Next Article in Journal
The Emerging Role of Epigenetic Mechanisms in the Causation of Aberrant MMP Activity during Human Pathologies and the Use of Medicinal Drugs
Next Article in Special Issue
Apolipoprotein Mimetic Peptides: An Emerging Therapy against Diabetic Inflammation and Dyslipidemia
Previous Article in Journal
The Role of Exosomes in Lysosomal Storage Disorders
Previous Article in Special Issue
Eicosapentaenoic Acid Inhibits KRAS Mutant Pancreatic Cancer Cell Growth by Suppressing Hepassocin Expression and STAT3 Phosphorylation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Impact of the Ca2+-Independent Phospholipase A2β (iPLA2β) on Immune Cells

Department of Cell, Developmental and Integrative Biology, Comprehensive Diabetes Center, University of Alabama at Birmingham, Birmingham, AL 35294, USA
*
Author to whom correspondence should be addressed.
These authors contributed equally.
Biomolecules 2021, 11(4), 577; https://doi.org/10.3390/biom11040577
Submission received: 5 March 2021 / Revised: 6 April 2021 / Accepted: 12 April 2021 / Published: 15 April 2021
(This article belongs to the Collection Bioactive Lipids in Inflammation, Diabetes and Cancer)

Abstract

:
The Ca2+-independent phospholipase A2β (iPLA2β) is a member of the PLA2 family that has been proposed to have roles in multiple biological processes including membrane remodeling, cell proliferation, bone formation, male fertility, cell death, and signaling. Such involvement has led to the identification of iPLA2β activation in several diseases such as cancer, cardiovascular abnormalities, glaucoma, periodontitis, neurological disorders, diabetes, and other metabolic disorders. More recently, there has been heightened interest in the role that iPLA2β plays in promoting inflammation. Recognizing the potential contribution of iPLA2β in the development of autoimmune diseases, we review this issue in the context of an iPLA2β link with macrophages and T-cells.

1. Introduction

Phospholipases A2 (PLA2s) hydrolyze the sn-2 substituent of glycerophospholipids to release a lysophospholipid and a free fatty acid [1]. Among the family of PLA2s are the group VI Ca2+-independent PLA2s (iPLA2s), which include iPLA2β (VIA), iPLA2γ (VIB) iPLA2δ (VIC), iPLA2ε (VID), iPLA2ζ (VIE), and iPLA2η (VIF) [2]. The cytosolic iPLA2β enzyme, along with the membrane-associated iPLA2γ, are the most studied of the group VI PLA2s. Because of its emerging link with inflammation, the focus of this review will be on iPLA2β.
Encoded by PLA2G6, the iPLA2β protein (84-88 kDa) has a serine lipase consensus sequence (GTSGT), contains eight N-terminal ankyrin repeats, an ATP binding cassette (GGGVKG), a caspase-3 cleavage site (DVDT), and two calmodulin (CAM) binding domains [2,3]. Crystal structure studies suggest that iPLA2β forms a dimer through the interaction of the catalytic domains and that CAM binds to the dimer to cause a closed state, denying the access of substrates to the active site [4]. Relief from this inhibitory state is achieved through activation of calmodulin kinase IIβ, which forms a signaling complex with iPLA2β [5]. Lipidomics-based LC/MS approaches have revealed that among glycerophospholipids, iPLA2β exhibited the greatest activity towards 1-palmitoyl-2-arachidonoyl-sn-glycero-3-phosphoethanolamine (PAPE) and, among sn-2 substituents, a greater selectivity for linoleate or myristate [6,7]. The currently available selective inhibitors of iPLA2β include the irreversible S-BEL (S-bromoenol lactone) and the reversible 1,1,1-trifluoro-6-(naphthalen-2-yl)hexan-2-one (FKGK18) [8,9]. iPLA2β is widely expressed and has many proposed roles, including signal transduction, and is recognized to contribute to neurodegenerative disorders, cancers, myocardial complications, and metabolic dysfunction (reviewed extensively elsewhere [2,3]).
While localized in the cytosol under basal conditions [10], iPLA2β mobilizes to subcellular organelles (e.g., the endoplasmic reticulum (ER), mitochondria, nucleus) upon stimulation. Its activation leads to the hydrolysis of the sn-2 fatty acid substituent from membrane glycerophospholipids [11,12,13,14,15,16] to yield a free fatty acid (e.g., arachidonic acid, eicosapentaenoic acid (EPA), docosahexaenoic acid (DHA)) and a 2-lysophospholipid [17]. Subsequent metabolism of arachidonic acid by cyclooxygenases (COX), lipoxygenase (LOX), and cytochrome P450 (CYP) pathways leads to the generation of bioactive oxidized eicosanoids, several of which are proinflammatory [18] and recognized contributors to inflammatory diseases [19,20,21,22,23,24,25,26,27]. Some of the most potent inflammatory eicosanoids [21] are prostaglandin E2 (PGE2), leukotrienes (LTs), hydroxyeicosatetraenoic acids (HETEs), and dihydroxyeicosatetraenoic acids (DHETEs), and they contribute to inflammation and autoimmune diseases [22]. It is not unexpected that iPLA2β activation can play critical roles in the initiation and progression of inflammatory pathways, which if not curbed can lead to the evolution of a variety of disorders. In contrast, other unsaturated fatty acids such as EPA and DHA can be metabolized to generate pro-resolving lipids, designated as specialized resolving mediators (SPMs) [28,29,30]. Among the proinflammatory lipids generated, PGs and LTs are the first to be produced [28]. When there is injury to the tissue, the generated PGs cause pain, swelling, and edema. To resolve the inflammation and clear antigenic debris, SPMs are generated in attempt to restore homeostasis, a process referred to as “lipid mediator class switching” [31,32].
Immune cells express iPLA2β [33,34,35,36] and iPLA2β-derived lipids (iDLs) contribute to cell proliferation [37], cell cycle progression [38], cell division [39], monocyte migration [40], and superoxide generation [41]. Inhibition of iPLA2β reduces reactive oxygen species (ROS) generation [42] and is reported to be effective against autoimmune- [43] and inflammation-based [44,45,46,47] diseases.
Although extensive literature exists linking a number of secretory PLA2s (sPLA2s) or cytosolic PLA2α (cPLA2α) with inflammatory responses, only a few studies have described a link between iPLA2β and macrophages, and even fewer have considered a link between iPLA2β and T-cells or B-cells. To provide context, in this paper, the basic functionality of these cells is discussed first, followed by a review of the specific impact of iPLA2β in these immune cells, in the context of different pathophysiologies.

2. Macrophages

Macrophages are recognized to play significant roles in the development of autoimmune diseases, including rheumatoid arthritis (RA), multiple sclerosis (MS), and type-1 diabetes (T1D) [48]. As fundamental components of the innate immune system, they contribute to tissue homeostasis, dead cell and antigen phagocytosis, and crucial induction of adaptive immunity [49]. They do so by recognizing pathogen-associated molecular patterns (PAMPs) of foreign microorganism or damage-associated molecular patterns (DAMPs) from damaged host cells through pattern recognition receptors (PRR). PPRs are found in all innate immune cells including macrophages. Following their recognition, macrophages internalize the PAMPs or DAMPs via phagocytosis and break them down for antigen presentation through major histocompatibility complex II (MHC-II) [50], which is another fundamental role of macrophages [51]. T lymphocytes recognize self-peptides that are presented on the MHC Class-II molecules and expressed mainly by antigen presenting cells (APC) like dendritic cells and macrophages [52]. All tissues have resident macrophages, which emerge from three different developmental stages—the yolk sac, fetal liver, and bone marrow [53,54]. Although fetal-liver and bone marrow-derived monocytes are the main source of macrophages after birth, yolk sac-derived macrophages give rise to microglia and a small fraction of tissue-derived macrophages [55].
Macrophages are essential in both inflammation and healing processes. However, the functionality of macrophages is subject to macrophage polarization, defined as the “M1-M2 paradigm” [56]. Tissue-resident macrophages are mainly of the M2 (anti-inflammatory) phenotype [57] and affect tissue homeostasis and the resolution of inflammation, which also involves debris clearance [58]. Polarization of macrophages to the M2 phenotype is induced by Th2 cytokines, including interleukins 4, 13, and 10 (IL-4, IL-13, and IL-10). This leads to upregulation of arginase-1 (Arg1), chemokine (C-C motif) ligands 17 and 24 (CCL17 and CCL24), and secretion of IL-10 and transforming growth factor beta (TGFβ) [59]. In contrast, the M1 macrophages express a proinflammatory phenotype and are induced by Th1 cytokines that include interferon gamma (IFNγ) and TNFα. This promotes upregulation of Arg2, iNOS, cyclooxygenase-2 (COX-2), and other proinflammatory factors, leading to increased production of nitric oxide (NO), TNFα, IL-1α, IL-1β, IL-6, and IL-12 [60]. Whereas M2-like macrophages represent the majority of tissue-resident macrophages [57,61], M1-like macrophages are found significantly in inflamed tissues [60].
Emerging studies highlight the importance of lipids in macrophage functionality [62,63]. Although macrophage signaling has long been studied in the context of cytokine and chemokine production, recent studies suggest a profound impact of eicosanoids and SPMs on macrophage functions [31,64]. Eicosanoids have been linked to modifying macrophage inflammatory response in multiple diseases [23,65,66]. In humans, PGE2 induces LOX-class switching from leukotriene B4 (LTB4) to lipoxins, which represents a stop signal for polymorphonuclear (PMN) recruitment and initiation of a resolution phase that promotes an anti-inflammatory macrophage phenotype and function (i.e., phagocytosis) [30]. Facilitating the resolution process are SPMs, which counter-regulate the early initiators (PGs and LTs) of acute inflammation, leading to inhibition of proinflammatory cytokines and upregulation of anti-inflammatory cytokines (e.g., IL-10) [29].

3. T-Cells and B-Cells

Lymphocyte development in mammals occurs in the central lymphoid organs such as the bone marrow (B-cell development) and the thymus (T-cell development) [67]. From these organs, T and B lymphocytes continue to migrate to other peripheral lymphoid tissues. Differences in both lymphocyte populations occurs in mature stages, where T-cell development slows down in the thymus, and they divide outside of the central lymphoid organs to maintain the number of mature T-cells [67].
The development of a cluster of either differentiation 4 or differentiation 8 (CD4+ or CD8+) T-cells depends on the thymocyte maturation process. This involves hematopoietic stem cells in the bone marrow migrating to the thymus, hence the “T” in T-cells, to mature into functional T-cells [68]. The earliest thymocytes are double-negative—they do not express either CD4 or CD8. In the double negative (DN) stage, these cells are negative for T-cell receptor (TCR) expression, are not able to bind to MHC, and do not carry out effector functions. The DN stage occurs in the cortex and has four subsets that are important for the full development of T-cells, designated as DN1–4. DN3 is a critical step in T-cell development and encompasses: (1) thymocytes becoming restricted to a T-lineage and generation of mature αβ and γδ T-cells, (2) T-cell receptor β-chain (TCRβ) V(D)J recombination, (3) expression of functional pre-TCR complex, and (4) thymocytes becoming early pre-T-cells that are committed to the αβ T-cell lineage. The DN4 cells start to express low levels of CD4 and CD8 and present themselves as double-positive (DP) [68]. During the DP stages, CD4 and CD8 expression levels are upregulated and the cells migrate to the medulla, where a negative selection process confers single positivity for CD4 or CD8.
Naïve CD4 T-cells are destined for four distinct fates that are determined by different cytokine patterns after interaction with antigens. These cytokine signals upregulate distinct transcription factors to generate specific cell populations, which include the T helper cells 1, 2, and 17 (Th1, Th2, Th17), and regulatory T-cells (Tregs) [69].
B lymphocytes also go through different stages, including (1) activation, (2) selection, and (3) maturation. Like T lymphocytes, B-cells are derived from hematopoietic stem cells (HSCs) that go through stages in the bone marrow. These stages include the pro-B-cell (progenitor B-cell), the pre-B-cell (precursor B-cell), the immature naïve B-cell, and lastly the mature naïve B-cell stage [70]. After maturation, the naïve B-cells migrate to a secondary lymphoid tissue (e.g., lymph nodes), where they can be activated by antigens to generate memory B cells and plasma antibody-secreting cells.
T and B lymphocytes are major cellular components for autoimmunity. B-cells are responsible for antigen presentation, antibody production, and cytokine production. However, T-cells take part in both “helper” and cytotoxic activity. T-helper cells, also known as CD4+ T-cells, are a type of T-cells that play an essential role in adaptive immune responses. Cytotoxic T-cells, CD8+ T-cells, are identified as T-cells that kill cancer cells or infected cells.
Both CD4 and CD8 T lymphocyte types contribute to the adaptive immunity component of the immune system. There are two distinct types of immunity based on B- and T-cell function: humoral and cell-mediated. Humoral immunity is based on the antibodies produced by plasma cells, which are derived from mature naïve B-cells. This allows for an individual exposed to a disease to be administered antibodies already generated by another individual previously exposed to the same disease. Cell-mediated immunity, however, involves T-cells, specifically the cytotoxic (CD8) T-cells. Similar to humoral immunity, T-helper and cytotoxic cells from an individual who was exposed to and survived a disease are administered to an individual who has been freshly exposed to the same disease. The helper T-cells are able to activate other immune cells and cytotoxic T-cells, initiating the removal process of those pathogens [71]. Hence, both humoral and cell-mediated immunity result in the binding of the antibodies to the antigen and affect their elimination.

4. Immune Cells and Lipids

Proinflammatory lipids promote CD4 Th1/Th17 differentiation [72], inhibit Tr1 differentiation [72], induce NF-κB [73], inhibit Treg cell differentiation [73], modulate local activation of T-cells [74], induce NO [75], participate in oxidative stress pathways [76], increase the expression of cytokine genes and CCL2, also known as monocyte chemoattractant protein-1 (MCP-1) [77,78], amplify ER stress [79,80,81], and reduce inflammation-resolving processes [82,83,84,85]. Furthermore, reduced cellular debris clearance [86] and apoptotic clearance defects in macrophages and dendritic cells of non-obese diabetic (NOD) mice, an autoimmune spontaneous T1D-prone model, have been attributed to PGE2 [87], which is significantly higher in infiltrated versus insulitis-free NOD islets [74]. Other lipids, in particular lysophosphatidic acid (LPA), are powerful chemoattractants [40,42,88,89]. LOX products induce migration of leukocytes, p38 mitogen-activated protein kinase pathway, ROS-generating NADPH oxidase (NOX), proinflammatory cytokine production, and β-cell apoptosis [90,91,92,93,94,95]. 12-LOX-deficiency protects mice from streptozotocin (STZ)-induced T1D [96] and reduces insulitis and T1D incidence in the NOD [97]. 12-LOX is detectable in islet cells in recent-onset T1D but absent in non-diabetic subjects [98].
Several publications over the years have introduced the concept that lipid mediators and fatty acids influence immune responses [99,100]. One of the most relevant T-cell discoveries has been the characterization of signaling platforms enriched in cholesterol and glycosphingolipids, which are named lipid rafts. These rafts are important for downstream signaling pathways for enzymes, scaffold proteins, and adaptors [101].

5. Deleterious Impact of iPLA2β Activation

5.1. Cellular Compass

Immune responses are propagated through the recruitment of peripheral blood leukocytes to inflamed areas. The chemokine CCL2 plays a critical role in this process by signaling through receptor CC chemokine receptor 2 (CCR2) [102]. The Cathcart group, examining the role of CCL2 in promoting directed migration of monocytes, identified distinct roles for iPLA2β and cPLA2α [40,103]. Upon CCL2 stimulation, cPLA2α translocated from the cytosol to the endoplasmic reticulum (ER) and regulated monocyte migration via arachidonic acid and its metabolites. In contrast, iPLA2β was recruited to the membrane-enriched pseudopod. Here, iPLA2β was proposed to catalyze hydrolysis of the sn-2 fatty acid from PLD-derived phosphatidic acid (PA) [104]. This phospholipid is among the negatively charged lipid substrates preferred by iPLA2β [7,105], and leads to the accumulation of LPA. Utilizing the iPLA2β-selective inhibitor (S-BEL) [106] and antisense oligodeoxyribucleotide (ODN), the authors demonstrated that iPLA2β-generated LPA regulates actin polymerization to facilitate directionality to the migrating monocytes. They go on to suggest that iPLA2β may manifest a cellular compass role or be an integral component of the cellular compass. Consistently, lipids derived from iPLA2β activation, but not other PLA2s, have been demonstrated to promote monocyte chemotaxis [42,88,107,108] (Figure 1).

5.2. Foam Cell Formation

Macrophages contribute to atherosclerosis development, and it requires their conversion to lipid-laden foam cells via a toll-like receptor (TLR)-mediated process [109]. Lipopolysaccharide (LPS) plays a critical role in this process through the generation of ROS [110,111]. Lee et al. [35] demonstrated that LPS binding to TLR4 induces NOX1 expression and, as a consequence, ROS production. Both S-BEL and small interfering RNA (siRNA) directed against iPLA2β, but not R-BEL or siRNA directed against membrane-associated iPLA2gamma (iPLA2γ), were found to decrease NOX1 expression, and consequently, ROS production. This was associated with mitigation of foam cell formation. They further reported that iPLA2β effects are signaled through the Akt pathway, and although the specific lipid signal involved was not identified, they hypothesized that iDLs promote Akt phosphorylation. However, this remains to be elucidated.

5.3. Vascular Injury

Neointima formation leads to several vascular-related pathologies and macrophages are critical contributors to vascular inflammation [112]. Liu et al. [107] using S-BEL and siRNA directed against iPLA2β, iPLA2β-deficient mice, and mice that selectively overexpress iPLA2β in smooth muscle cells, demonstrated that iPLA2β participates in ligation-induced neointima formation. They further demonstrated that this was associated with increased production of proinflammatory cytokines and vascular infiltration by macrophages. By comparing the effects of inhibiting the arachidonic acid metabolism pathways COX2 with indomethacin, 5-LOX with NDGA, 12-LOX with baicalein, CYP with 17-octadecynoic acid, and 12/15-LOX with luteolin, they deduced that the products of 12-LOX and 15-LOX are complicit in the formation of neointima. In view of reports suggesting that sterol regulatory element-binding protein 1 (SREBP-1) expression is induced in the injured vascular wall [95] and SREBP-1 induces iPLA2β [113,114], they speculated that SREBP-1-mediated induction of iPLA2β occurs in their model of ligation-induced neointima formation. Interestingly, they found that the overexpression of iPLA2β in the smooth muscle cells does not promote neointima formation and that it is only amplified upon injury, leading to increased TNFα production. As basally iPLA2β is inactivated through an interaction with calmodulin [5,115,116], they postulate that TNFα causes disassociation of the complex to unmask iPLA2β activity. Analogously, under basal conditions, β-cells overexpressing iPLA2β do not exhibit a difference in death rate; however, upon stimulation with ER stressors or pro-inflammatory cytokines, they exhibit exacerbated apoptosis [117,118].

5.4. Cigarette Smoke

An interesting link between iPLA2β and platelet activating factor (PAF) was recognized by McHowat’s group in their study of smoke-induced bladder inflammation [119]. This appears to involve a series of events initiated by the tethering of inflammatory cells to the apical endothelial cell surface, rolling of the inflammatory cells across the endothelium and tight adherence, and subsequent PAF-mediated transmigration between neighboring endothelial cells [120,121]. One route of PAF production, predominant during inflammation, is via remodeling initiated by the hydrolysis of the sn-2 substituent from phosphatidylcholine by PLA2 activation to generate alkyl-lysophosphatidyl choline (LPC). Subsequently, an acetyl group is added at the sn-2 position to generate PAF. McHowat’s group observed that upon exposure to cigarette smoke, human and rodent bladder endothelial cells (ECs) exhibited higher PAF accumulation, decreased activity of platelet-activating factor acetylhydrolase (PAF-AH), which degrades PAF, and increased inflammatory cell adherence. These outcomes were reversed by chemical inhibition or genetic knockout of iPLA2β, and by PAF receptor (PAFR) antagonism. Cellular adhesion is facilitated by adhesion molecules such as platelets/endothelial cells (P/E)-selectin and intercellular adhesion molecule 1 (ICAM-1) and these are produced by mast cells [120], which are key contributors to chronic inflammation. Mast cells express PAFR and also release tryptase, which can induce iPLA2β activation in bladder epithelial cells [122]. The authors proposed that mast-cell-induced PAF generation by endothelial cells via iPLA2β activation can prolong the inflammatory state, thus leading to chronic inflammation. Of interest, Ueno et al. [123] reported that cPLA2α, but not iPLA2β, contributes to arachidonic acid production by bone marrow-derived mast cells, facilitating their maturation. Curiously, although mast cells have been reported not to be involved in the development of T1D, which is a consequence of autoimmune destruction of β-cells, in the NOD rodent model [124], they have been suggested to be potentially important in promoting β-cell dysfunction and death in human T1D [125].

5.5. Early-Stage Disease

A common blood cancer is chronic lymphocytic leukemia (CLL), which is characterized by immune-incompetent B-CLL lymphocytes [126,127]. These cells express COX2 and have increased production of PGE2, PGF2, and LTB4. The latter signals through the BLT1 receptor to activate CD40-dependent chronic B lymphocytic leukemia cells to prolong B-CLL survival [128,129]. Greater expression of PLA2 enzymes has been noted in tumor cells, relative to healthy cells, supporting their role in tumor cell proliferation and metastasis [38,130,131]. The work of Guriec et al. suggests that both cPLA2α and iPLA2β are expressed in the tumor cells, with cPLA2α expression increasing higher in patients with advanced-stage disease and with iPLA2β expression being higher in the early disease stages [132]. The disease progression was observed to be associated with dysregulation of arachidonic acid metabolizing pathways (COX, 5-LOX) and of LTA4 hydrolase (LTA4H), which generates LTB4. The progressive increases in PGs and LTB4, the latter of which activates its receptors BLT1 and BLT2, are proposed to support tumor cell proliferation and survival.

5.6. Metabolic Stress

It is well recognized that type 2 diabetes (T2D) is associated with increased coronary artery disease and atherosclerosis [133] and that these involve recruitment of macrophages to inflamed vascular sites [134]. Additionally, cellular oxidative stress appears to be a compounding factor. ROS are generated as byproducts of oxygenases or primary products of NOX. Among the NOX variants [135], although both Nox4 and Nox2 are expressed in macrophages [136], Nox4 promotes monocyte chemotaxis and macrophage recruitment during diabetic metabolic stress (DMS) [137]. Tan et al. [42] reported that under conditions mimicking DMS (i.e., high glucose and LDL), overexpression of iPLA2β amplified macrophage NOX4 expression, ROS production, and CCL2-induced migration. Conversely, these outcomes were mitigated by S-BEL and siRNA directed against iPLA2β. The same interventions had no effect on NOX2. Interestingly, NOX2 is localized primarily in the plasma membrane, whereas NOX4 is localized in the mitochondria, ER and nuclear membranes [138,139], which are also the subcellular organelles that iPLA2β mobilizes to upon the induction of stress [11,12,14,15,140]. Tan et al. further suggested that iPLA2β-derived LPA manifested these effects, as LPA receptor antagonism prevented NOX4 induction and LPA addition was able to rescue outcomes inhibited by S-BEL.

5.7. Macrophage Polarization

In view of the reported impact of iPLA2β on macrophage function, we sought to identify the role of iDLs in this process [39,141]. We found that iPLA2β activation induces macrophage polarization, favoring the M1 inflammatory phenotype. Utilizing selective inhibitors, we identified that both COX and LOX products contribute to iPLA2β-modulated macrophage polarization. Such an outcome was mitigated in macrophages from iPLA2β-deficient mice upon classical activation (LPS+IFNγ). This was reflected by reductions in M1 markers (Arg2 and Nos2) and increases in several M2 markers, including Arg1 and Ccl2. Interestingly, Ptgs2 and Alox12, which encode COX2 and 12-LOX, also decreased, suggesting feedback regulation by products of these enzymes. Among the iDLs identified as inducers of the M1 phenotype were 6-keto PGF1α, PGE2, and LTB4. The inhibition of secretory PLA2s (sPLA2s) GIIA, GV, and GX or cPLA2α did not alter the classically activated macrophage production of eicosanoids, suggesting a select role for iDLs on macrophage polarization. However, other groups have reported a role for sPLA2 or cPLA2α in eicosanoid production from macrophages [142,143,144,145,146,147]. Those studies were performed with macrophage-like cells (RAW264.7 and P388D1) or peritoneal macrophages from rats and mice, with only LPS (10-1000 ng/mL) or zymosan stimulation for 6–48 h, and the involvement of various PLA2s was discerned with chemical inhibitors, some of which effected multiple PLA2s. In contrast, our studies were performed with primary mouse peritoneal macrophages, which have been demonstrated to behave differently from the macrophage-like cells [36], from wild-type and genetically- modified iPLA2β-deficient mice, and importantly, treated with LPS+IFNγ, conditions that induce a M1 macrophage phenotype. Interestingly, while iPLA2β-deficiency skewed macrophage polarization towards an M2 anti-inflammatory phenotype, inhibition of downstream oxygenases did not necessarily promote an M2 phenotype. We posit that a reduction in macrophage-iPLA2β activity, as a consequence of an attenuated inflammatory landscape, promotes a shift to a recovery/repair milieu.

5.8. T1D Development

Immune cells play a significant role in promoting β-cell death, which leads to T1D development [148,149]. In T1D-prone individuals, macrophages are among the first immune cells that migrate to pancreatic islets to initiate inflammatory responses and secrete proinflammatory cytokines and ROS, which lead to β-cell death [150,151]. Two different activation states of macrophages have been described—M1 proinflammatory macrophages [152], which are classically activated (e.g., by IFNγ, LPS, TNFα); and M2 anti-inflammatory macrophages, which are alternatively-activated (e.g., by IL-4 or IL-10) [153]. Although M1 macrophages are recognized to be causative factors in T1D development [154], M2 macrophages protect against T1D development [155].
Macrophages are the only resident myeloid cells found in pancreatic islets under normal conditions in all mice strains [156]. However, the cellular composition of pancreata from diabetic individuals and NOD mice revealed a considerable presence of macrophage infiltration into pancreatic islets, with a predominant M1-like phenotype [33,156,157]. Macrophages found in close contact with β-cells present diabetogenic antigens, predominantly insulin peptides [158,159,160]. The proinflammatory M1 macrophages produce ROS, which, along with proinflammatory cytokines such as IL-1β, IFNγ, and TNFα, cause β-cell destruction [49,161]. It is well-recognized that macrophages work in concert with other immune cells to promote β-cell destruction in T1D [149,162,163]. Furthermore, proinflammatory eicosanoids are linked to macrophage phagocytosis, adhesion, apoptosis, and amplifying macrophage-derived eicosanoid release [24,164,165,166]. In the context of T1D, 12-LOX-null NOD has a lowered T1D incidence, which is associated with reduced macrophage production of proinflammatory cytokines [97].
It is also well-established that in human T1D patients and NOD mice, both CD4 and CD8 T-cells are the major components of the islet infiltrate [162]. Different stages of disease onset show varying compositions of CD4 and CD8 T-cells, depending on the timeframe of progression. First, autoreactive T-cells are activated by β-cell antigens presented by antigen-presenting cells (APCs). The activated T-helper cells are required to activate CD8 T-cells. Second, the activated CD4 T-cells infiltrate the pancreas and are thought to contribute to β-cell destruction via the activation of macrophages [154].
CD4 T-cells recognize MHC class II peptides presented by APCs in order to carry out β-cell destruction. CD4 T-cell-mediated β-cell destruction can be caused by the production of proinflammatory cytokines such as IFNγ, which are toxic to the β-cell, and indirectly, by activating local innate cells such as macrophages and dendritic cells (DCs) to enhance infiltration [154]. CD4 T-cells are also thought to interact with CD8 T-cells by facilitating their activation [154]. CD8 T-cells can directly kill β-cells by interacting with MHC class I molecules and by means of perforin and granzyme secretion [149]. It has been suggested that the MHC class I/CD8 T-cell interaction is required for T1D in the early stages of development [162] and antigen presentation to CD4 T-cells within pancreatic islets is essential for β-cell destruction [154]. Other reports have suggested that the MHC class I/CD8+ T-cell interaction is required for T1D in the early stages of development [167] and that antigen presentation to CD4 T-cells within pancreatic islets are essential for β-cell destruction [154]. Taken together, the literature establishes that immune cells work in concert to promote β-cell destruction in T1D [21].
Evidence linking iDLs with T1D development (Figure 2). Several reports have provided evidence of an association between deleterious outcomes in experimental and clinical diabetes and iPLA2β activation [41,168].
In view of these observations, we explored the possibility that inhibition of iPLA2β can ameliorate T1D [33,169]. We found that administration of FKGK18 to female NOD mice promoted several positive outcomes. There was a significant reduction in insulitis, as reflected by reductions in islet abundances of CD4+ T-cells and B-cells. Glucose homeostasis was also improved, as reflected by β-cell preservation and higher circulating insulin. Consequentially, a significant reduction in T1D was achieved. Inhibition of iPLA2β resulted in decreased production of TNFα from CD4+ T-cells and antibodies from B-cells, suggesting that iDLs modulate immune cell responses. This was supported by the recapitulation of the mitigated TNFα production by select iPLA2β inhibitors, with COX and 12-LOX inhibition. TNFα acts as a powerful chemoattractant [89] and is produced by CD4 T-cells within inflamed islets during T1D development [170]. TNFα overexpression exacerbates insulitis, whereas the opposite occurs in TNFβ-receptor-null mice [171]. To date, ours are the first and only reports of the modulation of T- and B-cell functions by iDLs. Similar findings in a genetically modified NOD model with reduced iPLA2β expression [169] further support this possibility.
These findings prompted us to further examine iDL production by macrophages. Not surprisingly, a dramatically more profound proinflammatory landscape was evident in macrophages from the NOD, relative to the spontaneous diabetes-resistant C57BL/6J, mouse [169]. Lipidomic assessments in the NOD model identified select iDLs (PGE2, PGD2, hydroxyeicosatetraenoic acids 5 and 15 (5-HETE and 15-HETE), LTC4) that correlated with T1D development [169]. Importantly, a similar lipid signature was revealed in the plasma of human subjects at high risk of developing T1D [169]. We therefore posit that select iDLs contribute to T1D onset and that these could be targeted for therapeutics and, in conjunction with autoantibodies, serve as early biomarkers of pre-T1D.

6. Protective Consequences of iPLA2β Activation

6.1. Cancer Development

Inflammation is a key contributor to cancer development [172] and cytokines released by macrophages and T-cells are integral to this process [173]. In view of their earlier observations that iPLA2β-null mice are more susceptible to various inflammatory-based disorders [174,175,176], Inhoffen et al. [177] assessed the ability of immune cells from iPLA2β-null mice to produce cytokines following exposure to CD95/FasL, a trigger of proinflammatory cytokine production [178]. They found that iPLA2β-deficiency increased apoptosis in the liver, spleen, and mesenteric lymph nodes (MLN). Although Kupffer cells (i.e., satellite macrophages in the liver) generated a lower production of proinflammatory cytokines TNFα and IL-6, splenocytes became primed to release proinflammatory Th1-/Th17-related cytokines (IFNγ/IL-17α). These findings led to the suggestion that iPLA2β-deficiency can reduce age-related MLN lymphoma development. Mechanistically, they attributed this to the decreased availability of “find-me” and “eat-me” signals derived via iPLA2β activation. These include LPC, a find-me signal [88], which promotes the clearance of apoptotic debris, the accumulation of which triggers and amplifies subsequent immune responses [179,180] (Figure 3).

6.2. Inflammatory Bowel Disease (IBD)

An inflammatory disorder such as IBD is a consequence of the dysfunction of the intestinal epithelial barrier and the mucosal immune system [181]. Jiao et al. [175] further explored the role of “find-me” signals derived via iPLA2β activation in the context of dextran sodium sulfate-induced IBD. They found that iPLA2β-deficiency promotes the accumulation of infiltrating macrophages and dendritic cells in the colon lamina. These were associated with increased production of inflammatory cytokines (TNFα, IL-1β, and IL-6), macrophage inflammatory proteins (MIP-1α and MIP-1β), and CCL2, leading to intestinal epithelial cell apoptosis and mucus barrier damage. With a concurrent decrease in LPC levels, they speculated that iPLA2β-deficiency mitigated the availability of a “find-me” signal, which limited the clearance of apoptotic debris and the amplification of immune responses. Further, they predicted that the iPLA2β-deficiency also reduced LPA levels, impacting the cellular compass, and preventing the optimal phagocytic function of macrophages. Subsequently, Murase et al. [182], comparing the involvement of various PLA2s, suggested that iPLA2β-deficiency did not worsen the clinical scoring associated with IBD, relative to controls. As they did not reconcile the two studies, it is likely that differences between the source of the iPLA2β−/− mice and DSS concentrations used may be partly responsible. Furthermore, while a 7-day DSS regimen was employed by both groups, Jiao’s group maintained the mice for an additional 3 days without DSS, prior to analyses. It may also be noted that Murase’s study did demonstrate similar lower clinical scores in the wild-type (WT) and iPLA2β-deficient groups between days 1 and 6, relative to mice with deficiencies in cPLA2α or in a variety of sPLA2s; however, at day 7, the scores in the iPLA2β-deficient group appeared to be as high as in the other PLA2-deficient groups. As such, the role of iPLA2β in this model remains to be clarified.

6.3. Chagas Disease

A further link between iPLA2β and PAF was reported by McHowat’s group [183], in the context of Chagas disease. This disease is caused through infection by the protozoan parasite [184] Trypanosoma cruzi (T. cruizi) and can lead to various cardiac abnormalities [185]. The sequela of infection begins with induction of an inflammatory response and upregulation of endothelial adhesion molecules [186], followed by attempts at resolution through the generation of proinflammatory cytokines and the induction of signaling pathways to promote the chemotaxis of immune cells to mitigate the invasion. McHowat’s group reported that infection of iPLA2β-deficient mice resulted in lowered PAF and NO production by cardiac endothelial cells, but that neither the expression of adhesion molecules nor the development of myocardial inflammation was affected. However, significant increases in parasite pseudocysts were noted in the myocardium of iPLA2β-deficient mice. The authors suggested that this was due to an impairment in parasite clearance as a consequence of decreased iPLA2β-mediated LPA production and, as a result, Nox4 expression and NO production. Thus, they surmised that the absence of iPLA2β mitigates parasite clearance due to the reduced recruitment of inflammatory cells to the infected myocardial areas.

6.4. Negative Modulation of Inflammation by AAT and SLP1

In sequential reports, Grau’s group constructed events that participated in regulating IL-1β activation and release. IL-1β is critical to host defense against infections, but the generation of excessive active IL-1β can lead to deleterious inflammatory consequences [187]. The release of IL-1β occurs via two signals. The first is an external stimulus that induces the synthesis of pro-IL-1β. The second signal has been suggested to be ATP, which when released from damaged cells activates the purinergic receptor P2 × 7R, promoting the loss of K+ current and triggering the assembly of NLRP3 (NLR family pyrin domain containing 3)-containing inflammasome [188]. This leads to caspase-1 activation, cleavage of pro-IL-1β, and release of IL-1β from the immune cell. Present in inflammatory cells, alpha-1 antitrypsin (AAT) is a strong inducer of anti-inflammatory processes and its upregulation during systemic inflammation has been associated with the decreased production of proinflammatory cytokines, including IL-1β [189]. In examining the potential role of AAT in ATP-dependent regulation of IL-1β in their first study [190], Siebers et al. found that AAT signaling through the CD36 receptor activates iPLA2β, which leads to the release of a low molecular weight factor (LMWF). The LMWF is released from the cell and binds to nicotinic acetylcholine receptor (nAchR), leading to inhibition of P2X7R function, prevention of inflammasome assembly, and the processing of pro-IL-1β to active IL-1β, and its release from the cell. These outcomes were significantly mitigated with selective inhibition or knockdown of iPLA2β and were also not evident in PBMCs from iPLA2β-deficient mice. In the second study [191], Zakrzewicz et al. demonstrated that the secretory leukocyte protease inhibitor (SLP1) also interferes with ATP-dependent regulation of IL-1β via iPLA2β activation. SLP1 is also present in inflammatory cells [192] and its actions promote anti-inflammatory outcomes. Their results suggested that the SLP1 signals released through annexin 2, a membrane binding protein for SLP1 [193], activate iPLA2β, leading to the production of the LMWF and subsequent inhibition of IL-1β maturation and release. Although both pathways were found to mitigate inflammation, unfortunately neither study explored the identity of the LMWF. Interestingly, AAT administration to recent-onset T1D subjects improved β-cell function, which was correlated with reduced IL-1β production from monocytes and myeloid dendritic cells [194].

7. Summary

iPLA2β is a member of the family of PLA2s that hydrolyzes the sn-2 substituent from membrane glycerophospholipids. Thus, activation of iPLA2β can lead to the production of a variety of bioactive lipid mediators. As eicosanoids generated subsequently to iPLA2β-mediated hydrolysis of sn-2 arachidonic acid can exhibit profound proinflammatory effects and SPMs are produced from sn-2 substituents EPA and DHA, the impact of iPLA2β on the inflammatory sequelae is profound. In recent years, the impact of iPLA2β at the immune cell level is being recognized and those studies, as well as those from our laboratory, suggest that iPLA2β activation modulates the maturation, polarization, activation, and functionality of macrophages, T-cells, and B-cells. The continuation of studies addressing these actions of iPLA2β is important and warranted in order to gain a better understanding of the events that lead to the onset, maintenance, and amplification of inflammation. Identifying the relevant selective iDLs with effects on these processes could lead to the development of new strategies to treat autoimmune- and other inflammatory-based diseases.

Author Contributions

T.D.W., A.A., and S.R. contributed equally to the writing and editing, X.L. contributed to the editing, and Y.G.T. collected the reference materials. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by NIH/NIDDK R01DK110292, DK079626, and NIH/NIAID 1 R21 AI 146743.

Informed Consent Statement

Not Applicable.

Data Availability Statement

Not Applicable.

Acknowledgments

The authors would like to thank the Department of CDIB and the Comprehensive Diabetes Center for their administrative support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Gijon, M.A.; Leslie, C.C. Phospholipases A2. Semin. Cell Dev. Biol. 1997, 8, 297–303. [Google Scholar] [CrossRef] [PubMed]
  2. Turk, J.; White, T.D.; Nelson, A.J.; Lei, X.; Ramanadham, S. iPLA2b and its role in male fertility, neurological disorders, metabolic disorders, and inflammation. Biochim. Biophys. Acta Mol. Cell Biol. Lipids 2019, 1864, 846–860. [Google Scholar] [CrossRef] [PubMed]
  3. Ramanadham, S.; Ali, T.; Ashley, J.W.; Bone, R.N.; Hancock, W.D.; Lei, X. Calcium-independent phospholipases A2 and their roles in biological processes and diseases. J. Lipid Res. 2015, 56, 1643–1668. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Malley, K.R.; Koroleva, O.; Miller, I.; Sanishvili, R.; Jenkins, C.M.; Gross, R.W.; Korolev, S. The structure of iPLA2beta reveals dimeric active sites and suggests mechanisms of regulation and localization. Nat. Commun. 2018, 9, 765. [Google Scholar] [CrossRef] [Green Version]
  5. Wang, Z.; Ramanadham, S.; Ma, Z.A.; Bao, S.; Mancuso, D.J.; Gross, R.W.; Turk, J. Group VIA phospholipase A2 forms a signaling complex with the calcium/calmodulin-dependent protein kinase IIbeta expressed in pancreatic islet beta-cells. J. Biol. Chem. 2005, 280, 6840–6849. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Mouchlis, V.D.; Armando, A.; Dennis, E.A. Substrate-specific inhibition constants for phospholipase A2 acting on unique phospholipid substrates in mixed micelles and membranes using lipidomics. J. Med. Chem. 2019, 62, 1999–2007. [Google Scholar] [CrossRef]
  7. Mouchlis, V.D.; Chen, Y.; Mccammon, J.A.; Dennis, E.A. Membrane allostery and unique hydrophobic sites promote enzyme substrate specificity. J. Am. Chem. Soc. 2018, 140, 3285–3291. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Ali, T.; Kokotos, G.; Magrioti, V.; Bone, R.N.; Mobley, J.A.; Hancock, W.; Ramanadham, S. Characterization of FKGK18 as inhibitor of group VIA Ca2+-independent phospholipase A2 (iPLA2b): Candidate drug for preventing beta-cell apoptosis and diabetes. PLoS ONE 2013, 8, e71748. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Kokotos, G.; Hsu, Y.H.; Burke, J.E.; Baskakis, C.; Kokotos, C.G.; Magrioti, V.; Dennis, E.A. Potent and selective fluoroketone inhibitors of group VIA calcium-independent phospholipase A2. J. Med. Chem. 2010, 53, 3602–3610. [Google Scholar] [CrossRef] [Green Version]
  10. Gross, R.W.; Ramanadham, S.; Kruszka, K.K.; Han, X.; Turk, J. Rat and human pancreatic islet cells contain a calcium ion independent phospholipase A2 activity selective for hydrolysis of arachidonate which is stimulated by adenosine triphosphate and is specifically localized to islet beta-cells. Biochemistry 1993, 32, 327–336. [Google Scholar] [CrossRef]
  11. Bao, S.; Jin, C.; Zhang, S.; Turk, J.; Ma, Z.; Ramanadham, S. Beta-cell calcium-independent group VIA phospholipase A2 (iPLA2b): Tracking iPLA2b movements in response to stimulation with insulin secretagogues in INS-1 cells. Diabetes 2004, 53, S186–S189. [Google Scholar] [CrossRef] [Green Version]
  12. Lei, X.; Zhang, S.; Bohrer, A.; Ramanadham, S. Calcium-independent phospholipase A2 (iPLA2b)-mediated ceramide generation plays a key role in the cross-talk between the endoplasmic reticulum (ER) and mitochondria during ER stress-induced insulin-secreting cell apoptosis. J. Biol. Chem. 2008, 283, 34819–34832. [Google Scholar] [CrossRef] [Green Version]
  13. Ma, Z.; Ramanadham, S.; Wohltmann, M.; Bohrer, A.; Hsu, F.F.; Turk, J. Studies of insulin secretory responses and of arachidonic acid incorporation into phospholipids of stably transfected insulinoma cells that overexpress group VIA phospholipase A2 (iPLA2b) indicate a signaling rather than a housekeeping role for iPLA2b. J. Biol. Chem. 2001, 276, 13198–13208. [Google Scholar] [CrossRef] [Green Version]
  14. Ma, Z.; Zhang, S.; Turk, J.; Ramanadham, S. Stimulation of insulin secretion and associated nuclear accumulation of iPLA2b in INS-1 insulinoma cells. Am. J. Physiol. Endocrinol. Metab. 2002, 282, E820–E833. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Ramanadham, S.; Hsu, F.F.; Zhang, S.; Jin, C.; Bohrer, A.; Song, H.; Bao, S.; Ma, Z.; Turk, J. Apoptosis of insulin-secreting cells induced by endoplasmic reticulum stress is amplified by overexpression of group VIA calcium-independent phospholipase A2 (iPLA2b) and suppressed by inhibition of iPLA2b. Biochemistry 2004, 43, 918–930. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Song, H.; Bao, S.; Lei, X.; Jin, C.; Zhang, S.; Turk, J.; Ramanadham, S. Evidence for proteolytic processing and stimulated organelle redistribution of iPLA2b. Biochim. Biophys. Acta 2010, 1801, 547–558. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Ramanadham, S.; Hsu, F.F.; Bohrer, A.; Zhongmin, M.; Turk, J. Studies of the role of group VI phospholipase A2 in fatty acid incorporation, phospholipid remodeling, lysophosphatidylcholine generation, and secretagogue-induced arachidonic acid release in pancreatic islets and insulinoma cells. J. Biol. Chem. 1999, 275, 13915–13927. [Google Scholar] [CrossRef] [Green Version]
  18. Hanna, V.S.; Hafez, E.A.A. Synopsis of arachidonic acid metabolism: A review. J. Adv. Res. 2018, 11, 23–32. [Google Scholar] [CrossRef]
  19. Abdulkhaleq, L.A.; Assi, M.A.; Abdullah, R.; Zamri-Saad, M.; Taufiq-Yap, Y.H.; Hezmee, M.N.M. The crucial roles of inflammatory mediators in inflammation: A review. Vet. World 2018, 11, 627–635. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  20. Khanapure, S.P.; Garvey, D.S.; Janero, D.R.; Letts, L.G. Eicosanoids in inflammation: Biosynthesis, pharmacology, and therapeutic frontiers. Curr. Top. Med. Chem. 2007, 7, 311–340. [Google Scholar] [CrossRef]
  21. Luo, P.; Wang, M.H. Eicosanoids, beta-cell function, and diabetes. Prostaglandins Other Lipid Mediat. 2011, 95, 1–10. [Google Scholar] [CrossRef] [Green Version]
  22. Tessaro, F.H.; Ayala, T.S.; Martins, J.O. Lipid mediators are critical in resolving inflammation: A review of the emerging roles of eicosanoids in diabetes mellitus. Biomed. Res. Int. 2015, 2015, 568408. [Google Scholar] [CrossRef] [Green Version]
  23. Umamaheswaran, S.; Dasari, S.K.; Yang, P.; Lutgendorf, S.K.; Sood, A.K. Stress, inflammation, and eicosanoids: An emerging perspective. Cancer Metastasis Rev. 2018, 37, 203–211. [Google Scholar] [CrossRef]
  24. Zaitseva, L.; Vaisburd, M.; Shaposhnikova, G.; Mysyakin, E. Role of eicosanoids in regulation of macrophage phagocytic functions by platelet-activating factor during endotoxic shock. Bull. Exp. Biol. Med. 2000, 130, 879–881. [Google Scholar] [CrossRef] [PubMed]
  25. Kuhn, H.; O’donnell, V.B. Inflammation and immune regulation by 12/15-lipoxygenases. Prog. Lipid Res. 2006, 45, 334–356. [Google Scholar] [CrossRef] [PubMed]
  26. Issan, Y.; Hochhauser, E.; Guo, A.; Gotlinger, K.H.; Kornowski, R.; Leshem-Lev, D.; Lev, E.; Porat, E.; Snir, E.; Thompson, C.I.; et al. Elevated level of pro-inflammatory eicosanoids and EPC dysfunction in diabetic patients with cardiac ischemia. Prostaglandins Other Lipid Mediat. 2013, 100–101, 15–21. [Google Scholar] [CrossRef] [Green Version]
  27. Gilroy, D.W.; Newson, J.; Sawmynaden, P.; Willoughby, D.A.; Croxtall, J.D. A novel role for phospholipase A2 isoforms in the checkpoint control of acute inflammation. FASEB J. 2004, 18, 489–498. [Google Scholar] [CrossRef]
  28. De Jong, A.J.; Kloppenburg, M.; Toes, R.E.; Ioan-Facsinay, A. Fatty acids, lipid mediators, and T-cell function. Front. Immunol. 2014, 5, 483. [Google Scholar] [CrossRef] [Green Version]
  29. Serhan, C.N. Treating inflammation and infection in the 21st century: New hints from decoding resolution mediators and mechanisms. FASEB J. 2017, 31, 1273–1288. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Serhan, C.N.; Levy, B.D. Resolvins in inflammation: Emergence of the pro-resolving superfamily of mediators. J. Clin. Investig. 2018, 128, 2657–2669. [Google Scholar] [CrossRef] [PubMed]
  31. Serhan, C.N. Pro-resolving lipid mediators are leads for resolution physiology. Nature 2014, 510, 92–101. [Google Scholar] [CrossRef] [Green Version]
  32. Serhan, C.N.; Chiang, N.; Dalli, J.; Levy, B.D. Lipid mediators in the resolution of inflammation. Cold Spring Harb. Perspect. Biol. 2014, 7, a016311. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Bone, R.N.; Gai, Y.; Magrioti, V.; Kokotou, M.G.; Ali, T.; Lei, X.; Tse, H.M.; Kokotos, G.; Ramanadham, S. Inhibition of Ca2+-independent phospholipase A2beta (iPLA2b) ameliorates islet infiltration and incidence of diabetes in NOD mice. Diabetes 2015, 64, 541–554. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Gil-De-Gomez, L.; Astudillo, A.M.; Guijas, C.; Magrioti, V.; Kokotos, G.; Balboa, M.A.; Balsinde, J. Cytosolic group IVA and calcium-independent group VIA phospholipase A2s act on distinct phospholipid pools in zymosan-stimulated mouse peritoneal macrophages. J. Immunol. 2014, 192, 752–762. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Lee, S.H.; Park, D.W.; Park, S.C.; Park, Y.K.; Hong, S.Y.; Kim, J.R.; Lee, C.H.; Baek, S.H. Calcium-independent phospholipase A2beta-Akt signaling is involved in lipopolysaccharide-induced NADPH oxidase 1 expression and foam cell formation. J. Immunol. 2009, 183, 7497–7504. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Bao, S.; Li, Y.; Lei, X.; Wohltmann, M.; Jin, W.; Bohrer, A.; Semenkovich, C.F.; Ramanadham, S.; Tabas, I.; Turk, J. Attenuated free cholesterol loading-induced apoptosis but preserved phospholipid composition of peritoneal macrophages from mice that do not express group VIA phospholipase A2. J. Biol. Chem. 2007, 282, 27100–27114. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Ma, Z.; Bohrer, A.; Wohltmann, M.; Ramanadham, S.; Hsu, F.F.; Turk, J. Studies of phospholipid metabolism, proliferation, and secretion of stably transfected insulinoma cells that overexpress group VIA phospholipase A2. Lipids 2001, 36, 689–700. [Google Scholar] [CrossRef]
  38. Song, Y.; Wilkins, P.; Hu, W.; Murthy, K.S.; Chen, J.; Lee, Z.; Oyesanya, R.; Wu, J.; Barbour, S.E.; Fang, X. Inhibition of calcium-independent phospholipase A2 suppresses proliferation and tumorigenicity of ovarian carcinoma cells. Biochem. J. 2007, 406, 427–436. [Google Scholar] [CrossRef] [Green Version]
  39. Ashley, J.W.; Hancock, W.D.; Nelson, A.J.; Bone, R.N.; Tse, H.M.; Wohltmann, M.; Turk, J.; Ramanadham, S. Polarization of macrophages toward M2 phenotype is favored by reduction in iPLA2b (Group VIA Phospholipase A2). J. Biol. Chem. 2016, 291, 23268–23281. [Google Scholar] [CrossRef] [Green Version]
  40. Mishra, R.S.; Carnevale, K.A.; Cathcart, M.K. iPLA2beta: Front and center in human monocyte chemotaxis to MCP-1. J. Exp. Med. 2008, 205, 347–359. [Google Scholar] [CrossRef] [Green Version]
  41. Ayilavarapu, S.; Kantarci, A.; Fredman, G.; Turkoglu, O.; Omori, K.; Liu, H.; Iwata, T.; Yagi, M.; Hasturk, H.; Van Dyke, T.E. Diabetes-induced oxidative stress is mediated by Ca2+-independent phospholipase A2 in neutrophils. J. Immunol. 2010, 184, 1507–1515. [Google Scholar] [CrossRef] [Green Version]
  42. Tan, C.; Day, R.; Bao, S.; Turk, J.; Zhao, Q.D. Group VIA phospholipase A2 mediates enhanced macrophage migration in diabetes mellitus by increasing expression of nicotinamide adenine dinucleotide phosphate oxidase 4. Arter. Thromb. Vasc. Biol. 2014, 34, 768–778. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Kalyvas, A.; Baskakis, C.; Magrioti, V.; Constantinou-Kokotou, V.; Stephens, D.; Lopez-Vales, R.; Lu, J.Q.; Yong, V.W.; Dennis, E.A.; Kokotos, G.; et al. Differing roles for members of the phospholipase A2 superfamily in experimental autoimmune encephalomyelitis. Brain J. Neurol. 2009, 132, 1221–1235. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Mchowat, J.; Gullickson, G.; Hoover, R.G.; Sharma, J.; Turk, J.; Kornbluth, J. Platelet-activating factor and metastasis: Calcium-independent phospholipase A2beta deficiency protects against breast cancer metastasis to the lung. Am. J. Physiol. Cell Physiol. 2011, 300, C825–C832. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Nicotera, T.M.; Schuster, D.P.; Bourhim, M.; Chadha, K.; Klaich, G.; Corral, D.A. Regulation of PSA secretion and survival signaling by calcium-independent phopholipase A2beta in prostate cancer cells. Prostate 2009, 69, 1270–1280. [Google Scholar] [CrossRef] [PubMed]
  46. Scuderi, M.R.; Anfuso, C.D.; Lupo, G.; Motta, C.; Romeo, L.; Guerra, L.; Cappellani, A.; Ragusa, N.; Cantarella, G.; Alberghina, M. Expression of Ca2+-independent and Ca2+-dependent phospholipases A2 and cyclooxygenases in human melanocytes and malignant melanoma cell lines. Biochim. Biophys. Acta 2008, 1781, 635–642. [Google Scholar] [CrossRef] [PubMed]
  47. Thayer, T.C.; Delano, M.; Liu, C.; Chen, J.; Padgett, L.E.; Tse, H.M.; Annamali, M.; Piganelli, J.D.; Moldawer, L.L.; Mathews, C.E. Superoxide production by macrophages and T cells is critical for the induction of autoreactivity and type 1 diabetes. Diabetes 2011, 60, 2144–2151. [Google Scholar] [CrossRef] [Green Version]
  48. Jiang, Z.; Jiang, J.X.; Zhang, G.X. Macrophages: A double-edged sword in experimental autoimmune encephalomyelitis. Immunol. Lett. 2014, 160, 17–22. [Google Scholar] [CrossRef]
  49. Burg, A.R.; Tse, H.M. Redox-sensitive innate immune pathways during macrophage activation in type 1 diabetes. Antioxid. Redox Signal. 2018, 29, 1373–1398. [Google Scholar] [CrossRef]
  50. Hirayama, D.; Iida, T.; Nakase, H. The phagocytic function of macrophage-enforcing innate immunity and tissue homeostasis. Int. J. Mol. Sci. 2017, 19, 92. [Google Scholar] [CrossRef] [Green Version]
  51. Schenten, D.; Medzhitov, R. The control of adaptive immune responses by the innate immune system. Adv. Immunol. 2011, 109, 87–124. [Google Scholar] [PubMed]
  52. Colbert, J.D.; Cruz, F.M.; Rock, K.L. Cross-presentation of exogenous antigens on MHC I molecules. Curr. Opin. Immunol. 2020, 64, 1–8. [Google Scholar] [CrossRef] [PubMed]
  53. Epelman, S.; Lavine, K.J.; Randolph, G.J. Origin and functions of tissue macrophages. Immunity 2014, 41, 21–35. [Google Scholar] [CrossRef] [Green Version]
  54. Ginhoux, F.; Guilliams, M. Tissue-resident macrophage ontogeny and homeostasis. Immunity 2016, 44, 439–449. [Google Scholar] [CrossRef] [PubMed]
  55. Theret, M.; Mounier, R.; Rossi, F. The origins and non-canonical functions of macrophages in development and regeneration. Development 2019, 146. [Google Scholar] [CrossRef] [Green Version]
  56. Mantovani, A.; Sica, A.; Locati, M. Macrophage polarization comes of age. Immunity 2005, 23, 344–346. [Google Scholar] [CrossRef] [Green Version]
  57. Davies, L.C.; Jenkins, S.J.; Allen, J.E.; Taylor, P.R. Tissue-resident macrophages. Nat. Immunol. 2013, 14, 986–995. [Google Scholar] [CrossRef]
  58. Mantovani, A.; Biswas, S.K.; Galdiero, M.R.; Sica, A.; Locati, M. Macrophage plasticity and polarization in tissue repair and remodelling. J. Pathol. 2013, 229, 176–185. [Google Scholar] [CrossRef]
  59. Kurowska-Stolarska, M.; Stolarski, B.; Kewin, P.; Murphy, G.; Corrigan, C.J.; Ying, S.; Pitman, N.; Mirchandani, A.; Rana, B.; Van Rooijen, N.; et al. IL-33 amplifies the polarization of alternatively activated macrophages that contribute to airway inflammation. J. Immunol. 2009, 183, 6469–6477. [Google Scholar] [CrossRef] [Green Version]
  60. Shapouri-Moghaddam, A.; Mohammadian, S.; Vazini, H.; Taghadosi, M.; Esmaeili, S.A.; Mardani, F.; Seifi, B.; Mohammadi, A.; Afshari, J.T.; Sahebkar, A. Macrophage plasticity, polarization, and function in health and disease. J. Cell Physiol. 2018, 233, 6425–6440. [Google Scholar] [CrossRef]
  61. Davies, L.C.; Taylor, P.R. Tissue-resident macrophages: Then and now. Immunology 2015, 144, 541–548. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Weigert, A.; Strack, E.; Snodgrass, R.G.; Brüne, B. mPGES-1 and ALOX5/-15 in tumor-associated macrophages. Cancer Metastasis Rev. 2018, 37, 317–334. [Google Scholar] [CrossRef] [PubMed]
  63. Dalli, J.; Serhan, C.N. Pro-resolving mediators in regulating and conferring macrophage function. Front. Immunol. 2017, 8, 1400. [Google Scholar] [CrossRef] [Green Version]
  64. Norris, P.C.; Dennis, E.A. A lipidomic perspective on inflammatory macrophage eicosanoid signaling. Adv. Biol. Regul. 2014, 54, 99–110. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Li, X.; Bi, X.; Wang, S.; Zhang, Z.; Li, F.; Zhao, A.Z. Therapeutic Potential of ω-3 polyunsaturated fatty acids in human autoimmune diseases. Front. Immunol. 2019, 10, 2241. [Google Scholar] [CrossRef] [Green Version]
  66. Wei, J.; Gronert, K. Eicosanoid and specialized proresolving mediator regulation of lymphoid Cells. Trends Biochem. Sci. 2019, 44, 214–225. [Google Scholar] [CrossRef]
  67. Janeway, C.A., Jr.; Travers, P.; Walport, M. Generation of Lymphocytes in Bone Marrow and Thymus; Garland Science: New York, NY, USA, 2001. [Google Scholar]
  68. Li, M.O.; Rudensky, A.Y. T cell receptor signalling in the control of regulatory T cell differentiation and function. Nat. Rev. Immunol. 2016, 16, 220–233. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Zhu, J.; Paul, W.E. CD4 T cells: Fates, functions, and faults. Blood 2008, 5, 1557–1569. [Google Scholar] [CrossRef] [Green Version]
  70. Mak, T.W.; Saunders, M.E.; Jett, B.D. B cell development, activation and effector functions. Primer Immune Response 2014, 2, 111–142. [Google Scholar] [CrossRef]
  71. Bonilla, F.A.; Oettgen, H.C. Adaptive immunity. J. Allergy Clin. Immunol. 2010, 125, S33–S40. [Google Scholar] [CrossRef]
  72. Hooper, K.M.; Kong, W.; Ganea, D. Prostaglandin E2 inhibits Tr1 cell differentiation through suppression of c-Maf. PLoS ONE 2017, 12, e0179184. [Google Scholar] [CrossRef] [Green Version]
  73. Aoki, T.; Frosen, J.; Fukuda, M.; Bando, K.; Shioi, G.; Tsuji, K.; Ollikainen, E.; Nozaki, K.; Laakkonen, J.; Narumiya, S. Prostaglandin E2-EP2-NF-kappaB signaling in macrophages as a potential therapeutic target for intracranial aneurysms. Sci. Signal. 2017, 10. [Google Scholar] [CrossRef] [Green Version]
  74. Ganapathy, V.; Gurlo, T.; Jarstadmarken, H.O.; Von Grafenstein, H. Regulation of TCR-induced IFN-gamma release from islet-reactive non-obese diabetic CD8+ T cells by prostaglandin E2 receptor signaling. Int. Immunol. 2000, 12, 851–860. [Google Scholar] [CrossRef] [Green Version]
  75. Ling, J.J.; Sun, Y.J.; Zhu, D.Y.; Chen, Q.; Han, X. Potential role of NO in modulation of COX-2 expression and PGE2 production in pancreatic beta-cells. Acta Biochim. Biophys. Sin. 2005, 37, 139–146. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Boizel, R.; Bruttmann, G.; Benhamou, P.Y.; Halimi, S.; Stanke-Labesque, F. Regulation of oxidative stress and inflammation by glycaemic control: Evidence for reversible activation of the 5-lipoxygenase pathway in type 1, but not in type 2 diabetes. Diabetologia 2010, 53, 2068–2070. [Google Scholar] [CrossRef] [Green Version]
  77. Chakrabarti, S.K.; Cole, B.K.; Wen, Y.; Keller, S.R.; Nadler, J.L. 12/15-lipoxygenase products induce inflammation and impair insulin signaling in 3T3-L1 adipocytes. Obesity 2009, 17, 1657–1663. [Google Scholar] [CrossRef] [Green Version]
  78. Imai, Y.; Dobrian, A.D.; Morris, M.A.; Taylor-Fishwick, D.A.; Nadler, J.L. Lipids and immunoinflammatory pathways of beta cell destruction. Diabetologia 2016, 59, 673–678. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  79. Bauman, J.A.; Li, S.D.; Yang, A.; Huang, L.; Kole, R. Anti-tumor activity of splice-switching oligonucleotides. Nucleic Acids Res. 2010, 38, 8348–8356. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  80. Inceoglu, B.; Bettaieb, A.; Haj, F.G.; Gomes, A.V.; Hammock, B.D. Modulation of mitochondrial dysfunction and endoplasmic reticulum stress are key mechanisms for the wide-ranging actions of epoxy fatty acids and soluble epoxide hydrolase inhibitors. Prostaglandins Other Lipid Mediat. 2017, 133, 68–78. [Google Scholar] [CrossRef] [PubMed]
  81. Jung, T.W.; Hwang, H.J.; Hong, H.C.; Choi, H.Y.; Yoo, H.J.; Baik, S.H.; Choi, K.M. Resolvin D1 reduces ER stress-induced apoptosis and triglyceride accumulation through JNK pathway in HepG2 cells. Mol. Cell. Endocrinol. 2014, 391, 30–40. [Google Scholar] [CrossRef] [PubMed]
  82. Huang, H.; Weng, J.; Wang, M.H. EETs/sEH in diabetes and obesity-induced cardiovascular diseases. Prostaglandins Other Lipid Mediat. 2016, 125, 80–89. [Google Scholar] [CrossRef]
  83. Jouihan, S.A.; Zuloaga, K.L.; Zhang, W.; Shangraw, R.E.; Krasnow, S.M.; Marks, D.L.; Alkayed, N.J. Role of soluble epoxide hydrolase in exacerbation of stroke by streptozotocin-induced type 1 diabetes mellitus. J. Cereb. Blood Flow Metab. 2013, 33, 1650–1656. [Google Scholar] [CrossRef]
  84. Luther, J.M.; Brown, N.J. Epoxyeicosatrienoic acids and glucose homeostasis in mice and men. Prostaglandins Other Lipid Mediat. 2016, 125, 2–7. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Rodriguez, M.; Clare-Salzler, M. Eicosanoid imbalance in the NOD mouse is related to a dysregulation in soluble epoxide hydrolase and 15-PGDH expression. Ann. N. Y. Acad. Sci. 2006, 1079, 130–134. [Google Scholar] [CrossRef] [PubMed]
  86. Kim, R.; Emi, M.; Tanabe, K. Cancer cell immune escape and tumor progression by exploitation of anti-inflammatory and pro-inflammatory responses. Cancer Biol. Ther. 2005, 4, 924–933. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Atkinson, M.A. Mechanisms underlying the loss of self tolerance in NOD mice. Res. Immunol. 1997, 148, 301–306. [Google Scholar] [CrossRef]
  88. Lauber, K.; Bohn, E.; Krober, S.M.; Xiao, Y.J.; Blumenthal, S.G.; Lindemann, R.K.; Marini, P.; Wiedig, C.; Zobywalski, A.; Baksh, S.; et al. Apoptotic cells induce migration of phagocytes via caspase-3-mediated release of a lipid attraction signal. Cell 2003, 113, 717–730. [Google Scholar] [CrossRef] [Green Version]
  89. Ming, W.J.; Bersani, L.; Mantovani, A. Tumor necrosis factor is chemotactic for monocytes and polymorphonuclear leukocytes. J. Immunol. 1987, 138, 1469–1474. [Google Scholar]
  90. Ma, K.; Nunemaker, C.S.; Wu, R.; Chakrabarti, S.K.; Taylor-Fishwick, D.A.; Nadler, J.L. 12-Lipoxygenase products reduce insulin secretion and {beta}-cell viability in human islets. J. Clin. Endocrinol. Metab. 2010, 95, 887–893. [Google Scholar] [CrossRef] [Green Version]
  91. Mcduffie, M.; Maybee, N.A.; Keller, S.R.; Stevens, B.K.; Garmey, J.C.; Morris, M.A.; Kropf, E.; Rival, C.; Ma, K.; Carter, J.D.; et al. Nonobese diabetic (NOD) mice congenic for a targeted deletion of 12/15-lipoxygenase are protected from autoimmune diabetes. Diabetes 2008, 57, 199–208. [Google Scholar] [CrossRef] [Green Version]
  92. Prasad, K.M.; Thimmalapura, P.R.; Woode, E.A.; Nadler, J.L. Evidence that increased 12-lipoxygenase expression impairs pancreatic beta cell function and viability. Biochem. Biophys. Res. Commun. 2003, 308, 427–432. [Google Scholar] [CrossRef]
  93. Weaver, J.R.; Holman, T.R.; Imai, Y.; Jadhav, A.; Kenyon, V.; Maloney, D.J.; Nadler, J.L.; Rai, G.; Simeonov, A.; Taylor-Fishwick, D.A. Integration of pro-inflammatory cytokines, 12-lipoxygenase and NOX-1 in pancreatic islet beta cell dysfunction. Mol. Cell. Endocrinol. 2012, 358, 88–95. [Google Scholar] [CrossRef] [PubMed]
  94. Wen, Y.; Gu, J.; Vandenhoff, G.E.; Liu, X.; Nadler, J.L. Role of 12/15-lipoxygenase in the expression of MCP-1 in mouse macrophages. Am. J. Physiol. Heart Circ. Physiol. 2008, 294, H1933–H1938. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Zhou, R.H.; Pesant, S.; Cohn, H.I.; Eckhart, A.D. Enhanced sterol response element-binding protein in postintervention restenotic blood vessels plays an important role in vascular smooth muscle proliferation. Life Sci. 2008, 82, 174–181. [Google Scholar] [CrossRef] [PubMed]
  96. Bleich, D.; Chen, S.; Zipser, B.; Sun, D.; Funk, C.D.; Nadler, J.L. Resistance to type 1 diabetes induction in 12-lipoxygenase knockout mice. J. Clin. Investig. 1999, 103, 1431–1436. [Google Scholar] [CrossRef] [Green Version]
  97. Green-Mitchell, S.M.; Tersey, S.A.; Cole, B.K.; Ma, K.; Kuhn, N.S.; Cunningham, T.D.; Maybee, N.A.; Chakrabarti, S.K.; Mcduffie, M.; Taylor-Fishwick, D.A.; et al. Deletion of 12/15-lipoxygenase alters macrophage and islet function in NOD-Alox15(null) mice, leading to protection against type 1 diabetes development. PLoS ONE 2013, 8, e56763. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. Grzesik, W.J.; Nadler, J.L.; Machida, Y.; Nadler, J.L.; Imai, Y.; Morris, M.A. Expression pattern of 12-lipoxygenase in human islets with type 1 diabetes and type 2 diabetes. J. Clin. Endocrinol. Metab. 2015, 100, E387–E395. [Google Scholar] [CrossRef]
  99. Bennett, M.; Gilroy, D.W. Lipid mediators in inflammation. Microbiol. Spectr. 2016, 4. [Google Scholar] [CrossRef]
  100. Serhan, C.N.; Chiang, N.; Van Dyke, T.E. Resolving inflammation: Dual anti-inflammatory and pro-resolution lipid mediators. Nat. Rev. Immunol. 2008, 8, 349–361. [Google Scholar] [CrossRef] [Green Version]
  101. Tuosto, L.; Xu, C. Editorial: Membrane lipids in T cell functions. Front. Immunol. 2018, 9, 1608. [Google Scholar] [CrossRef] [Green Version]
  102. Gu, L.; Tseng, S.C.; Rollins, B.J. Monocyte chemoattractant protein-1. Chem. Immunol. 1999, 72, 7–29. [Google Scholar] [PubMed]
  103. Cathcart, M.K. Signal-activated phospholipase regulation of leukocyte chemotaxis. J. Lipid Res. 2009, 50, S231–S236. [Google Scholar] [CrossRef] [Green Version]
  104. Kam, Y.; Exton, J.H. Phospholipase D activity is required for actin stress fiber formation in fibroblasts. Mol. Cell. Biol. 2001, 21, 4055–4066. [Google Scholar] [CrossRef] [Green Version]
  105. Tang, J.; Kriz, R.W.; Wolfman, N.; Shaffer, M.; Seehra, J.; Jones, S.S. A novel cytosolic calcium-independent phospholipase A2 contains eight ankyrin motifs. J. Biol. Chem. 1997, 272, 8567–8575. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Jenkins, C.M.; Han, X.; Mancuso, D.J.; Gross, R.W. Identification of calcium-independent phospholipase A2 (iPLA2) beta, and not iPLA2gamma, as the mediator of arginine vasopressin-induced arachidonic acid release in A-10 smooth muscle cells. Enantioselective mechanism-based discrimination of mammalian iPLA2s. J. Biol. Chem. 2002, 277, 32807–32814. [Google Scholar] [PubMed] [Green Version]
  107. Liu, S.; Xie, Z.; Zhao, Q.; Pang, H.; Turk, J.; Calderon, L.; Su, W.; Zhao, G.; Xu, H.; Gong, M.C.; et al. Smooth muscle-specific expression of calcium-independent phospholipase A2beta (iPLA2b) participates in the initiation and early progression of vascular inflammation and neointima formation. J. Biol. Chem. 2012, 287, 24739–24753. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Zhao, X.; Wang, D.; Zhao, Z.; Xiao, Y.; Sengupta, S.; Xiao, Y.; Zhang, R.; Lauber, K.; Wesselborg, S.; Feng, L.; et al. Caspase-3-dependent activation of calcium-independent phospholipase A2 enhances cell migration in non-apoptotic ovarian cancer cells. J. Biol. Chem. 2006, 281, 29357–29368. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  109. Lee, J.G.; Lim, E.J.; Park, D.W.; Lee, S.H.; Kim, J.R.; Baek, S.H. A combination of Lox-1 and Nox1 regulates TLR9-mediated foam cell formation. Cell Signal. 2008, 20, 2266–2275. [Google Scholar] [CrossRef] [PubMed]
  110. Funk, J.L.; Feingold, K.R.; Moser, A.H.; Grunfeld, C. Lipopolysaccharide stimulation of RAW 264.7 macrophages induces lipid accumulation and foam cell formation. Atherosclerosis 1993, 98, 67–82. [Google Scholar] [CrossRef]
  111. Taniyama, Y.; Griendling, K.K. Reactive oxygen species in the vasculature: Molecular and cellular mechanisms. Hypertension 2003, 42, 1075–1081. [Google Scholar] [CrossRef] [Green Version]
  112. Shirai, T.; Hilhorst, M.; Harrison, D.G.; Goronzy, J.J.; Weyand, C.M. Macrophages in vascular inflammation. From atherosclerosis to vasculitis. Autoimmunity 2015, 48, 139–151. [Google Scholar] [CrossRef] [Green Version]
  113. Lei, X.; Zhang, S.; Barbour, S.E.; Bohrer, A.; Ford, E.L.; Koizumi, A.; Papa, F.R.; Ramanadham, S. Spontaneous development of endoplasmic reticulum stress that can lead to diabetes mellitus is associated with higher calcium-independent phospholipase A2 expression: A role for regulation by SREBP-1. J. Biol. Chem. 2010, 285, 6693–6705. [Google Scholar] [CrossRef] [Green Version]
  114. Seashols, S.J.; Del Castillo Olivares, A.; Gil, G.; Barbour, S.E. Regulation of group VIA phospholipase A2 expression by sterol availability. Biochim. Biophys. Acta 2004, 1684, 29–37. [Google Scholar] [CrossRef]
  115. Smani, T.; Zakharov, S.I.; Csutora, P.; Leno, E.; Trepakova, E.S.; Bolotina, V.M. A novel mechanism for the store-operated calcium influx pathway. Nat. Cell. Biol. 2004, 6, 113–120. [Google Scholar] [CrossRef] [PubMed]
  116. Wolf, M.J.; Wang, J.; Turk, J.; Gross, R.W. Depletion of intracellular calcium stores activates smooth muscle cell calcium-independent phospholipase A2. A novel mechanism underlying arachidonic acid mobilization. J. Biol. Chem. 1997, 272, 1522–1526. [Google Scholar] [CrossRef] [Green Version]
  117. Lei, X.; Bone, R.N.; Ali, T.; Wohltmann, M.; Gai, Y.; Goodwin, K.J.; Bohrer, A.E.; Turk, J.; Ramanadham, S. Genetic modulation of islet beta-cell iPLA2b expression provides evidence for its impact on beta-cell apoptosis and autophagy. Islets 2013, 5, 29–44. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Lei, X.; Bone, R.N.; Ali, T.; Zhang, S.; Bohrer, A.; Tse, H.M.; Bidasee, K.R.; Ramanadham, S. Evidence of contribution of iPLA2b-mediated events during islet beta-cell apoptosis due to proinflammatory cytokines suggests a role for iPLA2b in T1D development. Endocrinology 2014, 155, 3352–3364. [Google Scholar] [CrossRef] [PubMed]
  119. Marentette, J.; Kolar, G.; Mchowat, J. Increased susceptibility to bladder inflammation in smokers: Targeting the PAF-PAF receptor interaction to manage inflammatory cell recruitment. Physiol. Rep. 2015, 3. [Google Scholar] [CrossRef] [PubMed]
  120. Green, M.; Filippou, A.; Sant, G.; Theoharides, T.C. Expression of intercellular adhesion molecules in the bladder of patients with interstitial cystitis. Urology 2004, 63, 688–693. [Google Scholar] [CrossRef]
  121. Rastogi, P.; White, M.C.; Rickard, A.; Mchowat, J. Potential mechanism for recruitment and migration of CD133 positive cells to areas of vascular inflammation. Thromb. Res. 2008, 123, 258–266. [Google Scholar] [CrossRef] [Green Version]
  122. Rickard, A.; Portell, C.; Kell, P.J.; Vinson, S.M.; Mchowat, J. Protease-activated receptor stimulation activates a Ca2+-independent phospholipase A2 in bladder microvascular endothelial cells. Am. J. Physiol. Renal Physiol. 2005, 288, F714–F721. [Google Scholar] [CrossRef]
  123. Ueno, N.; Taketomi, Y.; Yamamoto, K.; Hirabayashi, T.; Kamei, D.; Kita, Y.; Shimizu, T.; Shinzawa, K.; Tsujimoto, Y.; Ikeda, K.; et al. Analysis of two major intracellular phospholipases A2 (PLA2) in mast cells reveals crucial contribution of cytosolic PLA2alpha, not Ca2+-independent PLA2beta, to lipid mobilization in proximal mast cells and distal fibroblasts. J. Biol. Chem. 2011, 286, 37249–37263. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Gutierrez, D.A.; Fu, W.; Schonefeldt, S.; Feyerabend, T.B.; Ortiz-Lopez, A.; Lampi, Y.; Liston, A.; Mathis, D.; Rodewald, H.R. Type 1 diabetes in NOD mice unaffected by mast cell deficiency. Diabetes 2014, 63, 3827–3834. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Martino, L.; Masini, M.; Bugliani, M.; Marselli, L.; Suleiman, M.; Boggi, U.; Nogueira, T.C.; Filipponi, F.; Occhipinti, M.; Campani, D.; et al. Mast cells infiltrate pancreatic islets in human type 1 diabetes. Diabetologia 2015, 58, 2554–2562. [Google Scholar] [CrossRef] [Green Version]
  126. Ryan, E.P.; Pollock, S.J.; Kaur, K.; Felgar, R.E.; Bernstein, S.H.; Chiorazzi, N.; Phipps, R.P. Constitutive and activation-inducible cyclooxygenase-2 expression enhances survival of chronic lymphocytic leukemia B cells. Clin. Immunol. 2006, 120, 76–90. [Google Scholar] [CrossRef] [PubMed]
  127. Secchiero, P.; Barbarotto, E.; Gonelli, A.; Tiribelli, M.; Zerbinati, C.; Celeghini, C.; Agostinelli, C.; Pileri, S.A.; Zauli, G. Potential pathogenetic implications of cyclooxygenase-2 overexpression in B chronic lymphoid leukemia cells. Am. J. Pathol. 2005, 167, 1599–1607. [Google Scholar] [CrossRef] [Green Version]
  128. Ray, D.M.; Akbiyik, F.; Bernstein, S.H.; Phipps, R.P. CD40 engagement prevents peroxisome proliferator-activated receptor gamma agonist-induced apoptosis of B lymphocytes and B lymphoma cells by an NF-kappaB-dependent mechanism. J. Immunol. 2005, 174, 4060–4069. [Google Scholar] [CrossRef] [Green Version]
  129. Ricciotti, E.; Fitzgerald, G.A. Prostaglandins and inflammation. Arterioscler. Thromb. Vasc. Biol. 2011, 31, 986–1000. [Google Scholar] [CrossRef]
  130. Roshak, A.K.; Capper, E.A.; Stevenson, C.; Eichman, C.; Marshall, L.A. Human calcium-independent phospholipase A2 mediates lymphocyte proliferation. J. Biol. Chem. 2000, 275, 35692–35698. [Google Scholar] [CrossRef] [Green Version]
  131. Patel, M.I.; Singh, J.; Niknami, M.; Kurek, C.; Yao, M.; Lu, S.; Maclean, F.; King, N.J.; Gelb, M.H.; Scott, K.F.; et al. Cytosolic phospholipase A2-alpha: A potential therapeutic target for prostate cancer. Clin. Cancer Res. 2008, 14, 8070–8079. [Google Scholar] [CrossRef] [Green Version]
  132. Guriec, N.; Le Jossic-Corcos, C.; Simon, B.; Ianotto, J.C.; Tempescul, A.; Dreano, Y.; Salaun, J.P.; Berthou, C.; Corcos, L. The arachidonic acid-LTB4-BLT2 pathway enhances human B-CLL aggressiveness. Biochim. Biophys. Acta 2014, 1842, 2096–2105. [Google Scholar] [CrossRef] [PubMed]
  133. Haffner, S.M.; Lehto, S.; Ronnemaa, T.; Pyorala, K.; Laakso, M. Mortality from coronary heart disease in subjects with type 2 diabetes and in nondiabetic subjects with and without prior myocardial infarction. N. Engl. J. Med. 1998, 339, 229–234. [Google Scholar] [CrossRef] [PubMed]
  134. Brownlee, M. Biochemistry and molecular cell biology of diabetic complications. Nature 2001, 414, 813–820. [Google Scholar] [CrossRef] [PubMed]
  135. Panday, A.; Sahoo, M.K.; Osorio, D.; Batra, S. NADPH oxidases: An overview from structure to innate immunity-associated pathologies. Cell Mol. Immunol. 2015, 12, 5–23. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Lee, C.F.; Qiao, M.; Schroder, K.; Zhao, Q.; Asmis, R. Nox4 is a novel inducible source of reactive oxygen species in monocytes and macrophages and mediates oxidized low density lipoprotein-induced macrophage death. Circ. Res. 2010, 106, 1489–1497. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Ullevig, S.; Zhao, Q.; Lee, C.F.; Seok Kim, H.; Zamora, D.; Asmis, R. NADPH oxidase 4 mediates monocyte priming and accelerated chemotaxis induced by metabolic stress. Arterioscler. Thromb. Vasc. Biol. 2012, 32, 415–426. [Google Scholar] [CrossRef] [Green Version]
  138. Ago, T.; Kuroda, J.; Pain, J.; Fu, C.; Li, H.; Sadoshima, J. Upregulation of Nox4 by hypertrophic stimuli promotes apoptosis and mitochondrial dysfunction in cardiac myocytes. Circ. Res. 2010, 106, 1253–1264. [Google Scholar] [CrossRef]
  139. Block, K.; Gorin, Y.; Abboud, H.E. Subcellular localization of Nox4 and regulation in diabetes. Proc. Natl. Acad. Sci. USA 2009, 106, 14385–14390. [Google Scholar] [CrossRef] [Green Version]
  140. Lei, X.; Zhang, S.; Bohrer, A.; Bao, S.; Song, H.; Ramanadham, S. The group VIA calcium-independent phospholipase A2 participates in ER stress-induced INS-1 insulinoma cell apoptosis by promoting ceramide generation via hydrolysis of sphingomyelins by neutral sphingomyelinase. Biochemistry 2007, 46, 10170–10185. [Google Scholar] [CrossRef] [Green Version]
  141. Nelson, A.J.; Stephenson, D.J.; Cardona, C.L.; Lei, X.; Almutairi, A.; White, T.D.; Tusing, Y.G.; Park, M.A.; Barbour, S.E.; Chalfant, C.E.; et al. Macrophage polarization is linked to Ca2+-independent phospholipase A2beta-derived lipids and cross-cell signaling in mice. J. Lipid Res. 2020, 61, 143–158. [Google Scholar] [CrossRef]
  142. Gil-De-Gomez, L.; Monge, P.; Rodriguez, J.P.; Astudillo, A.M.; Balboa, M.A.; Balsinde, J. Phospholipid arachidonic acid remodeling during phagocytosis in mouse peritoneal macrophages. Biomedicines 2020, 8, 274. [Google Scholar] [CrossRef] [PubMed]
  143. Goldsmith, M.; Daka, A.; Lamour, N.F.; Mashiach, R.; Glucksam, Y.; Meijler, M.M.; Chalfant, C.E.; Zor, T. A ceramide analog inhibits cPLA2 activity and consequent PGE2 formation in LPS-stimulated macrophages. Immunol. Lett. 2011, 135, 136–143. [Google Scholar] [CrossRef] [Green Version]
  144. Naraba, H.; Murakami, M.; Matsumoto, H.; Shimbara, S.; Ueno, A.; Kudo, I.; Oh-Ishi, S. Segregated coupling of phospholipases A2, cyclooxygenases, and terminal prostanoid synthases in different phases of prostanoid biosynthesis in rat peritoneal macrophages. J. Immunol. 1998, 160, 2974–2982. [Google Scholar]
  145. Qi, H.Y.; Shelhamer, J.H. Toll-like receptor 4 signaling regulates cytosolic phospholipase A2 activation and lipid generation in lipopolysaccharide-stimulated macrophages. J. Biol. Chem. 2005, 280, 38969–38975. [Google Scholar] [CrossRef] [Green Version]
  146. Ruiperez, V.; Casas, J.; Balboa, M.A.; Balsinde, J. Group V phospholipase A2-derived lysophosphatidylcholine mediates cyclooxygenase-2 induction in lipopolysaccharide-stimulated macrophages. J. Immunol. 2007, 1790, 631–638. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Zhu, C.; Zhou, J.; Li, T.; Mu, J.; Jin, L.; Li, S. Urocortin participates in LPS-induced apoptosis of THP-1 macrophages via S1P-cPLA2 signaling pathway. Eur. J. Pharmacol. 2020, 887, 173559. [Google Scholar] [CrossRef] [PubMed]
  148. Szablewski, L. Role of immune system in type 1 diabetes mellitus pathogenesis. Int. Immunopharmacol. 2014, 22, 182–191. [Google Scholar] [CrossRef]
  149. Mathis, D.; Vence, L.; Benoist, C. Beta-cell death during progression to diabetes. Nature 2001, 414, 792–798. [Google Scholar] [CrossRef]
  150. Pugliese, A. Autoreactive T cells in type 1 diabetes. J. Clin. Investig. 2017, 127, 2881–2891. [Google Scholar] [CrossRef]
  151. Padgett, L.E.; Broniowska, K.A.; Hansen, P.A.; Corbett, J.A.; Tse, H.M. The role of reactive oxygen species and proinflammatory cytokines in type 1 diabetes pathogenesis. Ann. N. Y. Acad. Sci. 2013, 1281, 16–35. [Google Scholar] [CrossRef] [Green Version]
  152. Mantovani, A.; Sica, A.; Sozzani, S.; Allavena, P.; Vecchi, A.; Locati, M. The chemokine system in diverse forms of macrophage activation and polarization. Trends Immunol. 2004, 25, 677–686. [Google Scholar] [CrossRef] [PubMed]
  153. Martinez, F.O.; Gordon, S. The M1 and M2 paradigm of macrophage activation: Time for reassessment. F1000Prime Rep. 2014, 6, 13. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Calderon, B.; Suri, A.; Unanue, E.R. In CD4+ T-cell-induced diabetes, macrophages are the final effector cells that mediate islet beta-cell killing: Studies from an acute model. Am. J. Pathol. 2006, 169, 2137–2147. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Parsa, R.; Andresen, P.; Gillett, A.; Mia, S.; Zhang, X.M.; Mayans, S.; Holmberg, D.; Harris, R.A. Adoptive transfer of immunomodulatory M2 macrophages prevents type 1 diabetes in NOD mice. Diabetes 2012, 61, 2881–2892. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  156. Zakharov, P.N.; Hu, H.; Wan, X.; Unanue, E.R. Single-cell RNA sequencing of murine islets shows high cellular complexity at all stages of autoimmune diabetes. J. Exp. Med. 2020, 217. [Google Scholar] [CrossRef] [Green Version]
  157. Chen, C.; Cohrs, C.M.; Stertmann, J.; Bozsak, R.; Speier, S. Human beta cell mass and function in diabetes: Recent advances in knowledge and technologies to understand disease pathogenesis. Mol. Metab. 2017, 6, 943–957. [Google Scholar] [CrossRef]
  158. Unanue, E.R. Macrophages in endocrine glands, with emphasis on pancreatic islets. Microbiol. Spectr. 2016, 4. [Google Scholar] [CrossRef] [Green Version]
  159. Calderon, B.; Suri, A.; Miller, M.J.; Unanue, E.R. Dendritic cells in islets of Langerhans constitutively present beta cell-derived peptides bound to their class II MHC molecules. Proc. Natl. Acad. Sci. USA 2008, 105, 6121–6126. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  160. Mohan, J.F.; Levisetti, M.G.; Calderon, B.; Herzog, J.W.; Petzold, S.J.; Unanue, E.R. Unique autoreactive T cells recognize insulin peptides generated within the islets of Langerhans in autoimmune diabetes. Nat. Immunol. 2010, 11, 350–354. [Google Scholar] [CrossRef] [Green Version]
  161. Feduska, J.M.; Tse, H.M. The proinflammatory effects of macrophage-derived NADPH oxidase function in autoimmune diabetes. Free Radic. Biol. Med. 2018, 125, 81–89. [Google Scholar] [CrossRef]
  162. Burrack, A.L.; Martinov, T.; Fife, B.T. T Cell-mediated beta cell destruction: Autoimmunity and alloimmunity in the context of type 1 diabetes. Front. Endocrinol. 2017, 8, 343. [Google Scholar] [CrossRef]
  163. Dimeglio, L.A.; Evans-Molina, C.; Oram, R.A. Type 1 diabetes. Lancet 2018, 391, 2449–2462. [Google Scholar] [CrossRef]
  164. Sun, Y.; Wang, Y.; Li, J.-H.; Zhu, S.-H.; Tang, H.-T.; Xia, Z.-F. Macrophage migration inhibitory factor counter-regulates dexamethasone-induced annexin 1 expression and influences the release of eicosanoids in murine macrophages. Immunology 2013, 140, 250–258. [Google Scholar] [CrossRef]
  165. Duan, L.; Gan, H.; Arm, J.; Remold, H.G. Cytosolic phospholipase A2 participates with TNF-alpha in the induction of apoptosis of human macrophages infected with Mycobacterium tuberculosis H37Ra. J. Immunol. 2001, 166, 7469–7476. [Google Scholar] [CrossRef] [Green Version]
  166. Nikolic, D.M.; Gong, M.C.; Turk, J.; Post, S.R. Class A scavenger receptor-mediated macrophage adhesion requires coupling of calcium-independent phospholipase A2 and 12/15-lipoxygenase to Rac and Cdc42 activation. J. Biol. Chem. 2007, 282, 33405–33411. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  167. Dilorenzo, T.; Graser, R.; Ono, T.; Christianson, G.; Chapman, H.; Roopenian, D.; Nathenson, S.; Serreze, D. Major histocompatibility complex class I-restricted T cells are required for all but the end stages of diabetes development in nonobese diabetic mice and use a prevalent T cell receptor alpha chain gene rearrangement. Proc. Natl. Acad. Sci. USA 1998, 95, 12438–12543. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  168. Rahnema, P.; Shimoni, Y.; Nygren, A. Reduced conduction reserve in the diabetic rat heart: Role of iPLA2 activation in the response to ischemia. Am. J. Physiol. Heart Circ. Physiol. 2011, 300, H326–H334. [Google Scholar] [CrossRef] [PubMed]
  169. Nelson, A.J.; Stephenson, D.J.; Bone, R.N.; Cardona, C.L.; Park, M.A.; Tusing, Y.G.; Lei, X.; Kokotos, G.; Graves, C.L.; Mathews, C.E.; et al. Lipid mediators and biomarkers associated with type 1 diabetes development. JCI Insight 2020, 5. [Google Scholar] [CrossRef] [PubMed]
  170. Held, W.; Macdonald, H.R.; Weissman, I.L.; Hess, M.W.; Mueller, C. Genes encoding tumor necrosis factor alpha and granzyme A are expressed during development of autoimmune diabetes. Proc. Natl. Acad. Sci. USA 1990, 87, 2239–2243. [Google Scholar] [CrossRef] [Green Version]
  171. Herrera, P.L.; Harlan, D.M.; Vassalli, P. A mouse CD8 T cell-mediated acute autoimmune diabetes independent of the perforin and Fas cytotoxic pathways: Possible role of membrane TNF. Proc. Natl. Acad. Sci. USA 2000, 97, 279–284. [Google Scholar] [CrossRef] [Green Version]
  172. Aggarwal, B.B.; Vijayalekshmi, R.V.; Sung, B. Targeting inflammatory pathways for prevention and therapy of cancer: Short-term friend, long-term foe. Clin. Cancer Res. 2009, 15, 425–430. [Google Scholar] [CrossRef] [Green Version]
  173. Lin, W.W.; Karin, M. A cytokine-mediated link between innate immunity, inflammation, and cancer. J. Clin. Investig. 2007, 117, 1175–1183. [Google Scholar] [CrossRef]
  174. Jiao, L.; Gan-Schreier, H.; Tuma-Kellner, S.; Stremmel, W.; Chamulitrat, W. Sensitization to autoimmune hepatitis in group VIA calcium-independent phospholipase A2-null mice led to duodenal villous atrophy with apoptosis, goblet cell hyperplasia and leaked bile acids. Biochim. Biophys. Acta 2015, 1852, 1646–1657. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  175. Jiao, L.; Inhoffen, J.; Gan-Schreier, H.; Tuma-Kellner, S.; Stremmel, W.; Sun, Z.; Chamulitrat, W. Deficiency of group VIA phospholipase A2 (iPLA2b) renders susceptibility for chemical-induced colitis. Dig. Dis. Sci. 2015, 60, 3590–3602. [Google Scholar] [CrossRef] [PubMed]
  176. Xu, W.; Tuma, S.; Katava, N.; Pathil-Warth, A.; Stremmel, W.; Chamultrat, W. Deficiencies of calcium-independent phospholipase A2 B in vivo causes reducedsystemic lipids and lipoproteins concomitant with increased hepatic apoptosis and inflammation. In Journal of Hepathology, Proceeding of The International Liver CongressTM (EASL), Vienna, Austria, 10–14 April 2012; EASL: Barcelona, Spain, 2012. [Google Scholar]
  177. Inhoffen, J.; Tuma-Kellner, S.; Straub, B.; Stremmel, W.; Chamulitrat, W. Deficiency of iPLA2beta primes immune cells for proinflammation: Potential involvement in age-related mesenteric lymph node lymphoma. Cancers 2015, 7, 2427–2442. [Google Scholar] [CrossRef] [PubMed]
  178. Park, D.R.; Thomsen, A.R.; Frevert, C.W.; Pham, U.; Skerrett, S.J.; Kiener, P.A.; Liles, W.C. Fas (CD95) induces proinflammatory cytokine responses by human monocytes and monocyte-derived macrophages. J. Immunol. 2003, 170, 6209–6216. [Google Scholar] [CrossRef] [PubMed]
  179. Albert, M.L. Death-defying immunity: Do apoptotic cells influence antigen processing and presentation? Nat. Rev. Immunol. 2004, 4, 223–231. [Google Scholar] [CrossRef] [PubMed]
  180. Scaffidi, P.; Misteli, T.; Bianchi, M.E. Release of chromatin protein HMGB1 by necrotic cells triggers inflammation. Nature 2002, 418, 191–195. [Google Scholar] [CrossRef]
  181. Xavier, R.J.; Podolsky, D.K. Unravelling the pathogenesis of inflammatory bowel disease. Nature 2007, 448, 427–434. [Google Scholar] [CrossRef]
  182. Murase, R.; Sato, H.; Yamamoto, K.; Ushida, A.; Nishito, Y.; Ikeda, K.; Kobayashi, T.; Yamamoto, T.; Taketomi, Y.; Murakami, M. Group X secreted phospholipase A2 releases omega3 polyunsaturated fatty acids, suppresses colitis, and promotes sperm fertility. J. Biol. Chem. 2016, 291, 6895–6911. [Google Scholar] [CrossRef] [Green Version]
  183. Sharma, J.; Blase, J.R.; Hoft, D.F.; Marentette, J.O.; Turk, J.; Mchowat, J. Mice with genetic deletion of group via phospholipase a2beta exhibit impaired macrophage function and increased parasite load in Trypanosoma cruzi-induced myocarditis. Infect. Immun. 2016, 84, 1137–1142. [Google Scholar] [CrossRef] [Green Version]
  184. Bern, C.; Verastegui, M.; Gilman, R.H.; Lafuente, C.; Galdos-Cardenas, G.; Calderon, M.; Pacori, J.; Del Carmen Abastoflor, M.; Aparicio, H.; Brady, M.F.; et al. Congenital Trypanosoma cruzi transmission in Santa Cruz, Bolivia. Clin. Infect. Dis. 2009, 49, 1667–1674. [Google Scholar] [CrossRef] [Green Version]
  185. Rassi, A., Jr.; Rassi, A.; Little, W.C. Chagas’ heart disease. Clin. Cardiol. 2000, 23, 883–889. [Google Scholar]
  186. Lannes-Vieira, J.; Silverio, J.C.; Pereira, I.R.; Vinagre, N.F.; Carvalho, C.M.; Paiva, C.N.; Silva Da, A.A. Chronic Trypanosoma cruzi-elicited cardiomyopathy: From the discovery to the proposal of rational therapeutic interventions targeting cell adhesion molecules and chemokine receptors–How to make a dream come true. Mem. Inst. Oswaldo Cruz 2009, 104, 226–235. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  187. Dinarello, C.A.; Simon, A.; Van Der Meer, J.W. Treating inflammation by blocking interleukin-1 in a broad spectrum of diseases. Nat. Rev. Drug Discov. 2012, 11, 633–652. [Google Scholar] [CrossRef] [Green Version]
  188. Lamkanfi, M.; Dixit, V.M. Mechanisms and functions of inflammasomes. Cell 2014, 157, 1013–1022. [Google Scholar] [CrossRef] [Green Version]
  189. Ehlers, M.R. Immune-modulating effects of alpha-1 antitrypsin. Biol. Chem. 2014, 395, 1187–1193. [Google Scholar] [CrossRef] [PubMed]
  190. Siebers, K.; Fink, B.; Zakrzewicz, A.; Agne, A.; Richter, K.; Konzok, S.; Hecker, A.; Zukunft, S.; Kullmar, M.; Klein, J.; et al. Alpha-1 Antitrypsin Inhibits ATP-Mediated Release of Interleukin-1beta via CD36 and Teceptors. Front. Immunol. 2018, 9, 877. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  191. Zakrzewicz, A.; Richter, K.; Zakrzewicz, D.; Siebers, K.; Damm, J.; Agne, A.; Hecker, A.; Mcintosh, J.M.; Chamulitrat, W.; Krasteva-Christ, G.; et al. SLPI Inhibits ATP-mediated maturation of IL-1beta in human monocytic leukocytes: A Tyer. Front. Immunol. 2019, 10, 664. [Google Scholar] [CrossRef] [Green Version]
  192. Moreau, T.; Baranger, K.; Dade, S.; Dallet-Choisy, S.; Guyot, N.; Zani, M.L. Multifaceted roles of human elafin and secretory leukocyte proteinase inhibitor (SLPI), two serine protease inhibitors of the chelonianin family. Biochimie 2008, 90, 284–295. [Google Scholar] [CrossRef]
  193. Ma, G.; Greenwell-Wild, T.; Lei, K.; Jin, W.; Swisher, J.; Hardegen, N.; Wild, C.T.; Wahl, S.M. Secretory leukocyte protease inhibitor binds to annexin II, a cofactor for macrophage HIV-1 infection. J. Exp. Med. 2004, 200, 1337–1346. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  194. Gottlieb, P.A.; Alkanani, A.K.; Michels, A.W.; Lewis, E.C.; Shapiro, L.; Dinarello, C.A.; Zipris, D. alpha1-Antitrypsin therapy downregulates toll-like receptor-induced IL-1beta responses in monocytes and myeloid dendritic cells and may improve islet function in recently diagnosed patients with type 1 diabetes. J. Clin. Endocrinol. Metab. 2014, 99, E1418–E1426. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Proposed mechanisms by which iPLA2β activation contributes to monocyte chemotaxis, foam cell formation, vascular and bladder inflammation, tumor cell survival, and macrophage polarization. LPA, lysophosphatidic acid; iDLs, iPLA2β-derived lipids; ROS, reactive oxygen species; EC, endothelial cell ( Biomolecules 11 00577 i001, increased expression/activity of iPLA2β; Biomolecules 11 00577 i002, indicates increase in).
Figure 1. Proposed mechanisms by which iPLA2β activation contributes to monocyte chemotaxis, foam cell formation, vascular and bladder inflammation, tumor cell survival, and macrophage polarization. LPA, lysophosphatidic acid; iDLs, iPLA2β-derived lipids; ROS, reactive oxygen species; EC, endothelial cell ( Biomolecules 11 00577 i001, increased expression/activity of iPLA2β; Biomolecules 11 00577 i002, indicates increase in).
Biomolecules 11 00577 g001
Figure 2. Proposed contribution of iDLs generated by immune cells to β-cell destruction and type-1 diabetes (T1D) development. ( Biomolecules 11 00577 i001, increased expression/activity of iPLA2β; Biomolecules 11 00577 i002, indicates increase in; and Biomolecules 11 00577 i004, indicates decrease in).
Figure 2. Proposed contribution of iDLs generated by immune cells to β-cell destruction and type-1 diabetes (T1D) development. ( Biomolecules 11 00577 i001, increased expression/activity of iPLA2β; Biomolecules 11 00577 i002, indicates increase in; and Biomolecules 11 00577 i004, indicates decrease in).
Biomolecules 11 00577 g002
Figure 3. Proposed mechanisms by which iPLA2β inactivation or activation can have protective outcomes in tumor development and promoting host defenses. ( Biomolecules 11 00577 i003, decreased expression/activity of iPLA2β; Biomolecules 11 00577 i001, increased expression/activity of iPLA2β; Biomolecules 11 00577 i002, indicates increase in; Biomolecules 11 00577 i004, indicates decrease in).
Figure 3. Proposed mechanisms by which iPLA2β inactivation or activation can have protective outcomes in tumor development and promoting host defenses. ( Biomolecules 11 00577 i003, decreased expression/activity of iPLA2β; Biomolecules 11 00577 i001, increased expression/activity of iPLA2β; Biomolecules 11 00577 i002, indicates increase in; Biomolecules 11 00577 i004, indicates decrease in).
Biomolecules 11 00577 g003
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

White, T.D.; Almutairi, A.; Tusing, Y.G.; Lei, X.; Ramanadham, S. The Impact of the Ca2+-Independent Phospholipase A2β (iPLA2β) on Immune Cells. Biomolecules 2021, 11, 577. https://doi.org/10.3390/biom11040577

AMA Style

White TD, Almutairi A, Tusing YG, Lei X, Ramanadham S. The Impact of the Ca2+-Independent Phospholipase A2β (iPLA2β) on Immune Cells. Biomolecules. 2021; 11(4):577. https://doi.org/10.3390/biom11040577

Chicago/Turabian Style

White, Tayleur D., Abdulaziz Almutairi, Ying Gai Tusing, Xiaoyong Lei, and Sasanka Ramanadham. 2021. "The Impact of the Ca2+-Independent Phospholipase A2β (iPLA2β) on Immune Cells" Biomolecules 11, no. 4: 577. https://doi.org/10.3390/biom11040577

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop