Next Article in Journal
Functional Analysis of Pheromone Biosynthesis Activating Neuropeptide Receptor Isoforms in Maruca vitrata
Next Article in Special Issue
Cardiovascular Functions of Ena/VASP Proteins: Past, Present and Beyond
Previous Article in Journal
The Haves and Have-Nots: The Mitochondrial Permeability Transition Pore across Species
Previous Article in Special Issue
Multiple Roles of cAMP in Vertebrate Retina
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Plethora of Functions Condensed into Tiny Phospholipids: The Story of PI4P and PI(4,5)P2

Laboratory of Hematopoiesis, Department of Biotechnology, University of Rijeka, R. Matejcic 2, 51000 Rijeka, Croatia
*
Author to whom correspondence should be addressed.
Cells 2023, 12(10), 1411; https://doi.org/10.3390/cells12101411
Submission received: 10 April 2023 / Revised: 8 May 2023 / Accepted: 9 May 2023 / Published: 17 May 2023
(This article belongs to the Special Issue Exclusive Review Papers in "Cell Signaling")

Abstract

:
Phosphoinositides (PIs) are small, phosphorylated lipids that serve many functions in the cell. They regulate endo- and exocytosis, vesicular trafficking, actin reorganization, and cell mobility, and they act as signaling molecules. The most abundant PIs in the cell are phosphatidylinositol-4-monophosphate (PI4P) and phosphatidylinositol-4,5-bisphosphate [PI(4,5)P2]. PI4P is mostly localized at the Golgi apparatus where it regulates the anterograde trafficking from the Golgi apparatus to the plasma membrane (PM), but it also localizes at the PM. On the other hand, the main localization site of PI(4,5)P2 is the PM where it regulates the formation of endocytic vesicles. The levels of PIs are regulated by many kinases and phosphatases. Four main kinases phosphorylate the precursor molecule phosphatidylinositol into PI4P, divided into two classes (PI4KIIα, PI4KIIβ, PI4KIIIα, and PI4KIIIβ), and three main kinases phosphorylate PI4P to form PI(4,5)P2 (PI4P5KIα, PI4P5KIβ, and PI4P5KIγ). In this review, we discuss the localization and function of the kinases that produce PI4P and PI(4,5)P2, as well as the localization and function of their product molecules with an overview of tools for the detection of these PIs.

Graphical Abstract

1. Introduction

Phospholipids are abundant, complex, and highly diverse components of cell architecture [1]. In addition to their main function as membrane building blocks, phospholipids have also been shown to be involved in intracellular trafficking and signal transduction, thus having a more dynamic role in cellular physiology [2]. A type of low-abundant phospholipids, called phosphoinositides (PIs), in addition to being a component of the cell membranes, interact with numerous effector proteins and were shown to serve as either signaling molecules themselves or to generate secondary messengers within different cells [3]. To date, PIs have been shown to serve important functions in processes such as cytoskeleton reorganization and membrane curvature generation upon endo-, exo-, or phagocytosis [4,5], polarized cell migration [6,7], cell adhesion [5,8,9], cellular transport utilizing either ion channels [5], concentration gradient, or protein and lipid transfer proteins, as well as receptor-mediated signaling [10,11,12] and gene expression [13]. Despite the range of different PI-dependent cellular processes, the extending spectra of evidence confirming the role of PIs in various physiological and pathophysiological states still manage both to surprise and excite the scientific community. PI metabolism is strictly spatially and temporally controlled by a pool of different kinases, phosphatases, and phospholipases, maintaining their levels and determining multiple aspects of cellular fate [8]. A recent mathematical model predicts that there are 19 kinases and 35 phosphatases involved in the PI pathway alone [14,15]. The distorted homeostasis in PI metabolism was shown to be involved in neurodegeneration [11,16] and neuroinflammation [17], oncogenesis [4,18], infection [19,20], and immune response [21,22]. In this review, we highlight the localization, function, and clinical relevance of the two most abundant PIs, phosphatidylinositol-4-monophosphate (PI4P) and phosphatidylinositol-4,5-bisphosphate [PI(4,5)P2] (Figure 1), as well as the kinases involved in their main synthesis pathways. We also provide a summary of known tools for the detection of those PIs.

The (re)Birth of PI4P and PI(4,5)P2

The parent molecule for PI synthesis is phosphatidylinositol (PtdIns), which consists of a myo-inositol ring linked by a phosphodiester bond to a diacylglycerol (DAG) backbone with two hydrophobic fatty acyl chains [19]. PtdIns is synthesized at the cytoplasmic leaflet of the endoplasmic reticulum (ER), but it can later either be flipped to the luminal leaflet and form glycosylphosphatidylinositol-linked proteins (GPIs) or redistributed across other membranes by vesicular transport or lipid transport proteins (LTPs) to generate other PIs [12,23]. Due to the fact of its steric properties, the inositol ring is available for phosphorylation at three of its hydroxyl groups (3, 4, or 5-OH) (Figure 1). Depending on the degree and position of phosphorylation, the inositol ring yields a total of seven different PIs (Figure 2). The number and site of the phosphorylation change the extent of their negative charge and modify their steric properties, both of which influence their distinctive binding affinities [24]. Interestingly, one distinguishing characteristic of mammalian PIs is their DAG composition, mostly enriched in polyunsaturated stearic (sn-1) and arachidonic acid (sn-2). Other phospholipids were not shown to be enriched in one specific fatty-acyl chain combination, which suggests that the PIs are being recycled [25]. The phosphorylation of PIs is a reversible process, meaning their levels are tightly controlled by multiple kinases and phosphatases [24] (Figure 2). Consequently, both PI4P and PI(4,5)P2 can be generated by three distinct pathways. PI4P mainly originates from the phosphorylation of PtdIn on the 4-OH by the action of phosphatidylinositol-4-kinases (PI4Ks) [26]. Two minor routes involve either dephosphorylation of PI(4,5)P2 on 5-OH by 5-phosphatases (OCRL, synaptojanin1/2, INPP5) or dephosphorylation of PI(3,4)P2 on 3-OH by 3-phosphatase PTEN [11,27]. PI4P can be further phosphorylated on 5-OH by type I phosphatidylinositol-4-phosphate 5-kinases (PI4P5KI), presenting a main route of PI(4,5)P2 formation. Alternative routes include PI5P phosphorylation on 4-OH by type II phosphatidylinositol-5-phosphate 4-kinases (PI5P4KII) or PI(3,4,5)P3 dephosphorylation on 3-OH by 3-phosphatases PTEN and TPIP [24,27]. Different isoforms of these enzymes reside in specific subcellular compartments, thus forming localized pools of their products [20]. We further discuss the structure, localization, and contribution to PI4P and PI(4,5)P2 pools, biological and clinical relevance, and finally the pharmacological potential of different PI4K and PI4P5KI isoforms, while phosphatases are reviewed elsewhere [27,28,29].

2. PI4Ks

PI4Ks are present as four different isozymes divided into two subfamilies: type II kinases (PI4KIIα and PI4KIIβ) and type III kinases (PI4Kα and PI4Kβ) [20]. The nomenclature lacks type I variants since they were later shown to be phosphatidylinositol-3-kinases (PI3Ks) [5]. The catalytic domain of type III PI4Ks is similar to those of PI3Ks, possibly having a role in interactions with other proteins [30,31]. Surprisingly, the structure of type II kinases varies significantly from other lipid kinases, resembling those of protein kinases [32]. Thus, the two types have distinct stimulation/inhibition routes [19]. For example, type II kinases are sensitive to adenosine and Ca2+ inhibition and insensitive to wortmannin (pan-anti-PI3K), while type III show the opposite susceptibility [33]. Furthermore, the smaller type II kinases mostly act as monomers [34] and are destined for membrane tethering due to the inserted palmitoylation sites [33]. In contrast, larger type III kinases have a dimerization region [34], are cytosolic, and shuffle among membranes of different cellular compartments or transiently associate with the PM [19].

2.1. PI4KIIα

PI4KIIα (55 kDa) is considered to be the most abundant and active isoform in mammalian cells, generating almost half of the total cellular PI4P levels [20,35]. PI4KIIα is constitutively incorporated within the membrane, possibly due to the distinctive highly hydrophobic pockets in addition to palmitoylation, thus acting as an integral membrane protein [36,37]. PI4KIIα is palmitoylated in the Golgi apparatus, and since the palmitoyltransferases require cholesterol, both PI4KIIα membrane association and activity depend on local cholesterol levels [38,39]. PI4KIIα is known to be recruited to the PM by Rac1 [37]. The structures of PI4KIIα and PI4KIIβ show a high level of similarity in their C-terminal but lower in their N-terminal regions [40]. PI4KIIα and PI4KIIβ share an N-terminal palmitoylation site, but only PI4KIIα has a clathrin adapter protein-3 (AP-3) binding site in the N-terminal proline-rich region [33,41], which is required for transport from late endosomes to lysosomes [39]. In addition to the greater trans-Golgi network (TGN) and endosomal PI4P pools, PI4KIIα also contributes to sub-pools in the PM, lysosomes [42], and multiple small post-Golgi vesicles [38]. It was shown to be associated with various cellular processes, such as autophagy [43,44,45], lysosomal delivery [41], lysosomal repair [46], endosomal receptor sorting [20], exocytosis [47], signal transduction [48,49], actin remodeling [47], and sphingomyelin synthesis [50].
Due to the fact of its abundance in mammalian cells, its role in biological processes and clinical perspective is of much interest. The role of PI4KIIα in oncogenesis is so far the most researched one, with multiple studies supposing its role as an oncoprotein, overexpressed in several different cancer types [38]. The loss of PI4KIIα was shown to be involved in tumor cell apoptosis by driving endosomal EGFR degradation [51] leading to impaired antiapoptotic Akt signaling [52], which makes PI4KIIα an interesting therapeutic target in EGFR-dependent tumors such as breast cancer [51], glioblastomas, and some subtypes of lung and colorectal cancers [53]. It was shown that the complex formed by PI4KIIα and AP-3 regulates lysosomal function in healthy cells, but forming a complex with RNA-dependent protein kinase R (PKR) promotes misfolded prion protein clearance and viability in cancerous cells [42]. Since the expression of cellular prion proteins is known to promote cancer proliferation and metastasis [54], targeting their accumulation via PI4KIIα complex destabilization could be an interesting therapeutic target. Another approach could be the disabling of angiogenesis in cancer cells by downregulating PI4KIIα, which was shown to inhibit human epidermal growth factor receptor 2 (HER-2) activity and lead to a decrease in hypoxia-inducible factor 1α (HIF1α) overexpression, as well as disruption of PI3K-mediated increase in pro-angiogenesis factors [49].
Although both type II kinases are ubiquitously expressed, PI4KIIα was shown to have higher expression in synaptic vesicles [33,55], posing the question of its role in neurological defects. Interestingly, PI4KIIα binding partner AP-3 is known to regulate endosomal cargo transport to axons and synaptic vesicles, as well as synaptic vesicle formation. It is yet to be investigated if this interaction is involved in progressive motor disabilities seen in PI4KIIα-deficient mice, a phenotype resembling human hereditary spastic paraplegia [56]. In addition, it was shown that the N-terminus of PI4KIIα is phosphorylated by glycogen synthase kinase 3 (GSK3), which is known to be essential for neurodevelopment and CNS function. The phosphorylation enables the binding of the kinase to AP-3, mediating neuronal receptor trafficking and expression [57]. Furthermore, PI4KIIα-mediated PI metabolism was also shown to be an interesting target in Alzheimer’s disease treatment by modulating γ-secretase activity, an enzyme responsible for amyloid β-peptide processing [58]. Recently, patients with biallelic deficiency of the enzyme were shown to suffer from severe encephalopathy and movement disorder, possibly due to Rab7-associated late endosome-lysosome trafficking defects [59]. Another intriguing case is of a patient diagnosed with cutis laxa, a severe connective-tissue disorder with common neurological symptoms, bearing a PI4KIIα mutation, for the first time posing the question of the importance of lipid metabolism in the pathophysiology of this disorder [60].
In addition to the rising evidence of its role in oncogenesis and neurological defects, there are reported functions of PI4KIIα in metabolic disorders, such as diabetes and Gaucher disease as well [61,62]. In diabetes, it was shown that PI4KIIα is involved in mediating exocytic insulin release by regulating protein kinase D (PKD) activity [61]. In Gaucher disease, PI4KIIα depletion leads to the failed transport of lysosomal integral membrane protein type 2 (LIMP-2) and β-glucocerebrosidase (GBA) enzyme secretion [62]. PI4KIIα was also shown to be important in Chlamydia species bacterial infection by mediating replication complex formation through ADP-ribosylation factor 1 (Arf-1) binding [35]. Finally, PI4KIIα was shown to form a tri-component complex with AP-3 and BLOC-1, both showing defects in Hermansky–Pudlak syndrome characterized by albinism and impaired platelet aggregation due to defects in endosomal sorting [63,64]. In addition, the role of PI metabolism in maintaining hemostatic function has been shown in human endothelial cells in which the PI4KIIα depletion caused the abnormal length of Weibel–Palade bodies and impaired folding of von Willebrand factor [65], an important component of primary platelet adhesion during vessel wall injury.
Altogether, the variety of PI4KIIα-related disorders, especially its supposed role as an EGFR-linked oncoprotein, made it an attractive target for the development of pharmacological inhibitors that could exceed the limitation of previously used nonselective inhibitors such as phenylarsine oxide (PAO) [35,66]. To date, there is one selective, reversible, and commercially available substrate-competitive PI4KIIα inhibitor, PI-273, shown to inhibit breast cancer cell growth [48]. Detailed information on PI-273 and other inhibitors is given in Table 1 (IC50, off-targets, research area, and state of development).

2.2. PI4KIIβ

PI4KIIβ (55 kDa) is the less studied and less active type II PI4K [20,35]. Unlike constitutively membrane-bound PI4KIIα, PI4KIIβ is mostly cytosolic or to a smaller extent associated with the PM, TGN, ER, and clathrin-coated vesicles (CCVs) [37,39]. Cytosolic PI4KIIβ is nonpalmitoylated, nonactive, and stabilized by the interaction with heat shock protein 90 (Hsp90) on its C-terminal site [33,36]. Upon platelet-derived growth factor (PDGF)-mediated cell activation, PI4KIIβ is recruited to the PM by Rac1 GTPase and this interaction is responsible for its activation as well. The membrane association potentiates its activity thus making it as active as PI4KIIα. PI4KIIβ has a different hydrophobic pocket structure, thus acting as a peripheral protein [36,37]. In addition, instead of the AP-3 binding site, PI4KIIβ has a binding site for clathrin adapter protein-1 (AP-1) on the N-terminal proline-rich region [33], important for trafficking from TGN to endosomes [47]. PI4KIIβ synthesizes PI4P pools mostly between the TGN and endosomes, as well as sub-pools in late, recycling, or enlarged early endosomes [47,88]. It is associated with vesicular cargo sorting, endosomal receptor trafficking [88], intracellular signaling, phagocytosis [89], and actin remodeling [47]. The role of PI4KIIβ in cancerogenesis is less investigated, but recent findings propose both its antioncogenic and prooncogenic effects. PI4KIIβ was shown to impair apoptosis in carcinoma cells by interaction with the tumor suppressor prostate apoptosis response-4 (PAR-4) [38]. In addition, the interaction of PI4KIIβ and AP-1 is required for Wnt receptor endosomal sorting, a pathway that was seen to be disturbed in human cancerogenesis [88]. In addition, PI4KIIβ was found to be downregulated in human tumor cells, which increases invadopodia formation and promotes cancer metastasis pointing to its antioncogenetic effect [47]. Another interesting role of PI4KIIβ is shown in inflammation, as observed by its upregulation during lipopolysaccharide (LPS) signaling. It was shown that its palmitoylation is a crucial step in the LPS-mediated production of proinflammatory cytokines [90]. In addition, PI4KIIβ was shown to be important in immune response mediating early T-cell activation through CD3 TCR [20]. Interestingly, PI4KIIβ is the more abundant type II isoform in the liver, but the reason for that remains to be elucidated [89]. There is no selective inhibitor for PI4KIIβ reported to date [33]; however, due to the presence of emerging evidence on the role of PI4Ks in a variety of different cellular processes, its crystal structure was recently solved and studies are underway to find selective inhibitors [91].

2.3. PI4Kα

Type III PI4Kα (230 kDa) has a nuclear localization sequence (NLS) and a pleckstrin homology domain (PH), possibly allowing its membrane interaction [33,89]. It is mainly located in the ER, transiently associates with the PM, and can also be found at early cis-Golgi compartments, nucleolus [92], multivesicular bodies (MVBs), and outer mitochondrial membrane [35]. It was shown to generate the largest hormone-sensitive PM PI4P pool, even though it is not bound to the membrane itself [89,93]. PI4Kα was shown to be transiently transported and held to the PM by a carrier complex [26]. This complex contains TTC7A (tetratricopeptide repeat domain 7A) for targeting to the PM, EFR3 for tethering, and FAM126A for stabilization [11,94]. The role of PI4Kα in the ER is somewhat less clear. PI4P produced by PI4Kα at the ER could be transported to the PM or the cis-Golgi. An unexpected but likely role of PI4Kα at the Golgi apparatus is the control of coat protein complex I (COPI) trafficking of protein and lipid cargo by altering PI4P levels and association with Rab1 [35]. PI4Kα can also be found in the nucleus where its PI4P pool is likely, at least in part, used for nuclear PI(4,5)P2 production [35].
PI4Kα was found to be a crucial component in hepatitis C virus (HCV) replication in the liver. It is recruited by NS5A viral protein to provide PI4P, needed for membranous web formation and its integrity preservation [19,39]. These findings resulted in the search for PI4Kα inhibitors in the hope of novel HCV antiviral therapy [20]. Interestingly, it was reported that an anti-HCV agent simeprevir, could make brain and breast tumor cells more prone to radiosensitization and delay their growth by downregulating antiprogrammed death-ligand 1 (PD-L1) and Akt signaling upon PI4Kα inhibition [95]. Further, PI4Kα loss of function was shown to result in impaired hematopoietic differentiation of myeloid and erythroid lineage through Akt and Erk signaling. It was found that in myeloid and erythroid lineage-linked leukemias, normal PI4Kα signaling is inhibited by upregulation of PI4KAP2, formerly considered a nonfunctional PI4Kα pseudogene [96]. In addition, PI4KA tethering to the PM by EFR3 was found to increase the levels of KRAS, the most common human oncogene at the PM and its downstream signaling, thus providing a novel intervention route in KRAS-dependent carcinomas such as pancreatic, colorectal, or lung carcinoma [97]. Finally, since PI4Kα was already shown to be upregulated in pancreatic cancer cell lines and it was linked with poor prognosis in hepatocellular carcinoma [98], it has been highlighted as an interesting anticancer agent.
PI4KA is highly expressed in the brain, and its defect was found to be associated with neurological phenotypes. PI4KA is thought to be important in early brain development since human fetuses carrying an inactivating mutation had fatal brain abnormalities [99]. It was also shown that the mice with PI4KA deletion in Schwann cells show symptoms of neuropathy, due to the fact of its distorted PI metabolism and actin reorganization, leading to impaired myelination [92]. In addition, hypomyelination was also observed when the PI4KA contact with the PM was destabilized by FAM126A depletion in oligodendrocytes [74,100]. Recently, patients with severe neurodevelopmental issues due to the fact of hypomyelination have been reported to carry the biallelic PI4KA variant [100]. Furthermore, an SNP in the promoter region of PI4Kα was associated with schizophrenia phenotype [19].
Interestingly, there was also a reported role of PI4Kα in intestines, its loss caused fatal intestinal lesions in mice, a phenotype previously seen in patients lacking TTC7A, a protein that targets PI4Kα to the membrane [94]. The intriguing data on the impact of PI4Kα complex partners emphasize the role of adequate PI4Kα targeting to the PM in health and disease. Since PI4Kα is an essential gene, its genetical targeting leads to early embryonic lethality [26,34,68]; therefore, its targeting switched a course to pharmacological inhibition. Nevertheless, a careful design of studies using PI4KA pharmacological inhibitors is needed, since it was shown that sometimes a high dose of PI4KA inhibition can lead to sudden death in mice due to the fact of cardiovascular collapse [68]. To date, there are two potent selective PI4Kα inhibitors reported, GSK-A1 and GSK-F1 [15,33,68,69] (detailed information provided in Table 1).

2.4. PI4Kβ

Type III PI4Kβ (92 kDa), in contrast to PI4Kα, lacks a PH domain but instead has a serine-rich region prone to protein kinase D (PKD) phosphorylation and NCS-1 activity-modulating binding site, as well as Rab11 binding site [33]. Along with PI4KIIα and PI4KIIβ, PI4Kβ is another PI4K isoform mainly generating PI4P at the Golgi apparatus [33]. It also greatly contributes to the nuclear PI4P pool and generates PI4P sub-pools in the ER, lysosomal, and outer mitochondrial membrane [35]. Its recruitment to the Golgi apparatus is regulated by Arf1 and ACBD3, and its activity is potentiated by PKD phosphorylation and then stabilized through interaction with the 14-3-3 protein [19,27]. PI4Kβ was shown to be important in regulating structural and functional aspects of the Golgi apparatus, vesicular fusion, cargo sorting, and endocytosis [35]. Like the other PI4K isoforms, PI4Kβ was also shown to be involved in cancerogenesis. It is upregulated in various malignancies such as skin (basal cell carcinoma), brain, as well as breast cancers where it increases Akt signaling through PI(3,4,5)P3 generation. Surprisingly, it was shown that this action is due to the fact of PI4Kβ direct binding of Rab11 and recruitment to recycling endosomes, independent of kinase activity [101]. It is also considered to be a potential antimalarial and antiviral therapeutic [33]. PI4Kβ inhibition was shown to impair poliovirus, picornavirus, coxsackievirus replication, and HCV secretion [33], as well as SARS-Cov entry [102]. Upon enteroviral infection, PI4Kβ is recruited by Arf-1, which is activated by Golgi brefeldin A resistant guanine nucleotide exchange factor 1 (GBF-1) interacting with a viral protein 3A. Another mechanism of PI4Kβ recruitment is utilized by the Aichi virus, whose 3A protein interacts with the ABCD3 kinase mobilizing protein [39]. PI4Kβ is a positive regulator of Hedgehog signaling, thus it is important in vertebrate embryonic development and tissue homeostasis and regeneration in adults [103]. Additionally, its inhibition leads to a suppressed immunological response to organ transplants in vivo [33]. Surprisingly, it is involved in Gaucher disease, where it controls the exit of β-glucocerebrosidase from the TGN [39]. Several potent selective PI4Kβ inhibitors are available, including IN-9, IN-10, T-00127-HEV1, BF738735, and enviroxime [15,33] (Table 1). Enviroxime and Plasmodium falciparum selective PI4KIIIβ inhibitor MMV390048 are the only PI4K inhibitors so far that entered the clinical studies but were eventually both discontinued or terminated, respectively [15].
It is worth noting that in the same compartments (such as the Golgi apparatus or the PM) different kinases produce the same product, e.g., PI4P. However, it is still not entirely clear why different kinases are performing the same reaction, if and how they work together. Interestingly, the kinases that produce PI4P (types II and III) are evolutionarily distinct from each other [34]. It is taught that since PI4P has many different roles in the cell (e.g., regulating anterograde transport from the Golgi apparatus or regulating lipid transport), this may explain why PI4P-producing kinases have evolved multiple times. Moreover, not all PI4P-generating kinases that produce PI4P at a given organelle are constantly present at that organelle. As explained above, PI4KIIα is constitutively incorporated into the membrane, whereas PI4KIIβ is mostly cytosolic and is recruited to the PM upon cell activation. Therefore, it is possible that the activation state of the cell or its requirements activate different kinases through numerous cell signaling pathways.

3. PI4P5KIs

PI4P5KIs present a major PI(4,5)P2 synthesis pathway due to the fact of significantly higher levels of available PI4P in contrast to PI5P [8]. Interestingly, even though PI4P5KIs preferably bind PI4P, they are not substrate-exclusive like PI4Ks. PI4P5KIs were also shown to bind PI3P or PI(3,4)P2, phosphorylate these substrates on 5-OH, and finally lead to PI(3,4,5)P3 generation [8,40]. PI4P5KIs are present in mammalian cells in three different isoforms: α, β, and γ [55]. PI4P5KIs show a high level of structural similarity in the core kinase domain (70–80%), mostly differ in the C-terminal domain and, to a lesser extent, in N-terminal domains or a kinase insert, altering their localization and interaction preference [3,40,87]. A great deal of its interaction partners can also bind PI(4,5)P2, modulating their function [104]. It is considered that the activation loop insert is responsible for sensing local PM levels of phosphatidylserine and cholesterol thus targeting PI4P5KIs to the PM and controlling their activity [12]. PI4P5KIs activity was also shown to be modulated by phosphatidic acid, small GTPases (Rac and Arf) [105], phosphorylation, or Wnt signaling [55], and suggested being mediated by homo- or hetero-dimers formation [12]. PI4P5KIs can all be found in the PM to a different extent, and some of them localize in the Golgi apparatus, nucleus, and endosomal compartments as well. They are known to be involved in various cellular processes such as endocytosis, exocytosis, apoptosis, and regulation of ion channels [20,40], but they are mostly known for their role in actin reorganization [106]. PI4P5KIs have been shown to be involved in multiple aspects of breast, pancreatic, and colorectal cancer oncogenesis [87].

3.1. PI4P5KIα

PI4P5KIα (68 kDa) was shown to mainly localize in the inner leaflet of the PM and the Golgi apparatus but was also found on the site of membrane ruffling and in nuclear speckles [8]. It was shown that the G protein-coupled receptor (GPCR) stimulation translocates the kinase across the membrane and potentiates its activity. PI4P5KIα was shown to regulate phagocytic microbe ingestion by creating a PI(4,5)P2 pool which facilitates WASP and ARP2/3-mediated actin reorganization [3,20]. For example, it was shown to facilitate actin reorganization in T cells, thus allowing HIV-1 entry [20]. Interestingly, it was suggested that its activity in the nucleus can be potentiated as a response to oxidative stress [8]. PI4P5KIα is upregulated in breast and prostate cancer cells, thus leading to poorer outcomes in KRAS/Akt- and p53-dependent oncogenesis [107]. So far the only known PI4P5KIα inhibitor, ISA-2011B, is considered to be a promising anticancer agent in advanced prostate cancer and triple-negative breast cancer [38] (Table 1).

3.2. PI4P5KIβ

PI4P5KIβ (68 kDa) is mostly found in the perinuclear region [3], vesicular structures, and to a lesser extent in the PM [5]. It was shown that the PI4P5KIβ interaction with Rac-1 GTPase regulates both its PM localization and activity. It forms a heterodimer with PI4P5KIα, which can compensate for its kinase activity if needed [20]. PI4P5KIβ was shown to be involved in various aspects of actin dynamics during cell migration and activation. It is targeted to the leading edge of the cell, where its product PI(4,5)P2 induces actin reorganization through N-WASP activation [8]. It was shown that the expression of PI4P5KIβ is needed for actin polymerization by mediating comet formation and elevating stress fibres formation [19]. In addition, the inactive form leads to disturbed EGFR-mediated endocytosis, responsible for clathrin and dynamin trafficking to the PM [40]. Loss of PI4P5KIβ was shown to have a major impact in mediating platelet aggregation by lowering the levels of inositol 1,4,5-triphosphate (IP3), a product of PI(4,5)P2 cleavage, whose generation represents one of the first steps in platelet activation [108].

3.3. PI4P5KIγ

PI4P5KIγ (~90 kDa) is considered to be the major source of PI(4,5)P2 at the PM [12]. Human PI4P5KIγ is a highly diverse family consisting of six splice variants, differing in C-terminal length and all having specific subcellular localizations and interaction partners [3,20]. PI4P5KIγ-v1 represents the most abundant splice variant in most tissues. It is mostly found in the PM, where it is involved in phagocytosis [20]. It was shown to be involved in the process of microbial attachment during macrophage phagocytosis by elevating actin depolymerization. PI4P5KIγ-v2 was shown to be the second most abundant isoform, the one mostly expressed in the brain. It mostly generates the PM PI(4,5)P2 pool but can be found in focal adhesion sites as well. Its role was observed in multiple aspects of cell migration. It is considered to be important for focal adhesions assembly by PI(4,5)P2 generation, focal adhesion kinase (FAK) activation, and interaction with talin and integrin receptors [3,4,19]. PI4P5KIγ-v2 also maintains cell polarity, another aspect of cell migration by promoting E-cadherin transport to the site of adherens junctions assembly. The cell-cell contact sites based on E-cadherin are an important component of epithelial cancer metastasis. Interestingly, PI4P5KIγ expression was linked to poor prognosis in breast cancer by regulating crucial steps in cancer metastasis, E-cadherin cell-cell contacts, and EGFR-stimulated cell migration [109]. The interaction of PI4P5KIγ-v2 with talin or clathrin adaptor AP2 was also shown to regulate endocytosis of synaptic vesicles, thus mediating neuronal activation. PI4P5Kγ-v3 was observed in the PM but only in neuronal cells across multiple brain regions [20]. PI4P5Kγ-v4 is found in nuclear speckles and PI4P5Kγ-v5 in the PM and endosomal compartments [3,19]. PI4P5Kγ-v5 is needed in endosomal EGFR sorting for lysosomal degradation. One of the mechanisms involves the interaction of PI4P5Kγ-v5 and its product PI(4,5)P2 with sorting nexin 5 (SNX5) leading to EGFR sorting and lysosomal degradation. On the contrary, the same interaction with SNX5 can prevent E-cadherin degradation [110]. PI4P5Kγ-v5 was also shown to have an important role in modulating endosomal maturation. It is needed for Rab7a recruitment to early endosomes, a step that is crucial for Rab5a to Rab7a replacement required for the maturation of early endosomes to late endosomes [104]. Finally, PI4P5Kγ-v6 is mostly found in the PM [111].
PI4P5KIγ was shown to be an essential enzyme during embryonic development since its depletion is lethal in mice [112,113]. The variant was shown to be enriched in the brain, its deficiency leading to neurotransmission defects. It was reported that a mutation in PI4P5KIγ leads to neonatally lethal congenital contracture syndrome type 3 (LCCS3), characterized by motoric neuron degeneration possibly due to the fact of PI(4,5)P2 deficiency in the brain [3,114,115]. Its loss disables the cytoskeleton linkage with the PM in platelets and their precursor cells megakaryocytes [19,116]. PI4P5KIγ is upregulated in colorectal cancer cells, being responsible for cell proliferation and growth through Akt signaling. In this study, the UNC3230 inhibitor was used [87], but it was later shown to lack selectivity. However, due to the lack of a more sensitive inhibitor, it is still widely used (Table 1).

4. PI4P: A Huge Burden on a Small Lipid

PI4P is mostly localized at the Golgi apparatus and the PM, but it can also be found in the ER [117] and cytoplasmic vesicles such as late endosomes and lysosomes [118] (Figure 3). The late endosomes/lysosomes localization was shown only recently with a novel PI4P probe that utilizes the P4M domain of SidM protein from Legionella pneumophila [118]. SidM is localized to bacteria-containing vacuoles where ER-derived materials are recruited through binding to PI4P.
At the Golgi apparatus, PI4P binds effector proteins and regulates the anterograde transport from the Golgi apparatus to the PM via the TGN [119,120,121], impacts protein targeting to endosomes, vesicle formation, and maintenance of Golgi resident enzymes [39,119,120,121]. On the PM it can serve as a regulator of the PM stability and PI(4,5)P2 [122] levels, as well as regulate ion channels [39,122].

4.1. PI4P at the Golgi Apparatus: Glycosylation and Anterograde Trafficking

The Golgi apparatus consists of cis-, medial-, and trans-Golgi cisternae and PI4P is mostly localized and the TGN [123]. It is believed that this is the case because Sac1, the enzyme that dephosphorylates PI4P, is mostly localized to the cis- and medial-Golgi. Not only that PI4P depletion causes a defect in the Golgi function, but abnormal levels of PI4P also lead to a defective Golgi and its enzymes. Sac1 knockdown (KD) by RNAi causes the PI4P levels to rise at the cis- and medial-Golgi. The Golgi becomes bigger and fragmented, PI4P is mislocalized at punctate and peripheral structures, cell proliferation is decreased, and mannosidase II and N-acetylglucosamine transferase-1 are translocated from the Golgi to other intracellular and periphery membranes. In addition, Sac1 KD selectively alters N- and O-glycosylation. Specifically, it decreases the addition of poly-N-acetylgalactosamine repeats on complex, multi-antennary N-glycans and reduces levels of Galβ1-3GalNacα-O-Ser/Thr. It was hypothesized that the exclusion of PI4P from the Golgi cisternae may be essential for preventing anterograde passage of Golgi enzymes to the cell periphery or that COP-1-mediated retrograde transport of Golgi enzymes may be impaired leading to their peripheral accumulation [123].
At the Golgi apparatus, PI4P recruits clathrin adaptors and Golgi-localized, Gamma-adaptin ear homology, Arf-binding proteins (GGAs), critical for lipid transport and cargo delivery from the Golgi apparatus to the PM [124]. Depletion of PI4P at the Golgi apparatus leads to a decline of the vesicular trafficking from the Golgi apparatus to the PM, almost completely shuts down the trafficking from the Golgi apparatus to the MVBs, and leads to the dissociation of GGA1 and GGA2 (but not GGA3) from the Golgi apparatus. Interestingly, early studies showed that the Golgi depletion of PI4P impairs PM PI(4,5)P2 pool, at least for a minor part. When PI(4,5)P2 levels are decreased in the PM, its recovery is slower when the cell is depleted of the Golgi pool of PI4P [124].

4.2. PI4P at the Plasma Membrane

More recent findings suggest that PI4P and PI(4,5)P2 have independent roles in the integrity and identity of the PM [125]. Hammond et al. reported that the selective depletion of PM pools of PI4P by recruitment of Sac1 to the PM (the rapamycin-inducible dimerization of FK506 binding protein, FKBP and the fragment of mTOR that binds rapamycin, FRB) does not affect the clathrin-mediated endocytosis, the formation of secondary messengers PI(3,4,5)P3 and PI(3,4)P2 as well as IP3 indicating that PI4P does not maintain the functionally relevant pool of PI(4,5)P2 at the PM [125]. If not as a precursor of PI(4,5)P2, what is the role of PI4P at the PM? Since it has been shown that PI(4,5)P2 can regulate the activity of some ion channels [126], it was hypothesized that PI4P could also have a similar role [125]. Indeed, the depletion of PI4P and PI(4,5)P2 but not PI4P or PI(4,5)P2 alone led to the inhibition of the capsaicin-activated transient receptor potential vanilloid 1 (TRPV1) cation channel [125]. These findings suggest that either lipid is sufficient for the channel activity and that PI4P at the PM has an autonomous contribution to the polyanionic lipid pool that defines the inner leaflet of the PM. This in turn would make PI(4,5)P2 free for its other roles.

4.3. PI4P at Membrane-Contact Sites

The localization and role of PI4P have also been implied at membrane contact sites (MCSs) [127]. MCSs are regions where membranes of distinct organelles are tethered and in close proximity, where nonvesicular lipid transfer occurs [128]. They have a role in intracellular signaling, lipid exchange and metabolism, and organelle function [128,129]. Lipid exchange is carried out by oxysterol-binding protein (OSBP) and OSBP-related proteins (ORPs) that are enriched at MCSs by binding to PI4P. At ER-Golgi MCSs, the hydrolysis of PI4P directs the exchange of PI4P from the Golgi to the ER and cholesterol from the ER to the Golgi in a four-step process [127]. At ER-PM MCSs, PI4P transfers from the PM to the ER while transferring PS in the opposite direction [117]. It is worth mentioning that the KD of Sac1 decreases the dissociation of PI4P from the PM. It was hypothesized that this implies that PI4P consumption by Sac1 at the ER is necessary for efficient counter-transport by allowing the OSBP-related domains (ORD) to favor PS binding over PI4P binding at the ER [117]. Furthermore, this exchange could help to control the levels of PI4P at the PM and selectively enrich PS in the PM.

5. PI(4,5)P2: A Dual Role

PI(4,5)P2 mainly accumulates at the PM but is also found in low abundance in intracellular membranes such as the ones of the Golgi apparatus, endosomes, ER, and electron-dense structures within the nucleus [130] (Figure 3). At the PM, PI(4,5)P2 is shown to be important for maintaining membrane curvature, clathrin-mediated endocytosis, focal adhesion assembly, and regulation of synaptic vesicle recycling [131], processes that include actin reorganization. PI(4,5)P2 has been also implicated in various nuclear processes such as RNA processing, nuclear export, regulation of nuclear actin, and chromatin remodeling [35]. In the nucleolus, PI(4,5)P2 was shown to colocalize and associate with transcription factor UBF, thus having a role as a transcriptional regulator as well [132]. Furthermore, PI(4,5)P2 is a signaling molecule and serves as a source of secondary messengers inositol 1,4,5-trisphosphate (IP3) and DAG [19].

5.1. PI(4,5)P2 in Clathrin-Mediated Endocytosis

CCVs are coordinately assembled at the PM and the polymerization of the outer clathrin layer is assisted by adaptor proteins which bind to clathrin, membrane lipids, and cargo proteins [133]. The central adaptor protein is adaptor complex 2 (AP-2) which binds all the above-mentioned parts needed for the CCV formation, and the most important membrane lipid is PI(4,5)P2. It has been shown that elevated levels of PI(4,5)P2 lead to the accumulation of CCVs at nerve endings [131] and to the increased levels of endocytosis of transferrin receptors, the association of AP-2 with the membranes, and the number of CCVs at the PM [134]. Adaptor proteins bind to PI(4,5)P2 by different domains such as PH, AP180 N-terminal homology (ANTH) domain, or epsin N-terminal homology domain (ENTH) but other proteins can also associate with PI(4,5)P2 [133]. These proteins mediate the function of the actin cytoskeleton, like Wiskott-Aldrich syndrome protein (WASP) and profilin. Once a CCV enters the cell, it goes through the endocytic pathway and PI(4,5)P2 in CCVs is exchanged for PI3P which allows the formation and maturation of early endosomes and the formation of recycling endosomes [135]. The inability of PI(4,5)P2 hydrolysis by the KD of the oculo-cerebro-renal syndrome of Lowe (OCRL) phosphatase results in the accumulation of PI(4,5)P2 at the CCVs. When PI(4,5)P2 is accumulated on CCVs, enlarged early endosomes form and the neuronal-WASP-dependent increase in endosomal F-actin is induced. Enlarged early endosomes with abundant actin cannot form recycling endosomes and the recycling machinery is shut down [135]. PI(4,5)P2 also plays a role in phagocytosis where the levels correlate with the formation of actin during pseudopod extension [19]. It has also been suggested that PI(4,5)P2 has a role in exocytosis, but the exact mechanisms yet need to be revealed [136]. However, it is assumed that it could be through the regulation of actin reorganization.

5.2. PI(4,5)P2 as a Signaling Molecule

The activation of phospholipase C (PLC) leads to PI(4,5)P2 hydrolysis to second messengers IP3 and DAG [137]. DAG then activates protein kinase C (PKC) which has an important role in several different signaling cascades. IP3 is soluble and binds to its receptor on the ER and increases intracellular calcium levels that activate calcium-sensitive signaling molecules. Furthermore, PI(4,5)P2 itself can act as a regulator of ion channels [126,138]. PI(4,5)P2 in most cases increases channel activity while its hydrolysis by PLC reduces the channel activity. PI(4,5)P2 can act on inward-rectifier K+ channels, voltage-gated K+ channels, voltage-gated Ca2+ channels, sensory transduction channels, and others [138].

6. Tools for Detection of Intracellular Phosphoinositides

To understand the characteristics and functions of PIs within cells, various tools for their visualization and detection have been developed. The visualization of PIs is usually performed using antibodies and/or genetically encoded probes that contain specific protein domains fused with fluorescent proteins [139]. These allow for the determination of the intracellular localization of PIs, relative levels in different compartments, and changes in their levels or distribution in response to genetic or pharmacological manipulation of PI metabolizing enzymes. Furthermore, in recent years researchers developed mass spectrometry analysis for PI detection. Every method for PI detection has its advantages and limitations, so their selection may vary owing to experimental design. Nevertheless, together they significantly contributed to our understanding of PIs and their specific intracellular roles. Here, we will focus on tools commonly used for the detection of PI4P and PI(4,5)P2.

6.1. Antibodies

Commercial antibodies are available against almost all PIs, including PI4P and PI(4,5)P2, and have been widely used for determining their localization and abundance within the cells. However, the antibody detection of these PIs has several caveats and limitations. One of them is that an antibody is much larger than a PI for which the antibody is supposed to be specific thus making it hard to distinguish between two PIs that differ only in one phosphate group [140]. PIs that are already bound to proteins in cells are not reachable to antibodies, therefore underestimating their amount or localization. In addition, lipids are not subject to formaldehyde fixation thus not being immobilized, and using standard detergents might result in the extraction of lipids [139]. Nevertheless, there are a few reliable commercially available antibodies that have been tested and their specificity proved, among which are antibodies for PI4P and PI(4,5)P2. However, antibody detection can provide information about the localization of these lipids only at a given time point and cannot be used to detect transient changes [139]. Next, for the antibody to gain access to intracellular compartments, cells need to be fixed and permeabilized which can lead to differences in the antibody accessibility between different cellular compartments. This has been observed with PI4P and PI(4,5)P2 where, although localized at the PM and/or the Golgi apparatus, they cannot be visualized in both at once [122]. Hammond et al. established specific immunocytochemical protocols for the preservation of different membranes and visualization of PM or intracellular PI4P and PI(4,5)P2 pools [122]. However, the established staining protocols do not stain confidently all cell types and need further optimization [141]. In addition, PIs show different localization in diverse cell types [141].

6.2. Genetically Encoded Probes

Genetically encoded probes contain a specific PI-binding domain fused with fluorescent proteins and are the most valuable tool for the visualization of PIs inside the cells [139]. They are compatible with live cells and allow the analysis of not only the intracellular localization of PIs but also the dynamic changes in their relative levels, as well as the changes in PI levels in response to stimuli. Although shown as useful, several aspects need to be considered when using genetically encoded probes to monitor PIs. The use of protein domains as probes depends on their specificity for the target PI, an affinity that allows the detection of the PI, and an understanding of secondary interactions that may interfere with the distribution of the probe. Some probes can recognize several types of PIs therefore not being selective [142]. Since PIs are bound to endogenous regulatory proteins in cells, their detection by expressed probes depends on the accessibility to those PI pools. Overexpression of probes can sequester their target lipid and, in this way, disturb lipid interaction with downstream effector proteins leading to a dominant negative effect. In addition, probes could be recruited to membranes by interacting with other proteins and in this way sequester protein function [142].
There are several families of PI-binding domains, the largest being the PH domain which is used primarily to monitor the distribution and levels of PI4P and PI(4,5)P2 within the cells (Table 2). The PH domain of PLCδ1 was not only useful for the assessment of the PI(4,5)P2 localization within the cell but also to determine the nanoscale spatial organization of this PI in the PM using single-molecule superresolution microscopy [143].
The limitation of genetically encoded probes is that their specificity and affinity can vary greatly. For example, pools of PI4P on the Golgi and the PM cannot be visualized using the same PH domain. The Golgi pool of PI4P can be detected with the PH domain of OSBP and FAPP1, whose association with the Golgi membrane requires the small GTPase Arf1 [121,145]. In contrast, the localization of the PH domain from OSH2 is not related to the Arf1 binding so this domain is used to visualize the PM pool of PI4P [144]. The localization of PI4P in the late endosomes/lysosomes, in addition to Golgi and PM, was shown only recently with a novel PI4P probe that utilizes P4M domain from SidM protein. As mentioned above, SidM is a secreted effector protein of Legionella pneumophila that binds PI4P [118].
PI(4,5)P2 distribution can be visualized using the PH domain from PLCδ1 protein that specifically labels PM pools of this lipid. The PLCδ1-PH domain has been used to monitor PI(4,5)P2 hydrolysis and to investigate its role in phagocytosis, calcium-dependent exocytosis, and clathrin-mediated endocytosis [150]. As mentioned earlier, when PLC is activated, it cleaves PI(4,5)P2, producing DAG and IP3 [151]. This results in the translocation of the PLCδ1-PH-containing probe from the PM to the cytosol, which reveals a decrease in PM PI(4,5)P2 levels. However, because PLCδ1-PH binds to the inositol group of PI(4,5)P2, it is unclear whether this decrease is a result of a PI(4,5)P2 depletion or an IP3 increase. Therefore, to confirm the presence of not only PI(4,5)P2 but also other PIs and to provide conclusions about their intracellular localization and levels, the use of multiple probes combined with other techniques is recommended.
The Tubby domain is used to detect intracellular PI(4,5)P2, but it can also recognize PI(3,4)P2 and PI(3,4,5)P3 [147]. However, it must be noted that the probe containing the Tubby domain is suitable for visualization of PI(4,5)P2 because it is the only PI that is present at sufficient concentration to drive the localization of the probe [11].
Finally, the ENTH domain of the Epsin1 protein was recently developed to overcome the limitations of the PLCδ1-PH and Tubby domains. It has been shown that the ENTH domain binds to PI(4,5)P2 without binding to IP3, it has a low affinity for PI(4,5)P2 and does not interfere with its dynamics if overexpressed [148]. Therefore, it has been shown as a highly sensitive probe able to detect PLCβ-dependent PI(4,5)P2 depletion. Furthermore, the overexpression of this probe did not attenuate GqPCR signaling (unlike the PLCδ1-PH domain) making it suitable for the detection of minute changes in PI(4,5)P2 levels.

6.3. Mass Spectrometry

The development of mass spectrometry analyses allowed the growth of a new branch called lipidomics [152], which allowed systems-level analysis of lipids in different human diseases. For the analysis of PIs in the cell, researchers have used fast atom bombarded mass spectrometry (FAB-MS), matrix-assisted laser desorption and ionization/time of flight mass spectrometry (MALDI-TOF-MS), and electrospray ionization mass spectrometry (ESI-MS). The use of ESI-MS was shown to be valuable in the sense it can distinguish acyl chain contents of PIs and it is suitable only for the investigation of high abundant PIs [152], suggesting that all the lipids could not be extracted or identified [153]. Indeed, it was discovered that the extraction method for PIs is quite challenging. However, the optimization of the extraction methods and using tandem mass spectrometry analyses (MS-MS) yielded better results in identifying PIs, although it still could not resolve for spatial isomers, only for the number of phosphate groups [153,154]. Today, using the high-performance ion chromatography-coupled selected reaction monitoring mass spectrometry (IC-MS/MS) it is possible to resolve PIs positional isomers for their quantification in both tissues and cells [155]. Using this method, it has been shown in human platelets that upon activation with the collagen-related peptide the levels of PI4P and PI(4,5)P2 significantly increase implicating that both PIs have an important role in platelet activation [155]. These, and other new methods that will be developed in the future will allow for a detailed analysis of the structure, localization, and function of all PIs.

7. Conclusions

The numerous localizations and roles of PI4P and PI(4,5)P2 in the cell clearly show that they are irreplaceable and valuable components of the cell membranes. The fact that mutations in the kinases that produce PI4P and PI(4,5)P2 result in many diseases enhances the importance of these lipids in both physiology and pathology. This drives the development of new and improved methods for their visualization and quantification that aim to clarify the exact localization and function of PI4P and PI(4,5)P2 with their kinases, which is of great importance for the identification of target molecules and the development of new drugs.

Author Contributions

Conceptualization, A.B. and A.J.B.; writing—original draft preparation, A.B., S.Č., I.Đ., T.B. and A.J.B.; writing—review and editing, A.B., S.Č. and A.J.B.; project administration, funding acquisition, A.J.B. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the University of Rijeka, support grant: 18-188-1343; a Croatian Science Foundation (ESF-HRZZ-01-2018), and an American Society of Hematology (ASH) Global Research Award (A.J.B.).

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. Cooper, G.M. The Cell: A Molecular Approach, 8th ed.; Oxford University Press: New York, NY, USA, 2019; p. 783. [Google Scholar]
  2. Dowham, W. Advances in Lipobiology; Gross, R.W., Ed.; Jai Press Inc.: London, UK, 1997; Volume 2, p. 355. [Google Scholar]
  3. Sun, Y.; Thapa, N.; Hedman, A.C.; Anderson, R.A. Phosphatidylinositol 4,5-bisphosphate: Targeted production and signaling. Bioessays 2013, 35, 513–522. [Google Scholar] [CrossRef] [PubMed]
  4. Mandal, K. Review of PIP2 in Cellular Signaling, Functions and Diseases. Int. J. Mol. Sci. 2020, 21, 8342. [Google Scholar] [CrossRef] [PubMed]
  5. Katan, M.; Cockcroft, S. Phosphatidylinositol(4,5)bisphosphate: Diverse functions at the plasma membrane. Essays Biochem. 2020, 64, 513–531. [Google Scholar] [CrossRef] [PubMed]
  6. Suetsugu, S.; Kurisu, S.; Takenawa, T. Dynamic shaping of cellular membranes by phospholipids and membrane-deforming proteins. Physiol. Rev. 2014, 94, 1219–1248. [Google Scholar] [CrossRef]
  7. Wu, C.-Y.; Lin, M.-W.; Wu, D.-C.; Huang, Y.-B.; Huang, H.-T.; Chen, C.-L. The role of phosphoinositide-regulated actin reorganization in chemotaxis and cell migration. Br. J. Pharmacol. 2014, 171, 5541–5554. [Google Scholar] [CrossRef]
  8. Bout, I.V.D.; Divecha, N. PIP5K-driven PtdIns(4,5)P2 synthesis: Regulation and cellular functions. J. Cell Sci. 2009, 122, 3837–3850. [Google Scholar] [CrossRef]
  9. Izard, T.; Brown, D.T. Mechanisms and Functions of Vinculin Interactions with Phospholipids at Cell Adhesion Sites. J. Biol. Chem. 2016, 291, 2548–2555. [Google Scholar] [CrossRef]
  10. Hammond, G.R.; Burke, J.E. Novel roles of phosphoinositides in signaling, lipid transport, and disease. Curr. Opin. Cell Biol. 2020, 63, 57–67. [Google Scholar] [CrossRef]
  11. Dickson, E.J.; Hille, B. Understanding phosphoinositides: Rare, dynamic, and essential membrane phospholipids. Biochem. J. 2019, 476, 1–23. [Google Scholar] [CrossRef]
  12. Pemberton, J.G.; Kim, Y.J.; Balla, T. Integrated regulation of the phosphatidylinositol cycle and phosphoinositide-driven lipid transport at ER-PM contact sites. Traffic 2019, 21, 200–219. [Google Scholar] [CrossRef]
  13. Castano, E.; Yildirim, S.; Fáberová, V.; Krausová, A.; Uličná, L.; Paprčková, D.; Sztacho, M.; Hozák, P. Nuclear Phosphoinositides—Versatile Regulators of Genome Functions. Cells 2019, 8, 649. [Google Scholar] [CrossRef] [PubMed]
  14. Olivença, D.V.; Uliyakina, I.; Fonseca, L.L.; Amaral, M.D.; Voit, E.O.; Pinto, F.R. A Mathematical Model of the Phosphoinositide Pathway. Sci. Rep. 2018, 8, 3904. [Google Scholar] [CrossRef] [PubMed]
  15. Burke, J.E.; Triscott, J.; Emerling, B.M.; Hammond, G.R.V. Beyond PI3Ks: Targeting phosphoinositide kinases in disease. Nat. Rev. Drug Discov. 2022, 22, 357–386. [Google Scholar] [CrossRef] [PubMed]
  16. Raghu, P.; Joseph, A.; Krishnan, H.; Singh, P.; Saha, S. Phosphoinositides: Regulators of Nervous System Function in Health and Disease. Front. Mol. Neurosci. 2019, 12, 208. [Google Scholar] [CrossRef] [PubMed]
  17. Ernest James Phillips, T.; Maguire, E. Phosphoinositides: Roles in the Development of Microglial-Mediated Neuroinflammation and Neurodegeneration. Front. Cell. Neurosci. 2021, 15, 652593. [Google Scholar] [CrossRef]
  18. Ijuin, T. Phosphoinositide phosphatases in cancer cell dynamics—Beyond PI3K and PTEN. Semin. Cancer Biol. 2019, 59, 50–65. [Google Scholar] [CrossRef]
  19. Balla, T. Phosphoinositides: Tiny Lipids with Giant Impact on Cell Regulation. Physiol. Rev. 2013, 93, 1019–1137. [Google Scholar] [CrossRef]
  20. Beziau, A.; Brand, D.; Piver, E. The Role of Phosphatidylinositol Phosphate Kinases during Viral Infection. Viruses 2020, 12, 1124. [Google Scholar] [CrossRef]
  21. O’donnell, V.B.; Rossjohn, J.; Wakelam, M.J. Phospholipid signaling in innate immune cells. J. Clin. Investig. 2018, 128, 2670–2679. [Google Scholar] [CrossRef]
  22. Okkenhaug, K. Signaling by the phosphoinositide 3-kinase family in immune cells. Annu. Rev. Immunol. 2013, 31, 675–704. [Google Scholar] [CrossRef]
  23. Zewe, J.P.; Miller, A.M.; Sangappa, S.; Wills, R.C.; Goulden, B.D.; Hammond, G.R. Probing the subcellular distribution of phosphatidylinositol reveals a surprising lack at the plasma membrane. J. Cell Biol. 2020, 219, e201906127. [Google Scholar] [CrossRef]
  24. Desale, S.E.; Chinnathambi, S. Phosphoinositides signaling modulates microglial actin remodeling and phagocytosis in Alzheimer’s disease. Cell Commun. Signal. 2021, 19, 28. [Google Scholar] [CrossRef] [PubMed]
  25. Traynor-Kaplan, A.; Kruse, M.; Dickson, E.J.; Dai, G.; Vivas, O.; Yu, H.; Whittington, D.; Hille, B. Fatty-acyl chain profiles of cellular phosphoinositides. Biochim. Biophys. Acta Mol. Cell Biol. Lipids 2017, 1862, 513–522. [Google Scholar] [CrossRef]
  26. Nakatsu, F.; Baskin, J.M.; Chung, J.; Tanner, L.B.; Shui, G.; Lee, S.Y.; Pirruccello, M.; Hao, M.; Ingolia, N.T.; Wenk, M.R.; et al. PtdIns4P synthesis by PI4KIIIα at the plasma membrane and its impact on plasma membrane identity. J. Cell Biol. 2012, 199, 1003–1016. [Google Scholar] [CrossRef] [PubMed]
  27. Posor, Y.; Jang, W.; Haucke, V. Phosphoinositides as membrane organizers. Nat. Rev. Mol. Cell Biol. 2022, 23, 797–816. [Google Scholar] [CrossRef] [PubMed]
  28. Hsu, F.; Mao, Y. The structure of phosphoinositide phosphatases: Insights into substrate specificity and catalysis. Biochim. Biophys. Acta 2015, 1851, 698–710. [Google Scholar] [CrossRef]
  29. Ramos, A.R.; Ghosh, S.; Erneux, C. The impact of phosphoinositide 5-phosphatases on phosphoinositides in cell function and human disease. J. Lipid Res. 2019, 60, 276–286. [Google Scholar] [CrossRef]
  30. Minogue, S.; Waugh, M.G.; De Matteis, M.A.; Stephens, D.J.; Berditchevski, F.; Hsuan, J.J. Phosphatidylinositol 4-kinase is required for endosomal trafficking and degradation of the EGF receptor. J. Cell Sci. 2006, 119, 571–581. [Google Scholar] [CrossRef]
  31. Dornan, G.L.; McPhail, J.A.; Burke, J.E. Type III phosphatidylinositol 4 kinases: Structure, function, regulation, signalling and involvement in disease. Biochem. Soc. Trans. 2016, 44, 260–266. [Google Scholar] [CrossRef]
  32. Boura, E.; Nencka, R. Phosphatidylinositol 4-kinases: Function, structure, and inhibition. Exp. Cell Res. 2015, 337, 136–145. [Google Scholar] [CrossRef]
  33. Li, Y.-P.; Mikrani, R.; Hu, Y.-F.; Faran Ashraf Baig, M.M.; Abbas, M.; Akhtar, F.; Xu, M. Research progress of phosphatidylinositol 4-kinase and its inhibitors in inflammatory diseases. Eur. J. Pharmacol. 2021, 907, 174300. [Google Scholar] [CrossRef] [PubMed]
  34. Burke, J.E. Structural Basis for Regulation of Phosphoinositide Kinases and Their Involvement in Human Disease. Mol. Cell 2018, 71, 653–673. [Google Scholar] [CrossRef] [PubMed]
  35. Tan, J.; Brill, J.A. Cinderella story: PI4P goes from precursor to key signaling molecule. Crit. Rev. Biochem. Mol. Biol. 2013, 49, 33–58. [Google Scholar] [CrossRef] [PubMed]
  36. Klíma, M.; Baumlova, A.; Chalupska, M.; Hřebabecký, H.; Dejmek, M.; Nencka, R.; Boura, E. The high-resolution crystal structure of phosphatidylinositol 4-kinase IIβ and the crystal structure of phosphatidylinositol 4-kinase IIα containing a nucleoside analogue provide a structural basis for isoform-specific inhibitor design. Acta Crystallogr. D Biol. Crystallogr. 2015, 71, 1555–1563. [Google Scholar] [CrossRef] [PubMed]
  37. Wei, Y.J.; Sun, H.Q.; Yamamoto, M.; Wlodarski, P.; Kunii, K.; Martinez, M.; Barylko, B.; Albanesi, J.P.; Yin, H.L. Type II phosphatidylinositol 4-kinase β is a cytosolic and peripheral membrane protein that is recruited to the plasma membrane and activated by Rac-GTP. J. Biol. Chem. 2002, 277, 46586–46593. [Google Scholar] [CrossRef]
  38. Waugh, M.G. The Great Escape: How phosphatidylinositol 4-kinases and PI4P promote vesicle exit from the Golgi (and drive cancer). Biochem. J. 2019, 476, 2321–2346. [Google Scholar] [CrossRef]
  39. De Matteis, M.A.; Wilson, C.; D’Angelo, G. Phosphatidylinositol-4-phosphate: The Golgi and beyond. Bioessays 2013, 35, 612–622. [Google Scholar] [CrossRef]
  40. Heath, C.M.; Stahl, P.D.; Barbieri, M. Lipid kinases play crucial and multiple roles in membrane trafficking and signaling. Histol. Histopathol. 2003, 18, 989–998. [Google Scholar]
  41. Craige, B.; Salazar, G.; Faundez, V. Phosphatidylinositol-4-kinase type II alpha contains an ap-3–sorting motif and a kinase domain that are both required for endosome traffic. Mol. Biol. Cell 2008, 19, 1415–1426. [Google Scholar] [CrossRef]
  42. Pataer, A.; Ozpolat, B.; Shao, R.; Cashman, N.R.; Plotkin, S.S.; Samuel, C.E.; Lin, S.H.; Kabil, N.N.; Wang, J.; Majidi, M.; et al. Therapeutic targeting of the PI4K2A/PKR lysosome network is critical for misfolded protein clearance and survival in cancer cells. Oncogene 2020, 39, 801–813. [Google Scholar] [CrossRef]
  43. Palamiuc, L.; Ravi, A.; Emerling, B.M. Phosphoinositides in autophagy: Current roles and future insights. FEBS J. 2019, 287, 222–238. [Google Scholar] [CrossRef]
  44. Wang, H.; Sun, H.-Q.; Zhu, X.; Zhang, L.; Albanesi, J.; Levine, B.; Yin, H. GABARAPs regulate PI4P-dependent autophagosome:lysosome fusion. Proc. Natl. Acad. Sci. USA 2015, 112, 7015–7020. [Google Scholar] [CrossRef] [PubMed]
  45. Baba, T.; Balla, T. Emerging roles of phosphatidylinositol 4-phosphate and phosphatidylinositol 4,5-bisphosphate as regulators of multiple steps in autophagy. J. Biochem. 2020, 168, 329–336. [Google Scholar] [CrossRef] [PubMed]
  46. Tan, J.X.; Finkel, T. A phosphoinositide signalling pathway mediates rapid lysosomal repair. Nature 2022, 609, 815–821. [Google Scholar] [CrossRef] [PubMed]
  47. Alli-Balogun, G.O.; Gewinner, C.A.; Jacobs, R.; Kriston-Vizi, J.; Waugh, M.; Minogue, S. Phosphatidylinositol 4-kinase IIβ negatively regulates invadopodia formation and suppresses an invasive cellular phenotype. Mol. Biol. Cell 2016, 27, 4033–4042. [Google Scholar] [CrossRef]
  48. Li, J.; Gao, Z.; Zhao, D.; Zhang, L.; Qiao, X.; Zhao, Y.; Ding, H.; Zhang, P.; Lu, J.; Liu, J.; et al. PI-273, a Substrate-Competitive, Specific Small-Molecule Inhibitor of PI4KIIα, Inhibits the Growth of Breast Cancer Cells. Cancer Res. 2017, 77, 6253–6266. [Google Scholar] [CrossRef] [PubMed]
  49. Li, J.; Lu, Y.; Zhang, J.; Kang, H.; Qin, Z.; Chen, C. PI4KIIα is a novel regulator of tumor growth by its action on angiogenesis and HIF-1α regulation. Oncogene 2010, 29, 2550–2559. [Google Scholar] [CrossRef] [PubMed]
  50. Banerji, S.; Ngo, M.; Lane, C.F.; Robinson, C.-A.; Minogue, S.; Ridgway, N.D. Oxysterol binding protein-dependent activation of sphingomyelin synthesis in the golgi apparatus requires phosphatidylinositol 4-kinase IIα. Mol. Biol. Cell 2010, 21, 4141–4150. [Google Scholar] [CrossRef]
  51. Li, J.; Zhang, L.; Gao, Z.; Kang, H.; Rong, G.; Zhang, X.; Chen, C. Dual inhibition of EGFR at protein and activity level via combinatorial blocking of PI4KIIα as anti-tumor strategy. Protein Cell 2014, 5, 457–468. [Google Scholar] [CrossRef]
  52. Chu, K.M.; Minogue, S.; Hsuan, J.J.; Waugh, M.G. Differential effects of the phosphatidylinositol 4-kinases, PI4KIIα and PI4KIIIβ, on Akt activation and apoptosis. Cell Death Dis. 2010, 1, e106. [Google Scholar] [CrossRef]
  53. Sigismund, S.; Avanzato, D.; Lanzetti, L. Emerging functions of theEGFRin cancer. Mol. Oncol. 2018, 12, 3–20. [Google Scholar] [CrossRef] [PubMed]
  54. Go, G.; Lee, S.H. The Cellular Prion Protein: A Promising Therapeutic Target for Cancer. Int. J. Mol. Sci. 2020, 21, 9208. [Google Scholar] [CrossRef] [PubMed]
  55. Tariq, K.; Luikart, B.W. Striking a balance: PIP2 and PIP3 signaling in neuronal health and disease. Explor. Neuroprot. Ther. 2021, 1, 86–100. [Google Scholar] [CrossRef] [PubMed]
  56. Simons, J.P.; Al-Shawi, R.; Minogue, S.; Waugh, M.G.; Wiedemann, C.; Evangelou, S.; Loesch, A.; Sihra, T.S.; King, R.; Warner, T.T.; et al. Loss of phosphatidylinositol 4-kinase 2α activity causes late onset degeneration of spinal cord axons. Proc. Natl. Acad. Sci. USA 2009, 106, 11535–11539. [Google Scholar] [CrossRef]
  57. Robinson, J.W.; Leshchyns’ka, I.; Farghaian, H.; Hughes, W.E.; Sytnyk, V.; Neely, G.G.; Cole, A.R. PI4KIIα phosphorylation by GSK3 directs vesicular trafficking to lysosomes. Biochem. J. 2014, 464, 145–156. [Google Scholar] [CrossRef]
  58. Kang, M.S.; Baek, S.-H.; Chun, Y.S.; Moore, A.Z.; Landman, N.; Berman, D.; Yang, H.O.; Morishima-Kawashima, M.; Osawa, S.; Funamoto, S.; et al. modulation of lipid kinase Pi4kIIα activity and lipid raft association of presenilin 1 underlies γ-secretase inhibition by ginsenoside (20S)-Rg3. J. Biol. Chem. 2013, 288, 20868–20882. [Google Scholar] [CrossRef]
  59. Dafsari, H.S.; Pemberton, J.G.; Ferrer, E.A.; Yammine, T.; Farra, C.; Mohammadi, M.H.; Ghayoor Karimiani, E.; Hashemi, N.; Souaid, M.; Sabbagh, S.; et al. PI4K2A deficiency causes innate error in intracellular trafficking with developmental and epileptic-dyskinetic encephalopathy. Ann. Clin. Transl. Neurol. 2022, 9, 1345–1358. [Google Scholar] [CrossRef]
  60. Mohamed, M.; Gardeitchik, T.; Balasubramaniam, S.; Guerrero-Castillo, S.; Dalloyaux, D.; van Kraaij, S.; Venselaar, H.; Hoischen, A.; Urban, Z.; Brandt, U.; et al. Novel defect in phosphatidylinositol 4-kinase type 2-alpha (PI4K2A) at the membrane-enzyme interface is associated with metabolic cutis laxa. J. Inherit. Metab. Dis. 2020, 43, 1382–1391. [Google Scholar] [CrossRef]
  61. Zhang, L.; Li, J.; Zhang, P.; Gao, Z.; Zhao, Y.; Qiao, X.; Chen, C. PI4KIIα regulates insulin secretion and glucose homeostasis via a PKD-dependent pathway. Biophys. Rep. 2018, 4, 25–38. [Google Scholar] [CrossRef]
  62. Jović, M.; Kean, M.J.; Szentpetery, Z.; Polevoy, G.; Gingras, A.-C.; Brill, J.A.; Balla, T. Two phosphatidylinositol 4-kinases control lysosomal delivery of the Gaucher disease enzyme, β-glucocerebrosidase. Mol. Biol. Cell 2012, 23, 1533–1545. [Google Scholar] [CrossRef]
  63. Salazar, G.; Zlatic, S.; Craige, B.; Peden, A.A.; Pohl, J.; Faundez, V. Hermansky-pudlak syndrome protein complexes associate with phosphatidylinositol 4-kinase type II α in neuronal and non-neuronal cells. J. Biol. Chem. 2009, 284, 1790–1802. [Google Scholar] [CrossRef] [PubMed]
  64. Ozdemir, N.; Çelik, E.; Baslar, Z.; Celkan, T.T. A rare cause of thrombocyte dysfunction: Hermansky-Pudlak syndrome. Türk Pediatri Arşivi 2014, 49, 163–166. [Google Scholar] [CrossRef] [PubMed]
  65. Lopes da Silva, M.; O’Connor, M.N.; Kriston-Vizi, J.; White, I.J.; Al-Shawi, R.; Simons, J.P.; Mössinger, J.; Haucke, V.; Cutler, D.F. Type II PI4-kinases control Weibel-Palade body biogenesis and von Willebrand factor structure in Human endothelial cells. J. Cell Sci. 2016, 129, 2096–2105. [Google Scholar] [CrossRef]
  66. Zhou, Q.; Li, J.; Yu, H.; Zhai, Y.; Gao, Z.; Liu, Y.; Pang, X.; Zhang, L.; Schulten, K.; Sun, F.; et al. Molecular insights into the membrane-associated phosphatidylinositol 4-kinase IIα. Nat. Commun. 2014, 5, 3552. [Google Scholar] [CrossRef] [PubMed]
  67. Shi, L.; Tan, X.; Liu, X.; Yu, J.; Bota-Rabassedas, N.; Niu, Y.; Luo, J.; Xi, Y.; Zong, C.; Creighton, C.J.; et al. Addiction to Golgi-resident PI4P synthesis in chromosome 1q21.3–amplified lung adenocarcinoma cells. Proc. Natl. Acad. Sci. USA 2021, 118, e2023537118. [Google Scholar] [CrossRef]
  68. Bojjireddy, N.; Botyanszki, J.; Hammond, G.; Creech, D.; Peterson, R.; Kemp, D.C.; Snead, M.; Brown, R.; Morrison, A.; Wilson, S.; et al. Pharmacological and genetic targeting of the pi4ka enzyme reveals its important role in maintaining plasma membrane phosphatidylinositol 4-phosphate and phosphatidylinositol 4,5-bisphosphate levels. J. Biol. Chem. 2014, 289, 6120–6132. [Google Scholar] [CrossRef]
  69. Dornan, G.L.; Dalwadi, U.; Hamelin, D.J.; Hoffmann, R.M.; Yip, C.K.; Burke, J.E. Probing the Architecture, Dynamics, and Inhibition of the PI4KIIIα/TTC7/FAM126 Complex. J. Mol. Biol. 2018, 430, 3129–3142. [Google Scholar] [CrossRef]
  70. Jiang, X.; Huang, X.; Zheng, G.; Jia, G.; Li, Z.; Ding, X.; Lei, L.; Yuan, L.; Xu, S.; Gao, N. Targeting PI4KA sensitizes refractory leukemia to chemotherapy by modulating the ERK/AMPK/OXPHOS axis. Theranostics 2022, 12, 6972–6988. [Google Scholar] [CrossRef]
  71. Dorobantu, C.M.; Harak, C.; Klein, R.; van der Linden, L.; Strating, J.R.P.M.; van der Schaar, H.M.; Lohmann, V.; van Kuppeveld, F.J.M. Tyrphostin AG1478 Inhibits Encephalomyocarditis Virus and Hepatitis C Virus by Targeting Phosphatidylinositol 4-Kinase IIIα. Antimicrob. Agents Chemother. 2016, 60, 6402–6406. [Google Scholar] [CrossRef]
  72. Zhang, Z.; Venditti, R.; Ran, L.; Liu, Z.; Vivot, K.; Schürmann, A.; Bonifacino, J.S.; De Matteis, M.A.; Ricci, R. Distinct changes in endosomal composition promote NLRP3 inflammasome activation. Nat. Immunol. 2023, 24, 30–41. [Google Scholar] [CrossRef]
  73. Guo, Z.; Jiang, C.-H.; Tong, C.; Yang, Y.; Wang, Z.; Lam, S.M.; Wang, D.; Li, R.; Shui, G.; Shi, Y.S.; et al. Activity-dependent PI4P synthesis by PI4KIIIα regulates long-term synaptic potentiation. Cell Rep. 2022, 38, 110452. [Google Scholar] [CrossRef] [PubMed]
  74. Alvarez-Prats, A.; Bjelobaba, I.; Aldworth, Z.; Baba, T.; Abebe, D.; Kim, Y.J.; Stojilkovic, S.S.; Stopfer, M.; Balla, T. Schwann-Cell-Specific Deletion of Phosphatidylinositol 4-Kinase Alpha Causes Aberrant Myelination. Cell Rep. 2018, 23, 2881–2890. [Google Scholar] [CrossRef] [PubMed]
  75. Govindarajan, B.; Sbrissa, D.; Pressprich, M.; Kim, S.; Vaishampayan, U.; Cher, M.L.; Chinni, S. Adaptor proteins mediate CXCR4 and PI4KA crosstalk in prostate cancer cells and the significance of PI4KA in bone tumor growth. Res. Sq. 2023, 1–13. [Google Scholar] [CrossRef]
  76. Rutaganira, F.U.; Fowler, M.L.; McPhail, J.A.; Gelman, M.A.; Nguyen, K.; Xiong, A.; Dornan, G.L.; Tavshanjian, B.; Glenn, J.S.; Shokat, K.M.; et al. Design and Structural Characterization of Potent and Selective Inhibitors of Phosphatidylinositol 4 Kinase IIIβ. J. Med. Chem. 2016, 59, 1830–1839. [Google Scholar] [CrossRef] [PubMed]
  77. Arita, M. Phosphatidylinositol-4 kinase III beta and oxysterol-binding protein accumulate unesterified cholesterol on poliovirus-induced membrane structure. Microbiol. Immunol. 2014, 58, 239–256. [Google Scholar] [CrossRef]
  78. Spickler, C.; Lippens, J.; Laberge, M.-K.; Desmeules, S.; Bellavance, E.; Garneau, M.; Guo, T.; Hucke, O.; Leyssen, P.; Neyts, J.; et al. Phosphatidylinositol 4-kinase III beta is essential for replication of human rhinovirus and its inhibition causes a lethal phenotype in vivo. Antimicrob. Agents Chemother. 2013, 57, 3358–3368. [Google Scholar] [CrossRef]
  79. Strating, J.R.; van der Linden, L.; Albulescu, L.; Bigay, J.; Arita, M.; Delang, L.; Leyssen, P.; van der Schaar, H.M.; Lanke, K.H.; Thibaut, H.J.; et al. Itraconazole inhibits enterovirus replication by targeting the oxysterol-binding protein. Cell Rep. 2015, 10, 600–615. [Google Scholar] [CrossRef]
  80. van der Schaar, H.M.; Leyssen, P.; Thibaut, H.J.; de Palma, A.; van der Linden, L.; Lanke, K.H.W.; Lacroix, C.; Verbeken, E.; Conrath, K.; MacLeod, A.M.; et al. A novel, broad-spectrum inhibitor of enterovirus replication that targets host cell factor phosphatidylinositol 4-kinase IIIβ. Antimicrob. Agents Chemother. 2013, 57, 4971–4981. [Google Scholar] [CrossRef]
  81. Phillpotts, R.J.; Wallace, J.; Tyrrell, D.; Tagart, V.B. Therapeutic activity of enviroxime against rhinovirus infection in volunteers. Antimicrob. Agents Chemother. 1983, 23, 671–675. [Google Scholar] [CrossRef]
  82. Paquet, T.; Le Manach, C.; Cabrera, D.G.; Younis, Y.; Henrich, P.P.; Abraham, T.S.; Lee, M.C.S.; Basak, R.; Ghidelli-Disse, S.; Lafuente-Monasterio, M.J.; et al. Antimalarial efficacy of MMV390048, an inhibitor of Plasmodium phosphatidylinositol 4-kinase. Sci. Transl. Med. 2017, 9, eaad9735. [Google Scholar] [CrossRef]
  83. Mohammed, R.; Asres, M.S.; Gudina, E.K.; Adissu, W.; Johnstone, H.; Marrast, A.C.; Donini, C.; Duparc, S.; Yilma, D. Efficacy, Safety, Tolerability, and Pharmacokinetics of MMV390048 in Acute Uncomplicated Malaria. Am. J. Trop. Med. Hyg. 2023, 108, 81–84. [Google Scholar] [CrossRef]
  84. Kunkl, M.; Porciello, N.; Mastrogiovanni, M.; Capuano, C.; Lucantoni, F.; Moretti, C.; Persson, J.L.; Galandrini, R.; Buzzetti, R.; Tuosto, L. ISA-2011B, a Phosphatidylinositol 4-Phosphate 5-Kinase α Inhibitor, Impairs CD28-Dependent Costimulatory and Pro-inflammatory Signals in Human T Lymphocytes. Front. Immunol. 2017, 8, 502. [Google Scholar] [CrossRef] [PubMed]
  85. Sarwar, M.; Semenas, J.; Miftakhova, R.; Simoulis, A.; Robinson, B.; Gjörloff Wingren, A.; Mongan, N.P.; Heery, D.M.; Johnson, H.; Abrahamsson, P.-A.; et al. Targeted suppression of AR-V7 using PIP5K1α inhibitor overcomes enzalutamide resistance in prostate cancer cells. Oncotarget 2016, 7, 63065–63081. [Google Scholar] [CrossRef]
  86. Semenas, J.; Hedblom, A.; Miftakhova, R.R.; Sarwar, M.; Larsson, R.; Shcherbina, L.; Johansson, M.E.; Härkönen, P.; Sterner, O.; Persson, J.L. The role of PI3K/AKT-related PIP5K1α and the discovery of its selective inhibitor for treatment of advanced prostate cancer. Proc. Natl. Acad. Sci. USA 2014, 111, E3689–E3698. [Google Scholar] [CrossRef] [PubMed]
  87. Peng, W.; Huang, W.; Ge, X.; Xue, L.; Zhao, W.; Xue, J. Type Iγ phosphatidylinositol phosphate kinase promotes tumor growth by facilitating Warburg effect in colorectal cancer. Ebiomedicine 2019, 44, 375–386. [Google Scholar] [CrossRef]
  88. Wieffer, M.; Cibrian Uhalte, E.; Posor, Y.; Otten, C.; Branz, K.; Schütz, I.; Mössinger, J.; Schu, P.; Abdelilah-Seyfried, S.; Krauß, M.; et al. PI4K2β/AP-1-Based TGN-Endosomal Sorting Regulates Wnt Signaling. Curr. Biol. 2013, 23, 2185–2190. [Google Scholar] [CrossRef] [PubMed]
  89. Sasaki, T.; Takasuga, S.; Sasaki, J.; Kofuji, S.; Eguchi, S.; Yamazaki, M.; Suzuki, A. Mammalian phosphoinositide kinases and phosphatases. Prog. Lipid Res. 2009, 48, 307–343. [Google Scholar] [CrossRef]
  90. Sobocińska, J.; Roszczenko-Jasinska, P.; Zaręba-Kozioł, M.; Hromada-Judycka, A.; Matveichuk, O.; Traczyk, G.; Łukasiuk, K.; Kwiatkowska, K. Lipopolysaccharide Upregulates Palmitoylated Enzymes of the Phosphatidylinositol Cycle: An Insight from Proteomic Studies. Mol. Cell. Proteom. 2018, 17, 233–254. [Google Scholar] [CrossRef]
  91. Misehe, M.; Klima, M.; Matoušová, M.; Chalupská, D.; Dejmek, M.; Šála, M.; Mertlíková-Kaiserová, H.; Boura, E.; Nencka, R. Structure-based design and modular synthesis of novel PI4K class II inhibitors bearing a 4-aminoquinazoline scaffold. Bioorg. Med. Chem. Lett. 2022, 76, 129010. [Google Scholar] [CrossRef]
  92. Clayton, E.L.; Minogue, S.; Waugh, M.G. Mammalian phosphatidylinositol 4-kinases as modulators of membrane trafficking and lipid signaling networks. Prog. Lipid Res. 2013, 52, 294–304. [Google Scholar] [CrossRef] [PubMed]
  93. Balla, A.; Kim, Y.J.; Varnai, P.; Szentpetery, Z.; Knight, Z.; Shokat, K.M.; Balla, T. Maintenance of hormone-sensitive phosphoinositide pools in the plasma membrane requires phosphatidylinositol 4-kinase IIIα. Mol. Biol. Cell 2008, 19, 711–721. [Google Scholar] [CrossRef] [PubMed]
  94. Jardine, S.; Dhingani, N.; Muise, A.M. TTC7A: Steward of Intestinal Health. Cell. Mol. Gastroenterol. Hepatol. 2019, 7, 555–570. [Google Scholar] [CrossRef] [PubMed]
  95. Park, Y.; Park, J.M.; Kim, D.H.; Kim, I.A. Inhibition of PI4K IIIα radiosensitizes in human tumor xenograft and immune-competent syngeneic murine tumor model. Oncotarget 2017, 8, 110392–110405. [Google Scholar] [CrossRef]
  96. Ziyad, S.; Riordan, J.D.; Cavanaugh, A.M.; Su, T.; Hernandez, G.E.; Hilfenhaus, G.; Morselli, M.; Huynh, K.; Wang, K.; Chen, J.-N.; et al. A Forward Genetic Screen Targeting the Endothelium Reveals a Regulatory Role for the Lipid Kinase Pi4ka in Myelo- and Erythropoiesis. Cell Rep. 2018, 22, 1211–1224. [Google Scholar] [CrossRef]
  97. Adhikari, H.; Kattan, W.E.; Kumar, S.; Zhou, P.; Hancock, J.F.; Counter, C.M. Oncogenic KRAS is dependent upon an EFR3A-PI4KA signaling axis for potent tumorigenic activity. Nat. Commun. 2021, 12, 5248. [Google Scholar] [CrossRef] [PubMed]
  98. Ilboudo, A.; Nault, J.-C.; Dubois-Pot-Schneider, H.; Corlu, A.; Zucman-Rossi, J.; Samson, M.; Le Seyec, J. Overexpression of phosphatidylinositol 4-kinase type IIIα is associated with undifferentiated status and poor prognosis of human hepatocellular carcinoma. BMC Cancer 2014, 14, 7. [Google Scholar] [CrossRef] [PubMed]
  99. Pagnamenta, A.T.; Howard, M.F.; Wisniewski, E.; Popitsch, N.; Knight, S.J.; Keays, D.A.; Quaghebeur, G.; Cox, H.; Cox, P.; Balla, T.; et al. Germline recessive mutations in PI4KA are associated with perisylvian polymicrogyria, cerebellar hypoplasia and arthrogryposis. Hum. Mol. Genet. 2015, 24, 3732–3741. [Google Scholar] [CrossRef] [PubMed]
  100. Verdura, E.; Rodríguez-Palmero, A.; Vélez-Santamaria, V.; Planas-Serra, L.; de la Calle, I.; Raspall-Chaure, M.; Roubertie, A.; Benkirane, M.; Saettini, F.; Pavinato, L.; et al. Biallelic PI4KA variants cause a novel neurodevelopmental syndrome with hypomyelinating leukodystrophy. Brain 2021, 144, 2659–2669. [Google Scholar] [CrossRef]
  101. Morrow, A.A.; Alipour, M.A.; Bridges, D.; Yao, Z.; Saltiel, A.R.; Lee, J.M. The lipid kinase PI4KIIIβ is highly expressed in breast tumors and activates akt in cooperation with Rab11a. Mol. Cancer Res. 2014, 12, 1492–1508. [Google Scholar] [CrossRef]
  102. Yang, N.; Ma, P.; Lang, J.; Zhang, Y.; Deng, J.; Ju, X.; Zhang, G.; Jiang, C. Phosphatidylinositol 4-kinase IIIβ is required for severe acute respiratory syndrome coronavirus spike-mediated cell entry. J. Biol. Chem. 2012, 287, 8457–8467. [Google Scholar] [CrossRef]
  103. Kremer, L.; Hennes, M.S.E.; Brause, A.; Ursu, A.; Robke, L.; Matsubayashi, H.T.; Nihongaki, Y.; Flegel, M.S.J.; Mejdrová, I.; Eickhoff, J.; et al. Discovery of the Hedgehog Pathway Inhibitor Pipinib that Targets PI4KIIIß. Angew. Chem. Int. Ed. Engl. 2019, 58, 16617–16628. [Google Scholar] [CrossRef] [PubMed]
  104. Li, S.; Ghosh, C.; Xing, Y.; Sun, Y. Phosphatidylinositol 4,5-bisphosphate in the Control of Membrane Trafficking. Int. J. Biol. Sci. 2020, 16, 2761–2774. [Google Scholar] [CrossRef] [PubMed]
  105. Wang, Y.; Zhao, L.; Suzuki, A.; Lian, L.; Min, S.H.; Wang, Z.; Litvinov, R.I.; Stalker, T.J.; Yago, T.; Klopocki, A.G.; et al. Platelets lacking PIP5KIγ have normal integrin activation but impaired cytoskeletal-membrane integrity and adhesion. Blood 2013, 121, 2743–2752. [Google Scholar] [CrossRef]
  106. Doughman, R.L.; Firestone, A.; Anderson, R.A. Phosphatidylinositol phosphate kinases put PI4,5P 2 in its place. J. Membr. Biol. 2003, 194, 77–89. [Google Scholar] [CrossRef]
  107. East, M.P.; Laitinen, T.; Asquith, C.R.M. PIP5K1A: A potential target for cancers with KRAS or TP53 mutations. Nat. Rev. Drug Discov. 2020, 19, 436. [Google Scholar] [CrossRef] [PubMed]
  108. Wang, Y.; Chen, X.; Lian, L.; Tang, T.; Stalker, T.J.; Sasaki, T.; Kanaho, Y.; Brass, L.F.; Choi, J.K.; Hartwig, J.H.; et al. Loss of PIP5KIβ demonstrates that PIP5KI isoform-specific PIP2 synthesis is required for IP3 formation. Proc. Natl. Acad. Sci. USA 2008, 105, 14064–14069. [Google Scholar] [CrossRef]
  109. Sun, Y.; Turbin, D.; Ling, K.; Thapa, N.; Leung, S.; Huntsman, D.G.; Anderson, R. Type I gamma phosphatidylinositol phosphate kinase modulates invasion and proliferation and its expression correlates with poor prognosis in breast cancer. Breast Cancer Res. 2010, 12, R6. [Google Scholar] [CrossRef]
  110. Schill, N.J.; Hedman, A.C.; Choi, S.; Anderson, R.A. Isoform 5 of PIPKIgamma regulates the endosomal trafficking and degradation of E-cadherin. J. Cell Sci. 2014, 127, 2189–2203. [Google Scholar] [CrossRef]
  111. Choi, S.; Thapa, N.; Tan, X.; Hedman, A.C.; Anderson, R.A. PIP kinases define PI4,5P2 signaling specificity by association with effectors. Biochim. Biophys. Acta 2015, 1851, 711–723. [Google Scholar] [CrossRef]
  112. Di Paolo, G.; Moskowitz, H.S.; Gipson, K.; Wenk, M.R.; Voronov, S.; Obayashi, M.; Flavell, R.; Fitzsimonds, R.M.; Ryan, T.A.; De Camilli, P. Impaired PtdIns(4,5)P2 synthesis in nerve terminals produces defects in synaptic vesicle trafficking. Nature 2004, 431, 415–422. [Google Scholar] [CrossRef]
  113. Wang, Y.; Lian, L.; Golden, J.A.; Morrisey, E.E.; Abrams, C.S. PIP5KIγ is required for cardiovascular and neuronal development. Proc. Natl. Acad. Sci. USA 2007, 104, 11748–11753. [Google Scholar] [CrossRef] [PubMed]
  114. McCrea, H.J.; De Camilli, P. Mutations in phosphoinositide metabolizing enzymes and human disease. Physiology 2009, 24, 8–16. [Google Scholar] [CrossRef] [PubMed]
  115. Narkis, G.; Ofir, R.; Landau, D.; Manor, E.; Volokita, M.; Hershkowitz, R.; Elbedour, K.; Birk, O.S. Lethal contractural syndrome type 3 (LCCS3) is caused by a mutation in PIP5K1C, which encodes PIPKIγ of the phophatidylinsitol pathway. Am. J. Hum. Genet. 2007, 81, 530–539. [Google Scholar] [CrossRef] [PubMed]
  116. Wang, Y.; Litvinov, R.I.; Chen, X.; Bach, T.L.; Lian, L.; Petrich, B.G.; Monkley, S.J.; Kanaho, Y.; Critchley, D.R.; Sasaki, T.; et al. Loss of PIP5KIγ, unlike other PIP5KI isoforms, impairs the integrity of the membrane cytoskeleton in murine megakaryocytes. J. Clin. Investig. 2008, 118, 812–819. [Google Scholar] [CrossRef]
  117. Chung, J.; Torta, F.; Masai, K.; Lucast, L.; Czapla, H.; Tanner, L.B.; Narayanaswamy, P.; Wenk, M.R.; Nakatsu, F.; De Camilli, P. PI4P/phosphatidylserine countertransport at ORP5- and ORP8-mediated ER–plasma membrane contacts. Science 2015, 349, 428–432. [Google Scholar] [CrossRef]
  118. Hammond, G.R.; Machner, M.P.; Balla, T. A novel probe for phosphatidylinositol 4-phosphate reveals multiple pools beyond the Golgi. J. Cell Biol. 2014, 205, 113–126. [Google Scholar] [CrossRef]
  119. Wang, Y.J.; Wang, J.; Sun, H.Q.; Martinez, M.; Sun, Y.X.; Macia, E.; Kirchhausen, T.; Albanesi, J.P.; Roth, M.G.; Yin, H.L. Phosphatidylinositol 4 Phosphate Regulates Targeting of Clathrin Adaptor AP-1 Complexes to the Golgi. Cell 2003, 114, 299–310. [Google Scholar] [CrossRef]
  120. Wang, J.; Sun, H.-Q.; Macia, E.; Kirchhausen, T.; Watson, H.; Bonifacino, J.S.; Yin, H.L. PI4P promotes the recruitment of the GGA adaptor proteins to the trans-golgi network and regulates their recognition of the ubiquitin sorting signal. Mol. Biol. Cell 2007, 18, 2646–2655. [Google Scholar] [CrossRef]
  121. Godi, A.; DI Campli, A.; Konstantakopoulos, A.; Di Tullio, G.; Alessi, D.; Kular, G.S.; Daniele, T.; Marra, P.; Lucocq, J.; De Matteis, M.A. FAPPs control Golgi-to-cell-surface membrane traffic by binding to ARF and PtdIns(4)P. Nat. Cell Biol. 2004, 6, 393–404. [Google Scholar] [CrossRef]
  122. Hammond, G.R.V.; Schiavo, G.; Irvine, R.F. Immunocytochemical techniques reveal multiple, distinct cellular pools of PtdIns4P and PtdIns(4,5)P2. Biochem. J. 2009, 422, 23–35. [Google Scholar] [CrossRef]
  123. Cheong, F.Y.; Sharma, V.; Blagoveshchenskaya, A.; Oorschot, V.M.J.; Brankatschk, B.; Klumperman, J.; Freeze, H.H.; Mayinger, P. Spatial Regulation of Golgi Phosphatidylinositol-4-Phosphate is Required for Enzyme Localization and Glycosylation Fidelity. Traffic 2010, 11, 1180–1190. [Google Scholar] [CrossRef] [PubMed]
  124. Szentpetery, Z.; Várnai, P.; Balla, T. Acute manipulation of Golgi phosphoinositides to assess their importance in cellular trafficking and signaling. Proc. Natl. Acad. Sci. USA 2010, 107, 8225–8230. [Google Scholar] [CrossRef] [PubMed]
  125. Hammond, G.R.V.; Fischer, M.J.; Anderson, K.E.; Holdich, J.; Koteci, A.; Balla, T.; Irvine, R.F. PI4P and PI(4,5)P2 Are Essential But Independent Lipid Determinants of Membrane Identity. Science 2012, 337, 727–730. [Google Scholar] [CrossRef] [PubMed]
  126. Suh, B.-C.; Hille, B. Regulation of ion channels by phosphatidylinositol 4,5-bisphosphate. Curr. Opin. Neurobiol. 2005, 15, 370–378. [Google Scholar] [CrossRef]
  127. Mesmin, B.; Bigay, J.; von Filseck, J.M.; Lacas-Gervais, S.; Drin, G.; Antonny, B. A four-step cycle driven by PI(4)P hydrolysis directs sterol/PI(4)P exchange by the ER-Golgi tether OSBP. Cell 2013, 155, 830–843. [Google Scholar] [CrossRef]
  128. Prinz, W.A. Bridging the gap: Membrane contact sites in signaling, metabolism, and organelle dynamics. J. Cell Biol. 2014, 205, 759–769. [Google Scholar] [CrossRef]
  129. Dickson, E.J.; Jensen, J.B.; Vivas, O.; Kruse, M.; Traynor-Kaplan, A.E.; Hille, B. Dynamic formation of ER–PM junctions presents a lipid phosphatase to regulate phosphoinositides. J. Cell Biol. 2016, 213, 33–48. [Google Scholar] [CrossRef]
  130. Watt, S.A.; Kular, G.; Fleming, I.N.; Downes, P.; Lucocq, J.M. Subcellular localization of phosphatidylinositol 4,5-bisphosphate using the pleckstrin homology domain of phospholipase C δ. Biochem. J. 2002, 363, 657–666. [Google Scholar] [CrossRef]
  131. Cremona, O.; Di Paolo, G.; Wenk, M.R.; Lüthi, A.; Kim, W.T.; Takei, K.; Daniell, L.; Nemoto, Y.; Shears, S.B.; Flavell, R.; et al. Essential Role of Phosphoinositide Metabolism in Synaptic Vesicle Recycling. Cell 1999, 99, 179–188. [Google Scholar] [CrossRef]
  132. Sobol, M.; Yildirim, S.; Philimonenko, V.V.; Marášek, P.; Castaño, E.; Hozák, P. UBF complexes with phosphatidylinositol 4,5-bisphosphate in nucleolar organizer regions regardless of ongoing RNA polymerase I activity. Nucleus 2013, 4, 478–486. [Google Scholar] [CrossRef]
  133. Haucke, V. Phosphoinositide regulation of clathrin-mediated endocytosis. Biochem. Soc. Trans. 2005, 33, 1285–1289. [Google Scholar] [CrossRef] [PubMed]
  134. Padrón, D.; Wang, Y.J.; Yamamoto, M.; Yin, H.; Roth, M.G. Phosphatidylinositol phosphate 5-kinase Iβ recruits AP-2 to the plasma membrane and regulates rates of constitutive endocytosis. J. Cell Biol. 2003, 162, 693–701. [Google Scholar] [CrossRef] [PubMed]
  135. Vicinanza, M.; Di Campli, A.; Polishchuk, E.; Santoro, M.; Di Tullio, G.; Godi, A.; Levtchenko, E.; De Leo, M.G.; Polishchuk, R.; Sandoval, L.; et al. OCRL controls trafficking through early endosomes via PtdIns4,5P2-dependent regulation of endosomal actin. EMBO J. 2011, 30, 4970–4985. [Google Scholar] [CrossRef] [PubMed]
  136. Martin, T.F. Role of PI(4,5)P2 in vesicle exocytosis and membrane fusion. Subcell Biochem 2012, 59, 111–130. [Google Scholar] [CrossRef] [PubMed]
  137. Gamper, N.; Shapiro, M.S. Target-specific PIP2 signalling: How might it work? J. Physiol. 2007, 582, 967–975. [Google Scholar] [CrossRef]
  138. Suh, B.-C.; Hille, B. PIP2 is a necessary cofactor for ion channel function: How and why? Annu. Rev. Biophys. 2008, 37, 175–195. [Google Scholar] [CrossRef]
  139. Idevall-Hagren, O.; De Camilli, P. Detection and manipulation of phosphoinositides. Biochim. Biophys. Acta 2015, 1851, 736–745. [Google Scholar] [CrossRef]
  140. Kalasova, I.; Fáberová, V.; Kalendová, A.; Yildirim, S.; Uličná, L.; Venit, T.; Hozák, P. Tools for visualization of phosphoinositides in the cell nucleus. Histochem. Cell Biol. 2016, 145, 485–496. [Google Scholar] [CrossRef]
  141. Bura, A.; Jurak Begonja, A. Imaging of Intracellular and Plasma Membrane Pools of PI(4,5)P2 and PI4P in Human Platelets. Life 2021, 11, 1331. [Google Scholar] [CrossRef]
  142. Wills, R.C.; Goulden, B.D.; Hammond, G.R.V. Genetically encoded lipid biosensors. Mol. Biol. Cell 2018, 29, 1526–1532. [Google Scholar] [CrossRef]
  143. Ji, C.; Zhang, Y.; Xu, P.; Xu, T.; Lou, X. Nanoscale Landscape of Phosphoinositides Revealed by Specific Pleckstrin Homology (PH) Domains Using Single-molecule Superresolution Imaging in the Plasma Membrane. J. Biol. Chem. 2015, 290, 26978–26993. [Google Scholar] [CrossRef] [PubMed]
  144. Roy, A.; Levine, T. Multiple pools of phosphatidylinositol 4-phosphate detected using the pleckstrin homology domain of Osh2p. J. Biol. Chem. 2004, 279, 44683–44689. [Google Scholar] [CrossRef] [PubMed]
  145. Levine, T.P.; Munro, S. The pleckstrin homology domain of oxysterol-binding protein recognises a determinant specific to Golgi membranes. Curr. Biol. 1998, 8, 729–739. [Google Scholar] [CrossRef] [PubMed]
  146. Stauffer, T.P.; Ahn, S.; Meyer, T. Receptor-induced transient reduction in plasma membrane PtdIns(4,5)P2 concentration monitored in living cells. Curr. Biol. 1998, 8, 343–346. [Google Scholar] [CrossRef]
  147. Santagata, S.; Boggon, T.J.; Baird, C.L.; Gomez, C.A.; Zhao, J.; Shan, W.S.; Myszka, D.G.; Shapiro, L. G-Protein Signaling through Tubby Proteins. Science 2001, 292, 2041–2050. [Google Scholar] [CrossRef]
  148. Leitner, M.G.; Thallmair, V.; Wilke, B.U.; Neubert, V.; Kronimus, Y.; Halaszovich, C.R.; Oliver, D. The N-terminal homology (ENTH) domain of Epsin 1 is a sensitive reporter of physiological PI(4,5)P2 dynamics. Biochim. Biophys. Acta Mol. Cell Biol. Lipids 2018, 1864, 433–442. [Google Scholar] [CrossRef]
  149. Ford, M.G.J.; Pearse, B.M.F.; Higgins, M.K.; Vallis, Y.; Owen, D.J.; Gibson, A.; Hopkins, C.R.; Evans, F.R.; McMahon, H.T. Simultaneous Binding of PtdIns(4,5)P2 and Clathrin by AP180 in the Nucleation of Clathrin Lattices on Membranes. Science 2001, 291, 1051–1055. [Google Scholar] [CrossRef]
  150. Hllet, G. Imaging phosphoinositide dynamics using GFP-tagged protein domain. Biol. Cell 2005, 97, 501–518. [Google Scholar] [CrossRef]
  151. Putney, J.W.; Tomita, T. Phospholipase C signaling and calcium influx. Adv. Biol. Regul. 2012, 52, 152–164. [Google Scholar] [CrossRef]
  152. Kim, Y.; Shanta, S.R.; Zhou, L.-H.; Kim, K.P. Mass spectrometry based cellular phosphoinositides profiling and phospholipid analysis: A brief review. Exp. Mol. Med. 2010, 42, 1–11. [Google Scholar] [CrossRef]
  153. Wakelam, M.J.; Clark, J. Methods for analyzing phosphoinositides using mass spectrometry. Biochim. Biophys. Acta 2011, 1811, 758–762. [Google Scholar] [CrossRef] [PubMed]
  154. Kielkowska, A.; Niewczas, I.; Anderson, K.E.; Durrant, T.N.; Clark, J.; Stephens, L.R.; Hawkins, P.T. A new approach to measuring phosphoinositides in cells by mass spectrometry. Adv. Biol. Regul. 2014, 54, 131–141. [Google Scholar] [CrossRef] [PubMed]
  155. Cheung, H.Y.F.; Coman, C.; Westhoff, P.; Manke, M.; Sickmann, A.; Borst, O.; Gawaz, M.; Watson, S.P.; Heemskerk, J.W.M.; Ahrends, R. Targeted Phosphoinositides Analysis Using High-Performance Ion Chromatography-Coupled Selected Reaction Monitoring Mass Spectrometry. J. Proteome Res. 2021, 20, 3114–3123. [Google Scholar] [CrossRef] [PubMed]
Figure 1. The structure of PI4P and PI(4,5)P2. Phosphoinositides can be phosphorylated on positions three, four, and five. They consist of an inositol ring, glycerol, and fatty acids. (A) PI4P is phosphorylated on position four of the inositol ring. (B) PI(4,5)P2 is phosphorylated on positions four and five of the inositol ring.
Figure 1. The structure of PI4P and PI(4,5)P2. Phosphoinositides can be phosphorylated on positions three, four, and five. They consist of an inositol ring, glycerol, and fatty acids. (A) PI4P is phosphorylated on position four of the inositol ring. (B) PI(4,5)P2 is phosphorylated on positions four and five of the inositol ring.
Cells 12 01411 g001
Figure 2. Types of phosphoinositides. There are seven different phosphoinositides depending on the positions of phosphorylation of the inositol ring. Their production is controlled by kinases (green arrows) and phosphatases (blue arrows).
Figure 2. Types of phosphoinositides. There are seven different phosphoinositides depending on the positions of phosphorylation of the inositol ring. Their production is controlled by kinases (green arrows) and phosphatases (blue arrows).
Cells 12 01411 g002
Figure 3. Phosphoinositides in the cell. PI4P (in red) is localized at the Golgi apparatus, endoplasmic reticulum (ER), plasma membrane (PM) and late endosomes/lysosomes (LE/LYS). PI(4,5)P2 (in green) is localized at the PM, early endosomes (EE), ER and the nucleus. RE-recycling endosomes, MVB-multivesicular body.
Figure 3. Phosphoinositides in the cell. PI4P (in red) is localized at the Golgi apparatus, endoplasmic reticulum (ER), plasma membrane (PM) and late endosomes/lysosomes (LE/LYS). PI(4,5)P2 (in green) is localized at the PM, early endosomes (EE), ER and the nucleus. RE-recycling endosomes, MVB-multivesicular body.
Cells 12 01411 g003
Table 1. Inhibitors of type II, type III PI4Ks, and PI4P5KIs.
Table 1. Inhibitors of type II, type III PI4Ks, and PI4P5KIs.
Target EnzymeInhibitorTarget
IC50 (nM)
Off-Targets,
IC50 (nM)
State of Clinical DevelopmentResearch AreaReferences
PI4KIIαPI-273470
[48]
no off-targets,
highly selective
in vitrocancer[48,67]
in vivo[48]
PI4KIIIαGSK-A13.1
[15]
>310
[15]
in vitrobasic[68,69]
cancer[70]
viral infection[71]
inflammation[72]
neuronal plasticity[73]
myelination[74]
GSK-F116
[15]
PI4KIIIβ, PI3Ks
>1600
[15]
in vitrocancer[75]
in vivobasic[68]
PI4KIIIβIN-97
[76]
PI4KIIIα, PI3Ks
>150
[76]
in vitrocancer[67]
inflammation[72]
IN-103.6
[15]
PI4KIIIα, PI3Ks
>720
[15,33]
in vitroinflammation[72]
T-00127-HEV160
[32]
PI4KIIIα, PIK3CD
10 000
[32]
in vitroviral infection[77,78,79]
in vivo[78]
BF7387355.7
[32]
PI4KIIIα
1700
[33]
in vitroviral infection[79,80]
Enviroxime120
[32]
PI4KIIIα
1400
[32]
discontinued in phase II clinical trialsviral infection[15,81]
Plasmodium
falciparum
PI4KIIIβ
MMV39004828
[82]
no off-targets,
highly selective
terminated phase II clinical trialsparasitic infection[15,83]
PI4P5KIαISA-2011Bn.d.
[15]
p110α
n.d.
[15]
in vitroinflammation[84]
cancer[85,86]
in vivocancer[85,86]
PI4P5KIγUNC323051 (Kd)
[15]
PI5P4Kγ
4 (Kd)
[15]
in vitrocancer[87]
Table 2. Phosphoinositide-binding protein domains commonly used for the visualization of PI4P and PI(4,5)P2 within cells.
Table 2. Phosphoinositide-binding protein domains commonly used for the visualization of PI4P and PI(4,5)P2 within cells.
PIProtein DomainIntracellular Localization
PI4POSH2-PH [144]PM
OSBP-PH [145]Golgi
FAPP1-PH [121]Golgi
P4M SidM [118]Golgi, PM, LE/Lys
PI(4,5)P2PLCδ1-PH [146]PM
Tubby-PX [147]PM
Epsin1-ENTH/AP180-ANTH [148,149]PM
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

Bura, A.; Čabrijan, S.; Đurić, I.; Bruketa, T.; Jurak Begonja, A. A Plethora of Functions Condensed into Tiny Phospholipids: The Story of PI4P and PI(4,5)P2. Cells 2023, 12, 1411. https://doi.org/10.3390/cells12101411

AMA Style

Bura A, Čabrijan S, Đurić I, Bruketa T, Jurak Begonja A. A Plethora of Functions Condensed into Tiny Phospholipids: The Story of PI4P and PI(4,5)P2. Cells. 2023; 12(10):1411. https://doi.org/10.3390/cells12101411

Chicago/Turabian Style

Bura, Ana, Sara Čabrijan, Iris Đurić, Tea Bruketa, and Antonija Jurak Begonja. 2023. "A Plethora of Functions Condensed into Tiny Phospholipids: The Story of PI4P and PI(4,5)P2" Cells 12, no. 10: 1411. https://doi.org/10.3390/cells12101411

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