Next Article in Journal
C-Methylated Spermidine Derivatives: Convenient Syntheses and Antizyme-Related Effects
Previous Article in Journal
Serum Urotensin II Levels Are Elevated in Patients with Obstructive Sleep Apnea
Previous Article in Special Issue
Single-Particle Tracking of Thermomyces lanuginosus Lipase Reveals How Mutations in the Lid Region Remodel Its Diffusion
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Activation Mechanisms and Diverse Functions of Mammalian Phospholipase C

Department of Applied Biological Science, Faculty of Science and Technology, Tokyo University of Science, Chiba 278-8510, Japan
*
Author to whom correspondence should be addressed.
Biomolecules 2023, 13(6), 915; https://doi.org/10.3390/biom13060915
Submission received: 29 April 2023 / Revised: 28 May 2023 / Accepted: 29 May 2023 / Published: 31 May 2023
(This article belongs to the Special Issue Advances in Lipases and Lipases Modification)

Abstract

:
Phospholipase C (PLC) plays pivotal roles in regulating various cellular functions by metabolizing phosphatidylinositol 4,5-bisphosphate in the plasma membrane. This process generates two second messengers, inositol 1,4,5-trisphosphate and diacylglycerol, which respectively regulate the intracellular Ca2+ levels and protein kinase C activation. In mammals, six classes of typical PLC have been identified and classified based on their structure and activation mechanisms. They all share X and Y domains, which are responsible for enzymatic activity, as well as subtype-specific domains. Furthermore, in addition to typical PLC, atypical PLC with unique structures solely harboring an X domain has been recently discovered. Collectively, seven classes and 16 isozymes of mammalian PLC are known to date. Dysregulation of PLC activity has been implicated in several pathophysiological conditions, including cancer, cardiovascular diseases, and neurological disorders. Therefore, identification of new drug targets that can selectively modulate PLC activity is important. The present review focuses on the structures, activation mechanisms, and physiological functions of mammalian PLC.

1. Introduction

Phospholipase C (PLC) hydrolyzes phosphatidylinositol 4,5–bisphosphate (PI(4,5)P2) to generate two second messengers, inositol 1,4,5 triphosphate (IP3) and diacylglycerol (DAG) [1,2], enabling eukaryotic cells to perform diverse functions such as cell proliferation, differentiation, and motility by spatially and temporally activating phosphoinositide turnover. Mammals possess 13 typical PLC isozymes, which can be categorized into six classes: PLCβ (β1–β4), PLCγ (γ1 and γ2), PLCδ (δ1, δ3, and δ4), PLCε, PLCζ, and PLCη (η1 and η2) [3,4,5]. The seventh family of PLC, referred to as PLCXD, has been identified in various eukaryotic species [6]. Thus, the PLC superfamily in mammalian cells comprises 16 members, with three PLCXDs (PLCXD1, PLCXD2, and PLCXD3). While it remains unclear why there is a need for such a multitude of PLC isozymes in mammalian cells despite catalyzing the same reaction, possible reasons could include distinct regulatory mechanisms and tissue distribution for each PLC isozyme (as described in Section 2 and Section 3, respectively).
Typical PLC isozymes possess a structure characterized by several conserved domains along with class-specific domains. The active sites and catalytic residues in typical PLC isozymes are located within the triosephosphateisomerase (TIM) barrel (X and Y) domains. While PLCζ is the only exception that lacks the pleckstrin homology (PH) domain, typical PLC isozymes harbor the PH domain, EF-hand motifs, and the C2 domain along with the X and Y domains. PLCβ possesses the C-terminal domain (CTD) of approximately 400 amino acids and the PSD-95, discs large, ZO-1 (PDZ)-binding motif. PLCγ bears the multidomain insertion between the X and Y domains, comprising the split PH domain, the N-terminal Src homology 2 (nSH2) domain, the C-terminal SH2 (cSH2) domain, and the Src homology 3 (SH3) domain. PLCε harbors the Cdc25 homology domain and Ras association domains. Contrary to typical PLC isozymes, the PLCXD family is a group of enzymes that contain a catalytic domain with a sequence that is similar to the X domain (Figure 1).
Upon exposure to various stimuli, typical PLC isozymes hydrolyze plasma membrane (PM) PI(4,5)P2 to produce two second messengers: IP3 and DAG [1,2]. IP3 binds IP3 receptors present in the endoplasmic reticulum (ER), inducing the release of Ca2+ into the cytosol from ER stores, while hydrophobic DAG binds proteins, including protein kinase C (PKC), for its membrane recruitment and activation. In addition, DAG activates transient receptor potential canonical (TRPC)3, TRPC6, and TRPC7, which are members of the TRP family of nonselective cation channels [7]. These channels are permeable to Ca2+ and can increase intracellular Ca2+ concentration. IP3 is metabolized to inositol 1,3,4,5-tetrakisphosphate (IP4) via phosphorylation by inositol 1,4,5-trisphosphate 3-kinase or inositol polyphosphate multikinase (IPMK). IP4 can be further metabolized to inositol 1,3,4,5,6-pentakisphosphate (IP5) by phosphorylation at the 6-position via IPMK and then to inositol hexakisphosphate (IP6) by phosphorylation at the 2-position via inositol 1,3,4,5,6-pentakisphosphate 2-kinase. IP5 and IP6 serve as substrates for the synthesis of inositol pyrophosphates (PP-InsPs) with high-energy phosphate bonds [8,9,10,11,12]. PP-InsPs are involved in various cellular processes, including chromatin remodeling, gene expression, membrane transport, insulin secretion, growth factor/cytokine signaling, apoptosis, and dopamine release [8,13,14]. In addition, IP3 is metabolized to inositol 1,4-bisphosphate (IP2) by dephosphorylation of the inositol ring at position 5 by inositol polyphosphate 5-phosphatase. IP2 is further dephosphorylated to myo-inositol by inositol monophosphatase or inositol polyphosphate 1-phosphatase. Myo-inositol is then re-incorporated into the phosphatidylinositol (PI) synthesis cycle by binding to CDP-DAG in the ER membrane. On the other hand, DAG is phosphorylated by DAG kinases to produce phosphatidic acid (PA) [15]. The specific acyl chain composition of PI(4,5)P2, with a high enrichment of stearic acid at the sn-1 position and arachidonic acid at the sn-2 position [16], is retained in the DAG generated by PLC. DAG lipases remove stearic acid, generating endocannabinoid 2-arachidonoyl glycerol, which acts as an agonist of endocannabinoid receptors [17,18]. DAG generated by the hydrolysis of PI(4,5)P2 is recycled into PI to maintain the total pool of phosphatidylinositol phosphates. This process involves the transport of the generated DAG and/or PA from the PM to the ER, where the PI synthetic enzymes CDP-DAG synthase and PI synthetase utilize them. This cycle is spatially confined to the PM–ER contact sites, where lipid transfer proteins transport lipid intermediates between the membranes. PLC also regulates the levels of its substrate, PI(4,5)P2. PI(4,5)P2 directly regulates various cellular functions, such as cytoskeletal remodeling, cytokinesis, phagocytosis, membrane dynamics, epithelial characterization, and ion channel activity [19,20,21,22,23]. PI(4,5)P2 also acts as a precursor to phosphatidylinositol 3,4,5-triphosphate (PI(3,4,5)P3), which triggers the activation of several other proteins, including AKT. This pathway plays a crucial role in numerous signaling processes, such as cell growth and survival [24]. Therefore, PLC-mediated hydrolysis of PI(4,5)P2 may exert multiple downstream effects (Figure 2).
Besides PI(4,5)P2, PLC enzymes have been reported to hydrolyze phosphatidylinositol 4-phosphate (PI(4)P) and, to a much lesser extent, PI in vitro [25]. Notably, PLCε could hydrolyze PI(4)P at the Golgi apparatus [26]. Several isozymes of PLC also hydrolyze nuclear PI(4,5)P2. Insulin-like growth factor 1 induces the activation of nuclear PLCβ1 and PI(4,5)P2 hydrolysis, thereby increasing nuclear DAG levels and inducing PKC nuclear translocation [27,28]. PLCβ1 isozyme has two splicing variants, PI-PLCβ1a and PI-PLCβ1b, which differ in their C-terminal sequences and intracellular localization [29]. Both variants carry a nuclear localization sequence (NLS); however, PI-PLCβ1a also possesses a nuclear export sequence (NES), allowing it to localize in the cytoplasm. Conversely, PI-PLCβ1b is primarily localized to the nucleus [30,31]. PLCγ1 induces nuclear generation of DAG [32]. PLCδ1 harbors an NES and an NLS, which contribute to nuclear–cytoplasmic shuttling [33]. Nuclear import of PLCδ1 is induced by increased cytoplasmic Ca2+ concentration [34]. PLCδ4 is primarily localized to the nucleus and responsible for regulating the transition between the G1 and S phases of the cell cycle [35]. PLCδ4 knockdown in adipose-derived mesenchymal stromal cells induced cell cycle arrest, with accumulation in the G1 phase [36].

2. Regulatory Mechanisms

Classical PLC enzymes have a shared regulatory mechanism where the enzyme’s active site is masked by the negatively charged X–Y linker and remains inactive. When PLC enzymes bind to the PM, the X–Y linker is pushed away by the negatively charged surface of the membrane, allowing the active site to become accessible and removing its auto-inhibition [37].

2.1. Regulatory Mechanisms of PLCβ

PLCβ isozymes act as downstream effectors of G protein-coupled receptors (GPCRs) and can be activated by either the Gαq family or Gβγ subunits [38,39]. The PH domain of PLCβ is involved in the activation of the enzyme by Gβγ and Rac [40,41]. Rac and Gβγ interact with the PH domain of PLCβ to optimize its orientation for substrate membranes [40]. PLCβ contains a CTD composed of approximately 400 amino acids, which bind to its catalytic core and inhibit enzymatic activity under resting conditions [42,43]. The CTD of PLCβ1 increases the curvature of the PM, thereby promoting efficient cleavage of PI(4,5)P2, which is present in highly curved membranes [44]. The activation of PLCβ by Gαq also requires the presence of a CTD. The PDZ-binding motif of PLCβ may facilitate selective binding to GPCRs via the PDZ scaffold proteins [45]. Furthermore, PLCβ functions as a GTPase-activating protein for Gαq in addition to its lipase activity [46]. Thus, PLCβ isozymes are activated by Gαq, Gβγ, and small GTPases of the Rho family, such as Rac (Figure 3).

2.2. Regulatory Mechanisms of PLCγ

PLCγ isoforms are regulated by both receptor tyrosine kinases (RTKs) and non-RTKs (Figure 3) [47,48,49,50,51]. Activation of PLCγ occurs via the binding of its nSH2 domain to phosphorylated tyrosine residues of RTKs, which induces the phosphorylation of a conserved tyrosine residue (Tyr783 in human PLCγ1 and Tyr759 in human PLCγ2) by RTKs [52]. The cSH2 domain inhibits PLCγ by interacting with residues around its catalytically active site under resting conditions. Phosphorylation of the conserved tyrosine residue removes the cSH2 domain from the active site via interaction with the cSH2 domain, allowing the binding of the active site of PLCγ to its substrate, PI(4,5)P2 [53,54]. Therefore, the PLCγ SH2 domain plays an essential role in RTK- and non-RTK-mediated PLCγ activation. PLCγ2, but not PLCγ1, interacts with Rac via the split PH domain, resulting in its recruitment to the PM and activation [55,56]. PI(3,4,5)P3 also recruits PLCγ isoforms to the PM and activates them [57,58,59]. Thus, the multidomain insertion located between the X and Y domains of PLCγ is essential for regulating its activity.

2.3. Regulatory Mechanisms of PLCδ

PLCδ activity can be stimulated by micromolar levels of Ca2+ within the physiological range through the activation of the other PLC isozymes or influx of Ca2+ through calcium channels (Figure 3) [60,61]. Ca2+ induces the translocation of PLCδ from the cytoplasm to the PM where it is activated. Therefore, PLCδ is thought to amplify elevated Ca2+ levels to concentrations sufficient for inducing downstream signaling. The PH domain also plays a critical role in activation of PLCδ. The PH domain of PLCδ binds specifically and with high affinity to PI(4,5)P2 [62,63], playing a crucial role in both the recruitment and activation of PLCδ on the PM. In vitro studies have suggested that the PH domain of PLCδ1 has a higher affinity for IP3 than for PI(4,5)P2 [64]. Since increased cytosolic IP3 levels inhibit the binding of PLCδ1 to PM PI(4,5)P2 [65], this may function as a negative feedback mechanism. Two putative positive regulators of PLCδ1, transglutaminase II and Ral, have been also identified [66,67].

2.4. Regulatory Mechanisms of PLCε

PLCε can be activated by GPCRs and RTKs, as well as by small GTPases (Figure 3) [68,69,70]. Binding to GTP-bound Rap and Ras results in differential localization of PLCε [68,71,72,73]. Ras-activating mutations and stimuli lead to PM localization of PLCε, whereas Rap activation results in its recruitment to the perinuclear region [74]. RhoA binds to PLCε through a specific region of the Y domain, resulting in its activation [75]. The Cdc25 homology domain functions as a guanine nucleotide exchange factor (GEF) for Ras and Rap1 [74,76]. The GEF activity of the Cdc25 homology domain for Rap1 can augment the lipase activity of PLCε, as activated Rap1 can stimulate PLCε. Thus, PLCε activity is regulated by various downstream signaling pathways.

2.5. Regulatory Mechanisms of PLCζ, PLCη, and PLCXD

PLCζ is activated by low concentrations of Ca2+, similar to the resting cytoplasmic Ca2+ concentration (Figure 3). Unlike other PLC isozymes, the X–Y linker of PLCζ exhibits distinct electrostatic features and is positively charged, which may enable it to bind to the PM or associate with the anionic substrate lipid PI(4,5)P2. Therefore, PLCζ is constitutively active [77]. The interaction of PLCζ with PI(4,5)P2 in membranes requires EF hands and the X–Y linker region, whereas its activity relies on the C2 domain [78,79].
PLCη is highly sensitive to Ca2+ and responds to elevated intracellular Ca2+ levels [80,81]. Since Gβγ also activates PLCη2, it may be activated upon GPCR stimulation (Figure 3) [82,83].
The regulatory mechanisms of PLCXDs remain unclear.

3. Physiological Functions of PLC

3.1. PLCβ

There are four isozymes of PLCβ (β1–β4), which are predominantly expressed in the brain and play essential roles in maintaining normal brain function. Several isozymes of PLCβ also play significant roles in blood cell types. PLCβ1-deficient mice experienced epileptic seizures due to impaired inhibitory neuronal circuitry; this was attributed to attenuated PKC activity, which leads to a deficit in GABAergic inhibition [84]. Similarly, human patients with PLCβ1 loss suffered from infantile epileptic encephalopathy [85,86]. In addition, PLCβ1 is crucial for glucose-stimulated insulin release in β-cells. Mice with conditional knockout (KO) of islet-expressed PLCβ1 displayed glucose intolerance, which is consistent with the observed in vitro defect [87,88]. PLCβ1 expression decreased in a malignancy-dependent manner in gliomas, and the level of PLCβ1 expression was significantly correlated with the survival rate [89]. PLCβ2 deficiency was found to inhibit Ca2+ release and superoxide production induced by chemoattractants in neutrophils of mice while paradoxically enhancing chemotactic activity via an unknown mechanism [90,91]. PLCβ2 also plays a central role in taste receptor signaling and is activated by βγ subunits released by various GPCRs [92,93,94]. In addition, PLCβ2 negatively regulates virus-induced pro-inflammatory responses by hydrolyzing PI(4,5)P2 and inhibiting PI(4,5)P2-mediated TGF-β-activated kinase 1 activation [95]. Loss of PLCβ3 inhibited the Src homology region 2 domain-containing phosphatase (SHP)-mediated suppression of Lyn, resulting in defective Fc epsilon Receptor I (FcεRI) signaling and mast cell-dependent immune responses in mice [96]. Loss of PLCβ3 impaired the formation of the signal transducer and activator of transcription (STAT)5–SHP-1–PLCβ3 protein complex, leading to STAT5 hyperactivation, mast cell hyperproliferation, and atopic dermatitis-like skin inflammation [97]. Interestingly, the lipase activity of PLCβ3 is not required for STAT5 regulation. In hematopoietic stem cells (HSCs), loss of PLCβ3 led to STAT5 hyperactivation, thereby increasing the number of HSCs with a myeloid differentiation ability and leading to the development of myeloproliferative neoplasms in PLCβ3-KO mice [98]. Mice lacking PLCβ3 exhibited increased sensitivity to apoptotic induction in their macrophages, which resulted in reduced atherosclerotic lesion size [99]. In humans, PLCβ3 mutations have been shown to either protect against cystic fibrosis or cause autosomal recessive spondylometaphyseal dysplasia [100,101,102]. Loss of PLCβ4 induced a range of phenotypic defects in mice, including impaired cerebellar development, which led to ataxia [103] and visual processing deficits [104]. PLCβ4 KO mice also exhibit absence seizures [105]. Studies involving human patients have shown that PLCB4 mutations are linked to the development of uveal melanomas, which are the most common type of eye tumors arising from melanocytes of the uveal tract [106]. Loss-of-function mutations in PLCB4 have also been implicated in auriculocondylar syndrome [107].

3.2. PLCγ

There are two isozymes of PLCγ (γ1 and γ2). PLCγ isozymes play essential roles in hematopoietic cell development and functions. Functional loss of PLCγ1 resulted in defective vasculogenesis and erythrogenesis, and PLCγ1-deficient mice died on embryonic day 9 [108,109]. Moreover, PLCγ1 is crucial for T-cell receptor (TCR) signaling, which is required for T-cell activation, development, and homeostasis. T-cell-specific deletion of PLCγ1 impaired the development of regulatory T cells [110]. PLCγ1 is also involved in the development of HSCs, as PLCγ1-KO cells failed to differentiate into hematopoietic cells in PLCγ1-KO chimeric mice [111]. Additionally, PLCγ1 has been implicated in various cancers in a number of studies, and these studies have highlighted the role of PLCγ1 in tumor progression and metastases [112,113,114,115,116,117,118]. Somatic mutations in PLCG1 have been reported in angiosarcoma [119]. PLCγ1 mutant plays a role in angiosarcoma by promoting invasiveness and influencing angiogenesis through vascular endothelial growth factor (VEGF) signaling [119,120,121,122]. PLCG1 mutations were also discovered in T-cell lymphomas, including cutaneous and adult T-cell leukemia/lymphoma. PLCG1 is the most commonly mutated gene in adult T-cell leukaemia/lymphoma, accounting for approximately 40% of all cases. Mutant forms of this isozyme are thought to contribute to the development of cancer by promoting phospholipase activity and subsequently enhancing nuclear factor of activated T-cells (NFAT)- and NF-κB-dependent transcription [123,124]. Mutations in the TCR signaling components and PLCG1 have been observed in T-cell lymphoma patients, and these mutations are associated with poorer overall and progression-free survival rates based on several clinical studies [125,126,127]. In contrast, some studies have indicated that reduced PLCγ1 expression is conducive to cancer cell survival and proliferation. For instance, PLCγ1 expression is downregulated during hypoxia in KRAS-mutant human lung adenocarcinoma cell lines, preventing lipid peroxidation, inhibiting apoptosis, and enhancing cancer cell proliferation [128]. Mice with a specific PLCγ1 KO in neuronal precursors exhibited deficiencies in midbrain axon guidance, resulting in structural alterations to the mesencephalic dopaminergic system, wherein axons fail to project to their appropriate locations [129,130,131]. Forebrain-selective PLCγ1 KO resulted in behavioral abnormalities such as hyperactivity [132]. Activation of PLCγ1 is triggered by the activation of tropomyosin-related kinase B (TrkB) receptors through binding to brain-derived neurotrophic factor (BDNF), which plays an essential role in the formation and function of inhibitory synapses that use gamma-aminobutyric acid (GABA) as a neurotransmitter. Selective PLCγ1 KO in inhibitory GABAergic neurons increased seizure susceptibility in aged mice [133]. In contrast, in a temporal lobe epilepsy model, hyperexcitation of excitatory neurons triggered the activation of cellular signaling pathways, including elevated phosphorylation of PLCγ1 via the BDNF-TrkB pathway. Uncoupling of the BDNF receptor TrkB from PLCγ1 prevented epilepsy, suggesting that the effects of PLCγ1 on epilepsy depend on the specific neuronal population involved [134]. PLCγ2 is a critical signaling effector of the pre-B-cell receptor and essential for B-cell development and maturation. PLCγ2-KO mice showed impaired B-cell maturation [135], whereas a gain-of-function mutation of PLCγ2 generated via ENU mutagenesis resulted in the hyperactivation of B-cells and innate immune cells [136]. Furthermore, PLCγ2 plays a role in the regulation of innate immune cells and platelets through the signaling of Fc receptors [137]. PLCγ2 deficiency also impaired receptor activator of NF-κB ligand (RANKL) signaling in hematopoietic cells, leading to defects in lymph node organogenesis and osteoclast differentiation [138]. Gain-of-function mutations in PLCG2 have been linked to a disorder called PLCγ2-associated antibody deficiency and immune dysregulation (PLAID), which is characterized by cold urticaria due to the spontaneous activation of mast cells expressing the mutant form of PLCγ2 when exposed to lower temperatures [139]. In addition, gain-of-function mutations in PLCG2 have been implicated in a complex immune disorder called autoinflammation, antibody deficiency, and immune dysregulation, which are predominantly inherited and resemble PLAID [140,141,142].

3.3. PLCδ

There are three isozymes of PLCδ (δ1, δ3, and δ4), which play critical roles in the normal function of the skin, osomosensitive neurons, placenta, heart, and sperm. Mice lacking PLCδ1 displayed sparse hair owing to an abnormal hair shaft structure and reduced hair keratin expression [143,144]. In addition, PLCδ1 plays a critical role in nail formation, as demonstrated by mutations in patients with hereditary leukonychia [145,146,147,148]. Furthermore, PLCδ1 is involved in the regulation of inflammatory skin diseases, such as psoriasis and contact hypersensitivity (CHS) [149,150]. PLCδ1 also plays a key role in the activation of deltaN-TRPV1 channels and osmosensory transduction in magnocellular neurosecretory cells [151,152]. Epigenetic silencing of PLCD1 has been observed in several cancers, suggesting its potential tumor-suppressive role [153,154,155,156]. Simultaneous loss of PLCδ1 and PLCδ3 in mice led to embryonic lethality due to decreased placental vascularization and excessive apoptosis of placental trophoblasts [157]. PLCδ1/PLCδ3 double-KO mice also exhibited impaired cardiac function, fibrosis, and spontaneous cardiac hypertrophy, possibly caused by excessive apoptosis of cardiomyocytes [158]. Male infertility is observed in mice lacking PLCδ4 due of their inability to initiate the acrosome reaction, which is essential for sperm penetration into the zona pellucida and fusion with the egg PM [159,160].

3.4. PLCε

There is only one isozyme of PLCε. Consistent with its high expression in cardiac tissues, PLCε plays a critical role in the regulation of cardiomyocyte development and function. Increased PLCε transcript levels were observed in the myocardial tissues of patients with idiopathic dilated cardiomyopathy, suggesting the potential involvement of PLCε in the pathogenesis of human cardiac diseases [161]. Studies on cardiomyocyte-specific PLCε-KO mice have demonstrated protection against pressure overload-induced hypertrophy. Mechanistically, PLCε catalyzes the hydrolysis of the noncanonical substrate PI(4)P in the perinuclear Golgi apparatus to generate DAG in cardiomyocytes. Subsequently, DAG activates the hypertrophic kinase protein kinase D [26,162]. PLCε also participates in cardiac development, as shown in mice lacking catalytically active PLCε displaying impaired cardiac semilunar valvulogenesis [163]. Furthermore, PLCε has been implicated in skin inflammation. PLCε overexpression in keratinocytes induced psoriasis-like skin inflammation [164], whereas lack of PLCε attenuated CHS [165]. Hence, PLCε appears to positively regulate skin inflammation. PLCε also plays a positive role in neuroinflammation [166]. Mutations in the X domain of PLCE1 in humans can lead to nephrotic syndrome, characterized by proteinuria due to disruption of the glomerular filtration barrier executed by podocytes [167,168].

3.5. PLCζ

PLCζ, which is specifically expressed in the sperm, plays a pivotal role in fertilization. It is a key molecule derived from sperm that induces Ca2+ oscillation, which is a crucial process for egg activation during fertilization [169]. Studies demonstrated that PLCζ downregulation in mouse sperm impaired Ca2+ oscillations and egg activation [170,171]. Conversely, broad ectopic PLCζ expression led to autonomous Ca2+ oscillations in unfertilized oocytes, resulting in egg activation and parthenogenetic development, highlighting the direct effect of PLCζ, which is analogous to fertilization [172]. In humans, loss-of-function mutations in many PLCζ variants found in patients have been identified and linked to the failure of oocyte activation, which is regulated by Ca2+ oscillations [173,174,175].

4. Chemical Inhibitors and Activators for PLC

There are several compounds that are known to modulate the activity of PLC. U73122 is a commonly used inhibitor of PLC, although it has been reported to affect other targets, such as ion channels, calcium pumps, and enzymes [176,177,178]. Similarly, m-3M3FBS is a commonly used pan-PLC activator; however, it interacts with unrelated targets in cells and there is no clear evidence that it directly binds PLC [179]. Thus, currently, no fully validated small-molecule inhibitors or activators of PLC suitable for research applications are available. This limitation is largely due to the lack of a powerful high-throughput screening system and difficulties associated with generating chemical probes based on the PLC substrate, PI(4,5)P2. Recent advances have been achieved to overcome these challenges. Although the half-life of IP3, a direct product of PLC, is short, the downstream metabolite IP1 can be stabilized by introducing lithium chloride (LiCl). Therefore, PLC activity can be evaluated by measuring IP1 accumulation in the presence of LiCl [180]. Furthermore, there is potential for in vitro assays utilizing PLC, as demonstrated by the use of fluorescently tagged PI(4,5)P2 analogs such as WH-15, which can be hydrolyzed by PLC isozymes to produce a fluorescent molecule [181]. A related compound, XY-69, has also been synthesized and used in vitro [182]. Recent advances in the development of high-throughput screening systems are expected to facilitate the identification of specific PLC inhibitors and activators.

5. Perspectives

PLC exerts its physiological functions primarily through generation of the second messengers IP3 and DAG. However, considering the involvement of PI(4,5)P2 in the regulation of diverse cellular functions, the reduction in PI(4,5)P2 levels caused by PLC is highly likely to play a role in its physiological functions. Further investigations are warranted to determine the impact of PLC-mediated PI(4,5)P2 metabolism on PI(4,5)P2 levels in various cellular contexts. Besides enzymatic activity, some PLC isozymes have multifunctional roles. For instance, PLCβ1 regulates caveolar invasion and membrane curvature in a lipase-independent manner. Future studies should explore the lipase-independent functions of PLC to elucidate their novel roles. Moreover, since the structure of PLCXD is distinct from that of typical PLC, investigation concerning its substrate specificity and activation mechanism would be intriguing. In addition, specific PLC isozymes that play critical roles in certain organs may serve as viable targets for the development of novel drugs. Structural data on PLC isozymes and the availability of fluorescent substrates can allow for the screening of specific PLC activators and inhibitors as potential drug candidates.

Author Contributions

Conceptualization, writing—original draft preparation, review, and editing, K.K. and Y.N.; visualization, K.K.; supervision, Y.N.; funding acquisition, K.K. and Y.N. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by Grant-in-Aid for Scientific Research (B) 23H03341, the Takeda Science Foundation, Kose Cosmetology Research Foundation, which was granted to Y.N., as well as Grant-in-Aid for Young Scientists 21K15109, Grant-in-Aid for Scientific Research (C) 23K06103, a Kishimoto Fund Research Grant, the ONO Medical Research Foundation, and the Kato Memorial Bioscience Foundation, which supported K.K.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Berridge, M.J.; Irvine, R.F. Inositol trisphosphate, a novel second messenger in cellular signal transduction. Nature 1984, 312, 315–321. [Google Scholar] [CrossRef] [PubMed]
  2. Nishizuka, Y. The molecular heterogeneity of protein kinase C and its implications for cellular regulation. Nature 1988, 334, 661–665. [Google Scholar] [CrossRef] [PubMed]
  3. Katan, M.; Cockcroft, S. Phospholipase C families: Common themes and versatility in physiology and pathology. Prog. Lipid Res. 2020, 80, 101065. [Google Scholar] [PubMed]
  4. Suh, P.G.; Park, J.I.; Manzoli, L.; Cocco, L.; Peak, J.C.; Katan, M.; Fukami, K.; Kataoka, T.; Yun, S.; Ryu, S.H. Multiple roles of phosphoinositide-specific phospholipase C isozymes. BMB Rep. 2008, 41, 415–434. [Google Scholar] [CrossRef]
  5. Nakamura, Y.; Fukami, K. Regulation and physiological functions of mammalian phospholipase C. J. Biochem. 2017, 161, 315–321. [Google Scholar] [CrossRef]
  6. Gellatly, S.A.; Kalujnaia, S.; Cramb, G. Cloning, tissue distribution and sub-cellular localisation of phospholipase C X-domain containing protein (PLCXD) isoforms. Biochem. Biophys. Res. Commun. 2012, 424, 651–656. [Google Scholar] [CrossRef]
  7. Hofmann, T.; Obukhov, A.G.; Schaefer, M.; Harteneck, C.; Gudermann, T.; Schultz, G. Direct activation of human TRPC6 and TRPC3 channels by diacylglycerol. Nature 1999, 397, 259–263. [Google Scholar] [CrossRef]
  8. Chakraborty, A.; Kim, S.; Snyder, S.H. Inositol pyrophosphates as mammalian cell signals. Sci. Signal. 2011, 4, re1. [Google Scholar] [CrossRef]
  9. Irvine, R.F.; Schell, M.J. Back in the water: The return of the inositol phosphates. Nat. Rev. Mol. Cell. Biol. 2001, 2, 327–338. [Google Scholar] [CrossRef]
  10. Laha, D.; Portela-Torres, P.; Desfougères, Y.; Saiardi, A. Inositol phosphate kinases in the eukaryote landscape. Adv. Biol. Regul. 2021, 79, 100782. [Google Scholar] [CrossRef]
  11. Lee, J.Y.; Kim, Y.R.; Park, J.; Kim, S. Inositol polyphosphate multikinase signaling in the regulation of metabolism. Ann. N. Y. Acad. Sci. 2012, 1271, 68–74. [Google Scholar] [CrossRef] [PubMed]
  12. Mulugu, S.; Bai, W.; Fridy, P.C.; Bastidas, R.J.; Otto, J.C.; Dollins, D.E.; Haystead, T.A.; Ribeiro, A.A.; York, J.D. A conserved family of enzymes that phosphorylate inositol hexakisphosphate. Science 2007, 316, 106–109. [Google Scholar] [CrossRef] [PubMed]
  13. Lee, Y.S.; Mulugu, S.; York, J.D.; O’Shea, E.K. Regulation of a cyclin-CDK-CDK inhibitor complex by inositol pyrophosphates. Science 2007, 316, 109–112. [Google Scholar] [CrossRef]
  14. Monserrate, J.P.; York, J.D. Inositol phosphate synthesis and the nuclear processes they affect. Curr. Opin. Cell. Biol. 2010, 22, 365–373. [Google Scholar] [CrossRef]
  15. Thakur, R.; Naik, A.; Panda, A.; Raghu, P. Regulation of membrane turnover by phosphatidic Acid: Cellular Functions and Disease Implications. Front. Cell. Dev. Biol. 2019, 7, 83. [Google Scholar] [CrossRef]
  16. Barneda, D.; Cosulich, S.; Stephens, L.; Hawkins, P. How is the acyl chain composition of phosphoinositides created and does it matter? Biochem. Soc. Trans. 2019, 47, 1291–1305. [Google Scholar] [CrossRef]
  17. Murataeva, N.; Straiker, A.; Mackie, K. Parsing the players: 2-arachidonoylglycerol synthesis and degradation in the CNS. Br. J. Pharmacol. 2014, 171, 1379–1391. [Google Scholar] [CrossRef]
  18. Tong, J.; Liu, X.; Vickstrom, C.; Li, Y.; Yu, L.; Lu, Y.; Smrcka, A.V.; Liu, Q.S. The Epac-Phospholipase Cε Pathway Regulates Endocannabinoid Signaling and Cocaine-Induced Disinhibition of Ventral Tegmental Area Dopamine Neurons. J. Neurosci. 2017, 37, 3030–3044. [Google Scholar] [CrossRef] [PubMed]
  19. Di Paolo, G.; De Camilli, P. Phosphoinositides in cell regulation and membrane dynamics. Nature 2006, 443, 651–657. [Google Scholar] [CrossRef]
  20. Janetopoulos, C.; Devreotes, P. Phosphoinositide signaling plays a key role in cytokinesis. J. Cell. Biol. 2006, 174, 485–490. [Google Scholar] [CrossRef]
  21. Martin, T.F. PI(4,5)P(2) regulation of surface membrane traffic. Curr. Opin. Cell. Biol. 2001, 13, 493–499. [Google Scholar] [CrossRef]
  22. Senju, Y.; Lappalainen, P. Regulation of actin dynamics by PI(4,5)P2 in cell migration and endocytosis. Curr. Opin. Cell. Biol. 2019, 56, 7–13. [Google Scholar] [CrossRef]
  23. Kanemaru, K.; Shimozawa, M.; Kitamata, M.; Furuishi, R.; Kayano, H.; Sukawa, Y.; Chiba, Y.; Fukuyama, T.; Hasegawa, J.; Nakanishi, H.; et al. Plasma membrane phosphatidylinositol (4,5)-bisphosphate is critical for determination of epithelial characteristics. Nat. Commun. 2022, 13, 2347. [Google Scholar] [CrossRef] [PubMed]
  24. Manning, B.D.; Cantley, L.C. AKT/PKB signaling: Navigating downstream. Cell 2007, 129, 1261–1274. [Google Scholar] [CrossRef] [PubMed]
  25. Ellis, M.V.; James, S.R.; Perisic, O.; Downes, C.P.; Williams, R.L.; Katan, M. Catalytic domain of phosphoinositide-specific phospholipase C (PLC). Mutational analysis of residues within the active site and hydrophobic ridge of plcdelta1. J. Biol. Chem. 1998, 273, 11650–11659. [Google Scholar] [CrossRef]
  26. Zhang, L.; Malik, S.; Pang, J.; Wang, H.; Park, K.M.; Yule, D.I.; Blaxall, B.C.; Smrcka, A.V. Phospholipase Cε hydrolyzes perinuclear phosphatidylinositol 4-phosphate to regulate cardiac hypertrophy. Cell 2013, 153, 216–227. [Google Scholar] [CrossRef]
  27. Divecha, N.; Banfić, H.; Irvine, R.F. The polyphosphoinositide cycle exists in the nuclei of Swiss 3T3 cells under the control of a receptor (for IGF-I) in the plasma membrane, and stimulation of the cycle increases nuclear diacylglycerol and apparently induces translocation of protein kinase C to the nucleus. EMBO J. 1991, 10, 3207–3214. [Google Scholar]
  28. Manzoli, L.; Billi, A.M.; Rubbini, S.; Bavelloni, A.; Faenza, I.; Gilmour, R.S.; Rhee, S.G.; Cocco, L. Essential role for nuclear phospholipase C beta1 in insulin-like growth factor I-induced mitogenesis. Cancer Res. 1997, 57, 2137–2139. [Google Scholar] [PubMed]
  29. Bahk, Y.Y.; Song, H.; Baek, S.H.; Park, B.Y.; Kim, H.; Ryu, S.H.; Suh, P.G. Localization of two forms of phospholipase C-beta1, a and b, in C6Bu-1 cells. Biochim. Biophys. Acta 1998, 1389, 76–80. [Google Scholar] [CrossRef]
  30. Martelli, A.M.; Fiume, R.; Faenza, I.; Tabellini, G.; Evangelista, C.; Bortul, R.; Follo, M.Y.; Falà, F.; Cocco, L. Nuclear phosphoinositide specific phospholipase C (PI-PLC)-beta 1: A central intermediary in nuclear lipid-dependent signal transduction. Histol. Histopathol. 2005, 20, 1251–1260. [Google Scholar]
  31. Martelli, A.M.; Follo, M.Y.; Evangelisti, C.; Falà, F.; Fiume, R.; Billi, A.M.; Cocco, L. Nuclear inositol lipid metabolism: More than just second messenger generation? J. Cell. Biochem. 2005, 96, 285–292. [Google Scholar] [CrossRef] [PubMed]
  32. Klein, C.; Gensburger, C.; Freyermuth, S.; Nair, B.C.; Labourdette, G.; Malviya, A.N. A 120 kDa nuclear phospholipase Cgamma1 protein fragment is stimulated in vivo by EGF signal phosphorylating nuclear membrane EGFR. Biochemistry 2004, 43, 15873–15883. [Google Scholar] [CrossRef] [PubMed]
  33. Yagisawa, H.; Okada, M.; Naito, Y.; Sasaki, K.; Yamaga, M.; Fujii, M. Coordinated intracellular translocation of phosphoinositide-specific phospholipase C-delta with the cell cycle. Biochim. Biophys. Acta 2006, 1761, 522–534. [Google Scholar] [CrossRef]
  34. Okada, M.; Ishimoto, T.; Naito, Y.; Hirata, H.; Yagisawa, H. Phospholipase Cdelta1 associates with importin beta1 and translocates into the nucleus in a Ca2+-dependent manner. FEBS Lett. 2005, 579, 4949–4954. [Google Scholar] [CrossRef] [PubMed]
  35. Liu, N.; Fukami, K.; Yu, H.; Takenawa, T. A new phospholipase C delta 4 is induced at S-phase of the cell cycle and appears in the nucleus. J. Biol. Chem. 1996, 271, 355–360. [Google Scholar] [CrossRef]
  36. Kunrath-Lima, M.; de Miranda, M.C.; Ferreira, A.D.F.; Faraco, C.C.F.; de Melo, M.I.A.; Goes, A.M.; Rodrigues, M.A.; Faria, J.A.Q.A.; Gomes, D.A. Phospholipase C delta 4 (PLCδ4) is a nuclear protein involved in cell proliferation and senescence in mesenchymal stromal stem cells. Cell. Signal. 2018, 49, 59–67. [Google Scholar] [CrossRef]
  37. Hicks, S.N.; Jezyk, M.R.; Gershburg, S.; Seifert, J.P.; Harden, T.K.; Sondek, J. General and versatile autoinhibition of PLC isozymes. Mol. Cell. 2008, 31, 383–394. [Google Scholar] [CrossRef]
  38. Philip, F.; Kadamur, G.; Silos, R.G.; Woodson, J.; Ross, E.M. Synergistic activation of phospholipase C-beta3 by Galpha(q) and Gbetagamma describes a simple two-state coincidence detector. Curr. Biol. 2010, 20, 1327–1335. [Google Scholar] [CrossRef]
  39. Smrcka, A.V.; Sternweis, P.C. Regulation of purified subtypes of phosphatidylinositol-specific phospholipase C beta by G protein alpha and beta gamma subunits. J. Biol. Chem. 1993, 268, 9667–9674. [Google Scholar] [CrossRef]
  40. Jezyk, M.R.; Snyder, J.T.; Gershberg, S.; Worthylake, D.K.; Harden, T.K.; Sondek, J. Crystal structure of Rac1 bound to its effector phospholipase C-beta2. Nat. Struct. Mol. Biol. 2006, 13, 1135–1140. [Google Scholar] [CrossRef]
  41. Illenberger, D.; Walliser, C.; Nurnberg, B.; Diaz Lorente, M.; Gierschik, P. Specificity and structural requirements of phospholipase C-beta stimulation by Rho GTPases versus G protein beta gamma dimers. J. Biol. Chem. 2003, 278, 3006–3014. [Google Scholar] [CrossRef]
  42. Lyon, A.M.; Dutta, S.; Boguth, C.A.; Skiniotis, G.; Tesmer, J.J. Full-length Gα(q)-phospholipase C-β3 structure reveals interfaces of the C-terminal coiled-coil domain. Nat. Struct. Mol. Biol. 2013, 20, 355–362. [Google Scholar] [CrossRef] [PubMed]
  43. Fisher, I.J.; Jenkins, M.L.; Tall, G.G.; Burke, J.E.; Smrcka, A.V. Activation of Phospholipase C β by Gβγ and Gαq Involves C-Terminal Rearrangement to Release Autoinhibition. Structure 2020, 28, 810–819.e5. [Google Scholar] [CrossRef] [PubMed]
  44. Inaba, T.; Kishimoto, T.; Murate, M.; Tajima, T.; Sakai, S.; Abe, M.; Makino, A.; Tomishige, N.; Ishitsuka, R.; Ikeda, Y.; et al. Phospholipase Cβ1 induces membrane tubulation and is involved in caveolae formation. Proc. Natl. Acad. Sci. USA 2016, 113, 7834–7839. [Google Scholar] [CrossRef]
  45. Oh, Y.S.; Jo, N.W.; Choi, J.W.; Kim, H.S.; Seo, S.W.; Kang, K.O.; Hwang, J.I.; Heo, K.; Kim, S.H.; Kim, Y.H.; et al. NHERF2 specifically interacts with LPA2 receptor and defines the specificity and efficiency of receptor-mediated phospholipase C-beta3 activation. Mol. Cell. Biol. 2004, 24, 5069–5079. [Google Scholar] [CrossRef]
  46. Berstein, G.; Blank, J.L.; Jhon, D.Y.; Exton, J.H.; Rhee, S.G.; Ross, E.M. Phospholipase C-beta 1 is a GTPase-activating protein for Gq/11, its physiologic regulator. Cell 1992, 70, 411–418. [Google Scholar] [CrossRef] [PubMed]
  47. Kim, H.K.; Kim, J.W.; Zilberstein, A.; Margolis, B.; Kim, J.G.; Schlessinger, J.; Rhee, S.G. PDGF stimulation of inositol phospholipid hydrolysis requires PLC-gamma 1 phosphorylation on tyrosine residues 783 and 1254. Cell 1991, 65, 435–441. [Google Scholar] [CrossRef] [PubMed]
  48. Wahl, M.I.; Daniel, T.O.; Carpenter, G. Antiphosphotyrosine recovery of phospholipase C activity after EGF treatment of A-431 cells. Science 1988, 241, 968–970. [Google Scholar] [CrossRef]
  49. Law, C.L.; Chandran, K.A.; Sidorenko, S.P.; Clark, E.A. Phospholipase C-gamma1 interacts with conserved phosphotyrosyl residues in the linker region of Syk and is a substrate for Syk. Mol. Cell. Biol. 1996, 16, 1305–1315. [Google Scholar] [CrossRef]
  50. Nakanishi, O.; Shibasaki, F.; Hidaka, M.; Homma, Y.; Takenawa, T. Phospholipase C-gamma 1 associates with viral and cellular src kinases. J. Biol. Chem. 1993, 268, 10754–10759. [Google Scholar] [CrossRef]
  51. Schaeffer, E.M.; Debnath, J.; Yap, G.; McVicar, D.; Liao, X.C.; Littman, D.R.; Sher, A.; Varmus, H.E.; Lenardo, M.J.; Schwartzberg, P.L. Requirement for Tec kinases Rlk and Itk in T cell receptor signaling and immunity. Science 1999, 284, 638–641. [Google Scholar] [CrossRef]
  52. Bae, J.H.; Lew, E.D.; Yuzawa, S.; Tomé, F.; Lax, I.; Schlessinger, J. The selectivity of receptor tyrosine kinase signaling is controlled by a secondary SH2 domain binding site. Cell 2009, 138, 514–524. [Google Scholar] [CrossRef] [PubMed]
  53. Hajicek, N.; Keith, N.C.; Siraliev-Perez, E.; Temple, B.R.; Huang, W.; Zhang, Q.; Harden, T.K.; Sondek, J. Structural basis for the activation of PLC-γ isozymes by phosphorylation and cancer-associated mutations. eLife 2019, 8, e51700. [Google Scholar] [CrossRef] [PubMed]
  54. Liu, Y.; Bunney, T.D.; Khosa, S.; Macé, K.; Beckenbauer, K.; Askwith, T.; Maslen, S.; Stubbs, C.; de Oliveira, T.M.; Sader, K.; et al. Structural insights and activating mutations in diverse pathologies define mechanisms of deregulation for phospholipase C gamma enzymes. EBioMedicine 2020, 51, 102607. [Google Scholar] [CrossRef] [PubMed]
  55. Piechulek, T.; Rehlen, T.; Walliser, C.; Vatter, P.; Moepps, B.; Gierschik, P. Isozyme-specific stimulation of phospholipase C-gamma2 by Rac GTPases. J. Biol. Chem. 2005, 280, 38923–38931. [Google Scholar] [CrossRef]
  56. Walliser, C.; Retlich, M.; Harris, R.; Everett, K.L.; Josephs, M.B.; Vatter, P.; Esposito, D.; Driscoll, P.C.; Katan, M.; Gierschik, P.; et al. rac regulates its effector phospholipase Cgamma2 through interaction with a split pleckstrin homology domain. J. Biol. Chem. 2008, 283, 30351–30362. [Google Scholar] [CrossRef]
  57. Bae, Y.S.; Cantley, L.G.; Chen, C.S.; Kim, S.R.; Kwon, K.S.; Rhee, S.G. Activation of phospholipase C-gamma by phosphatidylinositol 3,4,5-trisphosphate. J. Biol. Chem. 1998, 273, 4465–4469. [Google Scholar] [CrossRef]
  58. Falasca, M.; Logan, S.K.L.; Lehto, V.P.; Baccante, G.; Lemmon, M.A.; Schlessinger, J. Activation of phospholipase C gamma by PI 3-kinase-induced PH domain-mediated membrane targeting. EMBO J. 1998, 17, 414–422. [Google Scholar] [CrossRef]
  59. Gratacap, M.P.; Payrastre, B.; Viala, C.; Mauco, G.; Plantavid, M.; Chap, H. Phosphatidylinositol 3,4,5-trisphosphate-dependent stimulation of phospholipase C-gamma2 is an early key event in FcgammaRIIA-mediated activation of human platelets. J. Biol. Chem. 1998, 273, 24314–24321. [Google Scholar] [CrossRef] [PubMed]
  60. Allen, V.; Swigart, P.; Cheung, R.; Cockcroft, S.; Katan, M. Regulation of inositol lipid-specific phospholipase cdelta by changes in Ca2+ ion concentrations. Biochem. J. 1997, 327, 545–552. [Google Scholar] [CrossRef]
  61. Kim, Y.H.; Park, T.J.; Lee, Y.H.; Baek, K.J.; Suh, P.G.; Ryu, S.H.; Kim, K.T. Phospholipase C-delta1 is activated by capacitative calcium entry that follows phospholipase C-beta activation upon bradykinin stimulation. J. Biol. Chem. 1999, 274, 26127–26134. [Google Scholar] [CrossRef] [PubMed]
  62. Ferguson, K.M.; Lemmon, M.A.; Schlessinger, J.; Sigler, P.B. Structure of the high affinity complex of inositol trisphosphate with a phospholipase C pleckstrin homology domain. Cell 1995, 83, 1037–1046. [Google Scholar] [CrossRef]
  63. Harlan, J.E.; Hajduk, P.J.; Yoon, H.S.; Fesik, S.W. Pleckstrin homology domains bind to phosphatidylinositol-4,5-bisphosphate. Nature 1994, 371, 168–170. [Google Scholar] [CrossRef]
  64. Hirose, K.; Kadowaki, S.; Tanabe, M.; Takeshima, H.; Iino, M. Spatiotemporal dynamics of inositol 1,4,5-trisphosphate that underlies complex Ca2+ mobilization patterns. Science 1999, 284, 1527–1530. [Google Scholar] [CrossRef]
  65. Yagisawa, H.; Sakuma, K.; Paterson, H.F.; Cheung, R.; Allen, V.; Hirata, H.; Watanabe, Y.; Hirata, M.; Williams, R.L.; Katan, M. Replacements of single basic amino acids in the pleckstrin homology domain of phospholipase C-delta1 alter the ligand binding, phospholipase activity, and interaction with the plasma membrane. J. Biol. Chem. 1998, 273, 417–424. [Google Scholar] [CrossRef]
  66. Feng, J.F.; Rhee, S.G.; Im, M.J. Evidence that phospholipase delta1 is the effector in the Gh (transglutaminase II)-mediated signaling. J. Biol. Chem. 1996, 271, 16451–16454. [Google Scholar] [CrossRef]
  67. Sidhu, R.S.; Clough, R.R.; Bhullar, R.P. Regulation of phospholipase C-delta1 through direct interactions with the small GTPase Ral and calmodulin. J. Biol. Chem. 2005, 280, 21933–21941. [Google Scholar] [CrossRef] [PubMed]
  68. Kelley, G.G.; Reks, S.E.; Smrcka, A.V. Hormonal regulation of phospholipase Cepsilon through distinct and overlapping pathways involving G12 and Ras family G-proteins. Biochem. J. 2004, 378, 129–139. [Google Scholar] [CrossRef]
  69. Bunney, T.D.; Katan, M. Phospholipase C epsilon: Linking second messengers and small GTPases. Trends Cell. Biol. 2006, 16, 640–648. [Google Scholar] [CrossRef]
  70. Wing, M.R.; Houston, D.; Kelley, G.G.; Der, C.J.; Siderovski, D.P.; Harden, T.K. Activation of phospholipase C-epsilon by heterotrimeric G protein betagamma-subunits. J. Biol. Chem. 2001, 276, 48257–48261. [Google Scholar] [CrossRef] [PubMed]
  71. Bunney, T.D.; Harris, R.; Gandarillas, N.L.; Josephs, M.B.; Roe, S.M.; Sorli, S.C.; Paterson, H.F.; Rodrigues-Lima, F.; Esposito, D.; Ponting, C.P.; et al. Structural and mechanistic insights into ras association domains of phospholipase C epsilon. Mol. Cell. 2006, 21, 495–507. [Google Scholar] [CrossRef] [PubMed]
  72. Kelley, G.G.; Reks, S.E.; Ondrako, J.M.; Smrcka, A.V. Phospholipase C(epsilon): A novel Ras effector. EMBO J. 2001, 20, 743–754. [Google Scholar] [CrossRef]
  73. Song, C.; Hu, C.D.; Masago, M.; Kariyai, K.; Yamawaki-Kataoka, Y.; Shibatohge, M.; Wu, D.; Satoh, T.; Kataoka, T. Regulation of a novel human phospholipase C, PLCepsilon, through membrane targeting by Ras. J. Biol. Chem. 2001, 276, 2752–2757. [Google Scholar] [CrossRef]
  74. Jin, T.G.; Satoh, T.; Liao, Y.; Song, C.; Gao, X.; Kariya, K.; Hu, C.D.; Kataoka, T. Role of the CDC25 homology domain of phospholipase Cepsilon in amplification of Rap1-dependent signaling. J. Biol. Chem. 2001, 276, 30301–30307. [Google Scholar] [CrossRef] [PubMed]
  75. Wing, M.R.; Snyder, J.T.; Sondek, J.; Harden, T.K. Direct activation of phospholipase C-epsilon by Rho. J. Biol. Chem. 2003, 278, 41253–41258. [Google Scholar] [CrossRef] [PubMed]
  76. Lopez, I.; Mak, E.C.; Ding, J.; Hamm, H.E.; Lomasney, J.W. A novel bifunctional phospholipase c that is regulated by Galpha 12 and stimulates the Ras/mitogen-activated protein kinase pathway. J. Biol. Chem. 2001, 276, 2758–2765. [Google Scholar] [CrossRef]
  77. Nomikos, M.; Kashir, J.; Lai, F.A. The role and mechanism of action of sperm PLC-zeta in mammalian fertilization. Biochem. J. 2017, 474, 3659–3673. [Google Scholar] [CrossRef]
  78. Nomikos, M.; Sanders, J.R.; Parthimos, D.; Buntwal, L.; Calver, B.L.; Stamatiadis, P.; Smith, A.; Clue, M.; Sideratou, Z.; Swann, K.; et al. Essential Role of the EF-hand Domain in Targeting Sperm Phospholipase Cζ to Membrane Phosphatidylinositol 4,5-Bisphosphate (PIP2). J. Biol. Chem. 2015, 290, 29519–29530. [Google Scholar] [CrossRef]
  79. Yu, Y.; Nomikos, M.; Theodoridou, M.; Nounesis, G.; Lai, F.A.; Swann, K. PLCζ causes Ca(2+) oscillations in mouse eggs by targeting intracellular and not plasma membrane PI(4,5)P(2). Mol. Biol. Cell. 2012, 23, 371–380. [Google Scholar] [CrossRef]
  80. Nakahara, M.; Shimozawa, M.; Nakamura, Y.; Irino, Y.; Morita, M.; Kudo, Y.; Fukami, K. A novel phospholipase C, PLC(eta)2, is a neuron-specific isozyme. J. Biol. Chem. 2005, 280, 29128–29134. [Google Scholar] [CrossRef]
  81. Popovics, P.; Lu, J.; Nadia Kamil, L.; Morgan, K.; Millar, R.P.; Schmid, R.; Blindauer, C.A.; Stewart, A.J. A canonical EF-loop directs Ca(2+) -sensitivity in phospholipase C-η2. J. Cell. Biochem. 2014, 115, 557–565. [Google Scholar] [CrossRef] [PubMed]
  82. Zhou, Y.; Wing, M.R.; Sondek, J.; Harden, T.K. Molecular cloning and characterization of PLC-eta2. Biochem. J. 2005, 391, 667–676. [Google Scholar] [CrossRef]
  83. Zhou, Y.; Sondek, J.; Harden, T.K. Activation of human phospholipase C-eta2 by Gbetagamma. Biochemistry 2008, 47, 4410–4417. [Google Scholar] [CrossRef] [PubMed]
  84. Böhm, D.; Schwegler, H.; Kotthaus, L.; Nayernia, K.; Rickmann, M.; Köhler, M.; Rosenbusch, J.; Engel, W.; Flügge, G.; Burfeind, P. Disruption of PLC-beta 1-mediated signal transduction in mutant mice causes age-dependent hippocampal mossy fiber sprouting and neurodegeneration. Mol. Cell. Neurosci. 2002, 21, 584–601. [Google Scholar] [CrossRef]
  85. Desprairies, C.; Valence, S.; Maurey, H.; Helal, S.I.; Weckhuysen, S.; Soliman, H.; Mefford, H.C.; Spentchian, M.; Héron, D.; Leguern, E.; et al. Three novel patients with epileptic encephalopathy due to biallelic mutations in the PLCB1 gene. Clin. Genet. 2020, 97, 477–482. [Google Scholar] [CrossRef]
  86. Kurian, M.A.; Meyer, E.; Vassallo, G.; Morgan, N.V.; Prakash, N.; Pasha, S.; Hai, N.A.; Shuib, S.; Rahman, F.; Wassmer, E.; et al. Phospholipase C beta 1 deficiency is associated with early-onset epileptic encephalopathy. Brain 2010, 133, 2964–2970. [Google Scholar] [CrossRef]
  87. Hwang, H.J.; Yang, Y.R.; Kim, H.Y.; Choi, Y.; Park, K.S.; Lee, H.; Ma, J.S.; Yamamoto, M.; Kim, J.; Chae, Y.C.; et al. Phospholipase C-β1 potentiates glucose-stimulated insulin secretion. FASEB J. 2019, 33, 10668–10679. [Google Scholar] [CrossRef] [PubMed]
  88. Hwang, H.J.; Jang, H.J.; Cocco, L.; Suh, P.G. The regulation of insulin secretion via phosphoinositide-specific phospholipase Cβ signaling. Adv. Biol. Regul. 2019, 71, 10–18. [Google Scholar] [CrossRef]
  89. Ratti, S.; Marvi, M.V.; Mongiorgi, S.; Obeng, E.O.; Rusciano, I.; Ramazzotti, G.; Morandi, L.; Asioli, S.; Zoli, M.; Mazzatenta, D.; et al. Impact of phospholipase C β1 in glioblastoma: A study on the main mechanisms of tumor aggressiveness. Cell. Mol. Life Sci. 2022, 79, 195. [Google Scholar] [CrossRef]
  90. Jiang, H.; Kuang, Y.; Wu, Y.; Xie, W.; Simon, M.I.; Wu, D. Roles of phospholipase C beta2 in chemoattractant-elicited responses. Proc. Natl. Acad. Sci. USA 1997, 94, 7971–7975. [Google Scholar] [CrossRef]
  91. Li, Z.; Jiang, H.; Xie, W.; Zhang, Z.; Smrcka, A.V.; Wu, D. Roles of PLC-beta2 and -beta3 and PI3Kgamma in chemoattractant-mediated signal transduction. Science 2000, 287, 1046–1149. [Google Scholar] [CrossRef]
  92. Zhao, G.Q.; Zhang, Y.; Hoon, M.A.; Chandrashekar, J.; Erlenbach, I.; Ryba, N.J.; Zuker, C.S. The receptors for mammalian sweet and umami taste. Cell 2003, 115, 255–266. [Google Scholar] [PubMed]
  93. Damak, S.; Rong, M.; Yasumatsu, K.; Kokrashvili, Z.; Pérez, C.A.; Shigemura, N.; Yoshida, R.; Mosinger, B., Jr.; Glendinning, J.I.; Ninomiya, Y.; et al. Trpm5 null mice respond to bitter, sweet, and umami compounds. Chem. Senses 2006, 31, 253–264. [Google Scholar] [CrossRef]
  94. Hisatsune, C.; Yasumatsu, K.; Takahashi-Iwanaga, H.; Ogawa, N.; Kuroda, Y.; Yoshida, R.; Ninomiya, Y.; Mikoshiba, K. Abnormal taste perception in mice lacking the type 3 inositol 1,4,5-trisphosphate receptor. J. Biol. Chem. 2007, 282, 37225–37231. [Google Scholar]
  95. Wang, L.; Zhou, Y.; Chen, Z.; Sun, L.; Wu, J.; Li, H.; Liu, F.; Wang, F.; Yang, C.; Yang, J.; et al. PLCβ2 negatively regulates the inflammatory response to virus infection by inhibiting phosphoinositide-mediated activation of TAK1. Nat. Commun. 2019, 10, 746. [Google Scholar] [PubMed]
  96. Xiao, W.; Kashiwakura, J.; Hong, H.; Yasudo, H.; Ando, T.; Maeda-Yamamoto, M.; Wu, D.; Kawakami, Y.; Kawakami, T. Phospholipase C-β3 regulates FcɛRI-mediated mast cell activation by recruiting the protein phosphatase SHP-1. Immunity 2011, 34, 893–904. [Google Scholar] [CrossRef]
  97. Ando, T.; Xiao, W.; Gao, P.; Namiranian, S.; Matsumoto, K.; Tomimori, Y.; Hong, H.; Yamashita, H.; Kimura, M.; Kashiwakura, J.; et al. Critical role for mast cell Stat5 activity in skin inflammation. Cell. Rep. 2014, 6, 366–376. [Google Scholar] [CrossRef] [PubMed]
  98. Xiao, W.; Hong, H.; Kawakami, Y.; Kato, Y.; Wu, D.; Yasudo, H.; Kimura, A.; Kubagawa, H.; Bertoli, L.F.; Davis, R.S.; et al. Tumor suppression by phospholipase C-beta3 via SHP-1-mediated dephosphorylation of Stat5. Cancer Cell. 2009, 16, 161–171. [Google Scholar] [CrossRef] [PubMed]
  99. Wang, Z.; Liu, B.; Wang, P.; Dong, X.; Fernandez-Hernando, C.; Li, Z.; Hla, T.; Li, Z.; Claffey, K.; Smith, J.D.; et al. Phospholipase C beta3 deficiency leads to macrophage hypersensitivity to apoptotic induction and reduction of atherosclerosis in mice. J. Clin. Investig. 2008, 118, 195–204. [Google Scholar] [CrossRef] [PubMed]
  100. Rimessi, A.; Bezzerri, V.; Salvatori, F.; Tamanini, A.; Nigro, F.; Dechecchi, M.C.; Santangelo, A.; Prandini, P.; Munari, S.; Provezza, L.; et al. PLCB3 Loss of Function Reduces Pseudomonas aeruginosa-Dependent IL-8 Release in Cystic Fibrosis. Am. J. Respir. Cell. Mol. Biol. 2018, 59, 428–436. [Google Scholar] [CrossRef]
  101. Bezzerri, V.; d’Adamo, P.; Rimessi, A.; Lanzara, C.; Crovella, S.; Nicolis, E.; Tamanini, A.; Athanasakis, E.; Tebon, M.; Bisoffi, G.; et al. Phospholipase C-β3 is a key modulator of IL-8 expression in cystic fibrosis bronchial epithelial cells. J. Immunol. 2011, 186, 4946–4958. [Google Scholar] [CrossRef] [PubMed]
  102. Ben-Salem, S.; Robbins, S.M.; Lm Sobreira, N.; Lyon, A.; Al-Shamsi, A.M.; Islam, B.K.; Akawi, N.A.; John, A.; Thachillath, P.; Al Hamed, S.; et al. Defect in phosphoinositide signalling through a homozygous variant in PLCB3 causes a new form of spondylometaphyseal dysplasia with corneal dystrophy. J. Med. Genet. 2018, 55, 122–130. [Google Scholar] [CrossRef]
  103. Kim, D.; Jun, K.S.; Lee, S.B.; Kang, N.G.; Min, D.S.; Kim, Y.H.; Ryu, S.H.; Suh, P.G.; Shin, H.S. Phospholipase C isozymes selectively couple to specific neurotransmitter receptors. Nature 1997, 389, 290–293. [Google Scholar] [CrossRef] [PubMed]
  104. Jiang, H.; Lyubarsky, A.; Dodd, R.; Vardi, N.; Pugh, E.; Baylor, D.; Simon, M.I.; Wu, D. Phospholipase C beta 4 is involved in modulating the visual response in mice. Proc. Natl. Acad. Sci. USA 1996, 93, 14598–14601. [Google Scholar] [CrossRef]
  105. Cheong, E.; Zheng, Y.; Lee, K.; Lee, J.; Kim, S.; Sanati, M.; Lee, S.; Kim, Y.S.; Shin, H.S. Deletion of phospholipase C beta4 in thalamocortical relay nucleus leads to absence seizures. Proc. Natl. Acad. Sci. USA 2009, 106, 21912–21917. [Google Scholar] [CrossRef]
  106. Moore, A.R.; Ceraudo, E.; Sher, J.J.; Guan, Y.; Shoushtari, A.N.; Chang, M.T.; Zhang, J.Q.; Walczak, E.G.; Kazmi, M.A.; Taylor, B.S.; et al. Recurrent activating mutations of G-protein-coupled receptor CYSLTR2 in uveal melanoma. Nat. Genet. 2016, 48, 675–680. [Google Scholar] [CrossRef] [PubMed]
  107. Kido, Y.; Gordon, C.T.; Sakazume, S.; Ben Bdira, E.; Dattani, M.; Wilson, L.C.; Lyonnet, S.; Murakami, N.; Cunningham, M.L.; Amiel, J.; et al. Further characterization of atypical features in auriculocondylar syndrome caused by recessive PLCB4 mutations. Am. J. Med. Genet. A 2013, 161A, 2339–2346. [Google Scholar] [CrossRef]
  108. Ji, Q.S.; Winnier, G.E.; Niswender, K.D.; Horstman, D.; Wisdom, R.; Magnuson, M.A.; Carpenter, G. Essential role of the tyrosine kinase substrate phospholipase C-gamma1 in mammalian growth and development. Proc. Natl. Acad. Sci. USA 1997, 94, 2999–3003. [Google Scholar] [CrossRef]
  109. Liao, H.J.; Kume, T.; McKay, C.; Xu, M.J.; Ihle, J.N.; Carpenter, G. Absence of erythrogenesis and vasculogenesis in Plcg1-deficient mice. J. Biol. Chem. 2002, 277, 9335–9341. [Google Scholar] [CrossRef]
  110. Fu, G.; Chen, Y.; Yu, M.; Podd, A.; Schuman, J.; He, Y.; Di, L.; Yassai, M.; Haribhai, D.; North, P.E.; et al. Phospholipase C{gamma}1 is essential for T cell development, activation, and tolerance. J. Exp. Med. 2010, 207, 309–318. [Google Scholar] [CrossRef]
  111. Shirane, M.; Sawa, H.; Kobayashi, Y.; Nakano, T.; Kitajima, K.; Shinkai, Y.; Nagashima, K.; Negishi, I. Deficiency of phospholipase C-gamma1 impairs renal development and hematopoiesis. Development 2001, 128, 5173–5180. [Google Scholar] [CrossRef] [PubMed]
  112. Arteaga, C.L.; Johnson, M.D.; Todderud, G.; Coffey, R.J.; Carpenter, G.; Page, D.L. Elevated content of the tyrosine kinase substrate phospholipase C-gamma 1 in primary human breast carcinomas. Proc. Natl. Acad. Sci. USA 1991, 88, 10435–10439. [Google Scholar] [CrossRef] [PubMed]
  113. Park, J.G.; Lee, Y.H.; Kim, S.S.; Park, K.J.; Noh, D.Y.; Ryu, S.H.; Suh, P.G. Overexpression of phospholipase C-gamma 1 in familial adenomatous polyposis. Cancer Res. 1994, 54, 2240–2244. [Google Scholar]
  114. Noh, D.Y.; Lee, Y.H.; Kim, S.S.; Kim, Y.I.; Ryu, S.H.; Suh, P.G.; Park, J.G. Elevated content of phospholipase C-gamma 1 in colorectal cancer tissues. Cancer 1994, 73, 36–41. [Google Scholar] [CrossRef]
  115. Thomas, S.M.; Coppelli, F.M.; Wells, A.; Gooding, W.E.; Song, J.; Kassis, J.; Drenning, S.D.; Grandis, J.R. Epidermal growth factor receptor-stimulated activation of phospholipase Cgamma-1 promotes invasion of head and neck squamous cell carcinoma. Cancer Res. 2003, 63, 5629–5635. [Google Scholar] [PubMed]
  116. Sala, G.; Dituri, F.; Raimondi, C.; Previdi, S.; Maffucci, T.; Mazzoletti, M.; Rossi, C.; Iezzi, M.; Lattanzio, R.; Piantelli, M.; et al. Phospholipase Cgamma1 is required for metastasis development and progression. Cancer Res. 2008, 68, 10187–10196. [Google Scholar] [CrossRef] [PubMed]
  117. Lattanzio, R.; Marchisio, M.; La Sorda, R.; Tinari, N.; Falasca, M.; Alberti, S.; Miscia, S.; Ercolani, C.; Di Benedetto, A.; Perracchio, L.; et al. Overexpression of activated phospholipase Cγ1 is a risk factor for distant metastases in T1-T2, N0 breast cancer patients undergoing adjuvant chemotherapy. Int. J. Cancer. 2013, 132, 1022–1031. [Google Scholar] [CrossRef]
  118. Lattanzio, R.; Iezzi, M.; Sala, G.; Tinari, N.; Falasca, M.; Alberti, S.; Buglioni, S.; Mottolese, M.; Perracchio, L.; Natali, P.G.; et al. PLC-gamma- 1 phosphorylation status is prognostic of metastatic risk in patients with earlystage Luminal-A and -B breast cancer subtypes. BMC Cancer 2019, 19, 747. [Google Scholar] [CrossRef] [PubMed]
  119. Behjati, S.; Tarpey, P.S.; Sheldon, H.; Martincorena, I.; Van Loo, P.; Gundem, G.; Wedge, D.C.; Ramakrishna, M.; Cooke, S.L.; Pillay, N.; et al. Recurrent PTPRB and PLCG1 mutations in angiosarcoma. Nat. Genet. 2014, 46, 376–379. [Google Scholar] [CrossRef] [PubMed]
  120. Huang, S.C.; Zhang, L.; Sung, Y.S.; Chen, C.L.; Kao, Y.C.; Agaram, N.P.; Singer, S.; Tap, W.D.; D’Angelo, S.; Antonescu, C.R. Recurrent CIC Gene Abnormalities in Angiosarcomas: A Molecular Study of 120 Cases With Concurrent Investigation of PLCG1, KDR, MYC, and FLT4 Gene Alterations. Am. J. Surg. Pathol. 2016, 40, 645–655. [Google Scholar] [CrossRef]
  121. Kunze, K.; Spieker, T.; Gamerdinger, U.; Nau, K.; Berger, J.; Dreyer, T.; Sindermann, J.R.; Hoffmeier, A.; Gattenlöhner, S.; Bräuninger, A. A recurrent activating PLCG1 mutation in cardiac angiosarcomas increases apoptosis resistance and invasiveness of endothelial cells. Cancer Res. 2014, 74, 6173–6183. [Google Scholar] [CrossRef] [PubMed]
  122. Murali, R.; Chandramohan, R.; Möller, I.; Scholz, S.L.; Berger, M.; Huberman, K.; Viale, A.; Pirun, M.; Socci, N.D.; Bouvier, N.; et al. Targeted massively parallel sequencing of angiosarcomas reveals frequent activation of the mitogen activated protein kinase pathway. Oncotarget 2015, 6, 36041–36052. [Google Scholar] [CrossRef] [PubMed]
  123. Kataoka, K.; Nagata, Y.; Kitanaka, A.; Shiraishi, Y.; Shimamura, T.; Yasunaga, J.; Totoki, Y.; Chiba, K.; Sato-Otsubo, A.; Nagae, G.; et al. Integrated molecular analysis of adult T cell leukemia/lymphoma. Nat. Genet. 2015, 47, 1304–1315. [Google Scholar] [CrossRef]
  124. Vaqué, J.P.; Gómez-López, G.; Monsálvez, V.; Varela, I.; Martínez, N.; Pérez, C.; Domínguez, O.; Graña, O.; Rodríguez-Peralto, J.L.; Rodríguez-Pinilla, S.M.; et al. PLCG1 mutations in cutaneous T-cell lymphomas. Blood 2014, 123, 2034–2043. [Google Scholar] [CrossRef] [PubMed]
  125. Wang, L.; Ni, X.; Covington, K.R.; Yang, B.Y.; Shiu, J.; Zhang, X.; Xi, L.; Meng, Q.; Langridge, T.; Drummond, J.; et al. Genomic profiling of Sézary syndrome identifies alterations of key T cell signaling and differentiation genes. Nat. Genet. 2015, 47, 1426–1434. [Google Scholar] [CrossRef]
  126. Manso, R.; Rodríguez-Pinilla, S.M.; González-Rincón, J.; Gómez, S.; Monsalvo, S.; Llamas, P.; Rojo, F.; Pérez-Callejo, D.; Cereceda, L.; Limeres, M.A.; et al. Recurrent presence of the PLCG1 S345F mutation in nodal peripheral T-cell lymphomas. Haematologica 2015, 100, e25–e27. [Google Scholar] [CrossRef] [PubMed]
  127. Vallois, D.; Dobay, M.P.; Morin, R.D.; Lemonnier, F.; Missiaglia, E.; Juilland, M.; Iwaszkiewicz, J.; Fataccioli, V.; Bisig, B.; Roberti, A.; et al. Activating mutations in genes related to TCR signaling in angioimmunoblastic and other follicular helper T-cell-derived lymphomas. Blood 2016, 128, 1490–1502. [Google Scholar] [CrossRef] [PubMed]
  128. Saliakoura, M.; Rossi Sebastiano, M.; Pozzato, C.; Heidel, F.H.; Schnöder, T.M.; Savic Prince, S.; Bubendorf, L.; Pinton, P.A.; Schmid, R.; Baumgartner, J.; et al. PLCγ1 suppression promotes the adaptation of KRAS-mutant lung adenocarcinomas to hypoxia. Nat. Cell. Biol. 2020, 22, 1382–1395. [Google Scholar] [CrossRef] [PubMed]
  129. Kang, D.S.; Yang, Y.R.; Lee, C.; Park, B.; Park, K.I.; Seo, J.K.; Seo, Y.K.; Cho, H.; Lucio, C.; Suh, P.G. Netrin-1/DCC-mediated PLCγ1 activation is required for axon guidance and brain structure development. EMBO Rep. 2018, 19, e46250. [Google Scholar] [CrossRef]
  130. Xie, Y.; Ding, Y.Q.; Hong, Y.; Feng, Z.; Navarre, S.; Xi, C.X.; Zhu, X.J.; Wang, C.L.; Ackerman, S.L.; Kozlowski, D.; et al. Phosphatidylinositol transfer protein-alpha in netrin-1-induced PLC signalling and neurite outgrowth. Nat. Cell. Biol. 2005, 7, 1124–1132. [Google Scholar] [CrossRef]
  131. Xie, Y.; Hong, Y.; Ma, X.Y.; Ren, X.R.; Ackerman, S.; Mei, L.; Xiong, W.C. DCC-dependent phospholipase C signaling in netrin-1-induced neurite elongation. J. Biol. Chem. 2006, 281, 2605–2611. [Google Scholar] [CrossRef]
  132. Yang, Y.R.; Jung, J.H.; Kim, S.J.; Hamada, K.; Suzuki, A.; Kim, H.J.; Lee, J.H.; Kwon, O.B.; Lee, Y.K.; Kim, J.; et al. Forebrain-specific ablation of phospholipase Cγ1 causes manic-like behavior. Mol. Psychiatry 2017, 22, 1473–1482. [Google Scholar] [CrossRef] [PubMed]
  133. Kim, H.Y.; Yang, Y.R.; Hwang, H.; Lee, H.E.; Jang, H.J.; Kim, J.; Yang, E.; Kim, H.; Rhim, H.; Suh, P.G.; et al. Deletion of PLCγ1 in GABAergic neurons increases seizure susceptibility in aged mice. Sci. Rep. 2019, 9, 17761. [Google Scholar] [CrossRef] [PubMed]
  134. Gu, B.; Huang, Y.Z.; He, X.P.; Joshi, R.B.; Jang, W.; McNamara, J.O. A Peptide Uncoupling BDNF Receptor TrkB from Phospholipase Cγ1 Prevents Epilepsy Induced by Status Epilepticus. Neuron 2015, 88, 484–491. [Google Scholar] [CrossRef]
  135. Hashimoto, A.; Takeda, K.; Inaba, M.; Sekimata, M.; Kaisho, T.; Ikehara, S.; Homma, Y.; Akira, S.; Kurosaki, T. Cutting edge: Essential role of phospholipase C-gamma 2 in B cell development and function. J. Immunol. 2000, 165, 1738–1742. [Google Scholar] [CrossRef] [PubMed]
  136. Yu, P.; Constien, R.; Dear, N.; Katan, M.; Hanke, P.; Bunney, T.D.; Kunder, S.; Quintanilla-Martinez, L.; Huffstadt, U.; Schröder, A.; et al. Autoimmunity and inflammation due to a gain-of-function mutation in phospholipase C gamma 2 that specifically increases external Ca2+ entry. Immunity 2005, 22, 451–465. [Google Scholar] [CrossRef] [PubMed]
  137. Wang, D.; Feng, J.; Wen, R.; Marine, J.C.; Sangster, M.Y.; Parganas, E.; Hoffmeyer, A.; Jackson, C.W.; Cleveland, J.L.; Murray, P.J.; et al. Phospholipase Cgamma2 is essential in the functions of B cell and several Fc receptors. Immunity. 2000, 13, 25–35. [Google Scholar] [CrossRef]
  138. Chen, Y.; Wang, X.; Di, L.; Fu, G.; Chen, Y.; Bai, L.; Liu, J.; Feng, X.; McDonald, J.M.; Michalek, S.; et al. Phospholipase Cgamma2 mediates RANKL-stimulated lymph node organogenesis and osteoclastogenesis. J. Biol. Chem. 2008, 283, 29593–29601. [Google Scholar] [CrossRef]
  139. Bunney, T.D.; Baxendale, R.W.; Martins, M.S.; Romberg, N.; Komarow, H.; Aksentijevich, I.; Kim, H.S.; Ho, J.; Cruse, G.; Jung, M.Y.; et al. Cold urticaria, immunodeficiency, and autoimmunity related to PLCG2 deletions. N. Engl. J. Med. 2012, 366, 330–338. [Google Scholar]
  140. Novice, T.; Kariminia, A.; Del Bel, K.L.; Lu, H.; Sharma, M.; Lim, C.J.; Read, J.; Lugt, M.V.; Hannibal, M.C.; O’Dwyer, D.; et al. A Germline Mutation in the C2 Domain of PLCγ2 Associated with Gain-of-Function Expands the Phenotype for PLCG2-Related Diseases. J. Clin. Immunol. 2020, 40, 267–276. [Google Scholar] [CrossRef]
  141. Martín-Nalda, A.; Fortuny, C.; Rey, L.; Bunney, T.D.; Alsina, L.; Esteve-Solé, A.; Bull, D.; Anton, M.C.; Basagaña, M.; Casals, F.; et al. Severe Autoinflammatory Manifestations and Antibody Deficiency Due to Novel Hypermorphic PLCG2 Mutations. J. Clin. Immunol. 2020, 40, 987–1000. [Google Scholar] [CrossRef]
  142. Zhou, Q.; Lee, G.S.; Brady, J.; Datta, S.; Katan, M.; Sheikh, A.; Martins, M.S.; Bunney, T.D.; Santich, B.H.; Moir, S.; et al. A hypermorphic missense mutation in PLCG2, encoding phospholipase Cγ2, causes a dominantly inherited autoinflammatory disease with immunodeficiency. Am. J. Hum. Genet. 2012, 91, 713–720. [Google Scholar] [CrossRef]
  143. Nakamura, Y.; Fukami, K.; Yu, H.; Takenaka, K.; Kataoka, Y.; Shirakata, Y.; Nishikawa, S.I.; Hashimoto, K.; Yoshida, N.; Takenawa, T. Phospholipase Cdelta1 is required for skin stem cell lineage commitment. EMBO J. 2003, 22, 2981–2991. [Google Scholar] [CrossRef] [PubMed]
  144. Nakamura, Y.; Ichinohe, M.; Hirata, M.; Matsuura, H.; Fujiwara, T.; Igarashi, T.; Nakahara, M.; Yamaguchi, H.; Yasugi, S.; Takenawa, T.; et al. Phospholipase C-delta1 is an essential molecule downstream of Foxn1, the gene responsible for the nude mutation, in normal hair development. FASEB J. 2008, 22, 841–849. [Google Scholar] [CrossRef] [PubMed]
  145. Kiuru, M.; Kurban, M.; Itoh, M.; Petukhova, L.; Shimomura, Y.; Wajid, M.; Christiano, A.M. Hereditary leukonychia, or porcelain nails, resulting from mutations in PLCD1. Am. J. Hum. Genet. 2011, 88, 839–844. [Google Scholar] [CrossRef]
  146. Khan, A.K.; Khan, S.A.; Muhammad, N.; Muhammad, N.; Ahmad, J.; Nawaz, H.; Nasir, A.; Farman, S.; Khan, S. Mutation in Phospholipase C, δ1 ( PLCD1) Gene Underlies Hereditary Leukonychia in a Pashtun Family and Review of the Literature. Balkan J. Med. Genet. 2018, 21, 69–72. [Google Scholar] [CrossRef]
  147. Khan, T.; Khan, M.; Yousaf, A.; Khan, S.; Naeem, M.; Shah, A.; Murtaza, G.; Ali, A.; Jabeen, N.; Hussain, H.M.J.; et al. Whole exome sequencing identifies a novel dominant missense mutation underlying leukonychia in a Pakistani family. J. Hum. Genet. 2018, 63, 1071–1076. [Google Scholar] [CrossRef] [PubMed]
  148. Xue, K.; Zheng, Y.; Shen, C.; Cui, Y. Identification of a novel PLCD1 mutation in Chinese Han pedigree with hereditary leukonychia and koilonychia. J. Cosmet. Dermatol. 2019, 18, 912–915. [Google Scholar] [CrossRef]
  149. Kanemaru, K.; Nakamura, Y.; Sato, K.; Kojima, R.; Takahashi, S.; Yamaguchi, M.; Ichinohe, M.; Kiyonari, H.; Shioi, G.; Kabashima, K.; et al. Epidermal phospholipase Cδ1 regulates granulocyte counts and systemic interleukin-17 levels in mice. Nat. Commun. 2012, 3, 963. [Google Scholar] [CrossRef]
  150. Kanemaru, K.; Nakamura, Y.; Totoki, K.; Fukuyama, T.; Shoji, M.; Kaneko, H.; Shiratori, K.; Yoneda, A.; Inoue, T.; Iwakura, Y.; et al. Phospholipase Cδ1 regulates p38 MAPK activity and skin barrier integrity. Cell. Death Differ. 2017, 24, 1079–1090. [Google Scholar] [CrossRef]
  151. Park, S.J.; Haan, K.D.; Nakamura, Y.; Fukami, K.; Fisher, T.E. PLCδ1 Plays Central Roles in the Osmotic Activation of ΔN-TRPV1 Channels in Mouse Supraoptic Neurons and in Murine Osmoregulation. J. Neurosci. 2021, 41, 3579–3587. [Google Scholar] [CrossRef]
  152. Haan, K.D.; Park, S.J.; Nakamura, Y.; Fukami, K.; Fisher, T.E. Osmotically evoked PLCδ1-dependent translocation of ΔN-TRPV1 channels in rat supraoptic neurons. iScience 2023, 26, 106258. [Google Scholar] [CrossRef]
  153. Hu, X.T.; Zhang, F.B.; Fan, Y.C.; Shu, X.S.; Wong, A.H.; Zhou, W.; Shi, Q.L.; Tang, H.M.; Fu, L.; Guan, X.Y.; et al. Phospholipase C delta 1 is a novel 3p22.3 tumor suppressor involved in cytoskeleton organization, with its epigenetic silencing correlated with high-stage gastric cancer. Oncogene 2009, 28, 2466–2475. [Google Scholar] [CrossRef] [PubMed]
  154. Mu, H.; Wang, N.; Zhao, L.; Li, S.; Li, Q.; Chen, L.; Luo, X.; Qiu, Z.; Li, L.; Ren, G.; et al. Methylation of PLCD1 and adenovirus-mediated PLCD1 overexpression elicits a gene therapy effect on human breast cancer. Exp. Cell. Res. 2015, 332, 179–189. [Google Scholar] [CrossRef] [PubMed]
  155. Xiang, T.; Li, L.; Fan, Y.; Jiang, Y.; Ying, Y.; Putti, T.C.; Tao, Q.; Ren, G. PLCD1 is a functional tumor suppressor inducing G(2)/M arrest and frequently methylated in breast cancer. Cancer Biol. Ther. 2010, 10, 520–527. [Google Scholar] [CrossRef]
  156. Fu, L.; Qin, Y.R.; Xie, D.; Hu, L.; Kwong, D.L.; Srivastava, G.; Tsao, S.W.; Guan, X.Y. Characterization of a novel tumor-suppressor gene PLC delta 1 at 3p22 in esophageal squamous cell carcinoma. Cancer Res. 2007, 67, 10720–10726. [Google Scholar] [CrossRef] [PubMed]
  157. Nakamura, Y.; Hamada, Y.; Fujiwara, T.; Enomoto, H.; Hiroe, T.; Tanaka, S.; Nose, M.; Nakahara, M.; Yoshida, N.; Takenawa, T.; et al. Phospholipase C-delta1 and -delta3 are essential in the trophoblast for placental development. Mol. Cell. Biol. 2005, 25, 10979–10988. [Google Scholar] [CrossRef]
  158. Nakamura, Y.; Kanemaru, K.; Kojima, R.; Hashimoto, Y.; Marunouchi, T.; Oka, N.; Ogura, T.; Tanonaka, K.; Fukami, K. Simultaneous loss of phospholipase Cδ1 and phospholipase Cδ3 causes cardiomyocyte apoptosis and cardiomyopathy. Cell Death Dis. 2014, 5, e1215. [Google Scholar] [CrossRef] [PubMed]
  159. Fukami, K.; Nakao, K.; Inoue, T.; Kataoka, Y.; Kurokawa, M.; Fissore, R.A.; Nakamura, K.; Katsuki, M.; Mikoshiba, K.; Yoshida, N.; et al. Requirement of phospholipase Cdelta4 for the zona pellucida-induced acrosome reaction. Science 2001, 292, 920–923. [Google Scholar] [CrossRef]
  160. Fukami, K.; Yoshida, M.; Inoue, T.; Kurokawa, M.; Fissore, R.A.; Yoshida, N.; Mikoshiba, K.; Takenawa, T. Phospholipase Cdelta4 is required for Ca2+ mobilization essential for acrosome reaction in sperm. J. Cell. Biol. 2003, 161, 79–88. [Google Scholar] [CrossRef]
  161. Wang, H.; Oestreich, E.A.; Maekawa, N.; Bullard, T.A.; Vikstrom, K.L.; Dirksen, R.T.; Kelley, G.G.; Blaxall, B.C.; Smrcka, A.V. Phospholipase C epsilon modulates beta-adrenergic receptor-dependent cardiac contraction and inhibits cardiac hypertrophy. Circ. Res. 2005, 97, 1305–1313. [Google Scholar] [CrossRef] [PubMed]
  162. Nash, C.A.; Wei, W.; Irannejad, R.; Smrcka, A.V. Golgi localized β1-adrenergic receptors stimulate Golgi PI4P hydrolysis by PLCε to regulate cardiac hypertrophy. eLife 2019, 8, e48167. [Google Scholar] [CrossRef]
  163. Tadano, M.; Edamatsu, H.; Minamisawa, S.; Yokoyama, U.; Ishikawa, Y.; Suzuki, N.; Saito, H.; Wu, D.; Masago-Toda, M.; Yamawaki-Kataoka, Y.; et al. Congenital semilunar valvulogenesis defect in mice deficient in phospholipase C epsilon. Mol. Cell. Biol. 2005, 25, 2191–2199. [Google Scholar] [CrossRef] [PubMed]
  164. Takenaka, N.; Edamatsu, H.; Suzuki, N.; Saito, H.; Inoue, Y.; Oka, M.; Hu, L.; Kataoka, T. Overexpression of phospholipase Cε in keratinocytes upregulates cytokine expression and causes dermatitis with acanthosis and T-cell infiltration. Eur. J. Immunol. 2011, 41, 202–213. [Google Scholar] [CrossRef] [PubMed]
  165. Hu, L.; Edamatsu, H.; Takenaka, N.; Ikuta, S.; Kataoka, T. Crucial role of phospholipase Cepsilon in induction of local skin inflammatory reactions in the elicitation stage of allergic contact hypersensitivity. J. Immunol. 2010, 184, 993–1002. [Google Scholar] [CrossRef]
  166. Dusaban, S.S.; Purcell, N.H.; Rockenstein, E.; Masliah, E.; Cho, M.K.; Smrcka, A.V.; Brown, J.H. Phospholipase C epsilon links G protein-coupled receptor activation to inflammatory astrocytic responses. Proc. Natl. Acad. Sci. USA 2013, 110, 3609–3614. [Google Scholar] [CrossRef]
  167. Hinkes, B.; Wiggins, R.C.; Gbadegesin, R.; Vlangos, C.N.; Seelow, D.; Nürnberg, G.; Garg, P.; Verma, R.; Chaib, H.; Hoskins, B.E.; et al. Positional cloning uncovers mutations in PLCE1 responsible for a nephrotic syndrome variant that may be reversible. Nat. Genet. 2006, 38, 1397–1405. [Google Scholar] [CrossRef]
  168. Rao, J.; Ashraf, S.; Tan, W.; van der Ven, A.T.; Gee, H.Y.; Braun, D.A.; Fehér, K.; George, S.P.; Esmaeilniakooshkghazi, A.; Choi, W.I.; et al. Advillin acts upstream of phospholipase C ϵ1 in steroid-resistant nephrotic syndrome. J. Clin. Investig. 2017, 127, 4257–4269. [Google Scholar] [CrossRef]
  169. Saunders, C.M.; Larman, M.G.; Parrington, J.; Cox, L.J.; Royse, J.; Blayney, L.M.; Swann, K.; Lai, F.A. PLC zeta: A sperm-specific trigger of Ca(2+) oscillations in eggs and embryo development. Development 2002, 129, 3533–3544. [Google Scholar] [CrossRef]
  170. Knott, J.G.; Kurokawa, M.; Fissore, R.A.; Schultz, R.M.; Williams, C.J. Transgenic RNA interference reveals role for mouse sperm phospholipase Czeta in triggering Ca2+ oscillations during fertilization. Biol. Reprod. 2005, 72, 992–996. [Google Scholar] [CrossRef]
  171. Nozawa, K.; Satouh, Y.; Fujimoto, T.; Oji, A.; Ikawa, M. Sperm-borne phospholipase C zeta-1 ensures monospermic fertilization in mice. Sci. Rep. 2018, 8, 1315. [Google Scholar] [CrossRef] [PubMed]
  172. Yoshida, N.; Amanai, M.; Fukui, T.; Kajikawa, E.; Brahmajosyula, M.; Iwahori, A.; Nakano, Y.; Shoji, S.; Diebold, J.; Hessel, H.; et al. Broad, ectopic expression of the sperm protein PLCZ1 induces parthenogenesis and ovarian tumours in mice. Development 2007, 134, 3941–3952. [Google Scholar] [CrossRef]
  173. Escoffier, J.; Lee, H.C.; Yassine, S.; Zouari, R.; Martinez, G.; Karaouzène, T.; Coutton, C.; Kherraf, Z.E.; Halouani, L.; Triki, C.; et al. Homozygous mutation of PLCZ1 leads to defective human oocyte activation and infertility that is not rescued by the WW-binding protein PAWP. Hum. Mol. Genet. 2016, 25, 878–891. [Google Scholar] [CrossRef] [PubMed]
  174. Torra-Massana, M.; Cornet-Bartolomé, D.; Barragán, M.; Durban, M.; Ferrer-Vaquer, A.; Zambelli, F.; Rodriguez, A.; Oliva, R.; Vassena, R. Novel phospholipase C zeta 1 mutations associated with fertilization failures after ICSI. Hum. Reprod. 2019, 34, 1494–1504. [Google Scholar] [CrossRef]
  175. Heytens, E.; Parrington, J.; Coward, K.; Young, C.; Lambrecht, S.; Yoon, S.Y.; Fissore, R.A.; Hamer, R.; Deane, C.M.; Ruas, M.; et al. Reduced amounts and abnormal forms of phospholipase C zeta (PLCzeta) in spermatozoa from infertile men. Hum. Reprod. 2009, 24, 2417–2428. [Google Scholar] [CrossRef]
  176. Feisst, C.; Albert, D.; Steinhilber, D.; Werz, O. The aminosteroid phospholipase C antagonist U-73122 (1-[6-[[17-beta-3-methoxyestra-1,3,5(10)-trien-17-yl]amino]hexyl]-1H-pyrrole-2,5-dione) potently inhibits human 5-lipoxygenase in vivo and in vitro. Mol. Pharmacol. 2005, 67, 1751–1757. [Google Scholar] [CrossRef] [PubMed]
  177. Hollywood, M.A.; Sergeant, G.P.; Thornbury, K.D.; McHale, N.G. The PI-PLC inhibitor U-73122 is a potent inhibitor of the SERCA pump in smooth muscle. Br. J. Pharmacol. 2010, 160, 1293–1294. [Google Scholar] [CrossRef]
  178. Leitner, M.G.; Michel, N.; Behrendt, M.; Dierich, M.; Dembla, S.; Wilke, B.U.; Konrad, M.; Lindner, M.; Oberwinkler, J.; Oliver, D. Direct modulation of TRPM4 and TRPM3 channels by the phospholipase C inhibitor U73122. Br. J. Pharmacol. 2016, 173, 2555–2569. [Google Scholar] [CrossRef]
  179. Krjukova, J.; Holmqvist, T.; Danis, A.S.; Akerman, K.E.; Kukkonen, J.P. Phospholipase C activator m-3M3FBS affects Ca2+ homeostasis independently of phospholipase C activation. Br. J. Pharmacol. 2004, 143, 3–7. [Google Scholar] [CrossRef]
  180. Trinquet, E.; Fink, M.; Bazin, H.; Grillet, F.; Maurin, F.; Bourrier, E.; Ansanay, H.; Leroy, C.; Michaud, A.; Durroux, T.; et al. D-myo-inositol 1-phosphate as a surrogate of D-myo-inositol 1,4,5-tris phosphate to monitor G protein-coupled receptor activation. Anal. Biochem. 2006, 358, 126–135. [Google Scholar] [CrossRef]
  181. Huang, W.; Hicks, S.N.; Sondek, J.; Zhang, Q. A fluorogenic, small molecule reporter for mammalian phospholipase C isozymes. ACS Chem. Biol. 2011, 6, 223–228. [Google Scholar] [CrossRef] [PubMed]
  182. Huang, W.; Wang, X.; Endo-Streeter, S.; Barrett, M.; Waybright, J.; Wohlfeld, C.; Hajicek, N.; Harden, T.K.; Sondek, J.; Zhang, Q. A membrane-associated, fluorogenic reporter for mammalian phospholipase C isozymes. J. Biol. Chem. 2018, 293, 1728–1735. [Google Scholar] [CrossRef] [PubMed]
Figure 1. The domain structures of PLC. The active sites and catalytic residues in typical PLC isozymes are located within the X and Y domains (X and Y). While PLCζ is the only exception that lacks the PH domain, all typical PLC isozymes also possess the PH domain (PH), EF-hand motifs (EF), and the C2 domain (C2). PLCβ has the C-terminal domain (CTD), as well as the PDZ-binding motif. PLCγ features a multidomain insertion between the X and Y domains, consisting of the split PH domain (PH), N-terminal SH2 domain (nSH2), C-terminal SH2 domain (cSH2), and SH3 domain (SH3). PLCε has Ras association domains (RA) and a Cdc25 homology domain (CDC25). Atypical PLC isozymes, the PLCXD family contain a catalytic domain with a sequence that is similar to the X domain (X).
Figure 1. The domain structures of PLC. The active sites and catalytic residues in typical PLC isozymes are located within the X and Y domains (X and Y). While PLCζ is the only exception that lacks the PH domain, all typical PLC isozymes also possess the PH domain (PH), EF-hand motifs (EF), and the C2 domain (C2). PLCβ has the C-terminal domain (CTD), as well as the PDZ-binding motif. PLCγ features a multidomain insertion between the X and Y domains, consisting of the split PH domain (PH), N-terminal SH2 domain (nSH2), C-terminal SH2 domain (cSH2), and SH3 domain (SH3). PLCε has Ras association domains (RA) and a Cdc25 homology domain (CDC25). Atypical PLC isozymes, the PLCXD family contain a catalytic domain with a sequence that is similar to the X domain (X).
Biomolecules 13 00915 g001
Figure 2. The schematic pathway of PI turnover induced by PLC. PLC isozymes hydrolyze PM PI(4,5)P2 to produce two second messengers: IP3 and DAG. IP3 binds IP3 receptors (IP3R) present in the ER, inducing the release of Ca2+ into the cytosol, while DAG activates PKC. DAG also activates TRPC3, TRPC6, and TRPC7 and increases intracellular Ca2+ concentration. IP3 is metabolized to IP4, IP5, IP6, and PP-InsPs. IP3 is also metabolized to IP2. IP2 is further dephosphorylated to myo-inositol. Myo-inositol is then re-incorporated into the PI synthesis cycle by binding to CDP-DAG in the ER membrane. DAG is metabolized to PA. DAG and PA are transported from the PM to the ER, where the PI synthetic enzymes utilize them. Thus, DAG generated by the hydrolysis of PI(4,5)P2 is recycled into PI to maintain the total pool of phosphatidylinositol phosphates. DAG is metabolized to endocannabinoid 2-arachidonoyl glycerol (2-AG), which acts as an agonist of endocannabinoid receptors (CB). PI(4,5)P2 is also metabolized to PI(3,4,5)P3, which activates AKT.
Figure 2. The schematic pathway of PI turnover induced by PLC. PLC isozymes hydrolyze PM PI(4,5)P2 to produce two second messengers: IP3 and DAG. IP3 binds IP3 receptors (IP3R) present in the ER, inducing the release of Ca2+ into the cytosol, while DAG activates PKC. DAG also activates TRPC3, TRPC6, and TRPC7 and increases intracellular Ca2+ concentration. IP3 is metabolized to IP4, IP5, IP6, and PP-InsPs. IP3 is also metabolized to IP2. IP2 is further dephosphorylated to myo-inositol. Myo-inositol is then re-incorporated into the PI synthesis cycle by binding to CDP-DAG in the ER membrane. DAG is metabolized to PA. DAG and PA are transported from the PM to the ER, where the PI synthetic enzymes utilize them. Thus, DAG generated by the hydrolysis of PI(4,5)P2 is recycled into PI to maintain the total pool of phosphatidylinositol phosphates. DAG is metabolized to endocannabinoid 2-arachidonoyl glycerol (2-AG), which acts as an agonist of endocannabinoid receptors (CB). PI(4,5)P2 is also metabolized to PI(3,4,5)P3, which activates AKT.
Biomolecules 13 00915 g002
Figure 3. Activation mechanisms of PLC. PLCβ isozymes are activated by Gαq, Gβγ, and Rac. PLCγ isoforms are activated by RTKs. PLCδ activity can be stimulated by Ca2+ within the physiological range through the activation of the other PLC isozymes or influx of Ca2+ through calcium channels. PLCε can be activated by GPCRs and RTKs as well as by small GTPases. PLCζ and PLCη are highly sensitive to Ca2+ and respond to small elevations in intracellular Ca2+ levels. PLCη is also activated by GPCR.
Figure 3. Activation mechanisms of PLC. PLCβ isozymes are activated by Gαq, Gβγ, and Rac. PLCγ isoforms are activated by RTKs. PLCδ activity can be stimulated by Ca2+ within the physiological range through the activation of the other PLC isozymes or influx of Ca2+ through calcium channels. PLCε can be activated by GPCRs and RTKs as well as by small GTPases. PLCζ and PLCη are highly sensitive to Ca2+ and respond to small elevations in intracellular Ca2+ levels. PLCη is also activated by GPCR.
Biomolecules 13 00915 g003
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Kanemaru, K.; Nakamura, Y. Activation Mechanisms and Diverse Functions of Mammalian Phospholipase C. Biomolecules 2023, 13, 915. https://doi.org/10.3390/biom13060915

AMA Style

Kanemaru K, Nakamura Y. Activation Mechanisms and Diverse Functions of Mammalian Phospholipase C. Biomolecules. 2023; 13(6):915. https://doi.org/10.3390/biom13060915

Chicago/Turabian Style

Kanemaru, Kaori, and Yoshikazu Nakamura. 2023. "Activation Mechanisms and Diverse Functions of Mammalian Phospholipase C" Biomolecules 13, no. 6: 915. https://doi.org/10.3390/biom13060915

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