thematic review series Thematic Review Series: Phospholipases: Central Role in Lipid Signaling and Disease

Cytosolic phospholipase A2: physiological function and role in disease Christina C. Leslie1 Department of Pediatrics, National Jewish Health, Denver, CO 80206; and Departments of Pathology and Pharmacology, University of Colorado Denver, Aurora, CO 80045

Supplementary key words arachidonic acid • calcium • Golgi apparatus • inflammation • cyclooxygenase • leukotrienes • lipoxygenase • prostaglandins • protein kinases/MAP kinase

OVERVIEW OF THE MAMMALIAN PHOSPHOLIPASE A2 FAMILY The mammalian phospholipase A2 (PLA2) family is comprised of six types of diverse enzymes: GIV PLA2 [cytosolic PLA2 (cPLA2)], GVI PLA2 [calcium-independent

This work was supported by National Institutes of Health Grant HL34303. Manuscript received 11 January 2015 and in revised form 26 March 2015. Published, JLR Papers in Press, April 2, 2015 DOI 10.1194/jlr.R057588

PLA2 (iPLA2)], GVII and GVIII PLA2 [platelet-activating factor-acetylhydrolases (PAF-AHs)], GXV PLA2 (lysosomal PLA2), GXVI PLA2 (adipose PLA2), and several groups of secreted PLA2 (sPLA2). Several reviews have described in detail the structure, activities, and known functions of the different groups of PLA2 enzymes, many of which contain several subtypes (1–8). Most of these enzymes have no structural similarity, and have different catalytic mechanisms and regulation. cPLA2s are most similar to iPLA2s (also classified as patatin-like enzymes), because they utilize an active site Ser-Asp dyad for catalysis, unlike the catalytic mechanism of other PLA2s that involves a His residue (2–4). It has been shown that the active site of the plant lipid acylhydrolase patatin contains a Ser-Asp dyad with structural similarity to cPLA2 and has homology with the iPLA2 family (4, 9). Although they have only one active site, cPLA2s and iPLA2s exhibit multiple enzymatic activities, including PLA2, PLA1, lysophospholipase, and transacylase activities, and some iPLA2s are lipases (2, 10). Many types of mammalian PLA2 enzymes can release arachidonic acid for eicosanoid production (2). An intriguing observation is that different types of PLA2s can cooperate in cells (sPLA2s with either cPLA2 or iPLA2) in promoting arachidonic acid release, although the mechanism is not understood (1). Recent studies have shown specialized roles of several sPLA2 enzymes in mediating production of eicosanoids that regulate mast cell maturation, asthma, host defense, and metabolic disorders, as detailed in a recent review (11). However, cPLA2 is the only PLA2 that

Abbreviations: COX, cyclooxygenase; cPLA2, cytosolic phospholipase A2; DSS, dextran sodium sulfate; EAE, experimental autoimmune encephalomyelitis; ER, endoplasmic reticulum; FLAP, 5-lipoxygenaseactivating protein; GI, gastrointestinal; iPLA2, calcium-independent PLA2; 5-LO, 5-lipoxygenase; LPA, lysophosphatidic acid; LTB4, leukotriene B4; mPGES-1, microsomal prostaglandin E synthase-1; NOD, nonobese diabetic; OVA, ovalbumin; PAF-AH, platelet-activating factoracetylhydrolase; PG, prostaglandin; PLA2, phospholipase A2; SCI, spinal cord injury; sPLA2, secreted PLA2; TX, thromboxane. 1 To whom correspondence should be addressed. e-mail: [email protected] Copyright © 2015 by the American Society for Biochemistry and Molecular Biology, Inc.

1386

Journal of Lipid Research Volume 56, 2015

This article is available online at http://www.jlr.org

Downloaded from www.jlr.org by guest, on August 13, 2015

Abstract The group IV phospholipase A2 (PLA2) family is comprised of six intracellular enzymes (GIVA, -B, -C, -D, -E, and -F) commonly referred to as cytosolic PLA2 (cPLA2)␣, -␤, -␥, -␦, -␧, and -␨. They contain a Ser-Asp catalytic dyad and all except cPLA2␥ have a C2 domain, but differences in their catalytic activities and subcellular localization suggest unique regulation and function. With the exception of cPLA2␣, the focus of this review, little is known about the in vivo function of group IV enzymes. cPLA2␣ catalyzes the hydrolysis of phospholipids to arachidonic acid and lysophospholipids that are precursors of numerous bioactive lipids. The regulation of cPLA2␣ is complex, involving transcriptional and posttranslational processes, particularly increases in calcium and phosphorylation. cPLA2␣ is a highly conserved widely expressed enzyme that promotes lipid mediator production in human and rodent cells from a variety of tissues. The diverse bioactive lipids produced as a result of cPLA2␣ activation regulate normal physiological processes and disease pathogenesis in many organ systems, as shown using cPLA2␣ KO mice. However, humans recently identified with cPLA2␣ deficiency exhibit more pronounced effects on health than observed in mice lacking cPLA2␣, indicating that much remains to be learned about this interesting enzyme.—Leslie, C. C. Cytosolic phospholipase A2: physiological function and role in disease. J. Lipid Res. 2015. 56: 1386–1402.

exhibits preference for hydrolysis of arachidonic acid from phospholipid substrates that occurs in cells stimulated with diverse agonists (4). Most cells contain several types of PLA2 enzymes and studies have suggested that they have distinct functions. For example, a comparison of the function of cPLA2 and iPLA2 in mast cells and macrophages found that cPLA2, but not iPLA2, selectively hydrolyzes arachidonic acid-containing phospholipids in response to cell stimulation (12, 13).

transport (14, 19, 20). This is supported by a recent study showing that cPLA2 regulates formation of tubules involved in clathrin-independent endocytic trafficking (21). Therefore in studies designed to understand the role of cPLA2 in regulating cellular processes, it is important to consider that it has a functional role independent of eicosanoid production.

GROUP IVA cPLA2 (cPLA2) GROUP IV PLA2 FAMILY

REGULATION OF cPLA2 cPLA2 is widely expressed in cells throughout all tissues in mice and humans. It is a highly conserved enzyme with mouse and human sharing 95% amino acid identity, suggesting similarities in regulation and function (31). In addition to PLA2 activity, cPLA2 catalyzes other enzymatic reactions through the Ser-Asp catalytic dyad, including PLA1, lysophospholipase, and transacylase activity, although the physiological relevance of these activities in vivo is cPLA2: physiological function and role in disease

1387

Downloaded from www.jlr.org by guest, on August 13, 2015

The six members of the cPLA2 family (cPLA2, -, -, -, -, and -) share only about 30% homology, and have differences in enzymatic properties, tissue expression, and subcellular localization, suggesting that they are not redundant (4, 14). Analysis of the interfacial kinetic and binding properties of the cPLA2 family showed that they exhibit very different relative lysophospholipase, PLA2, and PLA1 activities, inhibitor sensitivities, calcium dependence, and activation by anionic phospholipids, highlighting potential differences in regulation and function (10). With the exception of the established role of cPLA2 in initiating production of lipid mediators, little is known about the in vivo function of other cPLA2 isoforms. However, there is evidence emerging from studies using cultured cells that members of the cPLA2 family, including cPLA2, play a role in regulating membrane trafficking that does not involve production of oxygenated metabolites of arachidonic acid. PLA2 (and PLA1) enzymes are implicated in regulating intracellular membrane trafficking by participating in the formation of transport carriers from donor membrane (15). Membrane budding requires alterations in membrane curvature that can occur by formation of lysophospholipids. Four cytoplasmic PLAs, including cPLA2, PAF-AH Ib, iPLA2, and iPLA1, are implicated in the formation of tubules from the endoplasmic reticulum (ER)-Golgi intermediate compartment and the Golgi apparatus, maintenance of Golgi structure, and transport through the Golgi. The details of the role of these enzymes have recently been reviewed (15, 16). cPLA2 regulates intra-Golgi transport by mediating the formation of tubules that connect the Golgi stacks (17). The localization of cPLA2 at the Golgi in endothelial cells has also been shown to regulate the transport of junction proteins (vascular endothelial-cadherin, occludin, claudin-5) from the Golgi to cell-cell junctions (18). In these studies it was shown that the enzymatic activity of cPLA2 is required for regulating transport, but oxygenated metabolites of arachidonic acid are not involved. Considering that these are fundamental processes important for cell function, it is surprising that the cPLA2 KO mouse does not exhibit any major phenotype. Golgi transport was found to be normal in cells deficient in cPLA2, but results of a siRNA screen suggested that GVIII PLA2 (PAF-AH) compensates for the loss of cPLA2 in regulating Golgi transport (17). Other group IV PLA2s have been shown to localize to endocytic vesicles, suggesting a possible role in regulating

It has been over 25 years since a soluble calcium-regulated high-molecular-weight PLA2 that exhibited arachidonic acid selectivity was identified in cells and purified (22–29). Cloning of PLA2 revealed the C2 domain for calcium regulation (30, 31). The identification of the active site residues, characterization of stimulus-dependent phosphorylation by MAPKs, and structural elucidation then followed (32–36). Over the years, we have learned a great deal about the regulation of cPLA2 and its role in initiating the production of bioactive lipid mediators. cPLA2 activation results in the production of diverse lipid mediators derived from its products, arachidonic acid and lysophospholipids. Leukotrienes are potent pro-inflammatory mediators produced by cells involved in the inflammatory response, particularly neutrophils, macrophages, and mast cells (37). Prostaglandins (PGs) are made by many cell types and have very diverse functions. In addition to their role in regulating homeostatic processes, they can promote all the cardinal signs of acute inflammation and participate in the maintenance of chronic inflammation (38, 39). However they facilitate the resolution of inflammation by altering the balance of pro-inflammatory and antiinflammatory cytokines, and by enhancing clearance of apoptotic neutrophils (38, 40–43). cPLA2 is also implicated in the production of the pro-resolving lipid mediators (lipoxins) and bioactive lysophospholipids (or derived from lysophospholipids) such as platelet-activating factor and lysophosphatidylinositol (44–46). The generation of mice deficient in cPLA2 has revealed its role in regulating normal physiological processes and disease pathogenesis in rodents (47, 48). More recently, the identification of humans with cPLA2 deficiency has emphasized its important role in human health (49–51). This review will provide a brief overview of the regulation of cPLA2 and primarily focus on reviewing its function in specific organs of mice and humans.

analysis of the human cPLA2 gene (73). This IL-1responsive distal regulatory element contains an AP-1 site, which regulates both basal and IL-1-induced expression of cPLA2. Binding of c-Jun to this site acts as a repressor in the basal state and an activator in response to IL-1, and an important role for C/EBP as a transcriptional activator of cPLA2 was demonstrated. cPLA2 is rapidly activated in cells by posttranslational processes, including increases in intracellular calcium and phosphorylation by MAPK (Fig. 1). These signaling pathways are activated in cells through engagement of many types of receptors, indicating that cPLA2 activation and arachidonic acid release occur commonly in response to cell stimulation. There have been a number of reviews that detail the regulation of cPLA2 by posttranslational processes (2–4, 74). In brief, increases in intracellular calcium promote the translocation of cPLA2 from the cytosol to intracellular membrane (75–78). Calcium binds to the Nterminal C2 domain of cPLA2 that increases the hydrophobicity of the calcium binding loops, which penetrate the membrane bilayer (79–83) (Fig. 2). Differences in phospholipid binding specificity of the C2 domains of protein kinase C (anionic phospholipids) and cPLA2 (phosphatidylcholine) play an important role in determining

Fig. 1. Regulation of cPLA2. cPLA2-mediated arachidonic acid (AA) release occurs following cell stimu2+ lation that results in increases in intracellular Ca and kinase activation. In the cytosol, cPLA2 is bound to p11/annexin A2 and is released from this complex when phosphorylated on S727. Calcium binds to the calcium binding loops (CBL, green) at the membrane binding face of the C2 domain. This increases the hydrophobicity of the C2 domain and promotes preferential binding of cPLA2 to the Golgi and, at higher levels of intracellular calcium, to the ER and nuclear envelope. Basic residues (orange patch) in the C2 domain are implicated in binding to ceramide-1-P in the Golgi. Association of the catalytic domain with membrane is mediated in part by a tryptophan residue in the catalytic domain (W464, red patch) that stabilizes cPLA2 on the membrane. Phosphorylation of S505 by MAPKs in the catalytic domain enhances cPLA2 activity. Interaction of basic residues in the catalytic domain (blue patch) with anionic components in the membrane, such as polyphosphoinositides, optimizes catalytic activity. AA released by cPLA2 is metabolized by COX-1 and -2, which are primarily localized to the ER, but also occur on the Golgi. Leukotriene production occurs at the nuclear envelope, the site of 5-LO translocation in response to increases in calcium. 5-LO binds to the scaffold protein FLAP, an AA binding protein, which also binds LTC4 synthase (syn) for production of LTC4.

1388

Journal of Lipid Research Volume 56, 2015

Downloaded from www.jlr.org by guest, on August 13, 2015

unknown (2, 4, 32, 35). The intense focus on cPLA2 stems from its preferential hydrolysis of sn-2 arachidonic acid, and its well-established role in initiating the release of arachidonic acid for the production of lipid mediators (52). cPLA2 KO mice have provided a source for cells, including (but not limited to) mast cells, neutrophils, macrophages, platelets, endothelial cells, and lung fibroblasts, to establish a role for cPLA2 in mediating arachidonic acid release and production of lipid mediators (47, 53–59). cPLA2 is regulated at the transcriptional level and by posttranslational mechanisms. It is basally expressed in many cells, due in part to transcription factor II D binding to the TATA-less promoter (60–62). cPLA2 expression is also induced transcriptionally through a number of signaling pathways involving Ras and MAPKs, and transcriptional activators NF-B, Krüppel-like factor, hypoxiainducible factor, Sp1, and c-Jun have been described (63–70). Pro-inflammatory cytokines induce expression of human cPLA2 mRNA that is blocked by glucocorticoids (71). However, glucocorticoids are not universally suppressive because they paradoxically enhance cPLA2 expression in human amnion fibroblasts (see Reproduction section below) (72). Recently the first cytokine-dependent enhancer element has been identified by DNase I hypersensitive site

their distinct subcellular targeting to the plasma membrane and intracellular membranes, respectively (84–86). The catalytic domain is then positioned on the membrane by calcium-independent mechanisms, in part through a tryptophan residue (W464) that stabilizes membrane binding (87–90). cPLA2 preferentially translocates to the Golgi apparatus and, at higher intracellular calcium concentrations, to the ER and nuclear envelope (76–78, 91, 92) (Fig. 1). Binding of cPLA2 to the Golgi may, in part, involve interaction of the C2 domain with sphingolipids, including ceramide-1-P and lactosylceramide (93–96). Translocation of cPLA2 to membrane is necessary but not sufficient for cPLA2 to release arachidonic acid (92). The catalytic activity of cPLA2 is enhanced by phosphorylation of S505 by the MAPKs, extracellular regulated protein kinases, and p38 (33, 34, 97, 98). More recent studies suggest that cPLA2 is also activated by c-jun N-terminal kinases (99, 100). Activation of cPLA2 in vascular smooth muscle cells by calcium/calmodulin-dependent kinase II that phosphorylates S515 has been described (101). Phosphorylation of cPLA2 on S727 by MAPK-interacting kinase does not enhance catalytic activity per se, but blocks its binding to an inhibitory complex in the cytosol composed of p11(S100A10)/annexin A2 (74, 102–105) (Fig. 1). In addition to phosphorylation, the activity of cPLA2 is regulated by basic residues in the catalytic domain that are the site for activation by polyphosphoinositides (89, 106– 108). The basic residues are required for the ability of cPLA2 to release arachidonic acid in cells but are not required for translocation of cPLA2 to the Golgi (92). Collectively, the

results show that the ability of cPLA2 to release arachidonic acid in cells involves calcium-dependent membrane binding and optimization of catalytic activity by phosphorylation and interaction of basic residues in the catalytic domain with anionic components, perhaps polyphosphoinositides, in the membrane (92). The translocation of cPLA2 to intracellular membranes raises the interesting question of how cPLA2-derived arachidonic acid couples to the downstream enzymes, 5-lipoxygenase (5-LO) and cyclooxygenase (COX)-1 and -2 (Fig. 1). There are general similarities in the regulation of cPLA2 and 5-LO involving both calcium and phosphorylation (109). 5-LO, a soluble enzyme (cytoplasmic or intranuclear) in resting cells, translocates to the nuclear envelope in response to calcium, which binds to an N-terminal C2like domain (110–115). Leukotriene production requires interaction of 5-LO with 5-LO-activating protein (FLAP), an integral nuclear envelope protein that functions as an arachidonic acid binding protein (116). It has recently been shown that leukotriene synthesis involves activationdependent formation of supramolecular complexes on the nuclear envelope composed of FLAP with 5-LO and FLAP with leukotriene C4 synthase (117, 118). In addition, cPLA2-derived arachidonic acid was shown to affect the conformation of FLAP and enhance its ability to recruit 5-LO for complex formation (119). It is not known whether cPLA2 physically associates with these complexes or whether arachidonic acid released in their vicinity diffuses through the membrane to initiate leukotriene production. cPLA2: physiological function and role in disease

1389

Downloaded from www.jlr.org by guest, on August 13, 2015

Fig. 2. Ribbon diagram of cPLA2 structure. Two ions of calcium bind to the calcium-binding loops (CBL) on the membrane binding face of the C2 domain. The catalytic domain with the active site nucleophile S228 contains residues involved in cPLA2 regulation (pink) that include phosphorylation sites (S505, S727), basic residues (K488, K544, K543), and W464 important for membrane interaction. The residues mutated in humans with cPLA2 deficiencies are shown in yellow. The residue V707 is predicted to be the start of the deletion in patients with cryptogenic multifocal ulcerating stenosing enteritis.

ROLE OF cPLA2 IN NORMAL PHYSIOLOGICAL PROCESSES AND DISEASE The first characterization of mice with cPLA2 gene disruption published in 1997 was followed by a large body of work that provides a comprehensive view of cPLA2 function in normal processes and disease pathogenesis in mice (47, 48). The cPLA2 KO mouse develops normally and lives a normal life span with no overt adverse health effects. Considering the evolutionary conservation of cPLA2 and its role in eicosanoid production, similarities in the regulation and function of cPLA2 in rodents and humans is expected. However, as discussed in this review, there are deleterious effects on the health of humans with cPLA2 deficiency, suggesting differences in the role of cPLA2 and eicosanoids in mouse models and humans. This could be due to a number of factors, including differences in expression of other types of PLA2 enzymes that could compensate and differences in the tissue-specific expression of receptors for lipid mediators that could influence their function. In many cases, the mechanisms responsible for observed phenotypes in human and mouse cPLA2 deficiencies are not well understood. This is understandable considering that cPLA2 is the first regulatory enzyme that releases the lipid mediators, arachidonic acid and lysophospholipids, which are precursors for an enormous number of diverse bioactive lipid metabolites that can have opposing effects. Gastrointestinal tract The consequences of inherited cPLA2 deficiency in several human patients have now been described (49–51). 1390

Journal of Lipid Research Volume 56, 2015

In all of these patients, a common clinical finding is abnormalities in the gastrointestinal (GI) tract. However, there are differences in the clinical presentations and severity of disease. In one subject, two rare heterozygous single bp mutations in PLA2G4A were found; one mutation (S111P) occurred in the C2 domain of cPLA2 and the other allele had a mutation in the catalytic domain (R485H) (Fig. 2) (49). The expression of cPLA2 protein and PLA2 activity in platelet lysates was reduced compared with normals. The patient exhibited a global reduction in eicosanoid production with reduced urinary metabolites of thromboxane (TX)A2, PGI2, PGD2, and PGE2, and undetectable leukotriene E4 in urine. The production of leukotriene B4 (LTB4) from leukocytes in A23187-stimulated blood was reduced by 97%. The patient had occult GI blood loss and anemia in childhood with more acute GI bleeding in his fourth decade with recurrent ulcerations in the ileum and jejunum of the small intestine. The beneficial effect of misoprostol suggests that cPLA2-mediated PGE2 production is important for function of the small intestine. Unlike the effect of nonsteroid anti-inflammatory drug toxicity, the patient did not exhibit gastroduodenal ulcerations, and the colon was normal. The results support a role for cPLA2 in providing arachidonic acid for a large proportion of eicosanoid production in humans. Other PLA2s may contribute to the residual eicosanoids produced, or the patient may retain a low level of cPLA2 activity. To investigate the effect of the mutations on cPLA2 function in more detail, we compared the catalytic activity of WT and mutant forms of purified cPLA2 in vitro, and compared their cellular function by expression in cells lacking endogenous cPLA2 (124). cPLA2 with the S111P mutation showed slightly reduced catalytic activity and interfacial binding in vitro at saturating calcium. When expressed in cells, there was less translocation of cPLA2 with the S111P mutation to the Golgi in response to serum stimulation and reduced arachidonic acid release. In contrast, cPLA2 with the R485H mutation translocated to the Golgi in serum-stimulated cells, but failed to release arachidonic acid and was inactive in vitro. Consequently, the mutations in the C2 and catalytic domains of cPLA2 compromise function, but do not result in a complete loss of protein or activity. Another patient, and his twin sister, with inherited cPLA2 deficiency also exhibited GI abnormalities from early childhood characterized by duodenal ulcers that were corrected by misoprostol treatment (51). The incidence of small intestinal ulcers was not evaluated. Homozygous bp mutations (D575H) in the cPLA2 catalytic domain were identified (Fig. 2). cPLA2 protein was not detected in lysates from patients with the homozygous mutations. Unlike the patients described above, the mutations resulted in global diathesis (see Hemostasis section below). The most severe clinical manifestation of inherited cPLA2 deficiency occurred in siblings diagnosed with cryptogenic multifocal ulcerating stenosing enteritis due to homozygous deletion mutations in PLA2G4A (50). The affected male had severe peptic ulcers and bleeding beginning at an early age (4 years) then over the next several years

Downloaded from www.jlr.org by guest, on August 13, 2015

COX-1 and -2 are also found in internal membranes, but unlike cPLA2 and 5-LO, they are integral membrane proteins primarily localized in the ER (120). However, it was recently shown that COX-2 localizes to the Golgi in cancer cell lines detected by using a COX-2-specific fluorescent probe (121). Another study investigating the posttranslational processing and degradation of COX-2 found catalytically active COX-2 (but not COX-1) localized to the Golgi as a result of anterograde trafficking from the ER, a step required for posttranslational modification prior to COX-2 degradation through the ER-associated degradation pathway (122). It is interesting that microsomal PGE synthase-1 (mPGES-1) was also found at the Golgi, suggesting that this site has a system for PGE2 production involving cPLA2/COX-2/mPGES-1 (122). Another report showed that COX-1, but not COX-2, is constitutively found at the Golgi in airway epithelial cells, where it colocalizes with cPLA2 (123). The localization of cPLA2, COX, mPGES-1, and leukotriene synthetic enzymes at intracellular membranes, including the ER, nuclear envelope, and Golgi, may efficiently coordinate the production of certain eicosanoids. However, several downstream enzymes involved in PG and leukotriene production are cytosolic, and not apparently subject to stimulus-dependent translocation, although the sub-cellular organization of enzymes involved in eicosanoid production is only beginning to be elucidated.

against colon cancer in mice induced by the colon carcinogen, azoxymethane, which induces increased tumor multiplicity in heterozygous and KO mice compared with WT mice (131). The increased colon tumorigenesis in cPLA2 mutant mice occurs despite a decrease in PGE2 production. In contrast, genetic ablation of cPLA2 protects against cancer development in mouse models of adenomatous polyposis (Apc 716 KO mice and ApcMin mice) (125, 132). Apc 716 KO mice develop small intestinal polyps that express high levels of COX-2 and cPLA2 mRNA compared with normal small intestine (125, 133). Genetic deletion of cPLA2 in Apc 716 KO mice leads to a reduction in size of the polyps but has no effect on the number of polyps unlike COX-2 KO mice (125, 133). In contrast, knocking out cPLA2 in ApcMin mice results in a reduction in size and a dramatic decrease in the number of small intestinal polyps, although there is no effect on tumor number in the large intestine (132). The results highlight differences in the role of cPLA2 in promoting tumorigenesis in the small and large intestine. Several lines of evidence suggest that tumorigenesis in the small intestine involves the cPLA2/COX-2/PGE2 pathway rather than a role for arachidonic acid itself in promoting apoptosis (132, 134). Understanding how cPLA2 regulates tumorigenesis is complex because it is upstream of many lipid mediators including arachidonic acid, diverse PGs, leukotrienes, and lysophospholipids. In addition, the large repertoire of receptors for lipid mediators is differentially expressed in tissues and promotes signaling pathways with distinct effects on cell function. This adds another level of complexity in understanding the role of cPLA2-derived bioactive lipids in organ-specific cancer development. Liver Studies have shown that PGs promote increased triglyceride storage in adipocytes and hepatocytes, prompting investigations on the role of cPLA2 in regulating lipid storage and liver damage (135–140). It is of interest that even under normal dietary conditions, cPLA2 regulates lipid storage in adipose tissue and liver (138). cPLA2 KO mice fed a normal diet have reduced adipose tissue mass compared with WT mice, and lower triglyceride deposition in the liver, that correlates with lower serum levels of PGE2. An extension of this work showed that cPLA2 KO mice are protected from fatty liver damage induced by a high-fat diet (139). Mice with cPLA2 deficiency have lower serum PGE2 levels, which may afford protection because the administration of PGE2 enhances hepatic triacylglycerol content, whereas inhibition of COX-2 prevents fat deposition in rodents (137, 141). Obesity is associated with development of nonalcoholic fatty liver disease with pathological manifestations including nonalcoholic steatohepatitis, fibrosis, or cirrhosis. Recently, it was shown that cPLA2 KO mice fed a high-fat diet are protected from hepatic liver deposition and development of hepatic fibrosis (140). Similar results are observed in mice subjected to repeated administration of carbon tetrachloride that induces fibrosis without lipid deposition. The mechanism is cPLA2: physiological function and role in disease

1391

Downloaded from www.jlr.org by guest, on August 13, 2015

developed more severe disease involving the ileum, stomach, esophagus, duodenum, biliary system, and liver, although the colon remained normal. He developed type 2 diabetes, peripheral neuropathy, and osteoporosis. His sister exhibited similar severe GI disease beginning at age 2 years that worsened over time. She experienced infections with Campylobacter enteritis and Salmonella enteriditis, and developed Candida septicemia, staphylococcus lung infection, and acute respiratory distress syndrome. Other extra-intestinal disease included acute renal failure, endometriosis, left ventricular concentric hypertrophy, fibrotic bladder with stones, and infertility. The siblings contained homozygous 4 bp deletions in the PLA2G4A gene predicted to result in a frameshift and premature stop codon with the loss of 43 amino acids at the C terminus of cPLA2 (from V707) (Fig. 2). cPLA2 was undetectable by immunohistochemical analysis of small bowel biopsy and could not be detected in peripheral blood mononuclear cells by Western blot. Many factors could contribute to the individual phenotypic differences in disease severity associated with inherited cPLA2 deficiency, but the studies show a critical role for cPLA2 and eicosanoids in maintaining the integrity of the gut, particularly the small intestine. The small intestine of the cPLA2 KO mouse has numerous small ulcers, but there appears to be little effect of the lesions on the health of mutant mice (125). However, there is a role for cPLA2 in protecting the GI tract of mice from chemical toxicity induced by COX inhibitors and dextran sodium sulfate (DSS) (126, 127). Injury induced by COX inhibitors in cPLA2 KO mice did not occur in the colon, but was extensive in the small intestine, the site of greatest drug absorption (126). Differences in toxicity of cPLA2 WT and KO mice did not correlate with PGE2 levels, but may have been due to greater mitochondrial toxicity in KO mice. In the DSS model, there was extensive damage in the colons of cPLA2 KO mice that correlated with lower levels of PGs in colon tissue compared with DSS-treated WT mice (127). Mice deficient in mPGES-1 also exhibited greater DSS toxicity in the colon, suggesting an important protective role for PGE2. In contrast to cPLA2 KO mice, which recovered from DSS toxicity to a similar extent as WT mice, the recovery of mPGES-1 KO mice was compromised. The studies in mice and humans show an important role for cPLA2 in protecting the GI tract. However, levels of eicosanoids in the GI tract of cPLA2 KO mice are only partially reduced, suggesting a role for other PLA2s. This may contribute to the less severe health effects of cPLA2 deficiency in mice compared with humans. There has not been a systematic comparison of the relative levels of the various types of PLA2 enzymes expressed in different tissues of mice and humans to speculate on the PLA2s that may account for eicosanoid production in the GI tract of mice. Although results support a role for PGs in protecting the integrity of the intestine, COX-2 expression and PGE2mediated inflammation are associated with development of human colorectal cancer, consistent with epidemiological evidence that nonsteroid anti-inflammatory drug use is protective (128–130). However, cPLA2 protects

thought to involve a role for cPLA2 in production of monocyte chemoattractant protein-1 and infiltration of macrophages in these models. Previous results have shown that leukotrienes and PGs regulate induction of monocyte chemoattractant protein-1, a chemotactic factor implicated in development of hepatic fibrosis (142–145). Eicosanoids secreted by cells influence gene expression through paracrine or autocrine interaction with their specific G proteincoupled receptors (146, 147). For example in an infectious disease model using macrophages from WT and cPLA2 KO mice, we found that cPLA2 activation promotes an autocrine loop involving production of PGs that engage prostanoid receptors, increase cAMP, and globally affect expression of genes involved in host defense and inflammation (43).

Reproduction PGs produced through the COX-1 and COX-2 pathways play an important role in regulating female reproduction, therefore it is not surprising that cPLA2 KO female mice 1392

Journal of Lipid Research Volume 56, 2015

Kidney cPLA2 KO female mice exhibit defects in renal function, possibly due to a deficiency in PGE2 production, which regulates various aspects of kidney physiology (161). Mice lacking cPLA2 have a urine-concentrating defect that develops with age. Based on abnormal aequorin staining in proximal tubules, the defect in cPLA2 KO female mice may be due to diminished water reabsorption. There is also evidence that cPLA2 regulates Na+-HCO3- transport by angiotensin II in kidney proximal tubules (162). The role of cPLA2 in kidney function of humans is poorly understood. The female sibling with the cPLA2 deletion developed acute renal failure requiring hemodialysis, although whether this is a direct effect of cPLA2 deletion is not clear (50). It has recently been

Downloaded from www.jlr.org by guest, on August 13, 2015

Hemostasis Considering the importance of eicosanoids in regulating hemostasis, the role of cPLA2 in TXA2 production, platelet activation, and hemostasis has been investigated. Platelets from cPLA2 KO mice stimulated with collagen produce much less TXA2, and bleeding time is increased compared with WT mice (59). However, the low level of TXA2 produced by other phospholipases in collagen-stimulated cPLA2 KO platelets is sufficient for platelet aggregation. In vivo, cPLA2 KO mice have increased bleeding time and are protected from thromboembolism that correlates with reduced TXA2 levels and vasoconstriction. cPLA2 appears to play a discrete role in the complex regulation of platelet activation in mice. Studies of humans with inherited cPLA2 deficiency have shown an important role of cPLA2 in regulating platelet function and hemostasis. In the patient with two heterologous bp mutations in the coding regions of PLA2G4A, the PLA2 activity in platelet lysates was reduced and levels of TXB2 and 12-hydroxyeicosatetraenoic acid in serum during blood clotting (platelet-derived) were much lower than normal levels (49). Platelet aggregation and ATP release were compromised indicating impaired function. In patients with homozygous deletion mutants of cPLA2 associated with cryptogenic multifocal ulcerating stenosing enteritis (50), similar abnormalities of platelet function were observed, including defects in collageninduced aggregation and reduced levels of TXB2 production. The results indicate that cPLA2 plays a dominant role in providing arachidonic acid for the COX-1 and 12-lipoxygenase products in platelets, although the patients described above did not exhibit generalized clotting deficiency. In contrast, the siblings with inherited cPLA2 deficiency due to homozygous bp mutations had lifelong bleeding diathesis characterized by spontaneous mucocutaneous bleeding in addition to GI bleeding. The bleeding defect correlated with defective platelet function and extremely low levels of serum TXB2 (51).

exhibit decreased fertility, small litter size, and delayed onset of labor (47, 48, 148–150). In cPLA2 KO mice, blastocyst attachment is delayed due to a loss of maternal cPLA2, but decidualization is unaffected (151). Implantation beyond the normal “window” results in defective embryo development, abnormal uterine spacing of the embryos and resorption contributing to the small litter size, and defects in parturition (151). A later study emphasized the importance of cPLA2 function early in pregnancy because the defect in parturition in cPLA2 KO mice could be attributed to the deferred embryo implantation (152). cPLA2 KO mice showed less pronounced defects in ovulation and fertilization compared with COX-2 KO mice (149, 151, 152). A more detailed study showed that cPLA2 modestly regulates ovulation and fertilization by promoting the induction of COX-2 for PG production (153). A role for cPLA2 in regulating male sexual maturation has also been described (154). Male cPLA2 KO mice are fertile but exhibit abnormal histology of the testes showing a reduced number of interstitial cells. Compared with WT mice, cPLA2 KO males have delayed spermatogenesis and reduced testosterone production, although these defects normalized with age. It is difficult to extrapolate from studies in mouse models in understanding the specific role of cPLA2 in human reproduction, which has many unique regulatory features not shared by other mammalian species (155). The female patient with homozygous deletion mutations in cPLA2 associated with cryptogenic multifocal ulcerating stenosing enteritis is infertile, although the basis for this is unknown, and may be due to the severe consequences of the disease (50). PGs regulate labor and birth in humans, and an important source of PGs is amnion fibroblasts in fetal membranes (156, 157). There is an increase in cortisol secretion and PG production during parturition, a paradoxical event because cortisol suppresses PG production in most tissues (158). Glucocorticoids have been shown to induce mRNA and protein expression of cPLA2 and COX-2 in human amnion fibroblasts (72, 159). The induction of cPLA2 by glucocorticoids involves activation of Gs leading to cyclic AMP response element-binding protein-1 phosphorylation, which interacts with the glucocorticoid receptor on the glucocorticoid response element in the cPLA2 promoter (160).

shown that cPLA2 plays a role in mediating the onset of nonobstructive neonatal hydronephrosis, induced in mice by treatment with 2,3,7,8-tetrachlorodibenzo-p-dioxin, which increases expression of cPLA2 in the kidneys (163). Results showed that cPLA2 KO neonates have reduced incidence and severity of hydronephrosis that correlates with reduced expression of COX-2, mPGES-1, and lower urinary PGE2 compared with 2,3,7,8-tetrachlorodibenzo-p-dioxintreated WT pups. The COX-2/PGE2 pathway is also implicated in the pathogenesis of dioxin-induced hydronephrosis (164). The toxic effects of dioxins are mediated by the arylhydrocarbon receptor, which binds to the second intron of Pla2g4a and mediates its induction (165, 166). The arylhydrocarbon receptor pathway also promotes signaling cascades for posttranslational activation of cPLA2 by increasing calcium and phosphorylation (167, 168).

cPLA2: physiological function and role in disease

1393

Downloaded from www.jlr.org by guest, on August 13, 2015

Nervous system Many studies have implicated cPLA2 in mediating several disease processes in the nervous system including postischemic brain injury, spinal cord injury (SCI), and neurodegenerative diseases, as will be discussed below. However, there is a role for cPLA2 in regulating neuronal homeostasis, particularly long-term depression and longterm potentiation involved in synaptic plasticity. Longterm depression is blocked in brain slices from cPLA2 KO mice by treatment with cPLA2 (pyrrolidine-1) and COX-2 inhibitors (169). The defect in long-term depression by cPLA2 deficiency or COX-2 inhibition attenuates optokinetic eye movement adaptation, suggesting that cPLA2regulated long-term depression is important for motor learning. Long-term potentiation is also important for motor learning and a recent report found that long-term potentiation in cerebellar parallel fiber-Purkinje cell synapses is abolished in cPLA2 KO mice (170, 171). The earliest work investigating the role of cPLA2 in brain injury found that cPLA2 KO mice are protected from postischemic brain injury induced by middle cerebral artery occlusion (47). Infarct volumes are reduced in KO mice after reperfusion; this is particularly evident in posterior brain slices. The KO mice also exhibit fewer neurological deficits compared with WT mice. No differences in the vascular anatomy of the brain in KO and WT mice were noted. An extension of this work found that cPLA2 contributes to the early events leading to injury during cerebral ischemia (172). cPLA2 WT mice exhibit greater COX-2 expression, PGE2 production, and reactive oxygen species after ischemia, and more disruption of neuron morphology than KO mice. Other studies have reported a role for cPLA2 in regulating the expression of COX-2 in the brain (173). After systemic LPS, there is less COX-2 expression and PGE2 production in cPLA2 KO compared with WT mice (174). COX-2 is induced in neuronal and nonneuronal cells of the brain during inflammation and promotes injury (175). A role for cPLA2 in regulating SCI has also been investigated. One report found that cPLA2 expression is increased in SCI, particularly in neurons and oligodendrocytes (176). It was shown that inhibiting cPLA2 (AACOCF3) or

genetic ablation (cPLA2 KO C57Bl/6 mice) results in improved motor deficiencies and less tissue damage from SCI. Another study reported a complex role for different PLA2 enzymes (cPLA2, sPLA2 GIIA, and iPLA2 GVIA) in protecting or exacerbating SCI (177). All of the PLA2s were upregulated after SCI. In contrast to Liu et al. (176), cPLA2 was found to protect from SCI, which was determined using specific inhibitors and cPLA2 KO mice (Balb/c), but sPLA2 GIIA and iPLA2 GVIA were detrimental. Results suggested that the beneficial effect of cPLA2 in SCI involves PGE2 signaling through the EP1 receptor. A confounding factor in evaluating the role of cPLA2 in these mouse models is the influence of genetic differences in mouse strains. For example, the commonly used mouse strains, BALB/c and C57BL/6, generally exhibit distinct polarization of helper T cell responses, with BALB/c predisposed to Th2 and C57BL/6 to Th1 responses (178). In addition, C57BL/6 mice lack sPLA2 GIIA due to a natural deletion in the gene that does not occur in BALB/c mice (179). A role for cPLA2 in mouse models of neurodegenerative disease such as Alzheimer’s disease has been investigated (180, 181). The A peptides that are implicated in Alzheimer’s neurodegeneration exert toxic effects in neurons, which are in part mediated by cPLA2 (182, 183). Genetic ablation of cPLA2 improves cognitive function in a mouse model of familial Alzheimer’s disease, and protects from the toxic effects of intra-cerebroventricular injection of A oligomers (184, 185). A recent study found that A peptides activate cPLA2 in rat primary cortical neurons, resulting in increased expression of -amyloid precursor protein by an autocrine pathway involving PGE2 production and increases in cAMP (186). Additional studies also suggest that cPLA2 activation is involved in promoting neurodegeneration that occurs in prion diseases. In primary cortical neurons, the prion-derived peptide (PrP82-146) induces cPLA2 activation and arachidonic acid release that correlates with synapse damage and cell death (187 ). Treatment of neurons with PLA2 inhibitors (AACOCF3, MAFP) and a PGE2 receptor antagonist (AH13205), protects from synapse degeneration (188). In a model of Parkinson’s disease, cPLA2 contributes to neurotoxicity induced by 1-methyl-4-phenyl-1,2,3,6tetrahydropyridine (189). The reduction in dopamine induced by 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine is significantly less in KO than in WT mice. It is not known how cPLA2 contributes to neurodegeneration in these models, but mechanisms involving free radical production, effects on intracellular membrane trafficking, and arachidonic acid metabolites have been proposed. It is important to consider underlying differences that have been found in the architecture and lipid composition of the brain in cPLA2 WT and KO mice. A recent study showed by microscopy that cortical neurons in cPLA2 KO mice exhibit abnormal architecture including larger nuclei, increased number of nucleoli, and aggregated intra-nuclear ribosomes (190). Abnormalities in the nuclear envelope, nuclear pore, and synapses were observed. The authors speculated that this could be due to the role of

Lung There is extensive information investigating the role of eicosanoids in regulating lung diseases, including fibrosis, acute lung injury, and cancer. Comparisons of cPLA2 KO and WT mice have provided interesting insight into how cPLA2 regulates these processes. When the cPLA2 KO mouse was initially developed, the role of cPLA2 in regulating type I allergic response (anaphylaxis) in the lung was investigated (48). To induce systemic anaphylaxis KO and WT mice were immunized by intraperitoneal injection of ovalbumin (OVA) followed by intravascular OVA injection to elicit anaphylaxis. In WT mice, there was narrowing of the airway and alveolar thickening that was not evident in KO mice. Lung resistance was lower and recovery was faster in KO mice than WT mice after OVA challenge. Bronchial reactivity to methacholine challenge was dramatically reduced in KO mice. Airway hyperreactivity is also reduced in 5-LO KO mice, suggesting that cPLA2mediated production of leukotrienes mediates allergic responses (192). In contrast, allergic lung responses are exacerbated in COX-1 and COX-2 KO mice compared with WT mice, suggesting PGs play a protective role (193). Over the past few years, Wyeth has developed indole compounds that selectively inhibit human cPLA2 and have been tested in a sheep model of allergic asthma (194). Because diverse lipid mediators are implicated in mediating asthmatic responses in this model, they reasoned that broad inhibition with a cPLA2 inhibitor would be beneficial. The indole cPLA2 inhibitors reduced late phase bronchoconstriction and airway hyperreactivity in allergen-challenged sheep. The indole inhibitor also blocked allergen-dependent induction of novel genes expressed in cells from asthmatics (195). One of the most potent indole 1394

Journal of Lipid Research Volume 56, 2015

inhibitors, PF-5212372, blocks eicosanoid production by stimulated human lung mast cells and lung homogenates. However, in human lung homogenates, which contain a complex cell mix, the inhibitor was less effective at blocking PGE2 production than other eicosanoids, suggesting that other PLA2s, particularly sPLA2s, may contribute to arachidonic acid release for eicosanoid production (1, 196). The inhibitor also blocked contraction of isolated human bronchial rings induced by AMP. The results provide convincing evidence for a role of cPLA2 in mediating responses in allergic asthma. Analysis of the promoter region of PLA2G4A has shown that microsatellite fragments are shorter in patients with severe asthma than healthy controls (197). A comparison of peripheral blood monocytes from severe asthmatics with short and long microsatellite sequences in the PLA2G4A promoter found a correlation between the shorter sequences with increased expression of cPLA2 mRNA and protein, and greater eicosanoid production (197, 198). Results of studies comparing cPLA2 WT and KO mice have also supported a role for cPLA2 in mediating acute lung injury (199). One model induced lung injury by intravenous administration of endotoxin and fungal cell walls, zymosan; the other model involved intratracheal administration of hydrochloric acid. These models simulate endotoxemia and acid aspiration, which are causes of acute respiratory distress syndrome. In both models of lung injury, the cPLA2 KO mice were protected, exhibiting less protein leakage and neutrophil recruitment into the lung than WT mice, that correlated with reduced levels of TXB2, LTB4, and cysteinyl leukotrienes. However, in a cecal ligation and puncture model of experimental sepsis, there was no difference in survival of cPLA2 WT and KO mice, although there was lower production of eicosanoids (with the exception of 12-hydroxy-eicosatetraenoic acid) and cytokines (IL-6 and CCL2) in the mutant mice (200). It is possible that lipid mediators do not play a dominant role in regulating this sepsis model, or that just targeting one PLA2 isoform is not sufficient (201). A study using an antisense approach to inhibit both sPLA2-IIa and cPLA2 showed improved survival of rats in a cecal ligation and puncture model (202). Another consideration is that cPLA2 mediates the production of diverse lipid mediators that may have opposing effects, some are pro-inflammatory and others anti-inflammatory (201). The opposing role of cPLA2-derived lipid mediators is illustrated in the regulation of pulmonary fibrosis, a chronic disease characterized by the accumulation of myofibroblasts and deposition of extracellular matrix (203). Chronic inflammation often leads to fibrosis due to aberrant wound healing resulting in permanent scarring and organ failure (204). A study of pulmonary fibrosis induced by bleomycin found that fibrosis, collagen synthesis, and inflammation were reduced in cPLA2 KO mice compared with WT mice (205). In WT mice, the effects of bleomycin were accompanied by increases in eicosanoids in bronchoalveolar lavage fluid that were markedly lower in bleomycin-treated cPLA2 KO mice. Although disease parameters were not completely attenuated in KO mice,

Downloaded from www.jlr.org by guest, on August 13, 2015

cPLA2 in regulating membrane structure and curvature. In addition, a study comparing the brain lipid content of cPLA2 WT and KO mice found altered brain phospholipid composition in mutant mice but no differences in the concentrations of nonesterified fatty acids in the plasma or brain (191). There were decreased levels of esterified arachidonic acid in phosphatidylinositol and decreases of several fatty acids including linoleic, arachidonic, and docosahexaenoic acids in phosphatidylcholine in cPLA2 KO mice compared with WT mice. However, an interesting finding was the increased incorporation and turnover of AA in phosphatidylinositol and phosphatidylcholine in cPLA2 KO mice, perhaps due to compensation from other PLA2s. Consequently the differences observed in phospholipid brain composition in cPLA2 WT and KO mice may reflect both a lack of cPLA2 and compensation from other enzymes with differences in lipid specificity as noted by the authors (191). The lipid alterations in KO mice may influence membrane properties contributing to the structural differences noted in cortical neurons, and could also influence the response of mutant mice to brain injury. This again raises the question of whether responses of cPLA2 KO mice are due to a primary effect of the loss of cPLA2 or secondary compensatory changes that occur.

Autoimmunity Chronic inflammation is a characteristic of certain autoimmune diseases, such as rheumatoid arthritis. Collageninduced arthritis in mice has pathological features resembling rheumatoid arthritis and the role of eicosanoids has been studied extensively. Interestingly, there is evidence that both leukotrienes and PGs contribute to disease pathogenesis based on studies using FLAP inhibitors, LTB4 receptor antagonist, and COX-2 KO mice (225–227). Therefore, it is not surprising that cPLA2 KO mice are also protected

from collagen-induced arthritis (228). Based on the collective involvement of eicosanoids in this autoimmune disease, it has been suggested that a cPLA2 inhibitor would have therapeutic value by simultaneously blocking eicosanoids from the COX and 5-LO pathways (228). In rheumatoid arthritis, bone homeostasis is disrupted due, in part, to chronic inflammation resulting in joint destruction and bone loss. The role of cPLA2 in regulating inflammatory bone resorption induced in mice with endotoxin has been investigated (229). In response to endotoxin (ip), cPLA2 KO mice exhibit less bone loss and reduced PGE2 production. The results indicate that cPLA2 mediates osteoclastic bone resorption due to endotoxininduced PGE2 production by osteoblasts. The role of cPLA2 in regulating autoimmune disease in nonobese diabetic (NOD) mice has also been investigated (230). NOD mice spontaneously develop autoimmune diabetes that has similarities to human type 1 diabetes. In contrast to collageninduced arthritis, cPLA2 plays a protective role in the development of diabetes. NOD mice with genetic ablation of cPLA2 have a higher incidence of diabetes than cPLA2 WT NOD mice. Results suggest that cPLA2 plays a protective role during the diabetic stage, possibly due to the ability of PGE2 produced by macrophages to suppress the production of the pro-inflammatory cytokine, TNF. Experimental autoimmune encephalomyelitis (EAE) is an inflammatory disease in mice that has features of the human autoimmune disease multiple sclerosis (231). There are conflicting data on the role of specific eicosanoids in regulating EAE based on studies with 5-LO KO mice, in which EAE is exacerbated, or using an LTB4 receptor antagonist, which protects (232). The role of PGs is not clear because COX-2 KO mice are not protected but a COX-2 inhibitor is beneficial in EAE (233, 234). A study of EAE induced in cPLA2 WT and KO mice on the susceptible C57Bl/6 background found that KO mice are resistant to disease (235). Results suggested that cPLA2 contributes in the induction and effector phase by contributing to the development of Th1 responses. Because the development of the immune system may be altered in KO mice, a follow up study tested the effect of cPLA2, 5-LO, and COX inhibitors on the development of EAE in WT mice (236). The cPLA2 inhibitor, WAY-196025, used at the time of immunization to promote EAE blocked the development of EAE, and when used at the start of disease reduced the duration of EAE relapses. Inhibitors of COX1/2 and 5-LO also reduced disease severity. Results using the selective cPLA2 inhibitor are consistent with the studies using the cPLA2 KO mice. A recent study also showed that another selective cPLA2 inhibitor (2-oxoamide, AX059) protected rats from EAE that correlated with increased production of Tregs (237). However, the conflicting studies on the role of specific eicosanoids in EAE highlight potential difficulties in interpreting results from KO mice, which can exhibit compensatory effects or changes that occur during embryogenesis. However, there are also complications in the use of inhibitors involving dosing and time of administration, as well as unknown off-target effects. cPLA2: physiological function and role in disease

1395

Downloaded from www.jlr.org by guest, on August 13, 2015

cPLA2 may contribute to the pathogenesis of fibrosis due to production of pro-inflammatory lipid mediators. This is supported by studies using mice with targeted disruption of 5-LO and the cysteinyl leukotriene receptors 1 and 2, which show reduced bleomycin-induced pulmonary fibrosis (206–208). In contrast, COX-2 deficiency promotes pulmonary fibrosis, and an important role for PGE2 and PGI2 in protecting the lung has emerged (209–212). Consequently cPLA2 is upstream of pathways that produce mediators with positive and negative effects on lung disease. There has been considerable interest in the role of lysophosphatidic acid (LPA) in promoting fibrosis in many organ systems, including the lung (213, 214). LPA levels increase during fibrosis and it exerts a profibrotic effect primarily through the LPA1 receptor. Several pathways for the synthesis of LPA have been described (215, 216). Although there is correlative data that cPLA2 may play a role in the initial production of lysophospholipids that are metabolized to LPA by autotaxin, there is as yet little definite data to support a role for cPLA2 in LPA production (217, 218). Mouse models have been used to investigate the role of cPLA2 in regulating lung cancer. Using the chemical carcinogen urethane to initiate lung cancer, cPLA2 KO mice developed fewer lung tumors than WT mice, and much lower levels of tumor-associated PGs (219). Another study investigated whether the presence of cPLA2 in the tumor microenvironment influenced lung cancer progression and metastasis (220). The injection of mouse lung tumor cells into mouse lungs induced primary tumors of similar size in cPLA2 WT and KO mice. However, KO mice had fewer secondary tumor metastases, less lymph node spread and increased survival. cPLA2 is thought to contribute to tumorigenesis, in part by enhancing the viability and proliferation of endothelial cells leading to formation of vascular networks in the tumor microenvironment (221, 222). This process diminishes the effectiveness of radiotherapy. It has recently been shown that a cPLA2 inhibitor (PLA-695) blocks pro-survival signaling in endothelial cells and renders cancer cells radiosensitive in a simulated tumor microenvironment (223). Combined therapy (PLA-695 and irradiation) also reduces lung tumor growth in mice. Results suggest that cPLA2-mediated production of lysophospholipid metabolites promotes vascular endothelial cell proliferation, migration, and tumor vascularization (224). Therefore a number of lipid mediators, including arachidonic acid, eicosanoids, and lysophospholipids, may contribute to cancer development.

Therefore multiple approaches are required to evaluate the role of these pathways in disease pathogenesis. As an example, a recent study used bone marrow transplantation to specifically block PGE2 synthesis, or EP receptor expression, in peripheral immune cells and showed that PGE2/EP4 signaling promotes development of EAE in mice (238). This approach demonstrates the importance of investigating the role of specific cell types and receptor pathways for understanding how lipid mediators regulate disease processes in vivo.

CONCLUDING REMARKS

1396

Journal of Lipid Research Volume 56, 2015

1. Lambeau, G., and M. H. Gelb. 2008. Biochemistry and physiology of mammalian secreted phopholipases A2. Annu. Rev. Biochem. 77: 495–520. 2. Dennis, E. A., J. Cao, Y. H. Hsu, V. Magrioti, and G. Kokotos. 2011. Phospholipase A2 enzymes: physical structure, biological function, disease implication, chemical inhibition, and therapeutic intervention. Chem. Rev. 111: 6130–6185. 3. Murakami, M., Y. Taketomi, Y. Miki, H. Sato, T. Hirabayashi, and K. Yamamoto. 2011. Recent progress in phospholipase A2 research: from cells to animals to humans. Prog. Lipid Res. 50: 152–192. 4. Ghosh, M., D. E. Tucker, S. A. Burchett, and C. C. Leslie. 2006. Properties of the group IV phospholipase A2 family. Prog. Lipid Res. 45: 487–510. 5. Shayman, J. A., R. Kelly, J. Kollmeyer, Y. He, and A. Abe. 2011. Group XV phospholipase A2, a lysosomal phospholipase A2. Prog. Lipid Res. 50: 1–13. 6. Stafforini, D. M. 2009. Biology of platelet-activating factor acetylhydrolase (PAF-AH, lipoprotein associated phospholipase A2). Cardiovasc. Drugs Ther. 23: 73–83. 7. Marathe, G. K., C. Pandit, C. L. Lakshmikanth, V. H. Chaithra, S. P. Jacob, and C. J. D’Souza. 2014. To hydrolyze or not to hydrolyze: the dilemma of platelet-activating factor acetylhydrolase. J. Lipid Res. 55: 1847–1854. 8. Abbott, M. J., T. Tang, and H. S. Sul. 2010. The role of phospholipase A2-derived mediators in obesity. Drug Discov. Today Dis. Mech. 7: e213–e218. 9. Rydel, T. J., J. M. Williams, E. Krieger, F. Moshiri, W. C. Stallings, S. M. Brown, J. C. Pershing, J. P. Purcell, and M. F. Alibhai. 2003. The crystal structure, mutagenesis, and activity studies reveal that patatin is a lipid acyl hydrolase with a Ser-Asp catalytic dyad. Biochemistry. 42: 6696–6708. 10. Ghomashchi, F., G. S. Naika, J. G. Bollinger, A. Aloulou, M. Lehr, C. C. Leslie, and M. H. Gelb. 2010. Interfacial kinetic and binding properties of mammalian group IVB phospholipase A2 (cPLA2B) and comparison with other cPLA2 isoforms. J. Biol. Chem. 285: 36100–36111. 11. Murakami, M., Y. Taketomi, Y. Miki, H. Sato, K. Yamamoto, and G. Lambeau. 2014. Emerging roles of secreted phospholipase A2 enzymes: the 3rd edition. Biochimie. 107: 105–113. 12. Ueno, N., Y. Taketomi, K. Yamamoto, T. Hirabayashi, D. Kamei, Y. Kita, T. Shimizu, K. Shinzawa, Y. Tsujimoto, K. Ikeda, et al. 2011. Analysis of two major intracellular phospholipases A2 (PLA2) in mast cells reveals crucial contribution of cytosolic PLA2alpha, not Ca(2+)-independent PLA2beta, to lipid mobilization in proximal mast cells and distal fibroblasts. J. Biol. Chem. 286: 37249–37263. 13. Gil-de-Gómez, L., A. M. Astudillo, C. Guijas, V. Magrioti, G. Kokotos, M. A. Balboa, and J. Balsinde. 2014. Cytosolic group IVA and calcium-independent group VIA phospholipase A2s act on distinct phospholipid pools in zymosan-stimulated mouse peritoneal macrophages. J. Immunol. 192: 752–762. 14. Ohto, T., N. Uozumi, T. Hirabayashi, and T. Shimizu. 2005. Identification of novel cytosolic phospholipase A2s, murine cPLA2delta, epsilon, and zeta, which form a gene cluster with cPLA2beta. J. Biol. Chem. 280: 24576–24583. 15. Bechler, M. E., P. de Figueiredo, and W. J. Brown. 2012. A PLA1-2 punch regulates the Golgi complex. Trends Cell Biol. 22: 116–124. 16. Ha, K. D., B. A. Clarke, and W. J. Brown. 2012. Regulation of the Golgi complex by phospholipid remodeling enzymes. Biochim. Biophys. Acta. 1821: 1078–1088. 17. San Pietro, E., M. Capestrano, E. V. Polishchuk, A. DiPentima, A. Trucco, P. Zizza, S. Mariggiò, T. Pulvirenti, M. Sallese, S. Tete, et al. 2009. Group IV phospholipase A2 controls the formation of inter-cisternal continuities involved in intra-Golgi transport. PLoS Biol. 7: e1000194. 18. Regan-Klapisz, E., V. Krouwer, M. Langelaar-Makkinje, L. Nallan, M. Gelb, H. Gerritsen, A. J. Verkleij, and J. A. Post. 2009. Golgiassociated cPLA2alpha regulates endothelial cell-cell junction integrity by controlling the trafficking of transmembrane junction proteins. Mol. Biol. Cell. 20: 4225–4234. 19. Ghosh, M., R. Loper, M. H. Gelb, and C. C. Leslie. 2006. Identification of the expressed form of human cytosolic phospholipase A2beta (cPLA2beta): cPLA2beta3 is a novel variant localized to mitochondria and early endosomes. J. Biol. Chem. 281: 16615–16624.

Downloaded from www.jlr.org by guest, on August 13, 2015

Due to the key role of cPLA2 in mediating lipid mediator production and its widespread tissue expression, it has been implicated in regulating homeostatic processes and disease pathogenesis throughout all organ systems. The function of cPLA2 in regulating certain physiological systems, such as female reproduction and hemostasis, is more straightforward and can be attributed to specific types of lipid mediators. However, elucidating the specific mechanism for cPLA2 action in disease processes in many cases is not known and understandably difficult to evaluate. The cPLA2 KO mouse has provided a wealth of information; however, multiple approaches are needed to determine cPLA2 function. Genetic differences in mouse strains lead to difficulties in interpretation of results, and compensatory changes in the conventional KO mouse are a complicating issue. The use of cPLA2 conditional KO mice with ablation in specific tissues would be useful to bypass changes occurring during embryogenesis and development. The availability of specific inhibitors is also important to understand cPLA2 function in animal models in vivo, and in human cells (2). The primary function of cPLA2 is due to its role in mediating production of eicosanoids, however, there is evidence that cPLA2 regulates processes such as membrane trafficking that are not linked to production of eicosanoids, adding an additional complexity to understanding its function (16, 17, 74). Many studies, primarily from comparisons of cPLA2 WT and KO mice, have suggested that cPLA2 is a useful therapeutic target for treating a plethora of diseases. There are a number of issues in considering cPLA2 as a therapeutic target, especially because most of the studies are based on disease models comparing cPLA2 WT and KO mice, which in some cases may not accurately reflect processes contributing to human disease. This is apparent from the differences observed between mice and humans as a consequence of cPLA2 deficiency. Another complication is that cPLA2 is the first regulatory enzyme in the pathway for production of numerous lipid mediators that often have opposing effects, although targeting cPLA2 may be beneficial in some diseases where both 5-LO and COX metabolites contribute to disease, as suggested in studies of asthma and CIA (194, 228). However, the essential role of cPLA2 in human health, particularly for maintenance of the small intestine and female reproduction, is also a consideration for targeting cPLA2 in disease.

REFERENCES

41. Kalinski, P. 2012. Regulation of immune responses by prostaglandin E2. J. Immunol. 188: 21–28. 42. Tilley, S. L., T. M. Coffman, and B. H. Koller. 2001. Mixed messages: modulation of inflammation and immune responses by prostaglandins and thromboxanes. J. Clin. Invest. 108: 15–23. 43. Suram, S., L. J. Silveira, S. Mahaffey, G. D. Brown, J. V. Bonventre, D. L. Williams, N. A. Gow, D. L. Bratton, R. C. Murphy, and C. C. Leslie. 2013. Cytosolic phospholipase A2alpha and eicosanoids regulate expression of genes in macrophages involved in host defense and inflammation. PLoS ONE. 8: e69002. 44. Shindou, H., S. Ishii, N. Uozumi, and T. Shimizu. 2000. Roles of cytosolic phospholipase A2 and platelet-activating factor receptor in the Ca-induced biosynthesis of PAF. Biochem. Biophys. Res. Commun. 271: 812–817. 45. Piñeiro, R., and M. Falasca. 2012. Lysophosphatidylinositol signalling: new wine from an old bottle. Biochim. Biophys. Acta. 1821: 694–705. 46. Norris, P. C., D. Gosselin, D. Reichart, C. K. Glass, and E. A. Dennis. 2014. Phospholipase A2 regulates eicosanoid class switching during inflammasome activation. Proc. Natl. Acad. Sci. USA. 111: 12746–12751. 47. Bonventre, J. V., Z. Huang, M. R. Taheri, E. O'Leary, E. Li, M. A. Moskowitz, and A. Sapirstein. 1997. Reduced fertility and postischaemic brain injury in mice deficient in cytosolic phospholipase A2. Nature. 390: 622–625. 48. Uozumi, N., K. Kume, T. Nagase, N. Nakatani, S. Ishii, F. Tashiro, Y. Komagata, K. Maki, K. Ikuta, Y. Ouchi, et al. 1997. Role of cytosolic phospholipase A2 in allergic response and parturition. Nature. 390: 618–622. 49. Adler, D. H., J. D. Cogan, J. A. Phillips 3rd, N. Schnetz-Boutaud, G. L. Milne, T. Iverson, J. A. Stein, D. A. Brenner, J. D. Morrow, O. Boutaud, et al. 2008. Inherited human cPLA2alpha deficiency is associated with impaired eicosanoid biosynthesis, small intestinal ulceration, and platelet dysfunction. J. Clin. Invest. 118: 2121–2131. 50. Brooke, M. A., H. J. Longhurst, V. Plagnol, N. S. Kirkby, J. A. Mitchell, F. Ruschendorf, T. D. Warner, D. P. Kelsell, and T. T. MacDonald. 2014. Cryptogenic multifocal ulcerating stenosing enteritis associated with homozygous deletion mutations in cytosolic phospholipase A2-. Gut. 63: 96–104. 51. Faioni, E. M., C. Razzari, A. Zulueta, E. A. Femia, L. Fenu, M. Trinchera, G. M. Podda, M. Pugliano, F. Marongiu, and M. Cattaneo. 2014. Bleeding diathesis and gastro-duodenal ulcers in inherited cytosolic phospholipase-A2 alpha deficiency. Thromb. Haemost. 112: 1182–1189. 52. Leslie, C. C. 2004. Regulation of the specific release of arachidonic acid by cytosolic phospholipase A2. Prostaglandins Leukot. Essent. Fatty Acids. 70: 373–376. 53. Nakatani, N., N. Uozumi, K. Kume, M. Murakami, I. Kudo, and T. Shimizu. 2000. Role of cytosolic phospholipase A2 in the production of lipid mediators and histamine release in mouse bone-marrow-derived mast cells. Biochem. J. 352: 311–317. 54. Fujishima, H., R. O. Sanchez Mejia, C. O. Bingham, B. K. Lam, A. Sapirstein, J. V. Bonventre, K. F. Austen, and J. P. Arm. 1999. Cytosolic phospholipase A2 is essential for both the immediate and the delayed phases of eicosanoid generation in mouse bone marrow-derived mast cells. Proc. Natl. Acad. Sci. USA. 96: 4803–4807. 55. Rubin, B. B., G. P. Downey, A. Koh, N. Degousee, F. Ghomashchi, L. Nallan, E. Stefanski, D. W. Harkin, C. Sun, B. P. Smart, et al. 2005. Cytosolic phospholipase A2-alpha is necessary for plateletactivating factor biosynthesis, efficient neutrophil-mediated bacterial killing, and the innate immune response to pulmonary infection. cPLA2-alpha does not regulate neutrophil NADPH oxidase activity. J. Biol. Chem. 280: 7519–7529. 56. Ghosh, M., A. Stewart, D. E. Tucker, J. V. Bonventre, R. C. Murphy, and C. C. Leslie. 2004. Role of cytosolic phospholipase A2 in prostaglandin E2 production by lung fibroblasts. Am. J. Respir. Cell Mol. Biol. 30: 91–100. 57. Gijón, M. A., D. M. Spencer, A. R. Siddiqi, J. V. Bonventre, and C. C. Leslie. 2000. Cytosolic phospholipase A2 is required for macrophage arachidonic acid release by agonists that do and do not mobilize calcium. Novel role of mitogen-activated protein kinase pathways in cytosolic phospholipase A2 regulation. J. Biol. Chem. 275: 20146–20156. 58. Tanaka, K., Y. Yamamoto, K. Ogino, S. Tsujimoto, M. Saito, N. Uozumi, T. Shimizu, and I. Hisatome. 2011. Cytosolic phospholipase

cPLA2: physiological function and role in disease

1397

Downloaded from www.jlr.org by guest, on August 13, 2015

20. Ghosh, M., R. Loper, F. Ghomashchi, D. E. Tucker, J. V. Bonventre, M. H. Gelb, and C. C. Leslie. 2007. Function, activity and membrane targeting of cytosolic phospholipase A2zeta in mouse lung fibroblasts. J. Biol. Chem. 282: 11676–11686. 21. Capestrano, M., S. Mariggio, G. Perinetti, A. V. Egorova, S. Iacobacci, M. Santoro, A. Di Pentima, C. Iurisci, M. V. Egorov, G. Di Tullio, et al. 2014. Cytosolic phospholipase A2 drives recycling through the clathrin-independent endocytic route. J. Cell Sci. 127: 977– 993. 22. Alonso, F., P. M. Henson, and C. C. Leslie. 1986. A cytosolic phospholipase in human neutrophils that hydrolyzes arachidonoyl-containing phosphatidylcholine. Biochim. Biophys. Acta. 878: 273–280. 23. Flesch, I., B. Schmidt, and E. Ferber. 1985. Acyl chain specificity and kinetic properties of phospholipase A1 and A2 of bone marrowderived macrophages. Z. Naturforsch. C. 40: 356–363. 24. Leslie, C. C., D. R. Voelker, J. Y. Channon, M. M. Wall, and P. T. Zelarney. 1988. Properties and purification of an arachidonoylhydrolyzing phospholipase A2 from a macrophage cell line, RAW 264.7. Biochim. Biophys. Acta. 963: 476–492. 25. Gronich, J. H., J. V. Bonventre, and R. A. Nemenoff. 1990. Purification of a high-molecular-mass form of phospholipase A2 from rat kidney activated at physiological calcium concentrations. Biochem. J. 271: 37–43. 26. Gronich, J. H., J. V. Bonventre, and R. A. Nemenoff. 1988. Identification and characterization of a hormonally regulated form of phospholipase A2 in rat renal mesangial cells. J. Biol. Chem. 263: 16645–16651. 27. Kramer, R. M., E. F. Roberts, J. Manetta, and J. E. Putnam. 1991. The Ca2+-sensitive cytosolic phospholipase A2 is a 100kDa protein in human monoblast U937 cells. J. Biol. Chem. 266: 5268–5272. 28. Kim, D. K., I. Kudo, and K. Inoue. 1991. Purification and characterization of rabbit platelet cytosolic phospholipase A2. Biochim. Biophys. Acta. 1083: 80–88. 29. Clark, J. D., N. Milona, and J. L. Knopf. 1990. Purification of a 110-kilodalton cytosolic phospholipase A2 from the human monocytic cell line U937. Proc. Natl. Acad. Sci. USA. 87: 7708–7712. 30. Sharp, J. D., D. L. White, X. G. Chiou, T. Goodson, G. C. Gamboa, D. McClure, S. Burgett, J. Hoskins, P. L. Skatrud, J. R. Sportsman, et al. 1991. Molecular cloning and expression of human Ca2+-sensitive cytosolic phospholipase A2. J. Biol. Chem. 266: 14850–14853. 31. Clark, J. D., L-L. Lin, R. W. Kriz, C. S. Ramesha, L. A. Sultzman, A. Y. Lin, N. Milona, and J. L. Knopf. 1991. A novel arachidonic acidselective cytosolic PLA2 contains a Ca2+-dependent translocation domain with homology to PKC and GAP. Cell. 65: 1043–1051. 32. Sharp, J. D., R. T. Pickard, X. G. Chiou, J. V. Manetta, S. Kovacevic, J. R. Miller, A. D. Varshavsky, E. F. Roberts, B. A. Strifler, D. N. Brems, et al. 1994. Serine 228 is essential for catalytic activities of 85-kDa cytosolic phospholipase A2. J. Biol. Chem. 269: 23250–23254. 33. Lin, L-L., M. Wartmann, A. Y. Lin, J. L. Knopf, A. Seth, and R. J. Davis. 1993. cPLA2 is phosphorylated and activated by MAP kinase. Cell. 72: 269–278. 34. Nemenoff, R. A., S. Winitz, N-X. Qian, V. Van Putten, G. L. Johnson, and L. E. Heasley. 1993. Phosphorylation and activation of a high molecular weight form of phospholipase A2 by p42 microtubule-associated protein 2 kinase and protein kinase C. J. Biol. Chem. 268: 1960–1964. 35. Dessen, A., J. Tang, H. Schmidt, M. Stahl, J. D. Clark, J. Seehra, and W. S. Somers. 1999. Crystal structure of human cytosolic phospholipase A2 reveals a novel topology and catalytic mechanism. Cell. 97: 349–360. 36. Pickard, R. T., X. G. Chiou, B. A. Strifler, M. R. DeFelippis, P. A. Hyslop, A. L. Tebbe, Y. K. Yee, L. J. Reynolds, E. A. Dennis, R. M. Kramer, et al. 1996. Identification of essential residues for the catalytic function of 85-kDa cytosolic phospholipase A2. J. Biol. Chem. 271: 19225–19231. 37. Peters-Golden, M., and W. R. Henderson, Jr. 2007. Leukotrienes. N. Engl. J. Med. 357: 1841–1854. 38. Ricciotti, E., and G. A. Fitzgerald. 2011. Prostaglandins and inflammation. Arterioscler. Thromb. Vasc. Biol. 31: 986–1000. 39. Aoki, T., and S. Narumiya. 2012. Prostaglandins and chronic inflammation. Trends Pharmacol. Sci. 33: 304–311. 40. Frasch, S. C., and D. L. Bratton. 2012. Emerging roles of lysophosphatidylserine in resolution of inflammation. Prog. Lipid Res. 51: 199–207.

59.

60. 61. 62.

63.

64.

65.

67. 68.

69.

70.

71.

72.

73.

74. 75.

76.

1398

Journal of Lipid Research Volume 56, 2015

77. Glover, S., M. S. de Carvalho, T. Bayburt, M. Jonas, E. Chi, C. C. Leslie, and M. H. Gelb. 1995. Translocation of the 85-kDa phospholipase A2 from cytosol to the nuclear envelope in rat basophilic leukemia cells stimulated with calcium ionophore or IgE/ antigen. J. Biol. Chem. 270: 15359–15367. 78. Evans, J. H., D. M. Spencer, A. Zweifach, and C. C. Leslie. 2001. Intracellular calcium signals regulating cytosolic phospholipase A2 translocation to internal membranes. J. Biol. Chem. 276: 30150–30160. 79. Nalefski, E. A., M. M. Slazas, and J. J. Falke. 1997. Ca2+-signaling cycle of a membrane-docking C2 domain. Biochemistry. 36: 12011–12018. 80. Ball, A., R. Nielsen, M. H. Gelb, and B. H. Robinson. 1999. Interfacial membrane docking of cytosolic phospholipase A2 C2 domain using electrostatic potential-modulated spin relaxation magnetic resonance. Proc. Natl. Acad. Sci. USA. 96: 6637–6642. 81. Davletov, B., O. Perisic, and R. L. Williams. 1998. Calciumdependent membrane penetration is a hallmark of the C2 domain of cytosolic phospholipase A2 whereas the C2A domain of synaptotagmin binds membranes electrostatically. J. Biol. Chem. 273: 19093–19096. 82. Perisic, O., H. F. Paterson, G. Mosedale, S. Lara-González, and R. L. Williams. 1999. Mapping the phospholipid-binding surface and translocation determinants of the C2 domain from cytosolic phospholipase A2. J. Biol. Chem. 274: 14979–14987. 83. Gijón, M. A., D. M. Spencer, A. L. Kaiser, and C. C. Leslie. 1999. Role of phosphorylation sites and the C2 domain in regulation of cytosolic phospholipase A2. J. Cell Biol. 145: 1219–1232. 84. Evans, J. H., S. H. Gerber, D. Murray, and C. C. Leslie. 2004. The calcium binding loops of the cytosolic phospholipase A2 C2 domain specify targeting to Golgi and ER in live cells. Mol. Biol. Cell. 15: 371–383. 85. Stahelin, R. V., J. D. Rafter, S. Das, and W. Cho. 2003. The molecular basis of differential subcellular localization of C2 domains of protein kinase C-alpha and group IVa cytosolic phospholipase A2. J. Biol. Chem. 278: 12452–12460. 86. Evans, J. H., D. Murray, C. C. Leslie, and J. J. Falke. 2006. Specific translocation of protein kinase Calpha to the plasma membrane requires both Ca2+ and PIP2 recognition by its C2 domain. Mol. Biol. Cell. 17: 56–66. 87. Evans, J. H., and C. C. Leslie. 2004. The cytosolic phospholipase A2 catalytic domain modulates association and residence time at Golgi membranes. J. Biol. Chem. 279: 6005–6016. 88. Nalefski, E. A., T. McDonagh, W. Somers, J. Seehra, J. J. Falke, and J. D. Clark. 1998. Independent folding and ligand specificity of the C2 calcium-dependent lipid binding domain of cytosolic phospholipase A2. J. Biol. Chem. 273: 1365–1372. 89. Das, S., and W. Cho. 2002. Roles of catalytic domain residues in interfacial binding and activation of group IV cytosolic phospholipase A2. J. Biol. Chem. 277: 23838–23846. 90. Burke, J. E., Y. H. Hsu, R. A. Deems, S. Li, V. L. Woods, Jr., and E. A. Dennis. 2008. A phospholipid substrate molecule residing in the membrane surface mediates opening of the lid region in group IVA cytosolic phospholipase A2. J. Biol. Chem. 283: 31227–31236. 91. Peters-Golden, M., K. Song, T. Marshall, and T. Brock. 1996. Translocation of cytosolic phospholipase A2 to the nuclear envelope elicits topographically localized phospholipid hydrolysis. Biochem. J. 318: 797–803. 92. Tucker, D. E., M. Ghosh, F. Ghomashchi, R. Loper, S. Suram, B. St. John, M. Girotti, J. G. Bollinger, M. H. Gelb, and C. C. Leslie. 2009. Role of phosphorylation and basic residues in the catalytic domain of cytosolic phospholipase A2alpha in regulating interfacial kinetics and binding and cellular function. J. Biol. Chem. 284: 9596–9611. 93. Pettus, B. J., A. Bielawska, P. Subramanian, D. S. Wijesinghe, M. Maceyka, C. C. Leslie, J. H. Evans, J. Freiberg, P. Roddy, Y. A. Hannun, et al. 2004. Ceramide 1-phosphate is a direct activator of cytosolic phospholipase A2. J. Biol. Chem. 279: 11320–11326. 94. Stahelin, R. V., P. Subramanian, M. Vora, W. Cho, and C. E. Chalfant. 2007. Ceramide-1-phosphate binds group IVA cytosolic phospholipase A2 via a novel site in the C2 domain. J. Biol. Chem. 282: 20467–20474. 95. Lamour, N. F., P. Subramanian, D. S. Wijesinghe, R. V. Stahelin, J. V. Bonventre, and C. E. Chalfant. 2009. Ceramide 1-phosphate is required for the translocation of group IVA cytosolic phospholipase A2 and prostaglandin synthesis. J. Biol. Chem. 284: 26897–26907.

Downloaded from www.jlr.org by guest, on August 13, 2015

66.

A2alpha contributes to blood pressure increases and endothelial dysfunction under chronic NO inhibition. Arterioscler. Thromb. Vasc. Biol. 31: 1133–1138. Wong, D. A., Y. Kita, N. Uozumi, and T. Shimizu. 2002. Discrete role for cytosolic phospholipase A2alpha in platelets: studies using single and double mutant mice of cytosolic and group IIA secretory phospholipase A2. J. Exp. Med. 196: 349–357. Wu, T., T. Ikezono, C. W. Angus, and J. H. Shelhamer. 1994. Characterization of the promoter for the human 85 kDa cytosolic phospholipase A2 gene. Nucleic Acids Res. 22: 5093–5098. Tay, A., P. Maxwell, Z. G. Li, H. Goldberg, and K. Skorecki. 1994. Isolation of promoter for cytosolic phospholipase A2 (cPLA2). Biochim. Biophys. Acta. 1217: 345–347. Cowan, M. J., X. L. Yao, R. Pawliczak, X. Huang, C. Logun, P. Madara, S. Alsaaty, T. Wu, and J. H. Shelhamer. 2004. The role of TFIID, the initiator element and a novel 5′ TFIID binding site in the transcriptional control of the TATA-less human cytosolic phospholipase A2-alpha promoter. Biochim. Biophys. Acta. 1680: 145–157. Heasley, L. E., S. Thaler, M. Nicks, B. Price, K. Skorecki, and R. A. Nemenoff. 1997. Induction of cytosolic phospholipase A2 by oncogenic Ras in human non-small cell lung cancer. J. Biol. Chem. 272: 14501–14504. Lee, C. W., C. C. Lin, I. T. Lee, H. C. Lee, and C. M. Yang. 2011. Activation and induction of cytosolic phospholipase A2 by TNF- mediated through Nox2, MAPKs, NF-B, and p300 in human tracheal smooth muscle cells. J. Cell. Physiol. 226: 2103–2114. Wick, M. J., S. A. Blaine, V. Van Putten, M. Saavedra, and R. A. Nemenoff. 2005. Lung Kruppel-like factor (LKLF) is a transcriptional activator of the cytosolic phospholipase A2 a promoter. Biochem. J. 387: 239–246. Buttar, N. S., C. J. DeMars, G. Lomberk, S. Rizvi, J. Bonilla-Velez, S. Achra, S. Rashtak, K. K. Wang, M. E. Fernandez-Zapico, and R. Urrutia. 2010. Distinct role of Kruppel-like factor 11 in the regulation of prostaglandin E2 biosynthesis. J. Biol. Chem. 285: 11433–11444. Alexandrov, P. N., J. G. Cui, and W. J. Lukiw. 2006. Hypoxiasensitive domain in the human cytosolic phospholipase A2 promoter. Neuroreport. 17: 303–307. Blaine, S. A., M. Wicks, C. Dessev, and R. A. Nemenoff. 2001. Induction of cPLA2 in lung epithelial cells and non-small cell lung cancer is mediated by Sp1 and c-Jun. J. Biol. Chem. 276: 42737–42743. Van Putten, V., Z. Refaat, C. Dessev, S. A. Blaine, M. J. Wick, L. Butterfield, S. Y. Han, L. E. Heasley, and R. A. Nemenoff. 2001. Induction of cytosolic phospholipase A2 by oncogenic Ras is mediated through the JNK and ERK pathways in rat epithelial cells. J. Biol. Chem. 276: 1226–1232. Tsou, J. H., K. Y. Chang, W. C. Wang, J. T. Tseng, W. C. Su, L. Y. Hung, W. C. Chang, and B. K. Chen. 2008. Nucleolin regulates c-Jun/Sp1-dependent transcriptional activation of cPLA2alpha in phorbol ester-treated non-small cell lung cancer A549 cells. Nucleic Acids Res. 36: 217–227. Dolan-O’Keefe, M., V. Chow, J. Monnier, G. A. Visner, and H. S. Nick. 2000. Transcriptional regulation and structural organization of the human cytosolic phospholipase A2 gene. Am. J. Physiol. Lung Cell. Mol. Physiol. 278: L649–L657. Sun, K., R. Ma, X. Cui, B. Campos, R. Webster, D. Brockman, and L. Myatt. 2003. Glucocorticoids induce cytosolic phospholipase A2 and prostaglandin H synthase type 2 but not microsomal prostaglandin E synthase (PGES) and cytosolic PGES expression in cultured primary human amnion cells. J. Clin. Endocrinol. Metab. 88: 5564–5571. Bickford, J. S., D. E. Beachy, K. J. Newsom, S. J. Barilovits, J. D. Herlihy, X. Qiu, J. N. Walters, N. Li, and H. S. Nick. 2013. A distal enhancer controls cytokine-dependent human cPLA2 gene expression. J. Lipid Res. 54: 1915–1926. Leslie, C. C., T. A. Gangelhoff, and M. H. Gelb. 2010. Localization and function of cytosolic phospholipase A2alpha at the Golgi. Biochimie. 92: 620–626. Channon, J. Y., and C. C. Leslie. 1990. A calcium-dependent mechanism for associating a soluble arachidonoyl-hydrolyzing phospholipase A2 with membrane in the macrophage cell line, RAW 264.7. J. Biol. Chem. 265: 5409–5413. Schievella, A. R., M. K. Regier, W. L. Smith, and L-L. Lin. 1995. Calcium-mediated translocation of cytosolic phospholipase A2 to the nuclear envelope and endoplasmic reticulum. J. Biol. Chem. 270: 30749–30754.

114. Chen, X. S., and C. D. Funk. 2001. The N-terminal “beta-barrel” domain of 5-lipoxygenase is essential for nuclear membrane translocation. J. Biol. Chem. 276: 811–818. 115. Rouzer, C. A., and S. Kargman. 1988. Translocation of 5-lipoxygenase to the membrane in human leukocytes challenged with ionophore A23187. J. Biol. Chem. 263: 10980–10988. 116. Evans, J. F., A. D. Ferguson, R. T. Mosley, and J. H. Hutchinson. 2008. What’s all the FLAP about?: 5-lipoxygenase-activating protein inhibitors for inflammatory diseases. Trends Pharmacol. Sci. 29: 72–78. 117. Mandal, A. K., P. B. Jones, A. M. Bair, P. Christmas, D. Miller, T. T. Yamin, D. Wisniewski, J. Menke, J. F. Evans, B. T. Hyman, et al. 2008. The nuclear membrane organization of leukotriene synthesis. Proc. Natl. Acad. Sci. USA. 105: 20434–20439. 118. Mandal, A. K., J. Skoch, B. J. Bacskai, B. T. Hyman, P. Christmas, D. Miller, T. D. Yamin, S. Xu, D. Wisniewski, J. F. Evans, et al. 2004. The membrane organization of leukotriene synthesis. Proc. Natl. Acad. Sci. USA. 101: 6587–6592. 119. Bair, A. M., M. V. Turman, C. A. Vaine, R. A. Panettieri, Jr., and R. J. Soberman. 2012. The nuclear membrane leukotriene synthetic complex is a signal integrator and transducer. Mol. Biol. Cell. 23: 4456–4464. 120. Smith, W. L., Y. Urade, and P. J. Jakobsson. 2011. Enzymes of the cyclooxygenase pathways of prostanoid biosynthesis. Chem. Rev. 111: 5821–5865. 121. Zhang, H., J. Fan, J. Wang, S. Zhang, B. Dou, and X. Peng. 2013. An off-on COX-2-specific fluorescent probe: targeting the Golgi apparatus of cancer cells. J. Am. Chem. Soc. 135: 11663–11669. 122. Yuan, C., and W. L. Smith. 2015. A cyclooxygenase-2 dependent prostaglandin E2 biosynthetic system in the Golgi apparatus. J. Biol. Chem. 290: 5606–5620. 123. Grewal, S., S. Ponnambalam, and J. H. Walker. 2003. Association of cPLA2-alpha and COX-1 with the Golgi apparatus of A549 human lung epithelial cells. J. Cell Sci. 116: 2303–2310. 124. Reed, K. A., D. E. Tucker, A. Aloulou, D. Adler, F. Ghomashchi, M. H. Gelb, C. C. Leslie, J. A. Oates, and O. Boutaud. 2011. Functional characterization of mutations in inherited human cPLA2 deficiency. Biochemistry. 50: 1731–1738. 125. Takaku, K., M. Sonoshita, N. Sasaki, N. Uozumi, Y. Doi, T. Shimizu, and M. M. Taketo. 2000. Suppression of intestinal polyposis in APC. J. Biol. Chem. 275: 34013–34016. 126. Montrose, D. C., K. Kadaveru, J. N. Ilsley, S. H. Root, T. V. Rajan, M. Ramesh, F. C. Nichols, B. T. Liang, D. Sonin, A. R. Hand, et al. 2010. cPLA2 is protective against COX inhibitor-induced intestinal damage. Toxicol. Sci. 117: 122–132. 127. Montrose, D. C., M. Nakanishi, R. C. Murphy, S. Zarini, J. P. McAleer, A. T. Vella, and D. W. Rosenberg. 2015. The role of PGE2 in intestinal inflammation and tumorigenesis. Prostaglandins Other Lipid Mediat. 116–117: 26–36. 128. Wang, D., and R. N. Dubois. 2010. The role of COX-2 in intestinal inflammation and colorectal cancer. Oncogene. 29: 781–788. 129. Wang, D., and R. N. DuBois. 2013. An inflammatory mediator, prostaglandin E2, in colorectal cancer. Cancer J. 19: 502–510. 130. Wang, D., and R. N. DuBois. 2013. The role of anti-inflammatory drugs in colorectal cancer. Annu. Rev. Med. 64: 131–144. 131. Ilsley, J. N., M. Nakanishi, C. Flynn, G. S. Belinsky, S. De Guise, J. N. Adib, R. T. Dobrowsky, J. V. Bonventre, and D. W. Rosenberg. 2005. Cytoplasmic phospholipase A2 deletion enhances colon tumorigenesis. Cancer Res. 65: 2636–2643. 132. Hong, K. H., J. C. Bonventre, E. O’Leary, J. V. Bonventre, and E. S. Lander. 2001. Deletion of cytosolic phospholipase A2 suppresses ApcMin-induced tumorigenesis. Proc. Natl. Acad. Sci. USA. 98: 3935–3939. 133. Oshima, M., J. E. Dinchuk, S. L. Kargman, H. Oshima, B. Hancock, E. Kwong, J. M. Trzaskos, J. F. Evans, and M. M. Taketo. 1996. Suppression of intestinal polyposis in Apc delta716 knockout mice by inhibitioon of cyclooxygenase 2 (COX-2). Cell. 87: 803–809. 134. Chan, T. A., P. J. Morin, B. Vogelstein, and K. W. Kinzler. 1998. Mechanisms underlying nonsteroidal antiinflammatory drugmediated apoptosis. Proc. Natl. Acad. Sci. USA. 95: 681–686. 135. Vassaux, G., D. Gaillard, C. Darimont, G. Ailhaud, and R. Negrel. 1992. Differential response of preadipocytes and adipocytes to prostacyclin and prostaglandin E2: physiological implications. Endocrinology. 131: 2393–2398. 136. Mazid, M. A., A. A. Chowdhury, K. Nagao, K. Nishimura, M. Jisaka, T. Nagaya, and K. Yokota. 2006. Endogenous 15-deoxyDelta(12,14)-prostaglandin J2 synthesized by adipocytes during

cPLA2: physiological function and role in disease

1399

Downloaded from www.jlr.org by guest, on August 13, 2015

96. Nakamura, H., Y. Moriyama, T. Makiyama, S. Emori, H. Yamashita, R. Yamazaki, and T. Murayama. 2013. Lactosylceramide interacts with and activates cytosolic phospholipase A2. J. Biol. Chem. 288: 23264–23272. 97. Bayburt, T., and M. H. Gelb. 1997. Interfacial catalysis by human 85 kDa cytosolic phospholipase A2 on anionic vesicles in the scooting mode. Biochemistry. 36: 3216–3231. 98. Kramer, R. M., E. F. Roberts, S. L. Um, A. G. Börsch-Haubold, S. P. Watson, M. J. Fisher, and J. A. Jakubowski. 1996. p38 mitogenactivated protein kinase phosphorylates cytosolic phospholipase A2 (cPLA2) in thrombin-stimulated platelets. J. Biol. Chem. 271: 27723–27729. 99. Casas, J., C. Meana, E. Esquinas, M. Valdearcos, J. Pindado, J. Balsinde, and M. A. Balboa. 2009. Requirement of JNK-mediated phosphorylation for translocation of group IVA phospholipase A2 to phagosomes in human macrophages. J. Immunol. 183: 2767–2774. 100. Gubern, A., M. Barcelo-Torns, D. Barneda, J. M. Lopez, R. Masgrau, F. Picatoste, C. E. Chalfant, J. Balsinde, M. A. Balboa, and E. Claro. 2009. JNK and ceramide kinase govern the biogenesis of lipid droplets through activation of group IVA phospholipase A2. J. Biol. Chem. 284: 32359–32369. 101. Pavicevic, Z., C. C. Leslie, and K. U. Malik. 2008. cPLA2 phosphorylation at serine-515 and serine-505 is required for arachidonic acid release in vascular smooth muscle cells. J. Lipid Res. 49: 724–737. 102. Hefner, Y., A. G. Borsch-Haubold, M. Murakami, J. I. Wilde, S. Pasquet, D. Schieltz, F. Ghomashchi, J. R. Yates 3rd, C. G. Armstrong, A. Paterson, et al. 2000. Serine 727 phosphorylation and activation of cytosolic phospholipase A2 by MNK1-related protein kinases. J. Biol. Chem. 275: 37542–37551. 103. Tian, W., G. T. Wijewickrama, J. H. Kim, S. Das, M. P. Tun, N. Gokhale, J. W. Jung, K. P. Kim, and W. Cho. 2008. Mechanism of regulation of group IVA phospholipase A2 activity by Ser727 phosphorylation. J. Biol. Chem. 283: 3960–3971. 104. Yao, X-L., M. J. Cowan, M. T. Gladwin, M. M. Lawrence, C. W. Angus, and J. H. Shelhamer. 1999. Dexamethasone alters arachidonate release from human epithelial cells by induction of p11 protein synthesis and inhibition of phospholipase A2 activity. J. Biol. Chem. 274: 17202–17208. 105. Wu, T., C. W. Angus, X-L. Yao, C. Logun, and J. H. Shelhamer. 1997. p11, a unique member of the S100 family of calcium-binding proteins, interacts with and inhibits the activity of the 85-kDa cytosolic phospholipase A2. J. Biol. Chem. 272: 17145–17153. 106. Mosior, M., D. A. Six, and E. A. Dennis. 1998. 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. 107. Casas, J., M. A. Gijon, A. G. Vigo, M. Sanchez Crespo, J. Balsince, and M. A. Balboa. 2006. Phosphatidylinositol 4,5-bisphosphate anchors cytosolic group IVA phospholipase A2 to perinuclear membranes and decreases its calcium requirement for translocation in live cells. Mol. Biol. Cell. 17: 155–162. 108. Leslie, C. C., and J. Y. Channon. 1990. Anionic phospholipids stimulate an arachidonoyl-hydrolyzing phospholipase A2 from macrophages and reduce the calcium requirement for activity. Biochim. Biophys. Acta. 1045: 261–270. 109. Rådmark, O., O. Werz, D. Steinhilber, and B. Samuelsson. 2015. 5-Lipoxygenase, a key enzyme for leukotriene biosynthesis in health and disease. Biochim. Biophys. Acta. 1851: 331–339. 110. Hammarberg, T., P. Provost, B. Persson, and O. Radmark. 2000. The N-terminal domain of 5-lipoxygenase binds calcium and mediates calcium stimulation of enzyme activity. J. Biol. Chem. 275: 38787–38793. 111. Brock, T. G., R. W. McNish, M. B. Bailie, and M. Peters-Golden. 1997. Rapid import of cytosolic 5-lipoxygenase into the nucleus of neutrophils after in vivo recruitment and in vitro adherence. J. Biol. Chem. 272: 8276–8280. 112. Woods, J. W., J. F. Evans, D. Ethier, S. Scott, P. J. Vickers, L. Hearn, J. A. Heibein, S. Charleson, and I. I. Singer. 1993. 5-lipoxygenase and 5-lipoxygenase-activating protein are localized in the nuclear envelope of activated human leukocytes. J. Exp. Med. 178: 1935–1946. 113. Kulkarni, S., S. Das, C. D. Funk, D. Murray, and W. Cho. 2002. Molecular basis of the specific subcellular localization of the C2-like domain of 5-lipoxygenase. J. Biol. Chem. 277: 13167– 13174.

137.

138.

139.

140.

141.

142.

144. 145. 146. 147.

148.

149. 150.

151.

152.

153.

154.

155. 156.

1400

Journal of Lipid Research Volume 56, 2015

157. Challis, J. R., D. M. Sloboda, N. Alfaidy, S. J. Lye, W. Gibb, F. A. Patel, W. L. Whittle, and J. P. Newnham. 2002. Prostaglandins and mechanisms of preterm birth. Reproduction. 124: 1–17. 158. Casey, M. L., P. C. MacDonald, and M. D. Mitchell. 1985. Despite a massive increase in cortisol secretion in women during parturition, there is an equally massive increase in prostaglandin synthesis. A paradox? J. Clin. Invest. 75: 1852–1857. 159. Guo, C., Z. Yang, W. Li, P. Zhu, L. Myatt, and K. Sun. 2008. Paradox of glucocorticoid-induced cytosolic phospholipase A2 group IVA messenger RNA expression involves glucocorticoid receptor binding to the promoter in human amnion fibroblasts. Biol. Reprod. 78: 193–197. 160. Guo, C., J. Li, L. Myatt, X. Zhu, and K. Sun. 2010. Induction of Galphas contributes to the paradoxical stimulation of cytosolic phospholipase A2alpha expression by cortisol in human amnion fibroblasts. Mol. Endocrinol. 24: 1052–1061. 161. Downey, P., A. Sapirstein, E. O’Leary, T. Sun, D. Brown, and J. V. Bonventre. 2001. Renal concentrating defect in mice lacking group IV cytosolic phospholipase A2. Am. J. Physiol. Renal Physiol. 280: F607–F618. 162. Li, Y., H. Yamada, Y. Kita, M. Kunimi, S. Horita, M. Suzuki, Y. Endo, T. Shimizu, G. Seki, and T. Fujita. 2008. Roles of ERK and cPLA2 in the angiotensin II-mediated biphasic regulation of Na+-HCO3(-) transport. J. Am. Soc. Nephrol. 19: 252–259. 163. Yoshioka, W., T. Kawaguchi, N. Fujisawa, K. Aida-Yasuoka, T. Shimizu, F. Matsumura, and C. Tohyama. 2014. Predominant role of cytosolic phospholipase A2 in dioxin-induced neonatal hydronephrosis in mice. Sci. Rep. 4: 4042. 164. Nishimura, N., F. Matsumura, C. F. Vogel, H. Nishimura, J. Yonemoto, W. Yoshioka, and C. Tohyama. 2008. Critical role of cyclooxygenase-2 activation in pathogenesis of hydronephrosis caused by lactational exposure of mice to dioxin. Toxicol. Appl. Pharmacol. 231: 374–383. 165. Kinehara, M., I. Fukuda, K. Yoshida, and H. Ashida. 2008. Highthroughput evaluation of aryl hydrocarbon receptor-binding sites selected via chromatin immunoprecipitation-based screening in Hepa-1c1c7 cells stimulated with 2,3,7,8-tetrachlorodibenzo-p-dioxin. Genes Genet. Syst. 83: 455–468. 166. Kinehara, M., I. Fukuda, K. Yoshida, and H. Ashida. 2009. Aryl hydrocarbon receptor-mediated induction of the cytosolic phospholipase A(2)alpha gene by 2,3,7,8-tetrachlorodibenzo-p-dioxin in mouse hepatoma Hepa-1c1c7 cells. J. Biosci. Bioeng. 108: 277–281. 167. Matsumura, F. 2009. The significance of the nongenomic pathway in mediating inflammatory signaling of the dioxin-activated Ah receptor to cause toxic effects. Biochem. Pharmacol. 77: 608–626. 168. Dong, B., N. Nishimura, C. F. Vogel, C. Tohyama, and F. Matsumura. 2010. TCDD-induced cyclooxygenase-2 expression is mediated by the nongenomic pathway in mouse MMDD1 macula densa cells and kidneys. Biochem. Pharmacol. 79: 487–497. 169. Le, T. D., Y. Shirai, T. Okamoto, T. Tatsukawa, S. Nagao, T. Shimizu, and M. Ito. 2010. Lipid signaling in cytosolic phospholipase A2alpha-cyclooxygenase-2 cascade mediates cerebellar longterm depression and motor learning. Proc. Natl. Acad. Sci. USA. 107: 3198–3203. 170. Wang, D. J., L. D. Su, Y. N. Wang, D. Yang, C. L. Sun, L. Zhou, X. X. Wang, and Y. Shen. 2014. Long-term potentiation at cerebellar parallel fiber-Purkinje cell synapses requires presynaptic and postsynaptic signaling cascades. J. Neurosci. 34: 2355–2364. 171. Su, L. D., D. J. Wang, D. Yang, Y. Shen, and Y. H. Hu. 2013. Retrograde cPLA2alpha/arachidonic acid/2-AG signaling is essential for cerebellar depolarization-induced suppression of excitation and long-term potentiation. Cerebellum. 12: 297–299. 172. Kishimoto, K., R. C. Li, J. Zhang, J. A. Klaus, K. K. Kibler, S. Dore, R. C. Koehler, and A. Sapirstein. 2010. Cytosolic phospholipase A2 alpha amplifies early cyclooxygenase-2 expression, oxidative stress and MAP kinase phosphorylation after cerebral ischemia in mice. J. Neuroinflammation. 7: 42. 173. Bosetti, F., and G. R. Weerasinghe. 2003. The expression of brain cyclooxygenase-2 is down-regulated in the cytosolic phospholipase A2 knockout mouse. J. Neurochem. 87: 1471–1477. 174. Sapirstein, A., H. Saito, S. J. Texel, T. A. Samad, E. O'Leary, and J. V. Bonventre. 2005. Cytosolic phospholipase A2alpha regulates induction of brain cyclooxygenase-2 in a mouse model of inflammation. Am. J. Physiol. Regul. Integr. Comp. Physiol. 288: R1774–R1782. 175. Andreasson, K. 2010. Emerging roles of PGE2 receptors in models of neurological disease. Prostaglandins Other Lipid Mediat. 91: 104–112.

Downloaded from www.jlr.org by guest, on August 13, 2015

143.

maturation phase contributes to upregulation of fat storage. FEBS Lett. 580: 6885–6890. Enomoto, N., K. Ikejima, S. Yamashina, A. Enomoto, T. Nishiura, T. Nishimura, D. A. Brenner, P. Schemmer, B. U. Bradford, C. A. Rivera, et al. 2000. Kupffer cell-derived prostaglandin E2 is involved in alcohol-induced fat accumulation in rat liver. Am. J. Physiol. Gastrointest. Liver Physiol. 279: G100–G106. Ii, H., S. Hatakeyama, K. Tsutsumi, T. Sato, and S. Akiba. 2008. Group IVA phospholipase A2 is associated with the storage of lipids in adipose tissue and liver. Prostaglandins Other Lipid Mediat. 86: 12–17. Ii, H., N. Yokoyama, S. Yoshida, K. Tsutsumi, S. Hatakeyama, T. Sato, K. Ishihara, and S. Akiba. 2009. Alleviation of high-fat diet-induced fatty liver damage in group IVA phospholipase A2knockout mice. PLoS ONE. 4: e8089. Ishihara, K., A. Miyazaki, T. Nabe, H. Fushimi, N. Iriyama, S. Kanai, T. Sato, N. Uozumi, T. Shimizu, and S. Akiba. 2012. Group IVA phospholipase A2 participates in the progression of hepatic fibrosis. FASEB J. 26: 4111–4121. Hsieh, P. S., J. S. Jin, C. F. Chiang, P. C. Chan, C. H. Chen, and K. C. Shih. 2009. COX-2-mediated inflammation in fat is crucial for obesity-linked insulin resistance and fatty liver. Obesity (Silver Spring). 17: 1150–1157. Huang, L., A. Zhao, F. Wong, J. M. Ayala, M. Struthers, F. Ujjainwalla, S. D. Wright, M. S. Springer, J. Evans, and J. Cui. 2004. Leukotriene B4 strongly increases monocyte chemoattractant protein-1 in human monocytes. Arterioscler. Thromb. Vasc. Biol. 24: 1783–1788. Efsen, E., A. Bonacchi, S. Pastacaldi, A. J. Valente, U. O. Wenzel, C. Tosti-Guerra, M. Pinzani, G. Laffi, H. E. Abboud, P. Gentilini, et al. 2001. Agonist-specific regulation of monocyte chemoattractant protein-1 expression by cyclooxygenase metabolites in hepatic stellate cells. Hepatology. 33: 713–721. Seki, E., S. de Minicis, S. Inokuchi, K. Taura, K. Miyai, N. van Rooijen, R. F. Schwabe, and D. A. Brenner. 2009. CCR2 promotes hepatic fibrosis in mice. Hepatology. 50: 185–197. Wynn, T. A. 2008. Cellular and molecular mechanisms of fibrosis. J. Pathol. 214: 199–210. Peters-Golden, M., C. Canetti, P. Mancuso, and M. J. Coffey. 2005. Leukotrienes: underappreciated mediators of innate immune responses. J. Immunol. 174: 589–594. Woodward, D. F., R. L. Jones, and S. Narumiya. 2011. International Union of Basic and Clinical Pharmacology. LXXXIII: classification of prostanoid receptors, updating 15 years of progress. Pharmacol. Rev. 63: 471–538. Lim, H., R. A. Gupta, W. G. Ma, B. C. Paria, D. E. Moller, J. D. Morrow, R. N. DuBois, J. M. Trzaskos, and S. K. Dey. 1999. Cyclooxygenase-2-derived prostacyclin mediates embryo implantation in the mouse via PPARdelta. Genes Dev. 13: 1561–1574. Lim, H., B. C. Paria, S. K. Das, J. E. Dinchuk, R. Langenbach, J. M. Trzaskos, and S. K. Dey. 1997. Multiple female reproductive failures in cyclooxygenase 2-deficient mice. Cell. 91: 197–208. Gross, G. A., T. Imamura, C. Luedke, S. K. Vogt, L. M. Olson, D. M. Nelson, Y. Sadovsky, and L. J. Muglia. 1998. Opposing actions of prostaglandins and oxytocin determine the onset of murine labor. Proc. Natl. Acad. Sci. USA. 95: 11875–11879. Song, H., H. Lim, B. C. Paria, H. Matsumoto, L. L. Swift, J. Morrow, J. V. Bonventre, and S. K. Dey. 2002. Cytosolic phospholipase A2alpha deficiency is crucial for ‘on-time’ embryo implantation that directs subsequent development. Development. 129: 2879–2889. Brown, N., J. D. Morrow, J. C. Slaughter, B. C. Paria, and J. Reese. 2009. Restoration of on-time embryo implantation corrects the timing of parturition in cytosolic phospholipase A2 group IVA deficient mice. Biol. Reprod. 81: 1131–1138. Kurusu, S., A. Sapirstein, and J. V. Bonventre. 2012. Group IVA phospholipase A2 optimizes ovulation and fertilization in rodents through induction of and metabolic coupling with prostaglandin endoperoxide synthase 2. FASEB J. 26: 3800–3810. Kurusu, S., A. Sapirstein, H. Sawada, M. Kawaminami, and J. V. Bonventre. 2011. Group IVA phospholipase A2 regulates testosterone biosynthesis by murine Leydig cells and is required for timely sexual maturation. Biochem. J. 439: 403–411. Smith, R. 2007. Parturition. N. Engl. J. Med. 356: 271–283. Challis, J. R. G., S. G. Matthews, W. Gibb, and S. J. Lye. 2000. Endocrine and paracrine regulation of birth at term and preterm. Endocr. Rev. 21: 514–550.

195.

196.

197.

198.

199. 200. 201. 202. 203.

204. 205.

206.

207.

208.

209. 210.

211.

212.

in vivo characterization of 4-{3-[5-chloro-2-(2-{[3,4-dichlorobenzyl)sulfonyl]amino}ethyl)-1-(diphenylmethyl)-1H-indol-3-yl]propyl}benzoic acid, efipladib. J. Med. Chem. 51: 3388–3413. Whalen, K. A., H. Legault, C. Hang, A. Hill, M. Kasaian, D. Donaldson, G. W. Gensch, G. Gensch, J. Baker, P. S. Reddy, et al. 2008. In vitro allergen challenge of peripheral blood induces differential gene expression in mononuclear cells of asthmatic patients: inhibition of cytosolic phospholipase A2alpha overcomes the asthma-associated response. Clin. Exp. Allergy. 38: 1590–1605. Granata, F., R. I. Staiano, S. Loffredo, A. Petraroli, A. Genovese, G. Marone, and M. Triggiani. 2010. The role of mast cell-derived secreted phospholipases A2 in respiratory allergy. Biochimie. 92: 588–593. Sokolowska, M., M. Borowiec, A. Ptasinska, M. Cieslak, J. H. Shelhamer, M. L. Kowalski, and R. Pawliczak. 2007. 85-kDa cytosolic phospholipase A2 group IValpha gene promoter polymorphisms in patients with severe asthma: a gene expression and case-control study. Clin. Exp. Immunol. 150: 124–131. Sokolowska, M., J. Stefanska, K. Wodz-Naskiewicz, M. Cieslak, and R. Pawliczak. 2010. Cytosolic phospholipase A2 group IVA is overexpressed in patients with persistent asthma and regulated by the promoter microsatellites. J. Allergy Clin. Immunol. 125: 1393–1395. Nagase, T., N. Uozumi, S. Ishii, K. Kume, T. Izumi, Y. Ouchi, and T. Shimizu. 2000. Acute lung injury by sepsis and acid aspiration: a key role for cytosolic phospholipase A2. Nat. Immunol. 1: 42–46. Uozumi, N., Y. Kita, and T. Shimizu. 2008. Modulation of lipid and protein mediators of inflammation by cytosolic phospholipase A2alpha during experimental sepsis. J. Immunol. 181: 3558–3566. Touqui, L. 2012. Antisense inhibition of phospholipase A2: a new approach for already tested therapeutic targets for the treatment of sepsis. Crit. Care Med. 40: 2250–2251. Liu, M. S., C. H. Liu, G. Wu, and Y. Zhou. 2012. Antisense inhibition of secretory and cytosolic phospholipase A2 reduces the mortality in rats with sepsis. Crit. Care Med. 40: 2132–2140. Selman, M., T. E. King, A. Pardo,; American Thoracic Society; European Respiratory Society; and American College of Chest Physicians. 2001. Idiopathic pulmonary fibrosis: prevailing and evolving hypotheses about its pathogenesis and implications for therapy. Ann. Intern. Med. 134: 136–151. Wynn, T. A., and T. R. Ramalingam. 2012. Mechanisms of fibrosis: therapeutic translation for fibrotic disease. Nat. Med. 18: 1028–1040. Nagase, T., N. Uozumi, S. Ishii, Y. Kita, H. Yamamoto, E. Ohga, Y. Ouchi, and T. Shimizu. 2002. A pivotal role of cytosolic phospholipase A2 in bleomycin-induced pulmonary fibrosis. Nat. Med. 8: 480–484. Beller, T. C., D. S. Friend, A. Maekawa, B. K. Lam, K. F. Austen, and Y. Kanaoka. 2004. Cysteinyl leukotriene 1 receptor controls the severity of chronic pulmonary inflammation and fibrosis. Proc. Natl. Acad. Sci. USA. 101: 3047–3052. Beller, T. C., A. Maekawa, D. S. Friend, K. F. Austen, and Y. Kanaoka. 2004. Targeted gene disruption reveals the role of the cysteinyl leukotriene 2 receptor in increased vascular permeability and in bleomycin-induced pulmonary fibrosis in mice. J. Biol. Chem. 279: 46129–46134. Peters-Golden, M., M. Bailie, T. Marshall, C. Wilke, S. H. Phan, G. B. Toews, and B. B. Moore. 2002. Protection from pulmonary fibrosis in leukotriene-deficient mice. Am. J. Respir. Crit. Care Med. 165: 229–235. Bozyk, P. D., and B. B. Moore. 2011. Prostaglandin E2 and the pathogenesis of pulmonary fibrosis. Am. J. Respir. Cell Mol. Biol. 45: 445–452. Lovgren, A. K., L. A. Jania, J. M. Hartney, K. K. Parsons, L. P. Audoly, G. A. Fitzgerald, S. L. Tilley, and B. H. Koller. 2006. COX-2-derived prostacyclin protects against bleomycin-induced pulmonary fibrosis. Am. J. Physiol. Lung Cell. Mol. Physiol. 291: L144–L156. Keerthisingam, C. B., R. G. Jenkins, N. K. Harrison, N. A. Hernandez-Rodriguez, H. Booth, G. J. Laurent, S. L. Hart, M. L. Foster, and R. J. McAnulty. 2001. Cyclooxygenase-2 deficiency results in a loss of the anti-proliferative response to transforming growth factor-beta in human fibrotic lung fibroblasts and promotes bleomycin-induced pulmonary fibrosis in mice. Am. J. Pathol. 158: 1411–1422. Bonner, J. C., A. B. Rice, J. L. Ingram, C. R. Moomaw, A. Nyska, A. Bradbury, A. R. Sessoms, P. C. Chulada, D. L. Morgan, D. C. Zeldin, et al. 2002. Susceptibility of cyclooxygenase-2-deficient mice to pulmonary fibrogenesis. Am. J. Pathol. 161: 459–470.

cPLA2: physiological function and role in disease

1401

Downloaded from www.jlr.org by guest, on August 13, 2015

176. Liu, N. K., L. X. Deng, Y. P. Zhang, Q. B. Lu, X. F. Wang, J. G. Hu, E. Oakes, J. V. Bonventre, C. B. Shields, and X. M. Xu. 2014. Cytosolic phospholipase A2 protein as a novel therapeutic target for spinal cord injury. Ann. Neurol. 75: 644–658. 177. López-Vales, R., N. Ghasemlou, A. Redensek, B. J. Kerr, E. Barbayianni, G. Antonopoulou, C. Baskakis, K. I. Rathore, V. ConstantinouKokotou, D. Stephens, et al. 2011. Phospholipase A2 superfamily members play divergent roles after spinal cord injury. FASEB J. 25: 4240–4252. 178. Watanabe, H., K. Numata, T. Ito, K. Takagi, and A. Matsukawa. 2004. Innate immune response in Th1- and Th2-dominant mouse strains. Shock. 22: 460–466. 179. Kennedy, B. P., P. Payette, J. Mudgett, P. Vadas, W. Pruzanski, M. Kwan, C. Tang, D. E. Rancourt, and W. A. Cromlish. 1995. A natural disruption of the secretory group II phospholipase A2 gene in inbred mouse strains. J. Biol. Chem. 270: 22378–22385. 180. Sun, G. Y., P. B. Shelat, M. B. Jensen, Y. He, A. Y. Sun, and A. Simonyi. 2010. Phospholipases A2 and inflammatory responses in the central nervous system. Neuromolecular Med. 12: 133–148. 181. Sun, G. Y., Y. He, D. Y. Chuang, J. C. Lee, Z. Gu, A. Simonyi, and A. Y. Sun. 2012. Integrating cytosolic phospholipase A2 with oxidative/ nitrosative signaling pathways in neurons: a novel therapeutic strategy for AD. Mol. Neurobiol. 46: 85–95. 182. Kriem, B., I. Sponne, A. Fifre, C. Malaplate-Armand, K. Lozac'hPillot, V. Koziel, F. T. Yen-Potin, B. Bihain, T. Oster, J. L. Olivier, et al. 2005. Cytosolic phospholipase A2 mediates neuronal apoptosis induced by soluble oligomers of the amyloid-beta peptide. FASEB J. 19: 85–87. 183. Shelat, P. B., M. Chalimoniuk, J. H. Wang, J. B. Strosznajder, J. C. Lee, A. Y. Sun, A. Simonyi, and G. Y. Sun. 2008. Amyloid beta peptide and NMDA induce ROS from NADPH oxidase and AA release from cytosolic phospholipase A2 in cortical neurons. J. Neurochem. 106: 45–55. 184. Sanchez-Mejia, R. O., J. W. Newman, S. Toh, G. Q. Yu, Y. Zhou, B. Halabisky, M. Cisse, K. Scearce-Levie, I. H. Cheng, L. Gan, et al. 2008. Phospholipase A2 reduction ameliorates cognitive deficits in a mouse model of Alzheimer’s disease. Nat. Neurosci. 11: 1311–1318. 185. Desbene, C., C. Malaplate-Armand, I. Youssef, P. Garcia, C. Stenger, M. Sauvee, N. Fischer, D. Rimet, V. Koziel, M. C. Escanye, et al. 2012. Critical role of cPLA2 in A oligomer-induced neurodegeneration and memory deficit. Neurobiol. Aging 33: 1123. e17–1123.e29. 186. Sagy-Bross, C., K. Kasianov, Y. Solomonov, A. Braiman, A. Friedman, N. Hadad, and R. Levy. 2015. The role of cytosolic phospholipase A2 in amyloid precursor protein induction by amyloid beta1-42: implication for neurodegeneration. J. Neurochem. 132: 559–571. 187. Last, V., A. Williams, and D. Werling. 2012. Inhibition of cytosolic phospholipase A2 prevents prion peptide-induced neuronal damage and co-localisation with beta III tubulin. BMC Neurosci. 13: 106. 188. Bate, C., M. Tayebi, and A. Williams. 2010. Phospholipase A2 inhibitors protect against prion and Abeta mediated synapse degeneration. Mol. Neurodegener. 5: 13. 189. Klivenyi, P., M. F. Beal, R. J. Ferrante, O. A. Andreassen, M. Wermer, M. Chin, and J. V. Bonventre. 1998. Mice deficient in group IV cytosolic phospholipase A2 are resistant to MPTP neurotoxicity. J. Neurochem. 71: 2634–2637. 190. Qu, B. X., Y. Gong, D. Sinclair, M. Fu, D. Perl, and R. DiazArrastia. 2013. cPLA2 knockout mice exhibit abnormalities in the architecture and synapses of cortical neurons. Brain Res. 1497: 101–105. 191. Rosenberger, T. A., N. E. Villacreses, M. A. Contreras, J. V. Bonventre, and S. I. Rapoport. 2003. Brain lipid metabolism in the cPLA2 knockout mouse. J. Lipid Res. 44: 109–117. 192. Irvin, C. G., Y. P. Tu, J. R. Sheller, and C. D. Funk. 1997. 5-Lipoxygenase products are necessary for ovalbumin-induced airway responsiveness in mice. Am. J. Physiol. 272: L1053–L1058. 193. Gavett, S. H., S. L. Madison, P. C. Chulada, P. E. Scarborough, W. Qu, J. E. Boyle, H. F. Tiano, C. A. Lee, R. Langenbach, V. L. Roggli, et al. 1999. Allergic lung responses are increased in prostaglandin H synthase-deficient mice. J. Clin. Invest. 104: 721–732. 194. McKew, J. C., K. L. Lee, M. W. Shen, P. Thakker, M. A. Foley, M. L. Behnke, B. Hu, F-W. Sum, S. Tam, Y. Hu, et al. 2008. Indole cytosolic phospholipase A2 a inhibitors: discovery and in vitro and

1402

Journal of Lipid Research Volume 56, 2015

226. Griffiths, R. J., M. A. Smith, M. L. Roach, J. L. Stock, E. J. Stam, A. J. Milici, D. N. Scampoli, J. D. Eskra, R. S. Byrum, B. H. Koller, et al. 1997. Collagen-induced arthritis is reduced in 5-lipoxygenaseactivating protein-deficient mice. J. Exp. Med. 185: 1123–1129. 227. Myers, L. K., A. H. Kang, A. E. Postlethwaite, E. F. Rosloniec, S. G. Morham, B. V. Shlopov, S. Goorha, and L. R. Ballou. 2000. The genetic ablation of cyclooxygenase 2 prevents the development of autoimmune arthritis. Arthritis Rheum. 43: 2687–2693. 228. Hegen, M., L. Sun, N. Uozumi, K. Kume, M. E. Goad, C. L. Nickerson-Nutter, T. Shimizu, and J. D. Clark. 2003. Cytosolic phospholipase A2alpha-deficient mice are resistant to collageninduced arthritis. J. Exp. Med. 197: 1297–1302. 229. Miyaura, C., M. Inada, C. Matsumoto, T. Ohshiba, N. Uozumi, T. Shimizu, and A. Ito. 2003. An essential role of cytosolic phospholipase A2alpha in prostaglandin E2-mediated bone resorption associated with inflammation. J. Exp. Med. 197: 1303–1310. 230. Oikawa, Y., E. Yamato, F. Tashiro, M. Yamamoto, N. Uozumi, A. Shimada, T. Shimizu, and J. Miyazaki. 2005. Protective role for cytosolic phospholipase A2alpha in autoimmune diabetes of mice. FEBS Lett. 579: 3975–3978. 231. Martin, R., and H. F. McFarland. 1995. Immunological aspects of experimental allergic encephalomyelitis and multiple sclerosis. Crit. Rev. Clin. Lab. Sci. 32: 121–182. 232. Emerson, M. R., and S. M. LeVine. 2004. Experimental allergic encephalomyelitis is exacerbated in mice deficient for 12/ 15-lipoxygenase or 5-lipoxygenase. Brain Res. 1021: 140–145. 233. Miyamoto, K., S. Miyake, M. Mizuno, N. Oka, S. Kusunoki, and T. Yamamura. 2006. Selective COX-2 inhibitor celecoxib prevents experimental autoimmune encephalomyelitis through COX-2independent pathway. Brain. 129: 1984–1992. 234. Ni, J., Y. Y. Shu, Y. N. Zhu, Y. F. Fu, W. Tang, X. G. Zhong, H. Wang, Y. F. Yang, J. Ren, M. W. Wang, et al. 2007. COX-2 inhibitors ameliorate experimental autoimmune encephalomyelitis through modulating IFN-gamma and IL-10 production by inhibiting T-bet expression. J. Neuroimmunol. 186: 94–103. 235. Marusic, S., M. W. Leach, J. W. Pelker, M. L. Azoitei, N. Uozumi, J. Cui, M. W. Shen, C. M. DeClercq, J. S. Miyashiro, B. A. Carito, et al. 2005. Cytosolic phospholipase A2alpha-deficient mice are resistant to experimental autoimmune encephalomyelitis. J. Exp. Med. 202: 841–851. 236. Marusic, S., P. Thakker, J. W. Pelker, N. L. Stedman, K. L. Lee, J. C. McKew, L. Han, X. Xu, S. F. Wolf, A. J. Borey, et al. 2008. Blockade of cytosolic phospholipase A2alpha prevents experimental autoimmune encephalomyelitis and diminishes development of Th1 and Th17 responses. J. Neuroimmunol. 204: 29–37. 237. Yang, D., H. F. Ji, X. M. Zhang, H. Yue, L. Lin, Y. Y. Ma, X. N. Huang, J. Fu, and W. Z. Wang. 2014. Protective effect of cytosolic phospholipase A2 inhibition against inflammation and degeneration by promoting regulatory T cells in rats with experimental autoimmune encephalomyelitis. Mediators Inflamm. 2014: 890139. 238. Schiffmann, S., A. Weigert, J. Mannich, M. Eberle, K. Birod, A. Haussler, N. Ferreiros, Y. Schreiber, H. Kunkel, M. Grez, et al. 2014. PGE2/EP4 signaling in peripheral immune cells promotes development of experimental autoimmune encephalomyelitis. Biochem. Pharmacol. 87: 625–635.

Downloaded from www.jlr.org by guest, on August 13, 2015

213. Rancoule, C., J. P. Pradere, J. Gonzalez, J. Klein, P. Valet, J. L. Bascands, J. P. Schanstra, and J. S. Saulnier-Blache. 2011. Lysophosphatidic acid-1-receptor targeting agents for fibrosis. Expert Opin. Investig. Drugs. 20: 657–667. 214. Tager, A. M., P. LaCamera, B. S. Shea, G. S. Campanella, M. Selman, Z. Zhao, V. Polosukhin, J. Wain, B. A. Karimi-Shah, N. D. Kim, et al. 2008. The lysophosphatidic acid receptor LPA1 links pulmonary fibrosis to lung injury by mediating fibroblast recruitment and vascular leak. Nat. Med. 14: 45–54. 215. Aoki, J., A. Inoue, and S. Okudaira. 2008. Two pathways for lysophosphatidic acid production. Biochim. Biophys. Acta. 1781: 513– 518. 216. Aikawa, S., T. Hashimoto, K. Kano, and J. Aoki. 2015. Lysophosphatidic acid as a lipid mediator with multiple biological actions. J. Biochem. 157: 81–89. 217. Ma, L., H. Uchida, J. Nagai, M. Inoue, J. Aoki, and H. Ueda. 2010. Evidence for de novo synthesis of lysophosphatidic acid in the spinal cord through phospholipase A2 and autotaxin in nerve injury-induced neuropathic pain. J. Pharmacol. Exp. Ther. 333: 540–546. 218. Torres, A. C., D. Boruszewska, M. Batista, I. Kowalczyk-Zieba, P. Diniz, E. Sinderewicz, J. S. Saulnier-Blache, I. Woclawek-Potocka, and L. Lopes-da-Costa. 2014. Lysophosphatidic acid signaling in late cleavage and blastocyst stage bovine embryos. Mediators Inflamm. 2014: 678968. 219. Meyer, A. M., L. D. Dwyer-Nield, G. L. Hurteau, R. L. Keith, E. O’Leary, M. You, J. V. Bonventre, R. A. Nemenoff, and A. M. Malkinson. 2004. Decreased lung tumorigenesis in mice genetically deficient in cytosolic phospholipase A2. Carcinogenesis. 25: 1517–1524. 220. Weiser-Evans, M. C., X. Q. Wang, J. Amin, V. Van Putten, R. Choudhary, R. A. Winn, R. Scheinman, P. Simpson, M. W. Geraci, and R. A. Nemenoff. 2009. Depletion of cytosolic phospholipase A2 in bone marrow-derived macrophages protects against lung cancer progression and metastasis. Cancer Res. 69: 1733–1738. 221. Linkous, A., L. Geng, A. Lyshchik, D. E. Hallahan, and E. M. Yazlovitskaya. 2009. Cytosolic phospholipase A2: targeting cancer through the tumor vasculature. Clin. Cancer Res. 15: 1635– 1644. 222. Yazlovitskaya, E. M., A. G. Linkous, D. K. Thotala, K. C. Cuneo, and D. E. Hallahan. 2008. Cytosolic phospholipase A2 regulates viability of irradiated vascular endothelium. Cell Death Differ. 15: 1641–1653. 223. Thotala, D., J. M. Craft, D. J. Ferraro, R. P. Kotipatruni, S. R. Bhave, J. J. Jaboin, and D. E. Hallahan. 2013. Cytosolic phospholipaseA2 inhibition with PLA-695 radiosensitizes tumors in lung cancer animal models. PLoS ONE. 8: e69688. 224. Linkous, A. G., E. M. Yazlovitskaya, and D. E. Hallahan. 2010. Cytosolic phospholipase A2 and lysophospholipids in tumor angiogenesis. J. Natl. Cancer Inst. 102: 1398–1412. 225. Griffiths, R. J., E. R. Pettipher, K. Koch, C. A. Farrell, R. Breslow, M. J. Conklyn, M. A. Smith, B. C. Hackman, D. J. Wimberly, A. J. Milici, et al. 1995. Leukotriene B4 plays a critical role in the progression of collagen-induced arthritis. Proc. Natl. Acad. Sci. USA. 92: 517–521.

Cytosolic phospholipase A₂: physiological function and role in disease.

The group IV phospholipase A2 (PLA2) family is comprised of six intracellular enzymes (GIVA, -B, -C, -D, -E, and -F) commonly referred to as cytosolic...
1MB Sizes 2 Downloads 7 Views