Next Article in Journal
Intracranial Solitary Fibrous Tumour Management: A French Multicentre Retrospective Study
Previous Article in Journal
C-Reactive Protein and Lymphocyte-to-Monocyte Ratio Predict Recurrence in Stage III Melanoma Patients with Microscopic Sentinel Lymph Node Metastasis
Previous Article in Special Issue
The Role of Genomics and Proteomics in Lung Cancer Early Detection and Treatment
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Current State and Future Challenges for PI3K Inhibitors in Cancer Therapy

1
Department of Medical Oncology, IRCCS Istituto Romagnolo per lo Studio dei Tumori (IRST) “Dino Amadori”, 47014 Meldola, Italy
2
Department of Life Sciences, University of Bath, Bath BA2 7AY, UK
3
Department of Oncology, Royal United Hospital, Bath BA1 3NG, UK
*
Author to whom correspondence should be addressed.
Cancers 2023, 15(3), 703; https://doi.org/10.3390/cancers15030703
Submission received: 2 January 2023 / Revised: 16 January 2023 / Accepted: 19 January 2023 / Published: 23 January 2023
(This article belongs to the Special Issue PI3K Pathway in Cancer)

Abstract

:

Simple Summary

Phosphatidylinositol 3-kinase (PI3K) is a key regulator of many cellular processes and its hyperactivation promotes tumor cell growth and survival. A broad evaluation of the upstream and downstream nodes of its pathway allowed the discovery of several PI3K inhibitors (PI3Ki) with anti-tumor activity. However, the highly intrinsic toxicity and the onset of therapeutic resistance can limit their clinical application. To increase the antitumor effect and the therapeutic index, combination strategies and new dosing schedules have been investigated. However, further efforts are necessary to discover potentially actionable genetic alterations towards the goal of precision medicine.

Abstract

The phosphoinositide 3 kinase (PI3K)-protein kinase B (PKB/AKT)-mammalian target of the rapamycin (mTOR) axis is a key signal transduction system that links oncogenes and multiple receptor classes which are involved in many essential cellular functions. Aberrant PI3K signalling is one of the most commonly mutated pathways in cancer. Consequently, more than 40 compounds targeting key components of this signalling network have been tested in clinical trials among various types of cancer. As the oncogenic activation of the PI3K/AKT/mTOR pathway often occurs alongside mutations in other signalling networks, combination therapy should be considered. In this review, we highlight recent advances in the knowledge of the PI3K pathway and discuss the current state and future challenges of targeting this pathway in clinical practice.

1. Introduction

Discovered in the late 1980s, the family of lipid kinases named phosphoinositide 3-kinase (PI3K) and the correlated PI3K/AKT signalling pathway have been shown to play a pivotal role in different oncogenic processes including cell survival, metabolism and metastasis [1]. The classical mechanisms behind the PI3K/AKT/mTOR pathway activation and its functions are described in Figure 1. Toward this goal, PI3K converts different signals from cytokines and growth factors into intracellular responses by producing phospholipids which, in turn, triggers the serine-threonine protein kinase AKT and downstream pathways [2]. While mTOR is one of the most common downstream effectors, the main critical regulator of the PI3K/AKT pathway is the phosphatase and tensin homologue (PTEN) tumour suppressor [3]. The PI3K/AKT pathway can be abnormally triggered in a wide range of cancers due to a plethora of mechanisms including somatic mutations and germline mutations in PIK3CA, AKT, PTEN and mTOR genes [4].
As a result, the PI3K/AKT pathway can be targeted by pharmacological molecules, thus making this pathway an interesting target for cancer intervention [5,6]. However, many issues regarding the use of pathway inhibitors, as well as the most effective drug to use in clinical practice, up to what cancer subtype might benefit the most from PI3K/Akt inhibitors, also due to the side effects, remain to be unsolved. Moreover, emerging evidence suggests that the PI3K/Akt pathway plays an immunomodulatory role [7]. In fact, several studies underlined how the PI3K pathway is involved in the differentiation of myeloid-derived suppressor cells (MDSCs) and Tregs into the tumor as well as the secretion of suppressive cytokines to impair stimulation of macrophages and dendritic cells, leading to an immunosuppressive tumour microenvironment (TME) [8]. This evidence suggests a potential synergy for combining PI3K inhibitors (PI3Kis) and immune-checkpoint inhibitors (ICIs).
In this review, we describe the critical role of the PI3K/AKT/mTOR pathway in tumorigenesis and the challenges in the clinical development of antitumour therapies targeting the PI3K/AKT/mTOR pathway, highlighting their limited clinical application. Finally, we provide an overview of the emerging data regarding PI3K/AKT/mTOR inhibitors in the most recent clinical trials, as well as their efficacy alone or in combinations for both solid and hematologic malignancies.

2. PI3K/AKT/mTOR Signalling in Cancer

The PI3K/Akt/mTOR pathway has been associated with the development and progression of different neoplastic diseases [9]. For instance, almost 70% of breast [10] and ovarian cancers [11] carry an alteration of PI3K/AKT; similarly, the aberrant activation of the PI3K/AKT/mTOR pathway has been identified in 90% of lung adenocarcinomas (ADCs) and 40% of squamous cell carcinomas (SCCs) [12], leading to its hyperactivation.
Physiologically, the activity and homeostasis of the PI3K/AKT/mTOR pathway are strictly controlled by regulatory mechanisms; nevertheless, this pathway can be constitutively activated in several cancers. There are different mechanisms underlying this abnormal activatio including inactivating mutations in tumor suppressors genes such as PTEN or INPP4B, genomic alterations in PIK3CA, PIK3R1 (p85α regulatory subunit) or PIK3R2 (p85β regulatory subunit) and Akt subunits [13]. Mutations or overexpression of growth factor receptor (EGFR) or human growth factor receptor 2 (HER2), inactivating mutations in mTOR regulators gene such as TSC1 and TSC2, as well as the activating mutations in mTOR itself, are also detected across cancer types [14,15,16,17].

2.1. PI3Ks

The phosphoinositide 3-kinase (PI3K) family has an important role in a wide range of aspects of cell and tissue biology and a crucial role in human cancer [18]. The majority of PI3K functions are mediated by phosphoinositides, the low-abundance phosphorylated forms of phosphatidylinositol [19]. There are three different classes (I-II-II) of PIK3 according to their structural and specificity features. Class I PI3Ks are the most investigated and clinically interesting as they can be directly activated by cell surface receptors including G protein-coupled receptors (GPCRs), receptor tyrosine kinases (RTKs) and oncogenes such as G protein RAS [20].

2.1.1. Class I

Activated as heterodimers, Class I PI3Ks are made of a regulatory (p85) and a catalytic subunit (p110), and they trigger downstream tyrosine kinases including GTPases and GPCRs [21]. It is worth mentioning that class I PI3Ks consist of four different catalytic isoforms (p110α, p110β, p110γ, p110δ) that are, respectively, expressed by PIK3CA, PIK3CB, PIK3CG and PIK3CD genes [22]. The most frequently mutated isoform in cancer is PIK3CA, whose mutations are an early event in colon and breast cancer [23]. While oncogenic mutations in PIK3CB are rare, reduced expression of PIK3CG expression has been associated with colon cancer development and progression [24]. On the contrary, PIK3CD is commonly found to be expressed in leukocytes and B cells, exerting a critical role in their growth and survival [25]. Of note, class I PI3Ks are further categorised into class IA and IB based on dissimilarity in the regulatory subunits. Several mechanisms are involved in the oncogenic activation of class IA PI3K including the inactivation of PTEN and p110 catalytic subunits, to mention a few [26].

2.1.2. Class II

Activated as monomers, Class II PI3Ks consist of three catalytic components (C2α, C2β and C2γ) and no regulatory subunit [27]. They currently work as important signalling proteins with major roles under normal and pathological circumstances [28]. Indeed, while it has been demonstrated that PI3KC2α and PI3KC2β are widely expressed in the human body, the former plays a critical role in breast cancer invasiveness by impairing mitotic spindle formation [28,29]. Noteworthy, class II PI3Ks are engaged in the unique lipid molecule expression, with a critical function in cellular processes [30,31].

2.1.3. Class III

Class III PI3K VPS34 exerts its role in regulating macrophage phagocytosis and autophagy by connecting itself to a protein complex consisting of a catalytic and a regulatory subunit [22,30]. It has been shown that once activated, VPS34 is involved in transducing signals by modulating different protein kinases rather than directly regulating signalling pathways [32]. Indeed, emerging evidence has shown that VPS34 can modulate the basal activity of mTOR complex 1 (mTORC1) in animal models and the glycogen synthetase kinase 3 (GSK3) pathway in breast cancer patients treated with AKT inhibitors [33]. According to these observations, strategies targeting VPS34 might be an effective clinical treatment approach.

2.2. AKT

Also known as protein kinase B (PKB), the serine and threonine kinase AKT has been investigated since the early 90s and its dysfunction was observed in several diseases including cancer [34,35]. AKT1, AKT2 and AKT3 are the isoforms that have been identified so far and, although being found broadly expressed in the human body, they exert different and critical roles in cancer: for example, while AKT2 expression has been observed to increase in pancreatic cancer with a major role in cell migration and invasion, AKT3 expression was found increased in prostate and breast cancer [36]. Once triggered by receptor tyrosine kinases (RTKs), AKT can recruit and engage single or multiple subtypes of class I PI3Ks on the cell surface. In turn, activated PI3K enables PIP2 to PIP3 conversion and the consequent phosphoinositide-dependent-kinase-1 (PDK1) activation [37]. However, the regulation of AKT protein is a very complex system that includes other modulating factors such as IGF1, TRAF6 and TBK1 [38]. A new research thread is currently focusing on AKT in order to control class I PI3K signalling.

2.3. PTEN

Originally discovered as a mutated lipid phosphatase protein in a plethora of cancers, PTEN is now widely considered a tumour suppressor with a crucial role in the PI3K signalling pathway by preserving physiological cell activity [39]. It has been also observed that PTEN exerts its role in modulating the PI3K pathway by suppressing PIP2 to PIP3 conversion [40]. When PTEN is mutated or its function impaired, PI3K effectors, including AKT, become activated with no need for external oncogenic stimulus [41]. is normally involved in tumour signalling by dephosphorylation of targets including PTEN itself, insulin receptor substrate 1 (IRS1) and focal adhesion kinase (FAK) [42]. In cancers lacking PTEN function, the increased activation of AKT is a major oncogenic strategy [43]. It has also been demonstrated that PTEN is actively involved in angiogenesis and cancer cell migration [44].

2.4. mTOR

One of the main downstream targets of the PI3K/AKT pathway is mTOR, a protein kinase reported to regulate tumour development, metabolism, survival, angiogenesis and immunity [45]. When assembled in major complexes (mTORC1 and mTORC2), mTOR exerts significant roles in different physiological and pathological biological activities. While mTORC2 consists of mTOR, Rictor, SIN1 and mLST8 subunits, mTORC1 consists of mTOR, PRAS40, raptor and mLST8 subunits with a major role in modulating cell growth by phosphorylation of elF-4E-binding protein 1 (4EBP1) and S6 kinase 1 (S6K1) [46]. However, Akt was observed to have a close collaboration with mTOR via activation of the latter by phosphorylating the tuberous sclerosis complex 2 (TSC2) [47]. Once the mTORC2 is assembled, AKT is phosphorylated and activated by mTORC2 [48]. For this reason, mTOR inhibition has raised significant interest in clinical cancer research. According to experimental data and computer simulations, PI3K/AKT and MERK/ERK pathways can interact, activating or inhibiting each other with a context-dependent cross-talk [49]. Furthermore, different studies have shown that the blockade of one pathway may activate the other signalling cascade. For instance, PI3K inhibition induces the ERK2-dependent reactivation of AKT, eliminating the anti-clonogenic effect of inhibitors [50]. Therefore, the block of both MEK and PI3K/AKT/mTOR pathways with a combination of different signalling inhibitors may be used to more effectively target tumor cells, as compared with treatment with a single agent (Figure 1).

3. Targeting PI3K Pathway in Cancer

The PI3K/AKT/mTOR pathway is frequently activated in a wide variety of cancers including breast, gastric, ovarian, colorectal, prostate, glioblastoma and endometrial cancers [27]. In addition, it plays a key role in cell survival, proliferation, differentiation and glucose transport [13], and its hyperactivation can induce resistance to antitumor treatment [51,52]. As a result, the PI3K pathway and associated components have become an attractive anticancer drug target.

3.1. PI3K Inhibitors

Recently, a plethora of PI3Kis have been investigated in clinical trials and subsequently approved as potential chemotherapeutic drugs for cancer therapy. PI3Kis are grouped into pan-PI3K inhibitors, isoform-selective PI3K inhibitors and dual PI3K/mTOR.

3.1.1. Pan-PI3K Inhibitors

The first generation of PI3Kis, defined as pan-PI3Ki, target all four catalytic isoforms of class I PI3Ks (α, β, γ, and δ) [53,54] and includes pictilisib (GDC-0941), buparlisib (BKM120) and copalinsib. These small molecules have shown a broad spectrum of activities, as well as a broader inhibition, leading to severe adverse events and treatment discontinuation [55]. Despite their high toxicity, multiple efforts have been carried out to develop agents accurately targeting PI3K isoforms to improve the therapeutic outcome.
Pictilisib (GDC-0941), a thienopyrimidine molecule, is a robust, selective, orally available PI3K inhibitor and the first PI3Ki to be assessed in a clinical trial. This drug demonstrated solid efficacy in human tumor xenografts murine models of U87MG glioblastoma and IGROV1 ovarian cancer, alone or in combination with other targeted therapy [20,56]. In a phase 1 dose-escalation clinical trial, a partial response was observed in a patient with V600E BRAF-mutant melanoma and in a patient with platinum-refractory epithelial ovarian cancer, exhibiting PTEN loss and PIK3CA amplification [57]. The most common toxicities were low-grade nausea, rash and fatigue, with one patient reporting grade 3 hyperglycemia [57].
Buparlisib (BKM120) is another potent and selective pan-class I PI3Ki which can cross the blood–brain barrier and potentially lead to PI3K inhibition in the brain [58]. Speranza et al. have found out that buparlisib has potent anti-invasive effects in glioblastoma cell lines and in in vitro patient-derived glioma cells, with no significant adverse effects [59]. As a consequence of PI3K inhibition in the central nervous system, a small number of patients experienced mood alterations such as anxiety, irritability or depression, which are generally mild and responsive to dose reductions [58,60].
In a very small subset of triple-negative breast cancer (TNBC) patients treated with buparlisib, Garrido-Castro et al. observed a prolonged stable disease (SD), although an objective response was not confirmed [61]. Given the emerging evidence regarding acquired endocrine resistance and PI3K activation [62,63], buparlisib has been evaluated in hormone receptor-positive/HER-2 negative metastatic breast cancer (HR+/HER2− MBC). In a phase 1b trial, Mayer et al. showed that the combination of buparlisib plus letrozole was safe with reversible toxicity in HR+/HER2− MBC refractory to endocrine therapy (ET) [60]. In the same setting, the BELLE 3 trial demonstrated that buparlisib plus fulvestrant was associated with better progression-free survival (PFS) compared to fulvestrant alone (median 3.9 months versus 1.8 months; HR 0.67 CI 0.53–0.84; p < 0.001), especially in patients with a real-time polymerase chain reaction (RT-PCR) or ctDNA PIK3CA mutations [64]. Alongside the positive findings with endocrine therapy, the addition of buparlisib to chemotherapy did not improve the median PFS [65]. Unfortunately, buparlisib administration was limited due to the high metabolic and psychiatric toxicity that emerged from clinical trials leading to the discontinuation of drug development. Additional data are required to confirm the clinical use of buparlisib.
Copanlisib is a potent, highly selective, pan-class I PI3Ki with a predominant activity against the isoforms p110α and p110δ [66]. It is administered intravenously while the other PI3K inhibitors, such as idelalisib and duvelisib, are normally administered orally [66]. Copanlisib was approved by FDA for the treatment of adult patients with relapsed follicular lymphoma (FL) who received at least two prior systemic therapies [67]. Interestingly, copanlisib is being studied in advanced HER2+ BC in addition to pertuzumab and trastuzumab to evaluate if this combination can overcome the resistance caused by the hyperactivation of the PI3K pathway [68]. Its use is associated with hypertension, diarrhoea and transient hyperglycemia, which is a common and predictable effect of PI3Kα inhibition due to the abrogation of downstream insulin receptor signalling [69].

3.1.2. Isoform-Selective PI3K Inhibitors

Although they require a stricter selection of patients, isoform-selective PI3K inhibitors are characterised by improved efficacy and fewer adverse events (AEs) compared to pan-PI3Ki [70]. The safer profile allowed isoform-selective PI3Ki to be developed and approved for clinical practice.
Alpelisib (BYL719) is the first oral isoform-selective PI3Ki targeting the p110α isoform of wild-type PI3Kα to be approved by the US Food and Drug Administration (FDA) and by the European Medicines Agency (EMA) [71]. Compared to the other isoform, alpelisib specificity induces a 50 times stronger activity against PI3Kα [72,73]. In 2019, Juric et al. conducted a phase 1b trial to assess the maximum tolerated dose of alpelisib (MTD) in patients with HR+/HER2− BC [74]. They observed that alpelisib plus fulvestrant lead to an improvement in PFS and OS in patients with PI3KCA alterations compared to the wild-type group, with manageable toxicity [74]. Similarly, Andrè et al. conducted a phase 3 clinical trial of alpelisib in combination with fulvestrant for the treatment of MBC [75]. They observed an increased PFS (7.4 months versus 5.6 months; HR: 0.85 95% CI 0.58–1.25) and objective response (OR) in patients treated with alpelisib and fulvestrant compared to the control arm (26% versus 23.8%). Nevertheless, patients in the experimental arm experienced a higher rate of hyperglycemia, rashes and diarrhoea compared to the placebo arm [75]. Following the aforementioned promising outcomes, alpelisib was approved by FDA in 2019 for the treatment of PIK3CA-mutant, HR+/HER2− MBC [65]. It is worthwhile to point out that alpelisib was approved in the USA with the companion diagnostic test Therascreen® PIK3CA RGQ PCR kit (Qiagen, Hilden, Germany).
In 2022, Rugo et al., using comprehensive genomic profiling (CGP), detected approximately 72% PIK3CA mutations (PIK3Cm) in tissue biopsies from 33,977 patients with MBC and demonstrated that up to 20% of patients carried PIK3CA mutations which have not been identified by Therascreen® PIK3CA [71]. Of note, this study found out that these patients with different PI3KCAm also have longer PFS when administered with alpelisib plus fulvestrant compared to fulvestrant alone [71].
So far, the optimal method to detect PIKCAm in clinical practice is not yet established, and prospective clinical trials are warranted to demonstrate the PI3K inhibitors benefit in patients with PI3KCA mutations, not included in the SOLAR-1 trial.
Taselisib (GDC-0032) is a novel potent inhibitor of PI3Kα, exerting its blocking activity onp110ɑ, p110γ and p110δ proteins. In a preclinical study, taselisib showed significant antiproliferative activity in head and neck squamous carcinomas (HNSCC) cell lines harbouring PIK3CA-activating mutations [76]. Moreover, in the same setting, the combination of taselisib and radiotherapy was more efficacious than treatment alone both in vitro and in vivo [76]. Following preclinical studies, taselisib showed clinical activity in a phase I dose-finding clinical trial in patients with PIK3CA-mutant solid tumors, especially in MBC with a 36% of overall response rate (ORR) [74]. Consequently, Baselga et al. evaluated the efficacy of taselisib and fulvestrant in a phase 3 trial for HR+/PI3KCA-mutated MBC patients. The study reported a modest PFS increase in the efficacy of the combination treatment compared to fulvestrant alone (median PFS 7.4 months versus 5.4 months; HR 0.70, p < 0.01). However, the modest PFS improvement was associated with significant toxicity, especially diarrhoea (grade 3/4 of 12% vs. <1% for hormonal therapy alone) and hyperglycemia (grade 3/4 of 11% for the taselisib arm vs. <1% for the control arm), resulting in discontinuation of drug development for this subgroup [77,78].
The greater selectivity for the mutant PI3Kɑ isoform and the stronger inhibitory effect may justify why taselisib is correlated with a worse toxicity profile compared to alpelisib [44].
Idelalisib (Zydelig) is an orally bioavailable ATP-competitive kinase inhibitor specifically designed to target the phosphoinositide 3-kinase p110 isoform δ (PI3Kδ) with accurate selectivity and potency [76]. Due to its hyperactivation in B-cell malignancies and its crucial role in the B-cell receptor (BCR) pathway, it has been approved by FDA in 2014 for the treatment of indolent B-cell malignancies including relapsed/refractory chronic lymphocytic leukaemia (CLL), in association with rituximab, as monotherapy for relapsed follicular lymphoma (FL) and relapsed small lymphocytic leukaemia (SLL), in patients who received at least two prior systemic therapies [79]. Several clinical trials are ongoing to determine the activity, efficacy, and toxicity profile of PIK3CA inhibitors alone or in combination (Table 1).

3.2. AKT Inhibitors

Due to its role as a key-molecular regulator of the PI3K/AKT/mTOR pathway, AKT could be an interesting target. Indeed, AKT inhibition induces the block of mTORC1 activation, leading to the control of the downstream effects of the PI3K/AKT/mTOR cascade [11,77].
The majority of AKT inhibitors investigated so far in clinical trials can inhibit all three AKT subunits and, for this reason, they are defined as pan-AKT inhibitors. Several Akt-inhibitors, such as MK2206, capivasertib (AZD5363), afuresertib (GSK2110183) and ipatasertib have been developed to target AKT signalling in vitro and in vivo; however, none of them has yet received FDA approval for cancer treatment.
MK220 is a first-in-class allosteric AKT1/2/3 inhibitor with evidence of preclinical efficacy when combined with cytotoxic agents including doxorubicin, gemcitabine, docetaxel and carboplatin in the lung NCI-H460 cell line [41].
In preclinical studies, this drug has been demonstrated to restore erlotinib activity in erlotinib-sensitive and resistant non-small cell lung cancer (NSCLC) cell lines [80]. Additionally, it showed encouraging anti-tumour activity in acute myeloid leukaemia (AML) [81] and an ability to inhibit both AKT and mTOR signalling in nasopharyngeal carcinoma (NPC) cell lines [82]. Moreover, MK220 has shown preliminary activity in different phase I trials [81,83], and it is being currently tested in phase II trials as a monotherapy in metastatic pancreatic cancer [84] or in combination with the MEK inhibitor selumetinib (+MK2206) in colon-rectal cancer [84].
Capivasertib (AZD5363) is a novel, selective ATP-competitive pan-AKT kinase inhibitor that exerts activity against the three AKT isoforms (AKT1, AKT2 and AKT3) [85]. It can potentially treat a wide range of solid and hematologic malignancies as a monotherapy or in combination, both in vivo and in vitro [86]. Tumor types carrying PTEN mutation, PIK3CA mutation, or HER2 amplification, without coincident RAS mutation, are strongly associated with preclinical sensitivity to capivasertib [87]. In preclinical BC models, capivasertib can overcome resistance or increase sensitivity to HER2 inhibitors and improve chemotherapy efficacy, leading to tumor regression [87]. Similarly, capivasertib as a monotherapy or combined with different drugs has demonstrated preclinical efficacy in castrate-resistant prostate cancer (CRPC) [88], PI3KCA-mutant gastric cancer [89], trastuzumab-resistant esophagal squamous cell carcinoma [90] and NSCLC [91]. Furthermore, in 2020, Smith et al. demonstrated that capivasertib alone or in combination with fulvestrant was well tolerated and showed promising anticancer activity in patients with AKT1E17K-mutant HR+/MBC in a phase I expansion study [92]. In the same setting, the phase 2 randomized FAKTION trial, demonstrated that the addition of capivasertib to fulvestrant resulted in a significantly longer PFS [93]. At the 2022 San Antonio Breast Cancer Symposium (SABCS), Turner presented the results from the CAPItello-291 phase III trial, showing a statistically significant and clinically meaningful improvement in PFS with the combination of fulvestrant and capivasertib in patients with HER2+/HER2-low or negative MBC, following recurrence or progression on, or after, endocrine therapy (with or without a CDK4/6 inhibitor) [94].
Given the synergy between poly(ADP-ribose) polymerase (PARP) and PI3Ki in preclinical data, Trap et al., evaluated capivasertib and the PARP inhibitor (PARPi) olaparib in a phase 1 study [91]. They observed that capivasertib was safe and well tolerated. Furthermore, antitumor activity was reported in both patients harbouring germline BRCA1/2 mutations and BRCA1/2 wild-type and with or without the somatic DNA damage repair gene (DDR) and/or PI3K/AKT pathway alterations [95].
It is worth mentioning that large-scale genomic studies of human cancer demonstrated that AKT1-E17K is the most common AKT mutation and improves the efficacy of AKT inhibitor therapy in solid tumors [87,96]. In a multicohort basket study, capivasertib obtained promising PFS outcomes in heavily pretreated AKT1 E17K-mutant breast and gynecologic cancer patients [96]. However, the response rate was lower than the response to those therapies targeting EGFR, ALK, ROS1 and BRAF. As a consequence, the full potential of capivasertib in AKT1-mutant cancers may require drug combination.
The large phase 2 screening trial MATCH (Molecular Analysis for Therapy Choice) (NCT02465060) is ongoing to match targeted therapy in 6452 patients with solid tumors or lymphomas, harbouring specific mutations, that have progressed to first-line standard treatment. Capivasertib and ipatasertib are eligible for patients with AKT mutations, while the PI3K inhibitor GSK2636771 is indicated for patients with PTEN mutation or deletion. On the other hand, the MyTACTIC trial is a phase II, multi-arm study investigating the safety and efficacy of targeted therapies in unresectable or metastatic solid tumors harbouring genomic alterations or protein expression patterns, predictive of response. This trial includes ipatasertib for patients with AKT1/2/3 mutations or PTEN loss of function and inavolisib for patients with PIK3CA mutations.
Afuresertib is another ATP-competitive AKT inhibitor that has been investigated in a phase Ib/II dose escalation study in combination with carboplatin and paclitaxel in recurrent platinum-resistant ovarian cancer. The study reported an ORR of 32% and a median PFS of 7.1 months [97]. Ultimately, ipatasertib is an additional highly selective oral ATP-competitive pan-AKT inhibitor showing encouraging activity, especially in tumors with markers of AKT activation, including high-basal phospho-AKT levels, PTEN loss and PIK3CA kinase domain mutations [98,99]. In a phase 1b trial, ipatasertib in combination with chemotherapy or hormone therapy was well tolerated and demonstrated radiographic responses in patients with MBC with a safe toxicity profile [100]. Similarly, the phase 2 randomized LOTUS trial showed an improvement in PFS with the addition of ipatasertib to paclitaxel in TNBC [101]. Conversely, in the phase 3 IPATunity130 trial, the addition of ipatasertib to paclitaxel did not improve efficacy in PIK3CA/AKT/PTEN-altered HR+/HER2− MBC [102]. These findings are consistent with the results of the BEECH trial where the combination of paclitaxel and an AKT inhibitor did not improve the PFS neither in the overall population nor in the PIK3CA-altered population [103]. A possible explanation can be the higher number of patients discontinuing paclitaxel due to ipatasertib adverse events (AEs) [102]. From the SOLAR1 and the FAKTION trial, it appears that endocrine blockade may be essential in order to obtain greater clinical benefit from AKT inhibition in HR+/HER2− MBC.
Different clinical trials are ongoing to assess the activity, efficacy, and toxicity profile of PIK3i alone or in combination with other target therapies (Table 2).

3.3. mTORC1 and mTORC2 Inhibitors

mTOR is a protein kinase which is extensively associated with cell growth, metabolism, survival, catabolism and autophagy [94], and is observed hyperactive in 40 to 90% of solid tumors [95]. mTOR is a downstream effector of the PI3K oncogenic pathway and it is the main reason behind the development of catalytic domain inhibitors, capable of blocking both mTOR and PI3K [96]. Rapamycin and its analogue (everolimus, temsirolimus and deforolimus) represent the first generation of mTOR inhibitors (mTORi), which are able to selectively inhibit the mTORC1 activity [97]. Rapamycin is a natural product that inhibits mTOR with high specificity [98]; however, its clinical application was limited due to its poor solubility and stability [97]. Therefore, rapamycin analogues with better solubility and metabolic properties have been developed. Water-soluble temsirolimus and deforolimus can be administered intravenously, while rapamycin and everolimus have lower solubility and can be administered orally [99].
In addition, rapamycin dosage may also affect mTOR activity [104]. Indeed, low nanomolar doses of rapamycin can impair S6K phosphorylation by mTORC1, delaying G1 cell-cycle progression [105,106]. On the other hand, micromolar doses of rapamycin might suppress the phosphorylation of both S6K and 4E-BP1 [107,108]. Unfortunately, this treatment can frequently result in a feedback activation of AKT phosphorylation by mTORC2 [109,110], which promotes cell survival [104,111].

3.3.1. ATP-Competitive mTOR Inhibitors

In order to more efficiently inhibit mTOR, a second generation of mTOR inhibitors targeting both mTORC1 and mTORC2 have been developed, also called selective mTOR kinase inhibitors (TORKIs) [112]. These small molecules, classified as ATP analogues, provide a robust inhibition of both mTORC1/2, and can reduce the resistance observed with rapamycin analogues [112]. Although ATP analogues showed a higher inhibitory effect in preclinical studies [113], large clinical trials have not been conducted yet and TORKIs are still not approved for clinical use.

3.3.2. Dual PI3K/mTOR Inhibitors

Even though the inhibition of mTORC1 and mTORC2 can lead to a downregulation of AKT S473 phosphorylation, mTOR inhibition may paradoxically induce the activation of the PI3K/PDK1 axis. Therefore, the inhibition of both PI3K and mTOR may enhance anti-tumor activity compared to the mTOR-block alone [114,115].
Dual PI3K/mTOR inhibitors (PI3K/mTORi) include SF1126, dactolisib (BEZ235), voxtalisib (XL765) and gedas (PKI-587) [57]. Dactolisib elicited antitumor activity in human glioblastoma (GBM) cell lines and an orthotopic xenograft model [101], whereas voxtalisib showed encouraging efficacy with an acceptable safety profile in patients with follicular lymphoma [102]. Moreover, in T-cell acute lymphoblastic leukaemia (T-ALL), the dual-specificity of PI3K/mTORi PKI-587 was the most selective for T-ALL cells dependent on the PI3K/mTOR pathway [103]. Finally, this class of drugs has the potential to treat tumors with a wide range of genetic abnormalities including PTEN and TSC1/2 loss of function and STK11 alterations [35], with the latter being found in a third of NSCLC and associated with KRAS mutations [116]. Additionally, they exhibit a broad activity profile and significantly higher toxicity [117] (Figure 2).
Several trials are currently ongoing to establish the efficacy of dual PI3K/mTORi (Table 3).

3.4. Combination Strategies

Acquired and intrinsic drug resistance with monotherapy is a major limit to PI3K inhibitors efficacy and it may be attributed to the complex feedback in the PI3K/AKT/mTOR signalling and its crosstalk with other pathways. Considering the well-establish evidence from preclinical studies, potential drug combinations may include chemotherapy, kinase inhibitors and ICIs.

3.4.1. Her-2 Inhibitors

Aberrant activation of the ErbB family of receptors is one of the most common causes of cancer [118]. EGFR and HER2 are members of the ErbB family of RTKs and they have a crucial role in cell proliferation and survival [119]. For instance, an important group of studies showed that her2/neu gene amplification is common in human BC and it is correlated with poor prognosis [120]. To date, targeting the HER2-receptor has significantly changed cancer therapy, preventing signal initiation and crosstalk with complementary pathways but also improving the sensitivity of tumor cells to both chemotherapy and radiation [121]. In a phase 1b trial, Zambrano et al. observed that buparlisib might be combined with paclitaxel trastuzumab in HER2+ MBC [122]. Similarly, Pistilli et al. showed that buparlisib plus trastuzumab regimen has an acceptable safety profile but limited efficacy in patients with heavily pretreated and trastuzumab-resistant HER2+ MBC, and patients with progressive brain metastases also receiving capecitabine [120].

3.4.2. MAPK Inhibitors

It is well established that PI3K/AKT/mTOR and RAS/RAF/MEK/ERK pathways interact with each other at several nodes, leading to a potential pathway convergence for the development of drug combinations [123]. Indeed, parallel activation of the PI3K/AKT/mTOR pathway may be responsible for primary and acquired resistance to BRAF-targeted therapy [124]. The results of a phase 1b trial by Shapiro et al. showed that the MEK inhibitors cobimetinib and pictilisib had limited tolerability and efficacy in solid tumors [125]. Increasing evidence suggests that dual blockade of both pathways has a critical role in tumors with a high frequency of RAS/RAF/MEK/ERK pathway activation and when double blockade is required to overcome drug resistance [126].
More specifically, melanoma and BC frequently exhibited hyperactivation of PI3K and PI3K/AKT/mTOR and MAPK/MEK/ERK pathways [125,127]. Numerous preclinical studies have demonstrated that dual pharmacological inhibition of PI3K and MAPK pathways (via both continuous and intermittent dosing) increased therapeutic efficacy in basal-like BC and melanoma models [128,129]. As a consequence, a phase Ib study has been conducted to test the MEK inhibitors pimasertib combined with voxtalisib in patients with advanced solid tumors, including TNBC and BRAFV600-mutant melanoma, who progressed on BRAF inhibitors [130]. However, the combination showed poor long-term tolerability and limited anti-tumour activity, preventing it from progressing into further testing [130]. Similar dose-limiting toxicities emerged in BRAFV600-mutant advanced melanoma patients treated with the combination of buparlisib and the BRAF inhibitor vemurafenib [131]. Regarding AKT inhibitors, uprosertib in combination with the oral MEK1/MEK2 inhibitor trametinib showed poor tolerability in patients with solid tumors and minimal clinical efficacy [132]. Similarly, the trametinib and afuresertib combination was poorly tolerated in patients with solid tumors and multiple myeloma [117]. It is worth mentioning that both RAS/MAPK and PI3K pathways play a key role in cancer metabolism. PI3K can directly reset cellular metabolism by phosphorylating metabolic enzymes and regulating metabolism-associated proteins such as sterol regulatory element-binding proteins (SREBP), thus enhancing the activities of nutrient transporters indirectly by controlling various transcriptional factors (TFs) [24,133].
Similarly, RAS/MAPK signalling is involved in glucose metabolism in different ways. Mutant KRAS upregulates the hexokinase 1 and 2 (HK1 and HK2) rate-limiting enzymes of glycolysis [134,135] and increases the expression of key glycolytic enzymes such as PFK1, ENO1, and LDHA [136], thus stimulating glycolytic flux and facilitating the synthesis of glycolytic intermediates [137,138,139].
Given the interest in combination treatment of MAPK and PI3K/AKT pathway inhibitors, further investigation may be warranted, especially in patients with coexisting PI3K pathway mutations and KRAS or BRAF mutations.

3.4.3. Chemotherapy

It is well-established that the PI3K pathway synergizes with various chemotherapeutic agents such as doxorubicin, etoposide, topotecan, cisplatin, vincristine and taxol, resulting in increased tumour sensitivity to chemotherapy [140]. Interestingly enough, PI3K inhibition was reported to induce apoptosis and suppress tumor growth in patients’ derived primary neuroblastoma cells and in an in vivo neuroblastoma model [141]. Additionally, preliminary clinical studies demonstrated that PI3Ki in combination with chemotherapy are safe and well tolerated [27]. Pictilisib, carboplatin and paclitaxel have demonstrated promising antitumor activity in patients with NSCLC [142]. In terms of clinical benefit, the addition of ipatasertib to mFOLFOX6 did not improve PFS in a phase 2 randomized trial enrolling metastatic gastric or gastroesophageal junction cancer patients [143]. According to preclinical models, PI3K signalling stabilizes and preserves DNA double-strand break (DSB) repair by interacting with the homologous recombination (HR) complex [144], and is fundamental for DNA repair during ionizing radiation [145]. It is jointly agreed that PI3K inhibition may induce DNA damage and subsequently increase the sensitivity of cell lines to PARPi [146,147]. Given this evidence, Ibrahim et al. investigated the effects of PI3K inhibition in BRCA-proficient TNBC’s preclinical models with PI3K-activating alterations [147]. They observed that PI3K blockade induces (HR) impairment and sensitization to PARP inhibition [147]. Batalini et al. have recently published the results of a phase 1b trial showing that alpelisib, in combination with olaparib, has antitumor activity in patients with pre-treated TNBC [148]. An additional clinical study investigating buparlisib is currently ongoing [149]. Results from the aforementioned trials will provide new insights into the efficacy of this combination, further promoting the use of PI3Ki as an emerging therapeutic strategy in TNBC.

3.4.4. Immunotherapy

The tumor microenvironment (TME) plays an essential role in tumor initiation, growth, invasion, metastasis and cancer treatment [24,150]. Specifically, TME allows cancer cells to become invasive and spread from the primary site to distant locations through a complex and multistep metastatic process [27]. Recently, the PI3K/AKT pathway has been shown to exert a pivotal role in regulating anti-tumor immunity by promoting an immunosuppressive TME and controlling the activity of the tumor infiltration cells associated [151]. Multiple studies have demonstrated how programmed death ligand-1 (PD-L1) and cytotoxic T lymphocyte-associated protein 4 (CTLA-4) interact with PI3K signalling. For instance, PI3K inhibition led to a reduction of tumor PD-L1 expression in PTEN-mutant TNBC and colorectal cancer (CRC) [152]. More specifically, the PI3Kα-specific or pan-PI3K inhibitor did not show an anti-tumor response over ICI alone in TNBC models, while the PI3K/mTOR dual inhibitor gedatolisib associated with ICIs induced a substantial cancer growth inhibition and a greater activation and response of T-cells, natural killer (NK)-cell, and dendritic cells (DC) [153].
On the other hand, in PTEN loss melanoma, preclinical and clinical studies have provided strong evidence that PI3Kβ inhibition, in combination with anti-CTLA-4 agent, improved the efficacy of immunotherapy [154,155]. Likewise, Lastwika et al. have demonstrated that in human lung adenocarcinomas and squamous cell carcinomas, PD-L1 expression was significantly correlated with mTOR activation [156]. Their findings were corroborated by studies using genetically engineered mouse models of lung cancer where an mTOR inhibitor, combined with a PD-1 antibody, reduced tumor growth, increased tumor-infiltrating T cells and diminished regulatory T cells [156].
At ASCO 2021, Schmid reported the preliminary results of the phase1b BEGONIA trial, evaluating the safety and efficacy of capivasertib with paclitaxel and durvalumab as a first-line treatment for PD-L1+ metastatic TNBC [124]. The addition of capivasertib resulted in an ORR similar to the paclitaxel/durvalumab arm, although the limited number of patients enrolled in the study does not allow robust conclusions to be drawn. Furthermore, the addition of capivasertib to durvalumab and paclitaxel regimen induced a relatively high rate of G3/4 treatment-related adverse events [121]. A phase 2 trial (NCT03961698) has investigated the triplet combination of eganelisib (PI3K-γ inhibitor) with atezolizumab and nab-paclitaxel as first-line therapy for locally advanced or metastatic TNBC patients and renal cell carcinoma (MARIO-3 trial) [122]. At the last update, this combination provided manageable toxicity and evidence of a long-term PFS benefit, in TNBC, with an ORR of 55.3% irrespective of PD-L1 expression [123]. Regarding ongoing clinical trials, a phase I/II trial (NCT03131908) is investigating the selective PI3K-beta inhibitor GSK2636771 in combination with pembrolizumab in patients with refractory metastatic PTEN-loss melanoma. Additionally, a phase I/II trial (NCT04317105) is evaluating copanlisib with nivolumab and ipilimumab in PI3K/AKT-mutated solid tumors. Another phase 2 trial (NCT03190174) is investigating the biological activity of the sequential administration of nivolumab and the mTOR inhibitor ABI-009 in multiple types of cancer. Another phase 1 trial (NCT03772561) is exploring capivasertib combined with durvalumab and olaparib in patients with advanced or metastatic solid tumors. Finally, a phase Ib study investigating the anti-PD1 antibody spartalizumab plus everolimus in TNBC patients (NCT02890069) has recently closed, although the outcome is not yet available. Recently, preclinical data have shown that PI3Ki may increase the efficacy of chimeric antigen receptor T cells (CAR-T) in vivo; however, these results are preliminary and further investigation is required to elucidate the underlying mechanism [157,158]. Several combination approaches with the PI3K/AKT/mTOR inhibitor and ICI are ongoing (Table 4).

4. Impact of Biomarkers

As mentioned above, differents oncogenic genomic alterations are responsible for PI3K/AKT/mTOR pathway hyperactivation. Somatic point mutations and gene amplifications are the two principal alterations promoting the PIK3CA functions [159]. PIK3CA gene status can vary among primary tumor and metastases [160]. This potential discordance can interest the gain or loss of the PIK3CA gene mutations or different levels of mutation [161,162]. All these aspects underline the importance of molecular characterization of metastatic sites on the activity of PI3Kis. The use of circulating tumor DNA (ctDNA) is an alternative when the biopsy of a metastatic site is difficult or cannot be obtained, selecting patients with adequate tumor burden or with disease progression to increase the probability of adequate ctDNA [163,164]. Beyond PIK3CA mutations, additional biomarkers have been evaluated as potential biomarkers of resistance to PI3Ki [165]. For instance, preclinical data showed that tumors characterized by PTEN expression loss are more sensitive to AKT/PI3K inhibitors and more dependent on PI3Kβ signalling, thereby benefiting from pictilisib [166,167]. On the other hand, the increase of insulin level induced by PI3Kis can lead to the re-activation of PI3K/AKT in murine tumour models [168], while PI3K inhibition may upregulate ER-dependent transcription due to the epigenomic crosstalk between PI3K and ER pathways [169].
Indeed, in a phase 1b trial of alpelisib and letrozole, the small subgroup of patients harbouring FGFR1/2 amplification, KRAS or TP53 mutations did not show any benefit [169]. It is important to underline that only PIK3CA mutation is currently approved as a predictive biomarker in clinical practice. Moreover, PIK3CA mutations have been recently included in the tier IA of genomic alterations in BC, of the ESMO Scale for Clinical Actionability of molecular Targets (ESCAT), as predictors of benefit from a-selective PI3Kis.
Ultimately, the PI3K pathway is involved in the differentiation of MDSCs and Tregs within TME, and suppressive cytokines can impair stimulation of macrophages and dendritic cells, suggesting a potential synergy for combining PI3Kis and immunotherapy [7,8]. The use of programmed PD-L1 and tumor-infiltrating lymphocytes (TILs) as predictive and prognostic biomarkers is well-recognized and associated with response to immunotherapy, although the latter biomarker is related to pathological specimens often found in the primary tumor site [170,171]. Several biomarkers of immunological state such as circulating tumor cells, circulating immunity cells and inflammatory indexes have been investigated as predictive and prognostic biomarkers [172,173,174,175].
Looking to the future, trials are being developed using baseline, on-treatment and post-treatment PI3Kis and immunotherapy which may improve our understanding of the complex interaction between host immunity and PI3KCA, ultimately improving our approach to patients.

5. Summary and Conclusions

Dysregulation of PI3K/AKT/mTOR signalling is frequently observed in human cancer and it is responsible for tumorigenesis, cancer progression, as well as intrinsic and acquired resistance to several treatments. This pathway is an attractive molecular target for therapeutic interventions and the development of novel anti-cancer molecules. The last two decades have seen exponential growth in the number of PI3K inhibitors investigated in pre-clinical studies, with approximately fifteen compounds that have progressed into clinical trials as new anticancer drugs. However, the high toxicity and the lack of selectivity have hampered the application and approval of PI3Ki in clinical practice. Clinical adverse events associated with these PI3K/AKT/mTORi such as hyperglycemia, pneumonitis, stomatitis, rashes and diarrhoea have a crucial impact on a patient’s quality of life leading to a high percentage of treatment discontinuation.
In order to optimize the efficacy of PI3Ki and limit toxicity, better management of side effects as well as well-designed studies for the identification and validation of actionable predictive biomarkers associated with the clinical activity are required.
Furthermore, additional investigations to establish the role of the PI3K pathway on the tumor microenvironment and a better patient selection and stratification will be crucial to pave the way for combination treatments with immunotherapies.
Due to their potential synergistic action, drug combination strategies with chemo or target therapies as well as novel dosing schedules may enhance the clinical benefit and potentially overcome intrinsic and acquired resistance with fewer AEs. In order to maximise combination therapies, high-throughput molecular profiling approaches will be essential to promote an accurate matching of patients with PIK3CA aberrations to specific tumor subtypes.

Author Contributions

Conceptualization, critical review of literature, writing—original draft preparation, writing—review and editing, validation: M.S., A.D., C.G., C.C., F.M., U.D.G. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Acknowledgments

This work was partly supported by “Ricerca Corrente”, by the Italian Ministry of Health within the research line "Precision, gender and ethnicity-based medicine and geroscience: genetic-molecular mechanisms in the development, characterization”.

Conflicts of Interest

Ugo De Giorgi received honoraria for advisory boards or speaker fees for Pfizer, BMS, MSD, PharmaMar, Astellas, Bayer, Ipsen, Roche, Novartis, Clovis, GSK, AstraZeneca, institutional research grants from AstraZeneca, Sanofi and Roche.

References

  1. Lawrence, M.S.; Stojanov, P.; Mermel, C.H.; Robinson, J.T.; Garraway, L.A.; Golub, T.R.; Meyerson, M.; Gabriel, S.B.; Lander, E.S.; Getz, G. Discovery and Saturation Analysis of Cancer Genes across 21 Tumour Types. Nature 2014, 505, 495–501. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Hennessy, B.T.; Smith, D.L.; Ram, P.T.; Lu, Y.; Mills, G.B. Exploiting the PI3K/AKT Pathway for Cancer Drug Discovery. Nat. Rev. Drug Discov. 2005, 4, 988–1004. [Google Scholar] [CrossRef] [PubMed]
  3. Vanhaesebroeck, B.; Guillermet-Guibert, J.; Graupera, M.; Bilanges, B. The Emerging Mechanisms of Isoform-Specific PI3K Signalling. Nat. Rev. Mol. Cell Biol. 2010, 11, 329–341. [Google Scholar] [CrossRef] [PubMed]
  4. Thorpe, L.M.; Yuzugullu, H.; Zhao, J.J. PI3K in Cancer: Divergent Roles of Isoforms, Modes of Activation and Therapeutic Targeting. Nat. Rev. Cancer 2015, 15, 7–24. [Google Scholar] [CrossRef] [Green Version]
  5. Engelman, J.A. Targeting PI3K Signalling in Cancer: Opportunities, Challenges and Limitations. Nat. Rev. Cancer 2009, 9, 550–562. [Google Scholar] [CrossRef]
  6. Hanker, A.B.; Kaklamani, V.; Arteaga, C.L. Challenges for the Clinical Development of PI3K Inhibitors: Strategies to Improve Their Impact in Solid Tumors. Cancer Discov. 2019, 9, 482–491. [Google Scholar] [CrossRef] [Green Version]
  7. Okkenhaug, K.; Graupera, M.; Vanhaesebroeck, B. Targeting PI3K in Cancer: Impact on Tumor Cells, Their Protective Stroma, Angiogenesis, and Immunotherapy. Cancer Discov. 2016, 6, 1090–1105. [Google Scholar] [CrossRef] [Green Version]
  8. O’Donnell, J.S.; Massi, D.; Teng, M.W.; Mandala, M. PI3K-AKT-mTOR Inhibition in Cancer Immunotherapy, Redux. Semin. Cancer Biol. 2018, 48, 91–103. [Google Scholar] [CrossRef] [Green Version]
  9. Ocaña, A.; Vera-Badillo, F.; Al-Mubarak, M.; Templeton, A.J.; Corrales-Sánchez, V.; Díez-González, L.; Cuenca-Lopez, M.D.; Seruga, B.; Pandiella, A.; Amir, E. Activation of the PI3K/mTOR/AKT Pathway and Survival in Solid Tumors: Systematic Review and Meta-Analysis. PLoS ONE 2014, 9, e95219. [Google Scholar] [CrossRef]
  10. Du Rusquec, P.; Blonz, C.; Frenel, J.S.; Campone, M. Targeting the PI3K/Akt/mTOR Pathway in Estrogen-Receptor Positive HER2 Negative Advanced Breast Cancer. Ther. Adv. Med. Oncol. 2020, 12, 1758835920940939. [Google Scholar] [CrossRef]
  11. Li, H.; Zeng, J.; Shen, K. PI3K/AKT/mTOR Signaling Pathway as a Therapeutic Target for Ovarian Cancer. Arch. Gynecol. Obstet. 2014, 290, 1067–1078. [Google Scholar] [CrossRef]
  12. Dobashi, Y.; Watanabe, Y.; Miwa, C.; Suzuki, S.; Koyama, S. Mammalian Target of Rapamycin: A Central Node of Complex Signaling Cascades. Int. J. Clin. Exp. Pathol. 2011, 4, 476–495. [Google Scholar]
  13. Janku, F.; Yap, T.A.; Meric-Bernstam, F. Targeting the PI3K Pathway in Cancer: Are We Making Headway? Nat. Rev. Clin. Oncol. 2018, 15, 273–291. [Google Scholar] [CrossRef]
  14. Saal, L.; Gruvberger-Saal, S.K.; Persson, C.; Lövgren, K.; Jumppanen, M.; Staaf, J.; Jönsson, G.; Pires, M.M.; Maurer, M.; Holm, K.; et al. Recurrent Gross Mutations of the PTEN Tumor Suppressor Gene in Breast Cancers with Deficient DSB Repair. Nat. Genet. 2008, 40, 102–107. [Google Scholar] [CrossRef]
  15. Stemke-Hale, K.; Gonzalez-Angulo, A.M.; Lluch, A.; Neve, R.M.; Kuo, W.-L.; Davies, M.; Carey, M.; Hu, Z.; Guan, Y.; Sahin, A.; et al. An Integrative Genomic and Proteomic Analysis of PIK3CA, PTEN, and AKT Mutations in Breast Cancer. Cancer Res. 2008, 68, 6084–6091. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Fedele, C.G.; Ooms, L.M.; Ho, M.; Vieusseux, J.; O’Toole, S.A.; Millar, E.K.; Lopez-Knowles, E.; Sriratana, A.; Gurung, R.; Baglietto, L.; et al. Inositol Polyphosphate 4-Phosphatase II Regulates PI3K/Akt Signaling and is Lost in Human Basal-Like Breast Cancers. Proc. Natl. Acad. Sci. USA 2010, 107, 22231–22236. [Google Scholar] [CrossRef] [Green Version]
  17. Cheung, L.W.; Mills, G.B. Targeting Therapeutic Liabilities Engendered by PIK3R1 Mutations for Cancer Treatment. Pharmacogenomics 2016, 17, 297–307. [Google Scholar] [CrossRef] [Green Version]
  18. Jean, S.; Kiger, A.A. Classes of Phosphoinositide 3-Kinases at a Glance. J. Cell Sci. 2014, 127, 923–928. [Google Scholar] [CrossRef] [Green Version]
  19. Toker, A.; Cantley, L.C. Signalling through the Lipid Products of Phosphoinositide-3-OH Kinase. Nature 1997, 387, 673–676. [Google Scholar] [CrossRef]
  20. Folkes, A.J.; Ahmadi, K.; Alderton, W.K.; Alix, S.; Baker, S.J.; Box, G.; Chuckowree, I.S.; Clarke, P.A.; Depledge, P.; Eccles, S.A.; et al. The Identification of 2-(1H-Indazol-4-Yl)-6-(4-Methanesulfonyl-Piperazin-1-Ylmethyl)-4-Morpholin-4-Yl-Thieno[3,2-d]Pyrimidine (GDC-0941) as a Potent, Selective, Orally Bioavailable Inhibitor of Class I PI3 Kinase for the Treatment of Cancer. J. Med. Chem. 2008, 51, 5522–5532. [Google Scholar] [CrossRef]
  21. 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]
  22. Kriplani, N.; Hermida, M.A.; Brown, E.R.; Leslie, N.R. Class I PI 3-Kinases: Function and Evolution. Adv. Biol. Regul. 2015, 59, 53–64. [Google Scholar] [CrossRef] [PubMed]
  23. Gerstung, M.; PCAWG Evolution & Heterogeneity Working Group; Jolly, C.; Leshchiner, I.; Dentro, S.C.; Gonzalez, S.; Rosebrock, D.; Mitchell, T.J.; Rubanova, Y.; Anur, P.; et al. The Evolutionary History of 2658 Cancers. Nature 2020, 578, 122–128. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. He, Y.; Sun, M.M.; Zhang, G.G.; Yang, J.; Chen, K.S.; Xu, W.W.; Li, B.B. Targeting PI3K/Akt Signal Transduction for Cancer Therapy. Signal Transduct. Target. Ther. 2021, 6, 425. [Google Scholar] [CrossRef]
  25. Herman, S.E.M.; Gordon, A.L.; Wagner, A.J.; Heerema, N.A.; Zhao, W.; Flynn, J.M.; Jones, J.; Andritsos, L.; Puri, K.D.; Lannutti, B.J.; et al. Phosphatidylinositol 3-Kinase-δ Inhibitor CAL-101 Shows Promising Preclinical Activity in Chronic Lymphocytic Leukemia by Antagonizing Intrinsic and Extrinsic Cellular Survival Signals. Blood 2010, 116, 2078–2088. [Google Scholar] [CrossRef] [Green Version]
  26. Tsolakos, N.; Durrant, T.N.; Chessa, T.; Suire, S.M.; Oxley, D.; Kulkarni, S.; Downward, J.; Perisic, O.; Williams, R.L.; Stephens, L.; et al. Quantitation of Class IA PI3Ks in Mice Reveals P110-Free-P85s and Isoform-Selective Subunit Associations and Recruitment to Receptors. Proc. Natl. Acad. Sci. USA 2018, 115, 12176–12181. [Google Scholar] [CrossRef] [Green Version]
  27. Mishra, R.; Patel, H.; Alanazi, S.; Kilroy, M.K.; Garrett, J.T. PI3K Inhibitors in Cancer: Clinical Implications and Adverse Effects. Int. J. Mol. Sci. 2021, 22, 3464. [Google Scholar] [CrossRef]
  28. Braccini, L.; Ciraolo, E.; Campa, C.C.; Perino, A.; Longo, D.L.; Tibolla, G.; Pregnolato, M.; Cao, Y.; Tassone, B.; Damilano, F.; et al. PI3K-C2γ Is a Rab5 Effector Selectively Controlling Endosomal Akt2 Activation Downstream of Insulin Signalling. Nat. Commun. 2015, 6, 7400. [Google Scholar] [CrossRef] [Green Version]
  29. Gulluni, F.; Martini, M.; De Santis, M.C.; Campa, C.C.; Ghigo, A.; Margaria, J.P.; Ciraolo, E.; Franco, I.; Ala, U.; Annaratone, L.; et al. Mitotic Spindle Assembly and Genomic Stability in Breast Cancer Require PI3K-C2α Scaffolding Function. Cancer Cell 2017, 32, 444–459.e7. [Google Scholar] [CrossRef] [Green Version]
  30. Gulluni, F.; De Santis, M.C.; Margaria, J.P.; Martini, M.; Hirsch, E. Class II PI3K Functions in Cell Biology and Disease. Trends Cell Biol. 2019, 29, 339–359. [Google Scholar] [CrossRef]
  31. Marat, A.L.; Haucke, V. Phosphatidylinositol 3-Phosphates—At the Interface between Cell Signalling and Membrane Traffic. EMBO J. 2016, 35, 561–579. [Google Scholar] [CrossRef]
  32. O’Farrell, F.; Lobert, V.H.; Sneeggen, M.; Jain, A.; Katheder, N.; Wenzel, E.M.; Schultz, S.W.; Tan, K.W.; Brech, A.; Stenmark, H.; et al. Class III Phosphatidylinositol-3-OH Kinase Controls Epithelial Integrity through Endosomal LKB1 Regulation. Nat. Cell Biol. 2017, 19, 1412–1423. [Google Scholar] [CrossRef] [Green Version]
  33. Stjepanovic, G.; Baskaran, S.; Lin, M.G.; Hurley, J.H. Vps34 Kinase Domain Dynamics Regulate the Autophagic PI 3-Kinase Complex. Mol. Cell 2017, 67, 528–534.e3. [Google Scholar] [CrossRef]
  34. Staal, S.P. Molecular Cloning of the Akt Oncogene and Its Human Homologues AKT1 and AKT2: Amplification of AKT1 in a Primary Human Gastric Adenocarcinoma. Proc. Natl. Acad. Sci. USA 1987, 84, 5034–5037. [Google Scholar] [CrossRef] [Green Version]
  35. Mayer, I.A.; Arteaga, C.L. The PI3K/AKT Pathway as a Target for Cancer Treatment. Annu. Rev. Med. 2016, 67, 11–28. [Google Scholar] [CrossRef]
  36. Jiang, N.; Dai, Q.; Su, X.; Fu, J.; Feng, X.; Peng, J. Role of PI3K/AKT Pathway in Cancer: The Framework of Malignant Behavior. Mol. Biol. Rep. 2020, 47, 4587–4629. [Google Scholar] [CrossRef] [Green Version]
  37. Cisse, O.; Quraishi, M.; Gulluni, F.; Guffanti, F.; Mavrommati, I.; Suthanthirakumaran, M.; Oh, L.C.R.; Schlatter, J.N.; Sarvananthan, A.; Broggini, M.; et al. Downregulation of Class II Phosphoinositide 3-Kinase PI3K-C2β Delays Cell Division and Potentiates the Effect of Docetaxel on Cancer Cell Growth. J. Exp. Clin. Cancer Res. 2019, 38, 472. [Google Scholar] [CrossRef]
  38. Wang, B.; Zhang, W.; Zhang, G.; Kwong, L.; Lu, H.; Tan, J.; Sadek, N.; Xiao, M.; Zhang, J.; Labrie, M.; et al. Targeting mTOR signaling overcomes acquired resistance to combined BRAF and MEK inhibition in BRAF-mutant melanoma. Oncogene 2021, 40, 5590–5599. [Google Scholar] [CrossRef]
  39. Shi, W.; Zhang, X.; Pintilie, M.; Ma, N.; Miller, N.; Banerjee, D.; Tsao, M.-S.; Mak, T.; Fyles, A.; Liu, F.-F. Dysregulated PTEN-PKB and Negative Receptor Status in Human Breast Cancer. Int. J. Cancer 2003, 104, 195–203. [Google Scholar] [CrossRef]
  40. Nunnery, S.; Mayer, I. Management of Toxicity to Isoform α-Specific PI3K Inhibitors. Ann. Oncol. Off. J. Eur. Soc. Med. Oncol. 2019, 30, x21–x26. [Google Scholar] [CrossRef]
  41. Papadimitrakopoulou, V. Development of PI3K/AKT/mTOR Pathway Inhibitors and Their Application in Personalized Therapy for Non-Small-Cell Lung Cancer. J. Thorac. Oncol. Off. Publ. Int. Assoc. Study Lung Cancer 2012, 7, 1315–1326. [Google Scholar] [CrossRef] [PubMed]
  42. Agoulnik, I.U.; Hodgson, M.C.; Bowden, W.A.; Ittmann, M.M. INPP4B: The New Kid on the PI3K Block. Oncotarget 2011, 2, 321–328. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Xie, Y.; Naizabekov, S.; Chen, Z.; Tokay, T. Power of PTEN/AKT: Molecular Switch between Tumor Suppressors and Oncogenes. Oncol. Lett. 2016, 12, 375–378. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Ngeow, J.; Eng, C. PTEN in Hereditary and Sporadic Cancer. Cold Spring Harb. Perspect. Med. 2019, 10, a036087. [Google Scholar] [CrossRef] [Green Version]
  45. Saxton, R.A.; Sabatini, D.M. mTOR Signaling in Growth, Metabolism, and Disease. Cell 2017, 168, 960–976. [Google Scholar] [CrossRef] [Green Version]
  46. Murugan, A.K. mTOR: Role in Cancer, Metastasis and Drug Resistance. Semin. Cancer Biol. 2019, 59, 92–111. [Google Scholar] [CrossRef]
  47. Hua, H.; Kong, Q.; Zhang, H.; Wang, J.; Luo, T.; Jiang, Y. Targeting mTOR for Cancer Therapy. J. Hematol. Oncol. 2019, 12, 71. [Google Scholar] [CrossRef]
  48. Krencz, I.; Sebestyen, A.; Khoor, A. mTOR in Lung Neoplasms. Pathol. Oncol. Res. 2020, 26, 35–48. [Google Scholar] [CrossRef]
  49. Niederst, M.J.; Engelman, J.A. Bypass Mechanisms of Resistance to Receptor Tyrosine Kinase Inhibition in Lung Cancer. Sci. Signal. 2013, 6, re6. [Google Scholar] [CrossRef] [Green Version]
  50. Toulany, M.; Minjgee, M.; Saki, M.; Holler, M.; Meier, F.; Eicheler, W.; Rodemann, H.P. ERK2-Dependent Reactivation of Akt Mediates the Limited Response of Tumor Cells with Constitutive K-RAS Activity to PI3K Inhibition. Cancer Biol. Ther. 2014, 15, 317–328. [Google Scholar] [CrossRef] [Green Version]
  51. Fekete, M.; Santiskulvong, C.; Eng, C.; Dorigo, O. Effect of PI3K/Akt Pathway Inhibition-Mediated G1 Arrest on Chemosensitization in Ovarian Cancer Cells. Anticancer. Res. 2012, 32, 445–452. [Google Scholar]
  52. Carden, C.P.; Stewart, A.; Thavasu, P.; Kipps, E.; Pope, L.; Crespo, M.; Miranda, S.; Attard, G.; Garrett, M.D.; Clarke, P.A.; et al. The Association of PI3 Kinase Signaling and Chemoresistance in Advanced Ovarian Cancer. Mol. Cancer Ther. 2012, 11, 1609–1617. [Google Scholar] [CrossRef]
  53. Garces, A.E.; Stocks, M.J. Class 1 PI3K Clinical Candidates and Recent Inhibitor Design Strategies: A Medicinal Chemistry Perspective. J. Med. Chem. 2019, 62, 4815–4850. [Google Scholar] [CrossRef]
  54. Yang, J.; Nie, J.; Ma, X.; Wei, Y.; Peng, Y.; Wei, X. Targeting PI3K in cancer: Mechanisms and Advances in Clinical Trials. Mol. Cancer 2019, 18, 26. [Google Scholar] [CrossRef] [Green Version]
  55. Akinleye, A.; Avvaru, P.; Furqan, M.; Song, Y.; Liu, D. Phosphatidylinositol 3-Kinase (PI3K) Inhibitors as Cancer Therapeutics. J. Hematol. Oncol. 2013, 6, 88. [Google Scholar] [CrossRef] [Green Version]
  56. Junttila, T.T.; Akita, R.W.; Parsons, K.; Fields, C.; Lewis Phillips, G.D.; Friedman, L.S.; Sampath, D.; Sliwkowski, M.X. Ligand-Independent HER2/HER3/PI3K Complex Is Disrupted by Trastuzumab and Is Effectively Inhibited by the PI3K Inhibitor GDC-0941. Cancer Cell 2009, 15, 429–440. [Google Scholar] [CrossRef] [Green Version]
  57. Sarker, D.; Ang, J.E.; Baird, R.; Kristeleit, R.; Shah, K.; Moreno, V.; Clarke, P.A.; Raynaud, F.I.; Levy, G.; Ware, J.A.; et al. First-in-Human Phase I Study of Pictilisib (GDC-0941), a Potent Pan-Class I Phosphatidylinositol-3-Kinase (PI3K) Inhibitor, in Patients with Advanced Solid Tumors. Clin. Cancer Res. Off. J. Am. Assoc. Cancer Res. 2015, 21, 77–86. [Google Scholar] [CrossRef] [Green Version]
  58. Mayer, I.A.; Abramson, V.G.; Isakoff, S.J.; Forero, A.; Balko, J.M.; Kuba, M.G.; Sanders, M.E.; Yap, J.; Van den Abbeele, A.D.; Li, Y.; et al. Stand up to Cancer Phase Ib Study of Pan-Phosphoinositide-3-Kinase Inhibitor Buparlisib with Letrozole in Estrogen Receptor-Positive/Human Epidermal Growth Factor Receptor 2-Negative Metastatic Breast Cancer. J. Clin. Oncol. Off. J. Am. Soc. Clin. Oncol. 2014, 32, 1202–1209. [Google Scholar] [CrossRef]
  59. Speranza, M.C.; Nowicki, M.O.; Behera, P.; Cho, C.-F.; Chiocca, E.A.; Lawler, S.E. BKM-120 (Buparlisib): A Phosphatidyl-Inositol-3 Kinase Inhibitor with Anti-Invasive Properties in Glioblastoma. Sci. Rep. 2016, 6, 20189. [Google Scholar] [CrossRef] [Green Version]
  60. Bendell, J.C.; Rodon, J.; Burris, H.A.; De Jonge, M.; Verweij, J.; Birle, D.; Demanse, D.; De Buck, S.S.; Ru, Q.C.; Peters, M.; et al. Phase I, Dose-Escalation Study of BKM120, an Oral Pan-Class I PI3K Inhibitor, in Patients with Advanced Solid Tumors. J. Clin. Oncol. Off. J. Am. Soc. Clin. Oncol. 2012, 30, 282–290. [Google Scholar] [CrossRef]
  61. Garrido-Castro, A.C.; Saura, C.; Barroso-Sousa, R.; Guo, H.; Ciruelos, E.; Bermejo, B.; Gavilá, J.; Serra, V.; Prat, A.; Paré, L.; et al. Phase 2 Study of Buparlisib (BKM120), a Pan-Class I PI3K Inhibitor, in Patients with Metastatic Triple-Negative Breast Cancer. Breast Cancer Res. 2020, 22, 120. [Google Scholar] [CrossRef] [PubMed]
  62. Miller, T.W.; Hennessy, B.T.; González-Angulo, A.M.; Fox, E.M.; Mills, G.B.; Chen, H.; Higham, C.; García-Echeverría, C.; Shyr, Y.; Arteaga, C.L. Hyperactivation of Phosphatidylinositol-3 Kinase Promotes Escape from Hormone Dependence in Estrogen Receptor—Positive Human Breast Cancer. J. Clin. Investig. 2010, 120, 2406–2413. [Google Scholar] [CrossRef] [PubMed]
  63. Miller, T.W.; Rexer, B.N.; Garrett, J.T.; Arteaga, C.L. Mutations in the Phosphatidylinositol 3-Kinase Pathway: Role in Tumor Progression and Therapeutic Implications in Breast Cancer. Breast Cancer Res. BCR 2011, 13, 224. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Di Leo, A.; Johnston, S.; Lee, K.S.; Ciruelos, E.; Lønning, P.E.; Janni, W.; O’Regan, R.; Mouret-Reynier, M.-A.; Kalev, D.; Egle, D.; et al. Buparlisib plus Fulvestrant in Postmenopausal Women with Hormone-Receptor-Positive, HER2-Negative, Advanced Breast Cancer Progressing on or After mTOR Inhibition (BELLE-3): A randomised, Double-Blind, Placebo-Controlled, Phase 3 Trial. Lancet Oncol. 2018, 19, 87–100. [Google Scholar] [CrossRef] [PubMed]
  65. Martín, M.; Chan, A.; Dirix, L.; O’Shaughnessy, J.; Hegg, R.; Manikhas, A.; Shtivelband, M.; Krivorotko, P.; Batista López, N.; Campone, M.; et al. A Randomized Adaptive Phase II/III Study of Buparlisib, a Pan-Class I PI3K Inhibitor, Combined with Paclitaxel for the Treatment of HER2- Advanced Breast Cancer (BELLE-4). Ann. Oncol. 2017, 28, 313–320. [Google Scholar] [CrossRef]
  66. Liu, N.; Rowley, B.R.; Bull, C.O.; Schneider, C.; Haegebarth, A.; Schatz, C.A.; Fracasso, P.R.; Wilkie, D.P.; Hentemann, M.; Wilhelm, S.M.; et al. BAY 80-6946 Is a Highly Selective Intravenous PI3K Inhibitor with Potent P110α and P110δ Activities in Tumor Cell Lines and Xenograft Models. Mol. Cancer Ther. 2013, 12, 2319–2330. [Google Scholar] [CrossRef] [Green Version]
  67. Commissioner, O. DA Approves New Treatment for Adults with Relapsed Follicular Lymphoma. Available online: https://www.fda.gov/news-events/press-announcements/fda-approves-new-treatment-adults-relapsed-follicular-lymphoma (accessed on 14 September 2017).
  68. National Cancer Institute (NCI). Phase Ib/II Trial of Copanlisib in Combination with Trastuzumab and Pertuzumab After Induction Treatment of HER2 Positive (HER2+) Metastatic Breast Cancer (MBC) with PIK3CA Mutation or PTEN Mutation. Available online: https://clinicaltrials.gov/ct2/show/NCT04108858 (accessed on 25 January 2022).
  69. Chauhan, A.F.; Cheson, B.D. Copanlisib in the Treatment of Relapsed Follicular Lymphoma: Utility and Experience from the Clinic. Cancer Manag. Res. 2021, 13, 677–692. [Google Scholar] [CrossRef]
  70. Ellis, H.; Ma, C.X. PI3K Inhibitors in Breast Cancer Therapy. Curr. Oncol. Rep. 2019, 21, 110. [Google Scholar] [CrossRef]
  71. FDA Approves Alpelisib for Metastatic Breast Cancer. FDA 2019. Available online: https://www.fda.gov/drugs/resources-information-approved-drugs/fda-approves-alpelisib-metastatic-breast-cancer (accessed on 24 May 2019).
  72. Li, H.; Prever, L.; Hirsch, E.; Gulluni, F. Targeting PI3K/AKT/mTOR Signaling Pathway in Breast Cancer. Cancers 2021, 13, 3517. [Google Scholar] [CrossRef]
  73. Mavratzas, A.; Marmé, F. Alpelisib in the Treatment of Metastatic HR+ Breast Cancer with PIK3CA Mutations. Future Oncol. 2021, 17, 13–36. [Google Scholar] [CrossRef]
  74. Juric, D.; Janku, F.; Rodón, J.; Burris, H.A.; Mayer, I.A.; Schuler, M.; Seggewiss-Bernhardt, R.; Gil-Martin, M.; Middleton, M.R.; Baselga, J.; et al. Alpelisib plus Fulvestrant in PIK3CA-Altered and PIK3CA-Wild-Type Estrogen Receptor—Positive Advanced Breast Cancer: A Phase 1b Clinical Trial. JAMA Oncol. 2019, 5, e184475. [Google Scholar] [CrossRef] [Green Version]
  75. André, F.; Ciruelos, E.; Rubovszky, G.; Campone, M.; Loibl, S.; Rugo, H.S.; Iwata, H.; Conte, P.; Mayer, I.A.; Kaufman, B.; et al. Alpelisib for PIK3CA-Mutated, Hormone Receptor—Positive Advanced Breast Cancer. N. Engl. J. Med. 2019, 380, 1929–1940. [Google Scholar] [CrossRef]
  76. Zumsteg, Z.S.; Morse, N.; Krigsfeld, G.; Gupta, G.; Higginson, D.S.; Lee, N.Y.; Morris, L.; Ganly, I.; Shiao, S.L.; Powell, S.N.; et al. Taselisib (GDC-0032), a Potent β-Sparing Small Molecule Inhibitor of PI3K, Radiosensitizes Head and Neck Squamous Carcinomas Containing Activating PIK3CA Alterations. Clin. Cancer Res. 2016, 22, 2009–2019. [Google Scholar] [CrossRef] [Green Version]
  77. Ndubaku, C.O.; Heffron, T.P.; Staben, S.T.; Baumgardner, M.; Blaquiere, N.; Bradley, E.; Bull, R.; Do, S.; Dotson, J.; Dudley, D.; et al. Discovery of 2-{3-[2-(1-Isopropyl-3-Methyl-1H-1,2–4-Triazol-5-Yl)-5,6-Dihydrobenzo[f]Imidazo[1,2-d][1,4]Oxazepin-9-Yl]-1H-Pyrazol-1-Yl}-2-Methylpropanamide (GDC-0032): A β-Sparing Phosphoinositide 3-Kinase Inhibitor with High Unbound Exposure and Robust in vivo Antitumor Activity. J. Med. Chem. 2013, 56, 4597–4610. [Google Scholar] [CrossRef]
  78. Baselga, J.; Dent, S.F.; Cortés, J.; Im, Y.-H.; Diéras, V.; Harbeck, N.; Krop, I.E.; Verma, S.; Wilson, T.R.; Jin, H.; et al. Phase III Study of Taselisib (GDC-0032) + Fulvestrant (FULV) v FULV in Patients (pts) with Estrogen Receptor (ER)-Positive, PIK3CA-Mutant (MUT), Locally Advanced or Metastatic Breast Cancer (MBC): Primary Analysis from SANDPIPER. J. Clin. Oncol. 2018, 36, LBA1006. [Google Scholar] [CrossRef]
  79. Ediriweera, M.K.; Tennekoon, K.H.; Samarakoon, S.R. Role of the PI3K/AKT/mTOR Signaling Pathway in Ovarian Cancer: Biological and Therapeutic Significance. Semin. Cancer Biol. 2019, 59, 147–160. [Google Scholar] [CrossRef]
  80. Lara, P.N.; Longmate, J.; Mack, P.C.; Kelly, K.; Socinski, M.A.; Salgia, R.; Gitlitz, B.; Li, T.; Koczywas, M.; Reckamp, K.L.; et al. Phase II Study of the AKT Inhibitor MK-2206 plus Erlotinib in Patients with Advanced Non-Small Cell Lung Cancer Who Previously Progressed on Erlotinib. Clin. Cancer Res. Off. J. Am. Assoc. Cancer Res. 2015, 21, 4321–4326. [Google Scholar] [CrossRef] [Green Version]
  81. Konopleva, M.Y.; Walter, R.B.; Faderl, S.H.; Jabbour, E.J.; Zeng, Z.; Borthakur, G.; Huang, X.; Kadia, T.M.; Ruvolo, P.P.; Feliu, J.B.; et al. Preclinical and Early Clinical Evaluation of the Oral AKT Inhibitor, MK-2206, for the Treatment of Acute Myelogenous Leukemia. Clin. Cancer Res. Off. J. Am. Assoc. Cancer Res. 2014, 20, 2226–2235. [Google Scholar] [CrossRef] [Green Version]
  82. Do, K.; Speranza, G.; Bishop, R.; Khin, S.; Rubinstein, L.; Kinders, R.J.; Datiles, M.I.B.; Eugeni, M.; Lam, M.H.; Doyle, L.A.; et al. Biomarker-Driven Phase 2 Study of MK-2206 and Selumetinib (AZD6244, ARRY-142886) in Patients with Colorectal Cancer. Investig. New Drugs 2015, 33, 720–728. [Google Scholar] [CrossRef]
  83. Ma, B.B.Y.; Lui, V.W.Y.; Hui, C.W.C.; Lau, C.P.Y.; Wong, C.-H.; Hui, E.P.; Ng, M.H.; Tsao, S.W.; Li, Y.; Chan, A.T.C. Preclinical evaluation of the AKT Inhibitor MK-2206 in Nasopharyngeal Carcinoma Cell Lines. Investig. New Drugs 2013, 31, 567–575. [Google Scholar] [CrossRef]
  84. Wang, Z.; Luo, G.; Qiu, Z. Akt Inhibitor MK-2206 Reduces Pancreatic Cancer Cell Viability and Increases the Efficacy of Gemcitabine. Oncol. Lett. 2020, 19, 1999–2004. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Andrikopoulou, A.; Chatzinikolaou, S.; Panourgias, E.; Kaparelou, M.; Liontos, M.; Dimopoulos, M.-A.; Zagouri, F. The Emerging Role of Capivasertib in Breast Cancer. Breast Off. J. Eur. Soc. Mastology 2022, 63, 157–167. [Google Scholar] [CrossRef] [PubMed]
  86. She, Q.-B.; Halilovic, E.; Ye, Q.; Zhen, W.; Shirasawa, S.; Sasazuki, T.; Solit, D.B.; Rosen, N. 4E-BP1 Is a Key Effector of the Oncogenic Activation of the AKT and ERK Signaling Pathways That Integrates Their Function in Tumors. Cancer Cell 2010, 18, 39–51. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Davies, B.R.; Greenwood, H.; Dudley, P.; Crafter, C.; Yu, D.-H.; Zhang, J.; Li, J.; Gao, B.; Ji, Q.; Maynard, J.; et al. Preclinical Pharmacology of AZD5363, an Inhibitor of AKT: Pharmacodynamics, Antitumor Activity, and Correlation of Monotherapy Activity with Genetic Background. Mol. Cancer Ther. 2012, 11, 873–887. [Google Scholar] [CrossRef] [Green Version]
  88. Shore, N.; Mellado, B.; Shah, S.; Hauke, R.; Costin, D.; Adra, N.; Cullberg, M.; Teruel, C.F.; Morris, T. A Phase I Study of Capivasertib in Combination with Abiraterone Acetate in Patients with Metastatic Castration-Resistant Prostate Cancer. Clin. Genitourin. Cancer 2022, 39, 85. [Google Scholar] [CrossRef]
  89. Li, J.; Davies, B.R.; Han, S.; Zhou, M.; Bai, Y.; Zhang, J.; Xu, Y.; Tang, L.; Wang, H.; Liu, Y.J.; et al. The AKT Inhibitor AZD5363 Is Selectively Active in PI3KCA Mutant Gastric Cancer, and Sensitizes a Patient-Derived Gastric Cancer Xenograft Model with PTEN Loss to Taxotere. J. Transl. Med. 2013, 11, 241. [Google Scholar] [CrossRef] [Green Version]
  90. Wu, X.; Zhang, J.; Zhen, R.; Lv, J.; Zheng, L.; Su, X.; Zhu, G.; Gavine, P.R.; Xu, S.; Lu, S.; et al. Trastuzumab Anti-Tumor Efficacy in Patient-Derived Esophageal Squamous Cell Carcinoma Xenograft (PDECX) Mouse Models. J. Transl. Med. 2012, 10, 180. [Google Scholar] [CrossRef] [Green Version]
  91. Puglisi, M.; Thavasu, P.; Stewart, A.; de Bono, J.; O’Brien, M.; Popat, S.; Bhosle, J.; Banerji, U. AKT Inhibition Synergistically Enhances Growth-Inhibitory Effects of Gefitinib and Increases Apoptosis in Non-Small Cell Lung Cancer Cell Lines. Lung Cancer 2014, 85, 141–146. [Google Scholar] [CrossRef]
  92. Smyth, L.M.; Tamura, K.; Oliveira, M.; Ciruelos, E.M.; Mayer, I.A.; Sablin, M.-P.; Biganzoli, L.; Ambrose, H.J.; Ashton, J.; Barnicle, A.; et al. Capivasertib, an AKT Kinase Inhibitor, as Monotherapy or in Combination with Fulvestrant in Patients with AKT1E17K-Mutant, ER-Positive Metastatic Breast Cancer. Clin. Cancer Res. Off. J. Am. Assoc. Cancer Res. 2020, 26, 3947–3957. [Google Scholar] [CrossRef] [Green Version]
  93. Jones, R.H.; Casbard, A.; Carucci, M.; Cox, C.; Butler, R.; Alchami, F.; Madden, T.-A.; Bale, C.; Bezecny, P.; Joffe, J.; et al. Fulvestrant plus Capivasertib versus Placebo after Relapse or Progression on an Aromatase Inhibitor in Metastatic, Oestrogen Receptor-Positive Breast Cancer (FAKTION): A Multicentre, Randomised, Controlled, Phase 2 Trial. Lancet Oncol. 2020, 21, 345–357. [Google Scholar] [CrossRef]
  94. Capivasertib Plus Faslodex Reduced the Risk of Disease Progression or Death by 40% versus Faslodex in Advanced HR-Positive Breast Cancer. Available online: https://www.astrazeneca.com/media-centre/press-releases/2022/capivasertib-pfs-in-hr-positive-breast-cancer.html (accessed on 8 December 2022).
  95. Yap, T.A.; Kristeleit, R.; Michalarea, V.; Pettitt, S.J.; Lim, J.S.; Carreira, S.; Roda, D.; Miller, R.; Riisnaes, R.; Miranda, S.; et al. Phase I Trial of the PARP Inhibitor Olaparib and AKT Inhibitor Capivasertib in Patients with BRCA1/2- and Non-BRCA1/2-Mutant Cancers. Cancer Discov. 2020, 10, 1528–1543. [Google Scholar] [CrossRef]
  96. Bhattarai, T.S.; Shamu, T.; Gorelick, A.N.; Chang, M.T.; Chakravarty, D.; Gavrila, E.I.; Donoghue, M.T.A.; Gao, J.; Patel, S.; Gao, S.P.; et al. AKT Mutant Allele-Specific Activation Dictates Pharmacologic Sensitivities. Nat. Commun. 2022, 13, 2111. [Google Scholar] [CrossRef]
  97. Blagden, S.P.; Hamilton, A.L.; Mileshkin, L.; Wong, S.; Michael, A.; Hall, M.; Goh, J.C.; Lisyanskaya, A.S.; DeSilvio, M.; Frangou, E.; et al. Phase IB Dose Escalation and Expansion Study of AKT Inhibitor Afuresertib with Carboplatin and Paclitaxel in Recurrent Platinum-Resistant Ovarian Cancer. Clin. Cancer Res. Off. J. Am. Assoc. Cancer Res. 2019, 25, 1472–1478. [Google Scholar] [CrossRef] [Green Version]
  98. Lin, J.; Sampath, D.; Nannini, M.A.; Lee, B.B.; Degtyarev, M.; Oeh, J.; Savage, H.; Guan, Z.; Hong, R.; Kassees, R.; et al. Targeting Activated Akt with GDC-0068, a Novel Selective Akt Inhibitor That Is Efficacious in Multiple Tumor Models. Clin. Cancer Res. Off. J. Am. Assoc. Cancer Res. 2013, 19, 1760–1772. [Google Scholar] [CrossRef] [Green Version]
  99. Saura, C.; Roda, D.; Roselló, S.; Oliveira, M.; Macarulla, T.; Pérez-Fidalgo, J.A.; Morales-Barrera, R.; Sanchis-García, J.M.; Musib, L.; Budha, N.; et al. A First-In-Human Phase I Study of the ATP-Competitive AKT Inhibitor Ipatasertib Demonstrates Robust and Safe Targeting of AKT in Patients with Solid Tumors. Cancer Discov. 2017, 7, 102–113. [Google Scholar] [CrossRef] [Green Version]
  100. Isakoff, S.; Tabernero, J.; Molife, L.; Soria, J.-C.; Cervantes, A.; Vogelzang, N.; Patel, M.; Hussain, M.; Baron, A.; Argilés, G.; et al. Antitumor Activity of Ipatasertib Combined with Chemotherapy: Results from a Phase Ib Study in Solid Tumors. Ann. Oncol. 2020, 31, 626–633. [Google Scholar] [CrossRef] [Green Version]
  101. Kim, S.-B.; Dent, R.; Im, S.-A.; Espié, M.; Blau, S.; Tan, A.R.; Isakoff, S.J.; Oliveira, M.; Saura, C.; Wongchenko, M.J.; et al. Ipatasertib plus Paclitaxel versus Placebo plus Paclitaxel as First-Line Therapy for Metastatic Triple-Negative Breast Cancer (LOTUS): A Multicentre, Randomised, Double-Blind, Placebo-Controlled, Phase 2 Trial. Lancet Oncol. 2017, 18, 1360–1372. [Google Scholar] [CrossRef]
  102. Turner, N.; Dent, R.A.; O’Shaughnessy, J.; Kim, S.-B.; Isakoff, S.J.; Barrios, C.; Saji, S.; Bondarenko, I.; Nowecki, Z.; Lian, Q.; et al. Ipatasertib plus Paclitaxel for PIK3CA/AKT1/PTEN-Altered Hormone Receptor-Positive HER2-Negative Advanced Breast Cancer: Primary Results from Cohort B of the IPATunity130 Randomized Phase 3 Trial. Breast Cancer Res. Treat. 2022, 191, 565–576. [Google Scholar] [CrossRef]
  103. Turner, N.C.; Alarcón, E.; Armstrong, A.C.; Philco, M.; Chuken, Y.L.; Sablin, M.-P.; Tamura, K.; Villanueva, A.G.; Pérez-Fidalgo, J.A.; Cheung, S.Y.A.; et al. BEECH: A Dose-Finding Run-In Followed by a Randomised Phase II Study Assessing the Efficacy of AKT Inhibitor Capivasertib (AZD5363) Combined with Paclitaxel in Patients with Estrogen Receptor-Positive Advanced or Metastatic Breast Cancer, and in a PIK3CA Mutant Sub-Population. Ann. Oncol. Off. J. Eur. Soc. Med. Oncol. 2019, 30, 774–780. [Google Scholar] [CrossRef]
  104. Mukhopadhyay, S.; Frias, M.A.; Chatterjee, A.; Yellen, P.; Foster, D.A. The Enigma of Rapamycin Dosage. Mol. Cancer Ther. 2016, 15, 347–353. [Google Scholar] [CrossRef] [Green Version]
  105. Chatterjee, A.; Mukhopadhyay, S.; Tung, K.; Patel, D.; Foster, D.A. Rapamycin-Induced G1 Cell Cycle Arrest Employs Both TGF-β and Rb Pathways. Cancer Lett. 2015, 360, 134–140. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Fingar, D.C.; Richardson, C.J.; Tee, A.R.; Cheatham, L.; Tsou, C.; Blenis, J. mTOR Controls Cell Cycle Progression through Its Cell Growth Effectors S6K1 and 4E-BP1/Eukaryotic Translation Initiation Factor 4E. Mol. Cell. Biol. 2004, 24, 200–216. [Google Scholar] [CrossRef] [PubMed]
  107. Yellen, P.; Chatterjee, A.; Preda, A.; Foster, D.A. Inhibition of S6 Kinase Suppresses the Apoptotic Effect of eIF4E Ablation by Inducing TGF-β-Dependent G1 Cell Cycle Arrest. Cancer Lett. 2013, 333, 239–243. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Yellen, P.; Saqcena, M.; Salloum, D.; Feng, J.; Preda, A.; Xu, L.; Rodrik-Outmezguine, V.; Foster, D.A. High-Dose Rapamycin Induces Apoptosis in Human Cancer Cells by Dissociating mTOR Complex 1 and Suppressing Phosphorylation of 4E-BP1. Cell Cycle 2011, 10, 3948–3956. [Google Scholar] [CrossRef] [Green Version]
  109. MTOR Inhibition Induces Upstream Receptor Tyrosine Kinase Signaling and Activates Akt—PubMed. Available online: https://pubmed.ncbi.nlm.nih.gov/16452206/ (accessed on 1 January 2023).
  110. Sun, S.-Y.; Rosenberg, L.M.; Wang, X.; Zhou, Z.; Yue, P.; Fu, H.; Khuri, F.R. Activation of Akt and eIF4E Survival Pathways by Rapamycin-Mediated Mammalian Target of Rapamycin Inhibition. Cancer Res. 2005, 65, 7052–7058. [Google Scholar] [CrossRef] [Green Version]
  111. Le Gendre, O.; Sookdeo, A.; Duliepre, S.-A.; Utter, M.; Frias, M.; Foster, D.A. Suppression of AKT Phosphorylation Restores Rapamycin-Based Synthetic Lethality in SMAD4-Defective Pancreatic Cancer Cells. Mol. Cancer Res. 2013, 11, 474–481. [Google Scholar] [CrossRef] [Green Version]
  112. Popova, N.V.; Jücker, M. The Role of mTOR Signaling as a Therapeutic Target in Cancer. Int. J. Mol. Sci. 2021, 22, 1743. [Google Scholar] [CrossRef]
  113. Tian, T.; Li, X.; Zhang, J. mTOR Signaling in Cancer and mTOR Inhibitors in Solid Tumor Targeting Therapy. Int. J. Mol. Sci. 2019, 20, 755. [Google Scholar] [CrossRef] [Green Version]
  114. Hall, C.P.; Reynolds, C.P.; Kang, M.H. Modulation of Glucocorticoid Resistance in Pediatric T-Cell Acute Lymphoblastic Leukemia by Increasing BIM Expression with the PI3K/mTOR Inhibitor BEZ235. Clin. Cancer Res. 2016, 22, 621–632. [Google Scholar] [CrossRef] [Green Version]
  115. Gazi, M.; Moharram, S.A.; Marhäll, A.; Kazi, J.U. The Dual Specificity PI3K/mTOR Inhibitor PKI-587 Displays Efficacy against T-Cell Acute Lymphoblastic Leukemia (T-ALL). Cancer Lett. 2017, 392, 9–16. [Google Scholar] [CrossRef] [Green Version]
  116. The Cancer Genome Atlas Research Network. Comprehensive Molecular Profiling of Lung Adenocarcinoma. Nature 2014, 511, 543–550. [Google Scholar] [CrossRef] [Green Version]
  117. Zhang, Y.; Yan, H.; Xu, Z.; Yang, B.; Luo, P.; He, Q. Molecular Basis for Class Side Effects Associated with PI3K/AKT/mTOR Pathway Inhibitors. Expert Opin. Drug Metab. Toxicol. 2019, 15, 767–774. [Google Scholar] [CrossRef]
  118. Arteaga, C.L.; Engelman, J.A. ERBB Receptors: From Oncogene Discovery to Basic Science to Mechanism-Based Cancer Therapeutics. Cancer Cell 2014, 25, 282–303. [Google Scholar] [CrossRef] [Green Version]
  119. Zhang, H.; Berezov, A.; Wang, Q.; Zhang, G.; Drebin, J.; Murali, R.; Greene, M.I. ErbB Receptors: From Oncogenes to Targeted Cancer Therapies. Available online: https://www.jci.org/articles/view/32278/pdf (accessed on 13 November 2022).
  120. Pistilli, B.; Pluard, T.; Urruticoechea, A.; Farci, D.; Kong, A.; Bachelot, T.; Chan, S.; Han, H.S.; Jerusalem, G.; Urban, P.; et al. Phase II Study of Buparlisib (BKM120) and Trastuzumab in Patients with HER2+ Locally Advanced or Metastatic Breast Cancer Resistant to Trastuzumab-Based Therapy. Breast Cancer Res. Treat. 2018, 168, 357–364. [Google Scholar] [CrossRef]
  121. Agus, D.B.; Akita, R.W.; Fox, W.D.; Lewis, G.D.; Higgins, B.; Pisacane, P.I.; Lofgren, J.A.; Tindell, C.; Evans, D.P.; Maiese, K.; et al. Targeting Ligand-Activated ErbB2 Signaling Inhibits Breast and Prostate Tumor Growth. Cancer Cell 2002, 2, 127–137. [Google Scholar] [CrossRef] [Green Version]
  122. Zambrano, C.C.; Schuler, M.H.; Machiels, J.-P.H.; Hess, D.; Paz-Ares, L.; Awada, A.; von Moos, R.; Steeghs, N.; Ahnert, J.R.; De Mesmaeker, P.; et al. Phase Ib Study of Buparlisib (BKM120) plus Either Paclitaxel (PTX) in Advanced Solid Tumors (aST) or PTX plus Trastuzumab (TZ) in HER2+ Breast Cancer (BC). J. Clin. Oncol. 2014, 32, 627. [Google Scholar] [CrossRef]
  123. Steelman, L.S.; Chappell, W.H.; Abrams, S.L.; Kempf, C.R.; Long, J.; Laidler, P.; Mijatovic, S.; Maksimovic-Ivanic, D.; Stivala, F.; Mazzarino, M.C.; et al. Roles of the Raf/MEK/ERK and PI3K/PTEN/Akt/mTOR Pathways in Controlling Growth and Sensitivity to Therapy-Implications for Cancer and Aging. Aging 2011, 3, 192–222. [Google Scholar] [CrossRef] [Green Version]
  124. Dong, C.; Wu, J.; Chen, Y.; Nie, J.; Chen, C. Activation of PI3K/AKT/mTOR Pathway Causes Drug Resistance in Breast Cancer. Front. Pharmacol. 2021, 12, 628690. [Google Scholar] [CrossRef]
  125. Shapiro, G.I.; LoRusso, P.; Kwak, E.; Pandya, S.; Rudin, C.M.; Kurkjian, C.; Cleary, J.M.; Pilat, M.J.; Jones, S.; de Crespigny, A.; et al. Phase Ib Study of the MEK Inhibitor Cobimetinib (GDC-0973) in Combination with the PI3K Inhibitor Pictilisib (GDC-0941) in Patients with Advanced Solid Tumors. Investig. New Drugs 2020, 38, 419–432. [Google Scholar] [CrossRef]
  126. Britten, C.D. PI3K and MEK Inhibitor Combinations: Examining the Evidence in Selected Tumor Types. Cancer Chemother. Pharmacol. 2013, 71, 1395–1409. [Google Scholar] [CrossRef]
  127. Asati, V.; Mahapatra, D.K.; Bharti, S.K. PI3K/Akt/mTOR and Ras/Raf/MEK/ERK Signaling Pathways Inhibitors as Anticancer Agents: Structural and Pharmacological Perspectives. Eur. J. Med. Chem. 2016, 109, 314–341. [Google Scholar] [CrossRef] [PubMed]
  128. Hoeflich, K.P.; Merchant, M.; Orr, C.; Chan, J.; Otter, D.D.; Berry, L.; Kasman, I.; Koeppen, H.; Rice, K.; Yang, N.-Y.; et al. Intermittent Administration of MEK Inhibitor GDC-0973 plus PI3K Inhibitor GDC-0941 Triggers Robust Apoptosis and Tumor Growth Inhibition. Cancer Res. 2012, 72, 210–219. [Google Scholar] [CrossRef] [PubMed]
  129. Hoeflich, K.P.; O’Brien, C.; Boyd, Z.; Cavet, G.; Guerrero, S.; Jung, K.; Januario, T.; Savage, H.; Punnoose, E.; Truong, T.; et al. In vivo Antitumor Activity of MEK and Phosphatidylinositol 3-Kinase Inhibitors in Basal-Like Breast Cancer Models. Clin. Cancer Res. Off. J. Am. Assoc. Cancer Res. 2009, 15, 4649–4664. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  130. Schram, A.M.; Gandhi, L.; Mita, M.M.; Damstrup, L.; Campana, F.; Hidalgo, M.; Grande, E.; Hyman, D.M.; Heist, R.S. A Phase Ib Dose-Escalation and Expansion Study of the Oral MEK Inhibitor Pimasertib and PI3K/MTOR Inhibitor Voxtalisib in Patients with Advanced Solid Tumours. Br. J. Cancer 2018, 119, 1471–1476. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  131. Algazi, A.P.; Rotow, J.; Posch, C.; Ortiz-Urda, S.; Pelayo, A.; Munster, P.N.; Daud, A. A Dual Pathway Inhibition Strategy Using BKM120 Combined with Vemurafenib Is Poorly Tolerated in BRAF V600 E/K Mutant Advanced Melanoma. Pigment. Cell Melanoma Res. 2019, 32, 603–606. [Google Scholar] [CrossRef]
  132. Tolcher, A.W.; Patnaik, A.; Papadopoulos, K.P.; Rasco, D.W.; Becerra, C.R.; Allred, A.J.; Orford, K.; Aktan, G.; Ferron-Brady, G.; Ibrahim, N.; et al. Phase I Study of the MEK Inhibitor Trametinib in Combination with the AKT Inhibitor Afuresertib in Patients with Solid Tumors and Multiple Myeloma. Cancer Chemother. Pharmacol. 2015, 75, 183–189. [Google Scholar] [CrossRef]
  133. Hoxhaj, G.; Manning, B.D. The PI3K-AKT Network at the Interface of Oncogenic Signalling and Cancer Metabolism. Nat. Rev. Cancer 2020, 20, 74–88. [Google Scholar] [CrossRef]
  134. Gaglio, D.; Metallo, C.M.; A Gameiro, P.; Hiller, K.; Danna, L.S.; Balestrieri, C.; Alberghina, L.; Stephanopoulos, G.; Chiaradonna, F. Oncogenic K-Ras Decouples Glucose and Glutamine Metabolism to Support Cancer Cell Growth. Mol. Syst. Biol. 2011, 7, 523. [Google Scholar] [CrossRef]
  135. Yun, J.; Rago, C.; Cheong, I.; Pagliarini, R.; Angenendt, P.; Rajagopalan, H.; Schmidt, K.; Willson, J.K.V.; Markowitz, S.; Zhou, S.; et al. Glucose Deprivation Contributes to the Development of KRAS Pathway Mutations in Tumor Cells. Science 2009, 325, 1555–1559. [Google Scholar] [CrossRef] [Green Version]
  136. Ying, H.; Kimmelman, A.C.; Lyssiotis, C.A.; Hua, S.; Chu, G.C.; Fletcher-Sananikone, E.; Locasale, J.W.; Son, J.; Zhang, H.; Coloff, J.L.; et al. Oncogenic Kras Maintains Pancreatic Tumors through Regulation of Anabolic Glucose Metabolism. Cell 2012, 149, 656–670. [Google Scholar] [CrossRef] [Green Version]
  137. Kole, H.K.; Resnick, R.J.; Van Doren, M.; Racker, E. Regulation of 6-Phosphofructo-1-Kinase Activity in Ras-Transformed Rat-1 Fibroblasts. Arch. Biochem. Biophys. 1991, 286, 586–590. [Google Scholar] [CrossRef]
  138. Racker, E.; Resnick, R.J.; Feldman, R. Glycolysis and Methylaminoisobutyrate Uptake in Rat-1 Cells Transfected with Ras or Myc Oncogenes. Proc. Natl. Acad. Sci. USA 1985, 82, 3535–3538. [Google Scholar] [CrossRef]
  139. Chun, S.Y.; Johnson, C.; Washburn, J.G.; Cruz-Correa, M.R.; Dang, D.T.; Dang, L.H. Oncogenic KRAS Modulates Mitochondrial Metabolism in Human Colon Cancer Cells by Inducing HIF-1α and HIF-2α Target Genes. Mol. Cancer 2010, 9, 293. [Google Scholar] [CrossRef] [Green Version]
  140. Bender, A.; Opel, D.; Naumann, I.; Kappler, R.; Friedman, L.; von Schweinitz, D.; Debatin, K.-M.; Fulda, S. PI3K Inhibitors Prime Neuroblastoma Cells for Chemotherapy by Shifting the Balance towards Pro-Apoptotic Bcl-2 Proteins and Enhanced Mitochondrial Apoptosis. Oncogene 2011, 30, 494–503. [Google Scholar] [CrossRef] [Green Version]
  141. Opel, D.; Naumann, I.; Schneider, M.; Bertele, D.; Debatin, K.-M.; Fulda, S. Targeting Aberrant PI3K/Akt Activation by PI103 Restores Sensitivity to TRAIL-Induced Apoptosis in Neuroblastoma. Clin. Cancer Res. Off. J. Am. Assoc. Cancer Res. 2011, 17, 3233–3247. [Google Scholar] [CrossRef] [Green Version]
  142. Yamamoto, N.; Fujiwara, Y.; Tamura, K.; Kondo, S.; Iwasa, S.; Tanabe, Y.; Horiike, A.; Yanagitani, N.; Kitazono, S.; Inatani, M.; et al. Phase Ia/Ib Study of the Pan-Class I PI3K Inhibitor Pictilisib (GDC-0941) Administered as a Single Agent in Japanese Patients with Solid Tumors and in Combination in Japanese Patients with Non-Squamous Non-Small Cell Lung Cancer. Investig. New Drugs 2017, 35, 37–46. [Google Scholar] [CrossRef]
  143. Bang, Y.-J.; Kang, Y.-K.; Ng, M.; Chung, H.; Wainberg, Z.; Gendreau, S.; Chan, W.; Xu, N.; Maslyar, D.; Meng, R.; et al. A Phase II, Randomised Study of mFOLFOX6 with or without the Akt Inhibitor Ipatasertib in Patients with Locally Advanced or Metastatic Gastric or Gastroesophageal Junction Cancer. Eur. J. Cancer 2019, 108, 17–24. [Google Scholar] [CrossRef]
  144. Kumar, A.; Fernandez-Capetillo, O.; Carrera, A.C. Nuclear Phosphoinositide 3-Kinase β Controls Double-Strand Break DNA Repair. Proc. Natl. Acad. Sci. USA 2010, 107, 7491–7496. [Google Scholar] [CrossRef] [Green Version]
  145. Kao, G.D.; Jiang, Z.; Fernandes, A.M.; Gupta, A.K.; Maity, A. Inhibition of Phosphatidylinositol-3-OH Kinase/Akt Signaling Impairs DNA Repair in Glioblastoma Cells following Ionizing Radiation. J. Biol. Chem. 2007, 282, 21206–21212. [Google Scholar] [CrossRef] [Green Version]
  146. Wang, D.; Li, C.; Zhang, Y.; Wang, M.; Jiang, N.; Xiang, L.; Li, T.; Roberts, T.M.; Zhao, J.J.; Cheng, H.; et al. Combined inhibition of PI3K and PARP Is Effective in the Treatment of Ovarian Cancer Cells with Wild-Type PIK3CA Genes. Gynecol. Oncol. 2016, 142, 548–556. [Google Scholar] [CrossRef] [Green Version]
  147. Ibrahim, Y.H.; García-García, C.; Serra, V.; He, L.; Torres-Lockhart, K.; Prat, A.; Anton, P.; Cozar, P.; Guzmán, M.; Grueso, J.; et al. PI3K Inhibition Impairs BRCA1/2 Expression and Sensitizes BRCA-Proficient Triple-Negative Breast Cancer to PARP Inhibition. Cancer Discov. 2012, 2, 1036–1047. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  148. Batalini, F.; Xiong, N.; Tayob, N.; Polak, M.; Eismann, J.; Cantley, L.C.; Shapiro, G.I.; Adalsteinsson, V.; Winer, E.P.; Konstantinopoulos, P.A.; et al. Phase 1b Clinical Trial with Alpelisib plus Olaparib for Patients with Advanced Triple-Negative Breast Cancer. Clin. Cancer Res. 2022, 28, 1493–1499. [Google Scholar] [CrossRef] [PubMed]
  149. Matulonis, U.; Wulf, G.M.; Birrer, M.J.; Westin, S.N.; Quy, P.; Bell-McGuinn, K.M.; Lasonde, B.; Whalen, C.; Aghajanian, C.; Solit, D.B.; et al. Phase I Study of Oral BKM120 and Oral Olaparib for High-Grade Serous Ovarian Cancer (HGSC) or Triple-negative Breast Cancer (TNBC). J. Clin. Oncol. 2014, 32, 2510. [Google Scholar] [CrossRef]
  150. Jin, M.-Z.; Jin, W.-L. The Updated Landscape of Tumor Microenvironment and Drug Repurposing. Signal Transduct. Target. Ther. 2020, 5, 166. [Google Scholar] [CrossRef] [PubMed]
  151. Chen, D.S.; Mellman, I. Oncology Meets Immunology: The Cancer-Immunity Cycle. Immunity 2013, 39, 1–10. [Google Scholar] [CrossRef] [Green Version]
  152. Song, M.; Chen, D.; Lu, B.; Wang, C.; Zhang, J.; Huang, L.; Wang, X.; Timmons, C.L.; Hu, J.; Liu, B.; et al. PTEN Loss Increases PD-L1 Protein Expression and Affects the Correlation between PD-L1 Expression and Clinical Parameters in Colorectal Cancer. PLoS ONE 2013, 8, e65821. [Google Scholar] [CrossRef]
  153. Zhang, Z.; Richmond, A.; Yan, C. Immunomodulatory Properties of PI3K/AKT/mTOR and MAPK/MEK/ERK Inhibition Augment Response to Immune Checkpoint Blockade in Melanoma and Triple-Negative Breast Cancer. Int. J. Mol. Sci. 2022, 23, 7353. [Google Scholar] [CrossRef]
  154. Peng, W.; Chen, J.Q.; Liu, C.; Malu, S.; Creasy, C.; Tetzlaff, M.T.; Xu, C.; McKenzie, J.A.; Zhang, C.; Liang, X.; et al. Loss of PTEN Promotes Resistance to T Cell-Mediated Immunotherapy. Cancer Discov. 2016, 6, 202–216. [Google Scholar] [CrossRef] [Green Version]
  155. Jiang, X.; Zhou, J.; Giobbie-Hurder, A.; Wargo, J.; Hodi, F.S. The Activation of MAPK in Melanoma Cells Resistant to BRAF Inhibition Promotes PD-L1 Expression That Is Reversible by MEK and PI3K Inhibition. Clin. Cancer Res. Off. J. Am. Assoc. Cancer Res. 2013, 19, 598–609. [Google Scholar] [CrossRef] [Green Version]
  156. Lastwika, K.J.; Wilson, W.; Li, Q.K.; Norris, J.; Xu, H.; Ghazarian, S.R.; Kitagawa, H.; Kawabata, S.; Taube, J.M.; Yao, S.; et al. Control of PD-L1 Expression by Oncogenic Activation of the AKT-mTOR Pathway in Non-Small Cell Lung Cancer. Cancer Res. 2016, 76, 227–238. [Google Scholar] [CrossRef] [Green Version]
  157. Zheng, W.; O’Hear, C.E.; Alli, R.; Basham, J.H.; Abdelsamed, H.A.; Palmer, L.E.; Jones, L.L.; Youngblood, B.; Geiger, T.L. PI3K Orchestration of the in vivo Persistence of Chimeric Antigen Receptor-Modified T Cells. Leukemia 2018, 32, 1157–1167. [Google Scholar] [CrossRef]
  158. Perkins, D.M.R.; Grande, S.; Hamel, B.A.; Horton, H.M.; Garrett, B.T.E.; Miller, S.M.; Latimer, I.H.J.; Horvath, D.C.J.; Kuczewski, M.M.; Friedman, K.M.; et al. Manufacturing an Enhanced CAR T Cell Product By Inhibition of the PI3K/Akt Pathway during T Cell Expansion Results in Improved in vivo Efficacy of Anti-BCMA CAR T Cells. Blood 2015, 126, 1893. [Google Scholar] [CrossRef]
  159. Alzahrani, A.S. PI3K/Akt/mTOR Inhibitors in Cancer: At the Bench and Bedside. Semin. Cancer Biol. 2019, 59, 125–132. [Google Scholar] [CrossRef]
  160. Schrijver, W.A.M.E.; Suijkerbuijk, K.P.M.; van Gils, C.H.; van der Wall, E.; Moelans, C.B.; van Diest, P.J. Receptor Conversion in Distant Breast Cancer Metastases: A Systematic Review and Meta-Analysis. J. Natl. Cancer Inst. 2018, 110, 568–580. [Google Scholar] [CrossRef] [Green Version]
  161. Thulin, A.; Andersson, C.; Rönnerman, E.W.; De Lara, S.; Chamalidou, C.; Schoenfeld, A.; Kovács, A.; Fagman, H.; Enlund, F.; Linderholm, B.K. Discordance of PIK3CA and TP53 Mutations between Breast Cancer Brain Metastases and Matched Primary Tumors. Sci. Rep. 2021, 11, 23548. [Google Scholar] [CrossRef]
  162. Fumagalli, C.; Ranghiero, A.; Gandini, S.; Corso, F.; Taormina, S.; De Camilli, E.; Rappa, A.; Vacirca, D.; Viale, G.; Guerini-Rocco, E.; et al. Inter-Tumor Genomic Heterogeneity of Breast Cancers: Comprehensive Genomic Profile of Primary Early Breast Cancers and Relapses. Breast Cancer Res. BCR 2020, 22, 107. [Google Scholar] [CrossRef]
  163. Toppmeyer, D.L.; Press, M.F. Testing Considerations for Phosphatidylinositol-3-Kinase Catalytic Subunit Alpha as an Emerging Biomarker in Advanced Breast Cancer. Cancer Med. 2020, 9, 6463–6472. [Google Scholar] [CrossRef]
  164. Dumbrava, E.; Call, S.; Huang, H.; Stuckett, A.; Madwani, K.; Adat, A.; Hong, D.; Piha-Paul, S.; Subbiah, V.; Karp, D.; et al. PIK3CA Mutations in Plasma Circulating Tumor DNA Predict Survival and Treatment Outcomes in Patients with Advanced Cancers. ESMO Open 2021, 6, 100230. [Google Scholar] [CrossRef]
  165. Brandão, M.; Caparica, R.; Eiger, D.; de Azambuja, E. Biomarkers of Response and Resistance to PI3K Inhibitors in Estrogen Receptor-Positive Breast Cancer Patients and Combination Therapies Involving PI3K Inhibitors. Ann. Oncol. Off. J. Eur. Soc. Med. Oncol. 2019, 30, x27–x42. [Google Scholar] [CrossRef] [Green Version]
  166. Wee, S.; Wiederschain, D.; Maira, S.-M.; Loo, A.; Miller, C.; Debeaumont, R.; Stegmeier, F.; Yao, Y.-M.; Lengauer, C. PTEN-Deficient Cancers Depend on PIK3CB. Proc. Natl. Acad. Sci. USA 2008, 105, 13057–13062. [Google Scholar] [CrossRef] [Green Version]
  167. O’Brien, C.; Wallin, J.J.; Sampath, D.; GuhaThakurta, D.; Savage, H.; Punnoose, E.A.; Guan, J.; Berry, L.; Prior, W.W.; Amler, L.C.; et al. Predictive Biomarkers of Sensitivity to the Phosphatidylinositol 3′ Kinase Inhibitor GDC-0941 in Breast Cancer Preclinical Models. Clin. Cancer Res. Off. J. Am. Assoc. Cancer Res. 2010, 16, 3670–3683. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  168. Hopkins, B.D.; Pauli, C.; Du, X.; Wang, D.G.; Li, X.; Wu, D.; Amadiume, S.C.; Goncalves, M.D.; Hodakoski, C.; Lundquist, M.R.; et al. Suppression of Insulin Feedback Enhances the Efficacy of PI3K Inhibitors. Nature 2018, 560, 499–503. [Google Scholar] [CrossRef] [PubMed]
  169. Bosch, A.; Li, Z.; Bergamaschi, A.; Ellis, H.; Toska, E.; Prat, A.; Tao, J.J.; Spratt, D.E.; Viola-Villegas, N.T.; Castel, P.; et al. PI3K Inhibition Results in Enhanced Estrogen Receptor Function and Dependence in Hormone Receptor—Positive Breast Cancer. Sci. Transl. Med. 2015, 7, 283ra51. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  170. Mayer, I.A.; Abramson, V.G.; Formisano, L.; Balko, J.M.; Estrada, M.V.; Sanders, M.E.; Juric, D.; Solit, D.; Berger, M.F.; Won, H.H.; et al. A Phase Ib Study of Alpelisib (BYL719), a PI3Kα-Specific Inhibitor, with Letrozole in ER+/HER2− Metastatic Breast Cancer. Clin. Cancer Res. Off. J. Am. Assoc. Cancer Res. 2017, 23, 26–34. [Google Scholar] [CrossRef] [Green Version]
  171. El Bairi, K.; Haynes, H.R.; Blackley, E.; Fineberg, S.; Shear, J.; Turner, S.; de Freitas, J.R.; Sur, D.; Amendola, L.C.; Gharib, M.; et al. The Tale of TILs in Breast Cancer: A Report from The International Immuno-Oncology Biomarker Working Group. NPJ Breast Cancer 2021, 7, 150. [Google Scholar] [CrossRef]
  172. Gagliato, D.D.M.; Cortes, J.; Curigliano, G.; Loi, S.; Denkert, C.; Perez-Garcia, J.; Holgado, E. Tumor-Infiltrating Lymphocytes in Breast Cancer and Implications for Clinical Practice. Biochim. Et Biophys. Acta (BBA) Rev. Cancer 2017, 1868, 527–537. [Google Scholar] [CrossRef]
  173. Mego, M.; Gao, H.; Cohen, E.; Anfossi, S.; Giordano, A.; Sanda, T.; Fouad, T.; De Giorgi, U.; Giuliano, M.; Woodward, W.; et al. Circulating Tumor Cells (CTC) Are Associated with Defects in Adaptive Immunity in Patients with Inflammatory Breast Cancer. J. Cancer 2016, 7, 1095–1104. [Google Scholar] [CrossRef] [Green Version]
  174. De Giorgi, U.; Mego, M.; Scarpi, E.; Giordano, A.; Giuliano, M.; Valero, V.; Alvarez, R.H.; Ueno, N.T.; Cristofanilli, M.; Reuben, J.M. Association between Circulating Tumor Cells and Peripheral Blood Monocytes in Metastatic Breast Cancer. Ther. Adv. Med Oncol. 2019, 11, 1758835919866065. [Google Scholar] [CrossRef] [Green Version]
  175. Gianni, C.; Palleschi, M.; Schepisi, G.; Casadei, C.; Bleve, S.; Merloni, F.; Sirico, M.; Sarti, S.; Cecconetto, L.; Di Menna, G.; et al. Circulating Inflammatory Cells in Patients with Metastatic Breast Cancer: Implications for Treatment. Front. Oncol. 2022, 12, 882896. [Google Scholar] [CrossRef]
Figure 1. The PI3K/AKT/mTOR pathway is involved in tumorigenesis and cancer progression. After being activated by RTKs, GPCR or RAS, PI3K catalyzes the phosphorylation of PIP2 to generate PIP3, which binds and recruits AKT and PDK1. Furthermore, by activating NF-κB and inducing the secretion of MMP, AKT promotes cell invasion while increasing the level of cyclin D1, leading to cell cycle progression. Ultimately, Akt promotes cell growth by phosphorylation of the downstream mTORC1, which activates p70S6K-S6 and inhibits 4E-BP1, resulting in protein synthesis and cell growth. Indeed, mTORC2 activates AKT itself. On the other hand, PTEN exerts its role in modulating the PI3K pathway by suppressing PIP2 to PIP3 conversion. Together with tuberous sclerosis protein 1 (TSC1) and TSC2, PTEN is the main negative regulator of the pathway. Simultaneously, activation of the growth factor receptor tyrosine kinases and G protein-coupled receptors induces RAS/RAF/MEK/ERK signalling, and ERK activation can further contribute to mTORC1 activation.
Figure 1. The PI3K/AKT/mTOR pathway is involved in tumorigenesis and cancer progression. After being activated by RTKs, GPCR or RAS, PI3K catalyzes the phosphorylation of PIP2 to generate PIP3, which binds and recruits AKT and PDK1. Furthermore, by activating NF-κB and inducing the secretion of MMP, AKT promotes cell invasion while increasing the level of cyclin D1, leading to cell cycle progression. Ultimately, Akt promotes cell growth by phosphorylation of the downstream mTORC1, which activates p70S6K-S6 and inhibits 4E-BP1, resulting in protein synthesis and cell growth. Indeed, mTORC2 activates AKT itself. On the other hand, PTEN exerts its role in modulating the PI3K pathway by suppressing PIP2 to PIP3 conversion. Together with tuberous sclerosis protein 1 (TSC1) and TSC2, PTEN is the main negative regulator of the pathway. Simultaneously, activation of the growth factor receptor tyrosine kinases and G protein-coupled receptors induces RAS/RAF/MEK/ERK signalling, and ERK activation can further contribute to mTORC1 activation.
Cancers 15 00703 g001
Figure 2. Summary of the complex phosphatidylinositol-3-kinase (PI3K)/AKT/mTOR signalling pathway and inhibitors.
Figure 2. Summary of the complex phosphatidylinositol-3-kinase (PI3K)/AKT/mTOR signalling pathway and inhibitors.
Cancers 15 00703 g002
Table 1. Summary of ongoing phases I–III trials with PI3k inhibitors in solid tumors.
Table 1. Summary of ongoing phases I–III trials with PI3k inhibitors in solid tumors.
Clinical TrialStudy DesignInterventionSettingsPrimary EndpointPhaseStatus
NCT0497595863 Participants
Interventional
Non-Randomized
Parallel Assignment
Open Label
Buparlisib
Atezolizumab
AN0025
Advanced
Solid tumors
DLTs1Recruiting
NCT04338399
(BURAN)
483 Participants
Interventional
Randomized
Parallel Assignment
Open Label
Buparlisib
Paclitaxel
mHNCCOS3Recruiting
NCT0410885812 Participants
Interventional
Parallel Assignment
Open Label
Copanlisib
Pertuzumab
Trastuzumab
PI3KCA/PTEN mutated
HER2+/HR-
MBC
AEs1/2Recruiting
NCT0457276348 Participants
Interventional
Non-Randomized
Single-Group Assignment
Open Label
Copanlisib
Venetoclax
Relapsed/refractory
DLBCL
MTD, ORR1/2Active
Not recruiting
NCT0371105818 Participants
Interventional
Non-Randomized
Sequential Assignment
Open Label
Copanlisib
Nivolumab
MSS CRCMTD, DLT1/2Active
Not recruiting
NCT0425326213 Participants
Interventional
Non-Randomized
Sequential Assignment
Open Label
Copanlisib
Rucaparib
mCRPCMTD1/2Active
Not recruiting
NCT0350273348 Participants
Interventional
Single-Group Assignment
Open Label
Copanlisib
Ipililumab
Nivolumab
Advanced cancer,
Lymphoma
RP2D1Active,
Not recruiting
NCT03484819106 Participants
Interventional
Single-Group Assignment
Open Label
Copanlisib Hydrochlorid NivolumabRefractory DLBCL
PMBCL
ORR2Active,
Not recruiting
NCT02367040
CHRONOS-3
458 Participants
Interventional
Randomized
Parallel Assignment
Copanlisib
Rituximab
Relapsed
iNHL
PFS2Active,
Not recruiting
NCT01660451227 Participants
Interventional
Non-Randomized
Parallel Assignment
Open Label
CopanlisibIndolent or aggressive NHLORR Active,
Not recruiting
NCT0514322918 Participants
Interventional
Non-Randomized
Sequential Assignment
Open Label
Alpelisib
Sacituzumab Govitecan
Stage III/Stage IV
HR+/HER2−
MBC
RP2D1Recruiting
NCT04208178
(EPIK-B2)
551 Participants
Interventional
Randomized
Parallel Assignment
Alpelisib
Trastuzumab
Pertuzumab
PIK3CA mutated
HER2+ MBC
PFS3Recruiting
NCT0476297944 Participants
Interventional
Single-Group Assignment
Open Label
Alpelisib
Fulvestrant
Aromatase inhibitor
PIK3CA mutated
HR+/HER2−
MBC
PFS2Recruiting
NCT0550890660 Participants
Interventional
Non-Randomized
Parallel Assignment
Open Label
Alpelisib
Ribociclib
OP-1250
HR+/HER2−
MBC
DLTsMTD1Recruiting
NCT04251533566 Participants
Interventional
Randomized
Parallel Assignment
Alpelisib
Nab paclitaxel
Placebo
PIK3CA mutated/PTEN loss
mTNBC
PFS, ORR3Recruiting
NCT0502573525 Participants
Interventional
Randomized
Single-Group Assignment
Open Label
Alpelisib Dapagliflozin FulvestrantPI3KCA mutated
HR+/HER2−
MBC
Incidence of all grade hyperglycemia2Recruiting
NCT0523081040 Participants
Interventional
Single-Group Assignment
Open Label
Alpelisib
Fulvestrant
Tucatinib
PIK3CA mutated
HER2+
MBC
AEs1/2Recruiting
NCT05501886
(VIKTORIA-1)
701 Participants
Interventional
Randomized
Parallel Assignment
Open Label
Alpelisib Gedatolisib
Palbociclib
Fulvestrant
HR+/HER2−
MBC
PFS3Recruiting
NCT0499790236 Participants
Interventional
Parallel Assignment
Open Label
Alpelisib
Tipifarnib
mHNCCDLTs1/2Recruiting
NCT05063786358 Participants
Interventional
Single-Group Assignment
Open Label
Alpelisib
Olaparib
Paclitaxel
PLD
metastatic
OC
PFS3Recruiting
NCT0452647055 Participants
Interventional
Single-Group Assignment
Open Label
Alpelisib
Paclitaxel
PIK3CA mutated
GA
MTD
RP2D
1/2Recruiting
NCT0320752928 Participants
Interventional
Single-Group Assignment
Open Label

Alpelisib
Enzalutamide
AR+/PTEN positive
MBC
MTD1Recruiting
NCT01872260255 Participants
Interventional
Randomized
Parallel Assignment
Open Label
Alpelisib
Letrozole
LEE011
HR+/HER2−
MBC
DLTs
Safety
1/2Active,
Not yet recruiting
NCT03284957
(AMEERA-1)
136 Participants
Interventional
Randomized
Parallel Assignment
Open Label
Amcenestrant
Palbociclib
Alpelisib
Everolimus
Abemaciclib
HR+/HER2−
MBC
DLTs Active,
Not yet recruiting
NCT04666038
(BRUIN CLL-321)
250 Participants
Interventional Randomized
Parallel Assignment
Open Label
Idelalisib
LOXO-305
Bendamustine
Rituximab
Chronic
CLL/SLL
PFS3Recruiting
NCT03890289
(GAUDEALIS)
5 Participants
Interventional
Single-Group Assignment
Open Label
Idelalisib
Obinutuzumab
Refractory
FL
ORR2Active
Not yet recruiting
NCT027873693 Participants
Interventional
Non-Randomized
Parallel Assignment
Open Label
Idelalisib ACY-1215
Ibrutinib
Refractory CLLMTD1Active
Not yet recruiting
NCT02970318311 Participants
Interventional
Randomized
Parallel Assignment
Idelalisib calabrutinib (ACP-196)
Rituximab
Bendamustine
Refractory CLLPFS3Active
Not yet recruiting
NCT0213513350 Participants
Interventional
Single-Group Assignment
Open Label
Idelalisib
Ofatumumab
CLL/SLLORR2Active,
Not recruiting
NCT04191499400 Participants
Interventional
Randomized
Parallel Assignment
Inavolisib Palbociclib
Fulvestrant
PIK3CA mutated
HR+/HER2+
MBC
PFS2/3Recruiting
Abbreviations: AEs: adverse events; AR: androgen receptor; CLL: chronic lymphocytic leukemia; CRC: colorectal cancer; DLBCL: diffuse large B-cell lymphoma; DLTs: dose-limiting toxicities; mHNCC: metastatic head and neck cancer; MBC: metastatic breast cancer; iNHL: indolent B-cell non-Hodgkin’s lymphoma; FL: follicular lymphoma; mCRPC: metastatic castration-resistant prostate cancer; MSS: microsatellite stable; MTD: maximum tolerated dose; NHL: non-Hodgkin’s lymphomas; OC: ovarian cancer; ORR: objective response rate; OS: overall survival; PMBCL: primary mediastinal large B-cell lymphoma; PLD: pegylated liposomal doxorubicin; PFS: progression-free survival; RP2D: recommended phase 2 dose; SLL: small lymphocytic lymphoma.
Table 2. Summary of ongoing phases I–III trials with AKT-inhibitors in tumors.
Table 2. Summary of ongoing phases I–III trials with AKT-inhibitors in tumors.
Clinical TrialStudy DesignInterventionSettingsPrimary EndpointPhaseStatus
NCT0331054112 Participants
Interventional
Parallel Assignment
Open Label
Capivasertib,
Enzalutamide,
Fulvestrant
Advanced solid tumors harboring mutations in AKT1, AKT2, or AKT3ORR1Active, not yet recruiting
NCT05593497
(SNARE)
30 Participants
Interventional
Single-Group Assignment
Open Label
Capivasertib,
Abiraterone Acetate
Leuprolide
PTEN loss
High-risk localized
PC
pCR
MRD
2Not recruiting
NCT04439123
(MATCH-Subprotocol Y)
35 Participants
Interventional
Single-Group Assignment
Open Label
CapivasertibCancers with AKT genetic changesORR2Active, not yet recruiting
NCT0485161320 Participants
Interventional
Non-Randomized
Single-Group Assignment
Open Label
Afuresertib,
Fulvestrant

Locally advanced or HR+/HER2− MBC
ORR1Recruiting
NCT04374630
(PROFECTA-II)
141 Participants
Interventional
Parallel Assisgnment
Open Label
Afuresertib
Paclitaxel
Platinum-resistant
ovarian cancer
rPFS2Recruiting
NCT05383482167 Participants
Non-Randomized
Sequential Assignment
Open Label
Afuresertib
Nab paclitaxel
Docetaxel
Sintilimab
Solid tumors
Resistant to prior anti-PD-1/PD-L1
AEs
DLTs
1/2Recruiting
NCT05390710101 Participants
Randomized
Sequential Assignment
Open Label
LAE005 + Afuresertib Nab-PaclitaxelMetastatic
TNBC
AEs
DLT
1/2Recruiting
NCT0406039474 Participants
Randomized
Sequential Assignment
Open Label
Afuresertib LAE001/prednisone +mCRPC
rPFS
1/2Recruiting
NCT04253561
(IPATHER)
25 Participants
Interventional
Single-Group Assignment
Open Label
Ipatasertib Trastuzumab
Pertuzumab
HER2+ PI3KCA
mutant
MBC
RP2D1Recruiting
NCT0517224536 Participants
Interventional
Single-Group Assignment
Open Label
Ipatasertib
Cisplatin
Radiation Therapy
Stage III-IVB
HNC
MTD
RP2D
1Recruiting
NCT04467801
(Ipat-Lung)
60 Participants
Interventional
Single-Group Assignment
Open Label
Ipatasertib
Docetaxel
mNSCLCPFS2Recruiting
NCT0367378787 Participants
Interventional
Non-Randomized
Parallel Assignment
Open Label
Ipatasertib
atezolizumab
Glioblastoma
Multiforme
mPC
MTD1/2Recruiting
NCT0527697324 Participants
Interventional
Single-Group Assignment
Open Label
Ipatasertib
Carboplatin
Paclitaxel
Stage III or IV
Epithelial
OC
MTD1Recruiting
NCT03959891
(TAKTIC)
60 Participants
Interventional
Non-Randomized
Parallel Assignment
Open Label
Ipatasertib
Fulvestrant
Aromatase Inhibitor
Palbociclib
HR+/HER2−
mBC
TEAE1Recruiting
NCT04650581
(FINER)
250 Participants
Interventional
Randomized
Parallel Assignment
Ipatasertib
Fulvestrant
HR+/HER2−
mBC
PFS3Recruiting
NCT04920708
Without ctDNA Suppression
(FAIM)
324 Participants
Interventional
Randomized
Parallel Assignment
Open Label
Ipatasertib
Fulvestrant
Palbociclib
HR+/HER2−
mBC
PFS2Not yet recruiting
NCT04464174
(PATHFINDER)
54 Participants
Interventional
Non-randomized
Parallel Assignment
Open Label
Ipatasertib
non-taxane chemotherapy
mTNBCSafety2Active, not yet recruiting
NCT0385370728 Participants
Interventional
Randomized
Parallel Assignment
Open Label
Ipatasertib
Atezolizumab
Capecitabine
Carboplatin
Ipatasertib
Paclitaxel
mTNBCRP2D,
PFS
1/2Active, not yet recruiting
NCT05498896
(BARBICAN)
146 Participants
Interventional
Non-randomized
Parallel Assignment
Open Label
Ipatasertib
Atezolizumab
Paclitaxel
Doxorubicin
Cyclophosphamide
mTNBCpCR2Active, not yet recruiting
NCT03072238
(IPATential150)
1101 Participants
Interventional
Parallel Assignment
Ipatasertib
Abiraterone
Placebo
mCRPCrPFS3Active, not yet recruiting
NCT0553889796 Participants
Interventional
Randomized
Parallel Assignment
Open Label
Ipatasertib
Megestrol Acetate
mECAEs1/2Not yet recruiting
NCT04739202
((IMMUNOGAST)
60 Participants
Interventional
Non-Randomized
Parallel Assignment
Open Label
Ipatasertib AtezolizumabmGAORR2Recruiting
Abbreviations: AEs: adverse events; DLT: dose-limiting toxicities; HNC: head and neck cancer; mBC: metastatic breast cancer; mCRPC: metastatic castration-resistant prostate cancer; mEC: metastatic endometrial cancer; mGA: metastatic gastric adenocarcinoma; MRD: minimal residual disease; mPC: metastatic prostate cancer; MTD: maximum tolerated dose; ORR: objective response rate; OC: ovarian cancer; PC: prostate cancer; pCR: pathologic complete response; PFS: progression free survival; rPFS: radiographic progression-free survival; RP2D: recommended phase 2 dose; TEAE: treatment-emergent adverse events.
Table 3. Summary of ongoing phases II-III trials with dual PI3K/mTOR inhibitors.
Table 3. Summary of ongoing phases II-III trials with dual PI3K/mTOR inhibitors.
Clinical TrialStudy DesignInterventionSettingsPrimary EndpointPhaseStatus
NCT0369838315 Participants
Interventional
Single-Group Assignment
Open Label
Trastuzumab biosimilars (Herzuma) GedatolisibHER2+
MBC
ORR2Recruiting
NCT0391197352 Participants
Interventional
Single-Group Assignment
Open Label
Talazoparib
Gedatolisib
mTNBCORR½Recruiting
NCT0306506296 Participants
Interventional
Single-Group Assignment
Open Label
Palbociclib
Gedatolisib
Solid tumorsMTD, RP2D1Recruiting
NCT05501886
(VIKTORIA-1)
141 Participants
Interventional Randomized
Parallel Assisgnment
Open Label
Palbociclib
Fulvestrant
Alpelisib
Gedatolisib
HR+/HER2
MBC
PFS3Recruiting
Abbreviations: MBC: metastatic breast cancer; MTD: maximum tolerated dose; ORR: objective response rate; PFS: progression-free survival; RP2D: recommended phase 2 dose (RP2D).
Table 4. Ongoing clinical trials combining immune checkpoint inhibitors with inhibitors of the PI3K/AKT/mTOR pathway.
Table 4. Ongoing clinical trials combining immune checkpoint inhibitors with inhibitors of the PI3K/AKT/mTOR pathway.
Clinical TrialStudy DesignInterventionSettingsPrimary EndpointPhaseStatus
NCT0443163535 Participants
Interventional
Non-Randomized
Single-Group Assignment
Open Label
Copanlisib Nivolumab RituximabRelapsed/refractory indolent follicular or marginal zone lymphomaMDT
CR rate
IbRecruiting
NCT0396169891 Participants
Interventional
Non-Randomized
Parallel Assignment
Eganelisib Atezolizumab Nab-paclitaxel
Bevacizumab
Metastatic TNBC or advanced RCCCR rateIIActive,
not recruiting
NCT04317105102 Participants
Interventional
Non-Randomized
Parallel Assignment
Open Label
Copanlisib Ipilimumab
Nivolumab
Advanced malignant solid neoplasmAEs,
DLT
I/IIRecruiting
NCT0313190836 Participants
Interventional
Non-Randomized
Parallel Assignment
Open Label
GSK2636771
Pembrolizumab
Metastatic PTEN loss
melanoma
MTD
ORR
I/IIActive,
not recruiting
NCT0377256140 Participants
Interventional
Non-Randomized
Single-Group Assignment
Open Label
Capivasertib
Olaparib
Durvalumab
Advanced or metastatic solid tumor malignanciesORRIRecruiting
NCT0538761698 Participants
Interventional
Non-Randomized
Single-Group Assignment
Open Label
Copanlisib
Obinutuzumab
Follicular lymphomaPFSIIRecruiting
NCT0371105854 Participants
Interventional
Non-Randomized
Sequential Assignment
Open Label
Copanlisib
Nivolumab
MSS relapsed/refractory solid tumors and CRCDLTs,
ORR
I/IIActive,
not recruiting
NCT0367378787 Participants
Non-Randomized
Parallel Assignment
Open Label
Ipatasertib AtezolizumabAdvanced solid tumours with PI3K pathway hyperactivationMTD, AEsI/IIRecruiting
NCT02637531219 Participants
Interventional
Non-Randomized
Single-Group Assignment
Open Label
Eganelisib NivolumabAdvanced
solid tumours
DLTs, AEsI/IbActive,
not recruiting
Abbreviations: AEs: adverse events; CLL: chronic lymphocytic leukemia; CR: complete response rate; CRC: colorectal cancer; DLTs: dose-limiting toxicities; TNBC: triple negative breast cancer; MSS: microsatellite stability; MTD: maximum tolerated dose; ORR: objective response rate; PFS: progression-free survival; RCC: renal cell carcinoma.
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

Sirico, M.; D’Angelo, A.; Gianni, C.; Casadei, C.; Merloni, F.; De Giorgi, U. Current State and Future Challenges for PI3K Inhibitors in Cancer Therapy. Cancers 2023, 15, 703. https://doi.org/10.3390/cancers15030703

AMA Style

Sirico M, D’Angelo A, Gianni C, Casadei C, Merloni F, De Giorgi U. Current State and Future Challenges for PI3K Inhibitors in Cancer Therapy. Cancers. 2023; 15(3):703. https://doi.org/10.3390/cancers15030703

Chicago/Turabian Style

Sirico, Marianna, Alberto D’Angelo, Caterina Gianni, Chiara Casadei, Filippo Merloni, and Ugo De Giorgi. 2023. "Current State and Future Challenges for PI3K Inhibitors in Cancer Therapy" Cancers 15, no. 3: 703. https://doi.org/10.3390/cancers15030703

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