Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2012 Aug 7.
Published in final edited form as: Mol Cancer Ther. 2011 Jan 7;10(3):395–403. doi: 10.1158/1535-7163.MCT-10-0905

Pushing the envelope in the mTOR pathway. The second generation of inhibitors

Eduardo Vilar 1, Jose Perez-Garcia 2, Josep Tabernero 2
PMCID: PMC3413411  NIHMSID: NIHMS262374  PMID: 21216931

Abstract

The phosphatydilinositol-3-kinase (PI3K)/mammalian target of rapamycin (mTOR) pathway has been a major focus of attention for cancer researchers in the past decade. A preliminary and not complete understanding of the molecular biology of this complex network has not only importantly conditioned the development of the first generation of mTOR inhibitors, but also the biomarker studies designed to identify the best responders to these agents. Most recently, research in this pathway has focused in the fact of the dual nature of mTOR that is integrated by the mTOR complex 1 (mTORC1) and complex 2 (mTORC2). These two complexes are formed and regulated by different proteins, and also driven by multiple different compensatory feedback loops. This deeper understanding has allowed the development of a promising second generation of inhibitors which are able to block simultaneously both complexes due to their catalytic activity over mTOR. Moreover, some of them also exert an inhibitory effect over PI3K that is a key player in the feedback loops. This article reviews the newest insights in the signaling of the mTOR pathway and then focuses in the development of the new wave of mTOR inhibitors.

Keywords: Dual inhibitors, mTOR, mTORC1, mTORC2, PI3K, Rapamycin, Rapalogs

Introduction

Since the discovery of mammalian target of rapamycin (mTOR) in the early 1990s the volume of research performed in this pathway has been substantial. These data have provided us with an increasingly detailed knowledge about the proteins and regulators involved in it, their different functions, and the genetic abnormalities that are present across different tumor types. Moreover, the interest among the scientific community for this pathway has been fostered by the development of a natural product derived from the bacterium Streptomyces hygroscopicus. This compound called Rapamycin (Sirolimus, Rapamune; Wyeth) has shown inhibitory activity against mTOR protein after coupling its intracellular receptor. Subsequently, several compounds have been synthesized with similar characteristics to Rapamycin integrating the family of Rapalogs.

However, the clinical results obtained by targeting this pathway have not been as straight forward as it was presumed at the beginning. Moreover, drug development against mTOR was started when the knowledge about its functions was very preliminary. Several key findings have changed the course of clinical research in this field. First, the fact that mTOR is constituted by two complexes: mTOR complex 1 (mTORC1) and mTOR complex 2 (mTORC2); those have a very intricate network of feedback loops, protein partners, substrates, and regulators that are specific to each. Second, the discovery that Rapamycin and Rapalogs exert an incomplete inhibition of mTORC1 and also are inactive against mTORC2. Finally, mTORC2 was shown to be one of the major regulators of the feedback loops associated with this pathway, thus explaining the limited activity of Rapalogs observed in clinical studies. Therefore, a closer analysis of the recent advances in the molecular biology of this pathway will help to correctly understand the results from previous in vitro studies and clinical trials.

In the present article we will review the data on the characterization of mTORC1 and mTORC2, their protein components, functions, and regulators emphasizing the role of the feedback loops recently described within this complex network. Then, the approved indications for the Rapalogs will be summarized. Finally, the last section will be devoted to a new class of compounds that are able to inhibit both mTOR complexes, and the new dual inhibitors that are also adding activity against the phosphatydilinositol-3-kinase (PI3K), a key component of the main feedback loop involved in this pathway.

Molecular biology of the mTOR pathway. A story of two complexes

The PI3K-AKT-mTOR pathway (Figure 1) is commonly altered in human cancers. Deregulation can be secondary to amplification or mutations in PIK3CA, which encodes the p110α catalytic subunit of the kinase complex and have been extensively described in several tumors (1); mutations and amplification in AKT; inactivation or mutations in phosphatase and tensin homolog (PTEN); and other less frequent events such as mutations in the insulin receptor-substrates (IRS) and the Ras homolog enriched in brain (RHEB) (24).

Figure 1.

Figure 1

PI3K-AKT-mTOR signaling pathway. A more detailed description of the biology of the mTOR has revealed the presence of two complexes and has led to the development of new drugs targeting specifically these complexes. In addition, feedback loops have been better characterized.

Phosphatydilinositol-3-kinase, PI3K; mammalian target of rapamycin, mTOR; mTOR complex, mTORC; tuberous sclerosis complex, TSC; eukaryotic initiation factor 4E binding protein 1, 4E-BP1; ribosomal S6 kinase 1, S6K1; Ras homolog enriched in brain, Rheb; serum- and glucocorticoid-induced protein kinase-1, SGK1; protein kinase C, PKC; insulin receptor-substrate 1, IRS-1; mitogen–activated protein kinase kinase, MEK; receptor tyrosine-kinase, RTK; regulatory-associated protein of mTOR, Raptor; proline-rich AKT substrate of 40 kDa, PRAS40; phosphatase and tensin homolog, PTEN; Phosphatidylinositol (4,5)-bisphosphate, PIP2; Phosphatidylinositol (3,4,5)-trisphosphate, PIP3; phosphoinositide-dependent kinase-1, PDK1.

mTOR is a serine/threonine kinase formed by two signaling complexes called mTORC1 and mTORC2 that contain common and specific partners proteins. Both complexes share the following proteins: mTOR, mLST8/GβL, and the negative regulator Deptor. On the other hand, they are integrated by distinct partner proteins and regulatory mechanisms acting on different substrates, and having specific effects on distinct cellular functions (5). mTORC1 is specifically composed by a regulatory-associated protein of mTOR (Raptor) and a proline-rich AKT substrate of 40 kDa (PRAS40). mTORC2 couples with the Rapamycin-insentive companion of mTOR (Rictor), mSin1, and PRR5/Protor (Figure 1).

mTORC1 enhances cell growth and proliferation by inducing protein and lipid synthesis, ribosome biogenesis, and reduction of autophagy (69). Growth factors and nutrients, such as energy and amino acids, promote mTORC1 signaling through the phosphorylation of eukaryotic initiation factor 4E binding protein 1 (4E-BP1) and ribosomal S6 kinase 1 (S6K1) which are the best-known downstream effectors of mTOR (10).

The tuberous sclerosis complex 2 (TSC2) is an essential link between growth-factor signaling and the mTORC1 activation triggered via PI3K-dependent or -independent pathways (1113). Growth-factor signaling phosphorylates and inhibits TSC2 (Tuberin), avoiding its association with TSC1 (Hamartin), thus activating the mTORC1 by releasing the inhibition of RHEB, a small guanosine triphosphatase (GTPase) necessary for the activation of mTORC1 (14). Likewise, inactivating mutations in the TSC1 or TSC2 genes cause hamartoma syndromes associated with elevated mTORC1 activity (15, 16). However, TSC2 is not required for the regulation of mTORC1 by amino acids; the Ras-related GTPase (RAG) proteins, a family of small GTPases, are the key regulators for mTORC1 amino acid activation (17). Other regulators of mTORC1 are Raptor, that positively regulates mTORC1 and functions as a scaffold for recruiting mTORC1 substrates (18), and PRAS40 and Deptor that act as negative regulators (19, 20). The function of mLST8/GβL remains yet unknown (21).

Much less is known about mTORC2 functions, substrates, and regulators. Unlike mTORC1, which is a direct-target of Rapamycin, mTORC2 was initially described as Rapamycin-insensitive (22), although it has been recently reported that continued exposure to Rapamycin also leads to its inhibition (23). mTORC2 promotes cell survival, actin cytoskeleton organization and is exclusively growth-factor responsive, being AKT its first recognized substrate protein. Full activation of AKT requires the phosphorylation of two residues: Ser473 by mTORC2 and Thr308 by phosphoinositide-dependent kinase-1 (PDK1) (24). Other mTORC2 substrates are serum- and glucocorticoid-induced protein kinase-1 (SGK1) and protein kinase C-alpha (PKCα) (25, 26).

The regulatory mechanisms of mTORC2 also remain partially unknown, although it has been shown that Rictor and mSin1 enhance mTORC2 signaling while Deptor appears to negatively regulate it. TSC1 and TSC2 have also been involved in promoting mTORC2 activation. However, the function of PRR5/Protor is still not well defined (21, 27).

Finally, the mTOR complex has also been suggested to play a crucial role integrating extracellular and intracellular signals that regulates cellular metabolism. This also includes the control of inflammatory and tolerance responses via regulation of TCRζ (28) and TGF-β-induced Foxp3, respectively (29). Although these physiologic functions need further mechanistic elucidation, they also open new avenues for development of biomarkers of mTOR inhibition through other alternative effects.

The mTOR pathway. An intrincated network with feedback loops

Development of resistance to mTORC1 inhibitors has been related with the presence of different feedback loops described within this complex network. Moreover, a better understanding of these mechanisms may help to identify novel therapeutic strategies to overcome the relative lack of efficacy of these compounds (5).

It is postulated that mTORC1 activation causes a negative feedback through S6K1 that reduce the activity of PI3K. The phosphorylation of S6K1 inactivates IRS-1 which is required for insulin signaling through PI3K (30). Therefore, mTOR inhibition will induce IRS-1 activation releasing the inhibition mediated by S6K1 and provoking the activation of AKT via an insulin growth factor receptor 1 (IGF-1R) dependent signaling process (31). O’Reilly et al published supporting evidence for this negative feedback loop. They have observed in a panel of cancer cell lines from different tumor types that Rapamycin was able to upregulate IRS-1 levels and promote AKT phosphorylation (32). Accordingly to these findings of a biomarker study developed in the context of the first phase I clinical trial with Everolimus (RAD001, Afinitor; Novartis) showed a dose- and schedule-dependent inhibition of mTOR and a subsequent upregulation of AKT. These effects were observed in 50% of the patients and were assessed in both tumor and skin biopsies, thus validating the in vitro observation (33). Moreover, Wan et al showed in human rhabdomyosarcoma cell lines and xenografts that blockade of IGF-1R led to an inhibition of the Rapamycin-induced AKT activation (31), providing evidence for a synergistic effect of mTOR and IGF-1R inhibition. This combination is currently under clinical evaluation in a phase I multiple-dose escalating study using Dalotuzumab, (a monoclonal antibody against IGF-1R; MK-0646; Merck) and Ridaforolimus (an mTORC1 small-molecule inhibitor analog of the Rapamycin; MK-8669, Deforolimus; Merck and ARIAD). Preliminary results have revealed important antitumor activity in estrogen receptor-positive and highly proliferative breast tumors, which frequently harbor PIK3CA mutations and IGF-1R overexpression (34). Other two studies of the combination of Cixutumumab (IGF-1R monoclonal antibody inhibitor; IMC-A12; Imclone) plus the Rapalog Temsirolimus (CCI-779, Torisel; Wyeth), and Figitumumab (IGF-1R monoclonal antibody inhibitor; CP-751871; Pfizer) plus Everolimus are underway (35, 36).

Furthermore, preclinical data have shown that mTORC1 inhibition results in a hyperactivation of the PI3K pathway and simultaneous increase of the signaling through the mitogen–activated protein kinase kinase (MAPK) pathway (37), thus proving the existence of another feedback loop that connect the PI3K-AKT-mTOR with the MAPK pathway. This observation has provided rationale for combining several ongoing phase I clinical trials combining mTOR, PI3K, or AKT inhibitors with MAP/ERK kinase (MEK) inhibitors. However, the most optimal combination of inhibitors deserves careful consideration due to dense cross-talk interactions among protein components of these complex pathways. Sophisticated systems biology analyses have recently predicted adverse effects in terms of reduction of citotoxicity with the combination of a MEK and a first generation mTOR inhibitor. Specifically, in vitro validation of this in silico data showed that Rapamycin, which led to significant activation of AKT, upon combination with a MEK inhibitor (U0126) rendered an increase in cell viability. In contrast, simultaneous inhibition of PI3K-AKT and MAPK pathways decreased cell viability and pointed towards as this combination as the most optimal way to effectively inhibit both pathways (38). On the other side, clinical studies have reported significant toxicities in a phase I trial which is testing the combination of an AKT inhibitor and a MEK inhibitor. Considering these preclinical and clinical results in conjunction, the combination of PI3K or second generation mTOR inhibitors with MEK inhibitors warrants further clinical validation.

First generation of mTOR inhibitors

The first generation inhibitors of mTOR are derivatives of Rapamycin that specifically inhibit mTORC1. This group of drugs is integrated by Rapamycin and its analogs also known as Rapalogs: Everolimus, Temsirolimus, and Ridaforolimus (previously known as Deforolimus). Rapamycin has been clinically approved several years ago for prophylaxis of organ rejection for renal transplant patients (Table 1 and Figure 2) (39).

Table 1.

Rapalogs and approved indications from the FDA and EMEA. Food and Drug Administration, FDA; European Medicines Agency, EMEA.

Compound Approved indication Agency Ref
Sirolimus Prophylaxis of organ rejection in renal transplant patients FDA/EMEA (39)
Everolimus Refractory advanced renal cell carcinoma FDA/EMEA (54)
Temsirolimus Poor-prognosis untreated advanced renal cell carcinoma FDA/EMEA (53)
Refractory mantle-cell lymphoma EMEA (55)
Ridaforolimus No approved indication. Phase I-II-III trials ongoing ClinicalTrials.gov

Figure 2.

Figure 2

Molecular structures of first and second generation of mTOR inhibitors. Rapalogs are displayed in the first row. Second generation inhibitors are displayed in the second and subsequent rows. Structures of NVP-BGT226, GDC-0980, SB-2312, INK-128, XL-388 have not been disclosed at the time of publication of this article.

The mechanism of action of Rapamycin has been very well described. This drug along with the FK506-binding protein (FKBP12) targets the FKBP12-Rapamycin binding (FRB) domain adjacent to the catalytic site of the mTOR protein (40). Several studies have shown that mTORC2 is Rapamycin-insensitive (22, 41), although long-term exposure to Rapamycin can also inhibit mTORC2 and then disrupt AKT signaling. Strikingly, this response has been shown to be tissue specific (23).

mTORC1-mediated 4E-BP1 phosphorylation induces the dissociation of 4E-BP1 from the eukaryotic initiation factor 4E (eIF4E), thus allowing the assembly of the eIF4F complex to initiate cap-dependent mRNA translation. 4E-BP1 is phosphorylated at multiple sites such as Thr36, Thr45, Ser64, Thr69, and Ser82 and needs to occur in a pre-specified order (42). The activation of Thr36 and Thr45 are the leading events necessary for phosphorylation of Thr69 that will be followed by Ser82 (43). Except for Ser82, all phosphorylation sites are sensitive to Rapamycin demonstrated by the complete inhibition of the initiation of cap-dependent mRNA translation by the treatment with Rapamycin in specific cellular and histological contexts (44). However, it has been recently observed in that Rapalogs may not fully block 4E-BP1 despite of a complete inhibition of S6K1 (45). This fact could be due to different reasons such as a relative lack of effect on the phosphorylation of Thr36 and Thr45 (46, 47), the existence of unknown feedback loops, and the inability to inhibit mTORC2; and it explains the unpredictable antitumor effect of Rapalogs across different cancer subtypes (4852). Mechanistic details are discussed in the section devoted to second generation inhibitors.

Despite of their limited cytotoxic activity, Rapalogs have demonstrated antiproliferative properties. Temsirolimus and Everolimus have been approved by the Food and Drug Administration (FDA) and the European Medicines Agency (EMEA) for treatment advanced renal cell carcinoma; Temsirolimus has been authorized for treatment of relapsed or refractory mantle-cell lymphoma by the EMEA only (Table 1 and Figure 2). The approval of Temsirolimus for treatment of previously untreated metastatic renal cell carcinoma was based on the results from a phase III clinical trial in which 626 patients randomly received Temsirolimus, Interferon-alfa, or combination therapy with Temsirolimus and Interferon-alfa. Temsirolimus alone rendered longer overall survival (10.9 vs 7.3 months; HR=0.73; P-value = 0.008) and progression-free survival than Interferon-alfa alone (5.5 vs 3.8 months; P-value < 0.001). In addition, no differences between the combination-therapy and the Interferon group were observed in terms of overall survival (53). After that, Everolimus was approved for the treatment of patients with advanced renal cell carcinoma who had progressed on Sorafenib, Sunitinib, or both. The authorization was supported by the data coming from a phase III clinical trial that randomized 410 patients to receive Everolimus or placebo in a 2:1 radtio. Everolimus showed a significant improvement in progression-free survival with mild adverse effects (4 vs 1.9 months; HR=0.30; P-value <0.001) (54). Finally, Temsirolimus showed improvement in progression-free survival and higher objective response rates compared with investigator’s choice treatment in patients with relapsed or refractory mantle-cell lymphoma leading to the approval by the EMEA (55).

Therefore, the next step in the development of Rapalogs will be the discovery of new biomarkers to predict what tumor subtypes and specific molecular features are more likely to respond to mTOR inhibitors. In this regard, responses to PI3K-AKT-mTOR pathway inhibitors may be higher among those tumors harboring PIK3CA mutations (56) and also those with loss of PTEN (57). Another example of response to Rapalogs in specific tumor subtypes is the case of Microsatellite Instable (MSI) colorectal cancers. PI3K-AKT-mTOR pathway has been involved in the pathogenesis of colorectal cancer. In fact, PIK3CA mutations have been identified in approximately 20–30% of colorectal tumors, and have been associated with shorter cancer-specific survival, poorer outcomes and resistance to Cetuximab (1, 58, 59). Although single-agent Everolimus has not achieved objective responses in refractory metastatic colorectal cancer (48), in vitro studies have suggested that colorectal tumors displaying MSI could potentially respond better to therapies against the PI3K-AKT-mTOR pathway (60). According to these results, dual PI3K-mTOR inhibitors may represent an interesting option to be evaluated in this specific tumor subtype.

Second generation of mTOR inhibitors

Whereas Rapamycin exerts its action almost exclusively through mTORC1 inhibition, a second generation of inhibitors targeting the adenosine triphosphate site of the kinase domain of mTOR has been developed. These compounds are able to block both mTORC1 and mTORC2. Theoretically, their most important advantages would be a significant decrease of AKT phosphorylation upon mTORC2 blockade and a better mTORC1 inhibition. In addition, the preclinical data of these agents have contributed to a better understanding of the functions of mTORC2 and the limitations of Rapalogs.

Due to the fact that the catalytic domain of mTOR and the p110α subunit of PI3K are structurally related, some of these second generation compounds have dual activity against both PI3K and mTOR. These drugs compared to single specific-mTORC1 and -PI3K inhibitors have the potential benefit of inhibiting mTORC1, mTORC2, and all the catalytic isoforms of PI3K (61). Therefore, targeting both kinases simultaneously should reduce the upregulation of PI3K that typically produced upon inhibition of mTORC1 (30).

The majority of dual PI3K-mTOR inhibitors have already entered into phase I–II clinical trials alone or in combination with other agents for different cancer subtypes (Table 2). NVP-BEZ235 (Novartis) is one of these dual kinase inhibitors and reversibly blocks the p110α catalytic subunit of PI3K and mTOR (62). Initial in vitro data analyzing pharmacodynamic endpoints in breast tumor xenografts treated with NVP-BEZ235 have shown a decrease in phosphorylation levels of AKT, 4E-BP1, and S6K1 following treatment with this drug and higher antiproliferative activity than Everolimus (63). A phase I of NVP-BEZ235 has been recently presented with promising efficacy. Among 51 evaluable and heavily pretreated patients, 14 achieved stable disease longer than 4 months and partial responses were observed in breast and lung tumors. However, pharmacokinetic studies showed that the area under the curve (AUC) increased non-proportionally with dose, so future studies will use a new formulation of the drug. No dose-limiting toxicities were reported and the maximum tolerated has not been reached (64). On the other hand, XL-765 (Exelixis) has exhibited potent pharmacodynamic effects on the inhibition of PI3K with a stable pharmacokinetic profile. In addition, durable disease stabilizations were observed in patients with different tumor types such as colorectal cancer, lung cancer, renal cell carcinoma, mesothelioma, and appendiceal cancer (65).

Table 2. Dual PI3K-mTOR, mTORC1 and mTORC2 inhibitors and status of drug development.

Phosphatydilinositol-3-kinase (PI3K); mammalian target of rapamycin (mTOR); mTOR complex (mTORC). Clinical Development, CD.

Compound Drug Company Targets Status
NVP-BGT226 Novartis PI3K/mTORC1/mTORC2 CD terminated
NVP-BEZ235 Novartis PI3K/mTORC1/mTORC2 Phase I/II
SF-1126 Semaphore Pharmaceuticals PI3K/mTORC1/mTORC2 Phase I
XL-765 Exelisis PI3K/mTORC1/mTORC2 Phase I/II
PKI-587/PF-05212384 Pfizer PI3K/mTOR Phase I
PF-04691502 Pfizer PI3K/mTOR Phase I
GDC-0980 Genentech PI3K/mTOR Phase I
SB-2312 S*Bio PI3K/mTOR Preclinical
PKI-402 Pfizer PI3K/mTOR Preclinical

OSI-027 Osi Pharmaceuticals mTORC1/mTORC2 Phase I
AZD-8055 Astra Zeneca mTORC1/mTORC2 Phase I
INK-128 Intellikine mTORC1/mTORC2 Phase I
XL-388 Exelixis mTORC1/mTORC2 Preclinical
PP-242 University of California mTORC1/mTORC2 Preclinical
PP-30 University of California mTORC1/mTORC2 Preclinical
Torin-1 Gray Laboratory Harvard mTORC1/mTORC2 Preclinical
KU-0063794 Kudos Pharmaceuticals mTORC1/mTORC2 Preclinical
WYE-125132 Wyeth mTORC1/mTORC2 Preclinical
Palomid-529 Paloma Pharmaceuticals mTORC1/mTORC2 Preclinical
WAY-600 Wyeth mTORC1/mTORC2 Preclinical
WYE-687 Wyeth mTORC1/mTORC2 Preclinical
WYE-354 Wyeth mTORC1/mTORC2 Preclinical

Regarding single specific mTOR catalytic inhibitors, several small molecules have also been identified (Table 2 and Figure 2), and three of them have entered into phase I clinical development [AZD-8055 (AstraZeneca), INK-128 (Intellikine), and OSI-027 (OSI Pharmaceuticals)]. Preclinical data with INK-128 have shown a potent inhibition of the phosphorylation of S6K1, 4E-BP1, and AKT at Ser473 in vitro, as well as important antiproliferative activity against multiple xenograft models and cells lines resistant to Rapamycin and pan-PI3K inhibitors (66). At the same time Feldman et al have reported the activity of two compounds PP-242 and PP-30 (University of California) with activity against both mTORC1 and mTORC2. These compounds are able to completely suppress 4E-BP1 and S6K1 along with a reduction of phosphorylation of AKT at Ser473, thus leading to a higher antiproliferative effect compared to Rapamycin. However, the inhibition of mTORC2 did not result in a total blockade of AKT, suggesting that additional mTORC1 inhibition by these compounds could be the basis for their superior antitumor activity (67). In this regard, Hsieh et al suggested that the therapeutic benefit of PP-242 is mediated through the inhibition of mTORC1-dependent 4E-BP1-eIF4E hyperactivation (68). Other preclinical studies with these ATP-competitive and -specific mTOR inhibitors have observed similar results and have confirmed its activity over those Rapamycin-resistant functions of mTORC1. In addition, these drugs induce a stronger G1 cell cycle arrest in several cancer lines and formidable autophagy activation (6973). Finally, a first-in-human phase I study exploring three schedules of OSI-027 has been recently presented with preliminary evidence of pharmacological activity. The maximum tolerated dose has not yet been defined and dose escalation is ongoing. Left ventricular ejection fraction and fatigue have been reported as dose-limiting toxicities (74). In the following years, we will obtain more detailed data from phase I studies regarding the pharmacokinetic profile, optimal dose, toxicity and preliminary activity of all of these compounds.

Conclusions

mTOR is one of the signaling pathways that has attracted more interest among basic and clinical researchers. Two main factors are the responsible for this phenomenon: mTOR is a downstream central effector of multiple pathways thus making it a very attractive target, and the drug Rapamycin which renders an incomplete inhibition of this protein complex became available in 1975. These facts have fostered the efforts of the pharmaceutical industry in order to synthesize newer and better compounds against it. In a relatively short period of time several companies have launched development programs of different drugs blocking the same target, including clinical trials to examine the activity of these compounds in solid and hematologic malignancies. In parallel, basic scientists continued exploring and trying to fill the gaps in the knowledge of the molecular biology of this pathway. At some point, biomarkers studies and clinical trials were developed without having a final clear portrait of the biology behind mTOR. Therefore, several unexpected and initially unexplainable results came back as a consequence of these studies.

Initial disappointment about preliminary clinical results decreased the excitement for targeting mTOR. It was later known that the mTOR pathway is almost a duality constituted by two complexes with different functions and many feedback loops, thus changing the original simplistic view of it. Now, a second generation of smarter compounds developed taking into account the latest biologic data is currently being developed. For one side, these compounds are able to inhibit both mTORC1 and mTORC2, and in the other side also incorporate activity against PI3K. Initial data from phase I clinical trials with these drugs have recently shown significant clinical activity, particularly in patients with deregulation of the PI3K-AKT-mTOR pathway.

Therefore, it is important to learn the lessons from the development of Rapamycin and Rapalogs. A complete understanding of the molecular biology of the pathway and its actors is needed in order to appropriately develop its targeted drugs and to correctly interpret the results from clinical studies. Finally, identification of biomarkers based on genetic, genomic, and systems biology approaches will allow defining what tumor subtypes may derive in a higher benefit with the use of mTOR inhibitors. These studies should be run in parallel to early clinical development trials, thus accelerating its implementation into phase III trials. In this way, biomarkers will be validated and ready to be approved simultaneously with drug indication.

Acknowledgments

Grant support: This work was supported in part by University of Michigan Comprehensive Cancer Center Core Support grant (NIH 5P30CA46592) and Michigan Institute for Clinical & Research Health (UL1RR024986).

References

  • 1.Samuels Y, Wang Z, Bardelli A, et al. High frequency of mutations of the PIK3CA gene in human cancers. Science. 2004 Apr 23;304(5670):554. doi: 10.1126/science.1096502. [DOI] [PubMed] [Google Scholar]
  • 2.Cantley LC. The phosphoinositide 3-kinase pathway. Science. 2002 May 31;296(5573):1655–1657. doi: 10.1126/science.296.5573.1655. [DOI] [PubMed] [Google Scholar]
  • 3.Yan L, Findlay GM, Jones R, Procter J, Cao Y, Lamb RF. Hyperactivation of mammalian target of rapamycin (mTOR) signaling by a gain-of-function mutant of the Rheb GTPase. J Biol Chem. 2006 Jul 21;281(29):19793–19797. doi: 10.1074/jbc.C600028200. [DOI] [PubMed] [Google Scholar]
  • 4.Wood LD, Parsons DW, Jones S, et al. The genomic landscapes of human breast and colorectal cancers. Science. 2007 Nov 16;318(5853):1108–1113. doi: 10.1126/science.1145720. [DOI] [PubMed] [Google Scholar]
  • 5.Guertin DA, Sabatini DM. The pharmacology of mTOR inhibition. Sci Signal. 2009;2(67):pe24. doi: 10.1126/scisignal.267pe24. [DOI] [PubMed] [Google Scholar]
  • 6.Inoki K, Ouyang H, Li Y, Guan KL. Signaling by target of rapamycin proteins in cell growth control. Microbiol Mol Biol Rev. 2005 Mar;69(1):79–100. doi: 10.1128/MMBR.69.1.79-100.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Chang YY, Juhasz G, Goraksha-Hicks P, et al. Nutrient-dependent regulation of autophagy through the target of rapamycin pathway. Biochem Soc Trans. 2009 Feb;37(Pt 1):232–236. doi: 10.1042/BST0370232. [DOI] [PubMed] [Google Scholar]
  • 8.Zhang HH, Huang J, Duvel K, et al. Insulin stimulates adipogenesis through the Akt-TSC2-mTORC1 pathway. PLoS One. 2009;4(7):e6189. doi: 10.1371/journal.pone.0006189. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Mayer C, Grummt I. Ribosome biogenesis and cell growth: mTOR coordinates transcription by all three classes of nuclear RNA polymerases. Oncogene. 2006 Oct 16;25(48):6384–6391. doi: 10.1038/sj.onc.1209883. [DOI] [PubMed] [Google Scholar]
  • 10.Wullschleger S, Loewith R, Hall MN. TOR signaling in growth and metabolism. Cell. 2006 Feb 10;124(3):471–484. doi: 10.1016/j.cell.2006.01.016. [DOI] [PubMed] [Google Scholar]
  • 11.Potter CJ, Pedraza LG, Xu T. Akt regulates growth by directly phosphorylating Tsc2. Nat Cell Biol. 2002 Sep;4(9):658–665. doi: 10.1038/ncb840. [DOI] [PubMed] [Google Scholar]
  • 12.Ma L, Chen Z, Erdjument-Bromage H, Tempst P, Pandolfi PP. Phosphorylation and functional inactivation of TSC2 by Erk implications for tuberous sclerosis and cancer pathogenesis. Cell. 2005 Apr 22;121(2):179–193. doi: 10.1016/j.cell.2005.02.031. [DOI] [PubMed] [Google Scholar]
  • 13.Manning BD, Tee AR, Logsdon MN, Blenis J, Cantley LC. Identification of the tuberous sclerosis complex-2 tumor suppressor gene product tuberin as a target of the phosphoinositide 3-kinase/akt pathway. Mol Cell. 2002 Jul;10(1):151–162. doi: 10.1016/s1097-2765(02)00568-3. [DOI] [PubMed] [Google Scholar]
  • 14.Tee AR, Manning BD, Roux PP, Cantley LC, Blenis J. Tuberous sclerosis complex gene products, Tuberin and Hamartin, control mTOR signaling by acting as a GTPase-activating protein complex toward Rheb. Curr Biol. 2003 Aug 5;13(15):1259–1268. doi: 10.1016/s0960-9822(03)00506-2. [DOI] [PubMed] [Google Scholar]
  • 15.van Slegtenhorst M, de Hoogt R, Hermans C, et al. Identification of the tuberous sclerosis gene TSC1 on chromosome 9q34. Science. 1997 Aug 8;277(5327):805–808. doi: 10.1126/science.277.5327.805. [DOI] [PubMed] [Google Scholar]
  • 16.Identification and characterization of the tuberous sclerosis gene on chromosome 16. Cell. 1993 Dec 31;75(7):1305–1315. doi: 10.1016/0092-8674(93)90618-z. [DOI] [PubMed] [Google Scholar]
  • 17.Sancak Y, Peterson TR, Shaul YD, et al. The Rag GTPases bind raptor and mediate amino acid signaling to mTORC1. Science. 2008 Jun 13;320(5882):1496–1501. doi: 10.1126/science.1157535. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Wang L, Lawrence JC, Jr, Sturgill TW, Harris TE. Mammalian target of rapamycin complex 1 (mTORC1) activity is associated with phosphorylation of raptor by mTOR. J Biol Chem. 2009 May 29;284(22):14693–14697. doi: 10.1074/jbc.C109.002907. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Peterson TR, Laplante M, Thoreen CC, et al. DEPTOR is an mTOR inhibitor frequently overexpressed in multiple myeloma cells and required for their survival. Cell. 2009 May 29;137(5):873–886. doi: 10.1016/j.cell.2009.03.046. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Wang L, Harris TE, Roth RA, Lawrence JC., Jr PRAS40 regulates mTORC1 kinase activity by functioning as a direct inhibitor of substrate binding. J Biol Chem. 2007 Jul 6;282(27):20036–20044. doi: 10.1074/jbc.M702376200. [DOI] [PubMed] [Google Scholar]
  • 21.Foster KG, Fingar DC. Mammalian target of rapamycin (mTOR): conducting the cellular signaling symphony. J Biol Chem. May 7;285(19):14071–14077. doi: 10.1074/jbc.R109.094003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Jacinto E, Loewith R, Schmidt A, et al. Mammalian TOR complex 2 controls the actin cytoskeleton and is rapamycin insensitive. Nat Cell Biol. 2004 Nov;6(11):1122–1128. doi: 10.1038/ncb1183. [DOI] [PubMed] [Google Scholar]
  • 23.Sarbassov DD, Ali SM, Sengupta S, et al. Prolonged rapamycin treatment inhibits mTORC2 assembly and Akt/PKB. Mol Cell. 2006 Apr 21;22(2):159–168. doi: 10.1016/j.molcel.2006.03.029. [DOI] [PubMed] [Google Scholar]
  • 24.Sarbassov DD, Guertin DA, Ali SM, Sabatini DM. Phosphorylation and regulation of Akt/PKB by the rictor-mTOR complex. Science. 2005 Feb 18;307(5712):1098–1101. doi: 10.1126/science.1106148. [DOI] [PubMed] [Google Scholar]
  • 25.Guertin DA, Stevens DM, Thoreen CC, et al. Ablation in mice of the mTORC components raptor, rictor, or mLST8 reveals that mTORC2 is required for signaling to Akt-FOXO and PKCalpha, but not S6K1. Dev Cell. 2006 Dec;11(6):859–871. doi: 10.1016/j.devcel.2006.10.007. [DOI] [PubMed] [Google Scholar]
  • 26.Ikenoue T, Inoki K, Yang Q, Zhou X, Guan KL. Essential function of TORC2 in PKC and Akt turn motif phosphorylation, maturation and signalling. Embo J. 2008 Jul 23;27(14):1919–1931. doi: 10.1038/emboj.2008.119. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Huang J, Dibble CC, Matsuzaki M, Manning BD. The TSC1-TSC2 complex is required for proper activation of mTOR complex 2. Mol Cell Biol. 2008 Jun;28(12):4104–4115. doi: 10.1128/MCB.00289-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Fernandez DR, Telarico T, Bonilla E, et al. Activation of mammalian target of rapamycin controls the loss of TCRzeta in lupus T cells through HRES-1/Rab4-regulated lysosomal degradation. J Immunol. 2009 Feb 15;182(4):2063–2073. doi: 10.4049/jimmunol.0803600. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Haxhinasto S, Mathis D, Benoist C. The AKT-mTOR axis regulates de novo differentiation of CD4+Foxp3+ cells. J Exp Med. 2008 Mar 17;205(3):565–574. doi: 10.1084/jem.20071477. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Harrington LS, Findlay GM, Gray A, et al. The TSC1-2 tumor suppressor controls insulin-PI3K signaling via regulation of IRS proteins. J Cell Biol. 2004 Jul 19;166(2):213–223. doi: 10.1083/jcb.200403069. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Wan X, Harkavy B, Shen N, Grohar P, Helman LJ. Rapamycin induces feedback activation of Akt signaling through an IGF-1R-dependent mechanism. Oncogene. 2007 Mar 22;26(13):1932–1940. doi: 10.1038/sj.onc.1209990. [DOI] [PubMed] [Google Scholar]
  • 32.O'Reilly KE, Rojo F, She QB, et al. mTOR inhibition induces upstream receptor tyrosine kinase signaling and activates Akt. Cancer Res. 2006 Feb 1;66(3):1500–1508. doi: 10.1158/0008-5472.CAN-05-2925. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Tabernero J, Rojo F, Calvo E, et al. Dose- and schedule-dependent inhibition of the mammalian target of rapamycin pathway with everolimus: a phase I tumor pharmacodynamic study in patients with advanced solid tumors. J Clin Oncol. 2008 Apr 1;26(10):1603–1610. doi: 10.1200/JCO.2007.14.5482. [DOI] [PubMed] [Google Scholar]
  • 34.Di Cosimo S, Bendell JC, Cervantes-Ruiperez A, et al. A phase I study of the oral mTOR inhibitor ridaforolimus (RIDA) in combination with the IGF-1R antibody dalotozumab (DALO) in patients (pts) with advanced solid tumors. J Clin Oncol (Meeting Abstracts) 2010 May 20;28(15_suppl):3008. [Google Scholar]
  • 35.Naing A, LoRusso P, Gupta S, et al. Dual inhibition of IGFR and mTOR pathways. J Clin Oncol (Meeting Abstracts) 2010 May 20;28(15_suppl):3007. [Google Scholar]
  • 36.Quek RH, Morgan JA, Shapiro G, et al. Combination mTOR+IGF-IR inhibition: Phase I trial of everolimus and CP-751871 in patients (pts) with advanced sarcomas and other solid tumors. J Clin Oncol (Meeting Abstracts) 2010 May 20;28(15_suppl):10002. [Google Scholar]
  • 37.Carracedo A, Ma L, Teruya-Feldstein J, et al. Inhibition of mTORC1 leads to MAPK pathway activation through a PI3K-dependent feedback loop in human cancer. J Clin Invest. 2008 Sep;118(9):3065–3074. doi: 10.1172/JCI34739. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Iadevaia S, Lu Y, Morales FC, Mills GB, Ram PT. Identification of optimal drug combinations targeting cellular networks: integrating phospho-proteomics and computational network analysis. Cancer Res. Sep 1;70(17):6704–6714. doi: 10.1158/0008-5472.CAN-10-0460. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Kahan BD. Efficacy of sirolimus compared with azathioprine for reduction of acute renal allograft rejection: a randomised multicentre study. The Rapamune US Study Group. Lancet. 2000 Jul 15;356(9225):194–202. doi: 10.1016/s0140-6736(00)02480-6. [DOI] [PubMed] [Google Scholar]
  • 40.Choi J, Chen J, Schreiber SL, Clardy J. Structure of the FKBP12-rapamycin complex interacting with the binding domain of human FRAP. Science. 1996 Jul 12;273(5272):239–242. doi: 10.1126/science.273.5272.239. [DOI] [PubMed] [Google Scholar]
  • 41.Sarbassov DD, Ali SM, Kim DH, et al. Rictor, a novel binding partner of mTOR, defines a rapamycin-insensitive and raptor-independent pathway that regulates the cytoskeleton. Curr Biol. 2004 Jul 27;14(14):1296–1302. doi: 10.1016/j.cub.2004.06.054. [DOI] [PubMed] [Google Scholar]
  • 42.Mothe-Satney I, Brunn GJ, McMahon LP, Capaldo CT, Abraham RT, Lawrence JC., Jr Mammalian target of rapamycin-dependent phosphorylation of PHAS-I in four (S/T)P sites detected by phospho-specific antibodies. J Biol Chem. 2000 Oct 27;275(43):33836–33843. doi: 10.1074/jbc.M006005200. [DOI] [PubMed] [Google Scholar]
  • 43.Gingras AC, Gygi SP, Raught B, et al. Regulation of 4E-BP1 phosphorylation: a novel two-step mechanism. Genes Dev. 1999 Jun 1;13(11):1422–1437. doi: 10.1101/gad.13.11.1422. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Beretta L, Gingras AC, Svitkin YV, Hall MN, Sonenberg N. Rapamycin blocks the phosphorylation of 4E-BP1 and inhibits cap-dependent initiation of translation. Embo J. 1996 Feb 1;15(3):658–664. [PMC free article] [PubMed] [Google Scholar]
  • 45.Choo AY, Yoon SO, Kim SG, Roux PP, Blenis J. Rapamycin differentially inhibits S6Ks and 4E-BP1 to mediate cell-type-specific repression of mRNA translation. Proc Natl Acad Sci U S A. 2008 Nov 11;105(45):17414–17419. doi: 10.1073/pnas.0809136105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.McMahon LP, Choi KM, Lin TA, Abraham RT, Lawrence JC., Jr The rapamycin-binding domain governs substrate selectivity by the mammalian target of rapamycin. Mol Cell Biol. 2002 Nov;22(21):7428–7438. doi: 10.1128/MCB.22.21.7428-7438.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Mothe-Satney I, Yang D, Fadden P, Haystead TA, Lawrence JC., Jr Multiple mechanisms control phosphorylation of PHAS-I in five (S/T)P sites that govern translational repression. Mol Cell Biol. 2000 May;20(10):3558–3567. doi: 10.1128/mcb.20.10.3558-3567.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Fuchs CS, Tabernero JM, Hwang J, et al. Multicenter phase II study of RAD001 in patients with chemotherapy-refractory metastatic colorectal cancer (mCRC); ASCO Gastrointestinal Cancers Symposium; San Francisco, CA. 2009. [Google Scholar]
  • 49.Yee KW, Zeng Z, Konopleva M, et al. Phase I/II study of the mammalian target of rapamycin inhibitor everolimus (RAD001) in patients with relapsed or refractory hematologic malignancies. Clin Cancer Res. 2006 Sep 1;12(17):5165–5173. doi: 10.1158/1078-0432.CCR-06-0764. [DOI] [PubMed] [Google Scholar]
  • 50.Yao JC, Lombard-Bohas C, Baudin E, et al. Daily oral everolimus activity in patients with metastatic pancreatic neuroendocrine tumors after failure of cytotoxic chemotherapy: a phase II trial. J Clin Oncol. Jan 1;28(1):69–76. doi: 10.1200/JCO.2009.24.2669. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Soria JC, Shepherd FA, Douillard JY, et al. Efficacy of everolimus (RAD001) in patients with advanced NSCLC previously treated with chemotherapy alone or with chemotherapy and EGFR inhibitors. Ann Oncol. 2009 Oct;20(10):1674–1681. doi: 10.1093/annonc/mdp060. [DOI] [PubMed] [Google Scholar]
  • 52.Galanis E, Buckner JC, Maurer MJ, et al. Phase II trial of temsirolimus (CCI-779) in recurrent glioblastoma multiforme: a North Central Cancer Treatment Group Study. J Clin Oncol. 2005 Aug 10;23(23):5294–5304. doi: 10.1200/JCO.2005.23.622. [DOI] [PubMed] [Google Scholar]
  • 53.Hudes G, Carducci M, Tomczak P, et al. Temsirolimus, interferon alfa, or both for advanced renal-cell carcinoma. N Engl J Med. 2007 May 31;356(22):2271–2281. doi: 10.1056/NEJMoa066838. [DOI] [PubMed] [Google Scholar]
  • 54.Motzer RJ, Escudier B, Oudard S, et al. Efficacy of everolimus in advanced renal cell carcinoma: a double-blind, randomised, placebo-controlled phase III trial. Lancet. 2008 Aug 9;372(9637):449–456. doi: 10.1016/S0140-6736(08)61039-9. [DOI] [PubMed] [Google Scholar]
  • 55.Hess G, Herbrecht R, Romaguera J, et al. Phase III study to evaluate temsirolimus compared with investigator's choice therapy for the treatment of relapsed or refractory mantle cell lymphoma. J Clin Oncol. 2009 Aug 10;27(23):3822–3829. doi: 10.1200/JCO.2008.20.7977. [DOI] [PubMed] [Google Scholar]
  • 56.Janku F, Garrido-Laguna I, Hong DS, et al. PIK3CA mutations in patients with advanced cancers treated in phase I clinical trials; Presented at: 2010 ASCO Annual Meeting; Chicago, IL, USA. Jun, Abstract B134. [Google Scholar]
  • 57.Garrido-Laguna I, Janku F, Tsimberidou A, et al. Phosphatase and tensin homologue (PTEN) loss and response to phase I trials targeting PI3K/AKT/mTOR pathway in patients with advanced cancer. J Clin Oncol (Meeting Abstracts) 2010 May 20;28(15_suppl):e13018. [Google Scholar]
  • 58.Ogino S, Nosho K, Kirkner GJ, et al. PIK3CA mutation is associated with poor prognosis among patients with curatively resected colon cancer. J Clin Oncol. 2009 Mar 20;27(9):1477–1484. doi: 10.1200/JCO.2008.18.6544. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.De Roock W, Claes B, Bernasconi D, et al. Effects of KRAS, BRAF, NRAS, and PIK3CA mutations on the efficacy of cetuximab plus chemotherapy in chemotherapy-refractory metastatic colorectal cancer: a retrospective consortium analysis. Lancet Oncol. Aug;11(8):753–762. doi: 10.1016/S1470-2045(10)70130-3. [DOI] [PubMed] [Google Scholar]
  • 60.Vilar E, Mukherjee B, Kuick R, et al. Gene expression patterns in mismatch repair-deficient colorectal cancers highlight the potential therapeutic role of inhibitors of the phosphatidylinositol 3-kinase-AKT-mammalian target of rapamycin pathway. Clin Cancer Res. 2009 Apr 15;15(8):2829–2839. doi: 10.1158/1078-0432.CCR-08-2432. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Courtney KD, Corcoran RB, Engelman JA. The PI3K Pathway As Drug Target in Human Cancer. J Clin Oncol. Feb 20;28(6):1075–1083. doi: 10.1200/JCO.2009.25.3641. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Maira SM, Stauffer F, Brueggen J, et al. Identification and characterization of NVP-BEZ235, a new orally available dual phosphatidylinositol 3-kinase/mammalian target of rapamycin inhibitor with potent in vivo antitumor activity. Mol Cancer Ther. 2008 Jul;7(7):1851–1863. doi: 10.1158/1535-7163.MCT-08-0017. [DOI] [PubMed] [Google Scholar]
  • 63.Serra V, Markman B, Scaltriti M, et al. NVP-BEZ235, a dual PI3K/mTOR inhibitor, prevents PI3K signaling and inhibits the growth of cancer cells with activating PI3K mutations. Cancer Res. 2008 Oct 1;68(19):8022–8030. doi: 10.1158/0008-5472.CAN-08-1385. [DOI] [PubMed] [Google Scholar]
  • 64.Burris H, Rodon J, Sharma S, et al. First-in-human phase I study of the oral PI3K inhibitor BEZ235 in patients (pts) with advanced solid tumors. J Clin Oncol (Meeting Abstracts) 2010 May 20;28(15_suppl):3005. [Google Scholar]
  • 65.Brana I, LoRusso P, Baselga J, et al. A phase I dose-escalation study of the safety, pharmacokinetics (PK), and pharmacodynamics of XL765 (SAR245409), a PI3K/TORC1/TORC2 inhibitor administered orally to patients (pts) with advanced malignancies. J Clin Oncol (Meeting Abstracts) 2010 May 20;28(15_suppl):3030. [Google Scholar]
  • 66.Jessen K, Wang S, Kessler L, et al. INK128 is a potent and selective TORC1/2 inhibitor with broad oral antitumor activity. Mol Cancer Ther. 2009;8(12 Suppl):B148. [Google Scholar]
  • 67.Feldman ME, Apsel B, Uotila A, et al. Active-site inhibitors of mTOR target rapamycin-resistant outputs of mTORC1 and mTORC2. PLoS Biol. 2009 Feb 10;7(2):e38. doi: 10.1371/journal.pbio.1000038. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Hsieh AC, Costa M, Zollo O, et al. Genetic dissection of the oncogenic mTOR pathway reveals druggable addiction to translational control via 4EBP-eIF4E. Cancer Cell. Mar 16;17(3):249–261. doi: 10.1016/j.ccr.2010.01.021. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Thoreen CC, Kang SA, Chang JW, et al. An ATP-competitive mammalian target of rapamycin inhibitor reveals rapamycin-resistant functions of mTORC1. J Biol Chem. 2009 Mar 20;284(12):8023–8032. doi: 10.1074/jbc.M900301200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Xue Q, Hopkins B, Perruzzi C, Udayakumar D, Sherris D, Benjamin LE. Palomid 529, a novel small-molecule drug, is a TORC1/TORC2 inhibitor that reduces tumor growth, tumor angiogenesis, and vascular permeability. Cancer Res. 2008 Nov 15;68(22):9551–9557. doi: 10.1158/0008-5472.CAN-08-2058. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Chresta CM, Davies BR, Hickson I, et al. AZD8055 is a potent, selective, and orally bioavailable ATP-competitive mammalian target of rapamycin kinase inhibitor with in vitro and in vivo antitumor activity. Cancer Res. Jan 1;70(1):288–298. doi: 10.1158/0008-5472.CAN-09-1751. [DOI] [PubMed] [Google Scholar]
  • 72.Garcia-Martinez JM, Moran J, Clarke RG, et al. Ku-0063794 is a specific inhibitor of the mammalian target of rapamycin (mTOR) Biochem J. 2009 Jul 1;421(1):29–42. doi: 10.1042/BJ20090489. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.Yu K, Toral-Barza L, Shi C, et al. Biochemical, cellular, and in vivo activity of novel ATP-competitive and selective inhibitors of the mammalian target of rapamycin. Cancer Res. 2009 Aug 1;69(15):6232–6240. doi: 10.1158/0008-5472.CAN-09-0299. [DOI] [PubMed] [Google Scholar]
  • 74.Tan DS, Dumez H, Olmos D, et al. First-in-human phase I study exploring three schedules of OSI-027, a novel small molecule TORC1/TORC2 inhibitor, in patients with advanced solid tumors and lymphoma. J Clin Oncol (Meeting Abstracts) 2010 May 20;28(15_suppl):3006. [Google Scholar]

RESOURCES