Europe PMC
Do data resources managed by EMBL-EBI and our collaborators make a difference to your work?
If so, please take 10 minutes to fill in our survey, and help us make the case for why sustaining open data resources is critical for life sciences research.

This website requires cookies, and the limited processing of your personal data in order to function. By using the site you are agreeing to this as outlined in our privacy notice and cookie policy.

Abstract 


Class I PI 3-kinases signal by producing the signaling lipid phosphatidylinositol(3,4,5) trisphosphate, which in turn acts by recruiting downstream effectors that contain specific lipid-binding domains. The class I PI 3-kinases comprise four distinct catalytic subunits linked to one of seven different regulatory subunits. All the class I PI 3-kinases produce the same signaling lipid, PIP3, and the different isoforms have overlapping expression patterns and are coupled to overlapping sets of upstream activators. Nonetheless, studies in cultured cells and in animals have demonstrated that the different isoforms are coupled to distinct ranges of downstream responses. This review focuses on the mechanisms by which the production of a common product, PIP3, can produce isoform-specific signaling by PI 3-kinases.

Free full text 


Logo of nihpaLink to Publisher's site
Bioessays. Author manuscript; available in PMC 2014 Jul 1.
Published in final edited form as:
PMCID: PMC3831874
NIHMSID: NIHMS525656
PMID: 23765576

PIP3: Tool of Choice for the Class I PI 3-kinases

Abstract

Class I PI 3-kinases signal by producing the signaling lipid phosphatidylinositol (3,4,5) trisphosphate, which in turn acts by recruiting downstream effectors that contain specific lipid-binding domains. The Class I PI 3-kinases comprise four distinct catalytic subunits linked to one of seven different regulatory subunits. All the Class I PI 3-kinases produce the same signaling lipid, PIP3, and the different isoforms have overlapping expression patterns and are coupled to overlapping sets of upstream activators. Nonetheless, studies in cultured cells and in animals have demonstrated that the different isoforms are coupled to distinct ranges of downstream responses. This review focuses on the mechanisms by which the production a common product, PIP3, can produce isoform-specific signaling by PI 3-kinases.

Phosphatidylinositol-3,4,5-trisphosphate (PIP3) is the exclusive product of the Class I PI 3-kinases, which are large heterodimeric proteins consisting of a 110 kDa catalytic subunit (p110α, p110β, p110δ, or p110γ) bound to a regulatory subunit (p85α, p85β, p50α, p55α, p55γ, p87 or p101)[1]. Depending on the isoform, the Class I PI 3-kinases are activated in response to ligand stimulation of both receptor tyrosine kinases and G-protein-coupled receptors. Although activated downstream of kinases, these enzymes are not activated by direct phosphorylation. Instead, their activity is regulated by transient binding to tyrosine phosphorylated proteins, Gβγ subunits from trimeric G-proteins, small GTPases in the Rho and Ras family, and SH3 domain-containing proteins.

Despite the fact that all of these enzymes utilize the same substrate, PI[4,5]P2, and produce the same lipid, PIP3, extensive studies in cell culture and in model organisms have demonstrated that the different PI 3-kinase isoforms produce a wide variety of distinct downstream responses. Isoform-specific signaling downstream from Class I PI 3-kinase has been reviewed in depth [2]. This article will instead focus on potential mechanisms by which the production of PIP3 by different PI 3-kinases can lead to divergent signals.

PIP3 works by recruiting downstream effectors containing lipid-binding domains

PIP3 was first identified by Cantley and colleagues in PDGF-stimulated fibroblasts [3]. The lipid was distinguished from other phosphoinositides by the mobility of its deacylated head group (gro-PIP3) on an anion-exchange HPLC column. The lipid can also be identified by its slow mobility, relative to phosphatidylinositol and mono- or bis-phosphoinositides, during thin layer chromatography [4]. More recent methods for the analysis of PIP3 have included displacement assays or blot-based assays with recombinant PIP3 binding proteins [5,6], and mass spectrometric methods [7].

Unlike PI[4,5]P2, which can be converted from a signaling lipid to the soluble second messengers IP3 and diacylglycerol by phospholipase C-mediated, 3-phosphoinositides are not substrates for phospholipase C [8]. Instead, they signal by binding to and recruiting effector proteins containing specific lipid binding domains. The first identified PIP3 effector was the protein kinase Akt/PKB, which contains a Pleckstrin Homology Domain (PH domain) that binds to PIP3 and PI[3,4]P2 and is required for Akt activation by PI 3- Additional PIP3-binding proteins include GRP-1, ARNO and centaurin-1, which are exchange factors for the ARF GTPases, cytohesin which regulates integrin signaling, and the Tec-family tyrosine kinase BTK[9]. The presence of a tandem Dbl-homology domain-PH domain was also identified as a common feature of guanine nucleotide exchange factors (GEFs) for Rho-family GTPases, including PIP3-regulated GEFs like Tiam1 and Vav [10,11]. In all of these cases, binding to PIP3 required an intact PH domain. It is important to note that only a subset of PH domains bind PIP3; others bind to different phosphoinositides, or do not bind lipids at all [9].

These studies led to a model in which production of PIP3 at the cytosolic face of cellular membranes leads to the recruitment and, in some cases, the allosteric activation of the effector proteins. The mechanism is most fully developed for Akt, where PIP3 binding to the Akt PH domain drives 3 related events: targeting of Akt to the cell membrane, which is coincident with membrane targeting of the upstream Akt activator PDK-1, followed by a conformational rearrangement of AKT that facilitates PDK-1 mediated phosphorylation on T308 in the Akt activation loop [12]. However, it must be noted that while lipid binding contributes to the membrane recruitment of some PH domain-containing proteins, protein-protein interactions may serve to further refine the site of recruitment, as well as the stability and duration of targeting. For example, the PH domains from Akt, ARNO, Btk and GRP1 all bind to PIP3, yet their expression in cells caused distinct inhibitory effects on cellular signaling, and mutation of residues not involved in PIP3 binding abolished these effects [13]. These data suggest that PIP3-independent interactions affect the location, stability or duration of PIP3-mediated targeting. In some cases, PIP3 binding to PH domains can regulate enzyme activity independently of membrane recruitment. For example, in the Arf GAP ARAP1, the PH domain is required for PIP3-stimulated GAP activity, yet does not bind tightly enough to PIP3 to mediate membrane binding [14]. In this case, membrane recruitment by a PIP3-independent mechanism, such as protein-protein interaction, facilitates allosteric regulation of the GAP by PIP3 binding to the PH domain. Overall, it seems likely that the specificity of PH domain-mediated signaling involves a combination of lipid mediated and protein mediated targeting.

In the repertoire of downstream signaling events initiated by production of PIP3, activation of the Akt kinase plays an outsized role [15]. The 3 isoforms of this serine/threonine protein kinase have major roles in the regulation of metabolism, and protein and lipid synthesis and degradation, through the FOXO family of transcription factors, the Glycogen Synthase Kinase 3 protein kinase, the AS160 regulator of glucose transport trafficking, and the mTOR protein kinase, among other downstream effectors. Akt also regulates survival pathways, inflammatory signaling, the cell cycle, and cardiovascular homeostasis. Other major effectors of PIP3 include guanine nucleotide exchange factors involved with vesicular trafficking and cytoskeletal regulation [16], and the Tec tyrosine kinases that regulate signaling in lymphocytes and mast cells [17]. The downstream effectors of PIP3 have been described in detail in recent reviews [18,19], and will not be discussed further here.

Class I PI 3-kinases are regulated by protein-protein interactions

The Class I PI 3-kinases are subdivided into the Class IA enzymes, in which the regulatory subunits have two SH2 domains that interact with tyrosine-phosphorylated proteins, and the Class IB enzymes, which interact with a distinct pair of structurally uncharacterized regulatory subunits (Figure 1). The Class IA heterodimer, which consists of an 85, 55 or 50 kDa regulatory subunit bound to one of three 110 kDa catalytic subunits, is extremely stable, and is essentially irreversible once formed [1]. The Class IB PI 3-kinase heterodimer is comprised of the p110γ catalytic subunit and either the p101 or p87 regulatory subunits [2]. We know of four types of regulatory interactions that activate heterodimeric PI 3-kinases.

An external file that holds a picture, illustration, etc.
Object name is nihms525656f1.jpg
Domain Organization and Activators of Class I PI 3-kinases

The Class IA PI 3-kinases consist of a p85 regulatory subunit containing SH3, BCR, proline-rich, SH2 and coiled-coil (iSH2) domains, and a p110 catalytic subunit (in yellow) containing an Adapter Binding Domain (ABD), a Ras Binding Domain (RBD), C2, helical and kinase domains. Shorter isoforms of p85 (p55α, p50α, p55γ) contain a truncated N-terminal segment linked to the first SH2 domain. The Class IB PI 3-kinase consists of structurally uncharacterized p87/84 or p101 regulatory subunits bound to a p110γ catalytic subunit (in yellow), whose domain organization is similar to that of Class IA PI 3-kinases. The mechanism of p87/p101 binding to p110γ is not yet known, but involves both the N- and C-terminal regions of p110γ [120]. Of the upstream activators shown, Ras binding regulates p85/p110α and p85/p110δ but not p85/p110β. Gγ regulates p85/p110β and p101/p110γ but not p85/p110α or p85/p110δ.

  1. Disruption of SH2 domain-mediated inhibition. For the Class IA PI 3-kinases, displacement of inhibitory contacts between the p85 SH2 domains and the p110 catalytic subunit activate the enzyme [2022]. Inhibitory interactions involving both the n-terminal and c-terminal SH2 domains have been demonstrated [23], although there is controversy as to whether the cSH2 domain is inhibitory for p110α [24]. The consensus is that SH2 domain binding to activated receptor tyrosine kinases or their substrates disrupt the inhibitory contacts and activates the p85/p110 dimer.

  2. Membrane targeting. Class I PI 3-kinases interact with RTKs, activated Ras, activated Rac, Gβγ subunits, and SH3 domains from Src-family tyrosine kinases [1]. All of these proteins are integral membrane proteins or are membrane associated, through the covalent attachment of myristoyl, palmitoyl and isoprenoid lipids. Given that the substrate of the PI 3-kinases is a membrane lipid, increased membrane targeting is sufficient to increase the rate of PIP3 production by PI 3-kinases. This was most clearly shown by the hyperactive phenotype induced by direct membrane targeting of PI 3-kinase catalytic subunits by addition of a C-terminal CAAX motif [25]. Using deuterium exchange-mass spectrometry methods, Williams and coworkers have noted that the binding of tyrosine phosphorylated peptides to p85 SH2 domains leads to conformational changes that enhance p110α and p110δ interactions with membranes[26]. Both p110β and p110γ bind to Gβγ subunits, which are released downstream from activated G-protein-coupled receptors (GPCRs). For p110β, the binding site has been localized to the C2 domain-helical domain linker [27]. Activation by Gβγ binding is synergistic with activation of phosphotyrosine peptide binding to the p85 N-terminal SH2 domain, and may act in part by disrupting the inhibitory contact formed by this domain. However, activation by Gβγ is also seen with monomeric p110β or p110γ[28], suggesting that a large component of Gβγ-mediated activation is due to membrane targeting. A similar mechanism has been suggested for the activation of p101/p110γ by Gβγ[29]

  3. Allosteric activation by small GTPases binding. As mentioned above, both small GTPases and trimeric G-proteins directly interact with Class I PI 3-kinase. All the p110 catalytic subunits contain a Ras-binding domain, although direct binding to Ras-GTP has only been demonstrated for p110α, p110δ and p110γ[30]. A crystal structure of the p110γ/Ras-GTP complex indicated conformational changes in the catalytic cleft, which presumably increases enzyme activity synergistically with membrane targeting [31]. Interestingly, it has been shown that activation by Ras preferentially occurs with the p87/p110γ dimer, as opposed to with the p101/ p110γ dimer [32], suggesting that the Class IB regulatory subunits influence the interactions of p110γ with Ras. Finally, binding of activated Rac or Cdc42 to the so-called BCR homology domain in the p85 subunit activates p85/p110α dimers in vitro[33], all this has not been demonstrated in intact cells. The mechanism of activation is not known, as there is little information on the conformation/structure of the N-terminal half of p85.

  4. Disruption of SH3/proline rich domain binding. The SH3 domain of p85, which resides at the extreme N-terminus of the protein, has been shown to bind to peptides derived from the internal proline-rich domains (PRDs) in p85. These data suggest that the N-terminus of p85 may exist in a closed conformation mediated by intramolecular SH3-PRD interactions. Binding events that would be predicted to disrupt such a conformation activate p85/p110 dimers. For example, p85/p110 dimers are activated by SH3 domains Src-family kinases [34], which can bind to p85 PRDs. While these data suggest a model in which exogenous proline-rich peptides or SH3 domains disrupt an inhibitory intramolecular SH3-PRD interaction, there is no direct evidence supporting this model. Given that the BCR domain lies between the two proline-rich domains, it is possible that Rac/Cdc42 binding to the BCR domain also functions by disrupting intramolecular SH3-propline-rich domain interactions.

Activating oncogenic mutations are commonly found in the p110α catalytic subunit [35], as well as in the p85 regulatory subunit [36,37]. The mechanisms by which these mutations activate PI 3-kinase are in some cases known. For example, the frequently mutated E545K residue in the p110α helical domain disrupts the inhibitory contact made by the N-terminal SH2 domain [20], and thus mimics phosphotyrosine-mediated activation. In contrast, a second commonly mutated site in p110α, H1047R, acts by increasing the basal association of p85/p1107aga; dimers with membranes [21]. The net effect of most activating mutants of p110α appears to be to induce conformational changes that increase membrane binding [26]. Mutants of both p85 and p110α in human tumors have been described; these mutants occur at an inhibitory contact between the C2 domain of p110α and the iSH2 domain of p85 [36,38,39]. It appears that these mutations mimic a loosening of the C2 domain-iSH2 domain interface that occurs in response to phosphotyrosine-mediated displacement of the inhibitory nSH2 domain [26].

PIP3 phosphatases shape the kinetics of PIP3 signaling

The importance of the timely elimination of PIP3-mediated signals, once they are initiated, is illustrated by the fact that the PIP3 3-phophatase, Phosphatase and Tensin Homology (PTEN), is frequently mutated or eliminated in human cancers [40,41]. The normal functioning of metazoan cells requires that the PIP3 signal be sharply constrained in both time and space.

The half-life of PIP3 in cells has been difficult to measure, but appears to be quite short. Using a PIP3 antibody to measure the decay in EGF-stimulated plasma membrane PIP3 induced by acute treatment with the PI 3-kinase inhibitor LY294002, the half-life for PIP3 was estimated to be less than 5 sec [42]. Of note, this study also found that a commonly used method to measure PIP3 levels in live cells, expression of GFP-labeled PH domains, led to an approximately 5-fold increase in the half-life of PIP3. This effect was presumably due to the sequestration of PIP3 by PH domain overexpression. This analysis suggests that measurements of PIP3 dynamics using PH domain probes, which have estimated the half-life of PIP3 to be approximately 1 min [43], likely overestimate the duration of the PIP3 signaling.

The overall dynamics of the cellular PIP3 response is stimulation dependent. In fMLP-stimulated neutrophils, PIP3 levels peaked by 20 sec and declined to close to basal levels by 2 min [44]. In PDGF-stimulated fibroblasts, PIP3 peaked by 40 sec but remained elevated at 5 min [45]. Other studies have suggested that sustained elevations of PIP3 occurs in response to acute stimulation of receptor tyrosine kinases [3,46]. The lifetime of the PIP3 signal in specialized cases is more transient. For example, in macrophages, there is an accumulation of PIP3 in the phagocytic cup during FcRγ-mediated phagocytosis; PIP3 disappears during closure of the phagosome, over a period of 2–3 min[47,48]. In Dictyostelium acutely exposed to a chemoattractant gradient, PIP3 accumulates at the leading edge in a biphasic manner, with a transient peak at 20 sec followed by a sustained elevation at 2 min[49].

The rapid turnover of newly synthesized PIP3 occurs either through removal of the 3-phosphate by PTEN, returning the lipid to the PI[4,5]P2 precursor, or through removal of the 5-phosphate by phosphatases such as SHIP or synaptojanin [50]. PTEN-mediated hydrolysis clearly terminates the PIP3 signal, and numerous studies have shown that inactivation or deletion of PTEN leads to increased levels of PIP3 [40]. The downstream effects of the activity of the 5-phosphatases is less clear, since their product is a lipid that can still signal via PH domains that bind to PI(3,4)P2, including those of Akt or Tapp1 [51,52]. Nonetheless, knockout of SHIP enhances the inflammatory response of myeloid and mast cells [53,54], clearly indicating its role as a negative regulators of PI 3-kinase signaling.

Interestingly, studies in Dictyostelium have shown that PTEN may contribute to certain PIP3-mediated cellular responses, particularly those that require an asymmetric spatial distribution of the lipid. cAMP-stimulation of Dictyostelium leads to an acute and spatially restricted accumulation of PIP3 at the leading edge of the cell [55]. This is accomplished at least in part by PTEN localization to the sides and rear of the cell, and its exclusion from the leading edge [56]. The positive role of PTEN in the chemotactic response is demonstrated by the impaired directionality of PTEN null cells, which produce poorly coordinated protrusions that reduce movement toward the chemoattractant source [57].

The biology of PTEN deletion involves phenotypes that are more complex than can be explained simply by an increase in PIP3 levels or lifetime. PTEN has a protein phosphatase activity whose disruption leads to distinct phenotypes [58,59]. PTEN also has signaling functions in the nucleus that do not appear to involve dephosphorylation of PIP3 [41,60].

How do different Class I PI 3-kinase isoforms achieve signaling specificity?

The distinct functions of the Class I PI 3-kinases have been studied in vitro, in cells and in animal models. Abundant evidence has accumulated showing that these lipid kinases perform quite different functions [2]. While isoform specific signaling by a protein kinase could easily be due to distinct substrate specificities, all the Class I PI 3-kinases make the same product, PIP3. In that case, how is isoform-specific signaling achieved? The question of how different p110 isoforms perform distinct tasks is especially interesting for the Class IA catalytic subunits, which share the same regulatory subunits and therefore should have similar interactions with tyrosine phosphorylated proteins.

Selective expression

To some extent, the dominant PI 3-kinase in a given response depends on which PI 3-kinase is expressed in the cell. While both p110β and p110α are ubiquitously expressed, the ratios of these two catalytic subunits may vary considerable. For example, mass spectroscopic analysis shows that p110β predominates in NIH3T3 cells, and in fat, liver and brain, whereas p110α and p110β are expressed at similar levels in muscle [61]. Similarly, the predominant expression of p110γ and p110δ in hematopoietic cells explains the anti-inflammatory phenotypes observed in kinase dead knock-in mice for these p110 isoforms [62,63]. Interestingly, these patterns of expression are disrupted in cancer. For example, increased expression of p110δ and p110γ have been observed in many non-hematopoietic cancers [2].

However, the expression pattern for the p110 isoforms is often insufficient to explain which isoform mediates a given cellular response. For example, 3T3 L1 adipocytes express 3–4 fold more p110β than p110α whereas L6 myotubes express similar levels of both isoforms. Selective inhibition of p110α decreased PIP3 production in response to insulin in both cell types, whereas an inhibitor of p110β had only a partial effect in 3T3L1 cells and no effect in L6 cells [64]. Similarly, rat adenocarcinoma cells express similar levels of p110α and p110β, yet p110α is selectively involved in EGF-stimulated protrusive responses [65].

One complication of this type of comparison is that the published data on the enzyme kinetic properties of the different Class I PI 3-kinases under different activation conditions is sparse. Using PIP2 as substrate, the maximal in vitro activity of purified p85/p110α from insect cells was 5-fold higher than that of p85/p110β or p85/p110δ or monomeric p110γ; dimers of p110γ with p101 or p87 were not tested [66]. Similarly, Shepherd and colleagues found that p85/p110α, expressed in mammalian cells, was 4-fold more active than p85/p110β using PI as a substrate, although activities were similar using PIP2. This study also found that the activity ratio of p110α versus p110β was reversed at low substrate concentrations, suggesting a lower Km for lipid for p110β [67]. In contrast, a comparison of the Class IA PI 3-kinases using a scintillation proximity assay found that the Km for ATP was 26–28 μM for all three p85/p110 dimers; the ATP Km increased to 43 μM p110δ/p55α [68]. All of these measurements were made under basal conditions; a careful comparison of the substrate Km and the maximal specific activity of each isoform in the presence of single and multiple activators (e.g. RTK plus Ras) is still lacking.

Mechanisms of activation: RTK/G-protein synergy

As discussed above, the three Class IA PI 3-kinase exhibit differences in their responsiveness to G-proteins, which could explain some of their distinct signaling outputs. For example, p110α and p110δ both interact directly with activated Ras in a manner that could synergize with activation via SH2 domain occupancy. In contrast, p110β has not been shown to be activated by Ras, but is activated synergistically by SH2 domain occupancy and direct binding to Gβγ subunits from trimeric G-proteins [28]. p110γ is activated by both Ras and Gβγ subunits [69], although it has been suggested that these are independent events mediated by the p101 versus p87 regulatory subunits [32]. Recent studies have also detected activation of p110γ downstream from selected receptor tyrosine kinase and Toll-like receptors [70]. Taken together, these data suggest that in response to a common ligand (e.g. insulin), the coordination of the insulin receptor/p-IRS-1 signal with concurrent signals from GPCRs or Ras could lead to distinct amplitudes or durations of signaling output (production of PIP3) from different PI 3-kinases, as well as distinct locations of PIP3 accumulation (see below).

Targeting of PI 3-kinases: plasma membrane, rafts and endomembranes, and the nucleus

The production of PIP3 at specific locations via PI3-kinase targeting is not well understood. A major stumbling block has been the absence of good antibodies for the detection of the endogenous enzymes. Overexpression studies on epitope- or fluorescent protein-tagged versions of the Class IA catalytic subunits are particularly difficult to interpret, since the overexpressed enzymes compete with endogenous enzymes for p85 adapter subunits. For example, in rat hepatoma cells that require p110α for EGF-stimulated cytoskeletal responses [65], overexpression of wild type p110β beta acts as a dominant negative [71], presumably by competing for a limiting pool of p85. It is hoped that these issues will be addressed in the future by arduous but definitive methods such as knocking in tagged PI 3-kinases into somatic cells or mice.

Despite these deficiencies in our knowledge, there is good evidence for the accumulation of PIP3 in a number of distinct subcellular locations. Most clear is the plasma membrane, where GFP-PH domain probes, anti-PIP3 immunostaining and electron microscopy using gold-labeled PH domains have shown the rapid accumulation of PIP3 in response to both RTK and GPCR ligands [42,43,60]. These data are consistent with the localization of the major allosteric activators of Class I PI 3-kinases: receptor tyrosine kinases like the PDGF and CSF-1 receptors, lipidated GTPases like Ras and Rac, and lipidated Gβγ subunits from trimeric G-proteins. Since the PI 3-kinases are largely cytosolic under basal conditions, the localized activation of distinct PI 3-kinase depends on the localization of their allosteric activators.

Within the plasma membrane, differential signaling by PI 3-kinase could occur via the partitioning of PI 3-kinase activators into and out of lipid rafts, which are regions characterized by relatively high levels of cholesterol, sphingolipids, glycophosphatidylinositol-anchored proteins and glycolipids [72]. Studies by Hope and Pike demonstrated enrichment of the Class I PI 3-kinase substrate PIP2 in lipid rafts [73], although this concept has been controversial [74]. Moreover, studies on the localization of fluorescent PH domains and FRET-based biosensors have documented the appearance of PIP3 in lipid rafts [7577]. Gao and Zhang measured the production of PIP3 and activation of Akt using fluorescent biosensors that could be specifically targeted to raft versus non-raft domains of the plasma membrane [78]. PI 3-kinase activation by insulin was highly raft-dependent, whereas responses to PDGF where only partially diminished by pharmacological disruption of rafts. These data suggested that insulin stimulated PI 3-kinase activation occurred primarily in rafts. In contrast, Lassere et al have suggested that the concentration of PIP3 in rafts is induced by PIP3 binding proteins, and does not imply that the PIP3 is actually synthesized in rafts[79].

How might rafts achieve isoform specific activation of PI 3-kinase? Some receptor tyrosine kinases have been reported to be concentrated in rafts [8082], and if these receptors preferentially bound a subset of Class IA isoforms, this could achieve specific signaling. Unfortunately, since all the Class IA isoforms share the same p85 subunits, it is not clear how this would generate specificity. Although numerous reports have shown that Gα subunits and Gβγ subunits from trimeric G-proteins are enriched in rafts [8386], isolated Gβγ subunits may actually be excluded from rafts [81]. Thus, it is not clear whether Gβγ interactions with PI 3-kinases would occur within rafts, or in non-raft regions after Gβγ dissociation from Gα. Furthermore, as far as we know, p110γ and p110β are activated by similar isoforms of Gβγ subunits [87], so this targeting would not lead to differential signaling between the two isoforms. The palmitoylated isoforms of Ras, like N-Ras and H-Ras, are raft associated, with GTP-loaded N-Ras showing increased partitioning into rafts [88]. In contrast, K-Ras is largely excluded from rafts [89]. If, for example, N-Ras preferentially interacts with p110α, p110δ or p110γ, then rafts could mediate isoform-specific activation. Although all Ras isoforms (N-, H- and K-Ras as well as R-Ras) bind to p110α, p110δ and p110γ in vitro (Julian Downward, personal communication), the net activation of PI 3-kinase signaling by distinct combinations of Ras and PI 3-kinase has not been systematically evaluated in vivo. One study did evaluate the effects of distinct Ras isoforms on the four p110 catalytic subunits in cells; p110α, p110δ and p110γ were activated by Ras, but only p110δ showed selectivity (for the Ras homologs R-Ras and TC21) [90]. However, since the catalytic subunits were expressed without their regulatory subunits, the findings are difficult to interpret.

In contrast to the clear evidence that the bulk of PIP3 production occurs in the plasma membrane, evidence for the production of PIP3 in endomembranes is more limited. The EM study of Lucocq and colleagues estimated that no more than 5% of total PIP3 was in any endomembrane compartment [60], but physiologically important signals could be generated from local concentrations of this signaling molecule. Several studies have shown that the PI 3-kinase substrate PI(4,5)P2 is present in endosomes [91,92]. Using a PIP3 biosensor that was directed to endomembranes via a palmitoylation-deficient Ras CAAX motif, PDGF-stimulated PIP3 production in endomembranes was observed [93]. The signal was abolished by mutant dynamin, suggesting that it was caused by the endocytosis of activated PDGF receptors. Similarly, Wang et al. used a combination of monensin and a PDGFR inhibitor to drive the accumulation of ligand-bound but inactive receptors in endosomes. Upon removal of the inhibitors, activation of the endosomal PDGF was sufficient to activate downstream effectors [82]. This suggests that endosomal PI 3-kinase signaling is feasible. Studies on the PIP3-stimulated kinase Akt in fact have suggested endosomes as an important compartment for Akt signaling, particularly in conjunction with the Rab5-associated endosomal protein APPL [94]. However, given that Akt activation continues after its dissociation from PIP3, it is possible that Akt is recruited from a plasma membrane site of activation, rather than being activated in situ at the endosome by PIP3.

In an interesting variation on the story of endosomal Class I PI 3-kinases, Zerial and colleagues demonstrated the presence of the p110β catalytic subunit in endosomes [95], and suggested that its colocalization with lipid phosphatases was an important source of endosomal PI[3]P rather than PIP3 [96]. However, the endosomal functions of p110β appear to be independent of its lipid kinase activity [97,98], so the presence of p110β in endosomes does not necessarily imply the presence of PIP3. In contrast, in degranulating mast cells, p101/p110γ was shown to produce a rapidly endocytosed pool of PIP3 that was not observed in cells expressing p87/p110γ[99]. In this case, endosomal PIP3 is derived from the plasma membrane, so the presence of endosomal PIP3 does not imply the presence of a class I PI 3-kinase.

How might distinct PI 3-kinases be differentially targeted to endosomes? The co-internalization of PDGF receptors and p85 was visualized in an early paper by Corvera and Cantley [100], but this study did not distinguish between p85-bound p110 isoforms. Similarly, activation of PI 3-kinase in insulin-stimulated cells is mediated by binding to the IRS-1 scaffold protein [101], which is localized to poorly defined intracellular membranes [102,103]. Although Vanhaesebroeck and colleagues suggested that IRS-1/2 preferentially recruits p85/p110α in insulin-stimulated liver, muscle and fat [104], the mechanism for this selectivity is not clear. As noted above, p85/p110β dimers directly interact with the early endosomal GTPase Rab5 [95], providing a unique localization signal for this isoform. We have recently defined residues in the helical domain of p110β, Q596 and I597, whose mutation abolishes p110β binding to Rab5 (R.S. Salamon, H.A. Dbouk and J.M. Backer, unpublished observations). Studies on the phenotype of cells expressing these mutants will directly evaluate the role of endosome-specific signaling from p110β.

Finally, Lucocq and Downes reported a PDGF-stimulated pool of nuclear PIP3 [60]. Studies have shown that both p110β and p110γ are present in the nucleus [105,106], and the CSF-1 receptor has been shown to traffic to the nuclear envelope, where is activates Akt in p110δ-dependent manner [107]. Nuclear localization would therefore not distinguish signaling by p110β and p110γ, but might distinguish signaling by p110δ versus p110α.

Isoform specific signaling by PI 3-kinases requires combinatorial inputs

Taken together, it does not seem as if any one activation or targeting mechanism can produce a specific or preferential activation of any given PI 3-kinase isoforms (Figure 2). In some cases, the predominance of one isoform over another may be explained by differences in the kinetic properties of the isoforms; this seems most likely to be the case for p110α, given its higher basal activity [66,67]. Alternatively, isoform-specific PI 3-kinase signaling, in a cell that contains multiple isoforms, may require a combination of localization constraints and synergistic activation by multiple activators (Table 1). For example, in a growth factor-stimulated cell, p85/p110α and p85/p110β might both reside in an endosome, via receptor tyrosine kinase binding for the former and Rab5 binding for the latter. If the growth factor were a potent activator of endosomal Ras, this would synergistically activate p85/p110α but not p85/p110β, and lead to a signal that was primarily p110α-dependent. A combined RTK/GPCR signal might preferentially activate p85/p110β, whereas a combined GPCR/Ras signal might preferentially activate p87/p110γ. Of all the isoforms, differential signaling by p110α and p110δ was the first to be described [108] and remains the hardest to understand. Both isoforms bind RTKs and Ras, and both undergo autoinhibitory autophosphorylation (although by different mechanisms [109,110]. The ability of RTKs to distinguish between these isoforms may reside in the still mysterious cellular code for the matching of p110 isoforms (p110α, p110β, and p110δ) with p85 isoforms (p85α, p55α, p50α, p85β and p55γ).

An external file that holds a picture, illustration, etc.
Object name is nihms525656f2.jpg
Pathways to isoform-specific PI 3-kinase activation

While no one PI 3-kinase activator yields specific or preferential activation of any one PI 3-kinase isoform, combinations of activators and distinct activator locations could allow differential activation of distinct PI 3-kinases.

Table 1

Activators of the Class I PI 3-kinases.

RTKs- pYGβγRas
p85/p110αXX
p85/p110βXX
p85/p110δXX
P87/p110γXX

Synergistic activation by RTKs plus GPCRS (for p85/p110β) and GPCRs plus Ras (for p87/p110γ) could produce specific activation.

Finally, it should be noted that PI 3-kinases have activities not directly related to PIP3 production, which may affect signaling output. All of the Class I PI 3-kinases have protein kinase activity. Although in most cases the activity seems to be limited to autophosphorylation [109111], a protein kinase-only mutant of p110γ was able to signal to Erk [112] and a limited number of the protein substrates have been described [113,114]. In addition, PI 3-kinases have been shown to have biological activities that do not require an active kinase domain [115]. For example, mice expressing kinase dead p110β are viable, although infertile, whereas homozygous knockout of p110β produces embryonic lethality [97,98,116]. p110γ, in particular, has been shown to regulate contractility in the heart by serving as a scaffold for phosphodiesterase 3B and G-protein-coupled receptors kinase 2, which regulates the β2-adrenergic receptor [117]. The scaffolding activity of the PI 3-kinase regulatory and catalytic subunits is likely to contribute to the spectrum of downstream responses to distinct PI 3-kinase isoforms.

Conclusions and outlook

The focus of much of current research on PIP3-mediated signaling is on the identification of novel downstream effectors, but there is still a great deal to be learned about the activation, targeting and specificity of the enzymes producing and regulating the production of PIP3 itself. A review of our knowledge of the activation mechanisms for Class I PI 3-kinasese suggests that differential outputs from the distinct PI 3-kinase isoforms depends on different combinations of simultaneous upstream inputs, different sites of action, as well as contributions from kinase-independent scaffolding functions. Given the compelling data linking aberrant regulation of Class I PI 3-kinase to pathological outcomes, a more precise understanding of how these enzymes are regulated in cells should hold considerable promise for novel interventions into human disease.

Box 1

PIP3 – Unexplored areas

Although PIP3 production and destruction is a critical component of PI 3-kinase signaling, it is not clear that all PIP3 molecules are the same. Mass spectrometry studies on the acyl chain distribution of PIP3 have been inconsistent. A study in RAW 264.7 and primary murine macrophages showed significant differences in the fatty acyl composition of PIP3 molecules produced in response to ligands for distinct receptor types (LPA, C5A, Zymosan and MCF) [118]. In contrast, Clark et al. found that the production of C18:0/C20:4 (stearoyl/arachidonyl) PIP3 predominated in fMLP-stimulated neutrophils, EGF-stimulated MCF10A breast cancer cells, and insulin stimulated fat and liver [7]. A striking finding in the latter study was that the production of C18:0/C20:4 PIP3 was greatly in excess of its proportion in the overall PIP2 pool, suggesting a significant role of the fatty acid chains in the ability of PI 3-kinase to utilize substrate. It will be interesting to test whether this preference is altered under conditions in which one or more PI 3-kinase isoform is silenced or inhibited.

A second unanswered question with regard to PIP3 is whether it stays where it is put, i.e. whether the lipid can move laterally within membranes. While the half-life of bulk PIP3 in a cell is short, its migration into membrane subdomains such as rafts could lead to prolonged lifetimes for local PIP3 pools. Finally, it is interesting to note a recent study suggesting that increases in intracellular divalent cations, including Ca+2 and Mg+2, can induce the clustering of PIP2 in model membranes [119]. The authors suggest that cation-induced clustering of PIP2 could affect the ability of PIP2-binding proteins to bind PIP2 at the plasma membrane. While the study did not examine PIP3, it seems possible that a similar mechanism might cause PIP3 clustering, locally increasing PIP3 concentrations and modulating interactions with PIP3 binding proteins.

Acknowledgments

We would like to thank Dr. Anne Bresnick for critically reading the manuscript, and Drs. Tamas Balla, Paul Janmey, Anant Menon, Julian Downward, Bart Vanhaesebroeck, and Bernd Nurnberg for helpful discussion. RSS was supported by NIH 5T32 GM007491 and by a National Research Service Award, 1 F31 AG040932-01. This work was funded by NIH grants GM55692 and PO1 CA 100324 (JMB).

List of Abbreviations

ARAP1Arf GAP with Rho GAP domain, ankyrin repeat, and PH domain 1
ArfADP ribosylation factor
ARNOADP-ribosylation factor (ARF) nucleotide-binding site opener
BTKBruton’s tyrosine kinase
GAPGTPase-Activating Proteins
GEFGuanine nucleotide exchange factors
GRP-1General receptor for 3-phosphoinositides 1
PDGFPlatelet-derived growth factor
PDK-1Phosphoinositide dependent kinase-1
PI(3,4)P2phosphatidylinositol (3,4) bisphosphate
PI(4,5,)2phosphatidylinositol (4,5) bisphosphate
PIP3phosphatidylinositol (3,4,5) trisphosphate

References

1. Backer JM. The regulation of class IA PI 3-kinases by inter-subunit interactions. Curr Top Microbiol Immunol. 2010;346:87–114. [Europe PMC free article] [Abstract] [Google Scholar]
2. Vanhaesebroeck B, Guillermet-Guibert J, Graupera M, Bilanges B. The emerging mechanisms of isoform-specific PI3K signalling. Nat Rev Mol Cell Biol. 2010;11:329–41. [Abstract] [Google Scholar]
3. Auger KR, Serunian LA, Soltoff SP, Libby P, Cantley LC. PDGF-dependent tyrosine phosphorylation stimulates production of novel polyphosphoinositides in intact cells. Cell. 1989;57:167–75. [Abstract] [Google Scholar]
4. Serunian LA, Auger KR, Cantley LC. Identification and quantification of polyphosphoinositides produced in response to platelet-derived growth factor stimulation. Methods Enzymol. 1991;198:78–87. [Abstract] [Google Scholar]
5. van der Kaay J, Batty IH, Cross DA, Watt PW, Downes CP. A novel, rapid, and highly sensitive mass assay for phosphatidylinositol 3,4,5-trisphosphate (PtdIns(3,4,5)P3) and its application to measure insulin-stimulated PtdIns(3,4,5)P3 production in rat skeletal muscle in vivo. J Biol Chem. 1997;272:5477–81. [Abstract] [Google Scholar]
6. Dowler S, Kular G, Alessi DR. Protein lipid overlay assay. Sci STKE. 2002:pl6. [Abstract] [Google Scholar]
7. Clark J, Anderson KE, Juvin V, Smith TS, Karpe F, et al. Quantification of PtdInsP3 molecular species in cells and tissues by mass spectrometry. Nat Methods. 2011;8:267–72. [Europe PMC free article] [Abstract] [Google Scholar]
8. Serunian LA, Haber MT, Fukui T, Kim JW, Rhee SG, et al. Polyphosphoinositides produced by phosphatidylinositol 3-kinase are poor substrates for phospholipase C from rat liver and bovine brain. J Biol Chem. 1989;264:17809–15. [Abstract] [Google Scholar]
9. Lemmon MA. Membrane recognition by phospholipid-binding domains. Nat Rev Mol Cell Biol. 2008;9:99–111. [Abstract] [Google Scholar]
10. Stam JC, Collard JG. The DH protein family, exchange factors for Rho-like GTPases. Prog Mol Subcell Biol. 1999;22:51–83. [Abstract] [Google Scholar]
11. Erickson JW, Cerione RA. Structural elements, mechanism, and evolutionary convergence of Rho protein-guanine nucleotide exchange factor complexes. Biochemistry. 2004;43:837–42. [Abstract] [Google Scholar]
12. Pearce LR, Komander D, Alessi DR. The nuts and bolts of AGC protein kinases. Nat Rev Mol Cell Biol. 2010;11:9–22. [Abstract] [Google Scholar]
13. Varnai P, Bondeva T, Tamas P, Toth B, Buday L, et al. Selective cellular effects of overexpressed pleckstrin-homology domains that recognize PtdIns(3,4,5)P3 suggest their interaction with protein binding partners. J Cell Sci. 2005;118:4879–88. [Abstract] [Google Scholar]
14. Campa F, Yoon HY, Ha VL, Szentpetery Z, Balla T, et al. A PH domain in the Arf GTPase-activating protein (GAP) ARAP1 binds phosphatidylinositol 3,4,5-trisphosphate and regulates Arf GAP activity independently of recruitment to the plasma membranes. J Biol Chem. 2009;284:28069–83. [Europe PMC free article] [Abstract] [Google Scholar]
15. Fayard E, Xue G, Parcellier A, Bozulic L, Hemmings BA. Protein kinase B (PKB/Akt), a key mediator of the PI3K signaling pathway. Curr Top Microbiol Immunol. 2010;346:31–56. [Abstract] [Google Scholar]
16. Mizuno-Yamasaki E, Rivera-Molina F, Novick P. GTPase networks in membrane traffic. Annu Rev Biochem. 2012;81:637–59. [Europe PMC free article] [Abstract] [Google Scholar]
17. Felices M, Falk M, Kosaka Y, Berg LJ. Tec kinases in T cell and mast cell signaling. Adv Immunol. 2007;93:145–84. [Abstract] [Google Scholar]
18. Engelman JA, Luo J, Cantley LC. The evolution of phosphatidylinositol 3-kinases as regulators of growth and metabolism. Nat Rev Genet. 2006;7:606–19. [Abstract] [Google Scholar]
19. Stephens L, Hawkins P. Signalling via class IA PI3Ks. Adv Enzyme Regul. 2011;51:27–36. [Abstract] [Google Scholar]
20. Miled N, Yan Y, Hon WC, Perisic O, Zvelebil M, et al. Mechanism of two classes of cancer mutations in the phosphoinositide 3-kinase catalytic subunit. Science. 2007;317:239–42. [Abstract] [Google Scholar]
21. Mandelker D, Gabelli SB, Schmidt-Kittler O, Zhu J, Cheong I, et al. A frequent kinase domain mutation that changes the interaction between PI3Kalpha and the membrane. Proc Natl Acad Sci U S A. 2009;106:16996–7001. [Europe PMC free article] [Abstract] [Google Scholar]
22. Zhang X, Vadas O, Perisic O, Anderson KE, Clark J, et al. Structure of lipid kinase p110beta/p85beta elucidates an unusual SH2-domain-mediated inhibitory mechanism. Mol Cell. 2011;41:567–78. [Europe PMC free article] [Abstract] [Google Scholar]
23. Yu JH, Wjasow C, Backer JM. Regulation of the p85/p110α Phosphatidylinositol 3′-kinase - Distinct roles for the N-terminal and C-terminal SH2 domains. J Biol Chem. 1998;273:30199–203. [Abstract] [Google Scholar]
24. Burke JE, Vadas O, Berndt A, Finegan T, Perisic O, et al. Dynamics of the phosphoinositide 3-kinase p110delta interaction with p85alpha and membranes reveals aspects of regulation distinct from p110alpha. Structure. 2011;19:1127–37. [Europe PMC free article] [Abstract] [Google Scholar]
25. Klippel A, Reinhard C, Kavanaugh WM, Apell G, Escobedo MA, et al. Membrane localization of phosphatidylinositol 3-kinase is sufficient to activate multiple signal-transducing kinase pathways. Mol Cell Biol. 1996;16:4117–27. [Europe PMC free article] [Abstract] [Google Scholar]
26. Burke JE, Perisic O, Masson GR, Vadas O, Williams RL. Oncogenic mutations mimic and enhance dynamic events in the natural activation of phosphoinositide 3-kinase p110alpha (PIK3CA) Proc Natl Acad Sci U S A. 2012;109:15259–64. [Europe PMC free article] [Abstract] [Google Scholar]
27. Dbouk HA, Vadas O, Shymanets A, Burke JE, Salamon RS. G Protein-Coupled Receptor-Mediated Activation of p110beta by Gbetagamma Is Required for Cellular Transformation and Invasiveness. Sci Signal. 2012;5:ra89. [Europe PMC free article] [Abstract] [Google Scholar]
28. Maier U, Babich A, Nürnberg B. Roles of non-catalytic subunits in Gβgamma-induced activation of class I phosphoinositide 3-kinase isoforms β and gamma. J Biol Chem. 1999;274:29311–7. [Abstract] [Google Scholar]
29. Brock C, Schaefer M, Reusch HP, Czupalla C, Michalke M, et al. Roles of G beta gamma in membrane recruitment and activation of p110 gamma/p101 phosphoinositide 3-kinase gamma. The Journal of cell biology. 2003;160:89–99. [Europe PMC free article] [Abstract] [Google Scholar]
30. Castellano E, Downward J. RAS Interaction with PI3K: More Than Just Another Effector Pathway. Genes & cancer. 2011;2:261–74. [Europe PMC free article] [Abstract] [Google Scholar]
31. Pacold ME, Suire S, Perisic O, Lara-Gonzalez S, Davis CT, et al. Crystal structure and functional analysis of Ras binding to its effector phosphoinositide 3-kinase gamma. Cell. 2000;103:931–43. [Abstract] [Google Scholar]
32. Kurig B, Shymanets A, Bohnacker T, Prajwal, Brock C, et al. Ras is an indispensable coregulator of the class IB phosphoinositide 3-kinase p87/p110gamma. Proc Natl Acad Sci U S A. 2009;106:20312–7. [Europe PMC free article] [Abstract] [Google Scholar]
33. Zheng Y, Bagrodia S, Cerione RA. Activation of phosphoinositide 3-kinase by Cdc42Hs binding to p85. J Biol Chem. 1994;269:18727–30. [Abstract] [Google Scholar]
34. Pleiman CM, Hertz WM, Cambier JC. Activation of phosphatidylinositol-3′ kinase by Src-family kinase SH3 binding to the p85 subunit. Science. 1994;263:1609–12. [Abstract] [Google Scholar]
35. Samuels Y, Wang Z, Bardelli A, Silliman N, Ptak J, et al. High frequency of mutations of the PIK3CA gene in human cancers. Science. 2004;304:554. [Abstract] [Google Scholar]
36. Jaiswal BS, Janakiraman V, Kljavin NM, Chaudhari S, Stern HM, et al. Somatic mutations in p85α promote tumorigenesis through class IA PI3K activation. Cancer Cell. 2009 In press. [Europe PMC free article] [Abstract] [Google Scholar]
37. Parsons DW, Jones S, Zhang X, Lin JC, Leary RJ, et al. An integrated genomic analysis of human glioblastoma multiforme. Science. 2008;321:1807–12. [Europe PMC free article] [Abstract] [Google Scholar]
38. Philp AJ, Campbell IG, Leet C, Vincan E, Rockman SP, et al. The phosphatidylinositol 3′-kinase p85alpha gene is an oncogene in human ovarian and colon tumors. Cancer Res. 2001;61:7426–9. [Abstract] [Google Scholar]
39. Wu H, Shekar SC, Flinn RJ, El-Sibai M, Jaiswal BS, et al. Regulation of Class IA PI 3-kinases: C2 domain-iSH2 domain contacts inhibit p85/p110alpha and are disrupted in oncogenic p85 mutants. Proc Natl Acad Sci U S A. 2009;106:20258–63. [Europe PMC free article] [Abstract] [Google Scholar]
40. Keniry M, Parsons R. The role of PTEN signaling perturbations in cancer and in targeted therapy. Oncogene. 2008;27:5477–85. [Abstract] [Google Scholar]
41. Song MS, Salmena L, Pandolfi PP. The functions and regulation of the PTEN tumour suppressor. Nat Rev Mol Cell Biol. 2012;13:283–96. [Abstract] [Google Scholar]
42. Yip SC, Eddy RJ, Branch AM, Pang H, Wu H, et al. Quantification of PtdIns(3,4,5)P(3) dynamics in EGF-stimulated carcinoma cells: a comparison of PH-domain-mediated methods with immunological methods. Biochem J. 2008;411:441–8. [Europe PMC free article] [Abstract] [Google Scholar]
43. Tengholm A, Meyer T. A PI3-kinase signaling code for insulin-triggered insertion of glucose transporters into the plasma membrane. Curr Biol. 2002;12:1871–6. [Abstract] [Google Scholar]
44. Stephens L, Eguinoa A, Corey S, Jackson T, Hawkins PT. Receptor stimulated accumulation of phosphatidylinositol (3,4,5)-trisphosphate by G-protein mediated pathways in human myeloid derived cells. EMBO J. 1993;12:2265–73. [Europe PMC free article] [Abstract] [Google Scholar]
45. Hawkins PT, Jackson TR, Stephens LR. Platelet-derived growth factor stimulates synthesis of PtdIns(3,4,5)P3 by activating a PtdIns(4,5)P2 3-OH kinase. Nature. 1992;358:157–9. [Abstract] [Google Scholar]
46. Kapeller R, Chem KS, Yoakim M, Schaffhausen BS, Backer JM, et al. Mutations in the juxtamembrane region of the insulin receptor impair activation of phosphatidylinositol 3-kinase by insulin. Mol Endocrin. 1991;5:769–77. [Abstract] [Google Scholar]
47. Bohdanowicz M, Cosio G, Backer JM, Grinstein S. Class I and class III phosphoinositide 3-kinases are required for actin polymerization that propels phagosomes. J Cell Biol. 2010;191:999–1012. [Europe PMC free article] [Abstract] [Google Scholar]
48. Marshall JG, Booth JW, Stambolic V, Mak T, Balla T, et al. Restricted accumulation of phosphatidylinositol 3-kinase products in a plasmalemmal subdomain during Fc gamma receptor-mediated phagocytosis. J Cell Biol. 2001;153:1369–80. [Europe PMC free article] [Abstract] [Google Scholar]
49. Xu X, Meier-Schellersheim M, Jiao X, Nelson LE, Jin T. Quantitative imaging of single live cells reveals spatiotemporal dynamics of multistep signaling events of chemoattractant gradient sensing in Dictyostelium. Mol Biol Cell. 2005;16:676–88. [Europe PMC free article] [Abstract] [Google Scholar]
50. Leslie NR, Dixon MJ, Schenning M, Gray A, Batty IH. Distinct inactivation of PI3K signalling by PTEN and 5-phosphatases. Adv Enzyme Regul 2011 [Abstract] [Google Scholar]
51. Dowler S, Currie RA, Campbell DG, Deak M, Kular G, et al. Identification of pleckstrin-homology-domain-containing proteins with novel phosphoinositide-binding specificities. Biochem J. 2000;351:19–31. [Europe PMC free article] [Abstract] [Google Scholar]
52. Manna D, Albanese A, Park WS, Cho W. Mechanistic basis of differential cellular responses of phosphatidylinositol 3,4-bisphosphate- and phosphatidylinositol 3,4,5-trisphosphate-binding pleckstrin homology domains. J Biol Chem. 2007;282:32093–105. [Abstract] [Google Scholar]
53. Liu QR, Sasaki T, Kozieradzki I, Wakeham A, Itie A, et al. SHIP is a negative regulator of growth factor receptor-mediated PKB/Akt activation and myeloid cell survival. Genes Dev. 1999;13:786–91. [Europe PMC free article] [Abstract] [Google Scholar]
54. Huber M, Helgason CD, Damen JE, Liu L, Humphries RK, et al. The src homology 2-containing inositol phosphatase (SHIP) is the gatekeeper of mast cell degranulation. Proc Natl Acad Sci U S A. 1998;95:11330–5. [Europe PMC free article] [Abstract] [Google Scholar]
55. Huang YE, Iijima M, Parent CA, Funamoto S, Firtel RA, et al. Receptor-mediated regulation of PI3Ks confines PI(3,4,5)P3 to the leading edge of chemotaxing cells. Mol Biol Cell. 2003;14:1913–22. [Europe PMC free article] [Abstract] [Google Scholar]
56. Ma L, Janetopoulos C, Yang L, Devreotes PN, Iglesias PA. Two complementary, local excitation, global inhibition mechanisms acting in parallel can explain the chemoattractant-induced regulation of PI(3,4,5)P3 response in dictyostelium cells. Biophys J. 2004;87:3764–74. [Europe PMC free article] [Abstract] [Google Scholar]
57. Iijima M, Devreotes P. Tumor suppressor PTEN mediates sensing of chemoattractant gradients. Cell. 2002;109:599–610. [Abstract] [Google Scholar]
58. Tibarewal P, Zilidis G, Spinelli L, Schurch N, Maccario H, et al. PTEN protein phosphatase activity correlates with control of gene expression and invasion, a tumor-suppressing phenotype, but not with AKT activity. Sci Signal. 2012;5:ra18. [Abstract] [Google Scholar]
59. Davidson L, Maccario H, Perera NM, Yang X, Spinelli L, et al. Suppression of cellular proliferation and invasion by the concerted lipid and protein phosphatase activities of PTEN. Oncogene. 2010;29:687–97. [Europe PMC free article] [Abstract] [Google Scholar]
60. Lindsay Y, McCoull D, Davidson L, Leslie NR, Fairservice A, et al. Localization of agonist-sensitive PtdIns(3,4,5)P3 reveals a nuclear pool that is insensitive to PTEN expression. J Cell Sci. 2006;119:5160–8. [Abstract] [Google Scholar]
61. Geering B, Cutillas PR, Nock G, Gharbi SI, Vanhaesebroeck B. Class IA phosphoinositide 3-kinases are obligate p85-p110 heterodimers. Proc Natl Acad Sci U S A. 2007;104:7809–14. [Europe PMC free article] [Abstract] [Google Scholar]
62. Patton DT, Garcon F, Okkenhaug K. The PI3K p110delta controls T-cell development, differentiation and regulation. Biochem Soc Trans. 2007;35:167–71. [Abstract] [Google Scholar]
63. Costa C, Martin-Conte EL, Hirsch E. Phosphoinositide 3-kinase p110gamma in immunity. IUBMB Life. 2011;63:707–13. [Abstract] [Google Scholar]
64. Knight ZA, Gonzalez B, Feldman ME, Zunder ER, Goldenberg DD, et al. A pharmacological map of the PI3-K family defines a role for p110alpha in insulin signaling. Cell. 2006;125:733–47. [Europe PMC free article] [Abstract] [Google Scholar]
65. Hill K, Welti S, Yu JH, Murray JT, Yip SC, et al. Specific requirement for the p85-p110α phosphatidylinositol 3-kinase during epidermal growth factor-stimulated actin nucleation in breast cancer cells. J Biol Chem. 2000;275:3741–4. [Abstract] [Google Scholar]
66. Meier TI, Cook JA, Thomas JE, Radding JA, Horn C, et al. Cloning, expression, purification, and characterization of the human Class Ia phosphoinositide 3-kinase isoforms. Protein expression and purification. 2004;35:218–24. [Abstract] [Google Scholar]
67. Beeton CA, Chance EM, Foukas LC, Shepherd PR. Comparison of the kinetic properties of the lipid- and protein-kinase activities of the p110alpha and p110beta catalytic subunits of class-Ia phosphoinositide 3-kinases. Biochem J. 2000;350(Pt 2):353–9. [Europe PMC free article] [Abstract] [Google Scholar]
68. Van Aller GS, Carson JD, Fernandes C, Lehr R, Sinnamon RH, et al. Characterization of PI3K class IA isoforms with regulatory subunit p55alpha using a scintillation proximity assay. Anal Biochem. 2008;383:311–5. [Abstract] [Google Scholar]
69. Suire S, Condliffe AM, Ferguson GJ, Ellson CD, Guillou H, et al. Gbetagammas and the Ras binding domain of p110gamma are both important regulators of PI(3)Kgamma signalling in neutrophils. Nat Cell Biol. 2006;8:1303–9. [Abstract] [Google Scholar]
70. Schmid MC, Avraamides CJ, Dippold HC, Franco I, Foubert P, et al. Receptor tyrosine kinases and TLR/IL1Rs unexpectedly activate myeloid cell PI3kgamma, a single convergent point promoting tumor inflammation and progression. Cancer Cell. 2011;19:715–27. [Europe PMC free article] [Abstract] [Google Scholar]
71. Yip SC, El-Sibai M, Hill KM, Wu H, Fu Z, et al. Over-expression of the p110beta but not p110alpha isoform of PI 3-kinase inhibits motility in breast cancer cells. Cell Motil Cytoskeleton. 2004;59:180–8. [Abstract] [Google Scholar]
72. Lingwood D, Simons K. Lipid rafts as a membrane-organizing principle. Science. 2010;327:46–50. [Abstract] [Google Scholar]
73. Hope HR, Pike LJ. Phosphoinositides and phosphoinositide-utilizing enzymes in detergent-insoluble lipid domains. Mol Biol Cell. 1996;7:843–51. [Europe PMC free article] [Abstract] [Google Scholar]
74. van Rheenen J, Achame EM, Janssen H, Calafat J, Jalink K. PIP2 signaling in lipid domains: a critical re-evaluation. EMBO J. 2005;24:1664–73. [Europe PMC free article] [Abstract] [Google Scholar]
75. Arcaro A, Aubert M, Espinosa del Hierro ME, Khanzada UK, Angelidou S, et al. Critical role for lipid raft-associated Src kinases in activation of PI3K-Akt signalling. Cell Signal. 2007;19:1081–92. [Abstract] [Google Scholar]
76. Bjorgo E, Tasken K. Novel mechanism of signaling by CD28. Immunol Lett. 2010;129:1–6. [Abstract] [Google Scholar]
77. Calay D, Vind-Kezunovic D, Frankart A, Lambert S, Poumay Y, et al. Inhibition of Akt signaling by exclusion from lipid rafts in normal and transformed epidermal keratinocytes. J Invest Dermatol. 2010;130:1136–45. [Abstract] [Google Scholar]
78. Gao X, Zhang J. Spatiotemporal analysis of differential Akt regulation in plasma membrane microdomains. Mol Biol Cell. 2008;19:4366–73. [Europe PMC free article] [Abstract] [Google Scholar]
79. Lasserre R, Guo XJ, Conchonaud F, Hamon Y, Hawchar O, et al. Raft nanodomains contribute to Akt/PKB plasma membrane recruitment and activation. Nat Chem Biol. 2008;4:538–47. [Abstract] [Google Scholar]
80. Chen X, Resh MD. Cholesterol depletion from the plasma membrane triggers ligand-independent activation of the epidermal growth factor receptor. J Biol Chem. 2002;277:49631–7. [Abstract] [Google Scholar]
81. Moffett S, Brown DA, Linder ME. Lipid-dependent targeting of G proteins into rafts. J Biol Chem. 2000;275:2191–8. [Abstract] [Google Scholar]
82. Wang Y, Pennock SD, Chen X, Kazlauskas A, Wang Z. Platelet-derived growth factor receptor-mediated signal transduction from endosomes. J Biol Chem. 2004;279:8038–46. [Abstract] [Google Scholar]
83. Patel HH, Murray F, Insel PA. G-protein-coupled receptor-signaling components in membrane raft and caveolae microdomains. Handbook of experimental pharmacology. 2008:167–84. [Abstract] [Google Scholar]
84. Oh P, Schnitzer JE. Segregation of heterotrimeric G proteins in cell surface microdomains. G(q) binds caveolin to concentrate in caveolae, whereas G(i) and G(s) target lipid rafts by default. Mol Biol Cell. 2001;12:685–98. [Europe PMC free article] [Abstract] [Google Scholar]
85. Foster LJ, De Hoog CL, Mann M. Unbiased quantitative proteomics of lipid rafts reveals high specificity for signaling factors. Proc Natl Acad Sci U S A. 2003;100:5813–8. [Europe PMC free article] [Abstract] [Google Scholar]
86. Lents NH, Irintcheva V, Goel R, Wheeler LW, Baldassare JJ. The rapid activation of N-Ras by alpha-thrombin in fibroblasts is mediated by the specific G-protein Galphai2-Gbeta1-Ggamma5 and occurs in lipid rafts. Cell Signal. 2009;21:1007–14. [Abstract] [Google Scholar]
87. Maier U, Babich A, Macrez N, Leopoldt D, Gierschik P, et al. Gbeta 5gamma 2 is a highly selective activator of phospholipid-dependent enzymes. J Biol Chem. 2000;275:13746–54. [Abstract] [Google Scholar]
88. Roy S, Plowman S, Rotblat B, Prior IA, Muncke C, et al. Individual palmitoyl residues serve distinct roles in H-ras trafficking, microlocalization, and signaling. Mol Cell Biol. 2005;25:6722–33. [Europe PMC free article] [Abstract] [Google Scholar]
89. Prior IA, Muncke C, Parton RG, Hancock JF. Direct visualization of Ras proteins in spatially distinct cell surface microdomains. J Cell Biol. 2003;160:165–70. [Europe PMC free article] [Abstract] [Google Scholar]
90. Rodriguez-Viciana P, Sabatier C, McCormick F. Signaling specificity by Ras family GTPases is determined by the full spectrum of effectors they regulate. Mol Cell Biol. 2004;24:4943–54. [Europe PMC free article] [Abstract] [Google Scholar]
91. Vicinanza M, Di Campli A, Polishchuk E, Santoro M, Di Tullio G, et al. OCRL controls trafficking through early endosomes via PtdIns4,5P(2)-dependent regulation of endosomal actin. EMBO J. 2011;30:4970–85. [Europe PMC free article] [Abstract] [Google Scholar]
92. Watt SA, Kular G, Fleming IN, Downes CP, Lucocq JM. Subcellular localization of phosphatidylinositol 4,5-bisphosphate using the pleckstrin homology domain of phospholipase C delta1. Biochem J. 2002;363:657–66. [Europe PMC free article] [Abstract] [Google Scholar]
93. Sato M, Ueda Y, Takagi T, Umezawa Y. Production of PtdInsP3 at endomembranes is triggered by receptor endocytosis. Nat Cell Biol. 2003;5:1016–22. [Abstract] [Google Scholar]
94. Miaczynska M, Christoforidis S, Giner A, Shevchenko A, Uttenweiler-Joseph S, et al. APPL proteins link Rab5 to nuclear signal transduction via an endosomal compartment. Cell. 2004;116:445–56. [Abstract] [Google Scholar]
95. Christoforidis S, Miaczynska M, Ashman K, Wilm M, Zhao L, et al. Phosphatidylinositol-3-OH kinases are Rab5 effectors. Nature Cell Biology. 1999;1:249–52. [Abstract] [Google Scholar]
96. Shin HW, Hayashi M, Christoforidis S, Lacas-Gervais S, Hoepfner S, et al. An enzymatic cascade of Rab5 effectors regulates phosphoinositide turnover in the endocytic pathway. J Cell Biol. 2005;170:607–18. [Europe PMC free article] [Abstract] [Google Scholar]
97. Ciraolo E, Iezzi M, Marone R, Marengo S, Curcio C, et al. Phosphoinositide 3-kinase p110beta activity: key role in metabolism and mammary gland cancer but not development. Sci Signal. 2008;1:ra3. [Europe PMC free article] [Abstract] [Google Scholar]
98. Jia S, Liu Z, Zhang S, Liu P, Zhang L, et al. Essential roles of PI(3)K-p110beta in cell growth, metabolism and tumorigenesis. Nature. 2008;454:776–9. [Europe PMC free article] [Abstract] [Google Scholar]
99. Bohnacker T, Marone R, Collmann E, Calvez R, Hirsch E, et al. PI3Kgamma adaptor subunits define coupling to degranulation and cell motility by distinct PtdIns(3,4,5)P3 pools in mast cells. Sci Signal. 2009;2:ra27. [Abstract] [Google Scholar]
100. Kapeller R, Chakrabarti R, Cantley L, Fay F, Corvera S. Internalization of activated platelet-derived growth factor receptor-phosphatidylinositol-3′ kinase complexes: Potential interactions with the microtubule cytoskeleton. Mol Cell Biol. 1993;13:6052–63. [Europe PMC free article] [Abstract] [Google Scholar]
101. Backer JM, Myers J, MG, Shoelson SE, Chin DJ, Sun XJ, et al. The Phosphatidylinositol 3′-Kinase is Activated by Association with IRS-1 During Insulin Stimulation. EMBO J. 1992;11:3469–79. [Europe PMC free article] [Abstract] [Google Scholar]
102. Kelly KL, Ruderman NB. Insulin-stimulated phosphatidylinositol 3-kinase. Association with a 185-kDa tyrosine-phosphorylated protein (IRS-1) and localization in a low density membrane vesicle. J Biol Chem. 1993;268:4391–8. [Abstract] [Google Scholar]
103. Inoue G, Cheatham B, Emkey R, Kahn CR. Dynamics of insulin signaling in 3T3-L1 adipocytes - Differential compartmentalization and trafficking of insulin receptor substrate (IRS)-1 and IRS-2. J Biol Chem. 1998;273:11548–55. [Abstract] [Google Scholar]
104. Foukas LC, Claret M, Pearce W, Okkenhaug K, Meek S, et al. Critical role for the p110alpha phosphoinositide-3-OH kinase in growth and metabolic regulation. Nature. 2006;441:366–70. [Abstract] [Google Scholar]
105. Metjian A, Roll RL, Ma AD, Abrams CS. Agonists cause nuclear translocation of phosphatidylinositol 3-kinase gamma. A Gbetagamma-dependent pathway that requires the p110gamma amino terminus. J Biol Chem. 1999;274:27943–7. [Abstract] [Google Scholar]
106. Kumar A, Redondo-Munoz J, Perez-Garcia V, Cortes I, Chagoyen M, et al. Nuclear but not cytosolic phosphoinositide 3-kinase beta has an essential function in cell survival. Mol Cell Biol. 2011;31:2122–33. [Europe PMC free article] [Abstract] [Google Scholar]
107. Zwaenepoel O, Tzenaki N, Vergetaki A, Makrigiannakis A, Vanhaesebroeck B, et al. Functional CSF-1 receptors are located at the nuclear envelope and activated via the p110delta isoform of PI3-kinase. FASEB journal: official publication of the Federation of American Societies for Experimental Biology. 2012;26:691–706. [Abstract] [Google Scholar]
108. Vanhaesebroeck B, Jones GE, Allen WE, Zicha D, Hooshmand-Rad R, et al. Distinct PI(3)Ks mediate mitogenic signalling and cell migration in macrophages. Nature Cell Biology. 1999;1:69–71. [Abstract] [Google Scholar]
109. Dhand R, Hiles I, Panayotou G, Roche S, Fry MJ, et al. PI 3-kinase is a dual specificity enzyme: Autoregulation by an intrinsic protein-serine kinase activity. EMBO J. 1994;13:522–33. [Europe PMC free article] [Abstract] [Google Scholar]
110. Vanhaesebroeck B, Higashi K, Raven C, Welham M, Anderson S, et al. Autophosphorylation of p110δ phosphoinositide 3-kinase: a new paradigm for the regulation of lipid kinases in vitro and in vivo. EMBO J. 1999;18:1292–302. [Europe PMC free article] [Abstract] [Google Scholar]
111. Czupalla C, Culo M, Muller EC, Brock C, Reusch HP, et al. Identification and characterization of the autophosphorylation sites of phosphoinositide 3-kinase isoforms beta and gamma. J Biol Chem. 2003;278:11536–45. [Abstract] [Google Scholar]
112. Bondeva T, Pirola L, Bulgarelli-Leva G, Rubio I, Wetzker R, et al. Bifurcation of lipid and protein kinase signals of PI3Kgamma to the protein kinases PKB and MAPK. Science. 1998;282:293–6. [Abstract] [Google Scholar]
113. Lam K, Carpenter CL, Ruderman NB, Friel JC, Kelly KL. The phosphatidylinositol 3-kinase serine kinase phosphorylates IRS-1. Stimulation by insulin and inhibition by Wortmannin. J Biol Chem. 1994;269:20648–52. [Abstract] [Google Scholar]
114. Naga Prasad SV, Jayatilleke A, Madamanchi A, Rockman HA. Protein kinase activity of phosphoinositide 3-kinase regulates beta-adrenergic receptor endocytosis. Nat Cell Biol. 2005;7:785–96. [Abstract] [Google Scholar]
115. Damilano F, Perino A, Hirsch E. PI3K kinase and scaffold functions in heart. Ann N Y Acad Sci. 2010;1188:39–45. [Abstract] [Google Scholar]
116. Bi L, Okabe I, Bernard DJ, Nussbaum RL. Early embryonic lethality in mice deficient in the p110beta catalytic subunit of PI 3-kinase. Mamm Genome. 2002;13:169–72. [Abstract] [Google Scholar]
117. Perino A, Ghigo A, Ferrero E, Morello F, Santulli G, et al. Integrating cardiac PIP3 and cAMP signaling through a PKA anchoring function of p110gamma. Mol Cell. 2011;42:84–95. [Europe PMC free article] [Abstract] [Google Scholar]
118. Milne SB, Ivanova PT, DeCamp D, Hsueh RC, Brown HA. A targeted mass spectrometric analysis of phosphatidylinositol phosphate species. J Lipid Res. 2005;46:1796–802. [Abstract] [Google Scholar]
119. Wang YH, Collins A, Guo L, Smith-Dupont KB, Gai F, et al. Divalent cation-induced cluster formation by polyphosphoinositides in model membranes. J Am Chem Soc. 2012;134:3387–95. [Europe PMC free article] [Abstract] [Google Scholar]
120. Leopoldt D, Hanck T, Exner T, Maier U, Wetzker R, et al. Gβgamma stimulates phosphoinositide 3-kinase-gamma by direct interaction with two domains of the catalytic p110 subunit. J Biol Chem. 1998;273:7024–9. [Abstract] [Google Scholar]

Citations & impact 


Impact metrics

Jump to Citations
Jump to Data

Citations of article over time

Smart citations by scite.ai
Smart citations by scite.ai include citation statements extracted from the full text of the citing article. The number of the statements may be higher than the number of citations provided by EuropePMC if one paper cites another multiple times or lower if scite has not yet processed some of the citing articles.
Explore citation contexts and check if this article has been supported or disputed.
https://scite.ai/reports/10.1002/bies.201200176

Supporting
Mentioning
Contrasting
0
38
0

Article citations


Go to all (21) article citations

Data 


Similar Articles 


To arrive at the top five similar articles we use a word-weighted algorithm to compare words from the Title and Abstract of each citation.

Funding 


Funders who supported this work.

NCI NIH HHS (2)

NIA NIH HHS (3)

NIGMS NIH HHS (4)