Next Article in Journal
The Influence of Synthesis Conditions on the Antioxidant Activity of Selenium Nanoparticles
Next Article in Special Issue
Outtakes from My Journey through the World of LIPID MAPS
Previous Article in Journal
Experimental and Computational Structural Studies of 2,3,5-Trisubstituted and 1,2,3,5-Tetrasubstituted Indoles as Non-Competitive Antagonists of GluK1/GluK2 Receptors
Previous Article in Special Issue
Phosphorylation of cPLA2α at Ser505 Is Necessary for Its Translocation to PtdInsP2-Enriched Membranes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Current Knowledge on Mammalian Phospholipase A1, Brief History, Structures, Biochemical and Pathophysiological Roles

Faculty of Pharmaceutical Sciences, The University of Tokyo, 7-3-1 Hongo, Bunkyo-ku, Tokyo 113, Japan
*
Author to whom correspondence should be addressed.
Molecules 2022, 27(8), 2487; https://doi.org/10.3390/molecules27082487
Submission received: 4 November 2021 / Revised: 28 March 2022 / Accepted: 29 March 2022 / Published: 12 April 2022

Abstract

:
Phospholipase A1 (PLA1) is an enzyme that cleaves an ester bond at the sn-1 position of glycerophospholipids, producing a free fatty acid and a lysophospholipid. PLA1 activities have been detected both extracellularly and intracellularly, which are well conserved in higher eukaryotes, including fish and mammals. All extracellular PLA1s belong to the lipase family. In addition to PLA1 activity, most mammalian extracellular PLA1s exhibit lipase activity to hydrolyze triacylglycerol, cleaving the fatty acid and contributing to its absorption into the intestinal tract and tissues. Some extracellular PLA1s exhibit PLA1 activities specific to phosphatidic acid (PA) or phosphatidylserine (PS) and serve to produce lysophospholipid mediators such as lysophosphatidic acid (LPA) and lysophosphatidylserine (LysoPS). A high level of PLA1 activity has been detected in the cytosol fractions, where PA-PLA1/DDHD1/iPLA1 was responsible for the activity. Many homologs of PA-PLA1 and PLA2 have been shown to exhibit PLA1 activity. Although much has been learned about the pathophysiological roles of PLA1 molecules through studies of knockout mice and human genetic diseases, many questions regarding their biochemical properties, including their genuine in vivo substrate, remain elusive.

1. Introduction

Phospholipase A1 (PLA1) is an enzyme that hydrolyzes an ester bond at the sn-1 position of glycerophospholipids (GPLs), usually producing a saturated or mono-unsaturated fatty acid and a 1-lyso-2-acyl-phospholipid (2-acyl-lysophospholipid, 2-acyl-LPL) (Figure 1). PLA1 has not attracted as much attention as other mammalian acyl hydrolases, such as phospholipase A2 (PLA2), which hydrolyzes fatty acids, mainly unsaturated fatty acids, and acts as a first step in producing enzymes for bioactive lipids such as eicosanoids and platelet-activating factor (PAF) [1,2]. Some PLA1s and PLA2 target neutral lipids such as triacylglycerol (TAG) and diacylglycerol (DAG) in addition to GPLs (Figure 1).
Much is known about the functions of PLA2, whereas those of PLA1 remain limited. However, because fatty acids at both the sn-1 and sn-2 positions of GPLs have a high turnover rate [3], PLA1 as well as PLA2, appears to be involved in the rapid turnover and remodeling of cellular GPLs (Figure 2). In addition, some PLA1s also have a specific role in the production of 2-acyl-1-lysophospholipids, which serve as lysophospholipid mediators. For example, one type of PLA1, membrane-associated phosphatidic acid-selective PLA1 (mPA-PLA1α in Table 1, Figure 3A), produces 2-acyl-1-lysophosphatidic acid (2-acyl-lysoPA (LPA)) with an unsaturated fatty acid residue [4]. The 2-acyl-LPA acts as a potent ligand for LPAR3/EDG7 and LPAR6/P2Y5, with LPA receptors preferring 2-acyl-LPA over 1-acyl-LPA [5,6]. Phosphatidylserine-specific PLA1 (PS-PLA1 in Table 1, Figure 3A) also acts as a producing enzyme of another lysophospholipid mediator, 2-acyl-lysophosphatidylserine (2-acyl-lysoPS (LysoPS)), which further supports the idea that PLA1s function as producing enzymes for lysophospholipid mediators.
Although PLA1 activity has been detected in many mammalian tissues and cells [7,8,9,10,11,12,13,14], only a few PLA1s have been purified and cloned. Some triacylglycerol (TAG)-hydrolyzing lipases (Figure 1) such as hepatic lipase, endothelial cell lipase, and lipoprotein lipase also show PLA1 activity [15].

2. History of PLA1 Research

The following is a brief history of mammalian PLA1 research. PLA1 as well as PLA2 activities have been detected in various tissues and plasma. In the 1990’s, two PLA1 molecules were biochemically purified and identified. These two PLA1s are phosphatidic acid-preferential PLA1 (PA-PLA1) and phosphatidylserine-specific PLA1 (PS-PLA1) (Table 1). In 1994, Higgs and Glomset purified a novel PLA1 from a cytosolic fraction of bovine testes that preferentially hydrolyzed PA [22] (Table 1). Shortly after, the same group cloned a cDNA of the PLA1, and PA-PLA1 was found to be an intracellular protein composed of 875-amino acids with a calculated molecular mass of approximately 100 kDa. Simultaneously, Exton’s group purified a very similar PLA1 from the bovine brain [37]. It was not clear whether the two intracellular PLA1s were identical, because the two groups used different PLA1 assay conditions and substrates. Later, Kudo’s group purified similar PLA1s from the brain and testes of mice, rats, and cows [38] and showed that they were identical to the PA-PLA1 mentioned above. They demonstrated that PA-PLA1 was predominantly hydrolyzed phosphatidic acid (PA) in the presence of Triton X-100 and phosphatidylethanolamine (PE) in its absence. The result showed clearly that the substrate specificity of PA-PLA1 in vitro is affected by the assay conditions, which makes it difficult to identify the natural substrates of PA-PLA1. This also implies that the name, PA-PLA1, is not suitable for the enzyme.
Horigome et al., detected two PLA activities in the supernatant of activated rat platelets [39,40]. One was identified as secretory PLA2, now known as Group IIA secretory phospholipase A2 (sPLA2IIA). Sato et al., succeeded in purifying and cloning another PLA that showed a high preference for serine-containing phospholipids [16]. The PLA is now known as PS-specific PLA1 (PS-PLA1) (Table 1). PS-PLA1 specifically hydrolyzed PS in vitro. In addition, it acted on the intact cell membrane, hydrolyzed PS and produced 2-acyl-LysoPS. Thus, PS-PLA1 is believed to be a LysoPS-producing enzyme.
The 3D structures of three PLA1 family members (extracellular PLA1/lipase, cPLA2 and PLAAT family members) were shown. For extracellular PLA1/lipase family members, β5 and the β9 loops and the lid domain are shown in yellow, cyan and green, respectively. The three residues forming a catalytic triad (Ser, Asp and His) are described as sticks (red). For cPLA2 family members (cPLA2α, cPLA2β and cPLA2ζ), the conserved lipase motifs (GXSGX and DXG) are shown in red, and the two residues forming a catalytic dyad (Ser and Asp) are shown as sticks. For PLAAT3/PLA2G16, the three residues forming a catalytic triad (two histidines and cysteine) are shown as sticks (red). The structures without asterisk were acquired from RCSB Protein Data Bank. (Reference PDB IDs; PL (1LPB), LPL (6OB0), PLRP1 (2PPL), cPLA2α (1CJY), PLA2G16 (4DOT)). The predicted structures of lipases denoted with an asterisk were acquired from AlphaFold Protein Structure Database. All the structures were visualized using PyMOL software.
After discovering the two PLA1 molecules, several researchers found similar PLA1s (homologs) in the nucleotide databases, expressed them as recombinant proteins and characterized them biochemically. These analyses identified the homologs as novel PLA1s. These include extracellular enzyme membrane-associated PA-selective PLA1α (mPA-PLA1α, a PS-PLA1 homologue) [4] (Table 1) and intracellular enzyme iPLA1 γ/DDHD2/KIAA0725 (a PA-PLA1 homologue) [24] and iPLA1β/SEC23IP/p125 (PLA1 activities have not been detected) [27] (Table 1). In addition, similar biochemical characterization was performed for PLA2 homologs. Interestingly, biochemical characterizations of the PLA2 homologs showed that some of them exhibited PLA1 activity in addition to their PLA2 activity (PLB activity (Figure 1)) (Table 1). It should also be noted here that certain lipases that hydrolyze TAG have PLA1 activity, as mentioned above (Table 1). Lipases hydrolyze fatty acids at the sn-1 and sn-3 positions of triacylglycerol (TAG) and diacylglycerol (DAG). They also hydrolyze fatty acids at the sn-1 or sn-3 positions of monoacylglycerol (MAG). This review will summarize current knowledge on PLA1 molecules reported thus far, discussing their discoveries, structures, tissue and cellular distributions, and possible biological functions.
PLA1s are roughly divided into two groups: (1) extracellular PLA1/lipase family and (2) intracellular PLA1 family (Table 1), depending on their cellular localization and primary amino acid sequences. The intracellular PLA1 family was further subdivided into several groups, consisting of an iPLA1 family, PNPLA family, cPLA2 family and PLAAT family.
In the following sections, we will summarize what has been known to date about each family member regarding the history of discovery, biochemical characterization, expression, structural feature and function at animal and cellular levels.

3. Structural Evaluation of PLA1 Molecules

Recently, a computational machine learning method named AlphaFold was developed enabling researchers to predict protein structures with high accuracy, even when no similar experimentally solved structure is available [41]. In this review, we utilized AlphaFold to generate structural predictions of PLA1 molecules. Figure 4 summarizes the AlphaFold-generated structures of some PLA1 molecules focused on in this review. Note that some structures have been determined, e.g., by X-ray crystallography, while others were structure-predicted. Since structures of any PNPLA and iPLA1 family members have not been determined, AlphaFold was unable to predict their structures. Thus, predicted structures were shown only for extracellular PLA1/lipase and for cPLA2 family members (Figure 4).

4. Extracellular PLA1/Lipase Family Members

4.1. Phosphatidylserine-Specific Phospholipase A1 (PS-PLA1)

4.1.1. Historical Aspects

PS-PLA1 was originally found as a lysophosphatidylserine-selective lysophospholipase (LysoPS-lysophospholipase) that was detected in the cell supernatant of activated rat platelets together with Group IIA secretory phospholipase A2 (sPLA2 IIA) [40,42]. The purified LysoPS-lysophospholipase was later found also to exhibit PS-PLA1 activity [16].

4.1.2. Biochemical Characterization and Tissue Distribution

Activated rat platelets secreted PS-PLA1 and sPLA2-IIA, both of which are stored in α−granules [40]. Sato et al., purified PS-PLA1 from a culture medium of thrombin–activated platelets prepared from approximately 1000 rats [16]. In rats, PS-PLA1 is expressed in platelets and in the heart and lung [17,43]. The expression of PS-PLA1 in platelets was species-dependent. For example, PS-PLA1 mRNA was poorly expressed in human platelets but highly expressed in rat platelets.
Interestingly, enhanced PS-PLA1 expression was detected in many tissues when rats were treated with bacterial lipopolysaccharide in vivo (our unpublished observation). Recently, Yatomi et al., established a PS-PLA1 immunoassay [44] and subsequently reported that PS-PLA1 antigen concentrations were variable in clinical samples from patients. Serum PS-PLA1 was significantly elevated in patients with autoimmune disorders, including systemic lupus erythematosus (SLE), rheumatoid arthritis, and Sjogren’s syndrome [45]. Interestingly, the level of PS-PLA1 in each SLE individual showed an excellent association with the SLE disease activity index and decreased after the commencement of medical therapy.
PS-PLA1 reacts specifically with PS and 1-acyl-LysoPS [39,46]. It cannot appreciably hydrolyze any other phospholipids, including phosphatidyl-D-serine. The enzyme is not a type of PLB because it exclusively hydrolyzes an acyl residue bound at the sn-1 position of either PS or 1-acyl-LysoPS. Incubation of the enzyme with 1-acyl-2-radioactive acyl-PS produced radioactive LysoPS but only a tiny amount of radioactive-free fatty acid, even after long-term incubation [16]. It can be concluded that PS-PLA1 recognizes a 1-acyl-glycerophospho-L-serine structure and hydrolyzes an acyl residue at the sn-1 position. PS-PLA1 may first interact with the bilayer structure’s surface to gain access to PS molecules buried in the bilayer.

4.1.3. Structural Characteristics

PS-PLA1 belongs to the lipase family, which is composed of classical lipases such as lipoprotein lipase (LPL), hepatic lipase (HL), endothelial lipase (EL), membrane-associated phosphatidic acid-selective PLA1α (mPA-PLA1α) and mPA-PLA1β (Table 1). They share approximately 30–50% amino acid identity. The sequence homology of PS-PLA1 was about 80.0% between humans [17] and rats [16]. The deduced amino acid sequence contains a catalytic triad composed of active Ser, Asp, His residues, and “lid” surface loops (Figure 4). Interestingly, conventional lipases have long “lid” (22–23 residues) and “β9” loops (18–19 residues), whereas PS-PLA1 has a shorter “lid” (12 residues) and a shorter (deleted) “β9” loop (13 residues) [16,17]. Both loops, which may be strengthened by disulfide bonds linking 14 cysteine residues, have been implicated in substrate recognition. These findings are compatible with the fact that PS-PLA1 activity was sensitive to diisopropylfluorophosphate and dithiothreitol [42].

4.1.4. Possible Functions

LysoPS has been shown to stimulate histamine release from rat peritoneal mast cells triggered by cross-linking of FcεRI, a high-affinity receptor for IgE [47,48,49,50]. Interestingly, recombinant PS-PLA1 protein had activity similar to that of LysoPS [51], indicating that PS-PLA1 interacts with PS in the mast cell plasma membrane to produce LysoPS. Because the PS-PLA1-induced degranulation was significantly enhanced in the presence of apoptotic cells [51], cells that surround the mast cells, such as neutrophils in the peritoneal cavity rather than mast cells, must be the source of PS. For example, a crude preparation of rat peritoneal mast cells (containing other cell types such as mononuclear leukocytes and neutrophils) released appreciable amounts of histamine in the presence of PS-PLA1 and an Fc receptor cross-linker [51].
PS-PLA1 also serves as a LysoPS-producing enzyme for cloned LysoPS receptors (Figure 3A). In a TGFα-shedding assay in which an ectodomain shedding of TGFα detects GPCR activation, PS-PLA1 and LysoPS can induce the TGFα shedding in LPSR1/GPR34 (a LysoPS receptor). Thus, it is likely that PS-PLA1 acts as a LysoPS-producing enzyme for LPSR1/GPR34. This hypothesis is further supported by the fact that PA-PLA1α/LIPH (see below) is a close homolog of PS-PLA1 and serves to produce LPA to activate an LPA receptor, LPAR6.

4.2. Membrane-Associated Phosphatidic Acid-Selective Phospholipase A1s (mPA-PLA1α/LIPH and mPA-PLA1β)

4.2.1. Historical Aspects

Lysophosphatidic acid (LPA or lysoPA) is a lipid mediator with multiple biological functions. These include physiological functions such as brain development (LPAR1), embryo implantation (LPAR3), blood vessel formation (LPAR4 and LPAR6), and hair follicle development (LPAR6), and has also been implicated in pathological conditions such as the development of fibrosis (LPAR1), endometriosis (LPAR3) and obesity (LPAR4). LPA induces platelet aggregation, smooth muscle contraction, stimulation of cell proliferation, formation of actin stress fibers in fibroblasts, and inhibition of neurite outgrowth in neuronal cells at the cellular level. Recent studies have identified a new family of receptor genes for LPA. A total of three G-protein-coupled receptors (GPCR) belonging to the EDG (endothelial differentiation gene) family (LPAR1/EDG2, LPAR2/EDG4, and LPAR3/EDG7) and another three belonging to the P2Y family (LPAR4/P2Y9, LPAR5/GPR192, LPAR6/P2Y5) have been identified [6,52,53]. LPA with an unsaturated fatty acid at the sn-2 position of the glycerol backbone preferentially activates LPAR3 and LPAR6 [5,6]. Interestingly, such LPA with an unsaturated fatty acid may have specific functions. For example, Hayashi et al., found that LPA with an unsaturated fatty acid (unsaturated LPA) but not LPA with a saturated fatty acid stimulated dedifferentiation of rat smooth muscle cells isolated from blood vessels [54]. Tokumura et al., also reported that unsaturated LPA stimulated the proliferation of smooth muscle cells in vitro [55]. Kurano et al., reported that unsaturated LPA such as DHA-LPA and arachidonic acid-containing LPA increased in plasma from acute coronary syndrome (ACS) patients. ACS is the rapid narrowing or occlusion of the coronary artery lumen caused by the collapse of an unstable plaque formed by atherosclerosis and thrombus formation [56].
In the process of the production of unsaturated LPA, PLA1 must be involved. There have been two pathways postulated [57]. In one pathway, phospholipids are subjected to the PLA1 reaction to produce lysophospholipids with unsaturated fatty acids. Then, the unsaturated lysophospholipids are converted to unsaturated LPA by phospholipase D (PLD) type enzymes. Little is known about the identity of PLA1, but the PLD enzyme has been identified as autotaxin (ATX). In the other pathway, phosphatidic acid (PA) is converted to unsaturated LPA by phosphatidic acid-selective phospholipase A1s. Two such PLA1s were cloned and characterized, mPA-PLA1α [4] and mPA-PLA1β [18], both of which are close homologs of PS-PLA1 (Table 1). PA-PLA1α is expressed explicitly in hair follicles and plays a critical role in forming proper structures of hair follicles by activating LPAR6.

4.2.2. Biochemical Characterization and Distribution

When expressed in insect Sf9 cells, neither mPA-PLA1α nor mPA-PLA1β protein was recovered from the culture medium, although both have signal sequences at the N-terminus. Instead, both were recovered from the detergent-resistant membrane domains, referred to as raft structures [18].
Human mPA-PLA1α is most abundantly expressed in the hair follicle, where, as stated above, it regulates the formation of hair follicles [58]. A closely homologous enzyme, mPA-PLA1β, was found in the gene bank. The characteristics of mPA-PLA1β were similar to those of mPA-PLA1α, except that it was exclusively detected in human testes. Although mPA-PLA1β was detected in sperm, only a small amount was detected in seminal fluids [18]. Interestingly, a high mPA-PLA1β expression was frequently observed in some rare testes cancers, including Ewing family tumor, although its role in cancer remains unclear [59].
Both mPA-PLA1α and mPA-PLA1β were found to exhibit PLA1 activity against PA. Other GPL substrates and TAG were not substrates for mPA-PLA1α and mPA-PLA1β [4,18].

4.2.3. Structural Characteristics

The deduced amino acid sequence of mPA-PLA1α shares 34.0% identity with that of human PS-PLA1 [4], and the N-terminal half of the molecule has about 40% identity with the N-terminal catalytic domain of PS-PLA1. In total, three of the amino acid residues in the sequences of mPA-PLA1α, Ser-154, Asp-178, and His-248, are completely conserved among lipases (Figure 4). The deduced amino acid sequences of mPA-PLA1α and mPA-PLA1β are 45.9% identical. The homologous regions were again most prominent in the first half of the molecule. The amino acid lengths of three loop structures, β5, β9 and lid, were exactly the same as in PS-PLA1 and mPA-PLA1β (Figure 4), while they were quite different in other lipase family members. Thus, these loop structures are believed to be involved in the substrate recognition.

4.2.4. Function

In vitro, both mPA-PLA1α and mPA-PLA1β showed PA-specific PLA1 activity [4,18]. Both enzymes appear to produce LPA at the cellular level. High LPA was detected in cells expressing mPA-PLA1α, especially when the cells were treated with bacterial phospholipase D to produce PA on the cell surface [4]. Of note, no appreciable change was observed in the level of any other lysophospholipids under the same conditions [4], indicating that PA-PLA1α has high substrate specificity for PA. In 2006, Kazantseva reported that a gene encoding mPA-PLA1α (also called Lipase H or LIPH) is a causative gene for an inherited form of hair loss and hair growth defect [60]. After the discovery, similar defects in the mPA-PLA1α/LIPH gene were identified all over the world. In addition, a gene whose mutation causes unusual hairiness in dogs and rabbits was reported to encode mPA-PLA1α/LIPH. Moreover, we showed that mPA-PLA1α/LIPH knockout mice displayed similar woolly or wavy hair [58]. Thus, it is now generally accepted that mPA-PLA1α/LIPH plays a critical role in hair growth in a wide range of mammals.
In 2008, Shimomura et al. and Pasternack et al., performed a genetic linkage analysis in several families with autosomal recessive woolly hair and showed linkage to chromosome 13q14.2–14.3 in which an orphan G protein-coupled receptor P2Y5 was mapped [61,62]. Later Yanagida et al., and Inoue et al., showed that P2Y5 was a receptor for LPA, now called LPAR6. Of note, LPAR6 prefers LPA with an unsaturated fatty acid at the sn-2 position. Thus, it is reasonable to assume that mPA-PLA1α/LIPH is an enzyme that supplies LPA to GPCR-type LPA receptors, also called LPAR6/P2Y5 receptors [63] (Figure 3A).

4.3. Extracellular Lipases

In addition to above-described PS- and PA-specific PLA1s, specific extracellular lipases exhibit PLA1 activity in addition to their intrinsic lipase activity to hydrolyze triacylglycerol (TAG) (Figure 1). These include pancreatic lipase (PL), lipoprotein lipase (LPL), hepatic lipase (HL), endothelial lipase (EL), and pancreatic lipase-related protein 2 (PLRP2) (Table 1), all of which belong to the pancreatic lipase family together with the PS-PLA1, mPA-PLA1α and mPA-PLA1β as mentioned earlier.

4.3.1. Historical Aspects

Lipases hydrolyze TAG present in food or the blood. PL and PLRP2 are secreted from the pancreas and hydrolyze TAG in the small intestine. By contrast, LPL, HL, and EL are mainly present in the blood, where they bind to the surfaces of endothelial cells via heparan sulfate proteoglycans. Both types of lipases catalyze a reaction to hydrolyze ester bonds of TAG at either sn-1 or sn-3 positions to produce diacylglycerols and fatty acids [15,19]. Interestingly, in addition to their lipase activities, all lipases exhibit considerable PLA1 and lysophospholipase activities toward PC and LPC. The precise substrate specificities of the lipases have not been demonstrated, but PC is the most probable phospholipid substrate since its levels are much higher than those of other phospholipid species in both blood and foods.

4.3.2. Structural Characteristics

As noted above, all lipases belong to the pancreatic lipase gene family. The family is conserved in a wide range of animals, from insects to mammals [15]. Presently eight members are known in humans (Table 1). Crystallographic studies of human pancreatic lipases [20] revealed that each of the lipases are composed of N- and C-terminal domains. The N-terminal domains are the sites of catalytic activity. Interestingly, all lipase-like molecules observed in insects (Drosophila melanogaster, Bombyx mori, and hornet) bear only the N-terminal domain [64], supporting the idea that the N-terminal domain is required for catalytic activity. Three amino acid residues in the N-terminal domains, Ser, Asp, and His, form a catalytic triad located in N-terminal domains and are conserved in the pancreatic lipase gene family (Figure 4). In addition, well-conserved cysteine residues that form intramolecular disulfide bonds are present in the N-terminal domains. These structural features indicate that common catalytic machinery is used for both PL and TAG hydrolysis. The crystallographic studies also revealed that human PL contains three surface loops called the lid, the β5 loop, and the β9 loop covering the active sites and have been implicated in the substrate specificity [20,21] (Figure 4). Interestingly, as mentioned above the length of the loop structures are quite different between lipases (PL, LPL, HL and EL) and PLA1s (PS-PLA1, mPA-PLA1α and mPA-PLA1β), indicating that these loops determine the substrate specificity of lipase family members.

4.3.3. Functional Characteristics

PL and PLRP2 are present in pancreatic juice and contribute to the intestinal absorption of lipids by hydrolyzing various types of glycerolipids, including TAG, diacylglycerol (DAG), monoacylglycerol (MAG), phospholipids, and lysophospholipids (Figure 3B). The resulting free fatty acids, MAG, and lysophospholipids are then absorbed by intestinal epithelial cells and re-build TAG in the form of lipoproteins (chylomicrons) in the cells.
LPL, HL, and EL share a common role in the metabolism of lipoproteins (Figure 3C). LPL is mainly synthesized by adipose tissues and other tissues such as heart muscle, and HL and EL are synthesized by hepatocytes and endothelial cells, respectively. LPL and HL play roles in the tissue uptake of free fatty acids from lipoproteins. LPL is bound to the capillary endothelium and supplies the underlying tissues with fatty acids derived from the TAG-rich chylomicrons and very-low-density lipoprotein (VLDL). EL and HL exhibit PLA1 activities toward PC on HDL in addition to their TAG-lipase activities. Their PLA1 activities are partly responsible for their actions in lipoprotein metabolism [65,66]. Interestingly, EL predominantly exhibits PLA1 activity, whereas HL exhibits both PLA1 and TAG-hydrolyzing activities. Of note, EL inhibition dramatically increases the level of HDL, which suggests that EL PLA1 activity has a role in the metabolism of HDL, which is rich in GPLs.

5. Intracellular PLA1 Families

5.1. iPLA1 Family

5.1.1. Historical Aspects

Most mammalian cells contain cytosolic PLA1s. Cytosolic PLA1 activities have been studied in various tissues, including the heart, brain, and testes. Accordingly, a PLA1 protein, designated phosphatidic acid-preferring phospholipase A1 (PA-PLA1, iPLA1a) (Table 1), was purified from the brain [37] and testes [22,67]. As with the activities of other lipolytic enzymes, the activities of the PA-PLA1s were affected considerably by assay conditions. Using a mixed micelle system, Glomset and colleagues [22] found that a 110 kDa enzyme from testes preferentially hydrolyzed PA, which provides the reason for its name PA-PLA1. They cloned a bovine cDNA for PA-PLA1 from bovine [23]. PA-PLA1 was shown to be a unique phospholipase since it lacked sequence similarity to any phospholipases identified thus far, including extracellular PLA1s, lysophospholipase, LCAT, and triacylglycerol lipases. Since PA-PLA1 has a DDHD domain, it is also called DDHD1 (Table 1).
Later, two orthologues of PA-PLA1 were identified in the database (Table 1). One was p125 which was also identified as a protein that interacted with mammalian Sec23p [27]. Sec23p is a component of the COPII coat that functions in the production of membrane traffic vesicles from the ER [68]. The p125 protein is localized in the ER Golgi intermediate compartment (ERGIC) and cis-Golgi, and its overexpression causes the dispersion of these membrane compartments, suggesting that it is involved in the early secretory pathway. Interestingly, p125 did not exhibit appreciable PLA1 activity. The other is KIAA0725p which is also known as DDHD2. KIAA0725p was shown to have PLA1 activity for various GPLs, including phosphatidic acid in an assay system containing Triton X-100.
The structures of neither iPLA1α/PA-PLA1/DDHD1, iPLA1β/SEC23IP nor iPLA1γ/KIAA0725p/DDHD2 have been determined. Therefore, AlphaFold poorly predicted the precise structures of these proteins, except near the active center, where they displayed weak structural similarities with other hydrolases (Figure 4).

5.1.2. Phosphatidic Acid-Preferring Phospholipase A1 (PA-PLA1)/DDHD1/iPLA1α

Characterization and Distribution

PA-PLA1 is highly expressed in the testes and brain in a wide range of mammals. It has multiple phosphorylated serine and threonine residues [69,70]. The molecular mass of the expressed PA-PLA1 was estimated to be about 110 kDa [22]. The PLA1 activity of PA-PLA1 was inhibited by methyl arachidonoyl fluorophosphonate (MAFP), an inhibitor of certain PLA2s, but not by diisopropylfluorophosphate (DFP) [71] a potent inhibitor of PA-PLA1. The activity of PA-PLA1 was Ca2+-independent at a neutral pH but Ca2+-dependent at an alkaline pH [71].

Substrate Specificity

Studies in vitro showed that the activity of PA-PLA1 against PA was 4- to 10-fold greater than the activities of PA-PLA1 against PI, PS, PE, and PC in a Triton X-100 mixed micelle system [71]. In the absence of Triton X-100, PE was the best substrate for PA-PLA1 [38]. In addition, the PLA1 activities were affected by the presence of divalent cations and GPL composition of the substrate [72]. Thus, further studies are needed to identify the real substrates in the cells and in vivo. Precise lipidomic analyses will clarify the substrate in the near future.

Function

Spastic paraplegias (SPGs) are neurological disorders characterized by spasticity and gait disturbance. More than 60 types of SPGs caused by mutations in different genes have been reported [73]. SPG type 28 (SPG28) is one type of SPG caused by recessive mutations in the gene encoding PA-PLA1α/DDHD1 [74]. How the deficiency of PA-PLA1α/DDHD1 leads to the development of the disease has not been identified, but it was shown recently, that the level of lysophosphatidylinositol (LPI) with arachidonic acid (20:4) at the sn-2 position decreased in the brains of aged PA-PLA1α/DDHD1 knockout mice [75,76], suggesting the potential role of 20:4-LPI in PA-PLA1α/DDHD1-mediated signals.
PA-PLA1α/DDHD1 is also highly expressed in the testes, where it plays a role in spermatogenesis. Baba et al., reported that PA-PLA1α/DDHD1-deficient mice have abnormally shaped sperm, possibly due to the dysregulation of mitochondria of the cells [77]. They speculated that PA metabolism is involved in mitochondrial regulation since PA has been implicated in the fusion and fission cycles of mitochondria. However, one should be cautious about identifying the substrates of PA-PLA1/DDHD1, because PA may not be an endogenous substrate for PA-PLA1/DDHD1 as discussed above.

5.1.3. KIAA0725p/DDHD2, a Second Member of Intracellular PLA1

Historical Aspects

A search of the databases revealed a protein encoded by the human expressed sequence tag clone KIAA0725p that exhibited strong sequence similarity (52.6%) to p125 throughout the entire sequence determined [24]. KIAA0725p has 711 amino acids with a calculated molecular mass of about 80 kDa, and similar to p125, it has a Gly-X-Ser-X-Gly consensus sequence. KIAA0725p also contains a DDHD domain and thus has also been called DDHD2 (Table 1).

Characterization and Distribution

KIAA0725p is a cytosolic protein. Both Northern and Western blot analyses showed a ubiquitous distribution of KIAA0725p [24]. However, unlike PA-PLA1/DDHD1 most of the KIAA0725p/DDHD2 protein was recovered in the membrane fraction. Immunostaining of the protein indicated KIAA0725p/DDHD2 was associated with membrane structures. The post-nuclear supernatant of cells transfected with KIAA0725p cDNA showed high hydrolytic activities against PE and PA and low activities against PS and PC in the absence of Triton X-100. In the presence of Triton X-100, however, it had activity against PA, but only weak activity against PE [24]. KIAA0725p/DDHD2 also exhibits TAG and DAG lipase activities [25,26].

Function

Overexpression of KIAA0725p was found to cause a morphological change of organelles, such as dispersion of the ERGIC and Golgi apparatus [24]. Morikawa et al., showed that KIAA0725p is one of the factors mediating a membrane transport pathway between the ER and the Golgi that is distinct from the previously characterized COPI- and Rab6-dependent pathways [78]. KIAA0725p/DDHD2 mutations cause autosomal recessive hereditary spastic paraplegia (SPG54) and abnormal mitochondrial morphology [79]. KIAA0725p/DDHD2 KO mice also show defects in locomotion and cognition, and KO MEF exhibits mitochondrial dysfunction [26,80]. Little is known about the lipids responsible for these abnormalities. Further studies are needed to identify the function of the protein.

5.2. Other Recently Identified Intracellular PLA1s

After the human genome project was completed, many putative PLA molecules were identified in the databases that showed significant homology to previously identified and characterized PLA1/PLA2 molecules. Of note, some molecules were indeed shown to display PLA1, lysophospholipase, and phospholipase B activities. An example is PLB, which hydrolyzes a fatty acid at both sn-1 and sn-2 positions. Here, we summarize the molecular features of these enzymes.

5.2.1. iPLA2/PNPLA Family

The iPLA2 (Ca2+-independent phospholipase A2) family members adopt a three-layer α/β/α architecture and harbor an enzymatically active site composed of a Ser-Asp catalytic dyad. They are also called PNPLA (patatin-like phospholipase domain-containing) family members [81] because they do not require Ca2+ ions for their activity and have a domain structure similar to that of plant lipase patatin family members. Some of these enzymes exhibit not only PLA2 activity but also PLA1 and lysophospholipase activity (PLB (phospholipase B) activity). Among the eight members of the iPLA2 family, at least three of them (PNPLA6/iPLA2δ/NTE, PNPLA7/iPLA2θ/NRE, and PNPLA8/iPLA2γ/group VIB PLA2) (Table 1) appear to have PLA1 activity.

PNPLA6/iPLA2δ/NTE

  • Historical Aspects
PNPLA6 is also known as NTE (neural target esterase) since it was identified as a target molecule for organophosphorus insecticides, which cause neurological disorders such as paralysis of the lower limbs due to degeneration of axons in the spinal cord and peripheral nerves [82]. Later the endogenous substrates for this enzyme were suggested to be GPLs, as shown in the following sections.
  • Biochemical Characterization and Tissue Distribution
PNPLA6 shows lysophospholipase activity toward lysophospholipids (especially LPC), producing fatty acids and glycerophosphocholine (GPC) [28]. The enzyme is also conserved in yeast and fruit flies, and these orthologues show phospholipase B activity, hydrolyzing fatty acids both at the sn-1 and sn-2 positions of PC to produce LPC and GPC [29,30]. PNPLA6 is highly expressed in the nervous system.
  • Structural Characteristics
PNPLA6 contains an extended N-terminal domain and three nucleotide-binding motifs in contrast to the common Patatin domain of the iPLA2 family. PNPLA6 has a transmembrane region near the N-terminus, which was not found in other family members except for PLPLA7.
  • Possible Functions
Systemic homozygous deficiency of PNPLA6 in mice has been reported to result in fetal lethality due to placental dysfunction. Heterozygous mice show an increased spontaneous locomotor activity and hyperactivity tendencies [82,83]. Progressive neurodegenerative abnormalities were observed in the hippocampus, thalamus, and cerebellum in mice with neuron-specific deficiency [84]. Mutations near the active center of PNPLA6 in humans have been reported to be responsible for hereditary diseases such as spasticity, ataxia, hypogonadism, and retinal degeneration (Boucher–Neuhäuser syndrome, Laurence–Moon syndrome, Oliver–McFarlane syndrome, and spastic paraplegia) [85,86,87]. The molecular mechanism underlying the PNPLA6-deficiency-induced neurodegeneration remains obscure. However, it has been proposed that glycerophosphocholine (GPC), the final product of PLB reaction from PC, produced in the kidney, may function as an osmotic regulator since its expression is induced by high NaCl concentrations [88].

PNPLA7/iPLA2θ/NRE

  • Historical Aspects
PNPLA7 is a close relative of PNPLA6/NTE and is therefore named NRE (NTE-related esterase).
  • Biochemical Characterization and Tissue Distribution
Similar to PNPLA6, PNPLA7 shows lysophospholipase activity for LPC, producing fatty acids and GPC [31]. While PNPLA6 is predominantly expressed in the nervous system, PNPLA7 is mainly found in peripheral tissues such as the testes, skeletal muscle, heart, and adipose tissue.
  • Structural Characteristics
PNPLA7 is a close homolog of PNPLA6 [31,89] and has a similar domain structure to PNPLA6, with an N-terminal transmembrane region and three nucleotide-binding motifs [31,89].
  • Possible Functions
PNPLA7 is involved in choline metabolism by hydrolyzing LPC [31]. Interestingly, PNPLA7 is localized to the ER and tiny lipid droplets (LDs) in a catalytic domain-dependent manner [90]. PNPLA7-deficient mice were recently shown to have a premature aging phenotype and perish before adulthood due to metabolic defects, particularly in the liver (Makoto Murakami personal communication). However, the underlying molecular mechanisms remain unclear.

PNPLA8/iPLA2γ/group VIB PLA2

  • Historical Aspects
PNPLA8 is synonymous with a Ca2+-independent PLA2γ (iPLA2γ) and was originally cloned from human heart cDNA [91].
  • Biochemical Characterization and Tissue Distribution
PNPLA8/iPLA2γ exhibits a unique enzymatic activity [91]. It hydrolyzes a fatty acid at the sn-2 position (PLA2 activity) when alkenyl PC is used as a substrate in vitro. It also hydrolyzes a fatty acid at the sn-1 position (PLA1 activity) when diacyl PC with a polyunsaturated fatty acid at the sn-2 position is used as a substrate [32]. In humans, PNPLA8/iPLA2γ mRNA expression is exceptionally high in the heart [91,92]. Immunohistochemical analysis in mouse myocardium demonstrated that PNPLA8/iPLA2γ is associated with mitochondria and peroxisomes, reflecting dual protein localization motifs.
  • Structural Characteristics
PNPLA8 is a close homolog of PNPLA9/iPLA2β and has a similar active center but less homology at the N-terminus [93]. PNPLA9/iPLA2β has four possible translation initiation sites, giving rise to proteins of 88, 77, 74, and 63 kDa. The N- and C-termini have mitochondrial and peroxisomal localization sequences, respectively, and the full-length 88- and 63 kDa translation products were observed to localize to mitochondria and peroxisomes, respectively, in the cell [94,95].
  • Possible Functions
Liu et al., suggested that PNPLA8 is involved in cardiolipin remodeling and is essential for maintaining mitochondrial function [96]. A variety of phenotypes have been reported in PNPLA8-deficient mice [33], including failure to thrive, reduced resistance to exercise and myocardial stress, muscle weakness, decreased body temperature, adipose tissue atrophy, and neurodegeneration. Mutations in PNPLA8 in humans are associated with mitochondrial myopathy [97]. In a mouse model of myocardial ischemia, the levels of potential lipid mediators, including eicosanoids and lysophospholipids were altered in PNPLA8-deficient and PNPLA8-overexpressing mice, suggesting that PNPLA8 also has a function in the production of these lipid mediators [98]. In the liver, PNPLA8 is a candidate enzyme that supplies LPC, the substrate of PNPLA7.

5.2.2. cPLA2 Family

The Group IV PLA2 family is comprised of six intracellular enzymes commonly called cytosolic PLA2α (cPLA2α), cPLA2β, cPLA2γ, cPLA2δ, cPLA2ε, and cPLA2ζ. Interestingly, they are mostly homologous to PLA and PLB/lysophospholipase of filamentous fungi, particularly in regions containing conserved residues involved in catalysis. Their primary functions appear to be exerted through their PLA2 activity providing arachidonic acid for synthesis of eicosanoids. However, some members, i.e., cPLA2α/PLA2G4A/GroupIVA PLA2, cPLA2β/PLA2G4B/GroupIVB PLA2, and cPLA2ζ/PLA2G4F/Group IVF PLA2 also exhibit PLA1 activity [34,35]. It should be noted that all these cPLA2s have their own abilities to carry out multiple reactions to varying degrees (PLA2, PLA1, lysophospholipase and transacylase activities). Thus, the conclusion that an enzyme has PLA1 activity must be evaluated carefully, as an enzyme with both PLA2 and lysophospholipase activity will apparently be judged to exhibit PLA1 activity in an in vitro assay. The function of the PLA1 activity of these enzymes remains completely unknown.
The X-ray crystal structures of cPLA2 family members were determined only for cPLA2α [99] (Figure 4). The structure explained the selectivity of the enzyme with arachidonic acid-containing GPLs using an unusual Ser-Asp dyad located in a deep cleft at the center of a predominantly hydrophobic funnel. AlphaFold predicted similar structures for cPLA2β and cPLA2ζ (Figure 4).

PLA2G16/group XVI PLA2/PLAAT3/HRASLS3/H-Rev107

  • Historical Aspects
PLAAT3 was found by subtraction cloning between sensitive and resistant fibroblasts to transformation by introducing H-Ras [100]. Overexpression of PLAAT3 in H-Ras-transformed fibroblasts suppressed cell proliferation and colony formation, indicating that PLAAT3 negatively regulates H-Ras’ function [101].
  • Biochemical Characterization and Tissue Distribution
PLAAT3 hydrolyses fatty acids at the sn-1 or sn-2 positions of PC and PE. Thus, it has both PLA1 and PLA2 activities, although the former is stronger than the latter. In vitro, PLAAT3 also shows weak N-acyltransferase and O-acyltransferase activities compared to other PLAAT family members [36,102]. PLAAT3 is highly expressed in adipocytes and is also known as adipose-specific PLA2 (Ad-PLA2) [103].
  • Structural Characteristics
PLAAT3 has homology to lecithin-retinol acyltransferase (LRAT) and belongs to the NlpC/P60 superfamily. PLAAT3 contains a histidine-containing three-amino acid residue called the H-box and a cysteine-containing domain called the NC domain, which is thought to form the active site. In addition, it bears a proline-rich domain at the N-terminal side and a membrane-binding domain with a cluster of hydrophobic amino acids at the C-terminal side. These domains were reported to be essential for the regulation of H-Ras function [101]. A recent crystal structure analysis of PLAAT3 (Figure 4) revealed the catalytic mechanism of the enzyme [104].
  • Possible Functions
PLAAT3-deficient mice are resistant to diet-induced obesity [105]. PLAAT3 is also involved in cancer invasion and metastasis [106] and is known to be involved in vitamin A metabolism [107], promotion of peroxisome formation, and production of ether GPLs [108]. Recently, Morishita et al., reported that PLAAT3 was involved in organelle degradation in the eye lens [109]. The eye lens of vertebrates is composed of fiber cells in which all membrane-bound organelles, including mitochondria, the endoplasmic reticulum, and lysosomes, undergo degradation during terminal differentiation to form an organelle-free zone. PLAAT3 in mammals and Plaat1 (functional homolog) in zebrafish were shown to be essential in the organelle degradation in the eye lens.
Interestingly, these enzymes translocate from the cytosol to various organelles immediately before organelle degradation. The C-terminal transmembrane domain of the enzymes was shown to be essential to the process. It is speculated that for the complete digestion of GPLs, both PLA1 and PLA2 activities (PLB activity) of PLAAT3 are needed.

6. Conclusions and Perspectives

In this article, we summarize our current understanding of PLA1 molecular structures and its activities. Genome projects have revealed a wide range of PLA1-like molecules in the human genome, and subsequent biochemical studies have also revealed their nature. Of note, extracellular PLA1 molecules such as mPA-PLA1α and PS-PLA1 are now more clearly accepted as enzymes that produce lysophospholipid mediators such as LPA and LysoPS. Intracellular PLA1s were also identified and an analysis of their mutants revealed that they exhibited essential physiological functions. However, the precise substrates and products and the physiological significance of the enzymatic reactions remain to be elucidated. Comprehensive lipidomics analysis of PLA1 mutants and their overexpressing cells will help elucidate the true substrates, products, and functions of PLA1 in future studies.

Author Contributions

J.A. wrote most of the manuscript. S.Y. and H.K. carried out a literature survey on the PLA1 activity of various PLA2 molecules. S.Y. created Table 1 and Figure 4, and H.K. created Figure 1, Figure 2, and Figure 3. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by AMED-LEAP (18gm0010004h0002; J.A.), KAKENHI (20K21379; J.A. and 20K15984; H.K.).

Acknowledgments

We are grateful to Makoto Murakami (The University of Tokyo) for sharing much information about the PLA2 molecules and their PLA1 activities.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

PLA1; phospholipase A1, PLA2; phospholipase A2, PLB; phospholipase B, GPL; glycerophospholipid, LPL; lysophospholipid, PC; phosphatidylcholine, PE; phosphatidylethanolamine, PS; phosphatidylserine, PI; phosphatidylinositol, PG; phosphatidylglycerol, PA; phosphatidic acid, LPC; lysophosphatidylcholine, LPA; lysophosphatidic acid, LysoPS; lysophosphatidylserine, TAG, triacylglycerol, DAG; diacylglycerol, MAG; monoacylglycerol, PS-PLA1; PS-specific PLA1, PA-PLA1; PA-preferential PLA1, mPA-PLA1α; membrane-associated PA-selective PLA1α, mPA-PLA1β; membrane-associated PA-selective PLA1β, PL; pancreatic lipase, LPL; lipoprotein lipase, HL; hepatic lipase, EL; endothelial lipase, ATX; autotaxin, iPLA1; intracellular PLA1, iPLA2; intracellular PLA2, sPLA2; secretory PLA2, cPLA2; cytosolic PLA2, PNPLA; patatin-like phospholipase, PLAAT; phospholipase A and acyltransferase.

References

  1. Dennis, E.A. The growing phospholipase A2 superfamily of signal transduction enzymes. Trends Biochem. Sci. 1997, 22, 1–2. [Google Scholar] [CrossRef]
  2. Murakami, M.; Nakatani, Y.; Atsumi, G.; Inoue, K.; Kudo, I. Regulatory functions of phospholipase A2. Crit. Rev. Immunol. 1997, 17, 225–283. [Google Scholar] [CrossRef] [PubMed]
  3. Lykidis, A.; Jackowski, S. Regulation of mammalian cell membrane biosynthesis. Prog. Nucleic Acid Res. Mol. Biol. 2001, 65, 361–393. [Google Scholar] [PubMed]
  4. Sonoda, H.; Aoki, J.; Hiramatsu, T.; Ishida, M.; Bandoh, K.; Nagai, Y.; Taguchi, R.; Inoue, K.; Arai, H. A novel phosphatidic acid-selective phospholipase A1 that produces lysophosphatidic acid. J. Biol. Chem. 2002, 277, 34254–34263. [Google Scholar] [CrossRef] [Green Version]
  5. Bandoh, K.; Aoki, J.; Tsujimoto, M.; Arai, H.; Inoue, K. Lysophosphatidic acid (LPA) receptors of the EDG family are differentially activated by LPA species. Structure-activity relationship of cloned LPA receptors. FEBS Lett. 2000, 478, 159–165. [Google Scholar] [CrossRef] [Green Version]
  6. Bandoh, K.; Aoki, J.; Hosono, H.; Kobayashi, S.; Kobayashi, T.; Murakami, M.K.; Tsujimoto, M.; Arai, H.; Inoue, K. Molecular cloning and characterization of a novel human G-protein-coupled receptor, EDG7, for lysophosphatidic acid. J. Biol. Chem. 1999, 274, 27776–27785. [Google Scholar] [CrossRef] [Green Version]
  7. Robinson, M.; Waite, M. Physical-chemical requirements for the catalysis of substrates by lysosomal phospholipase A1. J. Biol. Chem. 1983, 258, 14371–14378. [Google Scholar] [CrossRef]
  8. Kucera, G.L.; Miller, C.; Sisson, P.J.; Wilcox, R.W.; Wiemer, Z.; Waite, M. Hydrolysis of thioester analogs by rat liver phospholipase A1. J. Biol. Chem. 1988, 263, 12964–12969. [Google Scholar] [CrossRef]
  9. Ueda, H.; Kobayashi, T.; Kishimoto, M.; Tsutsumi, T.; Okuyama, H. A possible pathway of phosphoinositide metabolism through EDTA-insensitive phospholipase A1 followed by lysophosphoinositide-specific phospholipase C in rat brain. J. Neurochem. 1993, 61, 1874–1881. [Google Scholar] [CrossRef]
  10. Ueda, H.; Kobayashi, T.; Kishimoto, M.; Tsutsumi, T.; Watanabe, S.; Okuyama, H. The presence of Ca(2+)-independent phospholipase A1 highly specific for phosphatidylinositol in bovine brain. Biochem. Biophys. Res. Commun. 1993, 195, 1272–1279. [Google Scholar] [CrossRef]
  11. Ueda, H.; Kobayashi, T.; Kishimoto, M.; Tsutsumi, T.; Okuyama, H. EDTA-insensitive deacylation of phosphatidylinositol in porcine platelet membranes. Life Sci. 1993, 53, 629–634. [Google Scholar] [CrossRef]
  12. Hostetler, K.Y.; Gardner, M.F.; Giordano, J.R. Purification of lysosomal phospholipase A and demonstration of proteins that inhibit phospholipase A in a lysosomal fraction from rat kidney cortex. Biochemistry 1986, 25, 6456–6461. [Google Scholar] [CrossRef]
  13. Waite, M. The Phospholipases: Handbook of Lipid Research; Plenum Publishing Corp: New York, NY, USA, 1987; pp. 79–110. [Google Scholar]
  14. Van den Bosch, H. Phosphaolipases; Elsevier: Amsterdam, The Netherlands, 1982; pp. 313–351. [Google Scholar]
  15. Carriere, F.; Withers, M.C.T.H.; van Roussel, A.; Cambillau, C.; Verger, R. Structural basis for the substrate selectivity of pancreatic lipases and some related proteins. Biochim. Biophys. Acta 1998, 1376, 417–432. [Google Scholar] [CrossRef]
  16. Sato, T.; Aoki, J.; Nagai, Y.; Dohmae, N.; Takio, K.; Doi, T.; Arai, H.; Inoue, K. Serine phospholipid-specific phospholipase A that is secreted from activated platelets. A new member of the lipase family. J. Biol. Chem. 1997, 272, 2192–2198. [Google Scholar] [CrossRef] [Green Version]
  17. Nagai, Y.; Aoki, J.; Sato, T.; Amano, K.; Matsuda, Y.; Arai, H.; Inoue, K. An alternative splicing form of phosphatidylserine-specific phospholipase A1 that exhibits lysophosphatidylserine-specific lysophospholipase activity in humans. J. Biol. Chem. 1999, 274, 11053–11059. [Google Scholar] [CrossRef] [Green Version]
  18. Hiramatsu, T.; Sonoda, H.; Takanezawa, Y.; Morikawa, R.; Ishida, M.; Kasahara, K.; Sanai, Y.R.; Taguchi, J.; Aoki, H.A. Biochemical and molecular characterization of two phosphatidic acid-selective phospholipase A1s, mPA-PLA1alpha and mPA-PLA1beta. J. Biol. Chem. 2003, 278, 49438–49447. [Google Scholar] [CrossRef] [Green Version]
  19. Hide, W.A.; Chan, L.; Li, W.H. Structure and evolution of the lipase superfamily. J. Lipid. Res. 1992, 33, 167–178. [Google Scholar] [CrossRef]
  20. Winkler, F.K.; D’Arcy, A.; Hunziker, W. Structure of human pancreatic lipase. Nature 1990, 343, 771–774. [Google Scholar] [CrossRef]
  21. Carriere, F.; Thirstrup, K.; Hjorth, S.; Ferrato, F.; Nielsen, P.F.; Withers-Martinez, C.; Cambillau, C.; Boel, E.; Thim, L.; Vergerc, R. Pancreatic lipase structure-function relationships by domain exchange. Biochemistry 1997, 36, 239–248. [Google Scholar] [CrossRef]
  22. Higgs, H.N.; Glomset, J.A. Identification of a phosphatidic acid-preferring phospholipase A1 from bovine brain and testis. Proc. Natl. Acad. Sci. USA 1994, 91, 9574–9578. [Google Scholar] [CrossRef] [Green Version]
  23. Higgs, H.N.; Han, M.H.; Johnson, G.E.; Glomset, J.A. Cloning of a phosphatidic acid-preferring phospholipase A1 from bovine testis. J. Biol. Chem. 1998, 273, 5468–5477. [Google Scholar] [CrossRef] [Green Version]
  24. Nakajima, K.; Sonoda, H.; Mizoguchi, T.; Aoki, J.; Arai, H.; Nagahama, M.; Tagaya, M.; Tani, K. A novel phospholipase A1 with sequence homology to a mammalian Sec23p-interacting protein, p125. J. Biol. Chem. 2002, 277, 11329–11335. [Google Scholar] [CrossRef] [Green Version]
  25. Araki, M.; Ohshima, N.; Aso, C.; Konishi, A.; Obinata, H.; Tatei, K.; Izumi, T. Enzymatic characterization of recombinant rat DDHD2: A soluble diacylglycerol lipase. J. Biochem. 2016, 160, 269–279. [Google Scholar] [CrossRef]
  26. Inloes, J.M.; Hsu, K.L.; Dix, M.M.; Viader, A.; Masuda, K.; Takei, T.; Wood, M.R.; Cravatt, B.F. The hereditary spastic paraplegia-related enzyme DDHD2 is a principal brain triglyceride lipase. Proc. Natl. Acad. Sci. USA 2014, 111, 14924–14929. [Google Scholar] [CrossRef] [Green Version]
  27. Tani, K.; Mizoguchi, T.; Iwamatsu, A.; Hatsuzawa, K.; Tagaya, M. p125 is a novel mammalian Sec23p-interacting protein with structural similarity to phospholipid-modifying proteins. J. Biol. Chem. 1999, 274, 20505–20512. [Google Scholar] [CrossRef] [Green Version]
  28. Quistad, G.B.; Barlow, C.; Winrow, C.J.; Sparks, S.E.; Casida, J.E. Evidence that mouse brain neuropathy target esterase is a lysophospholipase. Proc. Natl. Acad. Sci. USA 2003, 100, 7983–7987. [Google Scholar] [CrossRef] [Green Version]
  29. Zaccheo, O.; Dinsdale, D.; Meacock, P.A.; Glynn, P. Neuropathy target esterase and its yeast homologue degrade phosphatidylcholine to glycerophosphocholine in living cells. J. Biol. Chem. 2004, 279, 24024–24033. [Google Scholar] [CrossRef] [Green Version]
  30. Muhlig-Versen, M.; da Cruz, A.B.; Tschape, J.A.; Moser, M.; Buttner, R.; Athenstaedt, K.; Glynn, P.; Kretzschmar, D. Loss of Swiss cheese/neuropathy target esterase activity causes disruption of phosphatidylcholine homeostasis and neuronal and glial death in adult Drosophila. J. Neurosci. 2005, 25, 2865–2873. [Google Scholar] [CrossRef] [Green Version]
  31. Kienesberger, P.C.; Lass, A.; Preiss-Landl, K.; Wolinski, H.; Kohlwein, S.D.; Zimmermann, R.; Zechner, R. Identification of an insulin-regulated lysophospholipase with homology to neuropathy target esterase. J. Biol. Chem. 2008, 283, 5908–5917. [Google Scholar] [CrossRef] [Green Version]
  32. Liu, X.; Moon, S.H.; Jenkins, C.M.; Sims, H.F.; Gross, R.W. Cyclooxygenase-2 Mediated Oxidation of 2-Arachidonoyl-Lysophospholipids Identifies Unknown Lipid Signaling Pathways. Cell Chem. Biol. 2016, 23, 1217–1227. [Google Scholar] [CrossRef] [Green Version]
  33. Mancuso, D.J.; Sims, H.F.; Han, X.; Jenkins, C.M.; Guan, S.P.; Yang, K.; Moon, S.H.; Pietka, T.; Abumrad, N.A.; Schlesinger, P.H.; et al. Genetic ablation of calcium-independent phospholipase A2gamma leads to alterations in mitochondrial lipid metabolism and function resulting in a deficient mitochondrial bioenergetic phenotype. J. Biol. Chem. 2007, 282, 34611–34622. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Pickard, R.T.; Strifler, B.A.; Kramer, R.M.; Sharp, J.D. Molecular cloning of two new human paralogs of 85-kDa cytosolic phospholipase A2. J. Biol. Chem. 1999, 274, 8823–8831. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Song, C.; Chang, X.J.; Bean, K.M.; Proia, M.S.; Knopf, J.L.; Kriz, R.W. Molecular characterization of cytosolic phospholipase A2-beta. J. Biol. Chem. 1999, 274, 17063–17067. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Uyama, T.; Morishita, J.; Jin, X.H.; Okamoto, Y.; Tsuboi, K.; Ueda, N. The tumor suppressor gene H-Rev107 functions as a novel Ca2+-independent cytosolic phospholipase A1/2 of the thiol hydrolase type. J. Lipid. Res. 2009, 50, 685–693. [Google Scholar] [CrossRef] [Green Version]
  37. Pete, M.J.; Ross, A.H.; Exton, J.H. Purification and properties of phospholipase A1 from bovine brain. J. Biol. Chem. 1994, 269, 19494–19500. [Google Scholar] [CrossRef]
  38. Uchiyama, S.; Miyazaki, Y.; Amakasu, Y.; Kuwata, H.; Nakatani, Y.; Atsumi, G.; Murakami, M.; Kudo, I. Characterization of heparin low-affinity phospholipase A1 present in brain and testicular tissue. J. Biochem. 1999, 125, 1001–1010. [Google Scholar] [CrossRef]
  39. Horigome, K.; Hayakawa, M.; Inoue, K.; Nojima, S. Selective release of phospholipase A2 and lysophosphatidylserine-specific lysophospholipase from rat platelets. J. Biochem. 1987, 101, 53–61. [Google Scholar] [CrossRef]
  40. Horigome, K.; Hayakawa, M.; Inoue, K.; Nojima, S. Purification and characterization of phospholipase A2 released from rat platelets. J. Biochem. 1987, 101, 625–631. [Google Scholar] [CrossRef]
  41. Jumper, J.; Evans, R.; Pritzel, A.; Green, T.; Figurnov, M.; Ronneberger, O.; Tunyasuvunakool, K.; Bates, R.; Žídek, A.; Potapenko, A.; et al. Highly accurate protein structure prediction with AlphaFold. Nature 2021, 596, 583–589. [Google Scholar] [CrossRef]
  42. Higashi, S.; Kobayashi, T.; Kudo, I.; Inoue, K. Purification and characterization of lysophospholipase released from rat platelets. J. Biochem. 1988, 103, 442–447. [Google Scholar] [CrossRef]
  43. Aoki, J.; Nagai, Y.; Hosono, H.; Inoue, K.; Arai, H. Structure and function of phosphatidylserine-specific phospholipase A1. Biochim. Biophys. Acta 2002, 1582, 26–32. [Google Scholar] [CrossRef]
  44. Nakamura, K.; Igarashi, K.; Ohkawa, R.; Saiki, N.; Nagasaki, M.; Uno, K.; Hayashi, N.; Sawada, T.; Syukuya, K.; Yokota, H.; et al. A novel enzyme immunoassay for the determination of phosphatidylserine-specific phospholipase A(1) in human serum samples. Clin. Chim. Acta 2010, 411, 1090–1094. [Google Scholar] [CrossRef]
  45. Sawada, T.; Kurano, M.; Shirai, H.; Iwasaki, Y.; Tahara, K.; Hayashi, H.; Igarashi, K.; Fujio, K.; Aoki, J.; Yatomi, Y. Serum phosphatidylserine-specific phospholipase A 1 as a novel biomarker for monitoring systemic lupus erythematosus disease activity. Int. J. Rheum. Dis. 2019, 22, 2059–2066. [Google Scholar] [CrossRef]
  46. Yokoyama, K.; Kudo, I.; Inoue, K. Phospholipid degradation in rat calcium ionophore-activated platelets is catalyzed mainly by two discrete secretory phospholipase As. J. Biochem. 1995, 117, 1280–1287. [Google Scholar] [CrossRef]
  47. Martin, T.W.; Lagunoff, D. Interactions of lysophospholipids and mast cells. Nature 1979, 279, 250–252. [Google Scholar] [CrossRef]
  48. Tamori, N.Y.; Horigome, K.; Inoue, K.; Nojima, S. Metabolism of lysophosphatidylserine, a potentiator of histamine release in rat mast cells. J. Biochem. 1986, 100, 581–590. [Google Scholar] [CrossRef]
  49. Horigome, K.; Tamori, N.Y.; Inoue, K.; Nojima, S. Effect of serine phospholipid structure on the enhancement of concanavalin A-induced degranulation in rat mast cells. J. Biochem. 1986, 100, 571–579. [Google Scholar] [CrossRef]
  50. Bruni, A.; Monastra, G.; Bellini, F.; Toffano, G. Autacoid properties of lysophosphatidylserine. Prog. Clin. Biol. Res. 1988, 282, 165–179. [Google Scholar]
  51. Hosono, H.; Aoki, J.; Nagai, Y.; Bandoh, K.; Ishida, M.; Taguchi, R.; Arai, H.; Inoue, K. Phosphatidylserine-specific phospholipase A1 stimulates histamine release from rat peritoneal mast cells through production of 2-acyl-1-lysophosphatidylserine. J. Biol. Chem. 2001, 276, 29664–29670. [Google Scholar] [CrossRef] [Green Version]
  52. Hecht, J.H.; Weiner, J.A.; Post, S.R.; Chun, J. Ventricular zone gene-1 (vzg-1) encodes a lysophosphatidic acid receptor expressed in neurogenic regions of the developing cerebral cortex. J. Cell Biol. 1996, 135, 1071–1083. [Google Scholar] [CrossRef] [Green Version]
  53. An, S.; Bleu, T.; Hallmark, O.G.; Goetzl, E.J. Characterization of a novel subtype of human G protein-coupled receptor for lysophosphatidic acid. J. Biol. Chem. 1998, 273, 7906–7910. [Google Scholar] [CrossRef] [Green Version]
  54. Hayashi, K.; Takahashi, M.; Nishida, W.; Yoshida, K.; Ohkawa, Y.; Kitabatake, A.; Aoki, J.; Arai, H.; Sobue, K. Phenotypic modulation of vascular smooth muscle cells induced by unsaturated lysophosphatidic acids. Circ. Res. 2001, 89, 251–258. [Google Scholar] [CrossRef] [Green Version]
  55. Tokumura, A.; Iimori, M.; Nishioka, Y.; Kitahara, M.; Sakashita, M.; Tanaka, S. Lysophosphatidic acids induce proliferation of cultured vascular smooth muscle cells from rat aorta. Am. J. Physiol. 1994, 267, C204–C210. [Google Scholar] [CrossRef]
  56. Kurano, M.; Suzuki, A.; Inoue, A.; Tokuhara, Y.; Kano, K.; Matsumoto, H.; Igarashi, K.; Ohkawa, R.; Nakamura, K.; Dohi, T.; et al. Possible involvement of minor lysophospholipids in the increase in plasma lysophosphatidic acid in acute coronary syndrome. Arterioscler. Thromb. Vasc. Biol. 2015, 35, 463–470. [Google Scholar] [CrossRef] [Green Version]
  57. Aoki, J.; Inoue, A.; Okudaira, S. Two pathways for lysophosphatidic acid production. Biochim. Biophys Acta 2008, 1781, 513–518. [Google Scholar] [CrossRef]
  58. Inoue, A.; Arima, N.; Ishiguro, J.; Prestwich, G.D.; Arai, H.; Aoki, J. LPA-producing enzyme PA-PLA1α regulates hair follicle development by modulating EGFR signalling. EMBO J. 2011, 30, 4248–4260. [Google Scholar] [CrossRef] [Green Version]
  59. Foell, J.L.; Hesse, M.; Volkmer, I.; Schmiedel, B.J.; Neumann, I.; Staege, M.S. Membrane-associated phospholipase A1 beta (LIPI) Is an Ewing tumour-associated cancer/testis antigen. Pediatr. Blood Cancer 2008, 51, 228–234. [Google Scholar] [CrossRef]
  60. Kazantseva, A.; Goltsov, A.; Zinchenko, R.; Grigorenko, A.P.; Abrukova, A.V.; Moliaka, Y.K.; Kirillov, A.G.; Guo, Z.; Lyle, S.; Ginter, E.K.; et al. Human hair growth deficiency is linked to a genetic defect in the phospholipase gene LIPH. Science 2006, 314, 982–985. [Google Scholar] [CrossRef] [Green Version]
  61. Shimomura, Y.; Wajid, M.; Ishii, Y.; Shapiro, L.; Petukhova, L.; Gordon, D.; Christiano, A.M. Disruption of P2RY5, an orphan G protein-coupled receptor, underlies autosomal recessive woolly hair. Nat. Genet. 2008, 40, 335–339. [Google Scholar] [CrossRef]
  62. Pasternack, S.M.; von Kügelgen, I.; al Aboud, K.; Lee, T.A.; Rüschendorf, F.; Voss, K.; Hillmer, A.M.; Molderings, G.J.; Franz, T.; Ramirez, A.; et al. G protein-coupled receptor P2Y5 and its ligand LPA are involved in maintenance of human hair growth. Nat. Genet. 2008, 40, 329–334. [Google Scholar] [CrossRef]
  63. Yanagida, K.; Masago, K.; Nakanishi, H.; Kihara, Y.; Hamano, F.; Tajima, Y.; Taguchi, R.; Shimizu, T.; Ishii, S. Identification and characterization of a novel lysophosphatidic acid receptor, p2y5/LPA6. J. Biol. Chem. 2009, 284, 17731–17741. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Soldatova, L.; Kochoumian, L.; King, T.P. Sequence similarity of a hornet (D. maculata) venom allergen phospholipase A1 with mammalian lipases. FEBS Lett. 1993, 320, 145–149. [Google Scholar] [CrossRef] [Green Version]
  65. Shamburek, R.D.; Zech, L.A.; Cooper, P.S.; Vandenbroek, J.M.; Schwartz, C.C. Disappearance of two major phosphatidylcholines from plasma is predominantly via LCAT and hepatic lipase. Am. J. Physiol. 1996, 271, E1073–E1082. [Google Scholar] [CrossRef] [PubMed]
  66. MMcCoy, G.; Sun, G.S.; Marchadier, D.; Maugeais, C.; Glick, J.M.; Rader, D.J. Characterization of the lipolytic activity of endothelial lipase. J. Lipid. Res. 2002, 43, 921–929. [Google Scholar] [CrossRef]
  67. Higgs, H.N.; Glomset, J.A. Purification and properties of a phosphatidic acid-preferring phospholipase A1 from bovine testis. Examination of the molecular basis of its activation. J. Biol. Chem. 1996, 271, 10874–10883. [Google Scholar] [CrossRef] [Green Version]
  68. Barlowe, C.; Schekman, R. SEC12 encodes a guanine-nucleotide-exchange factor essential for transport vesicle budding from the ER. Nature 1993, 365, 347–349. [Google Scholar] [CrossRef]
  69. Han, M.H.; Han, D.K.; Aebersold, R.H.; Glomset, J.A. Effects of protein kinase CK2, extracellular signal-regulated kinase 2, and protein phosphatase 2A on a phosphatidic acid-preferring phospholipase A1. J. Biol. Chem. 2001, 276, 27698–27708. [Google Scholar] [CrossRef] [Green Version]
  70. Matsumoto, N.; Nemoto-Sasaki, Y.; Oka, S.; Arai, S.; Wada, I.; Yamashita, A. Phosphorylation of human phospholipase A1 DDHD1 at newly identified phosphosites affects its subcellular localization. J. Biol. Chem. 2021, 297, 100851. [Google Scholar] [CrossRef]
  71. Miyazawa, D.; Ikemoto, A.; Fujii, Y.; Okuyama, H. Partial purification and characterization of phosphatidic acid-specific phospholipase A(1) in porcine platelet membranes. Biochim. Biophys. Acta 2003, 1631, 17–25. [Google Scholar] [CrossRef]
  72. Pete, M.J.; Exton, J.H. Phospholipid interactions affect substrate hydrolysis by bovine brain phospholipase A1. Biochim. Biophys. Acta 1995, 1256, 367–373. [Google Scholar] [CrossRef]
  73. Giudice, T.L.; Lombardi, F.; Santorelli, F.M.; Kawarai, T.; Orlacchio, A. Hereditary spastic paraplegia: Clinical-genetic characteristics and evolving molecular mechanisms. Exp. Neurol. 2014, 261, 518–539. [Google Scholar] [CrossRef]
  74. Tesson, C.; Nawara, M.; Salih, M.A.; Rossignol, R.; Zaki, M.S.; al Balwi, M.; Schule, R.; Mignot, C.; Obre, E.; Bouhouche, A.; et al. Alteration of fatty-acid-metabolizing enzymes affects mitochondrial form and function in hereditary spastic paraplegia. Am. J. Hum. Genet. 2012, 91, 1051–1064. [Google Scholar] [CrossRef] [Green Version]
  75. Yamashita, A.; Kumazawa, T.; Koga, H.; Suzuki, N.; Oka, S.; Sugiura, T. Generation of lysophosphatidylinositol by DDHD domain containing 1 (DDHD1): Possible involvement of phospholipase D/phosphatidic acid in the activation of DDHD1. Biochim. Biophys. Acta 2010, 1801, 711–720. [Google Scholar] [CrossRef]
  76. Inloes, J.M.; Jing, H.; Cravatt, B.F. The Spastic Paraplegia-Associated Phospholipase DDHD1 Is a Primary Brain Phosphatidylinositol Lipase. Biochemistry 2018, 57, 5759–5767. [Google Scholar] [CrossRef]
  77. Baba, T.; Kashiwagi, Y.; Arimitsu, N.; Kogure, T.; Edo, A.; Maruyama, T.; Nakao, K.; Nakanishi, H.; Kinoshita, M.; Frohman, M.A.; et al. Phosphatidic acid (PA)-preferring phospholipase A1 regulates mitochondrial dynamics. J. Biol. Chem. 2014, 289, 11497–11511. [Google Scholar] [CrossRef] [Green Version]
  78. Morikawa, R.K.; Aoki, J.; Kano, F.; Murata, M.; Yamamoto, A.; Tsujimoto, M.; Arai, H. Intracellular phospholipase A1gamma (iPLA1gamma) is a novel factor involved in coat protein complex I- and Rab6-independent retrograde transport between the endoplasmic reticulum and the Golgi complex. J. Biol. Chem. 2009, 284, 26620–26630. [Google Scholar] [CrossRef] [Green Version]
  79. Gonzalez, M.; Nampoothiri, S.; Kornblum, C.; Oteyza, A.C.; Walter, J.; Konidari, I.; Hulme, W.; Speziani, F.; Schöls, L.; Züchner, S.; et al. Mutations in phospholipase DDHD2 cause autosomal recessive hereditary spastic paraplegia (SPG54). Eur. J. Hum. Genet. 2013, 21, 1214–1218. [Google Scholar] [CrossRef] [Green Version]
  80. Maruyama, T.; Baba, T.; Maemoto, Y.; Hara-Miyauchi, C.; Hasegawa-Ogawa, M.; Enda, Y.O.; Matsumoto, K.; Arimitsu, N.; Nakao, K.; Hamamoto, H.; et al. Loss of DDHD2, whose mutation causes spastic paraplegia, promotes reactive oxygen species generation and apoptosis. Cell Death Dis. 2018, 9, 797. [Google Scholar] [CrossRef] [Green Version]
  81. Kienesberger, P.C.; Oberer, M.; Lass, A.; Zechner, R. Mammalian patatin domain containing proteins: A family with diverse lipolytic activities involved in multiple biological functions. J. Lipid. Res. 2009, 50, S63–S68. [Google Scholar] [CrossRef] [Green Version]
  82. Winrow, C.J.; Hemming, M.L.; Allen, D.M.; Quistad, G.B.; Casida, J.E.; Barlow, C. Loss of neuropathy target esterase in mice links organophosphate exposure to hyperactivity. Nat. Genet. 2003, 33, 477–485. [Google Scholar] [CrossRef] [Green Version]
  83. Moser, M.; Li, Y.; Vaupel, K.; Kretzschmar, D.; Kluge, R.; Glynn, P.; Buettner, R. Placental failure and impaired vasculogenesis result in embryonic lethality for neuropathy target esterase-deficient mice. Mol. Cell. Biol. 2004, 24, 1667–1679. [Google Scholar] [CrossRef] [Green Version]
  84. Akassoglou, K.; Malester, B.; Xu, J.; Tessarollo, L.; Rosenbluth, J.; Chao, M.V. Brain-specific deletion of neuropathy target esterase/swisscheese results in neurodegeneration. Proc. Natl. Acad. Sci. USA 2004, 101, 5075–5080. [Google Scholar] [CrossRef] [Green Version]
  85. Rainier, S.; Bui, M.; Mark, E.; Thomas, D.; Tokarz, D.; Ming, L.; Delaney, C.E.; Richardson, R.J.; Albers, J.W.; Matsunami, N.; et al. Neuropathy target esterase gene mutations cause motor neuron disease. Am. J. Hum. Genet. 2008, 82, 780–785. [Google Scholar] [CrossRef] [Green Version]
  86. Synofzik, M.; Gonzalez, M.A.; Lourenco, C.M.; Coutelier, M.; Haack, T.B.; Rebelo, A.; Hannequin, D.; Strom, T.M.; Prokisch, H.; Kernstock, C.; et al. PNPLA6 mutations cause Boucher-Neuhauser and Gordon Holmes syndromes as part of a broad neurodegenerative spectrum. Brain 2014, 137, 69–77. [Google Scholar] [CrossRef]
  87. Kmoch, S.; Majewski, J.; Ramamurthy, V.; Cao, S.; Fahiminiya, S.; Ren, H.; MacDonald, I.M.; Lopez, I.; Sun, V.; Keser, V.; et al. Mutations in PNPLA6 are linked to photoreceptor degeneration and various forms of childhood blindness. Nat. Commun. 2015, 6, 5614. [Google Scholar] [CrossRef] [Green Version]
  88. Gallazzini, M.; Ferraris, J.D.; Kunin, M.; Morris, R.G.; Burg, M.B. Neuropathy target esterase catalyzes osmoprotective renal synthesis of glycerophosphocholine in response to high NaCl. Proc. Natl. Acad. Sci. USA 2006, 103, 15260–15265. [Google Scholar] [CrossRef] [Green Version]
  89. Richardson, R.J.; Hein, N.D.; Wijeyesakere, S.J.; Fink, J.K.; Makhaeva, G.F. Neuropathy target esterase (NTE): Overview and future. Chem. Biol. Interact. 2013, 203, 238–244. [Google Scholar] [CrossRef]
  90. Heier, C.; Huang, B.K.F.; Eichmann, T.O.; Xie, H.; Zechner, R.; Chang, P. The phospholipase PNPLA7 functions as a lysophosphatidylcholine hydrolase and interacts with lipid droplets through its catalytic domain. J. Biol. Chem. 2017, 292, 19087–19098. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  91. Moon, S.H.; Jenkins, C.M.; Liu, X.; Guan, S.; Mancuso, D.J.; Gross, R.W. Activation of mitochondrial calcium-independent phospholipase A2γ (iPLA2γ) by divalent cations mediating arachidonate release and production of downstream eicosanoids. J. Biol. Chem. 2012, 287, 14880–14895. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Hazen, S.L.; Stuppy, R.J.; Gross, R.W. Purification and characterization of canine myocardial cytosolic phospholipase A2. A calcium-independent phospholipase with absolute f1-2 regiospecificity for diradyl glycerophospholipids. J. Biol. Chem. 1990, 265, 10622–10630. [Google Scholar] [CrossRef]
  93. Jabůrek, M.; Průchová, P.; Holendová, B.; Galkin, A.; Ježek, P. Antioxidant Synergy of Mitochondrial Phospholipase PNPLA8/iPLA2γ with Fatty Acid-Conducting SLC25 Gene Family Transporters. Antioxidants 2021, 10, 678. [Google Scholar] [CrossRef]
  94. Mancuso, D.J.; Jenkins, C.M.; Sims, H.F.; Cohen, J.M.; Yang, J.; Gross, R.W. Complex transcriptional and translational regulation of iPLAgamma resulting in multiple gene products containing dual competing sites for mitochondrial or peroxisomal localization. Eur. J. Biochem. 2004, 271, 4709–4724. [Google Scholar] [CrossRef]
  95. Murakami, M.; Masuda, S.; Ueda-Semmyo, K.; Yoda, E.; Kuwata, H.; Takanezawa, Y.; Aoki, J.; Arai, H.; Sumimoto, H.; Ishikawa, Y.; et al. Group VIB Ca2+-independent phospholipase A2gamma promotes cellular membrane hydrolysis and prostaglandin production in a manner distinct from other intracellular phospholipases A2. J. Biol. Chem. 2005, 280, 14028–14041. [Google Scholar] [CrossRef] [Green Version]
  96. Liu, G.Y.; Moon, S.H.; Jenkins, C.M.; Li, M.; Sims, H.F.; Guan, S.; Gross, R.W. The phospholipase iPLA 2 γ is a major mediator releasing oxidized aliphatic chains from cardiolipin, integrating mitochondrial bioenergetics and signaling. J. Biol. Chem. 2017, 292, 10672–10684. [Google Scholar] [CrossRef] [Green Version]
  97. Saunders, C.J.; Moon, S.H.; Liu, X.; Thiffault, I.; Coffman, K.; LePichon, J.B.; Taboada, E.; Smith, L.D.; Farrow, E.G.; Miller, N.; et al. Loss of function variants in human PNPLA8 encoding calcium-independent phospholipase A2 γ recapitulate the mitochondriopathy of the homologous null mouse. Hum. Mutat. 2015, 36, 301–306. [Google Scholar] [CrossRef] [Green Version]
  98. Moon, S.H.; Mancuso, D.J.; Sims, H.F.; Liu, X.; Nguyen, A.L.; Yang, K.; Guan, S.; Dilthey, B.G.; Jenkins, C.M.; Weinheimer, C.J.; et al. Cardiac Myocyte-specific Knock-out of Calcium-independent Phospholipase A2γ (iPLA2γ) Decreases Oxidized Fatty Acids during Ischemia/Reperfusion and Reduces Infarct Size. J. Biol. Chem. 2016, 291, 19687–19700. [Google Scholar] [CrossRef] [Green Version]
  99. Dessen, A.; Tang, J.; Schmidt, H.; Stahl, M.; Clark, J.D.; Seehra, J.; Somers, W.S. Crystal structure of human cytosolic phospholipase A2 reveals a novel topology and catalytic mechanism. Cell 1999, 97, 349–360. [Google Scholar] [CrossRef] [Green Version]
  100. Hajnal, A.; Klemenz, R.; Schäfer, R. Subtraction cloning of H-rev107, a gene specifically expressed in H-ras resistant fibroblasts. Oncogene 1994, 9, 479–490. [Google Scholar]
  101. Sers, C.; Emmenegger, U.; Husmann, K.; Bucher, K.; Andres, A.C.; Schäfer, R. Growth-inhibitory activity and downregulation of the class II tumor-suppressor gene H-rev107 in tumor cell lines and experimental tumors. J. Cell. Biol. 1997, 136, 935–944. [Google Scholar] [CrossRef]
  102. Shinohara, N.; Uyama, T.; Jin, X.H.; Tsuboi, K.; Tonai, T.; Houchi, H.; Ueda, N. Enzymological analysis of the tumor suppressor A-C1 reveals a novel group of phospholipid-metabolizing enzymes. J. Lipid. Res. 2011, 52, 1927–1935. [Google Scholar] [CrossRef] [Green Version]
  103. Duncan, E.; Sarkadi-Nagy, E.; Jaworski, K.; Ahmadian, M.; Sul, H.S. Identification and functional characterization of adipose-specific phospholipase A2 (AdPLA). J. Biol. Chem. 2008, 283, 25428–25436. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Golczak, M.; Kiser, P.D.; Sears, A.E.; Lodowski, D.T.; Blaner, W.S.; Palczewski, K. Structural basis for the acyltransferase activity of lecithin:retinol acyltransferase-like proteins. J. Biol. Chem. 2012, 287, 23790–23807. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Jaworski, K.; Ahmadian, M.; Duncan, R.E.; Sarkadi-Nagy, E.; Varady, K.A.; Hellerstein, M.K.; Lee, H.Y.; Samuel, V.T.; Shulman, G.I.; Kim, K.H.; et al. AdPLA ablation increases lipolysis and prevents obesity induced by high-fat feeding or leptin deficiency. Nat. Med. 2009, 15, 159–168. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Xiong, S.; Tu, H.; Kollareddy, M.; Pant, V.; Li, Q.; Zhang, Y.; Jackson, J.G.; Suh, Y.A.; Elizondo-Fraire, A.C.; Yang, P.; et al. Pla2g16 phospholipase mediates gain-of-function activities of mutant p53. Proc. Natl. Acad. Sci. USA 2014, 111, 11145–11150. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Golczak, M.; Sears, A.E.; Kiser, P.D.; Palczewski, K. LRAT-specific domain facilitates vitamin A metabolism by domain swapping in HRASLS3. Nat. Chem. Biol. 2015, 11, 26–32. [Google Scholar] [CrossRef] [Green Version]
  108. Uyama, T.; Kawai, K.; Kono, N.; Watanabe, M.; Tsuboi, K.; Inoue, T.; Araki, N.; Arai, H.; Ueda, N. Interaction of Phospholipase A/Acyltransferase-3 with Pex19p: A POSSIBLE INVOLVEMENT IN THE DOWN-REGULATION OF PEROXISOMES. J. Biol. Chem. 2015, 290, 17520–17534. [Google Scholar] [CrossRef] [Green Version]
  109. Morishita, H.; Eguchi, T.; Tsukamoto, S.; Sakamaki, Y.; Takahashi, S.; Saito, C.; Koyama-Honda, I.; Mizushima, N. Organelle degradation in the lens by PLAAT phospholipases. Nature 2021, 592, 634–638. [Google Scholar] [CrossRef]
Figure 1. Structures of glycerolipids and their metabolic enzymes. (A) Glycerophospholipids (GPL) and phospholipases. GPLs are composed of a polar head group (six major classes), a glycerol backbone and fatty acid moieties (esterified at the sn-1 and sn-2 positions). Phospholipase A1 (PLA1) hydrolyzes a fatty acid at the sn-1 position, generating sn-2-acyl-1-lyso-phospholipids (sn-2-acyl LPLs), while phospholipase A2 (PLA2) hydrolyzes a fatty acid at the sn-2 position generating sn-1-acyl-2-lyso-phospholipids (sn-1-acyl LPLs). Phospholipase B (PLB) hydrolyzes a fatty acid at both sn-1 and sn-2 positions. LysoPLA hydrolyzes a fatty acid of sn-2-acyl LPLs and sn-1-acyl LPLs, generating glycerophosphate compounds. (B) Triacylglycerol (TAG) has three fatty acids at the sn-1, sn-2 and sn-3 positions of glycerol backbone, diacylglycerol (DAG) has two fatty acids and monoacylglycerol (MAG) has one fatty acid. TAG lipase hydrolyzes a fatty acid of TAG, generating sn-1, 2, sn-2, 3 or sn-1, 3-diacylglycerols (DAGs). DAG lipase hydrolyzes a fatty acid of DAG and MAG lipase hydrolyzes a fatty acid of MAG.
Figure 1. Structures of glycerolipids and their metabolic enzymes. (A) Glycerophospholipids (GPL) and phospholipases. GPLs are composed of a polar head group (six major classes), a glycerol backbone and fatty acid moieties (esterified at the sn-1 and sn-2 positions). Phospholipase A1 (PLA1) hydrolyzes a fatty acid at the sn-1 position, generating sn-2-acyl-1-lyso-phospholipids (sn-2-acyl LPLs), while phospholipase A2 (PLA2) hydrolyzes a fatty acid at the sn-2 position generating sn-1-acyl-2-lyso-phospholipids (sn-1-acyl LPLs). Phospholipase B (PLB) hydrolyzes a fatty acid at both sn-1 and sn-2 positions. LysoPLA hydrolyzes a fatty acid of sn-2-acyl LPLs and sn-1-acyl LPLs, generating glycerophosphate compounds. (B) Triacylglycerol (TAG) has three fatty acids at the sn-1, sn-2 and sn-3 positions of glycerol backbone, diacylglycerol (DAG) has two fatty acids and monoacylglycerol (MAG) has one fatty acid. TAG lipase hydrolyzes a fatty acid of TAG, generating sn-1, 2, sn-2, 3 or sn-1, 3-diacylglycerols (DAGs). DAG lipase hydrolyzes a fatty acid of DAG and MAG lipase hydrolyzes a fatty acid of MAG.
Molecules 27 02487 g001
Figure 2. Fatty acid remodeling reactions of GPLs. Glycerophospholipids (GPL) in the cells are constantly subjected to two kinds of fatty acid hydrolyzing reactions mediated by phospholipase A1 (PLA1) and phospholipase A2 (PLA2), resulting in the production of 2-acyl-1-lysophospholipid (sn-2 LPL) and 1-acyl-2-lysophospholipid (sn-1 LPL). The LPLs thus produced are further subjected to acylation reactions to re-form the GPLs. Several kinds of lysophospholipid acyltransferases (LPLAT) are responsible for the introduction of fatty acids to lysophospholipids. By these sequential GPL remodeling reactions, the fatty acids of GPLs are constantly replaced.
Figure 2. Fatty acid remodeling reactions of GPLs. Glycerophospholipids (GPL) in the cells are constantly subjected to two kinds of fatty acid hydrolyzing reactions mediated by phospholipase A1 (PLA1) and phospholipase A2 (PLA2), resulting in the production of 2-acyl-1-lysophospholipid (sn-2 LPL) and 1-acyl-2-lysophospholipid (sn-1 LPL). The LPLs thus produced are further subjected to acylation reactions to re-form the GPLs. Several kinds of lysophospholipid acyltransferases (LPLAT) are responsible for the introduction of fatty acids to lysophospholipids. By these sequential GPL remodeling reactions, the fatty acids of GPLs are constantly replaced.
Molecules 27 02487 g002
Figure 3. Physiological roles of extracellular PLA1s. (A) PS-PLA1 and mPA-PLA1α serve as producing enzymes for lysophospholipid mediators. PS-PLA1 has a strict substrate specificity in that it only acts on serine containing GPLs such as phosphatidylserine (PS) and lysophosphatidylserine (LysoPS). LysoPS then acts on GPCR-type LysoPS receptors. Three such LysoPS receptors have been identified. These include LPSR1/GPR34, LPSR2/P2Y10, and LPSR3/GPR174. mPA-PLA1α acts on PA in a specific manner and produces sn-2 LPA, which then acts on GPCR-type LPA receptors, LPAR1-LPAR6 evoking various biological responses. (B) Pancreatic lipase (PL) is secreted from the pancreas into the lumen of the intestine, where it, with the aid of bile acids, hydrolyzes the fatty acids of triacylglycerol (TAG) and GPLs in the digestive juice yielding diacylglycerol (DAG), monoacylglycerol (MAG) and fatty acids. The liberated fatty acids are absorbed by intestinal cells as nutrients. (C) Lipoprotein lipase (LPL), hepatic lipase (HL), and endothelial lipase (EL), which are mainly present in the blood, are associated with endothelial cell surfaces in adipose tissues (LPL), heart (LPL), liver (HL) and various tissues. These lipases have both TAG lipase and PLA1 activities. They hydrolyze fatty acids of TAG and GPLs present in the circulating lipoproteins such as low-density lipoproteins (LDL) and high-density lipoproteins (HDL), yielding diacylglycerol (DAG), monoacylglycerol (MAG), lysophospholipids (LPL) and fatty acids. The free fatty acids are absorbed by corresponding cells for energy source and storage of neutral lipids.
Figure 3. Physiological roles of extracellular PLA1s. (A) PS-PLA1 and mPA-PLA1α serve as producing enzymes for lysophospholipid mediators. PS-PLA1 has a strict substrate specificity in that it only acts on serine containing GPLs such as phosphatidylserine (PS) and lysophosphatidylserine (LysoPS). LysoPS then acts on GPCR-type LysoPS receptors. Three such LysoPS receptors have been identified. These include LPSR1/GPR34, LPSR2/P2Y10, and LPSR3/GPR174. mPA-PLA1α acts on PA in a specific manner and produces sn-2 LPA, which then acts on GPCR-type LPA receptors, LPAR1-LPAR6 evoking various biological responses. (B) Pancreatic lipase (PL) is secreted from the pancreas into the lumen of the intestine, where it, with the aid of bile acids, hydrolyzes the fatty acids of triacylglycerol (TAG) and GPLs in the digestive juice yielding diacylglycerol (DAG), monoacylglycerol (MAG) and fatty acids. The liberated fatty acids are absorbed by intestinal cells as nutrients. (C) Lipoprotein lipase (LPL), hepatic lipase (HL), and endothelial lipase (EL), which are mainly present in the blood, are associated with endothelial cell surfaces in adipose tissues (LPL), heart (LPL), liver (HL) and various tissues. These lipases have both TAG lipase and PLA1 activities. They hydrolyze fatty acids of TAG and GPLs present in the circulating lipoproteins such as low-density lipoproteins (LDL) and high-density lipoproteins (HDL), yielding diacylglycerol (DAG), monoacylglycerol (MAG), lysophospholipids (LPL) and fatty acids. The free fatty acids are absorbed by corresponding cells for energy source and storage of neutral lipids.
Molecules 27 02487 g003
Figure 4. Structures of PLA1 molecules. 3D structures of three PLA1 family members (extracellular PLA1/lipase, cPLA2 and PLAAT family members) were shown. For extracellular PLA1/lipase family members, β5 and the β9 loops and the lid domain are shown in yellow, cyan and green, respectively. The three residues forming a catalytic triad (Ser, Asp and His) are described as sticks (red). For cPLA2 family members (cPLA2α, cPLA2β and cPLA2ζ), the conserved lipase motifs (GXSGX and DXG) are shown in red, and the two residues forming a catalytic dyad (Ser and Asp) are shown as sticks. For PLAAT3/PLA2G16, the three residues forming a catalytic triad (two histidines and cysteine) are shown as sticks (red). The structures without asterisk were acquired from RCSB Protein Data Bank. (Reference PDB IDs; PL (1LPB), LPL (6OB0), PLRP1 (2PPL), cPLA2α (1CJY), PLA2G16 (4DOT)). The predicted structures of lipases with asterisk were acquired from AlphaFold Protein Structure Database. All the structures were visualized using PyMOL software.
Figure 4. Structures of PLA1 molecules. 3D structures of three PLA1 family members (extracellular PLA1/lipase, cPLA2 and PLAAT family members) were shown. For extracellular PLA1/lipase family members, β5 and the β9 loops and the lid domain are shown in yellow, cyan and green, respectively. The three residues forming a catalytic triad (Ser, Asp and His) are described as sticks (red). For cPLA2 family members (cPLA2α, cPLA2β and cPLA2ζ), the conserved lipase motifs (GXSGX and DXG) are shown in red, and the two residues forming a catalytic dyad (Ser and Asp) are shown as sticks. For PLAAT3/PLA2G16, the three residues forming a catalytic triad (two histidines and cysteine) are shown as sticks (red). The structures without asterisk were acquired from RCSB Protein Data Bank. (Reference PDB IDs; PL (1LPB), LPL (6OB0), PLRP1 (2PPL), cPLA2α (1CJY), PLA2G16 (4DOT)). The predicted structures of lipases with asterisk were acquired from AlphaFold Protein Structure Database. All the structures were visualized using PyMOL software.
Molecules 27 02487 g004
Table 1. Mammalian PLA1s.
Table 1. Mammalian PLA1s.
Primary NameOther NameHuman GeneSubstrateReaction MediatedPDB ID
(H: Human, R: Rat)
Ref.
extracellular PLA1sPS-PLA1PLA1APLA1APSProducing enzyme for bioactive lysophospholipid, LysoPS-[16,17]
PA-PLA1αLIPH, mPA-PLA1αLIPHPAProducing enzyme for bioactive lysophospholipid, LPA-[4,18]
lipoprotein lipaseLPL, LIPDLIPDTAG, PLTAG lipase and PLA1 activityH: 6E7K, 6OAU, 6OAZ, 6OB0, 6WN4[19]
hepatic lipase HL, LIPCLIPCTAG, PLTAG lipase and PLA1 activity-[19]
endothelial cell-derived lipaseEDL, EL, LIPGLIPGPLPredominant PLA1 activity-
pacreatic lipasePL, PNLIPPNLIPTAG, PLTAG lipase and PLA1 activityH: 1GPL, 1LPA, 1LPB, 1N8S[15,20,21]
pancreatic lipase-related protein 1PLRP1PLRP1TAG, PLTAG lipase and PLA1 activityH: 2PPL
pancreatic lipase-related protein 2PLRP2PLRP2TAG, PLTAG lipase and PLA1 activityH: 2OXE, 2PVS; R: 1BU8[15,21]
intracellular PLA1sPA-PLA1DDHD1, iPLA1αDDHD1PLPLA1 activity-[22,23]
KIAA0725pDDHD2, iPLA1γDDHD2PLPE, DAG, CL-[24,25,26]
p125iPLA1βP125n.d.Enzymatic activity has not been detected-[27]
PNPLA6iPLA2δ, NTEPNPLA6PC, LPCPLB, LysoPLA activity cleaving FAs at both sn-1 and sn-2 positions-[28,29,30]
PNPLA7iPLA2θ, NREPNPLA7PC, LPCPLB, LysoPLA activity cleaving FAs at both sn-1 and sn-2 positions-[31]
PNPLA8iPLA2γ, Group VIB PLA2PNPLA8PCPLB activity cleaving FAs at both sn-1 and sn-2 positions-[32,33]
cPLA2αPLA2G4A, Group IVA PLA2PLA2G4APLPLB activity cleaving FAs at both sn-1 and sn-2 positionsH: 1BCI, 1CJY, 1RLW[34]
cPLA2βPLA2G4B, Group IVB PLA2PLA2G4BPLPLB activity cleaving FAs at both sn-1 and sn-2 positions-[34,35]
cPLA2ζPLA2G4F, Group IVF PLA2PLA2G4FPLPLB activity cleaving FAs at both sn-1 and sn-2 positions-
PLA2G16Group XVI PLA2, PLAAT3, HRASLS3, H-Rev107PLA2G16PLPLB activity cleaving FAs at both sn-1 and sn-2 positionsH: 2KYT, 4DOT, 4FA0, 4Q95, 7C3Z, 7C41[36]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Yaginuma, S.; Kawana, H.; Aoki, J. Current Knowledge on Mammalian Phospholipase A1, Brief History, Structures, Biochemical and Pathophysiological Roles. Molecules 2022, 27, 2487. https://doi.org/10.3390/molecules27082487

AMA Style

Yaginuma S, Kawana H, Aoki J. Current Knowledge on Mammalian Phospholipase A1, Brief History, Structures, Biochemical and Pathophysiological Roles. Molecules. 2022; 27(8):2487. https://doi.org/10.3390/molecules27082487

Chicago/Turabian Style

Yaginuma, Shun, Hiroki Kawana, and Junken Aoki. 2022. "Current Knowledge on Mammalian Phospholipase A1, Brief History, Structures, Biochemical and Pathophysiological Roles" Molecules 27, no. 8: 2487. https://doi.org/10.3390/molecules27082487

Article Metrics

Back to TopTop