LysoPlsEtn, lysoethanolamine plasmalogen; AA, arachidonic acid; FFA, free fatty acid; (+), stimulation; (-), inhibition.

Mitochondrial fractions from rat brain, PC12, and U251 astrocytoma cell cultures contain significant sPLA2 and iPLA2 activities (Macchioni et al., 2004). The mechanism for a secretory protein (like sPLA2) targeting an intracellular organelle (like mitochondria) remains unknown. However, it is proposed that at the molecular level, heparan sulfate, a glycosaminoglycan, may play an important role in internalization and attachment of PLA2 isoforms to intracellular organelles (Farooqui et al., 1994b; Boilard et al., 2003). A reduction in the mitochondrial membrane potential causes the release of sPLA2 and this sPLA2 along with other PLA2 isoforms may be involved in neural cell injury (Farooqui et al., 1997d; Macchioni et al., 2004).

Glutamate and its analogs stimulate sPLA2 activity in a dose- and time-dependent manner (Kim et al., 1995; Xu et al., 2003). The neurotoxicity of glutamate is synergistically increased with the addition of sPLA2 to cortical cultures. This observation suggests that glutamatergic synaptic activity may be modulated by sPLA2 and its receptors on the neuronal surface (DeCoster et al., 2002; Kolko et al., 2002). In PC12 cells, sPLA2 induces neurite outgrowth. Mutants with reduced sPLA2 activity exhibit a comparable reduction in neurite-inducing activity (Nakashima et al., 2003), indicating that sPLA2 performs a neurotrophin-like role in the central nervous system.

B. Cytosolic Phospholipases A2

Although brain tissue contains cPLA2 activity, it has never been purified to homogeneity and characterized from brain. The cytosolic fraction from rat brain contains two forms of PLA2 activity, PLA2-H and PLA2-L. PLA2-H has an apparent molecular mass of 200 to 500 kDa. Its activity is partially inhibited by Ca2+. In contrast, PLA2-L has a molecular mass of 100 kDa and requires Ca2+ (Yoshihara and Watanabe, 1990). Based on several enzymic properties such as Ca2+ sensitivity, molecular mass, and Ca2+-mediated translocation, PLA2-L seems to be identical to cPLA2 (Yoshihara et al., 1992). cPLA2 prefers arachidonic acid over other fatty acids and does not use Ca2+ for catalysis, although submicromolar Ca2+ concentrations are needed for membrane binding (Clark et al., 1987; Farooqui et al., 2000b). Owing to the presence of a Ca2+-dependent phospholipid-binding domain at the N-terminal region, cPLA2 is translocated in a Ca2+-dependent manner from cytosol to the nuclear or other cellular membranes (Clark et al., 1987; Hirabayashi et al., 2004), in which other downstream enzymes, including the cyclooxygenases and lipoxygenases responsible for the metabolism of arachidonic acid to eicosanoids, are located. This gives cPLA2 access to its membrane-associated phospholipid substrate. However, we do not know the manner in which cPLA2 is activated by extracellular stimuli and whether this activation occurs at specific sites. The C-terminal region of cPLA2 contains the phosphorylation site and catalytic site. These sites may be involved in regulation of the enzymic activity. The activation of cPLA2 can be through serine residues, notably Ser-505 and Ser-727, by mitogen-activated protein kinase and protein kinase C (PKC) (Hirabayashi and Shimizu, 2000; Hirabayashi et al., 2004). In neural membranes, cPLA2 activity and arachidonic acid release are linked to dopamine, glutamate, serotonin, P2-purinergic, muscarinic, cytokine, and growth factor receptors through different coupling mechanisms (Table 2). Some receptors involve G-proteins and others do not. The ligand-mediated stimulation of the above receptors modulates the release of arachidonic acid and levels of other second messengers in brain tissue (Farooqui et al., 2000b).

View this table:
  • View inline
  • View popup
TABLE 2

Coupling of PLA2 isoforms with various receptors in brain tissue.

cPLA2 activity can also be modulated through a cooperative binding mechanism with glycerophospholipids containing arachidonic acid (Burke et al., 1995) or through binding of anionic phospholipids, such as phosphatidylinositol 4,5-bisphosphate, phosphatidylinositol 3,4,5-trisphosphate, and ceramide 1-phosphate (Hirabayashi et al., 2004; Pettus et al., 2004), to a pleckstrin homology domain (Mosior et al., 1998a). Three paralogs of cPLA2 occur in brain and other non-neural tissues (Diaz-Arrastia and Scott, 1999; Farooqui et al., 2000b; Hirabayashi et al., 2004). They are cPLA2-α (molecular mass 85 kDa), cPLA2-β (molecular mass 114 kDa), and cPLA2-γ (molecular mass 61 kDa). cPLA2-β is found mainly in the cerebellum and shares more similarities with cPLA2-α than with cPLA2-γ. cPLA2-γ lacks the C2 domain, but contains a prenyl group-binding motif that behaves as a lipid anchor and allows binding of the enzyme to the membrane.

Recombinantly expressed cPLA2-γ liberates arachidonic acid from phosphatidylcholine. Unlike cPLA2-α, cPLA2-γ also acts on other fatty acid residues at the sn-2 and sn-1 positions of glycerophospholipids. cPLA2-α hydrolyzes fatty acids at the sn-2 position, cPLA2-β prefers to cleave fatty acids at the sn-1 position, and cPLA2-γ efficiently hydrolyzes fatty acid at the sn-1 as well as sn-2 positions of the glycerol moiety (Song et al., 1999). The overexpression of cPLA2-γ increases the proportions of polyunsaturated fatty acids in phosphatidylethanolamine, indicating that this paralog can modulate the phospholipid composition (Asai et al., 2003). cPLA2-γ is constitutively expressed in the endoplasmic reticulum where it is involved in remodeling and maintaining membrane phospholipid composition under oxidative stress.

cPLA2-β displays much lower activity with [2-arachidonyl]PtdCho than do the other two paralogs. The genes for human cPLA2-α, -β, and -γ map to chromosomes 1, 15, and 19, respectively. Mitogen-activated protein kinase phosphorylation sites are only present in cPLA2-α and are not conserved in cPLA2-β and cPLA2-γ. cPLA2-α activity is uniformly distributed in various regions of rat brain (Farooqui et al., 2000b). Recent studies have indicated the presence of a new paralog of cPLA2. This paralog is mainly found in skin and has been named cPLA2-δ (molecular mass 109 kDa) (Chiba et al., 2004). In contrast with other cPLA2 paralogs, cPLA2-δ has a preference for linoleic acid release instead of arachidonic acid release.

Considerable information is available on cPLA2-α. From immunolabeling and in situ hybridization studies, cPLA2-α is localized in somata and dendrites of Purkinje cells, whereas cPLA2-β is present in the granule cells of rat brain (Shirai and Ito, 2004). In addition, cPLA2-α is predominantly found in astrocytes of gray matter (Farooqui et al., 2000b; Pardue et al., 2003), as well as in hippocampal neurons (Sandhya et al., 1998; Kishimoto et al., 1999; Strokin et al., 2003), in which under physiological conditions cPLA2-α may be involved in second-messenger generation and long-term potentiation (LTP), a mechanism involved in memory storage. More recently, the mRNAs for cPLA2-β and cPLA2-δ have been identified by reverse transcription-polymerase chain reaction analysis in human brain tissue (Pickard et al., 1999; Song et al., 1999; Hirabayashi et al., 2004), but the role of these paralogs of cPLA2 in brain tissue remains speculative. Because cPLA2-δ mainly occurs in skin, it is proposed that this paralog plays a critical role in inflammation in psoriatic lesions (Chiba et al., 2004).

Even though cPLA2 isozymes are often considered to be the enzymes responsible for stimulus-mediated arachidonic acid release and eicosanoid formation, several studies also implicate sPLA2 activation. cPLA2 and sPLA2 may modulate arachidonic acid metabolism in the astrocytoma cell line 1321N1 via the mitogen-activated protein kinase pathway (Hernandez et al., 2000). At the cellular level, interactions and interplay among calcium mobilization, cPLA2 phosphorylation, and extracellular receptors of sPLA2 may be responsible for increased eicosanoid production (Fig. 1). cPLA2 may be a good candidate for triggering down-regulation of nitricoxide synthase activity and may thus be an important component of the cross-talk between calcium and nitric oxide-regulated signal transduction pathways in neuronal cells (Palomba et al., 2004). Under pathological conditions, the interactions among calcium, cPLA2, and nitric-oxide synthase may play an important role in the pathophysiology of neurological disorders associated with oxidative stress and inflammation (see below).

C. Plasmalogen-Selective Phospholipase A2

This enzyme hydrolyzes arachidonic acid and docosahexaenoic acid from the sn-2 position of plasmalogens, a special type of glycerophospholipid with a vinyl ether linkage at the sn-1 position of the glycerol backbone (Farooqui and Horrocks, 2001). This enzyme has been purified and characterized from bovine brain cytosol (Hirashima et al., 1992), rabbit kidney (Portilla and Dai, 1996), and rabbit heart (Hazen and Gross, 1993). Bovine brain PlsEtn-PLA2 has an apparent molecular mass of 39 kDa. Nonionic detergents, Triton X-100 and Tween 20, stimulate the enzymic activity. It is not inhibited by bromoenol lactone, an inhibitor that markedly inhibits iPLA2. Low micromolar concentrations of ATP have no effect on PlsEtn-PLA2 activity, but 2 mM ATP markedly inhibits its activity. Bovine brain PlsEtn-PLA2 is inhibited by 5,5′-dithiobis(2-nitrobenzoic acid, iodoacetate, and N-ethylmaleimide in a dose-dependent manner (Farooqui et al., 1995). Various polyvalent anions, citrate > sulfate > phosphate, and metal ions, Ag+, Hg2+, and Fe3+, also inhibit this enzyme in a dose-dependent manner. Glycosaminoglycans markedly inhibit bovine brain PlsEtn-PLA2 with an inhibition pattern of heparan sulfate > hyaluronic acid > chondroitin sulfate > heparin. This PLA2 is also inhibited by N-acetylneuraminic acid, gangliosides, and sialoglycoproteins (Yang et al., 1994b). Other glycosphingolipids, such as cerebrosides and sulfatides, have no effect. However, ceramide markedly stimulates PlsEtn-PLA2 activity in a time-and dose-dependent manner (Latorre et al., 2003). Treatment of rat brain slices with Staphylococcus aureus sphingomyelinase or C2-ceramide produces a marked decrease in PlsEtn levels, suggesting stimulation of PlsEtn-PLA2 activity. Bromoenol lactone, a potent inhibitor of iPLA2, does not affect this stimulation, but quinacrine and gangliosides, nonspecific inhibitors of PlsEtn-PLA2, completely block it (Latorre et al., 2003; Yang et al., 1994a,b). These studies have led to the suggestion that the degradation of plasmalogen by PlsEtn-PLA2 is a receptor-mediated process (Farooqui and Horrocks, 2001; Farooqui et al., 2003a; Latorre et al., 2003) and may involve an interaction between plasmalogen metabolism and sphingolipid metabolism.

PlsEtn-PLA2 has been localized immunochemically in neurons and astrocytes (Farooqui and Horrocks, 2001). The colocalization of PlsEtn-PLA2 with glial fibrillary acidic protein suggests that this PLA2 is predominantly associated with astrocytes. This is in contrast with cPLA2-α, which is present in neurons as well as astrocytes (Sandhya et al., 1998; Kishimoto et al., 1999). cPLA2 (molecular mass 85 kDa) and PlsEtn-PLA2 (molecular mass 39-42 kDa) release arachidonic acid and docosahexaenoic acid in rat brain astrocytes and cyclic AMP and Ca2+ regulate these enzymes differentially (Strokin et al., 2003). Because plasmalogens are major phospholipids of neural membranes, PlsEtn-PLA2 may be mainly involved in generating docosahexaenoic acid (DHA), a 22-carbon essential fatty acid with 6 double bonds. This fatty acid is highly enriched in synaptosomal membranes, synaptic vesicles, and growth cones and accounts for >17% by weight of the total fatty acids in the brain of adult rats (Hamano et al., 1996).

D. Calcium-Independent Phospholipases A2

The brain cytosolic fraction contains an 80-kDa Ca2+-independent PLA2 activity. This enzyme has been purified from rat brain to homogeneity using multiple column chromatographic procedures with a very low yield. The purified enzyme has a specific activity of 4.3 μmol/min/mg. The peptide sequence of this enzyme has considerable homology to sequences of the iPLA2 from P388D1 macrophages, Chinese hamster ovary cells, and human B lymphocytes (Yang et al., 1999). This iPLA2 hydrolyzes the sn-2 fatty acid from PtdCho with preferences linoleoyl > palmitoyl > oleoyl > arachidonyl group. iPLA2 has an unique amino acid sequence containing a lipase consensus sequence and eight ankyrin repeats. This enzyme is strongly inhibited by bromoenol lactone, and ATP augments its activity. iPLA2 is present in all brain regions with the highest activity in striatum, hypothalamus, and hippocampus. The gene encoding iPLA2 has been identified (Molloy et al., 1998). Alternative splicing can generate multiple iPLA2 isoforms with distinct tissue distribution and localization (Larsson et al., 1998). Truncated splice variant proteins that prevent the formation of active iPLA2 tetramers may negatively regulate iPLA2 (Larsson et al., 1998; Seashols et al., 2004). Native iPLA2 is a homotetramer that is potentially formed through interactions between N-terminal ankyrin repeats (Ackermann and Dennis, 1995). Five splice variants of iPLA2 occur in various tissues (Larsson et al., 1998; Shirai and Ito, 2004). One splice variant (iPLA2-1) lacks exon 9 (165 base pairs), whereas the other four variants (iPLA2-2, iPLA2-3, iPLA2-ankyrin-1, and iPLA2-ankyrin-2) contain exon 9. The presence of this exon makes these splice variants membrane-bound because exon 9 encodes hydrophobic amino acids (Larsson et al., 1998; Shirai and Ito, 2004). Rat cerebellum contains iPLA2-1 and iPLA2-2 or iPLA2-3, but not iPLA2-ankyrin-1 or iPLA2-ankyrin-2. From immunolabeling studies in rat brain, granule cells, stellate cells, and the nucleus of Purkinje cells contain iPLA2 (Shirai and Ito, 2004). Strong signals of iPLA2 immunoreactivity are observed in olfactory bulb, hippocampus CA1-3, dentate gyrus, and brain stem. In non-neural cells, the cleavage of iPLA2 by caspase-3 is associated with the execution of apoptosis (Atsumi et al., 1998, 2000). The proposed role of iPLA2 in phospholipid remodeling and apoptosis is based on the use of bromoenol lactone, thought to be a specific inhibitor, but this compound actually inhibits other enzymes such as diacylglycerol lipase and phosphatidate phosphohydrolase (see below). This makes it difficult to define the role of iPLA2 in phospholipid metabolism (Farooqui et al., 2000a; Pérez et al., 2004). iPLA2 may play an important role not only in long-term potentiation and long-term depression and activity-dependent changes in synaptic strength believed to underlie certain forms of learning and memory in the hippocampus (Fitzpatrick and Baudry, 1994; Wolf et al., 1995; Fujita et al., 2001) but also in neural cell proliferation, apoptosis, and differentiation (Farooqui et al., 2004a).

III. Involvement of Phospholipase A2 Activity in Brain Injury

A growing body of evidence suggests the involvement of isoforms of PLA2 in neurotransmitter release, LTP, long-term depression, membrane repair, neurodegeneration, neural cell proliferation, differentiation, and apoptosis (Fig. 2) (Farooqui et al., 1997c). It is not known which PLA2 isoform performs which central nervous system function. The multiplicity of PLA2 and the interplay among lipid mediators generated by PLA2 in brain tissue provide diversity in function and specificity of various isoforms of this family of enzymes in the regulation of enzymic activity in response to a wide range of extracellular signals. However, this diversity complicates the analysis of their function (Fig. 1). The complexity of this problem becomes obvious when one considers the coupling of various isoforms of PLA2 with different receptors in a single neural cell and tries to associate PLA2 activity with specific neuronal functions and disease processes.

Isoforms of PLA2 may not function interchangeably but act in parallel to the transducer signal (Farooqui et al., 1997d). Various isoforms of PLA2 probably act on different cellular pools of phospholipids located in different types of neural cells and these isoforms may be regulated by different coupling mechanisms involving common second messengers (Sun et al., 2004). The synthesis of eicosanoids depends not only on PLA2 activity for the generation of arachidonic acid, but also on cyclooxygenase or lipoxygenase activities (Sun et al., 2004). Tight regulation of PLA2 activity is necessary for normal brain function (Farooqui et al., 2000b; Phillis and O'Regan, 2004).

Fig. 2.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 2.

Proposed roles of PLA2 isoforms in brain.

Activities of PLA2 isoforms are also regulated through modulation of gene expression. Interleukin (IL)-1α, (Xu et al., 2003), tumor necrosis factor (TNF)-α (Pirianov et al., 1999; Tong et al., 1999; Jupp et al., 2003), interferon-γ (Xu et al., 2003), and several growth factors induce these enzymes (Jupp et al., 2003; Akiyama et al., 2004; Wang et al., 2004b). Regulatory interactions and coupling between PLA2 activity and the expression of genes encoding cyclooxygenase and lipoxygenases have been suggested (Doug et al., 2003). In brain tissue, the coupling between cPLA2 depends not only on the type of neural cell but also on the stimulatory status of neural cell involved.

Stimulation of PLA2 isoforms may contribute to brain damage in several ways:

  1. The loss of essential phospholipids with the accumulation of free fatty acids and lysophospholipids may have a detergent-like effect on neuronal membranes.

  2. Free fatty acids can uncouple oxidative phosphorylation, which results in mitochondrial dysfunction (Schapira, 1996). Arachidonic acid produces mitochondrial swelling in glial cells and induces changes in membrane permeability by regulating ion channels (Farooqui et al., 1997b,c). It also inhibits glutamate uptake. In the nucleus, arachidonic acid may also interact with elements of gene structure, such as promoters, enhancers, suppressors, and others, in a specific manner that is not shared by eicosanoids or other fatty acids. These interactions modulate gene expression (Farooqui et al., 1997c).

  3. Polyunsaturated fatty acid-induced oxidative stress is accompanied by an increase in AP-1 and NF-κB activity and gene expression (Mazière et al., 1999). Another consequence of increased polyunsaturated free fatty acid-induced oxidative stress is the activation and inactivation of redox-sensitive proteins (Wang, 2003).

  4. PAF, formed from the acetylation of lysophospholipids, not only can activate leukocytes and microglia but also can induce inflammation at endothelial and neuronal cell surfaces.

  5. The accumulation of free fatty acids can trigger an uncontrolled “arachidonic acid cascade”. This sets the stage for increased production of prostaglandins, leukotrienes, and thromboxanes. Cyclooxygenases and lipoxygenases catalyze these reactions. Prostaglandins, leukotrienes, and thromboxanes are collectively called eicosanoids. They play important roles in the generation and maintenance of inflammation in neural cells. The arachidonic acid cascade also produces 4-HNE and reactive oxygen species (ROS) such as superoxide anion, hydroxyl, alkoxyl, and peroxyl radicals and hydrogen peroxide.

  6. An uncontrolled sustained increase in calcium influx through increased phospholipid degradation can lead to increased membrane permeability and stimulation of many enzymes associated with lipolysis, proteolysis, and disaggregation of microtubules with a disruption of cytoskeleton and membrane structure (Farooqui and Horrocks, 1991, 1994). The most compelling evidence for the involvement of cPLA2 in neurodegeneration comes from reports indicating that mice with a targeted deletion of the gene that encodes cPLA2 show reduced infarct size after cerebral ischemia (Bonventre et al., 1997) and resistance to MPTP-induced neurotoxicity (Klivenyi et al., 1998). Studies on brain lipid metabolism in cPLA2 knockout mice indicate there is no net change in unesterified arachidonic acid. However, there was a 50% reduction in esterified arachidonic acid in phosphatidylcholine, indicating involvement of cPLA2 in wild-type mice (Rosenberger et al., 2003). The knockout mice also have reduced rates of arachidonic acid incorporation into ethanolamine and choline glycerophospholipids but elevated rates into phosphatidylinositol. cPLA2-deficient mice also show a 62% reduction in the rate of formation of prostaglandin (PG) E2, suggesting a coupling between cPLA2 and cyclooxygenase activities (Murakami et al., 1997; Bosetti and Weerasinghe, 2003). The deletion of cPLA2 also causes dysregulation ofinsulin-like growth factor-1 signaling and stimulates striated muscle growth (Obata et al., 2003).

4-HNE has three functional groups that confer to its molecule a very high reactivity toward thiol and amino groups. 4-HNE has been reported to cause a number of deleterious effects in cells including inhibition of DNA synthesis, disturbance in calcium homeostasis, and inhibition of mitochondrial respiration. All of these processes may result in neuronal injury in ischemia and glutamate-mediated neurotoxicity (Farooqui et al., 1997c). In brain tissue, 4-HNE also produces alterations in the function of key membrane proteins including glucose transporter, glutamate transporter, and sodium potassium ATPase (Friguet et al., 1994; Jamme et al., 1995). Thus, high levels of this metabolite are toxic for brain tissue (Farooqui et al., 1997c).

ROS inactivate membrane proteins and DNA (Berlett and Stadtman, 1997). The reaction between ROS and proteins or unsaturated lipids in the plasma membrane leads to a chemical cross-linking of membrane proteins and lipids and a reduction in membrane unsaturation. This depletion of unsaturation in membrane lipids is associated with decreased membrane fluidity and decreases in the activity of membrane-bound enzymes, ion channels, and receptors (Ray et al., 1994).

IV. Physiological and Pharmacological Effects of Phospholipase A2 Inhibitors

The stimulation of PLA2 isoforms, release of arachidonic acid, and generation of platelet-activating factor are important events in the inflammation and oxidative stress associated with acute neural trauma and chronic neurological disorders (Farooqui and Horrocks, 1994; Phillis and O'Regan, 2004). Treatment of these disorders requires potent and selective inhibitors of PLA2 activity that can be used as drugs. The problem with available inhibitors of PLA2 isoforms has been their specificity. Many PLA2 inhibitors originally thought to be selective for a specific PLA2 isoform are now known to not only inhibit other PLA2 isoforms but also block activities of different enzymes (Farooqui et al., 1999; Cummings et al., 2000; Fuentes et al., 2003). For example, in non-neural cells, arachidonyl trifluoromethyl ketone inhibits not only cPLA2 activity but also cyclooxygenase and acyltransferase activities (Cummings et al., 2000; Fuentes et al., 2003). Methyl arachidonyl fluorophosphonate, another inhibitor of brain cPLA2, also inhibits bovine brain iPLA2.

The discovery of potent and specific inhibitors of PLA2 isoforms is an important approach, not only for establishing functional roles of a given PLA2 isoform in a specific type of neural cell in brain tissue but also for treating oxidative stress and inflammation caused by neurodegenerative process. Studies on this important topic are beginning to emerge (Farooqui et al., 1999; Cummings et al., 2000; Miele, 2003a; Scott et al., 2003; Clark and Tam, 2004). These studies are complicated not only by the lack of information on the availability of specific inhibitors but also by the occurrence of isoforms of PLA2 activity and in vivo effects of PLA2 inhibitors on enzymic activity. Furthermore, the effect of inhibitors on the physical state of substrate aggregates in neural membranes remains unknown. In searching for good cPLA2 inhibitors, kinetic analysis is not enough to evaluate whether an inhibitor can block PLA2 activity by affecting the interfacial quality of phospholipid in lipid bilayer or by directly inhibiting the interaction between the phospholipids and the active site of the enzyme. An ideal inhibitor should have regional specificity and should be able to reach the site where cells are under oxidative stress and inflammatory and neurodegenerative processes are taking place.

Neurons are more susceptible to free radical-mediated neuroinflammation and oxidative stress than glial cells (Adibhatla et al., 2003; Ajmone-Cat et al., 2003). In fact, activated glial cells, including astroglia and microglia, sustain inflammatory processes initiated by arachidonic acid-generated metabolites. This suggests that signals modulating the induction, expression, and stimulation of PLA2 isoforms may play an important role in neurodegenerative diseases associated with neuroinflammation and oxidative stress (Farooqui and Horrocks, 1994; Farooqui et al., 2003b, 2004b). For the successful treatment of inflammatory and oxidative stress in neurological disorders, timely delivery of a well-tolerated, chronically active, and specific inhibitor of PLA2 that can bypass or cross the blood-brain barrier without harm is required. Some nonspecific PLA2 inhibitors (see below) have been used for the treatment of ischemia, spinal cord injury, and AD (Sano et al., 1997), but no compound with real clinical potential has emerged.

A. Arachidonyl Trifluoromethyl Ketone

Arachidonyl trifluoromethyl ketone (Fig. 3) is a potent inhibitor of cPLA2. NMR studies show that the carbon chain of AACOCF3 binds in a hydrophobic pocket and the carbonyl group of AACOCF3 forms a covalent bond with the serine 228 in the active site, generating a charged hemiketal oxoanion that interacts with a positively charged group of the enzyme (Street et al., 1993; Trimble et al., 1993). AACOCF3 is a 500-fold more potent inhibitor of cPLA2 than sPLA2 (Trimble et al., 1993), indicating that it is a more selective inhibitor of cPLA2 than sPLA2. This inhibitor also blocks cyclooxygenase activity (Riendeau et al., 1994).

Because of its physicochemical properties, AACOCF3 can readily penetrate into cell membranes. At 5 to 20 μM it essentially blocks all liberation of arachidonic acid in thrombin-stimulated platelets, in Ca2+ ionophore-stimulated human monocytic cells, and in interleukin 1-stimulated mesangial cells (Gronich et al., 1994). AACOCF3 inhibits bovine brain cPLA2 and iPLA2 in a dose-dependent manner with IC50 values of 1.5 and 6.0 μM, respectively.

The treatment of NG 108-15 cells with AACOCF3 decreases initial neurite formation in a concentration-dependent manner (Smalheiser et al., 1996). The pharmacological blockade of cPLA2 by a low concentration, 10 μM, of AACOCF3 significantly inhibits neuronal death in the CA1 region of rat hippocampus. In primary neuronal cultures, this PLA2 inhibitor prevents caspase-3 activation and neurodegeneration induced by β-amyloid peptide (Aβ) and human prion protein peptide (Bate et al., 2004), suggesting the role of PLA2 isoforms in neurodegenerative processes (Farooqui and Horrocks, 1994; Farooqui et al., 1997c). In primary neuronal cultures, AACOCF3 also abolishes methylmercury (MeHg2+)-mediated stimulation of cPLA2 and arachidonic acid release (Shanker et al., 2004), suggesting that cPLA2 plays an important role in MeHg2+-induced neurotoxicity. Similarly, in astrocytes MeHg2+-induced ROS generation is strongly inhibited by AACOCF3 (Shanker and Aschner, 2003). In neural cell cultures AACOCF3 also shortens the association of PKC-γ with plasma membrane indicating that this isoform of PKC may be involved in neuronal plasticity (Yagi et al., 2004). AACOCF3 also blocks l-buthionine sulfoximine toxicity in glutathione-depleted mesencephalic cultures (Kramer et al., 2004).

Fig. 3.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 3.

Chemical structures of PLA2 inhibitors. a, arachidonyl trifluoromethyl ketone; b, methyl arachidonyl fluorophosphonate; c, bromoenol lactone; d, 4-alkoxybenzamidine; e, long-chain 2-oxomide; f, quinacrine; and g, 3-(pyrrole)-2-propionic acid.

AACOCF3 induces the dispersal of Golgi stack and trans-Golgi network resident proteins. This suggests that cPLA2 isozymes play a crucial role in membrane trafficking and in maintenance of Golgi architecture. In non-neural cultured cells, AACOCF3 inhibits the expression of IL-2 at both the mRNA and protein levels, indicating that cPLA2 may have marked effects on T-cell function (Amandi-Burgermeister et al., 1997; Ouyang and Kaminski, 1999). AACOCF3 inhibited DNA fragmentation during apoptosis in U937 cells, but failed to affect the morphological changes that occur during apoptosis, suggesting that the use of AACOCF3 may distinguish between cytoplasmic and nuclear events that occur during apoptotic cell death (Vanags et al., 1997). AACOCF3 is described as a specific inhibitor of cPLA2, but recent studies in non-neural cells indicate it may also inhibit cyclooxygenases and 5-lipoxygenase (Cummings et al., 2000; Fuentes et al., 2003). These observations strongly suggest that AACOCF3 is not a specific inhibitor of cPLA2.

B. Methyl Arachidonyl Fluorophosphonate

MAFP (Fig. 3) is an irreversible inhibitor of bovine brain cPLA2 (IC50 0.5 μM) and has no effect on sPLA2.It inhibits enzymic activity by reacting with a serine residue at the active site. MAFP also inhibits bovine brain iPLA2 in a dose-dependent manner with an IC50 value of 0.75 μM. At 5.0 μM, MAFP completely inhibits bovine brain iPLA2 activity. In addition, MAFP inhibits Aβ-mediated stimulation of cPLA2 activity in cortical neuronal cultures (Kriem et al., 2005).

MAFP induces irreversible inhibition of the enzymic hydrolysis of arachidonyl ethanolamide (anandamide) by fatty acid amide hydrolase. Based on various pharmacological studies with cannabinoid CB1 receptors and MAFP, it has been suggested that MAFP is an irreversible cannabinoid CB1 receptor antagonist (Fernando and Pertwee, 1997).

Because MAFP interacts with several PLA2 isoforms and with fatty amide hydrolase, it cannot be considered as a specific inhibitor of cPLA2. MAFP also blocks l-buthionine sulfoximine toxicity in glutathione-depleted mesencephalic cultures (Kramer et al., 2004). Intrathecal injections of MAFP in rats produce antinociceptive effects (Ates et al., 2003), suggesting that PLA2 isoforms may play some role during pain states (see below).

C. Bromoenol Lactone

BEL (Fig. 3) is a potent inhibitor of bovine brain iPLA2 and PlsEtn-PLA2 with IC50 values of 60 and 40 nM, respectively. BEL has a structural resemblance to plasmalogen. It inhibits brain cPLA2 and sPLA2 at a very high concentration (500 μM). The injection of BEL (10 μM) into postsynaptic CA1 pyramidal neurons produces a robust increase in the amplitude of α-amino-3-hydroxy-5-methylisoxazole-4-propionate (AMPA) receptor-mediated excitatory postsynaptic currents, suggesting that iPLA2 plays an important role in AMPA-mediated synaptic plasticity (St-Gelais et al., 2004). AACOCF3 and palmityl trifluoromethyl ketone, which mainly interact with cPLA2, have no effect on AMPA-mediated synaptic transmission. The inhibition of iPLA2 by BEL and the enhancement of AMPA subunit immunoreactivity in brain homogenates and slices support the above electrophysiological studies. Taken together, these results support the hypothesis that BEL-mediated antagonism of AMPA receptors is involved in long-term potentiation and long-term depression during the regulation of synaptic plasticity. Based on the blockage of induction of hippocampal long-term potentiation, brain iPLA2 may be involved in learning and memory (Wolf et al., 1995; Fujita et al., 2001).

Intracerebroventricular injections of BEL, 3 nmol, markedly affect spatial performance in mice (Fujita et al., 2000), indicating that iPLA2 is involved in spatial memory formation. BEL also modulates intracellular membrane trafficking (Kuroiwa et al., 2001; Brown et al., 2003) by inhibiting iPLA2 activity in membrane tubule formation during reassembly of the Golgi complex. In addition, BEL treatment also interferes with membrane fusion events during endocytosis and exocytosis.

D. Benzenesulfonamides and Alkoxybenzamidines

Benzenesulfonamide (Fig. 3) and its piperidine derivative are the most potent inhibitors of membrane-bound heart PLA2 activity with IC50 values of 28.0 and 9.0 nM, respectively (Oinuma et al., 1991). Intravenous injections of these inhibitors protect rats against ischemic damage in acute myocardial infarction. These compounds are relatively metabolically stable in plasma with half-lives of 1 to 2 h. These compounds also inhibit brain cPLA2 activity in a dose-dependent manner with IC50 values of 23.0 and 10.0 nM, respectively. Nothing is known about their ability to cross the blood-brain barrier. The synthesis of 4-alkoxybenzamidines as PLA2 inhibitors has also been described previously (Aitdafoun et al., 1996). These compounds competitively inhibit bovine pancreatic and rabbit platelet PLA2 activities with IC50 values of 3.0 and 5.0 μM, respectively. It is interesting to note that 4-tetradecyloxybenzamidine has an anti-inflammatory effect in vivo on carrageenan-mediated rat paw edema.

E. 3-(Pyrrol)-2-propionic Acid

Lehr (1996, 1997a,b) has synthesized many compounds (Fig. 3) that inhibit cPLA2 activity in platelets with IC50 values varying from 0.5 to 10 μM. Effects of these compounds on other isoforms of PLA2 remain unknown. These inhibitors have not been used to block the activity of brain PLA2 isoforms and have not been injected into intact animals or animal models of neurological disorders; therefore, nothing is known about their tolerance, half-lives, and toxicity. The ability of these inhibitors to cross the blood-brain barrier also remains unknown.

F. 2-Oxoamide and 1,3-Disubstituted Propan-2-ones

Long-chain 2-oxoamides of γ-aminobutyric acid and γ-norleucine (AX006, AX007, and AX008) (Fig. 3) reversibly inhibit cPLA2 activity in a dose- and time-dependent manner (Kokotos et al., 2004). These inhibitors block the LPS-mediated release of arachidonic acid due to the stimulation of cPLA2 in murine P388 D1 macrophages with IC50 values of 8.0, 7.6, and 4.6 μM, respectively. These IC50 values are lower than the IC50 value reported for MAFP (25 μM). This strongly suggests that 2-oxoamides are more potent inhibitors of cPLA2 than MAFP. Anti-inflammatory and analgesic activities of 2-oxoamide have been tested in the rat paw carrageenan-mediated edema assay. Carrageenan-induced edema in rat paw can be prevented with 2-oxoamides (Kokotos et al., 2004). Effects of these inhibitors on brain cPLA2 remain unknown.

1,3-Disubstituted propan-2-ones (Fig. 4) were recently synthesized (Connolly et al., 2002). These compounds inhibit cPLA2 activity with an IC50 value of 0.03 μM in an in vitro assay. They are 10-fold more effective than 2-oxoamides and AACOCF3 in inhibiting cPLA2 activity. 1,3-Disubstituted propan-2-ones inhibit arachidonic acid production in HL60 cells with an IC50 value of 2.8 μM. These inhibitors have not been injected into intact animals or animal models of neurological disorders; therefore, nothing is known about their tolerance, half-lives, and toxicity. The ability of these inhibitors to cross the blood-brain barrier also remains unknown.

G. Choline Derivatives with a Long Aliphatic Chain

Recently, a new class of hydrophobic inhibitors that partition into the lipid bilayer and compete with monomers of the glycerophospholipid substrate was described previously (Burke et al., 1999). These compounds include 2-(2-benzyl-4-chlorophenoxy)ethyldimethyl-n-octadecyl-ammonium chloride and 2-(2-benzyl-4-chlorophenoxy)ethyldimethyl-n-octyl-ammonium bromide (Fig. 4). Both compounds inhibit cPLA2 activity in a competitive manner with IC50 values of 5 and 13 μM, respectively. The length of the N-alkyl chain plays an important role in the degree of inhibition. Shortening of the N-alkyl chain considerably decreases the percentage of inhibitor partitioned into the glycerophospholipid bilayer and increases the IC50 value. These effects may be due to diminished hydrophobic interaction between the shorter alkyl chain and the fatty acid tails of the glycerophospholipids making up the bilayer. In contrast, lengthening of the N-alkyl chain increases the percentage of inhibitor partitioned into the lipid bilayer and decreases the IC50 value. The synthesis of these inhibitors represents an important step in the development of potent in vivo cPLA2 inhibitors because these compounds inhibit enzymic activity at the interface and provide a weak interaction with the glycerophospholipid bilayer (Burke et al., 1999). Based on kinetic studies, the potential in vivo efficacy of these intracellular inhibitors can be more potent than other inhibitors such as AACOCF3. These inhibitors have not been used for in vivo studies so nothing is known about their tolerance, toxicity, and half-life. It is also not known whether these inhibitors can cross blood-brain barrier.

Fig. 4.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 4.

Chemical structures of PLA2 inhibitors. a, pyrrolidine-1; b, 2-(2-benzyl-4-chlorophenoxy)ethyldimethyl-n-octyl-ammonium chloride; c, 2-(2-benzyl-4-chlorophenoxy)ethyldimethyl-n-octadecyl-ammonium chloride; and d, AR-C70484XX [4-[3-4-(decyloxy)phenoxy)-2-oxopropoxy]benzoic acid; Connolly et al. (2002)].

H. Pyrrolidine-Based Inhibitors of Phospholipase A2

Pyrrolidine-containing compounds (Seno et al., 2000; Ghomashchi et al., 2001) markedly inhibit cPLA2-α in vitro and block arachidonate release in Ca2+ ionophore-stimulated non-neural cells (Seno et al., 2000). The structure of the most potent inhibitor, pyrrolidine-1, is shown in Fig. 4. In a fluorometric assay, pyrrolidine-1 inhibits cPLA2-α activity in a dose-dependent manner with an IC50 value of 0.18 μM. In a mixed-micelle assay system, this compound inhibits cPLA2-α activity with an IC50 value of 1.8 μM. This difference in the IC50 value may be due to the differences in interface concentrations of pyrrolidine-1 in the different assay systems. The treatment of Chinese hamster ovary cells with pyrrolidine-1 results in marked inhibition of A23187-induced arachidonic acid release with an IC50 value of 0.2 to 0.5 μM. The degree of inhibition approaches 100% with 4 to 10 μM pyrrolidine-1 (Ghomashchi et al., 2001). Similarly the treatment of Madin-Darby canine kidney cells with pyrrolidine-1 also results in marked inhibition of ATP-induced arachidonic acid release with an IC50 value of 0.8 μM. It must be stated here that pyrrolidine-1 also inhibits cPLA2-γ and iPLA2-β at a very high concentration but it does not inhibit sPLA2.

Pyrrophenone, a triphenylmethylthioether derivative of pyrrolidine, is a 39-fold more potent inhibitor (Ki 4.2 nM) of cPLA2 activity than pyrrolidine-1 (Ono et al., 2002). Pretreatment of rats with pyrrolidine dithiocarbamate, a powerful thiol antioxidant, protects against kainate (KA)-mediated neurotoxicity (Shin et al., 2004). In vitro studies indicate that pyrrophenone may be a potential therapeutic agent for inflammatory diseases (Ono et al., 2002). Pyrrolidine-containing PLA2 inhibitors have not been injected in animal models of neurological disorders so their half-life and side effects remain unknown.

I. Antimalarial Drugs

All isoforms of bovine brain PLA2 are strongly inhibited by antimalarial drugs in a dose-dependent manner with rank order potency of chloroquine > quinacrine > hydroxychloroquine > quinine (Lu et al., 2001b). Chloroquine, quinacrine, hydroxychloroquine, and quinine inhibit bovine brain cPLA2 with IC50 values of 125, 200, 185, and 250 μM. It is suggested that among PLA2 isoforms, PlsEtn-PLA2 and cPLA2 may be associated with proximal events involved in the induction and maintenance of inflammatory processes after ischemic and traumatic brain injuries (Farooqui et al., 2004b) and sPLA2 may be involved in intensification (Han et al., 2003) of inflammation during later stages of the inflammatory reaction. At low concentrations (<50 μM), these inhibitors have no effect on the growth of neuron-enriched cultures from rat brain cortex, but at high concentrations (>1000 μM), these inhibitors are toxic.

J. Lithium Ion and Carbamazepine

Lithium ion is a mood stabilizer. It has been used for the treatment of bipolar disorders for almost a half-century (Corbella and Vieta, 2003). Lithium ion has a neuroprotective effect on brain tissue. Chronic lithium ion administration in rats result in 50% reduction in mRNA and protein levels of cPLA2 with no changes observed in iPLA2 and sPLA2 protein during these studies (Chang and Jones, 1998; Rintala et al., 1999). Lithium ion does not reduce phosphorylated cPLA2 protein. Thus, the decrease in brain cPLA2 enzyme activity induced by chronic lithium ion treatment is due to down-regulation of cPLA2 transcription (Weerasinghe et al., 2004). This down-regulation of cPLA2 transcription may be responsible for a selective reduction of arachidonic acid turnover compared with docosahexaenoic acid in rat brain phospholipids (Basselin et al., 2003). The labeling of Purkinje cell dendrites with cPLA2 and cyclooxygenase (COX-2) antibodies is inhibited by lithium ion, indicating the functional coupling at brain synapses between cPLA2 and COX-2 enzymes (Weerasinghe et al., 2004).

An anticonvulsant drug, carbamazepine (CBZ), has been used for the treatment of bipolar disorders for many years. Chronic administration of CBZ not only inhibits cPLA2 activity but also alters its protein and mRNA levels. In contrast, it did not affect iPLA2 and sPLA2 activities and protein levels (Ghelardoni et al., 2004). These effects are accompanied by a decrease in COX-2 activity and prostaglandin E2 levels in brain tissue, suggesting that CBZ blocks the cPLA2-mediated release of arachidonic acid and its conversion via COX-2 to prostaglandin E2 (Ghelardoni et al., 2004). The protein levels of other arachidonic acid-metabolizing enzymes such as 5-lipoxygenase and cytochrome P450 and their reaction product leukotriene B4 are not affected by CBZ. Thus, nonspecific PLA2 inhibitors, such as lithium ion and CBZ, may have beneficial effects, not only in neurological diseases (see below), but also in bipolar disorders and manic-depressive patients.

K. Vitamin E and Gangliosides

Vitamin E (α-tocopherol) is another nonspecific inhibitor of cPLA2 activity (Douglas et al., 1986). It modulates the production of arachidonic acid and eicosanoids (Tran et al., 1996). It inhibits bovine brain cPLA2 and iPLA2 activities in a dose- and time-dependent manner with IC50 values of 500 and 750 μM, respectively. Vitamin E crosses the blood-brain barrier and with time accumulates in brain (Pentland et al., 1992). It reduces lipid peroxidation and stabilizes neuronal membranes. Several reports indicate that ischemia is accompanied by an increase in PLA2 activity and a reduction in the levels of vitamin E and glutathione (Farooqui et al., 1994a). A deficiency of vitamin E and selenium in rats leads to a biphasic increase in iPLA2 activity in non-neural cells (Burgess and Kuo, 1996), once again supporting the view that PLA2 activity is regulated by vitamin E. Vitamin E may modulate activities of PLA2 isozymes by two mechanisms: first, by direct incorporation into substrate vesicles; and second, by stimulation of the activity of PLA2 isoforms by up-regulating the rate of their synthesis at the transcriptional or translation level (Tran et al., 1996).

GM1 and GM3 gangliosides also inhibit cPLA2 and PlsEtn-PLA2 activities in a dose-dependent manner (Yang et al., 1994a,b). With PlsEtn-PLA2 the IC50 values for GM1 and GM3 gangliosides were 150.0 and 75.0 μg/ml, respectively. The IC50 values for brain cPLA2 for GM1 and GM3 were 250.0 and 100.0 μg/ml, respectively. The mechanism of inhibition by gangliosides remains unknown. However, the orientation of N-acetylneuraminic acid residues in glycoconjugates is important for inhibitory activity (Yang et al., 1994b). Gangliosides not only stabilize neural membranes but also regulate calcium influx and enzyme activities associated with signal transduction.

L. Cytidine 5-Diphosphoamines

CDP-amines are key intermediates in the biosynthesis of phosphatidylcholine and phosphatidylethanolamine. CDP-choline (citicoline) decreases cPLA2 stimulation and hydroxyl radical generation after transient cerebral ischemia (Adibhatla and Hatcher, 2003). This process results in lowering of the concentration of free fatty acids in a dose- and time-dependent manner. CDP-choline protects neural membranes, not only by accelerating the re-synthesis of phospholipids but also by quenching free radicals generated by PLA2 isozymes (Rao et al., 2001; Adibhatla et al., 2002). A mixture of CDP-ethanolamine and CDP-choline may be more effective than CDP-choline alone (Murphy and Horrocks, 1993).

M. Long-Chain Polyunsaturated Fatty Acids

Long-chain polyunsaturated fatty acids are normal constituents of neural membrane phospholipids and products of the PLA2-catalyzed reaction. They include arachidonic acid (belonging to the n-6 class), eicosapentaenoic acid (EPA), and DHA (belonging to the n-3 class). Arachidonic acid is released by the action of cPLA2, and EPA and DHA are released by the action of PlsEtn-PLA2 on neural membrane phospholipids. In vitro, the addition of these fatty acids to the reaction mixture inhibits the PLA2-catalyzed reaction in a dose- and time-dependent manner. This inhibition can be reversed by the addition of bovine serum albumin. The in vivo effects of these fatty acids on brain metabolism are quite complex. Arachidonic acid is metabolized to prostaglandins, leukotrienes, and thromboxanes by cyclooxygenases and lipoxygenases. These metabolites can cause vasoconstriction and hence compromise blood flow and oxygen delivery to brain tissue. EPA competes with arachidonic acid at the cyclooxygenase level to produce the 3-series prostaglandins (PGE3, PGI3, and TXA3) or the 5-series leukotrienes (LTB5, LTC5, and LTD5). These metabolites are less active than the corresponding arachidonic acid-derived compounds (Anderle et al., 2004). For example, TXA3 is less active than TXA2 in aggregating platelets and constricting blood vessels (James et al., 2000; Calder and Grimble, 2002). DHA is not a substrate for cyclooxygenase. It inhibits cyclooxygenase activity and is metabolized to docosanoids (Fig. 5). Docosanoids include 10,17S-docosatrienes and 17S-resolvins (Hong et al., 2003; Marcheselli et al., 2003; Serhan et al., 2004). They not only antagonize the effects of arachidonic acid-generated metabolites but also act potently on leukocyte trafficking as well as down-regulating the expression of cytokines in glial cells (Hong et al., 2003; Marcheselli et al., 2003; Mukherjee et al., 2004; Serhan et al., 2004). The specific receptors for these bioactive lipid metabolites occur in neural and non-neural tissues. These receptors include resolvin D receptors (ResoDR1), resolvin E receptors (resoER1), and neuroprotectin D receptors (NPDR). Characterization of these receptors in brain tissue is in progress. The generation of docosanoids may be an internal protective mechanism for preventing brain damage (Horrocks and Farooqui, 2004; Mukherjee et al., 2004; Serhan et al., 2004).

Fig. 5.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 5.

Chemical structures of docosahexaenoic acid and its metabolites. a, docosahexaenoic acid; b, eicosapentaenoic acid; c, 10,17S-docosatriene; d, 16,17S-docosatriene; e, 7,16,17S-resolvin; and f, 4,5,17S-resolvin. These metabolites retard the actions of arachidonic acid and its eicosanoid metabolites.

N. Phospholipase A2 Antisense Oligonucleotides and Interfering RNA

Antisense oligonucleotides have been synthesized to inhibit isoforms of PLA2 (Locati et al., 1996; Yoo et al., 2001; Laktionova et al., 2004). These antisense oligonucleotides efficiently inhibit the activity of various isoforms of PLA2 and also block their expression. They may be used for the treatment of inflammation and oxidative stress in neurological disorders. It must be emphasized here that the in vivo effects of antisense oligonucleotides may not be predictable from in vitro studies, partly because of the potential for the activation of the immune system by nucleotides (Endres and von Schacky, 1996). Furthermore, for neurological disorders there may be additional problems, including the efficient delivery of antisense oligonucleotides to a specific brain region in vivo and the high cost in manufacturing large quantities of antisense oligonucleotides.

An important development for inhibiting the enzymic activity of PLA2 isoforms has been the discovery of the RNA interference (RNAi) technique. This technology takes advantage of the evolutionary adaptation of neural cells to silence a gene whose corresponding double-stranded RNA molecule is present in the cell (Hannon, 2002). Introduction of a synthesized double-stranded RNA specific for a PLA2 isoform gene may cause rapid and prolonged reduction of mRNA and protein expression of that PLA2 isoform in brain tissue. RNAi has been developed for iPLA2 in non-neural cells (Shinzawa and Tsujimoto, 2003). It inhibits protein expression and iPLA2 activity in a dose- and time-dependent manner. In vivo injections of RNAi have not been made in intact animals, so the therapeutic importance of RNAi remains unknown.

V. Phospholipase A2 Activity in Kainic Acid-Induced Neural Cell Injury

KA is a nondegradable analog of glutamate. It is 30- to 100-fold more potent than glutamate as a neuronal excitant. Systemic administration of KA in adult rats induces persistent seizures and seizure-mediated brain damage (Lerma, 1997; Ben-Ari and Cossart, 2000). KA produces selective degeneration of neurons, especially in striatal and hippocampal areas of brain after intraventricular and intracerebral injections (Ben-Ari and Cossart, 2000). Axons and nerve terminals are more resistant to the destructive effects of KA than the cell soma. The mechanisms underlying KA-induced neuronal damage are quite complex. These mechanisms not only include inhibition of mitochondrial function and increases in intracellular calcium ion, due to depolarization, but also free radical generation (Farooqui et al., 2001). Our nuclear microscopic studies have clearly indicated a steady increase in calcium ion at 1, 2, and 3 weeks after KA injection (Ong et al., 1999). This increase in calcium ion may be responsible for the stimulation of many calcium-dependent enzymes including isoforms of PLA2 (Sandhya et al., 1998), calpain II (Ong et al., 1997), and endonucleases during KA-induced neurotoxicity.

Systemic administration of KA into adult rats markedly increases cPLA2 immunoreactivity in neurons at 1 and 3 days after injection. The cPLA2 immunoreactivity is also increased in astrocytes at 1, 2, 4, and 11 weeks after KA injection (Sandhya et al., 1998). Collective evidence suggests that an increase of cPLA2, PlsEtn-PLA2, and sPLA2 activities in KA-induced toxicity may be involved in neurodegeneration, whereas the elevation of cPLA2 activity in astrocytes may be associated with gliosis (Sandhya et al., 1998). We have also observed a decrease in 4-HNE and glutathione levels in KA-mediated toxicity.

Our immunocytochemical studies are supported by neurochemical studies on neuronal cell cultures (Kim et al., 1995; Thwin et al., 2003; Farooqui et al., 2003a,b). Treatment of neuron-enriched cultures with KA results in the stimulation of cPLA2, PlsEtn-PLA2, and sPLA2 activities in a dose- and time-dependent manner (Fig. 6) (Farooqui et al., 2003a,b; Thwin et al., 2003). This stimulation of cPLA2 activity in neuron-enriched cultures can be blocked by a nonspecific cPLA2 inhibitor, quinacrine (Farooqui et al., 2003b), a PlsEtn-PLA2 inhibitor, BEL (Farooqui et al., 2003a), a sPLA2 inhibitor, 12-episclaradial (Thwin et al., 2003), or by a KA/AMPA receptor antagonist, 6-cyano-7-nitroquinoxaline-2,3-dione (Farooqui et al., 2003a). During KA neurotoxicity, the increased cPLA2 immunoreactivity is accompanied by an increase in the 4-HNE immunoreactivity and a decrease in glutathione immunoreactivity (Ong et al., 2000) in hippocampus, indicating alterations in cellular redox. The reduction in glutathione levels may be due either to the formation of a 4-HNE-glutathione adduct or to decrease in cysteine uptake caused by KA-mediated neurotoxicity. Both processes increase the vulnerability of neurons to oxidative stress. All of these events, along with the decrease in ATP levels, are closely associated with cell death (Farooqui et al., 2004b).

Fig. 6.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 6.

Effect of kainic acid on cPLA2 (□) and PlsEtn-PLA2 (▪) activities of neuronal cultures. Data modified from Farooqui et al. (2003a,b) and Thwin et al. (2003).

The molecular mechanism(s) by which KA stimulates cPLA2 activity are not fully understood. However, excessive stimulation of the KA receptors may produce a marked increase in the level of intracellular Ca2+ due to membrane depolarization. Overstimulation of KA receptors may also induce the release of proinflammatory cytokines. Ca2+ stimulates the enzymic activity through translocation to neural membranes, whereas cytokines may activate cPLA2 activity through phosphorylation.

Like KA, intracerebroventricular injections of carrageenan, a high molecular weight sulfated polygalactose, and lipopolysaccharide, a major glycolipid component of Gram-negative bacterial outer membranes, also produce an increase in cPLA2 activity and neuronal death in the hippocampus (Ong et al., 2003a; Rosenberger et al., 2004). Unlike KA, intracerebroventricular injections of carrageenan do not produce convulsions.

During KA-induced neurotoxicity, neuronal loss, as shown by a decrease in GluR1 staining, is visible in Nissl sections at 3 days after injection. In contrast, carrageenan injections do not produce a decrease in GluR1 staining until the 3rd day after injection, and cell death can be seen in Nissl sections only from 1 week after injection. Finally, during carrageenan neurotoxicity glial cell induction of cPLA2 occurs earlier than KA-mediated neurotoxicity (Ong et al., 2003a). Like KA-mediated neurotoxicity, carrageenan injections induce the expression of cyclooxygenase immunoreactivity (Sandhya et al., 1998; Ibuki et al., 2003). This response appears at 3 h and becomes most prominent at 6 h after carrageenan injection. Intrathecal administration of a cyclooxygenase-2 inhibitor at 2 h after carrageenan injection produces a prominent therapeutic effect on hyperalgesia. These studies once again support the coupling between PLA2 and cyclooxygenase enzymes. Similarly, intracerebroventricular injections of lipopolysaccharide produce elevations in cPLA2 activity but not in sPLA2 and iPLA2 activities (Rosenberger et al., 2004).

The molecular mechanism of neuronal loss in carrageenan- and lipopolysaccharide-mediated toxicity is not fully understood. These agents may produce neuroinflammation by activating microglial cells and releasing proinflammatory and cytotoxic factors such as interleukin-1, tumor necrosis factor-α, and nitric oxide (Serhan et al., 2004; Wang et al., 2004a). All these factors are known to up-regulate cPLA2 activity in a neurotoxin-mediated model of neurodegeneration and represent early steps in generation and maintenance of inflammatory processes and oxidative stress (Farooqui et al., 2002).

Our recent studies have indicated a significant increase in the number of 4-HNE-immunoreactive pyramidal neurons and in cPLA2 activity in the hippocampus at 1 day, 1 week, and 2 weeks after a single intracerebroventricular injection of 1 μl of 10 mM ferrous ammonium citrate, an agent that produces severe oxidative stress (Ong et al., 2005). Intense 4-HNE labeling is observed at 1 day postinjection. 4-HNE immunoreactivity is markedly decreased after 2 weeks postinjection. A significant increase in cPLA2 immunoreactivity is also observed in pyramidal neurons at 1 week postinjection. Despite the increased 4-HNE and cPLA2 labeling, no loss of neurons is observed. Electron microscopic studies showed that 4-HNE- or cPLA2-positive neurons have features of injured neurons, but they were still viable. The reduction of 4-HNE immunoreactivity in neurons at 2 weeks after oxidative injury, and the lack of cell loss at any of the time intervals, indicates that the hippocampal pyramidal neurons have a remarkable ability to recover from a single episode of severe oxidative injury (W.-Y. Ong et al., 2005, manuscript submitted for publication).

VI. Protection of Kainic Acid-Induced Neural Cell Injury by Phospholipase A2 Inhibitors

In our studies on KA-induced toxicity in rat brain hippocampal slices and primary neuronal cells in culture, activities of PLA2 isoforms and their immunoreactivities are markedly increased (Lu et al., 2001a,b; Farooqui et al., 2003a,b). Quinacrine and other inhibitors of cPLA2 activity, such as AACOCF3 and BEL, could block this increase in PLA2 activities and immunoreactivity (Fig. 7). Detailed investigations on the effect of quinacrine indicate that this drug not only blocks the cPLA2 enzymic activity but also inhibits the expression of its mRNA (Ong et al., 2003b). Thus, 3 days after KA injection into the right lateral ventricle of rats, cPLA2 mRNA levels were up-regulated 6- to 13-fold in the injected right hippocampus and 5- to 9-fold in the noninjected contralateral hippocampus. cPLA2 mRNA levels remained at the low basal level in phosphate-buffered saline-injected control rat hippocampus. One week later, cPLA2 mRNA was up-regulated 3- to 6-fold in the injected right hippocampus and 3- to 8-fold in the noninjected contralateral side. Administration of quinacrine prevented the KA-mediated increases in cPLA2 mRNA levels. In rat brain hippocampal slices, other antimalarial drugs, including chloroquine, hydroxychloroquine, and quinine, also produce a significant reduction of the increased cPLA2 immunoreactivity. This suggests that the increased cPLA2 immunoreactivity after KA-induced neuronal damage is regulated at the mRNA level and antimalarial drugs inhibit this increase (Ong et al., 2003b).

Similarly surfactin, a surfactant lipopeptide with antibiotic, antifungal, and anticoagulant properties, inhibits cPLA2 activity in a dose-and time-dependent manner. Our studies on KA-induced neurotoxicity in rat hippocampal slice cultures indicate that surfactin significantly reduces neurodegeneration, up-regulation of cPLA2 activity, and 4-HNE formation. This cPLA2 inhibitor can be used to prevent KA-induced neurotoxicity (Farooqui et al., 2004b). Vitamin E and ganglioside, the nonspecific inhibitors of cPLA2, protect neurons from KA-mediated neurotoxicity (Ortiz et al., 2001), supporting our proposal that PLA2 inhibitors act as neuroprotective agents in KA-mediated neurotoxicity.

Fig. 7.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 7.

Histograms showing the effect of 100 μM kainate alone or 100 μM kainate followed 3 h later by a PLA2 inhibitor. Slices were fixed and processed for immunocytochemistry 1 week after treatment. The y-axes indicate the percentage of a line traced along the stratum pyramidale of the CA fields of the hippocampus that contains immunoreactivity to the AMPA receptor subunit GluR1 (used here as a neuronal marker) (A) or cPLA2 (B). The results were analyzed with Student's t test. Asterisks indicate statistically significant differences between the kainate-treated group and the other groups (P < 0.05). n = 6 to 12 slices in each group. A significantly greater number of GluR1-positive neurons remain in slices treated with kainate and quinacrine or AACOCF3 than after kainate treatment alone (A). Treatment with kainate and quinacrine or AACOCF3 also resulted in less induction of cPLA2 immunoreactivity, compared with treatment with kainate alone (B). NORM, untreated slices; KA, kainate; KA + QUIN or AACOCF3 or SURF or BEL, slices treated with kainate, followed by quinacrine, AACOCF3, surfactin, or bromoenol lactone. Data modified from Lu et al. (2001b).

The molecular mechanism of action of PLA2 inhibitors in KA-induced neurotoxicity may include the following possibilities:

  1. PLA2 inhibitors may inhibit cPLA2 activity by binding to phospholipid substrates (Blackwell and Flower, 1983).

  2. PLA2 inhibitors may block cPLA2 activity by interacting with the active site (Lu et al., 2001b).

  3. PLA2 inhibitors, such as quinacrine, may inhibit cPLA2 by blocking the expression of the mRNA for cPLA2 in kainic acid-induced neurotoxicity, as indicated by our northern blot studies (Ong et al., 2003b).

  4. PLA2 inhibitors might modulate KA receptors by controlling the levels of eicosanoids. The latter are known to be involved in the regulation of hippocampal glutamate receptors (Chabot et al., 1998). PLA2 inhibitors can modulate the expression of cytokines, NF-κB, adhesion molecules, and brain proteases (Burgermeister et al., 1999). All these molecules are associated with the induction and maintenance of inflammatory processes after neuronal injury.

  5. PLA2 inhibitors may exert their neuroprotective effects by binding to an AP-1 site and suppressing or lowering transcription of genes for cPLA2 isozymes. In addition to their effect on cPLA2 activity and gene expression, PLA2 inhibitors, such as quinacrine, have also been shown to have anesthetic and anti-inflammatory effects. These effects may be important in preventing neuronal degeneration after KA-induced neurotoxicity. Collective evidence strongly suggests that cPLA2 inhibitors can be used as potential therapeutic agents for treatment of inflammation and oxidative stress in the KA-mediated model of neural cell injury.

VII. Phospholipase A2 Activity in Neurological Disorders

It is now well established that activities of PLA2 isoforms are up-regulated in acute neural trauma, ischemia, spinal cord injury, and head injury, and many neurodegenerative diseases (Fig. 8) with substantial inflammatory and oxidative components such as AD, PD, and MS (Kalyvas and David, 2004; Phillis and O'Regan, 2004; Sun et al., 2004). Many studies indicate that neuroinflammation in these neurological disorders is accompanied by the activation of astrocytes and microglia and the release of inflammatory cytokines. These cytokines propagate inflammation in turn by a number of mechanisms including the up-regulation of PLA2 isoforms, generation of eicosanoids and platelet-activating factor, stimulation of nitric-oxide synthase, and calpain activation (Farooqui and Horrocks, 1991, 1994; Farooqui et al., 2000b). In this context, among the various approaches under investigation as therapies for neurotrauma and neurodegenerative diseases, there is growing support for strategies that prevent inflammatory reactions during neurodegeneration.

A. Ischemia

Ischemic injury produces the stimulation of cPLA2 and PlsEtn-PLA2 in brain tissue (Edgar et al., 1982; Rordorf et al., 1991; Clemens et al., 1996). The stimulation of these enzymes results in the massive release of free fatty acids in brain, the Bazan effect (Farooqui et al., 1994a). The formation of free fatty acids, depletion of ATP, and alterations in ion homeostasis induce membrane dysfunction that may lead to cellular injury (Farooqui and Horrocks, 1991, 1994). The stimulation of PlsEtn-PLA2 may occupy a proximal position in the injury pathway, initiating neural cell injury, whereas cPLA2 may participate by hydrolyzing PtdCho and amplifying the injury process (Sapirstein and Bonventre, 2000; Farooqui et al., 2003a,b). The mechanisms of stimulation of isoforms of PLA2 in ischemic injury are not known. Covalent modification, such as phosphorylation, may be involved in the stimulation process (Edgar et al., 1982). An increased expression of cPLA2 mRNA and sPLA2 mRNA also occurs after transient forebrain ischemia (Lin et al., 2004; Sun et al., 2004). The role of cPLA2 in neuronal damage is strongly supported by studies on cPLA2 knockout mice (Bonventre et al., 1997; Sapirstein and Bonventre, 2000; Tabuchi et al., 2003). After transient middle cerebral artery occlusion, cPLA2 knockout mice develop smaller infarcts, less brain edema, and less neurological deficits than control mice, indicating a reduced susceptibility of cPLA2 knockout mice to ischemic neurodegeneration. Primary neural cell cultures prepared from cPLA2-deficient mice generate significantly smaller amounts of prostaglandins and leukotrienes (Uozumi and Shimizu, 2002). This suggests that the cPLA2 deletion contributes to a decrease in arachidonic acid supply to cyclooxygenase-2 and a resultant decrease in prostaglandins synthesis (Hong et al., 2001).

Fig. 8.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 8.

Involvement of PLA2 isoforms in neurological disorders.

Fig. 9.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 9.

Activities of cPLA2 (A) and PlsEtn-PLA2 (B) in different regions of brains from normal subjects (□) and AD patients (▪). Specific activity is expressed as picomoles per minute per milligram of protein. FC, frontal cortex; PC, parietal cortex; OC, occipital cortex; NB, nucleus basalis; HP, hippocampus; CC, corpus callosum. Data modified from Farooqui et al. (2003a,b).

B. Alzheimer's Disease

Activities of cPLA2 and PlsEtn-PLA2 are markedly higher in nucleus basalis and hippocampal regions of AD brain compared with age-matched control brains (Fig. 9) (Stephenson et al., 1996, 1999; Farooqui et al., 1997a, 2003a,b). A similar elevation in COX-2 activity is also found in hippocampal neurons of AD patients compared with age-matched control subjects (Sugaya et al., 2000). In an earlier study, a significantly lower Ca2+-dependent PLA2 activity was found in the parietal and frontal cortices in AD patients compared with age-matched, nondementia patients and control subjects (Gattaz et al., 1995). The method for determining enzymic activity is probably responsible for this discrepancy. Recent studies by this group (Gattaz et al., 2004) also indicate that iPLA2 activity of platelets from AD patients is markedly lower than that from control subjects. This suggests that more studies are required on the determination of activities of PLA2 isoforms not only in different regions of AD brain but also on various stages of AD (initial stage, moderately advanced stage, and advanced AD) in human populations. It should be kept in mind that the progress of neurodegeneration varies considerably during the development of AD. This process is strongly influenced by genetic factors. The level of the inflammatory response within the brain is a central factor influencing the neurodegenerative process through the release of inflammatory mediators that are supplied by PLA2 and COX-2 catalyzed reactions.

The elevation of phospholipid degradation metabolites, phosphomonoesters and phosphodiesters, in AD brain supports our finding of increased cPLA2 and PlsEtn-PLA2 activities. The increase in phosphomonoester and phosphodiester correlates with pathological markers of AD, such as neurofibrillary tangles and senile plaques (Pettegrew, 1989). Changes in brain membrane phospholipid and high-energy phosphate metabolism occur before any clinical manifestation of AD (Pettegrew et al., 1995), suggesting that abnormal signal transduction due to disturbed phospholipid metabolism may be an important feature of AD. The aldehydic product of arachidonic acid metabolism, 4-HNE, colocalizes with intraneuronal neurofibrillary tangles and may contribute to the cytoskeletal derangement found in AD. Alterations in phospholipid metabolism may be closely associated with the loss of synapses and neurons and the formation of senile plaques and neurofibrillary tangles in AD (Pettegrew, 1989; Farooqui and Horrocks, 1994). We believe that the synapse loss in AD may be specifically related to lower levels of plasmalogens (phospholipids containing vinyl ether bonds) in autopsy brain samples (Wells et al., 1995; Ginsberg et al., 1998; Guan et al., 1999; Han et al., 2001; Pettegrew et al., 2001), which indicate that plasmalogen deficiency may be related to the stimulation of PlsEtn-PLA2 (Farooqui et al., 1997a).

The exact cause of increased cPLA2 and PlsEtn-PLA2 activities in AD brain is not fully understood. However, there are several possibilities. Aβ, which accumulates in AD, activates cPLA2 activity (Lehtonen et al., 1996; Kanfer et al., 1998). Treatment of cortical cultures with Aβ stimulates cPLA2 activity. This stimulation is blocked not only by the cPLA2 inhibitor MAFP but also by cPLA2 antisense oligonucleotides (Kriem et al., 2005), strongly suggesting the involvement of cPLA2 in the pathogenesis of AD (Kriem et al., 2005). The second possibility is that the activation of astrocytes and microglia in AD may result in expression of cytokines, TNF-α, IL-1β, and IL-6, that are known to stimulate cPLA2 activity (Xu et al., 2003; Rosales-Corral et al., 2004). Finally, the Aβ-mediated influx of calcium ions promotes the activation and translocation of cPLA2 from cytosol to neural membranes. This results in breakdown of the membranes and abnormal signal transduction in AD (Farooqui and Horrocks, 1994). β-Amyloid-mediated changes in calcium ion signaling underlie not only its action on long-term potentiation but also on many calcium-dependent enzymes (Farooqui and Horrocks, 1994), which may render neurons vulnerable to neurodegenerative process. At this stage it is not known whether elevation of cPLA2 and PlsEtn-PLA2 activities is the cause or the consequence of neurodegenerative process and whether changes in activities of PLA2 isoforms are primary or secondary. Thus, more studies on the involvement of PLA2 isoforms in pathogenesis of AD are required.

C. Experimental Model of Parkinson's Disease

PD is characterized by a selective degeneration of the dopaminergic neurons of the substantia nigra. Free radicals and lipid peroxides also play an important role in the pathogenesis of PD. They are generated by the action of PLA2 isoforms and produce oxidative stress in dopaminergic neurons in the substantia nigra. Mice deficient in cPLA2 activity are resistant to MPTP neurotoxicity. This resistance strongly suggests that cPLA2 is closely associated with the pathophysiology of PD (Klivenyi et al., 1998). In brain, MPTP is converted to its toxic metabolite, 1-methyl-4-phenylpyridinium ion (MPP+), in the presence of monoamine oxidase B. MPP+ is actively taken up into nigrostriatal neurons where it inhibits mitochondrial oxidative phosphorylation, leading to neuronal cell death (Singer et al., 1987).

The involvement of cPLA2 in PD pathogenesis is supported by MPP+-mediated toxicity in GH3 cell cultures (Yoshinaga et al., 2000). MPP+-mediated neurodegeneration is accompanied by the stimulation of PLA2 and arachidonic acid release in GH3 cells. This release of arachidonic acid can be blocked by AACOCF3, a specific inhibitor of cPLA2. Once again, this finding suggests the involvement of cPLA2 in MPP+-mediated neurodegeneration. It is interesting to note that in the MPTP-induced model of parkinsonism, quinacrine protects dopaminergic neurons from neurodegeneration (Tariq et al., 2001). In this system quinacrine may act not only as a PLA2 inhibitor but also as a membrane stabilizer and antioxidant (Tariq et al., 2001; Turnbull et al., 2003).

D. Multiple Sclerosis and Experimental Autoimmune Encephalomyelitis

MS and EAE, an animal model for MS, are inflammatory demyelinating diseases of the brain and spinal cord that result in motor and sensory deficits. A marked increase in PLA2 activity occurs in brain tissue from MS patients (Huterer et al., 1995). cPLA2 is highly expressed in EAE lesions (Trigueros et al., 2003; Kalyvas and David, 2004; Phillis and O'Regan, 2004) and inhibition of this enzyme results in a remarkable reduction in the onset and progression of EAE. The reduction in EAE severity correlates with cPLA2 activity and its downstream mediators such as COX-2, the LTB4 receptor 2, and various chemokines and cytokines (Kalyvas and David, 2004). The induction and maintenance of neuroinflammation in MS patients may involve cPLA2. sPLA2 activity is markedly elevated in EAE and LPS-mediated neurotoxicity (Pinto et al., 2003). An extracellular sPLA2 inhibitor, N-derivatized phosphatidylethanolamine linked to a polymeric carrier, blocks central nervous system inflammation under both in vivo and in vitro conditions. These interesting observations suggest that more studies are required on the involvement of PLA2 isoforms in neurodegenerative processes in MS.

E. Prion Diseases

Prion diseases include scrapie, found in goats and sheep, bovine spongiform encephalopathy (mad cow disease) in cattle, and fetal familial insomnia, Creutzfeldt-Jakob disease (CJD), kuru, and Gerstmann-Sträussler-Scheinker syndrome in humans (Prusiner, 2001; Grossman et al., 2003). Neuronal loss, spongiform degeneration, and glial cell proliferation are the pathological hallmarks of prion diseases. Human prion protein (PrPc) contains 209 amino acids, a disulfide bridge between residues 179 and 214, a glycosylphosphatidylinositol (GPI) anchor, and two sites of nonobligatory N-linked glycosylation at amino acids 181 and 197 (DeArmond and Prusiner, 2003). In prion diseases, the soluble PrPc, which consists of an α-helix and random coil structures, is refolded into a β-pleated sheet, the insoluble protease resistant isoform, PrPsc (DeArmond and Prusiner, 2003). The accumulation of PrPsc, which occurs in the cytoplasm and in secondary lysosomes as well as in neuronal plasmalemma and synaptic regions, may be responsible for the loss of cognitive function in prion diseases (Lazarewicz et al., 1990; Jeffrey et al., 1992).

PrPsc and PrP106-126, a neurotoxic prion peptide, are known to stimulate NMDA receptors. This stimulation is blocked by MK-801, memantine, and flupirtine (Muller et al., 1993; Perovic et al., 1995; Peyrin et al., 1999). PrP106-126 peptide-mediated stimulation of the NMDA receptor is accompanied with the release of arachidonic acid in cerebellar granule neurons, suggesting the association of PLA2 isoforms in the pathogenesis of prion diseases (Stewart et al., 2001). The involvement of PLA2 in the pathogenesis of prion diseases is also supported by recent neuronal cell culture studies (Bate and Williams, 2004). In this model of prion disease, neuronal PLA2 is activated by GPI isolated from PrPc or PrPsc. The ability of GPI to activate PLA2 is lost by either removal of the acyl chains or cleavage of the phosphatidylinositol-glycan linkage and inhibited by a monoclonal antibody that recognizes phosphatidylinositol (Bate and Williams, 2004). Furthermore, the treatment of neuronal cultures with inositol monophosphate or sialic acid provides resistance to the toxic effects of prion neurotoxic peptides. These observations strongly implicate PLA2 in the pathogenesis of prion diseases. The involvement of PLA2 in prion-mediated diseases may be a primary event or a secondary effect of an abnormal signal transduction process related to inflammation and oxidative stress in degenerating neurons. However, it is interesting to note that quinacrine, an acridine-based PLA2 antagonist, inhibits PrPsc formation (IC50 300 nM), and can be used for the treatment of the human prion disease, CJD (Doh-ura et al., 2000; Korth et al., 2001; Love, 2001; Follette, 2003; May et al., 2003). The molecular mechanism involved in the inhibition of PrPsc formation by quinacrine remains unknown. However, it has been suggested that quinacrine blocks PrP106-126 formed channels (Farrelly et al., 2003). NMR spectroscopic studies indicate that the PLA2 inhibitor, quinacrine, binds to human prion protein at the Tyr-225, Tyr-226, and Gln-227 residues of helix α3 (Vogtherr et al., 2003). Similarly, other antimalarial drugs, such as chloroquine and the phenothiazine derivatives acepromazine, chlorpromazine, and promazine, also bind to prion protein between residues 121 to 230, suggesting that Tyr-225, Tyr-226, and Gln-227 residues are necessary for the binding of antimalarial drugs and phenothiazine derivatives (Vogtherr et al., 2003) to PrPc. It has also been reported that quinacrine acts an antioxidant and reduces the toxicity of PrP106-126 (Turnbull et al., 2003). This finding once again suggests that the release of arachidonic acid and the oxidative stress generated by altered arachidonic acid metabolism may play an important role in the pathogenesis of prion diseases (Guentchev et al., 2000; Milhavet et al., 2000).

F. Spinal Cord Injury

Spinal cord injury triggers a series of secondary neurochemical processes that result in apoptotic as well as necrotic cell death (Beattie et al., 2000). The neurochemical changes include a rise in glutamate and intracellular calcium, degradation of membrane phospholipids with generation of free fatty acids, diacylglycerols, eicosanoids, and lipid peroxides (Demediuk et al., 1985), and activation of phospholipases and lipases (Anderson et al., 1985; Taylor, 1988). During the 1st min of compression trauma to the spinal cord, 10% of the PlsEtn is lost with an overall loss of 18% found at 30 min after the compression injury (Horrocks et al., 1985). Similar results have been reported in another model of spinal cord injury in rabbits (Lukácová et al., 1996).

The loss of PlsEtn after compression and ischemic injuries can be explained by the stimulation of PlsEtn-PLA2 due to shear stress (Taylor, 1988). Stimulation of the PlsEtn-PLA2 may result in changes in membrane fluidity and permeability, resulting in increased calcium influx, impaired mitochondrial function, and the subsequent generation of ROS. Low levels of ROS act as second messengers and produce neurodegeneration by apoptosis whereas high levels of ROS produce irreversible damage to cellular components and cause cell death by necrosis (Denecker et al., 2001). Necrosis normally occurs at the core of the injury site, whereas neural cells, including oligodendroglia, undergo apoptosis several hours or days after injury in the surrounding area.

G. Head Injury

The neurochemical changes in head injury are accompanied by widespread neuronal depolarization, accumulation of glutamate in the extracellular space, and increased levels of arachidonic acid and eicosanoids as well as leukotrienes (McIntosh et al., 1998). The release of arachidonic acid and its metabolites is due to activation of cPLA2 (Shohami et al., 1989), as well as the phospholipase C/diacylglycerol-lipase pathway (Wei et al., 1982). The pathological consequences of alterations in phospholipid metabolism in neural trauma and neurodegenerative diseases may include release of glutamate, energy failure, stimulation of PLA2 and diacylglycerol lipase activities, free radical damage, and alteration of membrane fluidity and permeability. These changes may influence the pattern of membrane-bound enzymes, receptors, and ion channels (Farooqui and Horrocks, 1994).

Glutamate-mediated abnormal phospholipid metabolism may be a common mechanism involved in ischemia and neurodegenerative diseases such as AD. The acute neural trauma in ischemia is accompanied by a rapid release of glutamate and sustained calcium influx at the core of the injury site but not in the surrounding area. This process may result in necrotic cell death at the core of the injury, whereas penumbral region neurons may die by apoptosis. In contrast in AD, there is not an excessive release of glutamate. However, the number of NMDA and other glutamate receptors is decreased in neocortex and hippocampal regions compared with that in age-matched control subjects (Geddes et al., 1992). This decrease results in an alteration in calcium homeostasis and breakdown of membrane phospholipids due to the stimulation of cPLA2 and PlsEtn-PLA2 activities. In ischemic injury at the core site, the glutamate-mediated neurodegeneration may be rapid (days), because of the sudden lack of oxygen and a quick drop in ATP and alteration in ion homeostasis. In AD, oxygen, nutrients, and ATP are available to the nerve cell and ion homeostasis is maintained to a limited extent, so neural cells take a much longer time period (years) to die (Farooqui and Horrocks, 1994).

H. Epilepsy

Epileptic seizures are known to stimulate cPLA2 activity and its expression with the accumulation of arachidonic acid (Visioli et al., 1994; Kajiwara et al., 1996). Studies on the pentylenetetrazol-induced model of epilepsy in rat brain indicate significant elevations in sPLA2 activity in cortical, hippocampal, and cerebellar regions compared with the control group. The increase in sPLA2 activity is more pronounced in hippocampal and cortical regions than in the cerebellar region (Yegin et al., 2002). At present, information on activities of PLA2 isoforms is not available for different regions of epileptic human brain. More studies are needed on the involvement of PLA2 isoforms in the pathogenesis of epilepsy.

I. Schizophrenia and Depressive Disorders

Marked elevations are found in iPLA2 activity in brain tissue of schizophrenic patients (Hudson et al., 1996; Ross et al., 1997, 1999). This results in accelerated phospholipid metabolism in schizophrenia. Levels of phosphatidylcholine and phosphatidylethanolamine are decreased, whereas the levels of lysophosphatidylcholine are increased in brain, erythrocytes, platelets, and skin fibroblasts of patients with schizophrenia (Yao et al., 2000; Ross, 2003). The cause of the increased iPLA2 activity is not known. However, abnormalities in dopamine and retinoid metabolism along with alterations in cytokines in schizophrenic patients may be responsible for the stimulation of PLA2 isoforms (Laruelle et al., 1999; Farooqui et al., 2004a; Yao and Van Kammen, 2004). In contrast, cPLA2 activity is markedly decreased in schizophrenic patients (Ross, 2003). Abnormalities in the promoter region of the gene encoding cPLA2 may be responsible for the altered cPLA2 activity in schizophrenic subjects (Hudson et al., 1996; Rybakowski et al., 2003). A dimorphic site within the first intron of cPLA2 may also be responsible for abnormal PLA2 activity (Ross, 2003). The wide range of phenotypic variability compounds the problem of analyzing and identifying vulnerable genes in schizophrenia. Some clinical or biochemical markers might be used to diminish this problem. Thus, more studies are needed to understand the involvement of PLA2 polymorphism in the pathophysiology of schizophrenia (Junqueira et al., 2004). Cocaine users also have reduced cPLA2 activity in their brain tissue (Ross and Turenne, 2002). As in schizophrenia, we suggest that dopaminergic hyperactivity in cocaine users may be responsible for the decreased cPLA2 activity.

Studies on the association between cPLA2 and mood disorders indicate a potential involvement of cPLA2 polymorphism in mood disorders (Pae et al., 2004). It has been reported that genotyping and allele distributions in patients with major depressive disorders are significantly different from those of the control human subjects and a BanI polymorphism of the cPLA2 gene may be related to the pathogenesis of major depressive disorder (Pae et al., 2004).

VIII. Use of Phospholipase A2 Inhibitors for the Treatment of Neurological Disorders

In brain, quinacrine (Fig. 3) protects gerbil CA1 hippocampal pyramidal cells during 5 min of forebrain ischemia (Estevez and Phillis, 1997). After intravascular injections, quinacrine appears in monkey brain after 24 h (Dubin et al., 1982), indicating that this inhibitor is metabolically stable and can cross the blood-brain barrier. It is localized in neurons. It has also been administered (5 mg/kg) to rats that underwent 2 h of middle cerebral artery occlusion (Estevez and Phillis, 1997). The administration of quinacrine results in a marked reduction in neurological deficits after 24 h of reperfusion. Importantly, the effects of quinacrine persist even after 7 days. These findings are supported by biochemical and histopathological analysis that indicate a significant decrease in infarct size in quinacrine-treated rats compared with saline-treated controls. Based on these studies (Estevez and Phillis, 1997), it has been proposed that PLA2 inhibitors have cerebroprotective effects in focal as well as global models of cerebral ischemia. In organotypic hippocampal cultures oxygen/glucose deprivation produces a 2-fold increase in PLA2 activity with significant cell death. This increase in PLA2 can be blocked by AACOCF3 in a dose- and time-dependent manner (Arai et al., 2001). sPLA2 and iPLA2 inhibitors were ineffective in blocking cell death. In an attempt to evaluate the contribution of PLA2 isoforms to the release of free fatty acids, rat cerebral cortex was superfused with inhibitors of PLA2 activity. AACOCF3 markedly inhibited the efflux of arachidonic, docosahexaenoic, linoleic, palmitic, and oleic acids from the ischemic/reperfused rat cerebral cortex (Phillis and O'Regan, 2004). Exposure to the sPLA2 and iPLA2 inhibitors has minimal effect on the efflux of free fatty acids. These observations strongly suggest that cPLA2 plays an important role in ischemic injury and that PLA2 inhibitors can be used for the treatment of ischemic injury.

In addition, quinacrine has been proposed as a therapeutic agent for prion diseases (Korth et al., 2001). Quinacrine blocks prion protein peptide (PrP106-126)-mediated caspase-3 activation, supporting the involvement of cPLA2 in apoptotic cell death (Stewart et al., 2001). Quinacrine and a combination of quinacrine with chlorpromazine, a phenothiazine derivative, were used for the treatment of CJD using compassionate use as a justification (Love, 2001; Follette, 2003; Kobayashi et al., 2003). During quinacrine treatment, the initial responses of CJD patients have been positive, but within days of starting treatment, patients returned back to their previous states, indicating a transient recovery (Love, 2001; Follette, 2003; Kobayashi et al., 2003). The reason for the transient effect of quinacrine on CJD patients remains unknown. The transient effect may be due to the advanced stage of CJD. If quinacrine treatment could be started at the onset of CJD, patients would probably respond to this drug in a positive manner (Follette, 2003; Kobayashi et al., 2003). PLA2 activities were not determined during these studies. Thus, no comments can be made about the levels of arachidonic acid and its metabolites, inflammatory reactions, and oxidative stress that occur in prion-mediated neurodegeneration in CJD. More studies are required on the involvement of PLA2 isoforms and generation of proinflammatory mediators in the pathogenesis of prion diseases in animal and cell culture models.

Recent in vitro studies on prion-infected cell lines ScN2a, SMB, and ScGT1 also indicate that daily treatment of these cells with CDP, aristolochic acid, BEL, and AACOCF3 causes a significant decrease in protease-resistant prion protein compared with untreated control cells. The treatment with PLA2 inhibitors decreases protease-resistant prion protein but also reduces prostaglandin E2 levels. This observation strongly suggests that PLA2 activity may be closely related to the pathogenesis of prion diseases (Bate et al., 2004). Furthermore, corticosteroids that induce the formation of lipocortins (annexins), a family of PLA2 inhibitory proteins (Kaetzel and Dedman, 1995), also reduce the content of protease-resistant prion protein in prion-infected cell lines. This finding again supports an involvement of PLA2 in prion diseases (Bate et al., 2004). However, the use of glucocorticoids in prevention of prion diseases should be treated with caution because the chronic administration of glucocorticoids is known to produce neuronal atrophy (Abraham et al., 2001).

PAF, which is generated by the acetylation of lysophosphatidylcholine, another product of PLA2-catalyzed reactions, increases the generation of protease-resistant prion protein, and PAF antagonists block it. The mechanism by which PAF antagonists inhibit the formation of protease-resistant prion protein remains unknown. However, PAF antagonists and PLA2 inhibitors may act by altering intracellular trafficking of cellular prion protein. These observations suggest the pivotal role of PLA2 and PAF in modulating formation of protease-resistant prion protein, an agent that is suggested to be the main cause of prion diseases (Bate et al., 2004).

Vitamin E protects from neuronal damage induced by cerebral ischemia by inhibiting apoptosis in hippocampal neurons (Tagami et al., 1999; Zhang et al., 2004). This finding suggests that vitamin E reacts with the free radicals and prevents neuronal apoptosis produced by cerebral ischemia and reperfusion. Vitamin E also protects neurons from a toxic concentration of sodium nitroprusside, a nitric oxide donor, in a dose-dependent manner indicating that this vitamin protects brain tissue by inhibiting free radical generation and oxidative stress. Double-blind human trials of vitamin E have been performed. It slows the progression of Alzheimer's disease (Sano et al., 1997) or has only symptomatic effects with no alteration of the progression of AD (Sano et al. 1997). In cat spinal cord, preloading with vitamin E and selenium promotes recovery after spinal cord injury (Anderson et al., 1985). Collective evidence suggests that vitamin E may be useful for the treatment of inflammation and oxidative stress in acute neural trauma and neurodegenerative diseases. Negative results in human trials are probably due to the lack of a range of redox inhibitors.

CDP-choline (citicoline) inhibits cPLA2 activity and lowers the concentration of free fatty acids in a dose- and time-dependent manner (Adibhatla et al., 2002). This compound is an intermediate in PtdCho biosynthesis that has been used for the treatment of ischemic and head injuries (Andersen et al., 1999; Dempsey and Rao, 2003). It not only restores the concentration of PtdCho after ischemic injury by increasing PtdCho synthesis from diacylglycerol but also blocks the activation of cPLA2 activity (Adibhatla et al., 2002). The decrease in cPLA2 activity may lead to a reduction in levels of arachidonic acid and reactive oxygen species, with stabilization of neural membranes. CDP-choline also protects cerebellar granule neurons from glutamate-mediated neurotoxicity (Mir et al., 2003), suggesting that CDP-choline may protect neurons from excitotoxicity. CDP-choline has been used in phase III clinical trials for stroke and is being evaluated for the treatment of AD and PD. It also improves the verbal memory of aged human subjects. These observations suggest that this cPLA2 inhibitor can be used for treating acute neural trauma as well as neurodegenerative diseases.

Neurotrophic effects of gangliosides have been demonstrated in AD, PD, spinal cord injury, and stroke (Geisler et al., 1991; Svennerholm, 1994). The mechanism underlying the ganglioside action is not fully understood. However, gangliosides are known to inhibit neurotransmitter release and cPLA2 and PlsEtn-PLA2 activities. Gangliosides also rescue neuronal cultures from death after neurotrophic factor deprivation (Ferrari et al., 1993). NMR studies indicate that GM1 ganglioside binds tightly with β-amyloid peptide and inhibits the α-helix to β-sheet conformational change in β-amyloid peptide (Mandal and Pettegrew, 2004). The interaction with GM1 ganglioside may release the inhibition of PLA2 isoforms by GM1 and β-amyloid, resulting in delay of neurodegeneration in AD. In ischemic brain, gangliosides protect neural cells by scavenging free radicals generated during reperfusion (Fighera et al., 2004).

In the MPP+-mediated cell culture model of PD (Yoshinaga et al., 2000), neurodegeneration is accompanied by the stimulation of PLA2 and arachidonic acid release in GH3 cells. This release of arachidonic acid can be blocked by arachidonyl trifluoromethyl ketone, a potent inhibitor of cPLA2, suggesting the involvement of cPLA2-mediated oxidative stress in MPP+-mediated neurodegeneration. Similarly in the MPTP-induced model of parkinsonism, quinacrine protects dopaminergic neurons from neurodegeneration (Tariq et al., 2001). In this system quinacrine may act not only as a PLA2 inhibitor but also as a membrane stabilizer and antioxidant (Tariq et al., 2001; Turnbull et al., 2003).

Polyunsaturated fatty acids of the n-3 series have many beneficial effects in the central nervous system (Farooqui and Horrocks, 2004a). Thus, EPA prevents LPS-mediated TNF-α expression by preventing NF-κB activation and protects rat hippocampus from LPS-mediated neurotoxicity (Lonergan et al., 2004; Zhao et al., 2004). In C6 glioma cells, EPA modulates myelin proteolipid gene expression (Salvati et al., 2004). This fatty acid is used for the treatment of schizophrenia (Peet and Ryles, 2001; Horrobin, 2003).

Chronic preadministration of DHA prevents β-amyloid-induced impairment of an avoidance ability-related memory function in a rat model of AD (Hashimoto et al., 2002) and protects mice from synaptic loss and dendritic pathological changes in another model of AD (Calon et al., 2004). Thus, DHA is beneficial in preventing the learning deficiencies in these AD models. DHA also affects amyloid precursor protein processing by inhibiting α- and β-secretase activities (de Wilde et al., 2003; Walsh and Selkoe, 2004). Supplements of DHA produce a neuroprotective effect on β-amyloid deposition without significant toxic effects. DHA reverses the age-related impairment in LTP and depolarization-induced glutamate release. It also inhibits the production of TNF-α, interleukin-1, and interleukin-6. DHA protects the brain against ischemic and excitotoxic damage in rats (Gamoh et al., 1999; Terano et al., 1999). DHA may act as an antioxidant (Hossain et al., 1998). DHA induces antioxidant defenses by enhancing cerebral activities of catalase, glutathione peroxidase, and levels of glutathione (Hossain et al., 1999). Thus, EPA and DHA exert their neuroprotective effects by modulating cytokines, inflammation, and oxidative stress.

Treatment of Zellweger syndrome patients with purified DHA partially improves visual function, increases levels of plasmalogens, and reduces levels of saturated very long-chain fatty acids (Hossain et al., 1999). The level of DHA is also low in patients with multiple sclerosis. Fish oil supplements with vitamins improve the clinical outcome in MS patients (Nordvik et al., 2000). Collective evidence from many studies indicates that DHA supplementation restores signal transduction processes associated with behavioral deficits, learning activity in Alzheimer disease, schizophrenia, depression, hyperactivity, stroke, and peroxisomal disorders (Farooqui and Horrocks, 2004a).

All of the cPLA2 inhibitors used in these studies are nonspecific. Thus, the design and synthesis of specific cPLA2 inhibitors is urgently needed to make progress in this important area of research. The reaction catalyzed by cPLA2 is a rate-limiting step for the generation of eicosanoids, lysophospholipids, and platelet-activating factor. High levels of these metabolites are responsible for oxidative stress, inflammation, and neuronal death at the injury site. Potent PLA2 inhibitors can effectively block the above events associated with neurodegenerative processes and rescue neural cells from cell death. It is hoped that the synthesis of a new generation of cPLA2 inhibitors would have regional specificity. The inhibitors should be able to reach the injury site where neural cells are under oxidative stress and where neurodegenerative processes are taking place.

IX. Prevention of Pain by Phospholipase A2 Inhibitors

Proinflammatory cytokines released at the site of neural trauma and nerve injury may be involved in sensitization of nociceptors leading to hyperalgesia (Walters, 1994). Injections of carrageenan into the paw or face have been widely used as a model to induce pain sensitization (Ng and Ong, 2001). We have recently studied the effect of intracerebroventricular injections of a sPLA2 inhibitor, 12-epi-scalaradial, a cPLA2 inhibitor, AACOCF3, and a iPLA2 inhibitor, bromoenol lactone, on the development of allodynia after facial carrageenan injections in two strains of mice (Yeo et al., 2004) (Fig. 10). C57BL/6J (B6) mice show an increase in allodynia from 8 h to 3 days after facial carrageenan injection. On the other hand, the BALB/c strain did not show an increase in allodynia at any time point. In both B6 and BALB/c mice, all PLA2 inhibitors significantly reduced responses to von Frey hair stimulation at 8 h and 1 day, but at 3 days only the sPLA2 inhibitor had an effect. Because BALB/c mice do not show increases in allodynia after carrageenan injection, the reduction in responses seen with PLA2 inhibitors actually means that these inhibitors produce a loss of normal sensitivity to von Frey hair stimulation. The effects of PLA2 inhibitors are unlikely to be due simply to inhibition of arachidonic acid generation, because intracerebroventricular injection of arachidonic acid also had an antinociceptive effect (Yeo et al., 2004). It is proposed that lysophosphatidylcholine mediates pain transmission in the central nervous system. The pronounced and long-lasting antinociceptive effect of 12-epi-scalaradial is consistent with our recent finding that sPLA2 induces exocytosis and neurotransmitter release in neurons and supports a key role of central nervous system sPLA2 in synaptic and pain transmission. These results suggest that PLA2 isoforms play an important role not only in pain transmission but also in nonpainful, touch, or pressure sensation. Our studies on the antinociceptive effect of PLA2 inhibitors are supported by recent studies on intrathecal injections of MAFP and AACOCF3 in rats. MAFP has a significant antinociceptive effect in the rat formalin test (Ates et al., 2003). Based on our studies, we suggest that PLA2 inhibitors may reduce allodynia and hyperalgesia in inflammation-mediated central pain.

X. Perspective and Direction for Future Studies

PLA2 isoforms along with cyclooxygenases have emerged as major players in modulating inflammation and oxidative stress in brain tissue. Elucidation of the mechanism of action of PLA2 inhibitors in vivo is a critical area of research because of the potential pharmacological benefits of these compounds as therapeutic agents for the treatment of inflammation and oxidative stress in neurotrauma and neurodegenerative diseases (Farooqui et al., 1999). Although several inhibitors of PLA2 activity have been reported in the literature (Farooqui et al., 1999; Cummings et al., 2000; Miele, 2003b), little information is available on the mechanism of their action. Different mechanisms of action are possible; e.g., an inhibitor can produce alterations in enzymic activity by perturbing the physicochemical properties of phospholipid bilayers. A PLA2 inhibitor can directly interact with the active site of an isoform, as AACOCF3, MAFP, and BEL, or it can act on an allosteric site on the enzyme molecule to bring about changes in enzymic activity. An inhibitor may also possess a detergent-like structure that can induce nonspecific changes in membrane properties in vivo through the interaction of its amphiphilic groups with other membrane components to produce changes in enzymic activity.

Fig. 10.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 10.

Total responses in mice treated with dimethyl sulfoxide (DMSO), PLA2 inhibitors or arachidonic acid, and carrageenan at 8 h after injection. The y-axes indicate total scratching/withdrawal/attack/escape responses in mice that received intracerebroventricular injection of DMSO (5 μl, vehicle control), PLA2 inhibitors (0.01 μmol in DMSO), or arachidonic acid (0.01 μmol in DMSO) and a facial injection of carrageenan, 8 h after injections. Data were analyzed by one-way analysis of variance with Bonferroni's multiple comparison post hoc test. *, statistically significant differences (P < 0.05). n = 4 mice in each group. Both B6 mice (A) and BALB/c mice (B) showed significantly fewer total responses to von Frey hair stimulation of the face after injection with inhibitors to each of the three isoforms of PLA2 or arachidonic acid compared with DMSO. D+C, S+C, A+C, B+C, AA+C indicate DMSO, 12-epi-scalaradial, AACOCF3, BEL, or arachidonic acid, plus carrageenan, respectively. Data modified from Yeo et al. (2004).

Inhibitors of cPLA2 modulate the expression of cytokines, growth factors, NF-κB, and adhesion molecules and thus can block endogenous inflammatory reactions. Some nonspecific inhibitors of PLA2 have been used for the treatment of ischemia, spinal cord injury, and AD in animal models and in humans. It is clear from the above discussion that novel, potent, and specific inhibitors of PLA2 are now emerging as potential therapeutic agents for neuroinflammation and oxidative stress in cell culture systems, but they have not yet been used in animal models of neurological disorders.

In brain, PLA2 occurs in multiple forms. Thus, specific PLA2 inhibitors must be designed for individual PLA2 isozymes to define their roles in brain metabolism. The design of PLA2 inhibitors should be focused on our rapidly emerging understanding of the role of signal transduction pathways in neurological disorders. Because PLA2-catalyzed reactions are the rate-limiting steps for the production of prostaglandins, leukotrienes, and thromboxanes, the identification of PLA2-coupled receptors and their endogenous regulatory pathways that mediate proliferative, metabolic, and inflammatory signals can provide better targets for designing PLA2 inhibitors. Specificity, selectivity, harmlessness, and the ability to cross the blood-brain barrier are important qualities of a PLA2 inhibitor as a potential therapeutic agent for neurological disorders.

At this time, it is quite difficult to predict the potential side effects of the chronic use of cell-permeable, specific, or nonspecific inhibitors of PLA2. Hence, studies on the availability of specific, nontoxic potent inhibitors with greater blood-brain barrier permeability in animal models of neurodegenerative diseases are urgently needed. Surely, PLA2 isoforms from different sources will have different inhibitor sensitivities and reactivity in vivo. Nevertheless, pharmacological studies in animal models of acute trauma and neurodegenerative diseases will provide directions that should be taken to develop better PLA2 inhibitors for targeting isoforms of brain PLA2 activities.

In recent years advanced molecular biology procedures have been used in a number of studies to overcome some of the problems associated with the specificity of chemical inhibitors of PLA2. For example, antisense oligonucleotides that inhibit specific cPLA2 or iPLA2 have been developed. Transgenic mice that are deficient in or overexpress cPLA2 isoforms are now available (Bonventre et al., 1997). Overexpression of cPLA2 allows one to study the effect of increased cPLA2 activity, whereas a deficiency of cPLA2 can be helpful in studying the consequences of reduced cPLA2 activity on cellular metabolism in normal and diseased cells. RNAi for iPLA2 has also been developed. Transfection studies with RNAi of iPLA2 indicate that the levels of iPLA2 protein and iPLA2 activity are decreased in a dose-dependent manner in transfected non-neural cells (Shinzawa and Tsujimoto, 2003). Such studies are needed in neuronal cell cultures and animal models to understand the neurophysiological importance of the RNAi technique. We propose that a comparison of activities of PLA2 isoforms and intensity of signal transduction process between normal and genetically manipulated mice may provide further insight into the role of PLA2 isoforms, their ligands, and their lipid mediators in neurodegenerative processes.

Thus, the development of specific inhibitors for different PLA2 isoforms should be an important goal for future research on brain PLA2 activities. The chemical approach, together with molecular biological procedures such as RNAi and alterations in signal transduction processes in knockout mice, may provide the important information needed to develop specific PLA2 inhibitors that can be used to retard oxidative stress and inflammatory reactions during neurodegeneration in neurological disorders.

Acknowledgments

We thank Dr. Tahira Farooqui, Department of Entomology, The Ohio State University, Columbus, OH, for useful discussion and Siraj A Farooqui for providing figures and help with the preparation of this manuscript.

Footnotes

  • ↵1 Abbreviations: PLA2, phospholipase A2; PAF, platelet activating factor; AD, Alzheimer's disease; PD, Parkinson's disease; MS, multiple sclerosis; 4-HNE, 4-hydroxynonenal; sPLA2, secretory phospholipase A2; cPLA2, cytosolic phospholipase A2; PlsEtnPLA2, plasmalogen-selective phospholipase A2; iPLA2, calcium-independent phospholipase A2; PKC, protein kinase C; PtdCho, phosphatidylcholine; LTP, long-term potentiation; DHA, docosahexaenoic acid; IL, interleukin; TNF, tumor necrosis factor; AP-1, activator protein-1; NF-κB, nuclear factor-κB; ROS, reactive oxygen species; MPTP, 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine; PG, prostaglandin; NMR, nuclear magnetic resonance; AACOCF3, arachidonyl trifluoromethyl ketone; Aβ, β-amyloid peptide; MeHg2+, methylmercury; MAPF, methyl arachidonyl fluorophosphonate; BEL, bromoenol lactone; AMPA, α-amino-3-hydroxy-5-methylisoxazole-4-propionate; LPS, lipopolysaccharide; KA, kainate; COX-2, cyclooxygenase-2; CBZ, carbamazepine; CDP, cytidine 5-diphospho; EPA, eicosapentaenoic acid; TX, thromboxane; LT, leukotriene; RNAi, RNA interference; MMP+, 1-methyl-4-phenylpyridinium ion; MS, multiple sclerosis; EAE, experimental autoimmune encephalomyelitis; CJD, Creutzfeldt-Jakob disease; PrP, prion disease protein; GPI, glycosylphosphatidylinositol; NMDA, N-methyl-d-aspartate; MK-801, 5H-dibenzo[a,d]cyclohepten-5,10-imine (dizocilpine maleate).

  • Article, publication date, and citation information can be found at http://pharmrev.aspetjournals.org.

  • doi:10.1124/pr.58.3.7.

  • The American Society for Pharmacology and Experimental Therapeutics

References

  1. ↵
    Abraham IM, Harkany T, Horvath KM, and Luiten PG (2001) Action of glucocorticoids on survival of nerve cells: promoting neurodegeneration or neuroprotection? J Neuroendocrinol 13: 749-760.
    OpenUrlCrossRefPubMed
  2. ↵
    Ackermann EJ and Dennis EA (1995) Mammalian calcium-independent phospholipase A2. Biochim Biophys Acta 1259: 125-136.
    OpenUrlPubMed
  3. ↵
    Adibhatla RM and Hatcher JF (2003) Citicoline decreases phospholipase A2 stimulation and hydroxyl radical generation in transient cerebral ischemia. J Neurosci Res 73: 308-315.
    OpenUrlCrossRefPubMed
  4. ↵
    Adibhatla RM, Hatcher JF, and Dempsey RJ (2002) Citicoline: neuroprotective mechanisms in cerebral ischemia. J Neurochem 80: 12-23.
    OpenUrlCrossRefPubMed
  5. ↵
    Adibhatla RM, Hatcher JF, and Dempsey RJ (2003) Phospholipase A2, hydroxyl radicals, and lipid peroxidation in transient cerebral ischemia. Antioxid Redox Signal 5: 647-654.
    OpenUrlCrossRefPubMed
  6. ↵
    Aitdafoun M, Mounier C, Heymans F, Binisti C, Bon C, and Godfroid JJ (1996) 4-Alkoxybenzamidines as new potent phospholipase A2 inhibitors. Biochem Pharmacol 51: 737-742.
    OpenUrlCrossRefPubMed
  7. ↵
    Ajmone-Cat MA, Nicolini A, and Minghetti L (2003) Prolonged exposure of microglia to lipopolysaccharide modifies the intracellular signaling pathways and selectively promotes prostaglandin E-2 synthesis. J Neurochem 87: 1193-1203.
    OpenUrlCrossRefPubMed
  8. ↵
    Akiyama N, Hatori Y, Takashiro Y, Hirabayashi T, Saito T, and Murayama T (2004) Nerve growth factor-induced up-regulation of cytosolic phospholipase A2α level in rat PC12 cells. Neurosci Lett 365: 218-222.
    OpenUrlCrossRefPubMed
  9. ↵
    Amandi-Burgermeister E, Tibes U, Kaiser BM, Friebe WG, and Scheuer WV (1997) Suppression of cytokine synthesis, integrin expression and chronic inflammation by inhibitors of cytosolic phospholipase A2. Eur J Pharmacol 326: 237-250.
    OpenUrlCrossRefPubMed
  10. ↵
    Anderle P, Farmer P, Berger A, and Roberts MA (2004) Nutrigenomic approach to understanding the mechanisms by which dietary long-chain fatty acids induce gene signals and control mechanisms involved in carcinogenesis. Nutrition 20: 103-108.
    OpenUrlCrossRefPubMed
  11. ↵
    Andersen M, Overgaard K, Meden P, and Boysen G (1999) Effects of citicoline combined with thrombolytic therapy in a rat embolic stroke model. Stroke 30: 1464-1470.
    OpenUrlAbstract/FREE Full Text
  12. ↵
    Anderson DK, Saunders RD, Demediuk P, Dugan LL, Braughler JM, Hall ED, Means ED, and Horrocks LA (1985) Lipid hydrolysis and peroxidation in injured spinal cord: Partial protection with methylprednisolone or vitamin E and selenium. Cent Nerv Syst Trauma 2: 257-267.
    OpenUrlPubMed
  13. ↵
    Arai K, Ikegaya Y, Nakatani Y, Kudo I, Nishiyama N, and Matsuki N (2001) Phospholipase A2 mediates ischemic injury in the hippocampus: a regional difference of neuronal vulnerability. Eur J Neurosci 13: 2319-2323.
    OpenUrlCrossRefPubMed
  14. ↵
    Asai K, Hirabayashi T, Houjou T, Uozumi N, Taguchi R, and Shimizu T (2003) Human group IVC phospholipase A2 (cPLA2γ)—roles in the membrane remodeling and activation induced by oxidative stress. J Biol Chem 278: 8809-8814.
    OpenUrlAbstract/FREE Full Text
  15. ↵
    Ates M, Hamza M, Seidel K, Kotalla CE, Ledent C, and Guhring H (2003) Intrathecally applied flurbiprofen produces an endocannabinoid-dependent antinociception in the rat formalin test. Eur J Neurosci 17: 597-604.
    OpenUrlCrossRefPubMed
  16. ↵
    Atsumi G, Murakami M, Kojima K, Hadano A, Tajima M, and Kudo I (2000) Distinct roles of two intracellular phospholipase A2s in fatty acid release in the cell death pathway: proteolytic fragment of type IVA cytosolic phospholipase A2α inhibits stimulus-induced arachidonate release, whereas that of type VI Ca2+-independent phospholipase A2 augments spontaneous fatty acid release. J Biol Chem 275: 18248-18258.
    OpenUrlAbstract/FREE Full Text
  17. ↵
    Atsumi G, Tajima M, Hadano A, Nakatani Y, Murakami M, and Kudo I (1998) Fas-induced arachidonic acid release is mediated by Ca2+-independent phospholipase A2 but not cytosolic phospholipase A2 which undergoes proteolytic inactivation. J Biol Chem 273: 13870-13877.
    OpenUrlAbstract/FREE Full Text
  18. ↵
    Balboa MA, Varela-Nieto I, Lucas KK, and Dennis EA (2002) Expression and function of phospholipase A2 in brain. FEBS Lett 531: 12-17.
    OpenUrlCrossRefPubMed
  19. ↵
    Basselin M, Chang L, Seemann R, Bell JM, and Rapoport SI (2003) Chronic lithium administration potentiates brain arachidonic acid signaling at rest and during cholinergic activation in awake rats. J Neurochem 85: 1553-1562.
    OpenUrlCrossRefPubMed
  20. ↵
    Bate C, Reid S, and Williams A (2004) Phospholipase A2 inhibitors or platelet-activating factor antagonists prevent prion replication. J Biol Chem 279: 36405-36411.
    OpenUrlAbstract/FREE Full Text
  21. ↵
    Bate C and Williams A (2004) Role of glycosylphosphatidylinositols in the activation of phospholipase A2 and the neurotoxicity of prions. J Gen Virol 85: 3797-3804.
    OpenUrlAbstract/FREE Full Text
  22. Bayón Y, Hernández M, Alonso A, Nunez L, Garcia-Sancho J, Leslie C, Crespo MS, and Nieto ML (1997) Cytosolic phospholipase A2 is coupled to muscarinic receptors in the human astrocytoma cell line 1321N1: characterization of the transducing mechanism. Biochem J 323: 281-287.
    OpenUrlAbstract/FREE Full Text
  23. ↵
    Beattie MS, Farooqui AA, and Bresnahan JC (2000) Review of current evidence for apoptosis after spinal cord injury. J Neurotrauma 17: 915-925.
    OpenUrlCrossRefPubMed
  24. ↵
    Ben-Ari Y and Cossart R (2000) Kainate, a double agent that generates seizures: two decades of progress. Trends Neurosci 23: 580-587.
    OpenUrlCrossRefPubMed
  25. ↵
    Berlett BS and Stadtman ER (1997) Protein oxidation in aging, disease, and oxidative stress. J Biol Chem 272: 20313-20316.
    OpenUrlFREE Full Text
  26. ↵
    Blackwell GJ and Flower RJ (1983) Inhibition of phospholipase. Br Med Bull 39: 260-264.
    OpenUrlFREE Full Text
  27. ↵
    Boilard E, Bourgoin SG, Bernatchez C, Poubelle PE, and Surette ME (2003) Interaction of low molecular weight group IIA phospholipase A2 with apoptotic human T cells: role of heparan sulfate proteoglycans. FASEB J 17: 1068-1080.
    OpenUrlAbstract/FREE Full Text
  28. ↵
    Bonventre JV, Huang ZH, Taheri MR, O'Leary E, Li E, Moskowitz MA, and Sapirstein A (1997) Reduced fertility and postischaemic brain injury in mice deficient in cytosolic phospholipase A2. Nature (Lond) 390: 622-625.
    OpenUrlCrossRefPubMed
  29. ↵
    Bosetti F and Weerasinghe GR (2003) The expression of brain cyclooxygenase-2 is down-regulated in the cytosolic phospholipase A2 knockout mouse. J Neurochem 87: 1471-1477.
    OpenUrlCrossRefPubMed
  30. ↵
    Brown WJ, Chambers K, and Doody A (2003) Phospholipase A2 (PLA2) enzymes in membrane trafficking: mediators of membrane shape and function. Traffic 4: 214-221.
    OpenUrlCrossRefPubMed
  31. ↵
    Burgermeister E, Tibes U, Stockinger H, and Scheuer WV (1999) Activation of nuclear factor-κB by lipopolysaccharide in mononuclear leukocytes is prevented by inhibitors of cytosolic phospholipase A2. Eur J Pharmacol 369: 373-386.
    OpenUrlCrossRefPubMed
  32. ↵
    Burgess JR and Kuo CF (1996) Increased calcium-independent phospholipase A2 activity in vitamin E and selenium-deficient rat lung, liver, and spleen cytosol is time-dependent and reversible. J Nutr Biochem 7: 366-374.
    OpenUrlCrossRef
  33. ↵
    Burke JR, Witmer MR, Tredup J, Micanovic R, Gregor KR, Lahiri J, Tramposch KM, and Villafranca JJ (1995) Cooperativity and binding in the mechanism of cytosolic phospholipase A2. Biochemistry 34: 15165-15174.
    OpenUrlCrossRefPubMed
  34. ↵
    Burke JR, Witmer MR, Zusi FC, Gregor KR, Davern LB, Padmanabha R, Swann RT, Smith D, Tredup JA, Micanovic R, et al. (1999) Competitive, reversible inhibition of cytosolic phospholipase A2 at the lipid-water interface by choline derivatives that partially partition into the phospholipid bilayer. J Biol Chem 274: 18864-18871.
    OpenUrlAbstract/FREE Full Text
  35. ↵
    Calder PC and Grimble RF (2002) Polyunsaturated fatty acids, inflammation and immunity. Eur J Clin Nutr 56: S14-S19.
    OpenUrlCrossRefPubMed
  36. ↵
    Calon F, Lim GP, Yang FS, Morihara T, Teter B, Ubeda O, Rostaing P, Triller A, Salem NJ, Ashe KH, et al. (2004) Docosahexaenoic acid protects from dendritic pathology in an Alzheimer's disease mouse model. Neuron 43: 633-645.
    OpenUrlCrossRefPubMed
  37. ↵
    Chabot C, Gagne J, Giguere C, Bernard J, Baudry M, and Massicotte G (1998) Bidirectional modulation of AMPA receptor properties by exogenous phospholipase A2 in the hippocampus. Hippocampus 8: 299-309.
    OpenUrlCrossRefPubMed
  38. ↵
    Chakraborti S (2003) Phospholipase A2 isoforms: a perspective. Cell Signal 15: 637-665.
    OpenUrlCrossRefPubMed
  39. ↵
    Chang MCJ and Jones CR (1998) Chronic lithium treatment decreases brain phospholipase A2 activity. Neurochem Res 23: 887-892.
    OpenUrlCrossRefPubMed
  40. ↵
    Chiba H, Michibata H, Wakimoto K, Seishima M, Kawasaki S, Okubo K, Mitsui H, Torii H, and Imai Y (2004) Cloning of a gene for a novel epithelium-specific cytosolic phospholipase A2, cPLA2δ, induced in psoriatic skin. J Biol Chem 279: 12890-12897.
    OpenUrlAbstract/FREE Full Text
  41. ↵
    Clark JD and Tam S (2004) Potential therapeutic uses of phospholipase A2 inhibitors. Expert Opin Ther Patents 14: 937-950.
    OpenUrlCrossRef
  42. ↵
    Clark MA, Conway TM, Shorr RGL, and Crooke ST (1987) Identification and isolation of a mammalian protein which is antigenically and functionally related to the phospholipase A2 stimulatory peptide melittin. J Biol Chem 262: 4402-4406.
    OpenUrlAbstract/FREE Full Text
  43. ↵
    Clemens JA, Stephenson DT, Smalstig EB, Roberts EF, Johnstone EM, Sharp JD, Little SP, and Kramer RM (1996) Reactive glia express cytosolic phospholipase A2 after transient global forebrain ischemia in the rat. Stroke 27: 527-535.
    OpenUrlAbstract/FREE Full Text
  44. ↵
    Connolly S, Bennion C, Botterell S, Croshaw PJ, Hallam C, Hardy K, Hartopp P, Jackson CG, King SJ, Lawrence L, et al. (2002) Design and synthesis of a novel and potent series of inhibitors of cytosolic phospholipase A2 based on a 1,3-disubstituted propan-2-one skeleton. J Med Chem 45: 1348-1362.
    OpenUrlCrossRefPubMed
  45. ↵
    Corbella B and Vieta E (2003) Molecular targets of lithium action. Acta Neuropsychiatry 15: 316-340.
    OpenUrlCrossRef
  46. ↵
    Cummings BS, McHowat J, and Schnellmann RG (2000) Phospholipase A2s in cell injury and death. J Pharmacol Exp Ther 294: 793-799.
    OpenUrlAbstract/FREE Full Text
  47. ↵
    de Wilde MC, Leenders I, Broersen LM, Kuipers AAM, van der Beek EM, and Kiliaan AJ (2003) The omega-3 fatty acid docosahexaenoic acid (DHA) inhibits the formation of beta amyloid in CHO7PA2 cells, in Society for Neuroscience 33rd Annual Meeting Abstracts; 2003 Nov 8-12; New Orleans, LA; Abstract nr 730.11.
  48. ↵
    DeArmond SJ and Prusiner SB (2003) Perspectives on prion biology, prion disease pathogenesis, and pharmacologic approaches to treatment. Clin Lab Med 23: 1-41.
    OpenUrlCrossRefPubMed
  49. ↵
    DeCoster MA, Lambeau G, Lazdunski M, and Bazan NG (2002) Secreted phospholipase A2 potentiates glutamate-induced calcium increase and cell death in primary neuronal cultures. J Neurosci Res 67: 634-645.
    OpenUrlCrossRefPubMed
  50. ↵
    Demediuk P, Saunders RD, Anderson DK, Means ED, and Horrocks LA (1985) Membrane lipid changes in laminectomized and traumatized cat spinal cord. Proc Natl Acad Sci USA 82: 7071-7075.
    OpenUrlAbstract/FREE Full Text
  51. ↵
    Dempsey RJ and Rao VLR (2003) Cytidinediphosphocholine treatment to decrease traumatic brain injury-induced hippocampal neuronal death, cortical contusion volume, and neurological dysfunction in rats. J Neurosurg 98: 867-873.
    OpenUrlPubMed
  52. ↵
    Denecker G, Vercammen D, Declercq W, and Vandenabeele P (2001) Apoptotic and necrotic cell death induced by death domain receptors. Cell Mol Life Sci 58: 356-370.
    OpenUrlCrossRefPubMed
  53. ↵
    Diaz-Arrastia R and Scott KS (1999) Expression of cPLA2-β and cPLA2-γ, novel paralogs of group IV cytosolic phospholipase A2 in mammalian brain. Soc Neurosci Abs 25: 2206.
    OpenUrl
  54. ↵
    Doh-ura K, Iwaki T, and Caughey B (2000) Lysosomotropic agents and cysteine protease inhibitors inhibit scrapie-associated prion protein accumulation. J Virol 74: 4894-4897.
    OpenUrlAbstract/FREE Full Text
  55. ↵
    Doug M, Guda K, Nambiar PR, Rezaie A, Belinsky GS, Lambeau G, Giardina C, and Rosenberg DW (2003) Inverse association between phospholipase A2 and COX-2 expression during mouse colon tumorigenesis. Carcinogenesis 24: 307-315.
    OpenUrlAbstract/FREE Full Text
  56. ↵
    Douglas CE, Chan AC, and Choy PC (1986) Vitamin E inhibits platelet phospholipase A2. Biochim Biophys Acta 876: 639-645.
    OpenUrlPubMed
  57. ↵
    Dubin NH, Blake DA, DiBlasi MC, Parmley TH, and King TM (1982) Pharmacokinetic studies on quinacrine following intrauterine administration to cynomolgus monkeys. Fertil Steril 38: 735-740.
    OpenUrlPubMed
  58. ↵
    Edgar AD, Strosznajder J, and Horrocks LA (1982) Activation of ethanolamine phospholipase A2 in brain during ischemia. J Neurochem 39: 1111-1116.
    OpenUrlPubMed
  59. ↵
    Endres S and von Schacky C (1996) n-3 polyunsaturated fatty acids and human cytokine synthesis. Curr Opin Lipidol 7: 48-52.
    OpenUrlCrossRefPubMed
  60. ↵
    Estevez AY and Phillis JW (1997) The phospholipase A2 inhibitor, quinacrine, reduces infarct size in rats after transient middle cerebral artery occlusion. Brain Res 752: 203-208.
    OpenUrlCrossRefPubMed
  61. ↵
    Farooqui AA, Antony P, Ong WY, Horrocks LA, and Freysz L (2004a) Retinoic acid-mediated phospholipase A2 signaling in the nucleus. Brain Res Rev 45: 179-195.
    OpenUrlCrossRefPubMed
  62. ↵
    Farooqui AA, Haun SE, and Horrocks LA (1994a) Ischemia and hypoxia, in Basic Neurochemistry (Siegel GJ, Agranoff BW, Albers RW, and Molinoff PB eds) pp 867-883, Raven Press, New York.
  63. ↵
    Farooqui AA and Horrocks LA (1991) Excitatory amino acid receptors, neural membrane phospholipid metabolism and neurological disorders. Brain Res Rev 16: 171-191.
    OpenUrlCrossRefPubMed
  64. ↵
    Farooqui AA and Horrocks LA (1994) Excitotoxicity and neurological disorders: involvement of membrane phospholipids. Int Rev Neurobiol 36: 267-323.
    OpenUrlPubMed
  65. ↵
    Farooqui AA and Horrocks LA (2001) Plasmalogens: workhorse lipids of membranes in normal and injured neurons and glia. Neuroscientist 7: 232-245.
    OpenUrlAbstract/FREE Full Text
  66. ↵
    Farooqui AA and Horrocks LA (2004a) Beneficial effects of docosahexaenoic acid on health of the human brain. Agro Food Ind Hi-Tech 15: 52-53.
    OpenUrl
  67. ↵
    Farooqui AA and Horrocks LA (2004b) Plasmalogens, platelet activating factor, and other ether lipids, in Bioactive Lipids (Nicolaou A and Kokotos G eds) pp 107-134, Oily Press, Bridgwater, UK.
  68. ↵
    Farooqui AA, Horrocks LA, and Farooqui T (2000a) Deacylation and reacylation of neural membrane glycerophospholipids. J Mol Neurosci 14: 123-135.
    OpenUrlCrossRefPubMed
  69. ↵
    Farooqui AA, Litsky ML, Farooqui T, and Horrocks LA (1999) Inhibitors of intracellular phospholipase A2 activity: their neurochemical effects and therapeutical importance for neurological disorders. Brain Res Bull 49: 139-153.
    OpenUrlCrossRefPubMed
  70. ↵
    Farooqui AA, Ong WY, and Horrocks LA (2003a) Plasmalogens, docosahexaenoic acid, and neurological disorders, in Peroxisomal Disorders and Regulation of Genes (Roels F, Baes M, and de Bies S eds) pp 335-354, Kluwer Academic/Plenum Publishers, London.
  71. ↵
    Farooqui AA, Ong WY, and Horrocks LA (2003b) Stimulation of lipases and phospholipases in Alzheimer disease, in Nutrition and Biochemistry of Phospholipids (Szuhaj B and van Nieuwenhuyzen W eds) pp 14-29, AOCS Press, Champaign, IL.
  72. ↵
    Farooqui AA, Ong WY, and Horrocks LA (2004b) Neuroprotection abilities of cytosolic phospholipase A2 inhibitors in kainic acid-induced neurodegeneration. Curr Drug Targets Cardiovasc Haematol Disord 4: 85-96.
    OpenUrlCrossRefPubMed
  73. ↵
    Farooqui AA, Ong WY, Horrocks LA, and Farooqui T (2000b) Brain cytosolic phospholipase A2: Localization, role, and involvement in neurological diseases. Neuroscientist 6: 169-180.
    OpenUrlPubMed
  74. ↵
    Farooqui AA, Ong WY, Lu XR, Halliwell B, and Horrocks LA (2001) Neurochemical consequences of kainate-induced toxicity in brain: involvement of arachidonic acid release and prevention of toxicity by phospholipase A2 inhibitors. Brain Res Rev 38: 61-78.
    OpenUrlCrossRefPubMed
  75. ↵
    Farooqui AA, Ong WY, Lu XR, and Horrocks LA (2002) Cytosolic phospholipase A2 inhibitors as therapeutic agents for neural cell injury. Curr Med Chem - Anti-Inflammatory & Anti-Allergy Agents 1: 193-204.
    OpenUrlCrossRef
  76. ↵
    Farooqui AA, Pendley CE II, Taylor WA, and Horrocks LA (1985) Studies on diacylglycerol lipases and lysophospholipases of bovine brain, in Phospholipids in the Nervous System, Vol. II: Physiological Role (Horrocks LA, Kanfer JN, and Porcellati G eds) pp 179-192, Raven Press, New York.
  77. ↵
    Farooqui AA, Rapoport SI, and Horrocks LA (1997a) Membrane phospholipid alterations in Alzheimer disease: deficiency of ethanolamine plasmalogens. Neurochem Res 22: 523-527.
    OpenUrlCrossRefPubMed
  78. ↵
    Farooqui AA, Rosenberger TA, and Horrocks LA (1997b) Arachidonic acid, neurotrauma, and neurodegenerative diseases, in Handbook of Essential Fatty Acid Biology (Yehuda S and Mostofsky DI eds) pp 277-295, Humana Press, Totowa, NJ.
  79. ↵
    Farooqui AA, Yang H-C, and Horrocks LA (1994b) Purification of lipases, phospholipases and kinases by heparin-Sepharose chromatography. J Chromatogr 673: 149-158.
    OpenUrlCrossRefPubMed
  80. ↵
    Farooqui AA, Yang H-C, and Horrocks LA (1995) Plasmalogens, phospholipases A2, and signal transduction. Brain Res Rev 21: 152-161.
    OpenUrlCrossRefPubMed
  81. ↵
    Farooqui AA, Yang H-C, and Horrocks LA (1997c) Involvement of phospholipase A2 in neurodegeneration. Neurochem Int 30: 517-522.
    OpenUrlCrossRefPubMed
  82. ↵
    Farooqui AA, Yang HC, Rosenberger TA, and Horrocks LA (1997d) Phospholipase A2 and its role in brain tissue. J Neurochem 69: 889-901.
    OpenUrlPubMed
  83. ↵
    Farrelly PV, Kenna BL, Laohachai KL, Bahadi R, Salmona M, Forloni G, and Kourie JI (2003) Quinacrine blocks PrP (106-126)-formed channels. J Neurosci Res 74: 934-941.
    OpenUrlCrossRefPubMed
  84. ↵
    Fernando SR and Pertwee RG (1997) Evidence that methyl arachidonyl fluorophosphate is an irreversible cannabinoid receptor antagonist. Br J Pharmacol 121: 1716-1720.
    OpenUrlCrossRefPubMed
  85. ↵
    Ferrari G, Batistatou A, and Greene LA (1993) Gangliosides rescue neuronal cells from death after trophic factor deprivation. J Neurosci 13: 1879-1887.
    OpenUrlAbstract
  86. ↵
    Fighera MR, Bonini JS, Frussa R, Dutra CS, Hagen MEK, Rubin MA, and Mello CF (2004) Monosialoganglioside increases catalase activity in cerebral cortex of rats. Free Radic Res 38: 495-500.
    OpenUrlCrossRefPubMed
  87. ↵
    Fitzpatrick JS and Baudry M (1994) Blockade of long-term depression in neonatal hippocampal slices by a phospholipase A2 inhibitor. Dev Brain Res 78: 81-86.
    OpenUrlPubMed
  88. ↵
    Follette P (2003) New perspectives for prion therapeutics meeting: prion disease treatment's early promise unravels. Science (Wash DC) 299: 191-192.
    OpenUrlAbstract/FREE Full Text
  89. ↵
    Friguet B, Stadtman ER, and Szweda LI (1994) Modification of glucose-6-phosphate dehydrogenase by 4-hydroxy-2-nonenal. J Biol Chem 269: 21639-21643.
    OpenUrlAbstract/FREE Full Text
  90. ↵
    Fuentes L, Pérez R, Nieto ML, Balsinde J, and Balboa MA (2003) Bromoenol lactone promotes cell death by a mechanism involving phosphatidate phosphohydrolase-1 rather than calcium-independent phospholipase A2. J Biol Chem 278: 44683-44690.
    OpenUrlAbstract/FREE Full Text
  91. ↵
    Fujita S, Ikegaya Y, Nishikawa M, Nishiyama N, and Matsuki N (2001) Docosahexaenoic acid improves long-term potentiation attenuated by phospholipase A2 inhibitor in rat hippocampal slices. Br J Pharmacol 132: 1417-1422.
    OpenUrlCrossRefPubMed
  92. ↵
    Fujita S, Ikegaya Y, Nishiyama N, and Matsuki N (2000) Ca2+-independent phospholipase A2 inhibitor impairs spatial memory of mice. Jpn J Pharmacol 83: 277-278.
    OpenUrlCrossRefPubMed
  93. ↵
    Gamoh S, Hashimoto M, Sugioka K, Hossain MS, Hata N, Misawa Y, and Masumura S (1999) Chronic administration of docosahexaenoic acid improves reference memory-related learning ability in young rats. Neuroscience 93: 237-241.
    OpenUrlCrossRefPubMed
  94. ↵
    Gattaz WF, Maras A, Cairns NJ, Levy R, and Förstl H (1995) Decreased phospholipase A2 activity in Alzheimer brains. Biol Psychiatry 37: 13-17.
    OpenUrlCrossRefPubMed
  95. ↵
    Gattaz WF, Forlenza OV, Talib LL, Barbosa NR, and Bottino CM (2004) Platelet phospholipase A2 activity in Alzheimer's disease and mild cognitive impairment. J Neural Transm 111: 591-601.
    OpenUrlCrossRefPubMed
  96. ↵
    Geddes JW, Ulas J, Brunner LC, Choe W, and Cotman CW (1992) Hippocampal excitatory amino acid receptors in elderly, normal individuals and those with Alzheimer's disease: non-N-methyl-d-aspartate receptors. Neuroscience 50: 23-34.
    OpenUrlCrossRefPubMed
  97. ↵
    Geisler FH, Dorsey FC, and Coleman WP (1991) Recovery of motor function after spinal-cord injury—a randomized, placebo-controlled trial with GM-1 ganglioside. N Engl J Med 324: 1829-1887.
    OpenUrlPubMed
  98. ↵
    Ghelardoni S, Tomita YA, Bell JM, Rapoport SI, and Bosetti F (2004) Chronic carbamazepine selectively downregulates cytosolic phospholipase A2 expression and cyclooxygenase activity in rat brain. Biol Psychiat 56: 248-254.
    OpenUrlCrossRefPubMed
  99. ↵
    Ghomashchi F, Stewart A, Hefner Y, Ramanadham S, Turk J, Leslie CC, and Gelb MH (2001) A pyrrolidine-based specific inhibitor of cytosolic phospholipase A2α blocks arachidonic acid release in a variety of mammalian cells. Biochim Biophys Acta 1513: 160-166.
    OpenUrlPubMed
  100. ↵
    Ginsberg L, Xuereb JH, and Gershfeld NL (1998) Membrane instability, plasmalogen content, and Alzheimer's disease. J Neurochem 70: 2533-2538.
    OpenUrlPubMed
  101. ↵
    Gronich J, Konieczkowski M, Gelb MH, Nemenoff RA, and Sedor JR (1994) Interleukin 1α causes rapid activation of cytosolic phospholipase A2 by phosphorylation in rat mesangial cells. J Clin Investig 93: 1224-1233.
    OpenUrlCrossRefPubMed
  102. ↵
    Grossman A, Zeiler B, and Sapirstein V (2003) Prion protein interactions with nucleic acid: possible models for prion disease and prion function. Neurochem Res 28: 955-963.
    OpenUrlCrossRefPubMed
  103. ↵
    Guan ZZ, Wang YA, Cairns NJ, Lantos PL, Dallner G, and Sindelar PJ (1999) Decrease and structural modifications of phosphatidylethanolamine plasmalogen in the brain with Alzheimer disease. J Neuropathol Exp Neurol 58: 740-747.
    OpenUrlCrossRefPubMed
  104. ↵
    Guentchev M, Voigtlander T, Haberler C, Groschup MH, and Budka H (2000) Evidence for oxidative stress in experimental prion disease. Neurobiol Dis 7: 270-273.
    OpenUrlCrossRefPubMed
  105. ↵
    Hamano H, Nabekura J, Nishikawa M, and Ogawa T (1996) Docosahexaenoic acid reduces GABA response in substantia nigra neuron of rat. J Neurophysiol 75: 1264-1270.
    OpenUrlAbstract/FREE Full Text
  106. ↵
    Han WK, Sapirstein A, Hung CC, Alessandrini A, and Bonventre JV (2003) Cross-talk between cytosolic phospholipase A2α (cPLA2α) and secretory phospholipase A2 (sPLA2) in hydrogen peroxide-induced arachidonic acid release in murine mesangial cells—sPLA2 regulates cPLA2α activity that is responsible for arachidonic acid release. J Biol Chem 278: 24153-24163.
    OpenUrlAbstract/FREE Full Text
  107. ↵
    Han XL, Holtzman DM, and McKeel DW Jr (2001) Plasmalogen deficiency in early Alzheimer's disease subjects and in animal models: molecular characterization using electrospray ionization mass spectrometry. J Neurochem 77: 1168-1180.
    OpenUrlCrossRefPubMed
  108. ↵
    Hanasaki K and Arita H (2002) Phospholipase A2 receptor: a regulator of biological functions of secretory phospholipase A2. Prostaglandins Other Lipid Mediat 68-69: 71-82.
    OpenUrlCrossRefPubMed
  109. ↵
    Hannon GJ (2002) RNA interference. Nature (Lond) 418: 244-251.
    OpenUrlCrossRefPubMed
  110. ↵
    Hashimoto M, Hossain S, Shimada T, Sugioka K, Yamasaki H, Fujii Y, Ishibashi Y, Oka JI, and Shido O (2002) Docosahexaenoic acid provides protection from impairment of learning ability in Alzheimer's disease model rats. J Neurochem 81: 1084-1091.
    OpenUrlCrossRefPubMed
  111. ↵
    Hazen SL and Gross RW (1993) The specific association of a phosphofructokinase isoform with myocardial calcium-independent phospholipase A2: implications for the coordinated regulation of phospholipolysis and glycolysis. J Biol Chem 268: 9892-9900.
    OpenUrlAbstract/FREE Full Text
  112. ↵
    Hernandez M, Nieto ML, and Crespo MS (2000) Cytosolic phospholipase A2 and the distinct transcriptional programs of astrocytoma cells. Trends Neurosci 23: 259-264.
    OpenUrlCrossRefPubMed
  113. ↵
    Hirabayashi T, Murayama T, and Shimizu T (2004) Regulatory mechanism and physiological role of cytosolic phospholipase A2. Biol Pharm Bull 27: 1168-1173.
    OpenUrlCrossRefPubMed
  114. ↵
    Hirabayashi T and Shimizu T (2000) Localization and regulation of cytosolic phospholipase A2. Biochim Biophys Acta 1488: 124-138.
    OpenUrlPubMed
  115. ↵
    Hirashima Y, Farooqui AA, Mills JS, and Horrocks LA (1992) Identification and purification of calcium-independent phospholipase A2 from bovine brain cytosol. J Neurochem 59: 708-714.
    OpenUrlCrossRefPubMed
  116. ↵
    Hong KH, Bonventre JC, O'Leary E, Bonventre JV, and Lander ES (2001) Deletion of cytosolic phospholipase A2 suppresses ApcMin-induced tumorigenesis. Proc Natl Acad Sci USA 98: 3935-3939.
    OpenUrlAbstract/FREE Full Text
  117. ↵
    Hong S, Gronert K, Devchand PR, Moussignac RL, and Serhan CN (2003) Novel docosatrienes and 17S-resolvins generated from docosahexaenoic acid in murine brain, human blood, and glial cells—autacoids in anti-inflammation. J Biol Chem 278: 14677-14687.
    OpenUrlAbstract/FREE Full Text
  118. ↵
    Horrobin DF (2003) Eicosapentaenoic acid derivatives in the management of schizophrenia, in Phospholipid Spectrum Disorders in Psychiatry and Neurology (Peet M, Glen L, and Horrobin DF eds) pp 371-376, Marius Press, Carnforth, Lancashire, UK.
  119. ↵
    Horrocks LA, Demediuk P, Saunders RD, Dugan L, Clendenon NR, Means ED, and Anderson DK (1985) The degradation of phospholipids, formation of metabolites of arachidonic acid, and demyelination following experimental spinal cord injury. Cent Nerv Syst Trauma 2: 115-120.
    OpenUrlPubMed
  120. ↵
    Horrocks LA and Farooqui AA (2004) Docosahexaenoic acid in the diet: its importance in maintenance and restoration of neural membrane function. Prostaglandins Leukotrienes Essent Fatty Acids 70: 361-372.
    OpenUrlCrossRefPubMed
  121. ↵
    Hossain MS, Hashimoto M, Gamoh S, and Masumura S (1999) Antioxidative effects of docosahexaenoic acid in the cerebrum versus cerebellum and brainstem of aged hypercholesterolemic rats. J Neurochem 72: 1133-1138.
    OpenUrlCrossRefPubMed
  122. ↵
    Hossain MS, Hashimoto M, and Masumura S (1998) Influence of docosahexaenoic acid on cerebral lipid peroxide level in aged rats with and without hypercholes-terolemia. Neurosci Lett 244: 157-160.
    OpenUrlCrossRefPubMed
  123. ↵
    Hudson CJ, Kennedy JL, Gotowiec A, Lin A, King N, Gojtan K, Macciardi F, Skorecki K, Meltzer HY, Warsh JJ, et al. (1996) Genetic variant near cytosolic phospholipase A2 associated with schizophrenia. Schizophr Res 21: 111-116.
    OpenUrlCrossRefPubMed
  124. ↵
    Huterer SJ, Tourtellotte WW, and Wherrett JR (1995) Alterations in the activity of phospholipases A2 in post-mortem white matter from patients with multiple sclerosis. Neurochem Res 20: 1335-1343.
    OpenUrlCrossRefPubMed
  125. ↵
    Ibuki T, Matsumura K, Yamazaki Y, Nozaki T, Tanaka Y, and Kobayashi S (2003) Cyclooxygenase-2 is induced in the endothelial cells throughout the central nervous system during carrageenan-induced hind paw inflammation; its possible role in hyperalgesia. J Neurochem 86: 318-328.
    OpenUrlCrossRefPubMed
  126. ↵
    James MJ, Gibson RA, and Cleland LG (2000) Dietary polyunsaturated fatty acids and inflammatory mediator production. Am J Clin Nutr 71: 343S-348S.
    OpenUrlAbstract/FREE Full Text
  127. ↵
    Jamme I, Petit E, Divoux D, Gerbi A, Maxient JM, and Nouvelot A (1995) Modulation of mouse cerebral Na+, K+-ATPase activity by oxygen free radicals. NeuroReport 7: 333-337.
    OpenUrlPubMed
  128. ↵
    Jeffrey M, Goodsir CM, Bruce ME, McBride PA, Scott JR, and Halliday WG (1992) Infection specific prion protein (PrP) accumulates on neuronal plasmalemma in scrapie infected mice. Neurosci Lett 147: 106-109.
    OpenUrlCrossRefPubMed
  129. ↵
    Junqueira R, Cordeiro Q, Meira-Lima I, Gattaz WF, and Vallada H (2004) Allelic association analysis of phospholipase A2 genes with schizophrenia. Psychiatr Genet 14: 157-160.
    OpenUrlCrossRefPubMed
  130. ↵
    Jupp OJ, Vandenabeele P, and MacEwan DJ (2003) Distinct regulation of cytosolic phospholipase A2 phosphorylation, translocation, proteolysis and activation by tumour necrosis factor-receptor subtypes. Biochem J 374: 453-461.
    OpenUrlCrossRefPubMed
  131. ↵
    Kaetzel MA and Dedman JR (1995) Annexins: novel Ca2+-dependent regulators of membrane function. News Physiol Sci 10: 171-176.
    OpenUrlAbstract/FREE Full Text
  132. ↵
    Kajiwara K, Nagawawa H, Shimizu-Nishikawa S, Ookuri T, Kimura M, and Sugaya E (1996) Molecular characterization of seizure-related genes isolated by differential screening. Biochem Biophys Res Commun 219: 795-799.
    OpenUrlCrossRefPubMed
  133. ↵
    Kalyvas A and David S (2004) Cytosolic phospholipase A2 plays a key role in the pathogenesis of multiple sclerosis-like disease. Neuron 41: 323-335.
    OpenUrlCrossRefPubMed
  134. ↵
    Kanfer JN, Sorrentino G, and Sitar DS (1998) Phospholipases as mediators of amyloid β-peptide neurotoxicity: an early event contributing to neurodegeneration characteristic of Alzheimer's disease. Neurosci Lett 257: 93-96.
    OpenUrlCrossRefPubMed
  135. ↵
    Katsuki H and Okuda S (1995) Arachidonic acid as a neurotoxic and neurotrophic substance. Prog Neurobiol 46: 607-636.
    OpenUrlCrossRefPubMed
  136. ↵
    Kim DK, Rordorf G, Nemenoff RA, Koroshetz WJ, and Bonventre JV (1995) Glutamate stably enhances the activity of two cytosolic forms of phospholipase A2 in brain cortical cultures. Biochem J 310: 83-90.
    OpenUrlAbstract/FREE Full Text
  137. ↵
    Kishimoto K, Matsumura K, Kataoka Y, Morii H, and Watanabe Y (1999) Localization of cytosolic phospholipase A2 messenger RNA mainly in neurons in the rat brain. Neuroscience 92: 1061-1077.
    OpenUrlCrossRefPubMed
  138. ↵
    Klivenyi P, Beal MF, Ferrante RJ, Andreassen OA, Wermer M, Chin MR, and Bonventre JV (1998) Mice deficient in group IV cytosolic phospholipase A2 are resistant to MPTP neurotoxicity. J Neurochem 71: 2634-2637.
    OpenUrlPubMed
  139. ↵
    Kobayashi Y, Hirata K, Tanaka H, and Yamada T (2003) Quinacrine administration to a patient with Creutzfeldt-Jakob disease who received a cadaveric dura mater graft—an EEG evaluation. Rinsho Shinkeigaku 43: 403-408.
    OpenUrlPubMed
  140. ↵
    Kokotos G, Six DA, Loukas V, Smith T, Constantinou-Kokotou V, Hadjipavlou-Litina D, Kotsovolou S, Chiou A, Beltzner CC, and Dennis EA (2004) Inhibition of group IVA cytosolic phospholipase A2 by novel 2-oxoamides in vitro, in cells, and in vivo. J Med Chem 47: 3615-3628.
    OpenUrlCrossRefPubMed
  141. Kolko M, DeCoster MA, Rodriguez de Turco EB, and Bazan NG (1996) Synergy by secretory phospholipase A2 and glutamate on inducing cell death and sustained arachidonic acid metabolic changes in primary cortical neuronal cultures. J Biol Chem 271: 32722-32728.
    OpenUrlAbstract/FREE Full Text
  142. ↵
    Kolko M, Rodriguez de Turco EB, Diemer NH, and Bazan NG (2002) Secretory phospholipase A2-mediated neuronal cell death involves glutamate ionotropic receptors. NeuroReport 13: 1963-1966.
    OpenUrlCrossRefPubMed
  143. ↵
    Korth C, May BC, Cohen FE, and Prusiner SB (2001) Acridine and phenothiazine derivatives as pharmacotherapeutics for prion disease. Proc Natl Acad Sci USA 98: 9836-9841.
    OpenUrlAbstract/FREE Full Text
  144. ↵
    Kramer BC, Yabut JA, Cheong J, Jnobaptiste R, Robakis T, Olanow CW, and Mytilineou C (2004) Toxicity of glutathione depletion in mesencephalic cultures: a role for arachidonic acid and its lipoxygenase metabolites. Eur J Neurosci 19: 280-286.
    OpenUrlCrossRefPubMed
  145. ↵
    Kriem B, Sponne I, Fifre A, Malaplate-Armand C, Lozac'h-Pillot K, Koziel V, Yen-Potin FT, Bihain B, Oster T, Olivier JL, et al. (2005) Cytosolic phospholipase A2 mediates neuronal apoptosis induced by soluble oligomers of the amyloid-β peptide. FASEB J 19: 85-87.
    OpenUrlAbstract/FREE Full Text
  146. ↵
    Kuroiwa N, Nakamura M, Tagaya M, and Takatsuki A (2001) Arachidonyltrifluoromethyl ketone, a phospholipase A2 antagonist, induces dispersal of both Golgi stack- and trans Golgi network-resident proteins throughout the cytoplasm. Biochem Biophys Res Commun 281: 582-588.
    OpenUrlCrossRefPubMed
  147. Kurrasch-Orbaugh DM, Parrish JC, Watts VJ, and Nichols DE (2003) A complex signaling cascade links the serotonin2A receptor to phospholipase A2 activation: the involvement of MAP kinases. J Neurochem 86: 980-991.
    OpenUrlCrossRefPubMed
  148. ↵
    Laktionova P, Rykova E, Toni M, Spisni E, Griffoni C, Bryksin A, Volodko N, Vlassov V, and Tomasi V (2004) Knock down of cytosolic phospholipase A2: an antisense oligonucleotide having a nuclear localization binds a C-terminal motif of glyceraldehyde-3-phosphate dehydrogenase. Biochim Biophys Acta 1636: 129-135.
    OpenUrlPubMed
  149. ↵
    Larsson PKA, Claesson HE, and Kennedy BP (1998) Multiple splice variants of the human calcium-independent phospholipase A2 and their effect on enzyme activity. J Biol Chem 273: 207-214.
    OpenUrlAbstract/FREE Full Text
  150. ↵
    Laruelle M, Abi-Dargham A, Gil R, Kegeles L, and Innis R (1999) Increased dopamine transmission in schizophrenia: relationship to illness phases. Biol Psychiatry 46: 56-72.
    OpenUrlCrossRefPubMed
  151. ↵
    Latorre E, Collado MP, Fernández I, Aragonés MD, and Catalán RE (2003) Signaling events mediating activation of brain ethanolamine plasmalogen hydrolysis by ceramide. Eur J Biochem 270: 36-46.
    OpenUrlPubMed
  152. ↵
    Lazarewicz JW, Wroblewski JT, and Costa E (1990) N-Methyl-d-aspartate-sensitive glutamate receptors induce calcium-mediated arachidonic acid release in primary cultures of cerebellar granule cells. J Neurochem 55: 1875-1881.
    OpenUrlCrossRefPubMed
  153. ↵
    Lehr M (1996) 3-(3,5-Dimethyl-4-octadecanoylpyrrol-2-yl)propionic acids as inhibitors of 85 kDa cytosolic phospholipase A2. Arch Pharm (Weinheim) 329: 483-488.
    OpenUrlPubMed
  154. ↵
    Lehr M (1997a) Structure-activity relationship studies on (4-acylpyrrol-2-yl)alkanoic acids as inhibitors of the cytosolic phospholipase A2: variation of the alkanoic acid substituent, the acyl chain and the position of the pyrrole nitrogen. Eur J Med Chem 32: 805-814.
    OpenUrlCrossRef
  155. ↵
    Lehr M (1997b) Synthesis, biological evaluation, and structure-activity relationships of 3-acylindole-2-carboxylic acids as inhibitors of the cytosolic phospholipase A2. J Med Chem 40: 2694-2705.
    OpenUrlCrossRefPubMed
  156. ↵
    Lehtonen JYA, Holopainen JM, and Kinnunen PKJ (1996) Activation of phospholipase A2 by amyloid β-peptides in vitro. Biochemistry 35: 9407-9414.
    OpenUrlCrossRefPubMed
  157. ↵
    Lerma J (1997) Kainate reveals its targets. Neuron 19: 1155-1158.
    OpenUrlCrossRefPubMed
  158. ↵
    Leslie CC (2004) Regulation of arachidonic acid availability for eicosanoid production. Biochem Cell Biol 82: 1-17.
    OpenUrlCrossRefPubMed
  159. ↵
    Lin TN, Wang Q, Simonyi A, Chen JJ, Cheung WM, He YY, Xu J, Sun AY, Hsu CY, and Sun GY (2004) Induction of secretory phospholipase A2 in reactive astrocytes in response to transient focal cerebral ischemia in the rat brain. J Neurochem 90: 637-645.
    OpenUrlCrossRefPubMed
  160. ↵
    Locati M, Lamorte G, Luini W, Introna M, Bernasconi S, Mantovani A, and Sozzani S (1996) Inhibition of monocyte chemotaxis to C-C chemokines by antisense oligonucleotide for cytosolic phospholipase A2. J Biol Chem 271: 6010-6016.
    OpenUrlAbstract/FREE Full Text
  161. ↵
    Lonergan PE, Martin DSD, Horrobin DF, and Lynch MA (2004) Neuroprotective actions of eicosapentaenoic acid on lipopolysaccharide-induced dysfunction in rat hippocampus. J Neurochem 91: 20-29.
    OpenUrlCrossRefPubMed
  162. ↵
    Love R (2001) Old drugs to treat new variant Creutzfeldt-Jakob disease. Lancet 358: 563.
    OpenUrlCrossRefPubMed
  163. ↵
    Lu XR, Ong WY, and Halliwell B (2001a) The phospholipase A2 inhibitor quinacrine prevents increased immunoreactivity to cytoplasmic phospholipase A2 (cPLA2) and hydroxynonenal (HNE) in neurons of the lateral septum following fimbria-fornix transection. Exp Brain Res 138: 500-508.
    OpenUrlCrossRefPubMed
  164. ↵
    Lu XR, Ong WY, Halliwell B, Horrocks LA, and Farooqui AA (2001b) Differential effects of calcium-dependent and calcium-independent phospholipase A2 inhibitors on kainate-induced neuronal injury in rat hippocampal slices. Free Radic Biol Med 30: 1263-1273.
    OpenUrlCrossRefPubMed
  165. ↵
    Lukácová N, Halát G, Chavko M, and Marsala J (1996) Ischemia-reperfusion injury in the spinal cord of rabbits strongly enhances lipid peroxidation and modifies phospholipid profiles. Neurochem Res 21: 869-873.
    OpenUrlPubMed
  166. ↵
    Macchioni L, Corazzi L, Nardicchi V, Mannucci R, Arcuri C, Porcellati S, Sposini T, Donato R, and Goracci G (2004) Rat brain cortex mitochondria release group II secretory phospholipase A2 under reduced membrane potential. J Biol Chem 279: 37860-37869.
    OpenUrlAbstract/FREE Full Text
  167. ↵
    Mandal PK and Pettegrew JW (2004) Alzheimer's disease: NMR studies of asialo (GM1) and trisialo (GT1b) ganglioside interactions with A β(1-40) peptide in a membrane mimic environment. Neurochem Res 29: 447-453.
    OpenUrlCrossRefPubMed
  168. ↵
    Marcheselli VL, Hong S, Lukiw WJ, Tian XH, Gronert K, Musto A, Hardy M, Gimenez JM, Chiang N, Serhan CN, et al. (2003) Novel docosanoids inhibit brain ischemia-reperfusion-mediated leukocyte infiltration and pro-inflammatory gene expression. J Biol Chem 278: 43807-43817.
    OpenUrlAbstract/FREE Full Text
  169. ↵
    Matsuzawa A, Murakami M, Atsumi G, Imai K, Prados P, Inoue K, and Kudo I (1996) Release of secretory phospholipase A2 from rat neuronal cells and its possible function in the regulation of catecholamine secretion. Biochem J 318: 701-709.
    OpenUrlAbstract/FREE Full Text
  170. ↵
    May BCH, Fafarman AT, Hong SB, Rogers M, Deady LW, Prusiner SB, and Cohen FE (2003) Potent inhibition of scrapie prion replication in cultured cells by bisacridines. Proc Natl Acad Sci USA 100: 3416-3421.
    OpenUrlAbstract/FREE Full Text
  171. ↵
    Mazière C, Conte MA, Degonville J, Ali D, and Mazière JC (1999) Cellular enrichment with polyunsaturated fatty acids induces an oxidative stress and activates the transcription factors AP1 and NFκB. Biochem Biophys Res Commun 265: 116-122.
    OpenUrlCrossRefPubMed
  172. ↵
    McIntosh TK, Saatman KE, Raghupathi R, Graham DI, Smith DH, Lee VM, and Trojanowski JQ (1998) The molecular and cellular sequelae of experimental traumatic brain injury: pathogenetic mechanisms. Neuropathol Appl Neurobiol 24: 251-267.
    OpenUrlCrossRefPubMed
  173. ↵
    Miele L (2003a) New weapons against inflammation: dual inhibitors of phospholipase A2 and transglutaminase. J Clin Investig 111: 19-21.
    OpenUrlCrossRefPubMed
  174. ↵
    Miele L (2003b) New weapons against inflammation: dual inhibitors of phospholipase A2 and transglutaminase. J Clin Investig 111: 19-21.
    OpenUrlCrossRefPubMed
  175. ↵
    Milhavet O, McMahon HE, Rachidi W, Nishida N, Katamine S, Mange A, Arlotto M, Casanova D, Riondel J, Favier A, et al. (2000) Prion infection impairs the cellular response to oxidative stress. Proc Natl Acad Sci USA 97: 13937-13942.
    OpenUrlAbstract/FREE Full Text
  176. ↵
    Mir C, Clotet J, Aledo R, Durany N, Argemi J, Lozano R, Cervos-Navarro J, and Casals N (2003) CDP-choline prevents glutamate-mediated cell death in cerebellar granule neurons. J Mol Neurosci 20: 53-59.
    OpenUrlCrossRefPubMed
  177. ↵
    Molloy GY, Rattray M, and Williams RJ (1998) Genes encoding multiple forms of phospholipase A2 are expressed in rat brain. Neurosci Lett 258: 139-142.
    OpenUrlCrossRefPubMed
  178. ↵
    Mosior M, Six DA, and Dennis EA (1998a) Group IV cytosolic phospholipase A2 binds with high affinity and specificity to phosphatidylinositol 4,5-bisphosphate resulting in dramatic increases in activity. J Biol Chem 273: 2184-2191.
    OpenUrlAbstract/FREE Full Text
  179. ↵
    Moskowitz N, Puszkin S, and Schook W (1983) Characterization of brain synaptic vesicle phospholipase A2 activity and its modulation by calmodulin, prostaglandin E2, prostaglandin F2α, cyclic AMP and ATP. J Neurochem 41: 1576-1586.
    OpenUrlPubMed
  180. ↵
    Mukherjee PK, Marcheselli VL, Serhan CN, and Bazan NG (2004) Neuroprotectin D1: a docosahexaenoic acid-derived docosatriene protects human retinal pigment epithelial cells from oxidative stress. Proc Natl Acad Sci USA 101: 8491-8496.
    OpenUrlAbstract/FREE Full Text
  181. ↵
    Muller WEG, Ushijima H, Schroder HC, Forrest JMS, Schatton WFH, Rytik PG, and Heffner-Lauc M (1993) Cytoprotective effect of NMDA receptor antagonists on prion protein (PrionSc)-induced toxicity in rat cortical cell cultures. Eur J Pharmacol 246: 261-267.
    OpenUrlCrossRefPubMed
  182. ↵
    Murakami M, Nakatani Y, Atsumi G, Inoue K, and Kudo I (1997) Regulatory functions of phospholipase A2. Crit Rev Immunol 17: 225-283.
    OpenUrlCrossRefPubMed
  183. ↵
    Murphy EJ and Horrocks LA (1993) CDPcholine, CDPethanolamine, lipid metabolism and disorders of the central nervous system, in Phospholipids and Signal Transmission (Massarelli R, Horrocks L, Kanfer JN, and Löffelholz K eds) pp 353-372, Springer-Verlag GmbH, Heidelberg.
  184. ↵
    Nakashima S, Ikeno Y, Yokoyama T, Kuwana M, Bolchi A, Ottonello S, Kitamoto K, and Arioka M (2003) Secretory phospholipases A2 induce neurite outgrowth in PC12 cells. Biochem J 376: 655-666.
    OpenUrlCrossRefPubMed
  185. ↵
    Ng CH and Ong WY (2001) Increased expression of γ-aminobutyric acid transporters GAT-1 and GAT-3 in the spinal trigeminal nucleus after facial carrageenan injections. Pain 92: 29-40.
    OpenUrlCrossRefPubMed
  186. ↵
    Nordvik I, Myhr KM, Nyland H, and Bjerve KS (2000) Effect of dietary advice and n-3 supplementation in newly diagnosed MS patients. Acta Neurol Scand 102: 143-149.
    OpenUrlCrossRefPubMed
  187. ↵
    Obata T, Sakurai Y, Kase Y, Tanifuji Y, Horiguchi T, Haq S, Kilter H, Michael A, Tao J, O'Leary E, et al. (2003) Simultaneous determination of endocannabinoids (arachidonylethanolamide and 2-arachidonylglycerol) and isoprostane (8-epiprostaglandin F2α) by gas chromatography-mass spectrometry-selected ion monitoring for medical samples. J Chromatogr B 792: 131-140.
    OpenUrlCrossRef
  188. ↵
    Oinuma H, Takamura T, Hasegawa T, Nomoto KI, Naitoh T, Daiku Y, Hamano S, Kakisawa H, and Minami N (1991) Synthesis and biological evaluation of substituted benzenesulfonamides as novel potent membrane-bound phospholipase A2 inhibitors. J Med Chem 34: 2260-2267.
    OpenUrlCrossRefPubMed
  189. ↵
    Ong WY, He Y, Suresh S, and Patel S (1997) Differential expression of apoD and apoE in the kainate lesioned rat hippocampus. Neuroscience 79: 359-367.
    OpenUrlCrossRefPubMed
  190. ↵
    Ong WY, Hu CY, Hjelle OP, Ottersen OP, and Halliwell B (2000) Changes in glutathione in the hippocampus of rats injected with kainate: depletion in neurons and upregulation in glia. Exp Brain Res 132: 510-516.
    OpenUrlCrossRefPubMed
  191. ↵
    Ong WY, Ling SF, Yeo JF, Chiueh CC, and Farooqui AA (2005) Injury and recovery of pyramidal neurons in the rat hippocampus after a single episode of oxidative stress induced by intracerebroventricular injection of ferrous ammonium citrate. Reprod Nutr Dev 45: 647-662.
    OpenUrlCrossRefPubMed
  192. ↵
    Ong WY, Lu XR, Horrocks LA, Farooqui AA, and Garey LJ (2003a) Induction of astrocytic cytoplasmic phospholipase A2 and neuronal death after intracerebroventricular carrageenan injection, and neuroprotective effects of quinacrine. Exp Neurol 183: 449-457.
    OpenUrlCrossRefPubMed
  193. ↵
    Ong WY, Lu XR, Ong BKC, Horrocks LA, Farooqui AA, and Lim SK (2003b) Quinacrine abolishes increases in cytoplasmic phospholipase A2 mRNA levels in the rat hippocampus after kainate-induced neuronal injury. Exp Brain Res 148: 521-524.
    OpenUrlPubMed
  194. ↵
    Ong WY, Ren MQ, Makjanic J, Lim TM, and Watt F (1999) A nuclear microscopic study of elemental changes in the rat hippocampus after kainate-induced neuronal injury. J Neurochem 72: 1574-1579.
    OpenUrlCrossRefPubMed
  195. ↵
    Ono T, Yamada K, Chikazawa Y, Ueno M, Nakamoto S, Okuno T, and Seno K (2002) Characterization of a novel inhibitor of cytosolic phospholipase A2α, pyrrophenone. Biochem J 363: 727-735.
    OpenUrlCrossRefPubMed
  196. ↵
    Ortiz GG, Sánchez-Ruiz MY, Tan DX, Reiter RJ, Benítez-King G, and Beas-Zárate C (2001) Melatonin, vitamin E, and estrogen reduce damage induced by kainic acid in the hippocampus: potassium-stimulated GABA release. J Pineal Res 31: 62-67.
    OpenUrlCrossRefPubMed
  197. ↵
    Ouyang Y and Kaminski NE (1999) Phospholipase A2 inhibitors p-bromophenacyl bromide and arachidonyl trifluoromethyl ketone suppressed interleukin-2 (IL-2) expression in murine primary splenocytes. Arch Toxicol 73: 1-6.
    OpenUrlCrossRefPubMed
  198. ↵
    Pae CU, Yu HS, Kim JJ, Lee CU, Lee SJ, Lee KU, Jun TY, Paik IH, Serretti A, and Lee C (2004) BanI polymorphism of the cytosolic phospholipase A2 gene and mood disorders in the Korean population. Neuropsychobiology 49: 185-188.
    OpenUrlCrossRefPubMed
  199. ↵
    Palomba L, Bianchi M, Persichini T, Magnani M, Colasanti M, and Cantoni O (2004) Downregulation of nitric oxide formation by cytosolic phospholipase A2-released arachidonic acid. Free Radic Biol Med 36: 319-329.
    OpenUrlCrossRefPubMed
  200. ↵
    Pardue S, Rapoport SI, and Bosetti F (2003) Co-localization of cytosolic phospholipase A2 and cyclooxygenase-2 in rhesus monkey cerebellum. Mol Brain Res 116: 106-114.
    OpenUrlCrossRefPubMed
  201. ↵
    Peet M and Ryles S (2001) Eicosapentaenoic acid—a potential new treatment for schizophrenia?, in Fatty Acids: Physiological and Behavioral Functions (Mostofsky DI, Yehuda S, and Salem N eds) pp 345-356, Humana Press Inc, Totowa, NJ.
  202. ↵
    Pentland AP, Morrison AR, Jacobs SC, Hruza LL, Hebert JS, and Packer L (1992) Tocopherol analogs suppress arachidonic acid metabolism via phospholipase inhibition. J Biol Chem 267: 15578-15584.
    OpenUrlAbstract/FREE Full Text
  203. ↵
    Pérez R, Melero R, Balboa MA, and Balsinde J (2004) Role of group VIA calcium-independent phospholipase A2 in arachidonic acid release, phospholipid fatty acid incorporation, and apoptosis in U937 cells responding to hydrogen peroxide. J Biol Chem 279: 40385-40391.
    OpenUrlAbstract/FREE Full Text
  204. ↵
    Perovic S, Pergande G, Ushijima H, Kelve M, Forrest J, and Muller WEG (1995) Flupirtine partially prevents neuronal injury induced by prion protein fragment and lead acetate. Neurodegeneration 4: 369-374.
    OpenUrlCrossRefPubMed
  205. ↵
    Pettegrew JW (1989) Molecular insights into Alzheimer disease. Ann NY Acad Sci 568: 5-28.
    OpenUrlPubMed
  206. ↵
    Pettegrew JW, Klunk WE, Kanal E, Panchalingam K, and McClure RJ (1995) Changes in brain membrane phospholipid and high-energy phosphate metabolism precede dementia. Neurobiol Aging 16: 973-975.
    OpenUrlCrossRefPubMed
  207. ↵
    Pettegrew JW, Panchalingam K, Hamilton RL, and McClure RJ (2001) Brain membrane phospholipid alterations in Alzheimer's disease. Neurochem Res 26: 771-782.
    OpenUrlCrossRefPubMed
  208. ↵
    Pettus BJ, Bielawska A, Subramanian P, Wijesinghe DS, Maceyka M, Leslie CC, Evans JH, Freiberg J, Roddy P, Hannun YA, et al. (2004) Ceramide 1-phosphate is a direct activator of cytosolic phospholipase A2. J Biol Chem 279: 11320-11326.
    OpenUrlAbstract/FREE Full Text
  209. ↵
    Peyrin JM, Lasmezas CI, Haik S, Tagliavini F, Salmona M, Williams A, Richie D, Deslys JP, and Dormont D (1999) Microglial cells respond to amyloidogenic PrP peptide by the production of inflammatory cytokines. NeuroReport 10: 723-729.
    OpenUrlCrossRefPubMed
  210. ↵
    Phillis JW and O'Regan MH (2004) A potentially critical role of phospholipases in central nervous system ischemic, traumatic, and neurodegenerative disorders. Brain Res Rev 44: 13-47.
    OpenUrlCrossRefPubMed
  211. ↵
    Pickard RT, Strifler BA, Kramer RM, and Sharp JD (1999) Molecular cloning of two new human paralogs of 85-kDa cytosolic phospholipase A2. J Biol Chem 274: 8823-8831.
    OpenUrlAbstract/FREE Full Text
  212. ↵
    Pinto F, Brenner T, Dan P, Krimsky M, and Yedgar S (2003) Extracellular phospholipase A2 inhibitors suppress central nervous system inflammation. Glia 44: 275-282.
    OpenUrlCrossRefPubMed
  213. ↵
    Pirianov G, Danielsson C, Carlberg C, James SY, and Colston KW (1999) Potentiation by vitamin D analogs of TNFα and ceramide-induced apoptosis in MCF-7 cells is associated with activation of cytosolic phospholipase A2. Cell Death Differ 6: 890-901.
    OpenUrlCrossRefPubMed
  214. ↵
    Portilla D and Dai G (1996) Purification of a novel calcium-independent phospholipase A2 from rabbit kidney. J Biol Chem 271: 15451-15457.
    OpenUrlAbstract/FREE Full Text
  215. ↵
    Prusiner SB (2001) Shattuck lecture—Neurodegenerative diseases and prions. N Engl J Med 344: 1516-1526.
    OpenUrlCrossRefPubMed
  216. Qu Y, Chang L, Klaff J, Seeman R, Balbo A, and Rapoport SI (2003a) Imaging of brain serotonergic neurotransmission involving phospholipase A2 activation and arachidonic acid release in unanesthetized rats. Brain Res Brain Res Protoc 12: 16-25.
    OpenUrlCrossRefPubMed
  217. Qu Y, Chang L, Klaff J, Seemann R, and Rapoport SI (2003b) Imaging brain phospholipase A2-mediated signal transduction in response to acute fluoxetine administration in unanesthetized rats. Neuropsychopharmacology 28: 1219-1226.
    OpenUrlPubMed
  218. ↵
    Rao AM, Hatcher JF, and Dempsey RJ (2001) Does CDP-choline modulate phospholipase activities after transient forebrain ischemia? Brain Res 893: 268-272.
    OpenUrlCrossRefPubMed
  219. ↵
    Rapoport SI (1999) In vivo fatty acid incorporation into brain phospholipids in relation to signal transduction and membrane remodeling. Neurochem Res 24: 1403-1415.
    OpenUrlCrossRefPubMed
  220. ↵
    Ray P, Ray R, Broomfield CA, and Berman JD (1994) Inhibition of bioenergetics alters intracellular calcium, membrane composition, and fluidity in a neuronal cell line. Neurochem Res 19: 57-63.
    OpenUrlCrossRefPubMed
  221. ↵
    Riendeau D, Guay J, Weech PK, Laliberté F, Yergey J, Li C, Desmarais S, Perrier H, Liu S, Nicoll-Griffith D, et al. (1994) Arachidonyl trifluoromethyl ketone, a potent inhibitor of 85-kDa phospholipase A2, blocks production of arachidonate and 12-hydroxyeicosatetraenoic acid by calcium ionophore-challenged platelets. J Biochem 269: 15619-15624.
    OpenUrl
  222. ↵
    Rintala J, Seemann R, Chandrasekaran K, Rosenberger TA, Chang L, Contreras MA, Rapoport SI, and Chang MCJ (1999) 85 kDa cytosolic phospholipase A2 is a target for chronic lithium in rat brain. NeuroReport 10: 3887-3890.
    OpenUrlCrossRefPubMed
  223. ↵
    Rordorf G, Uemura Y, and Bonventre JV (1991) Characterization of phospholipase A2 (PLA2) activity in gerbil brain: enhanced activities of cytosolic, mitochondrial, and microsomal forms after ischemia and reperfusion. J Neurosci 11: 1829-1836.
    OpenUrlAbstract
  224. ↵
    Rosales-Corral S, Tan DX, Reiter RJ, Valdivia-Velázquez M, Acosta-Martiínez JP, and Ortiz GG (2004) Kinetics of the neuroinflammation-oxidative stress correlation in rat brain following the injection of fibrillar amyloid-β onto the hippocampus in vivo. J Neuroimmunol 150: 20-28.
    OpenUrlCrossRefPubMed
  225. ↵
    Rosenberger TA, Villacreses NE, Contreras MA, Bonventre JV, and Rapoport SI (2003) Brain lipid metabolism in the cPLA2 knockout mouse. J Lipid Res 44: 109-117.
    OpenUrlAbstract/FREE Full Text
  226. ↵
    Rosenberger TA, Villacreses NE, Hovda JT, Bosetti F, Weerasinghe G, Wine RN, Harry GJ, and Rapoport SI (2004) Rat brain arachidonic acid metabolism is increased by a 6-day intracerebral ventricular infusion of bacterial lipopolysaccharide. J Neurochem 88: 1168-1178.
    OpenUrlCrossRefPubMed
  227. ↵
    Ross BM (2003) Phospholipase A2-associated processes in the human brain and their role in neuropathology and psychopathology, in Phospholipid Spectrum Disorders in Psychiatry and Neurology (Peet M, Glen L, and Horrobin DF eds) pp 163-182, Marius Press, Carnforth, Lancashire, UK.
  228. ↵
    Ross BM, Hudson C, Erlich J, Warsh JJ, and Kish SJ (1997) Increased phospholipid breakdown in schizophrenia— evidence for the involvement of a calcium-independent phospholipase A2. Arch Gen Psychiatry 54: 487-494.
    OpenUrlCrossRefPubMed
  229. ↵
    Ross BM, Kim DK, Bonventre JV, and Kish SJ (1995) Characterization of a novel phospholipase A2 activity in human brain. J Neurochem 64: 2213-2221.
    OpenUrlPubMed
  230. ↵
    Ross BM and Turenne SD (2002) Chronic cocaine administration reduces phospholipase A2 activity in rat brain striatum. Prostaglandins Leukotrienes Essent Fatty Acids 66: 479-483.
    OpenUrlCrossRefPubMed
  231. ↵
    Ross BM, Turenne S, Moszczynska A, Warsh JJ, and Kish SJ (1999) Differential alteration of phospholipase A2 activities in brain of patients with schizophrenia. Brain Res 821: 407-413.
    OpenUrlCrossRefPubMed
  232. ↵
    Rybakowski JK, Borkowska A, Czerski PM, Dmitrzak-Weglarz M, and Hauser J (2003) The study of cytosolic phospholipase A2 gene polymorphism in schizophrenia using eye movement disturbances as an endophenotypic marker. Neuropsychobiology 47: 115-119.
    OpenUrlCrossRefPubMed
  233. ↵
    Salvati S, Natali F, Attorri L, Raggi C, Di Biase A, and Sanchez M (2004) Stimulation of myelin proteolipid protein gene expression by eicosapentaenoic acid in C6 glioma cells. Neurochem Int 44: 331-338.
    OpenUrlCrossRefPubMed
  234. ↵
    Sandhya TL, Ong WY, Horrocks LA, and Farooqui AA (1998) A light and electron microscopic study of cytoplasmic phospholipase A2 and cyclooxygenase-2 in the hippocampus after kainate lesions. Brain Res 788: 223-231.
    OpenUrlCrossRefPubMed
  235. ↵
    Sano M, Ernesto C, Thomas RG, Klauber MR, Schafer K, Grundman M, Woodbury P, Growdon J, Cotman DW, Pfeiffer E, et al. (1997) A controlled trial of selegiline, α-tocopherol, or both as treatment for Alzheimer's disease. N Engl J Med 336: 1216-1222.
    OpenUrlCrossRefPubMed
  236. ↵
    Sapirstein A and Bonventre JV (2000) Phospholipases A2 in ischemic and toxic brain injury. Neurochem Res 25: 745-753.
    OpenUrlCrossRefPubMed
  237. ↵
    Schapira AH (1996) Oxidative stress and mitochondrial dysfunction in neurodegeneration. Curr Opin Neurol 9: 260-264.
    OpenUrlCrossRefPubMed
  238. ↵
    Scott KF, Graham GG, and Bryant KJ (2003) Secreted phospholipase A2 enzymes as therapeutic targets. Expert Opin Ther Targets 7: 427-440.
    OpenUrlCrossRefPubMed
  239. ↵
    Seashols SJ, del Castillo Olivares A, Gil G, and Barbour SE (2004) Regulation of group VIA phospholipase A2 expression by sterol availability. Biochim Biophys Acta 1684: 29-37.
    OpenUrlPubMed
  240. ↵
    Seno K, Okuno T, Nishi K, Murakami Y, Watanabe F, Matsuura T, Wada M, Fujii Y, Yamada M, Ogawa T, et al. (2000) Pyrrolidine inhibitors of human cytosolic phospholipase A2. J Med Chem 43: 1041-1044.
    OpenUrlCrossRefPubMed
  241. ↵
    Serhan CN, Gotlinger K, Hong S, and Arita M (2004) Resolvins, docosatrienes, and neuroprotectins, novel ω-3-derived mediators, and their aspirin-triggered endogenous epimers: an overview of their protective roles in catabasis. Prostaglandins Other Lipid Mediat 73: 155-172.
    OpenUrlCrossRefPubMed
  242. ↵
    Shanker G and Aschner M (2003) Methylmercury-induced reactive oxygen species formation in neonatal cerebral astrocytic cultures is attenuated by antioxidants. Mol Brain Res 110: 85-91.
    OpenUrlPubMed
  243. ↵
    Shanker G, Hampson RE, and Aschner M (2004) Methylmercury stimulates arachidonic acid release and cytosolic phospholipase A2 expression in primary neuronal cultures. Neurotoxicology 25: 399-406.
    OpenUrlCrossRefPubMed
  244. ↵
    Shin EJ, Jhoo JH, Kim WK, Jhoo WK, Lee C, Jung BD, and Kim HC (2004) Protection against kainate neurotoxicity by pyrrolidine dithiocarbamate. Clin Exp Pharmacol Physiol 31: 320-326.
    OpenUrlCrossRefPubMed
  245. ↵
    Shinzawa K and Tsujimoto Y (2003) PLA2 activity is required for nuclear shrinkage in caspase-independent cell death. J Cell Biol 163: 1219-1230.
    OpenUrlAbstract/FREE Full Text
  246. ↵
    Shirai Y and Ito M (2004) Specific differential expression of phospholipase A2 subtypes in rat cerebellum. J Neurocytol 33: 297-307.
    OpenUrlCrossRefPubMed
  247. ↵
    Shohami E, Shapira Y, Yadid G, Reisfeld N, and Yedgar S (1989) Brain phospholipase A2 is activated after experimental closed head injury in the rat. J Neurochem 53: 1541-1546.
    OpenUrlCrossRefPubMed
  248. ↵
    Singer TP, Castagnoli N Jr, Ramsay RR, and Trevor AJ (1987) Biochemical events in the development of parkinsonism induced by 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine. J Neurochem 49: 1-8.
    OpenUrlPubMed
  249. ↵
    Smalheiser NR, Dissanayake S, and Kapil A (1996) Rapid regulation of neurite outgrowth and retraction by phospholipase A2-derived arachidonic acid and its metabolites. Brain Res 721: 39-48.
    OpenUrlCrossRefPubMed
  250. ↵
    Song C, Chang XJ, Bean KM, Proia MS, Knopf JL, and Kriz RW (1999) Molecular characterization of cytosolic phospholipase A2-β. J Biol Chem 274: 17063-17067.
    OpenUrlAbstract/FREE Full Text
  251. ↵
    St-Gelais F, Ménard C, Congar P, Trudeau LE, and Massicotte G (2004) Postsynaptic injection of calcium-independent phospholipase A2 inhibitors selectively increases AMPA receptor-mediated synaptic transmission. Hippocampus 14: 319-325.
    OpenUrlCrossRefPubMed
  252. ↵
    Stephenson DT, Lemere CA, Selkoe DJ, and Clemens JA (1996) Cytosolic phospholipase A2 (cPLA2) immunoreactivity is elevated in Alzheimer's disease brain. Neurobiol Dis 3: 51-63.
    OpenUrlCrossRefPubMed
  253. ↵
    Stephenson D, Rash K, Smalstig B, Roberts E, Johnstone E, Sharp J, Panetta J, Little S, Kramer R, and Clemens J (1999) Cytosolic phospholipase A2 is induced in reactive glia following different forms of neurodegeneration. Glia 27: 110-128.
    OpenUrlCrossRefPubMed
  254. ↵
    Stewart LR, White AR, Jobling MF, Needham BE, Maher F, Thyer J, Beyreuther K, Masters CL, Collins SJ, and Cappai R (2001) Involvement of the 5-lipoxygenase pathway in the neurotoxicity of the prion peptide PrP106-126. J Neurosci Res 65: 565-572.
    OpenUrlCrossRefPubMed
  255. Stone SJ, Cui Z, and Vance JE (1998) Cloning and expression of mouse liver phosphatidylserine synthase-1 cDNA— overexpression in rat hepatoma cells inhibits the CDP-ethanolamine pathway for phosphatidylethanolamine biosynthesis. J Biol Chem 273: 7293-7302.
    OpenUrlAbstract/FREE Full Text
  256. ↵
    Street IP, Lin HK, Laliberté F, Ghomashchi F, Wang Z, Perrier H, Tremblay NM, Huang Z, Weech PK, and Gelb MH (1993) Slow- and tight-binding inhibitors of the 85-kDa human phospholipase A2. Biochemistry 32: 5935-5940.
    OpenUrlCrossRefPubMed
  257. ↵
    Strokin M, Sergeeva M, and Reiser G (2003) Docosahexaenoic acid and arachidonic acid release in rat brain astrocytes is mediated by two separate isoforms of phospholipase A2 and is differently regulated by cyclic AMP and Ca2+. Br J Pharmacol 139: 1014-1022.
    OpenUrlCrossRefPubMed
  258. ↵
    Sugaya K, Uz T, Kumar V, and Manev H (2000) New anti-inflammatory treatment strategy in Alzheimer's disease. Jpn J Pharmacol 82: 85-94.
    OpenUrlCrossRefPubMed
  259. ↵
    Sun GY, Xu JF, Jensen MD, and Simonyi A (2004) Phospholipase A2 in the central nervous system: implications for neurodegenerative diseases. J Lipid Res 45: 205-213.
    OpenUrlAbstract/FREE Full Text
  260. ↵
    Svennerholm L (1994) Gangliosides-a new therapeutic agent against stroke and Alzheimer's disease. Life Sci 55: 2125-2134.
    OpenUrlCrossRefPubMed
  261. ↵
    Tabuchi S, Uozumi N, Ishii S, Shimizu Y, Watanabe T, and Shimizu T (2003) Mice deficient in cytosolic phospholipase A2 are less susceptible to cerebral ischemia/reperfusion injury, in Brain Edema XII (Kuroiwa T, Baethmann A, Czernicki Z, Hoff JT, Ito U, Katayama Y, Mararou A, Mendelow AD, and Reulen HJ eds) pp 169-172, Springer-Verlag Wien, Vienna.
  262. ↵
    Tagami M, Ikeda K, Yamagata K, Nara Y, Fujino H, Kubota A, Numano F, and Yamori Y (1999) Vitamin E prevents apoptosis in hippocampal neurons caused by cerebral ischemia and reperfusion in stroke-prone spontaneously hypertensive rats. Lab Invest 79: 609-615.
    OpenUrlPubMed
  263. ↵
    Tariq M, Khan HA, Al Moutaery K, and Al Deeb S (2001) Protective effect of quinacrine on striatal dopamine levels in 6-OHDA and MPTP models of Parkinsonism in rodents. Brain Res Bull 54: 77-82.
    OpenUrlCrossRefPubMed
  264. ↵
    Taylor WA (1988) Effects of Impact Injury of Rat Spinal Cord on Activities of Some Enzymes of Lipid Hydrolysis. Doctoral Dissertation, The Ohio State University, Columbus, OH.
  265. ↵
    Terano T, Fujishiro S, Ban T, Yamamoto K, Tanaka T, Noguchi Y, Tamura Y, Yazawa K, and Hirayama T (1999) Docosahexaenoic acid supplementation improves the moderately severe dementia from thrombotic cerebrovascular diseases. Lipids 34 (Suppl): S345-S346.
    OpenUrlCrossRefPubMed
  266. ↵
    Thwin MM, Ong WY, Fong CW, Sato K, Kodama K, Farooqui AA, and Gopalakrishnakone P (2003) Secretory phospholipase A2 activity in the normal and kainate injected rat brain, and inhibition by a peptide derived from python serum. Exp Brain Res 150: 427-433.
    OpenUrlPubMed
  267. ↵
    Tong W, Shah D, Xu JF, Diehl JA, Hans A, Hannink M, and Sun GY (1999) Involvement of lipid mediators on cytokine signaling and induction of secretory phospholipase A2 in immortalized astrocytes (DITNC). J Mol Neurosci 12: 89-99.
    OpenUrlPubMed
  268. ↵
    Tran K, Wong JT, Lee E, Chan AC, and Choy PC (1996) Vitamin E potentiates arachidonate release and phospholipase A2 activity in rat heart myoblastic cells. Biochem J 319: 385-391.
    OpenUrlAbstract/FREE Full Text
  269. Trevisi L, Bova S, Cargnelli G, Ceolotto G, and Luciani S (2002) Endothelin-1-induced arachidonic acid release by cytosolic phospholipase A2 activation in rat vascular smooth muscle via extracellular signal-regulated kinases pathway. Biochem Pharmacol 64: 425-431.
    OpenUrlCrossRefPubMed
  270. ↵
    Trigueros SDA, Kalyvas A, and David S (2003) Phospholipase A2 plays an important role in myelin breakdown and phagocytosis during Wallerian degeneration. Mol Cell Neurosci 24: 753-765.
    OpenUrlCrossRefPubMed
  271. ↵
    Trimble LA, Street IP, Perrier H, Tremblay NM, Weech PK, and Bernstein MA (1993) NMR structural studies of the tight complex between a trifluoromethyl ketone inhibitor and the 85-kDa human phospholipase A2. Biochemistry 32: 12560-12565.
    OpenUrlCrossRefPubMed
  272. ↵
    Turnbull S, Tabner BJ, Brown DR, and Allsop D (2003) Quinacrine acts as an antioxidant and reduces the toxicity of the prion peptide PrP106-126. NeuroReport 14: 1743-1745.
    OpenUrlCrossRefPubMed
  273. ↵
    Uozumi N and Shimizu T (2002) Roles for cytosolic phospholipase A2α as revealed by gene-targeted mice. Prostaglandins Other Lipid Mediat 68-69: 59-69.
    OpenUrlCrossRefPubMed
  274. ↵
    Vanags DM, Larsson P, Feltenmark S, Jakobsson PJ, Orrenius S, Claesson HE, and Aguilar-Santelises M (1997) Inhibitors of arachidonic acid metabolism reduce DNA and nuclear fragmentation induced by TNF plus cycloheximide in U937 cells. Cell Death Differ 4: 479-486.
    OpenUrlCrossRefPubMed
  275. ↵
    Visioli F, Rodriguez de Turco EB, Kreisman NR, and Bazan NG (1994) Membrane lipid degradation is related to interictal cortical activity in a series of seizures. Metab Brain Dis 9: 161-170.
    OpenUrlCrossRefPubMed
  276. ↵
    Vogtherr M, Grimme S, Elshorst B, Jacobs DM, Fiebig K, Griesinger C, and Zahn R (2003) Antimalarial drug quinacrine binds to C-terminal helix of cellular prion protein. J Med Chem 46: 3563-3564.
    OpenUrlCrossRefPubMed
  277. ↵
    Walsh DM and Selkoe DJ (2004) Deciphering the molecular basis of memory failure in Alzheimer's disease. Neuron 44: 181-193.
    OpenUrlCrossRefPubMed
  278. ↵
    Walters ET (1994) Injury-related behavior and neuronal plasticity: an evolutionary perspective on sensitization, hyperalgesia, and analgesia. Int Rev Neurobiol 36: 325-427.
    OpenUrlPubMed
  279. ↵
    Wang TG, Qin L, Liu B, Liu YX, Wilson B, Eling TE, Langenbach R, Taniura S, and Hong JS (2004a) Role of reactive oxygen species in LPS-induced production of prostaglandin E-2 in microglia. J Neurochem 88: 939-947.
    OpenUrlCrossRefPubMed
  280. ↵
    Wang XH, Yan GT, Wang LH, Hao XH, Zhang K, and Xue H (2004b) The mediating role of cPLA2 in IL-1β and IL-6 release in LPS-induced HeLa cells. Cell Biochem Funct 22: 41-44.
    OpenUrlCrossRefPubMed
  281. ↵
    Wang ZG (2003) Role of redox state in modulation of ion channel function by fatty acids and phospholipids. Br J Pharmacol 139: 681-683.
    OpenUrlCrossRefPubMed
  282. ↵
    Weerasinghe GR, Rapoport SI, and Bosetti F (2004) The effect of chronic lithium on arachidonic acid release and metabolism in rat brain does not involve secretory phospholipase A2 or lipoxygenase/cytochrome P450 pathways. Brain Res Bull 63: 485-489.
    OpenUrlCrossRefPubMed
  283. ↵
    Wei EP, Lamb RG, and Kontos HA (1982) Increased phospholipase C activity after experimental brain injury. J Neurosurg 56: 695-698.
    OpenUrlPubMed
  284. ↵
    Wei S, Ong WY, Thwin MM, Fong CW, Farooqui AA, Gopalakrishnakone P, and Hong WJ (2003) Differential activities of secretory phospholipase A2 (sPLA2)inrat brain and effects of sPLA2 on neurotransmitter release. Neuroscience 121: 891-898.
    OpenUrlCrossRefPubMed
  285. ↵
    Wells K, Farooqui AA, Liss L, and Horrocks LA (1995) Neural membrane phospholipids in Alzheimer disease. Neurochem Res 20: 1329-1333.
    OpenUrlCrossRefPubMed
  286. ↵
    Wolf MJ, Izumi Y, Zorumski CF, and Gross RW (1995) Long-term potentiation requires activation of calcium-independent phospholipase A2. FEBS Lett 377: 358-362.
    OpenUrlCrossRefPubMed
  287. Xing M, Wilkins PL, McConnell BK, and Mattera R (1994) Regulation of phospholipase A2 activity in undifferentiated and neutrophil-like HL60 cells: linkage between impaired responses to agonists and absence of protein kinase C-dependent phosphorylation of cytosolic phospholipase A2. J Biol Chem 269: 3117-3124.
    OpenUrlAbstract/FREE Full Text
  288. ↵
    Xu JF, Yu S, Sun AY, and Sun GY (2003) Oxidant-mediated AA release from astrocytes involves cPLA2 and iPLA2. Free Radic Biol Med 34: 1531-1543.
    OpenUrlCrossRefPubMed
  289. ↵
    Yagi K, Shirai Y, Hirai M, Sakai N, and Saito N (2004) Phospholipase A2 products retain a neuron specific γ isoform of PKC on the plasma membrane through the C1 domain—a molecular mechanism for sustained enzyme activity. Neurochem Int 45: 39-47.
    OpenUrlCrossRefPubMed
  290. ↵
    Yang H-C, Farooqui AA, and Horrocks LA (1994a) Effects of glycosaminoglycans and glycosphingolipids on cytosolic phospholipases A2 from bovine brain. Biochem J 299: 91-95.
    OpenUrlAbstract/FREE Full Text
  291. ↵
    Yang H-C, Farooqui AA, and Horrocks LA (1994b) Effects of sialic acid and sialoglycoconjugates on cytosolic phospholipases A2 from bovine brain. Biochem Biophys Res Commun 199: 1158-1166.
    OpenUrlCrossRefPubMed
  292. ↵
    Yang HC, Mosior M, Ni B, and Dennis EA (1999) Regional distribution, ontogeny, purification, and characterization of the Ca2+-independent phospholipase A2 from rat brain. J Neurochem 73: 1278-1287.
    OpenUrlCrossRefPubMed
  293. ↵
    Yao JK and Van Kammen DP (2004) Membrane phospholipids and cytokine interaction in schizophrenia, in Disorders of Synaptic Plasticity and Schizophrenia (Smythies J ed) pp 297-326, Academic Press, San Diego, CA.
  294. ↵
    Yao JK, Leonard S, and Reddy RD (2000) Membrane phospholipid abnormalities in postmortem brains from schizophrenic patients. Schizophr Res 42: 7-17.
    OpenUrlCrossRefPubMed
  295. ↵
    Yegin A, Akbas SH, Ozben T, and Korgun DK (2002) Secretory phospholipase A2 and phospholipids in neural membranes in an experimental epilepsy model. Acta Neurol Scand 106: 258-262.
    OpenUrlCrossRefPubMed
  296. ↵
    Yeo JF, Ong WY, Ling SF, and Farooqui AA (2004) Intracerebroventricular injection of phospholipases A2 inhibitors modulates allodynia after facial carrageenan injection in mice. Pain 112: 148-155.
    OpenUrlCrossRefPubMed
  297. ↵
    Yoo MH, Woo CH, You HJ, Cho SH, Kim BC, Choi JE, Chun JS, Jhun BH, Kim TS, and Kim JH (2001) Role of the cytosolic phospholipase A2-linked cascade in signaling by an oncogenic, constitutively active Ha-Ras isoform. J Biol Chem 276: 24645-24653.
    OpenUrlAbstract/FREE Full Text
  298. ↵
    Yoshihara Y and Watanabe Y (1990) Translocation of phospholipase A2 from cytosol to membranes in rat brain induced by calcium ions. Biochem Biophys Res Commun 170: 484-490.
    OpenUrlCrossRefPubMed
  299. ↵
    Yoshihara Y, Yamaji M, Kawasaki M, and Watanabe Y (1992) Ontogeny of cytosolic phospholipase A2 activity in rat brain. Biochem Biophys Res Commun 185: 350-355.
    OpenUrlCrossRefPubMed
  300. ↵
    Yoshinaga N, Yasuda Y, Murayama T, and Nomura Y (2000) Possible involvement of cytosolic phospholipase A2 in cell death induced by 1-methyl-4-phenylpyridinium ion, a dopaminergic neurotoxin, in GH3 cells. Brain Res 855: 244-251.
    OpenUrlCrossRefPubMed
  301. ↵
    Zanassi P, Paolillo M, and Schinelli S (1998) Coexpression of phospholipase A2 isoforms in rat striatal astrocytes. Neurosci Lett 247: 83-86.
    OpenUrlCrossRefPubMed
  302. ↵
    Zhang B, Tanaka J, Yang L, Yang L, Sakanaka M, Hata R, Maeda N, and Mitsuda N (2004) Protective effect of vitamin E against focal brain ischemia and neuronal death through induction of target genes of hypoxia-inducible factor-1. Neuroscience 126: 433-440.
    OpenUrlCrossRefPubMed
  303. ↵
    Zhao Y, Joshi-Barve S, Barve S, and Chen LH (2004) Eicosapentaenoic acid prevents LPS-induced TNF-α expression by preventing NF-κB activation. J Am Coll Nutr 23: 71-78.
    OpenUrlCrossRefPubMed
PreviousNext
Back to top

In this issue

Pharmacological Reviews: 58 (3)
Pharmacological Reviews
Vol. 58, Issue 3
1 Sep 2006
  • Table of Contents
  • Table of Contents (PDF)
  • About the Cover
  • Index by author
  • Back Matter (PDF)
  • Editorial Board (PDF)
  • Front Matter (PDF)
Download PDF
Article Alerts
Sign In to Email Alerts with your Email Address
Email Article

Thank you for sharing this Pharmacological Reviews article.

NOTE: We request your email address only to inform the recipient that it was you who recommended this article, and that it is not junk mail. We do not retain these email addresses.

Enter multiple addresses on separate lines or separate them with commas.
Inhibitors of Brain Phospholipase A2 Activity: Their Neuropharmacological Effects and Therapeutic Importance for the Treatment of Neurologic Disorders
(Your Name) has forwarded a page to you from Pharmacological Reviews
(Your Name) thought you would be interested in this article in Pharmacological Reviews.
CAPTCHA
This question is for testing whether or not you are a human visitor and to prevent automated spam submissions.
Citation Tools
OtherReview Article

Inhibitors of Brain Phospholipase A2 Activity: Their Neuropharmacological Effects and Therapeutic Importance for the Treatment of Neurologic Disorders

Akhlaq A. Farooqui, Wei-Yi Ong and Lloyd A. Horrocks
Pharmacological Reviews September 1, 2006, 58 (3) 591-620; DOI: https://doi.org/10.1124/pr.58.3.7

Citation Manager Formats

  • BibTeX
  • Bookends
  • EasyBib
  • EndNote (tagged)
  • EndNote 8 (xml)
  • Medlars
  • Mendeley
  • Papers
  • RefWorks Tagged
  • Ref Manager
  • RIS
  • Zotero

Share
OtherReview Article

Inhibitors of Brain Phospholipase A2 Activity: Their Neuropharmacological Effects and Therapeutic Importance for the Treatment of Neurologic Disorders

Akhlaq A. Farooqui, Wei-Yi Ong and Lloyd A. Horrocks
Pharmacological Reviews September 1, 2006, 58 (3) 591-620; DOI: https://doi.org/10.1124/pr.58.3.7
Reddit logo Twitter logo Facebook logo Mendeley logo
  • Tweet Widget
  • Facebook Like
  • Google Plus One

Jump to section

  • Article
    • Abstract
    • I. Introduction
    • II. Multiplicity of Phospholipases A2 in Brain
    • III. Involvement of Phospholipase A2 Activity in Brain Injury
    • IV. Physiological and Pharmacological Effects of Phospholipase A2 Inhibitors
    • V. Phospholipase A2 Activity in Kainic Acid-Induced Neural Cell Injury
    • VI. Protection of Kainic Acid-Induced Neural Cell Injury by Phospholipase A2 Inhibitors
    • VII. Phospholipase A2 Activity in Neurological Disorders
    • VIII. Use of Phospholipase A2 Inhibitors for the Treatment of Neurological Disorders
    • IX. Prevention of Pain by Phospholipase A2 Inhibitors
    • X. Perspective and Direction for Future Studies
    • Acknowledgments
    • Footnotes
    • References
  • Figures & Data
  • Info & Metrics
  • eLetters
  • PDF

Related Articles

Cited By...

More in this TOC Section

  • Review of Natural Language Processing in Pharmacology
  • PHARMACOGENOMICS: Driving Personalized Medicine
  • Potassium Channels in Parkinson's Disease
Show more Review Articles

Similar Articles

Advertisement
  • Home
  • Alerts
Facebook   Twitter   LinkedIn   RSS

Navigate

  • Current Issue
  • Latest Articles
  • Archive
  • Search for Articles
  • Feedback
  • ASPET

More Information

  • About Pharmacological Reviews
  • Editorial Board
  • Instructions to Authors
  • Submit a Manuscript
  • Customized Alerts
  • RSS Feeds
  • Subscriptions
  • Permissions
  • Terms & Conditions of Use

ASPET's Other Journals

  • Drug Metabolism and Disposition
  • Journal of Pharmacology and Experimental Therapeutics
  • Molecular Pharmacology
  • Pharmacology Research & Perspectives
ISSN 1521-0081 (Online)

Copyright © 2023 by the American Society for Pharmacology and Experimental Therapeutics