Download as pdf or txt
Download as pdf or txt
You are on page 1of 20

Seminars in Cancer Biology xxx (xxxx) xxx–xxx

Contents lists available at ScienceDirect

Seminars in Cancer Biology


journal homepage: www.elsevier.com/locate/semcancer

Review

mTOR: Role in cancer, metastasis and drug resistance


Avaniyapuram Kannan Murugan
Department of Molecular Oncology, King Faisal Specialist Hospital & Research Centre, PO Box 3354, Research Center (MBC 03), Riyadh, 11211, Saudi Arabia

A R T I C LE I N FO A B S T R A C T

Keywords: Mammalian target of rapamycin (mTOR) is a serine/threonine kinase that gets inputs from the amino acids,
Cancer nutrients, growth factor, and environmental cues to regulate varieties of fundamental cellular processes which
mTOR include protein synthesis, growth, metabolism, aging, regeneration, autophagy, etc. The mTOR is frequently
Mutation deregulated in human cancer and activating somatic mutations of mTOR were recently identified in several types
Oncogene
of human cancer and hence mTOR is therapeutically targeted. mTOR inhibitors were commonly used as im-
Molecular target
munosuppressors and currently, it is approved for the treatment of human malignancies. This review briefly
Resistance
Metastasis focuses on the structure and biological functions of mTOR. It extensively discusses the genetic deregulation of
Precision medicine mTOR including amplifications and somatic mutations, mTOR-mediated cell growth promoting signaling,
Inhibitors therapeutic targeting of mTOR and the mechanisms of resistance, the role of mTOR in precision medicine and
other recent advances in further understanding the role of mTOR in cancer.

1. Introduction including lysosome biogenesis and autophagy and it is a master reg-


ulator of cellular metabolism. The mTOR integrates these wide ranges
Rapamycin is a natural macrolide with antifungal, antiproliferative, of environmental cues-induced anabolic processes to modulate meta-
and immunosuppressive properties which was originally discovered bolic pathways for cell proliferation, growth, and metabolism. The
from the bacteria Streptomyces hygroscopicus in soil samples by Surendra mTOR also plays a central role in the regulation of autophagy [7].
Nath Sehgal and his group when they traveled to Easter Island (it was Cancer is a multifactorial and complex disease and known to be the
known as Rapa Nui by locals) in 1964 [1]. Three decades later, two most leading causes of human deaths worldwide. Phosphatidylinositol
genes (TOR1 and TOR2) were identified as a direct target of the im- 3-kinase (PI3K)-Akt-mTOR pathway is one of the most deregulated
munosuppressant, rapamycin toxicity in yeast cells by MN Hall and his signaling pathways in human cancer. The mTOR is considered as a
colleagues in 1991 [2]. After three years, the mTOR protein was dis- master regulator of this signaling pathway and recent findings reported
covered as a direct target of rapamycin-FKBP12 (12 kDa FK506-binding that the mTOR has a pivotal role in human cancer when it is activated.
protein) complex in mammalian cells, called mammalian target of ra- Further, mTOR signaling is often reported to be hyperactivated in the
pamycin (mTOR). This discovery came independently from three dif- majority of human cancers particularly implicated in the cell transfor-
ferent labs, SL Schreiber and SH Snyder’s lab demonstrated in mid-1994 mation, growth, survival, etc [7–9].
while RT Abraham lab reported in early 1995 [3–5]. These studies
confirmed the previously-identified yeast TOR to be the homolog of 2. Structure of mTOR
mTOR. Since then pioneering discoveries in this area were made by
many labs from all over the world. In particular, the molecular and mTOR is a 288.892 kDa serine/threonine protein kinase that com-
biochemical approach uncovered various binding partners, down- prises 2550 amino acids (including stop codon) which are encoded by
stream/upstream effectors, signaling mediators of mTOR and its role in 7650 nucleotides and the mTOR gene is located in chromosome 1
the biological process which was recently reviewed by DM Sabatini [6]. (1p36.22). The mTOR basically belongs to PI3K-related kinases (PIKK)
In a normal cell, mTOR gets various environmental stimuli from amino and most of the PIKK members have conserved domain architecture
acids, growth factors, oxygen, stress, redox sensors, and energy. In re- [10,11]. The mTOR protein molecule has a large N-terminal α-solenoid
sponse to these diverse environmental cues, the active mTOR promotes (HEAT repeats), a FAT (FRAP, ATM, TRRAP) domain, KD (a protein
cellular anabolism to generate various macromolecules such as nucleic kinase domain), and an FRB (FKBP12-rapamycin binding) domain and
acids, proteins, lipids which build cellular biomass, ribosome biogen- a FATC (C-terminal FAT) domain. The mTOR plays an essential role as a
esis, and protein translation. mTOR also blocks catabolic processes core component of two functionally different multi-subunit protein

E-mail address: akmurugan@gmail.com.

https://doi.org/10.1016/j.semcancer.2019.07.003
Received 16 March 2019; Received in revised form 14 June 2019; Accepted 3 July 2019
1044-579X/ © 2019 Elsevier Ltd. All rights reserved.

Please cite this article as: Avaniyapuram Kannan Murugan, Seminars in Cancer Biology, https://doi.org/10.1016/j.semcancer.2019.07.003
A.K. Murugan Seminars in Cancer Biology xxx (xxxx) xxx–xxx

Fig. 1. Functions of mTORC1 and mTORC2 and their cellular signaling. A. mTORC1 and its signaling pathway. The capsulated diagram shows the schematic picture of
mTOR protein and its various multiprotein subunits of mTORC1 and related binding sites on the mTOR. The lower panel shows the 3D ribbon view of the cryo-EM
structure of human mTOR complex1 with 4.4-Aº resolution (PDB ID: 5H64). Arrow indicates mTORC1-mediated various vital cellular function-based pathways and
their signaling proteins. B. mTORC2 and its signaling pathway. The capsulated diagram shows the schematic picture of mTOR protein and its various multiprotein
subunits of mTORC2 and related binding sites on the mTOR. The arrow indicates the mTORC2-mediated vital cellular function-based pathways and their signaling
proteins. Lower panel shows the 3D ribbon view of the cryo-EM structure of human mTOR complex 2 with 4.9-Aº resolution (PDB ID: 5ZCS).

2
A.K. Murugan Seminars in Cancer Biology xxx (xxxx) xxx–xxx

complexes named as mTORC1 (mTOR complex 1) and mTORC2 (mTOR 3. Upstream and downstream signaling of mTORC1 and mTORC2
complex 2) [12,13].
Growth factor-mediated RTKs (receptor tyrosine kinases)/PI3K/Akt
signaling pathway is an important upstream signaling pathway of the
2.1. mTORC1 mTOR protein molecule [28]. In response to various extracellular sti-
muli such as growth factors, nutrients, and amino acids, mTOR strongly
As illustrated in Fig. 1A, in addition to the mTOR - a core protein, associates to various protein molecules and to form two distinct mul-
the mTORC1 contains Raptor (regulatory protein associated with tiprotein molecular complexes such as mTORC1 and mTORC2, and
mTOR) and mLST8 (mammalian lethal with sec 13 protein 8, otherwise regulates various growth signals by directly phosphorylating the im-
called as GβL). Apart from these three proteins, mTORC1 also consists mediate substrates [29]. In a normal cell, various RTKs, such as epi-
of an additional two inhibitory protein subunits including PRAS40 dermal growth factor receptor (EGFR), insulin receptor (IR), and G-
(proline-rich Akt substrate of 40 kDa) and DEPTOR (DEP domain con- protein coupled receptor (GPCR) avails extracellular stimuli from var-
taining mTOR interacting protein) [13]. The Raptor helps to recruit ious growth factors. Upon the growth factor stimulation, the RTKs
substrate to mTORC1 via binding to the TOR signaling motif and also trigger input signals to PI3K. Generally, PI3K activity is tightly con-
essential for the subcellular localization of mTORC1. The mLST8 sta- trolled to a basal level in normal condition. Then, the PI3K catalyzes the
bilizes the kinase activation loop by associating with the kinase domain production of phosphatidylinositol 3,4,5-triphosphate (PIP3) by phos-
of mTORC1. When the FKBP12-rapamycin complex binds to the FRB phorylating phosphatidylinositol 4,5-bisphosphate (PIP2). This activity
domain, the substrate entry could be blocked that results in inhibition is antagonized by PTEN (phosphatase and tensin homolog deleted on
of mTORC1 kinase activity [14]. Recent three-dimensional structural chromosome 10), the tumor suppressor which reverses the PIP3 to PIP2
studies have revealed various structural inputs, assembly, and functions [30]. Accumulation of the PIP3, the intracellular second messenger, at
of mTORC1. The cryo-EM structures of mTORC1 and TORC1 found this the plasma membrane recruits the protein serine-threonine kinases, Akt
complex as a dimer with a “lozenge” shape [14–17]. A recent report on (a cellular homolog of the murine thymoma virus Akt8 oncogene),
the atomic resolution structure of mTORC1 showed many structural SGK1 (serum and glucocorticoid-induced kinase 1), and PDK1/2 (3-
insights of Rheb-mediated activation of mTORC1 and inhibition by phosphatidylinositol dependent kinase-1 and 2), the PH (pleckstrin-
PRAS40 [17]. homology) domain-containing proteins to the membrane [31] where,
PDK1 phosphorylates Akt at threonine 308 that results in the activation
of Akt and its signaling pathway [32]. Further, increased protein
2.2. mTORC2 synthesis and cell proliferation signals from Akt reach the GTPase-ac-
tivating protein TSC1/2 (tuberous sclerosis complex1/2) and active Akt
The mTORC2 constitutes an mTOR, mLST8, and Rictor (rapamycin- directly phosphorylates that results in inactivation of TSC2 (known also
insensitive companion of mTOR). The Rictor is a unique component of as tuberin), which makes an enhanced Rheb (Ras homolog enriched in
the mTORC2 while the Raptor is a part of mTORC1 [18]. In addition to brain) activity [33]. Activated Rheb (GTP-bound) directly switches on
these basic protein subunits, mTORC2 has DEPTOR, mSin1 (stress-ac- the mTOR thus mTOR assembles its components of mTORC1. The
tivated map kinase interacting protein 1), a regulatory subunit and mobilized mTORC1 then relays signaling by phosphorylating two key
Protor1/2. Although the PH domain of mSin1 inhibits the kinase ac- substrate protein molecules, p70S6K, and 4EBP1 (4E-binding protein
tivity of mTORC2, availability of the PIP3 (phosphatidylinositol 3,4,5- 1), resulting in activation of p70S6K and inactivation of 4EBP1 [34].
triphosphate) releases the auto-inhibition [19]. Nevertheless, the me- This signaling results in enhanced S6 kinase activity that promotes
chanism is not clearly understood how the PH domain of mSin1 inhibits translation, and protein synthesis. The mTORC1 is reported to regulate
the mTORC2 activity. The mTORC2 is insensitive to rapamycin many downstream effectors including the transcriptional activity of
[12,20,21,7] however it is not known why the FKBP12-rapamycin various protein molecules such as ATF4 (activating transcription factor
complexes do not abrogate the mTORC2 activity. Further, a prolonged 4), SREBP (sterol regulatory element-binding proteins), HIF1α (hy-
rapamycin treatment was shown to inhibit the mTORC2 signaling that poxia-inducible factor 1α), PGC1α (peroxisome proliferator-activated
may be because of the binding of rapamycin-FKBP12 complex to “free receptor gamma coactivator 1-alpha), and TFEB (Transcription factor
mTOR” (mTOR, unbound with Raptor and Rictor), could cause deple- EB), which promotes metabolic homeostasis, lipid synthesis, glycolysis,
tion of the “free mTOR” as all the synthesized mTOR molecules are mitochondrial biogenesis, lysosome biogenesis, respectively [35]. Fur-
gradually trapped by rapamycin-FKBP12 complex which prohibits ther, mTORC1-mediated signaling to HIF1α, LIPIN1, p70S6K, and ATF4
binding of mTOR with Rictor and this process inhibits assembly of enhance glucose metabolism, lipid synthesis, nucleotide synthesis, re-
mTORC2. Particularly, more than 24 h treatments make this saturation spectively. Moreover, various inhibitory signaling to ERK5 (extra-
of rapamycin-FKB12 and newly-synthesized mTOR binding, which re- cellular-signal-regulated kinase 5), TFEB, ULK1 (unc-51 like autophagy
sults in inhibition of mTORC2-mediated Akt signaling. Still, this effect activating kinase 1), and UVRAG (UV radiation resistance associated
has been observed as a differential effect among various cell types that gene) promotes proteasome assembly, lysosome biogenesis, and in-
may be due to the cross-talk between the two mTOR complexes [22,23]. hibits autophagy, respectively. The RAS is an important member of the
Moreover, it has been shown that the downstream effector of mTORC1, RTK-mediated PI3K/Akt and MAPK (mitogen-activated protein kinase)
p70S6K (p70 ribosomal protein S6 kinase 1) could phosphorylate Rictor signaling. Growth factor-mediated activation of RTKs enhances RAS
to inhibit mTORC2 [24]. Structural analyses that include the negative signaling thus result in activation of PI3K/Akt and MAPK pathway. In
stain and cryo-EM (electron microscopy) structures of yeast TORC2 addition, accumulation of PIP3 was also shown to stimulate RAS ac-
discovered various conformational patterns and subunit architectures tivity [36–38].
of the TORC2 (TOR complex2) [25,26]. More recently, a cryo-EM On the other hand, an activated mTOR associates with its compo-
structure of human mTOR complex was reported that displayed nents of protein subunits and forms mTORC2. Active mTORC2 phos-
mTORC2 as a dimerized form with hollow rhombohedral fold and ex- phorylates Akt at the hydrophobic motif (HM) site serine 473 and turn
hibited a 2-fold symmetry which is compact than mTORC1. Further, the motif (TM) site threonine 450 when the mTORC2 is activated by growth
study also revealed the mechanism for rapamycin insensitivity of factors and associated with translating ribosomes, respectively. Besides,
mTORC2 showing how Rictor and mSin1 jointly make steric hindrance the mTORC2 also phosphorylates SGK at HM site serine 422. In addi-
for inhibiting the binding of FKBP12-rapamycin to mTOR (Fig. 1B) tion, mTORC2 phosphorylates cPKC (conventional PKC) and nPKC
[27]. (novel PKC) isotypes. It phosphorylates at the TM site threonine 638/
641 of PKCα/βII and at the HM site serine 657/660 of all cPKC) and

3
A.K. Murugan Seminars in Cancer Biology xxx (xxxx) xxx–xxx

some nPKC also requires mTORC2. These phosphorylations of the key in human samples. Researches on human genetics identified several de
substrate protein molecules critically modulate cellular apoptosis, au- novo heterozygous mutations in the mTOR gene in Smith-Kingsmore
tophagy, glucose metabolism, cytoskeleton rearrangement and motility syndrome. mTOR mutations were detected in many different functional
[39]. Moreover, Akt either directly or indirectly regulates the tran- codons, for example, C1483F 49, E1799K [50–53] F1888C and M2327I
scription factors that control many vital cellular processes. It has been [54] and the majority of them were harbored in the FAT domain.
demonstrated that mTOR inhibition could induce IRS-1 (insulin re- Structural analysis based on molecular modeling identified that the
ceptor substrate-1) expression and abolishes feedback inhibition of the amino acid residue E1799 of FAT domain binds on the kinase domain
pathway, thus results in the activation of Akt. Inhibition of IGF-I re- that results in negative regulation of mTOR. Genetic mutations which
ceptor suppressed rapamycin-induced Akt activation and tumor cells alter this critical amino acid residue (E1799) causes destabilization of
become sensitive to mTOR inhibition suggests a rationale for combi- the mTOR protein molecule thus the protein could switch to an active
nation therapy [40]. The mTOR was delineated to function as a tyrosine state as the negative regulation machinery is functionally impaired
protein kinase because it could activate IRs (insulin receptors) and IGF- [51]. Consistent with this observation, subsequent functional analyses
IR1 (insulin-like growth factor receptor 1) when it is a core component of all these mutations exhibited a gain-of-function effect [49–51,54,53].
of mTORC2 [41]. In order to turn on the downstream signaling, the Akt These observations suggested that mutation in these amino acid re-
directly phosphorylates FOXO proteins (forkhead box proteins) for sidues may critically impair the protein conformation that may activate
downregulating the transcriptional activity whenever the transcrip- the mTOR activity thus resulting in enhanced protein synthesis during
tional activity is enhanced [42]. Akt indirectly regulates p53 in- the early stages of brain development and this may likely to ignite an
dependent of mTOR through MDM2 (murine double minute) and hy- inflammatory reaction in the brain.
peractivates c-Myc and β-catenin, thereby inhibits GSK3β and this In addition to whole-exome sequencing, deep sequencing, conven-
cascade provides an advantage for cell proliferation and regulating tional PCR amplification-based Sanger sequencing of mTOR gene in
apoptosis [35.43]. Recently, BioID (proximity-dependent biotin la- diverse neurological diseases found mTOR mutations in 15.6% (12/77)
beling)-mediated analysis identified known interactors for wild-type of patients in focal cortical dysplasia, type II (FCD type II). These mTOR
and mutant RAS (H, K, and NRAS) which includes RAF and PI3K and a mutations were found to be somatic origin as the brain tissue-derived
set of 130 unknown novel proximal proteins. Further CRISPR-based mutations were not detected in the blood samples of the same patients.
screen of these 130 proteins for RAS dependence identified mTOR. The Many mTOR mutations such as W1456G 55, L1460P 56, L2427P [57],
oncogenic RAS was shown to bind directly with two components of S2215Y [56,54,52] were identified in FCD type II and those mutations
mTORC2 which includes mTOR and MAPKAP1 and that resulted in were mostly distributed both in the FAT and kinase domain of mTOR.
enhanced mTORC2 kinase activity at the cell membrane. The mTORC2 Inoculation of three of these mutations into the HEK293T cells ex-
promoted RAS-induced pro-proliferative cascade of the cell and in- hibited constitutive activation of mTOR kinase activity while compared
hibition of RAS-mTORC2 interaction impaired RAS-mediated tumor- to that of wild-type control cells. Treatment of mice carrying these
igenesis in vivo [44]. Identification of mTORC2 as a direct RAS effector mutants with rapamycin inhibited cytomegalic neuron and seizures
paves an alternative way for RASopathy for the undruggable oncogenic [57]. Consistent with these findings, another study also found de novo
RAS mutants in human cancer that could be achieved by mTOR- somatic missense mTOR mutations in 46% (6/13) of FCD type IIb [56].
mediated inhibitors mainly in combination with BRAF/ERK inhibitors Further, functional studies on these mutations observed hyperactivation
as RAS also activates the MAPK pathway (Fig. 1). of mTOR resulting in phosphorylation of its immediate substrate,
4EBP1.
4. Biological functions of mTOR
5.2. mTOR and human cancer
The mTOR is a multifaceted protein molecule and is a catalytic core
component of Raptor-mTOR (mTORC1) and Rictor-mTOR (mTORC2) Activated mTOR signaling is a key for growth promotion and is
complexes known to function both as a serine/threonine kinase and implicated in human cancer, playing various roles in cell survival, cy-
tyrosine kinase, respectively [41]. The mTOR is demonstrated to play a toskeleton rearrangement, invasion, metastasis, anti-apoptosis and in-
key role in early embryonic development particularly in pluripotency hibition of autophagy (Fig. 2). The mTOR signaling pathway is altered
and T-cell transdifferentiation [45,46]. As seen in Fig. 1, the mTORC1 by a variety of mechanisms. Firstly, as a mediator of the PI3K signaling
plays a central role in metabolic homeostasis, protein, and lipid pathway, it gets hyperactivated by upstream RTKs/PI3K/Akt signaling.
synthesis, glycolysis, mitochondrial biogenesis, lysosome biogenesis by Secondly, deregulation of closely related and its immediate signal
regulating various transcription factors and directly controls translation transducers of mTOR such as TSC1/2, Rheb and p70S6K/4EBP1 acti-
[39,7,9]. It modulates glucose metabolism and nucleotide synthesis vate mTOR signaling. Thirdly, recent research revealed that mTOR itself
through metabolic pathway proteins. It also regulates the proteasome was found to be genetically deregulated in human cancers by amplifi-
assembly and autophagy [47]. The mTORC2 mainly regulates cell cation, downregulation, and somatic mutation. These genetic altera-
proliferation, survival, cytoskeletal rearrangement, cell migration, tions cause hyperactivation of mTOR that results in activation of its
glucose metabolism, and apoptosis [39,7,9,48]. signaling pathway and these mechanisms are detailed below.

5. Deregulated mTOR in human diseases 6. Mechanisms of mTOR activation in human cancer

The mTOR signaling has a broader impact on basic vital cellular 6.1. Activation of mTOR by upstream signaling molecules in human cancer
functions and its deregulated signaling alters normal physiological
functions that result in pathogenesis in humans. Moreover, deregulated The RTKs-mediated PI3K activation phosphorylates and switches
mTOR is displayed to be implicated in human growth and metabolic PIP2 to PIP3 (directly antagonized by the tumor suppressor, PTEN by
diseases including neuronal degeneration, obesity, type 2 diabetes, and dephosphorylating the PIP3 to PIP2) that drives PI3K signaling cascade
cancer [7,9,49]. which feeds onto both mTORC1 and mTORC2 complexes.
Amplifications and genetic mutations are the common genetic changes
5.1. mTOR and brain diseases that constitutively activate the protein molecules. Either activation of
RTKs or the other upstream members such as PIK3CA, RAS (H, K, and
Recently, various advanced next-generation sequencing technolo- NRAS), Akt and loss of PTEN molecules could activate this signaling
gies paved the way for identifying many disease-associated mutations axis that in turn trigger mTOR which result in activation of the mTOR

4
A.K. Murugan Seminars in Cancer Biology xxx (xxxx) xxx–xxx

Fig. 2. mTOR-mediated signaling pathways in cancer. Growth factors, stress, amino acids, energy, and oxygen give inputs to mTOR which activate the mTORC1.
Activated mTORC1 regulate stress/DNA damage, enhance nucleotide synthesis, protein synthesis and metabolism that promotes, cell proliferation, cell survival
metastasis, etc. Growth factor-induced RTKs activate mTORC2 which then increase cell proliferation, survival, and cytoskeleton rearrangements by phosphorylating
various downstream effectors.

5
A.K. Murugan Seminars in Cancer Biology xxx (xxxx) xxx–xxx

signaling. Receptor tyrosine kinase, EGFR is frequently shown to be oncogenesis [97]. Recently, it has been demonstrated that p70S6K1-
genetically altered in human cancers. The EGFR member, HER2 mediated phosphorylation of PIPKIγ90 is critical for the development of
(ERBB2) is found to be overexpressed in breast cancer in ∼10-30% of focal adhesions and invadopodia formations, the key machinery for
patients and associated with poor survival [58]. EGFR gene mutation migration and invasion [98]. Consistent with this finding, a subsequent
results in overexpression of this gene and its mutations are often found study also found that this molecule plays a leading role in motility of
in human cancers. EGFR mutations are more frequently reported in cells, a preclinical testing of a newly developed small molecule inhibitor
∼25% to 82% of colorectal cancers [59], ∼30-50% glioblastomas [60], of p70S6K1 (FS-115) showed that the compound could potentially in-
and in 5%–20% in non-small-cell lung cancer, etc [61]. Though these hibit the colony formation and anchorage-independent growth of breast
tumors benefitted from the monoclonal antibody-based (Herceptin) cancer cells suggesting that opportunity for repressing loco-regional
molecular targeted therapy, but only achieved for patients with the relapse and metastasis using p70S6K inhibitor in breast cancer [99].
wild-type RAS molecule. PDGFRα mutations were identified in ∼5% of Analyses of phosphorylated 4EBP1 (p-4EBP1) revealed that this protein
GIST (gastrointestinal stromal cancer) [62] and a 5–10% of amplifica- molecule was overexpressed and its expression was directly associated
tions were found in esophageal cancer, brain malignancies (glio- with progressive malignancies and an unfavorable prognosis in-
blastoma multiforme and oligodendrocytoma), and artery intimal sar- dependent of the upstream oncogenic aberrations in major malig-
comas [63–67]. Some gene rearrangements have also been reported in nancies including breast, ovary, and prostate cancers. This study sug-
leukemia [68]. TK inhibitors were developed and used for the GIST- gested that p-4EBP1 likely to function as a funnel factor for oncogenic
positive KIT; despite a significant increase in overall survival, patients signaling particularly in oncogenic ability and self-sufficiency in pro-
develop resistance over time. FGFR mutations have been reported in a moting proliferating signals and may be a prominent molecular marker
wide spectrum of human cancers [69]. Somatic mutations of FGFR1/2/ of malignant ability [100]. Moreover, it has been found that cell-type
3 were reported and shown to have constitutive activation of this specific 4EBP1 abundance level could determine the sensitivity or re-
protein in head and neck, melanoma, brain, breast, lung, stomach, sistance to inhibitors of the PI3K/Akt pathway [101]. A further study
uterus, and colon cancers [70–80]. Amplifications of FGFR1/2 have also also confirms that the overexpression of p-4EBP1 associates with poor
been reported in breast cancer [81–84]. The FGFR mutation in gastric prognosis in human cancer [102]. A recent study found that the levels
cancer was shown to be associated with poor prognosis [85]. Moreover, of 4EBP1/eIF4e activation could readily predict early and late recur-
chromosomal translocation in the FGFR produces oncogenic fusions and rence in renal cell carcinomas [103]. The eIF4E has been shown to be
this was frequently reported in several blood cancers including multiple implicated in cell proliferation, oncogenesis, and metastasis of human
myelomas and myeloproliferative disorder syndrome [86–88]. FGFR cancer including prostate cancer [104,105]. Moreover, recently it is
inhibitors are in clinical trials and these inhibitors (Brivanib) are ef- revealed that the p-4EBP1 is a novel independent predictor of overall
fective both on wild-type and activated FGFR3 isoforms. Further, and progression-free survival and suggested that it could be a potential
PIK3CA is an oncogenic activator of mTOR signaling and shown to be biomarker for prognostic stratification and treatment decisions in
amplified and mutated in human cancers [89,43]. Otherwise, the loss- ovarian cancer patients [106].
of-function of PTEN also gives a similar effect on this signaling
pathway. The loss-of-function mutations/deletions of PTEN were often 6.3. Genetic deregulation of mTOR in human cancer
found in human malignancies [90]. The RAS proto-oncogenes such as
HRAS, KRAS and NRAS were often altered in human cancers. In Copy gain, amplification and/or mutation of a gene are the typical
sporadic cancer, activating KRAS mutations were more frequent at mechanism for an oncogenic activation which exerts growth and shut
21.6%, when compared to either NRAS (8%) or HRAS (3.3%) [91]. off the apoptosis and autophagy. The mTOR is shown to be deregulated
Activated RAS could directly turn on MAPK pathway and binds to via different molecular mechanisms which ultimately lead to human
p110α via the RBD (RAS-binding domain) that results in activation of pathogenesis. The recent literature shows that mTOR is frequently ge-
PI3K/Akt pathway which has the capacity to inactivate TSC1/TSC2 via netically deregulated that includes gene amplification and mutation.
phosphorylation of TSC2 by activated MAPK and PI3K downstream
members such as ERK (extracellular signal–regulated kinase) [92], RSK 6.3.1. Copy number alterations (CNAs): gain/loss of copy number
(ribosomal s6 kinase) [93], and Akt [94], respectively thus results in the alterations
conversion of Rheb to a GTP-bound form and activation of mTORC1. The mTOR gene is located at chromosome 1p36.22 and this locus is
Further, the Raptor is directly phosphorylated by RSK in order to fur- often genetically altered in human cancers. The copy number variations
ther enhance the activity of mTORC1 [95]. Gene amplification of for this locus are clearly recorded in the COSMIC (Catalogue Of Somatic
members of RTKs which are upstream of both PI3K and RAS are also Mutations in Cancer, UK) database. The prevalence of both the gain and
commonly detected in cancer that also ultimately leads to aberrant loss of copy number variation is not significantly high in the observed
signal transduction through both mTOR complexes (Fig. 2). cases as all the copy number variations shown in different types of
cancers invariably reveals a small fraction of percentage. For example,
6.2. Activation of downstream members of the mTOR signaling pathway in though copy gain was observed in stomach cancer, 0.4% was the
human cancer highest prevalence among the various malignancies with the copy gain.
On the other hand, the highest prevalence of copy number loss was
The downstream effectors of mTOR including p70S6K, 4EBP1, and observed in adrenal gland with a prevalence of 1.49% while the other
eIF4 are reported to play a key role in cancer by regulating protein cancers exhibited even much lower than this tumor type. These ob-
synthesis, survival, and cell growth (Fig. 2). A microarray-mediated servations clearly show that loss or gain of copy number alterations-
analysis used 4209 complementary DNA clones from breast cancer cell mediated genetic deregulation may contribute to only a small fraction
lines, primary tumors with matched normal tissues to determine the of human cancers and suggest the importance of obtaining a clear ge-
amplification of p70S6K1. They found that the p70S6K1 was amplified netic data before the therapeutic strategy which could facilitate ca-
and highly overexpressed in MCF-7 cells. Moreover, compared with the tering the cases for precision medicine (Table 1).
normal tissues, p70S6K1 was identified to be amplified in 8% (59/668)
of primary breast cancers and the amplification was significantly as- 6.3.2. Amplification: overexpression /underexpression of mTOR gene
sociated with poor prognosis (p = 0.0021) while combined with HER-2 A detailed analysis of the COSMIC database revealed a frequent
found poor survival (p = 0.0001) thus suggested a crucial role in on- overexpression of mTOR gene in human cancer (Table 1). Over-
cogenesis and prognosis of breast cancer patients [96]. Further, the expression of mTOR gene was more prevalent in the skin cancers
p70S6K1 was shown to be implicated in glial cell transformation and (8.46%), urinary tract cancers (6.62%) followed by soft tissue cancers

6
A.K. Murugan Seminars in Cancer Biology xxx (xxxx) xxx–xxx

Table 1
Genetic alterations of the mTOR gene in human cancers.
S. No Primary tumor tissues Mutations (%)* Copy number variations (%)* Gene expression (%)*

Gain (Chro. 1p36.22) Loss (Chro. 1p36.22) High Low

1 Adrenal gland 0/657 (0%) – 4/268 (1.49%) 1/79 (1.27%) –


2 Anatomic ganglia 0/1221 (0%) – – – –
3 Biliary tract 19/1056 (1.8%) – – – –
4 Bone 3/734 (0.4%) – – – –
5 Breast 87/4641 (1.87%) – 2/1544 (0.13%) 39/1104 (3.53%) 37/1104 (3.35%)
6 Central nervous system 28/3160 (0.89%) 1/1093 (0.09%) – 21/697 (3.01%) 90/697 (12.91%)
7 Cervix 16/384 (4.17%) – – 17/307 (5.54%) 5/307 (1.63%)
8 Endometrium 59/942 (6.26%) – – 28/602 (4.65%) 9/602 (1.5%)
9 Eye 1/177 (0.56%) – – – –
10 Fallopian tube 1/3 (33.3%) – – – –
11 Gastrointestinal tract 0/1 (0%) – – – –
12 Genital tract 14/249 (5.62%) – – – –
13 Hematopoietic & lymphoid 38/4809 (0.79%) – – 5/221 (2.26%) 6/221 (2.71%)
14 Kidney 134/2879 (4.65%) – – 17/600 (2.83%) 7/600 (1.17%)
15 Large intestine 244/3839 (6.37%) – 1/773 (0.13%) 13/610 (2.13%) 27/610 (4.43%)
16 Liver 53/2274 (2.23%) – – 5/373 (1.34%) –
17 Lung 142/4448 (3.19%) 4/1185 (0.34%) – 53/1019 (5.2%) 21/1019 (2.06%)
18 Meninges 0/95 (0%) - – – –
19 Nervous system 24/382 (6.28%) – – – –
20 Oesophagus 31/1633 (1.9%) 1/546 (0.18%) – 6/125 (4.8%) –
21 Ovary 28/1373 (2.04%) 1/729 (0.14%) – – 28/266 (10.53%)
22 Pancreas 23/2447 (0.94%) – – – 12/179 (6.7%)
23 Parathyroid 4/45 (8.89%) – – – –
24 Penis 2/7 (28.57%) – – – –
25 Perineum 0/1 (0%) – – – –
26 Peritoneum 1/33 (3.03%) – – – –
27 Pituitary 0/60 (0%) – – – –
28 Pleura 0/463 (0%) – – – –
29 Prostate 26/2828 (0.92%) – 1/1037 (0.1%) – 15/498 (3.01%)
30 Salivary gland 7/388 (1.81%) – – – –
31 Skin 140/1844 (7.49%) 2/650 (0.31%) – 40/473 (8.46%) 24/473 (5.07%)
32 Small intestine 4/111 (3.64%) – – – –
33 Soft tissue 9/1322 (0.68%) – – 17/263 (6.46%) –
34 Stomach 44/1260 (3.49%) 2/501 (0.4%) – 18/285 (6.32%) 4/285 (1.4%)
35 Testis 5/476 (1.05%) – – – –
36 Thymus 0/172 (0%) – – – –
37 Thyroid 36/1846 (1.95%) – – 17/513 (3.31%) 7/513 (1.36%)
38 Upper aero-digestive tract 25/1703 (1.47%) – – 26/522 (4.98%) 7/522 (1.34%)
39 Urinary tract 58/1137 (5.11%) – – 27/408 (6.62%) –
40 Vagina 0/2 (0%) – – – –
41 Vulva 1/30 (3.33%) – – – –

1307/51,132 (2.55%) 11/4704 (0.233%) 8/3622 (0.22%) 350/8201 (4.26%) 299/7896 (3.78%)

*
Mutated samples/Total number of sample analyzed (%); -, not analyzed. The data is derived from COSMIC as on February 24, 2019.

(6.46%) and stomach cancer (6.32%). Further, to date, analysis of the RTP801 shows no physical interaction with both the TSC1/2 [111].
PubMed database revealed no evidence of individual report on over- Interestingly, it has been observed for the first time that down-
expression of mTOR gene in human cancer. regulation of mTOR and P70S6Kβ2 expression is associated with poor
As seen in Table 1, the mTOR gene was recorded to be down- prognosis in T-cell acute lymphoblastic leukemia (T-ALL) [112]. These
regulated in many human cancers such as breast, central nervous observations warrant future studies on mTOR gene amplification in
system, cervix, endometrium, hematopoietic, lymphoid, kidney, etc. human cancers.
The most prevalence of mTOR downregulation was detected in malig-
nancies of the central nervous system (12.91%) followed by the ovary
(10.53%), and pancreas (6.7%). As a result of upstream signaling ac- 7. Hyperactivating mutations of mTOR gene
tivation or genetic mutations in mTOR may achieve gain-of-function
and hyperactivate the mTOR signaling, thus it is clearly expected that Initially, the hyperactivation of mTOR was observed by Abraham
mTOR is more likely to show overexpression rather than under- and his group. When the mTOR mutant (named as ΔTOR-deletion
expression. These features show that mTOR has split personalities to- mutant) was generated with unimpaired phosphorylation site for Akt
wards triggering oncogenic signaling and warrant further studies on (2430–2450 amino acids close to the carboxyl terminus of the mTOR)
how these downregulated gene expressions could contribute to the and overexpressed in HEK293 cells, they found that mutant could en-
oncogenesis. In normal physiological condition, downregulation of hance a 3.5-fold of its basal protein kinase activity resulting in phos-
mTOR has been shown in response to maternal nutrient restriction in phorylation of p70S6K and mTOR-mediated in vivo signaling. This
the placental tissues of the baboon [107]. In general, hypoxia, energy group also found that mTOR is the direct target of the PI3K/Akt sig-
stress or exposure to dopaminergic neurotoxins causes downregulation naling pathway [113]. Four years later, CB Thomson and his group
of mTOR if the cell is under extreme stress and it is exclusively de- observed that expression of the ΔTOR-deletion mutant in the p53-/-
pendent on intact TSC1/2 and the expression of RTP80 [108–110]. mouse embryonic fibroblasts (MEFs) could readily enhance the colony
Although RTP801 requires the intact TSC1/2 to downregulate mTOR, formation efficiency and addition of adenoviral E1A protein further
increase the number of colony formation suggested that the deletion

7
A.K. Murugan Seminars in Cancer Biology xxx (xxxx) xxx–xxx

mutant of mTOR may contribute to cell survival and transformation sequencing of mTOR gene in thyroid cancer samples revealed no mTOR
[114]. In line with these studies, two other activating mutations were mutation in differentiated thyroid cancer [126]. It is possible to spec-
found when analyzed for caffeine target in yeast TOR complex1 [115]. ulate that mTOR mutation likely to exist in human cancer but they
Urano et al identified 22 single point mutations in the Tor2 protein might have been missed due to poor sensitivity of Sanger sequencing
which could promote growth in the absence of rhb1 (the Rheb homolog and/or common intra-tumor heterogeneity of mTOR mutations in
in yeast). It has also been found that mTOR harboring analogous to the human cancer [127]. Notwithstanding, whole-exome sequencing of the
mutations identified in the yeast (L1460P or E2419K) exhibit con- ATC (anaplastic thyroid cancer), most aggressive and deadly thyroid
stitutive activation even in the nutrient-starved conditions though cancer samples revealed ∼10% (2/22) of mTOR mutations (R164Q and
sensitive to rapamycin. Moreover, it has been found that a heterodimer M2327I), despite uncharacterized [128]. A recent study detected mTOR
comprising a mutant with a wild-type form of mTOR is active in nu- somatic mutations in melanoma with a prevalence of ∼11%, (43/412)
trient-starved cells [116]. Subsequently, two other mTOR mutations which was associated with increased melanoma-specific mortality
were shown to be hyperactive [117]. Even though these mTOR muta- [129]. The prevalence of mTOR mutation is profiled in Table 1 and the
tions were derived from non-tumors, they were hyperactive, all mutants frequently occurred mutations are shown in Fig. 3.
activated the immediate substrate, p70S6K, and enhanced colony for-
mation. In 2010, for the first time, Sato et al reported that some cancer- 9. Oncogenic mutations of mTOR in human cancers
associated mTOR mutations derived from COSMIC database possessed
gain-of-function activities [118]. Nevertheless, till then, whether mTOR Although somatic mutations were earlier identified in the mTOR
mutation itself could transform the normal fibroblasts and make tumor gene, Sato et al were the first to functionally characterize the cancer-
in vivo was completely unclear until M Xing and his colleagues de- associated mTOR mutation in vitro. Their study selected human cancer-
monstrated in 2013 [119]. When mutations were introduced in the associated mTOR mutations (A8S, S2215Y, P2476L and R2505P) from
critical domains of mTOR and experimented both in vitro and in vivo, the COSMIC database and introduced them in the mTOR gene using
found those mTOR mutations could increase protein kinase activities site-directed mutagenesis. Transient overexpression of each of these
many folds, activate mTOR/p70S6K and Akt signaling pathways. Fur- mutants in HEK293T cells revealed enhanced protein kinase activity,
ther those mutations induced cell transformation, invasion in NIH3T3 and hyperactivation of mTOR/p70S6K signaling in cells transfected
cells and remarkably rapid tumor formation in nude mice [119]. These with S2215Y and R2505P mutants. Interestingly, even under nutrient-
results suggested for the first time that mTOR is tumorigenic upon starved conditions, the phosphorylation of mTORC1 substrates (p70S6K
mutation (Table 2). and 4EBP1) was intact with higher phosphorylation. Further, compared
to the wild-type, enhanced focus formation was observed in Rat1 cells
8. Prevalence of mTOR mutations in human cancer when the mTOR mutant (S2215Y) was co-transfected with mutant
KRAS yet, it was insignificant when mTOR mutant alone expressed
Analysis of the COSMIC database displayed that the incidence of [130]. Later, it has been demonstrated that intratumor heterogeneity
mTOR mutations greatly varies among the different types of cancer. The was identified for mTOR mutation which harbored within the auto-
highest frequency was detected in 28.57% (2/7) and 33.3% (1/3) of the inhibitory region. The renal clear cell carcinoma tumor-bearing mTOR
fallopian tube, penis cancer, respectively. Nonetheless, the number of mutant (L2431P) could display enhanced staining of p-p70S6K and p-
tumors analyzed for mTOR mutation was relatively very low and hence 4EBP1, but tumor regions with wild-type mTOR displayed no staining.
raise the skepticism about the incidence of mutation. Several other In addition, expression of this mutant in a clear cell carcinoma cell lines
cancer types follow with a considerably higher incidence of mTOR showed increased phosphorylation of p70S6K even after serum-starved
mutations consisting relatively more tumor samples particularly, condition and was sensitive to everolimus suggesting that the mTOR
parathyroid, skin, large intestine, and endometrium having a pre- mutant could constitutively activate mTOR signaling [127]. Functional
valence of 8.89% (4/45), 7.49% (138/1830), 6.37% (244/3830) and analyses of a large cell neuroendocrine carcinoma-associated mTOR
6.26% (58/938), respectively. The incidence observed in the later cases mutation (L2209V) with 28 other non-synonymous mutations identified
shows higher reliability than the former cases which studied only a few in different human cancer types revealed that 12 of 28 mutations in-
tumors. The mTOR mutations are only rarely found in several other cluding the L2209V could activate p70S6K signaling and transform the
types of cancer. The mTOR mutational spectrum clearly reveals the mouse 3T3 cell phenotypes. Structural analysis displayed that these
possible segregation towards a specific cell/tissue types. It seems that mutations were accumulated between the kinase and FAT domain. All
mTOR mutations occur frequently in tumors of reproductive and ex- these transforming mTOR mutants were shown to accelerate cell sur-
cretory organs and their subset of tumors derived from cells/tissues vival in 3T3 cells, and they were sensitive to rapamycin inhibition
with high rates of self-renewal capacity. Further, a subset of the tumors [131]. Furthermore, clear cell renal cell carcinoma (ccRCC)-associated
which are negative for mTOR mutations may have other genetic events FAT domain mTOR mutations derived from the cBioPortal/COSMIC
with mutual exclusivity that could give a similar downstream effect. database were demonstrated to be gain-of-function mutations as they
Moreover, mTOR mutations might have missed because of intra-tumor could activate both mTORC1 and mTORC2. Likewise, these mutations
heterogeneity, poor sensitivity of Sanger sequencing particularly when enhanced the cell proliferation in HEK293T cells and shown to confer
malignant tumor dissected with more normal surrounding tissue, and relatively resistant to rapamycin-mediated mTOR inhibition. In addi-
may be away from the analyzed hotspot exons. Whole-genome/exome tion, it has been found that FAT domain point mutations of mTOR
sequencing of human cancer has significantly identified the mTOR differentially activate the mTORC1 as it shows nutrient sensitive (p-
mutations along with other somatic mutations suggesting that mTOR is p70S6K at Thr389 and p-4EBP1 at Ser65) and resistant (p-4EBP1 at
significantly mutated in human cancers and required to study and Thr37/47) for the output of phosphorylation for different sites [132].
analyze through more sensitive methods such as deep sequencing, Analyses of cancer-associated recurrent and non-recurrent mTOR mu-
pyrosequencing, and next-generation sequencing including Illumina, tations revealed that all the mTOR mutants tested were activating
etc [120–123]. Whole-exome sequencing of 10 non-familial pancreatic mutations as they displayed enhanced p70S6K phosphorylation. Fur-
neuroendocrine tumors revealed that mutation harbored in the genes of ther, functional characterization of recurrent mutations conferred ac-
the mTOR pathway in 14% of tumors posing a potential treatment with tivation of respective substrates of both mTORC1 and mTORC2 sig-
mTOR inhibitors [124]. Targeted sequencing analysis of 504 cancer- naling such as p70S6K/4EBP1 and Akt at a different strength.
related genes in 29 lymph node metastases of cutaneous squamous cell Moreover, it has been shown that some of those mutants (L1460P,
carcinomas (SCC) found 6.8% (2/29) of mTOR mutations with other C1483F, S2215Y, and R2505P) preferably phosphorylated p70S6K/
somatic mutations and genetic alterations [125]. Initially, Sanger 4EBP1 while V2006I preferred the Akt and suggested that those

8
A.K. Murugan Seminars in Cancer Biology xxx (xxxx) xxx–xxx

Table 2
Functional characterization of activating/oncogenic mutations of mTOR gene.
Type of mutation Mutant Domain Artificial/Tumor* Function References

Missense H419R HEAT Tumor Tumorigenic Murugan et al. 2019 [230]


Missense G2359E Kinase Tumor Tumorigenic Murugan et al, 2019 [230]
Missense Y1974H FAT Tumor Activating Rodríguez et al. 2017 [229]
Missense H1968Y FAT Tumor Activating Kong et al. 2016 [129]
Missense P2213S Kinase Tumor Activating Kong et al. 2016 [129]
Missense M2327I Kinase Tumor Activating Xu et al. 2016 [134]
Missense L2230V Kinase Tumor Activating Xu et al. 2016 [134]
Missense L2185A Kinase Artificial Activating Wu et al. 2015 [209]
Missense L2185C Kinase Artificial Activating Wu et al. 2015 [209]
Missense A1519T FAT Tumor Activating Ghosh et al. 2015 [132]
Missense Y1463S FAT Tumor Activating Ghosh et al. 2015 [132]
Missense K1452N FAT Tumor Activating Ghosh et al. 2015 [132]
Missense V2406A Kinase Tumor Transforming Yamaguchi et al. 2015 [131]
Missense V2006F FRB Tumor Transforming Yamaguchi et al. 2015 [131]
Missense T1971I FAT Tumor Transforming Yamaguchi et al. 2015 [131]
Missense D2512H Kinase Tumor Hyperactive Grabiner et al. 2014 [133]
Missense I2500F/M Kinase Tumor Hyperactive Grabiner et al. 2014 [133]
Missense A2226S Kinase Tumor Hyperactive Grabiner et al. 2014 [133]
Missense Q2223K Kinase Tumor Hyperactive Grabiner et al. 2014 [133]
Missense L2220F Kinase Tumor Hyperactive Grabiner et al. 2014 [133]
Missense R2217W Kinase Tumor Hyperactive Grabiner et al. 2014 [133]
Missense L2216P Kinase Tumor Hyperactive Grabiner et al. 2014 [133]
Missense S2215F/P Kinase Tumor Hyperactive Grabiner et al. 2014 [133]
Missense A2210P Kinase Tumor Hyperactive Grabiner et al. 2014 [133]
Missense L2209V Kinase Tumor Hyperactive Grabiner et al. 2014 [133]
Missense N2206S Kinase Tumor Hyperactive Grabiner et al. 2014 [133]
Missense V2006I FRB Tumor Hyperactive Grabiner et al. 2014 [133]
Missense T1977R/K FAT Tumor Hyperactive Grabiner et al. 2014 [133]
Missense F1888I/L FAT Tumor Hyperactive Grabiner et al. 2014 [133]
Missense E1799K FAT Tumor Hyperactive Grabiner et al. 2014 [133]
Missense E1519T FAT Tumor Hyperactive Grabiner et al. 2014 [133]
Missense C1483F/R/W/Y FAT Tumor Hyperactive Grabiner et al. 2014 [133]
Missense A1459P FAT Tumor Hyperactive Grabiner et al. 2014 [133]
Missense L1433S FAT Tumor Hyperactive Grabiner et al. 2014 [133]
Missense F2108L FRB Tumor Activating Wagle et al. 2014 [135]
Missense E2419K Kinase Tumor Activating Wagle et al. 2014 [135]
Missense E2014K FRB Tumor Activating Wagle et al. 2014 [135]
Missense E2288K Kinase Artificial Tumorigenic Murugan et al. 2013 [119]
Missense P2273S Kinase Artificial Tumorigenic Murugan et al. 2013 [119]
Missense T2294I Kinase Artificial Activating Murugan et al. 2013 [119]
Missense V2291I Kinase Artificial Activating Murugan et al. 2013 [119]
Missense V2284M Kinase Artificial Activating Murugan et al. 2013 [119]
Missense G1479N FAT Artificial Activating Murugan et al. 2013 [119]
Missense W1456R FAT Artificial Activating Murugan et al. 2013 [119]
Missense L2431P Kinase Tumor Activating Gerlinger et al. 2012 [127]
Missense R2505P Kinase Tumor Constitutive Sato et al. 2010 [210]
Missense S2215Y Kinase Tumor Constitutive Sato et al. 2010 [210]
Missense A2020V FRB Artificial Hyperactive Ohne et al. 2008 [117]
Missense I2017T FRB Artificial Hyperactive Ohne et al. 2008 [117]
Missense E2419K Kinase Artificial Constitutive Urano et al. 2007 [116]
Missense L1460P FAT Artificial Constitutive Urano et al. 2007 [116]
Missense A1957V FAT Artificial Enhanced Reinke et al. 2006 [115]
Missense I1954V FAT Artificial Enhanced Reinke et al. 2006 [115]
Deletion ΔTOR 2430-50 Kinase Artificial Transforming Edinger et al. 2004 [114]
Deletion ΔTOR 2430-50 Kinase Artificial Activating Sekulić et al. 2000 [113]

*Artificial, created artificially inserting mutation into the mTOR gene; Tumor, originally derived from primary malignant tumor; #, mTOR mutations without
significant substrate activation by Grabiner et al, 2014 (A41P/S/T, F1888V, T1977S, V2006L, S2215T, R2505Q/*, D2512G/Y). For each mutant, initial report is
listed in the table to avoid redundancy.

mutants selectively activate mTORC1 or mTORC2. These mTOR mu- conferred gain-of-function, sensitive to PI3K-Akt-mTOR pathway in-
tations were shown to be dramatically sensitive to mTOR inhibitors hibitors as these mutants (H1968Y and P2213S) expressing HEK293T
both in vitro and in vivo but partially resistant to nutrient deprivation cells were more sensitive and proliferation was effectively inhibited by
[133]. It has been shown that 22 kidney cancer-derived somatic mu- LY294002 and AZD5363 rather than temsirolimus or BYL719 [129].
tations of mTOR which clustered in FAT and kinase domains were These results collectively suggest that mTOR mutations are oncogenic,
hyperactivating mutations by their increased kinase activity, increase which strongly activate the p70S6K/4EBP1 and Akt, the downstream
cell size, and resistant to glucose and serum starvation withal this does effector signaling at different levels to promote tumorigenesis (Table 2).
not reflect in amino acid deprivation. Further, these mTOR mutations Moreover, all the functionally characterized cancer-associated mTOR
were shown to be resistant to REDD1 (resistant to DNA damage-in- mutations either identified in different types of cancer or derived from
ducible transcript 1)-mediated suppression, induce tumor formation the COSMIC/cBioPortal database were demonstrated to be oncogenic
and those tumors were remarkably sensitive to rapamycin [134]. and sensitive to the mTOR inhibitor, rapamycin except - an mTOR
Functional characterization of melanoma-associated mTOR mutations mutation identified in an ATC. It was shown to confer resistance to

9
A.K. Murugan Seminars in Cancer Biology xxx (xxxx) xxx–xxx

Fig. 3. Genetic deregulation of mTOR in human cancers. A. Human cancer-associated hotspot somatic mutations and drug-resistant mutations of mTOR. Shown is the
schematic diagram of human mTOR protein structure which consists of various functional domains. Mutations plotted on top of mTOR indicate human cancer-
associated somatic mutational hotspots (frequency of 5 and more) and the mutations plotted in the bottom shows the drug-resistant mutations. Boxes indicated in the
bottom shows various mTOR inhibitors and their binding domains. B. Types of mTOR mutations. Tables show various types of mTOR mutations and different types of
nucleotide changes identified to date in human cancer, respectively. This data was derived from the COSMIC database. C.A 3D ribbon diagram of mTOR. It shows the
FRB domain with the drug-resistant mTOR mutations and other domain (PDB ID: 4jsv).

allosteric mTOR inhibitor but sensitive to kinase inhibitors [135]. conclusion that these mutants directly disturb the DEPTOR interacting
site on mTOR. This observation was also consistent with the other two
activating mTOR mutants localized in C1483 cluster of the FAT domain
10. Mechanisms of mutation-mediated activation of mTOR
(A1459P and C1483Y). These findings show that the possible me-
chanism of a subset of activating mutation particularly of FAT domain
The mechanism of mutant-induced mTOR activation is not ex-
and suggest other possibilities such as mutations in other domains may
clusively studied and clearly understood. However, it has been specu-
have different mechanisms to enhance the activity of kinase which
lated that mutations harboring in the kinase domain and close to the
could result in activation the mTORC1 [133].
activation loop are likely to alter the conformation that may result in
Xu et al divided 28 activating mutations of mTOR double mutants
enhanced kinase activities [119]. Further, Grabiner et al contemplated
into 6 groups and demonstrated that those mutants fall in the category
that mutation might distort the mTORC1 or mTORC2 assembly which
of synergistic Rheb independent mTORC1 activation. Moreover, it has
results in enhanced p70S6K1/Akt1 phosphorylation, respectively or
been shown that mTOR mutants were resistant to REDD1-mediated
mutant mTOR may mislay the binding efficiency to DEPTOR, the en-
inhibition suggesting that in the absence of the tumor suppressor, VHL
dogenous inhibitor of mTOR which binds to the FAT domain [136].
(von Hippel–Lindau) expression, activating mTOR mutations could es-
Experiments on protein-protein interactions revealed that mutant
cape from the negative feedback loop of mTORC1 and HIF1. Notably,
mTOR proteins showed reduced binding to DEPTOR compared with
activating mTOR mutants expressing VHL-deficient cells could grow
wild-type mTOR. On the other hand, like wild-type, all mTOR mutant
tumors which were inhibited by rapamycin. It additionally suggests that
proteins could bind to their binding partners including Raptor and
mTOR mutations harboring in cells with VHL loss and HIF activation is
Rictor towards the formation of the two mTOR complexes. These results
highly pathogenic in clear cell kidney cells [134].
suggested that mutations have no role in the complex formation and it
was consistent with the previous finding that even artificially created
mutants could activate both mTORC1 and mTORC2 [119]. These 11. The mTOR is a driver of invasion and metastasis
findings suggest that reduced DEPTOR binding could enhance mTOR
activity in mTOR mutants. Most of the cancer deaths occur because of the metastasis of the
Further, it has been demonstrated that two mTOR mutations in the primary tumor to the multiple organs. Acquiring the migratory and
FAT domain (L1460P and C1483Y) could activate mTORC1 and bind invasive potential is the initial step and that is the rate-limiting primary
relatively very low to DEPTOR compared to other mutants and draw the step in the metastasis cascade. Epithelial-mesenchymal transition

10
A.K. Murugan Seminars in Cancer Biology xxx (xxxx) xxx–xxx

(EMT) is found to be the critical mechanism in the primary step in the attenuate migration and invasion potentially suppressing EMT in gall-
metastatic process because during this molecular program epithelial bladder cancer [155]. In addition, inhibition of PI3K/mTOR and sonic
cells undergo various steps by losing polarized. Differentiated pheno- hedgehog (SHH) pathways could effectively inhibit the cancer stem cell
types have many cell-cell junctions to mesenchymal phenotype which features and tumor progression in human pancreatic cancer stem cells
includes non-polarized; loss of cell-cell junctions often results in in- [156]. These reports collectively suggest that the mTOR and its sig-
creased motility and invasive potential. Moreover, these phenotypes naling pathways are critically involved in regulating EMT and the
exhibit more resistance to chemotherapy [137]. The flexibility of this metastasis cascade and treating with rapamycin potentially inhibit the
cascade is well known by the fact that EMT is reversible and the process mTOR and its signaling partners and reverse the invasive phenotypes
is named mesenchymal-epithelial transition (MET). In addition to this, represent an opportunity for rapamycin analogs in therapeutically tar-
during the initial step of the metastasis cascade, small GTPases, that geting various metastatic human cancers.
includes RhoA and Rac1, are activated that results in actin cytoskeletal
rearrangement, cell migration, and invasion. Particularly, invasive 12. The mTOR inhibitors: rapamycin
phenotypes are made when the RhoA protein gives the stimulus to form
actin stress fibers and cell-cell adhesions, and Rac1 induces the for- The PI3K/Akt/mTOR signaling pathway is one of the most com-
mation of lamellipodia [138]. This programmed-change in the cell monly deregulated pathways found in human cancers that supplied the
phenotype enables the cell to be more invasive and that leads to being rationale for therapeutically targeting mTOR in cancer [47,39]. Sub-
involved in metastasis machinery. The contribution of mTOR and its sequent research activities led to synthesize and develop several small-
signaling members in the invasion and metastasis were demonstrated molecule inhibitors that could target many active molecules in the
when ribosome profiling was performed in prostate cancer [139]. pathway. Indeed, as of today, only the mTOR inhibitors were the first
Further studying the functional role of mTORC2 shed more lights on the small molecules which were translated from bench to bedside to target
role of mTOR in the metastasis cascade. the PI3K/Akt/mTOR pathway [157,158]. To date, three generations of
Moreover, it has been found that TGF-β could induce enhanced cell mTOR inhibitors were developed. In the early phase, rapamycin was
size and protein content during EMT that resulted in activation of discovered with inhibiting the proliferation property in yeast [1], after
mTOR via the PI3K/Akt leading to the phosphorylation of the direct long time research demonstrated that rapamycin has im-
regulators of translation initiation molecules, p70S6K1, and 4EBP1. munosuppressive effects and many years later, FDA-approved for pre-
Rapamycin inhibited the TGF-β-induced EMT, translation signaling and vention of transplant rejection mainly as an immunosuppressive agent
decreased the migratory and invasive behavior of cells. In addition, the [159]. Currently, a high dose of rapamycin is clinically used (e.g.,
results revealed that the TGF-β-induced canonical transcription 5 mg/day), and enrolled in many clinical trials for many types of can-
pathway via Smads could have complemented by the mTOR-mediated cers where mTOR and its signaling are hyperactivated and critical for
translation signaling [140]. The same group further delineated the role promoting and/or sustaining oncogenic transformation. It has been
of mTOR in the EMT transition and found that TGF-β-induced EMT discovered that rapamycin mechanistically acts by binding with the
transition and invasion was driven solely by the mTORC2 and suggested protein FKBP12 [2], producing a complex that can bind the FRB region
that mTORC2 is a necessary member in the downstream of TGF-β sig- of mTOR that could partially block the active site of mTOR resulting in
naling pathway [141]. Nevertheless, it has also been shown that both suppression of the mTORC1 activity [160]. This rapamycin-mediated
the mTORC1 and mTORC2 were demonstrated to play a major role in inhibition of mTORC1 results in suppression of protein synthesis, in-
regulating EMT, migration, invasion, and metastasis mostly through hibition of cellular growth and enhances autophagy and these processes
RhoA and Rac1 signaling mechanisms in colon cancer [142]. On the facilitate tumor regression [161]. Although the rapamycin effectively
other hand, this was further confirmed by a later study in prostate targets the mTORC1, it does not equally occlude the phosphorylation of
cancer model that mTOR could regulate EMT through RhoA and Rac1 all its substrates. The p70S6K1 phosphorylation is exclusively inhibited
signaling pathways [143]. Furthermore, the direct substrate of mTOR, but the 4EBP1 phosphorylation is partially blocked [162]. On the other
p70S6K was shown to promote the IL-6-induced EMT and metastasis in hand, this phenomenon was also truly observed in activated mTOR
HNSCC (head and neck squamous cell carcinomas) [144]. Further, the mediated phosphorylation as it phosphorylates more preferentially
mTOR signaling pathway was reported to be involved in playing a p70S6K1 than 4EBP1. This effect was may be largely dependent on
major role in regulating the growth and dissemination of metastatic species and cell-type [119]. Although differential substrate quality
brain tumors through the classical EMT mechanisms [145]. Infiltrating could also be a cause, mTOR, rapamycin, and FKBP12 crystal structure
macrophages were shown to enhance the EMT and stem cell-like po- analysis predicted that this may be due to differential substrate access
pulations through Akt and mTOR signaling in renal cell carcinomas to the active kinase site of mTOR which is exclusively controlled by the
[146]. Consistent with these reports, use of dual PI3K/mTOR inhibitor mTOR FRB domain [163].
NVP-BEZ235 was shown to strongly inhibit the EMT transition which
was induced by hypoxia and TGF-β1 in both ovarian (SKOV-3) and 12.1. First generation mTOR inhibitors: allosteric inhibitors
prostate (PC-3) cancer cell lines suggest a role of mTOR in the EMT
transition [147]. Apart from the direct involvement of the mTOR and its Rapalogues, the structural and functional analogs of rapamycin
signaling, upregulation of various protein molecules such as TRIM29 were developed particularly to improve bioavailability and pharmaco-
(tripartite motif-containing protein 29), BMP2 (bone morphogenetic kinetics and these allosteric inhibitors were the first generation in-
protein 2), TRIM44 (tripartite motif containing protein 44), UBE2Q1 hibitors. As seen in Table 3 and Fig. 4A, temsirolimus is a prodrug form
(ubiquitin conjugating enzyme E2 Q1), and PGAM1 (phosphoglycerate of rapamycin which inhibits mTORC1 by forming a complex with the
mutase 1) were demonstrated to promote proliferation and metastasis peptidyl-prolyl cis-trans isomerase FKBP12 that is integrated into
via mTOR and its signaling pathway in many different kinds of human mTORC1 and suppress the mTOR. This drug was approved in 2007 as a
cancers that includes nasopharyngeal carcinoma, non-small cell lung monotherapy for the treatment of renal cell carcinoma (RCC) with
cancer, hepatocellular carcinomas, pancreatic-ductal adenocarcinomas, advanced stage [157]. Everolimus also act allosterically by recruiting
respectively [148–152]. Downregulation of ROC-1 (regulator of cullins- the FKBP12 to mTORC1. Though the drug has a profound effect on
1) was demonstrated to inhibit malignant progression via mTOR/ mTORC1, the mTORC2 does not show any sensitivity to inhibition as
DEPTOR signaling in muscle-invasive transitional cell carcinoma [153]. observed in temsirolimus. This drug was approved by FDA as mono-
Fibrinogen-mediated EMT and malignant tumor phenotypes were also therapy for the therapeutic use of several human malignancies such as
shown to be regulated through the Akt/mTOR signaling in esophageal hormone receptor-positive, and HER2-negative breast cancer in post-
cell carcinoma [154]. Inhibition of mTOR was demonstrated to menopause cases, selective neuroendocrine tumors, pediatric/adult

11
A.K. Murugan Seminars in Cancer Biology xxx (xxxx) xxx–xxx

SEGA (subependymal giant cell astrocytoma) with TSC (tuberous

FDA approved in 2003-18


sclerosis complex), pancreatic neuroendocrine tumors, renal angio-

(for a series of diseases)


FDA approved in 2007-
FDA approved in 1999 myolipoma associated with TSC, advanced stage RCC [164–169]. Ri-
daforolimus, another analog of rapamycin progressed to the phase III
SUCCEED trial in metastatic soft-tissue sarcomas despite it was dropped
as no effective activity was observed [170]. Further, nab-rapamycin
Status

(nanoparticle albumin-bound rapamycin) named ABI-009 exhibited


2008

better bioavailability compared with oral rapamycin in RCC, NET, and


mesothelioma and it has progressed to phase II studies [171].
Advanced renal cell carcinoma, neuroendocrine tumors of

Rapalogs produced no effect on mTORC2 however, sustained


pancreatic, gastrointestinal, & tumors of lung origin as

treatment was shown to have an impact on mTORC2 as well [22]. This


may be via feedback mechanism through insulin signaling and/or via
the key subunit FKBP12-mediated binding competition, that could be
the limiting factor for the sensitivity that results in differential sensi-
tivity to rapalogues between the two complexes [172]. In human cancer
cell lines, the inhibitory effect of mTORC2 was found only after 24 h of
treatment. In practical, for human beings, it may be required for more
Acute lymphocytic leukemia
Cancer & Immuno rejection

than many weeks and months. The mTORC2 inhibition was demon-
strated to have a bigger role in dysregulated glucose metabolism, in-
sulin insensitivity which leads to diabetes. Various tissue-specific RNAi
studies on invertebrate animals particularly annelids assumed that it
Disease setting

monotherapy

was because of mTORC2 activity loss mainly in the intestinal region


that ultimately resulted in impaired glucose metabolism. [173]. It is
noteworthy that most of these studies evaluate based on phosphoryla-
tion of Akt on serine 473 by mTORC2 as a reflection of activity of
mTORC2 nevertheless this Akt phosphorylation site is may also be
PI3K/Akt/mTOR pathway
Activated mTOR pathway
PIK3CA mutation and/

targeted particularly by other kinases such as DNA-PK [174], IKKε, and


TBK1 suggesting an alternative mechanism for the mTORC2-specific
target site in Akt at serine 473 [175]. Nonetheless, compared with
amplifications

mTOR kinase inhibitors (TORKi), allosteric mTOR inhibitors (rapa-


alterations
Biomarker

mycin), partially suppress mTORC1 [176].

12.2. Second generation mTOR inhibitors: mTOR kinase inhibitors


Interaction is similar to rapamycin, hydrogen at C-40-O
position is replaced by dihydroxymethyl propionic acid

Similar to rapamycin, hydrogen at C-40-O position is

The mTOR has been reported to be hyperactivated in several types


Interacts with FKBP12 and blocks mTOR substrate

of human cancers and treatment of these cases with rapamycin caused


mTORC2-mediated activation of Akt as the rapamycin inhibits only the
mTORC1 thus leading to failure of treatment strategies [39,177–179].
This urged to develop second-generation mTOR inhibitors as anti-
replaced with hydroxyethyl group

cancer agents to target activated mTOR, particularly to inhibit both the


mTORC1 and mTORC2 [180]. As illustrated in Fig. 4B, the mTOR ki-
nase inhibitors potentially could inhibit both the mTORC1 and
mTORC2 and these inhibitors compete with ATP for binding to the
mTOR kinase active site. Most of these small molecules dramatically
exhibited high specificity and selectivity towards mTOR kinase. Among
Mode of action

various mTOR kinase inhibitors, vistusertib (AZD2014) was tested in


recognition

advanced stage malignancies including an acinar cell carcinoma [181].


mTOR inhibitors approved for the treatment of human cancers.

ester

Regardless of the unfavorable results, being clinically further evaluated


in multiple phase I and II studies [182]. AZD8055 has been reported to
have many hundred fold better inhibiting effect on the mTORC1 than
Generation

the effect of other PI3Ks [183]. Further, a dual inhibitory effect, both on
First

First

First

mTOR complex and PI3K was observed with the inhibitor, dactolisib
(BEZ235) [184] and exhibited a more than five-fold higher Kd for ATR,
mTORC1

mTORC1

mTORC1

a damage response kinase [185]. Notably, these inhibitors were shown


Class

to have a biphasic effect on Akt. Inhibiting mTORC2 results in depho-


sphorylation of Akt at serine 473 and expeditious yet transient inhibi-
tion of Akt phosphorylation at threonine 308 thus leading to a sup-
pression of Akt signaling. Nevertheless, mTOR inhibition using kinase
inhibitors also alleviate RTK-mediated feedback inhibition resulting in
Rapamycin (Sirolimus)

activation of PI3K leading to activation of Akt by rephosphorylating at


Everolimus (RAD001)
779,NSC683864)
Temsirolimus (CCI-

threonine 308. It clearly revealed that mTOR kinase inhibition could


lead to a different steady state profound mTORC1 abolishment still
Drug name

accumulate activated Akt phosphorylation on threonine 308 exempting


at serine 473. Besides, it was demonstrated that a combination of mTOR
Table 3

kinase inhibitors with RTK inhibitors could potentially suppress bi-


phasic Akt signaling that results in cell death and tumor regression in

12
A.K. Murugan Seminars in Cancer Biology xxx (xxxx) xxx–xxx

Fig. 4. mTOR inhibitors in human cancers. A. mTORC1 inhibitors (rapalogs). The mTORC1 inhibitors fail to inhibit the mTORC2 and that keep the activation of Akt
and mTORC2 signaling-mediated negative feedback loop intact. B. mTOR kinase inhibitors. The mTOR kinase inhibitors inhibit both the mTORC1 and mTORC2 that
exclusively interrupt the PI3K/Akt/mTOR pathway-mediated oncogenic signaling. C. Dual pan-PI3K and mTOR (mTORC1 and mTORC2) inhibitors. These inhibitors
potentially inhibit all four PI3K isoforms, mTORC1 and mTORC2 of the activated PI3K-Akt-mTOR signaling in all human cancers.

vivo. These results reveal the adaptive capabilities of oncogenic sig- inhibitors. This drug was shown effectively to reduce the levels of p-
naling and suggest that the importance of combination therapy to es- 4EBP1 and cell growth. Compared with rapamycin or MLN0128,
cape from the biphasic effect [186]. In the majority of the in vitro stu- RapaLink-1 inhibited the growth and arrested the cell at G0/G1 phase
dies, compared with rapalogues, the ATP-competitive inhibitors more potentially upon RapaLink-1 treatment. Moreover, in vivo anti-
showed significantly higher inhibitory effects. Although various such tumor efficacy of RapaLink-1 was more significant. Interestingly,
small molecules were tested in clinical trials for safety, large scale trials RapaLink-1 treated tumors presented initial regression and followed by
were not yet conducted to show greater and convincing efficacy than tumor size stabilization in xenograft mice nonetheless vehicle, rapa-
the currently available best treatment options and hence, the inhibitors mycin, or MLN0128 treated tumors could steadily grow. This drug ef-
such as WYE354, AZD2014, and AZD8055 were still could not get ap- fectively blocked mTORC1 and further associated with mTOR-inter-
proval from FDA (Table 4). Despite their beneficial effects, many side acting protein, FKBP12, and enabled accumulation of RapaLink-1.
effects were also observed when these drugs were chronically ad- Furthermore, it inhibited cancer-derived, activating mutants of mTOR
ministered which includes ulceration of mucosal tissues, hematological with better efficacy than rapamycin or TORKi [187]. In glioblastoma,
abnormalities, induction of insulin insensitivity, obesity, and diabetes, PI3K-Akt-mTOR signaling is deregulated yet inhibition of PI3K or Akt
howbeit these adverse side effects may mainly be dependent on dose displayed a modest effect on downstream mTOR activity. A compara-
and case dependent [180]. tive analysis of the drug effect of RapaLink-1 to the earlier generation
mTOR inhibitors such as rapamycin and TORKi - produced poor dur-
ability of TORKi compared with rapamycin and Rapalink-1. As Rapa-
12.3. Third generation mTOR inhibitors: RapaLink Link-1 bound FKBP12 enable accumulation of RapaLink-1 molecules in
the cell, it could show better efficacy than earlier generation mTOR
The RapaLink-1, the TORKi linked to rapamycin is a third-genera- inhibitors thus potently could block cancer-associated activating mTOR
tion mTOR inhibitor. This molecule exploits the unique juxtaposition of mutants (Fig. 4A and B) [176]. The dual pan-PI3K inhibitors could
the first and second-generation inhibitors-binding pockets to create a potentially inhibit all four PI3K isoforms (α, β, γ, and δ), mTORC1 and
bivalent interaction which exclusively allows inhibiting the drug-re- mTORC2 of the activated PI3K-Akt-mTOR signaling pathway (Fig. 4C).
sistant mTOR mutants that exhibited resistance to the mTOR kinase
inhibitors (TORKi). Various experiments were conducted and compared
the inhibiting potency of rapamycin, RapaLink-1, and MLN0128 in 13. Drug resistance in human cancers: the pivotal role of mTOR
LN229 and U87MG cells. Overall, RapaLink-1 showed the more potent
inhibiting effect on mTOR than the first and second-generation Precision medicine facilitates molecular targeted therapy and eases

13
A.K. Murugan Seminars in Cancer Biology xxx (xxxx) xxx–xxx

the therapeutic management. Tumors which tend to develop drug re-

Rodrik-Outmezguine et al. 2016


sistance during the course of treatment that complicates the subsequent
management. Understanding and resolving the resistance mechanism is
more complex as most resistance arises due to the activation of up-
stream/downstream molecules upon molecular targeted therapy.
ClinicalTrials.gov*

Further, resistance also arises due to the multi-clonal origin of tumors


NCT00499486

NCT01595009
NCT02352844
NCT02244463

NCT02091531

NCT03166904

NCT02549989

NCT03127020
and heterogeneity in targeted mutation. Tumor cell acquires this sur-
vival mechanism by inducing mutations during the course of the evo-

[187]
lution of malignancies and metastasis. Gene set enrichment analyses
identified activation of mTOR signaling in oxaliplatin-treated early
passaged cell lines and patient-derived xenografts of colon cancer and a
Phase IV
Phase II

Phase II
Phase II

Phase II

Phase II

Phase II

Phase II
Status

combination of oxaliplatin with an mTOR inhibitor in the in vitro and in


vivo experiments demonstrated a synergistic effect [188]. It has been

reported that AT406, the inhibitor of apoptosis proteins (IAPs) an-


Thyroid cancer, Metastatic anaplastic thyroid cancer, Head and
Solid cancers with advanced stage. Advanced pancreatic cancer.

Recurrent or persistent endometrial cancer Endometrial cancer.

tagonist could effectively induce cytotoxicity and pro-apoptotic cascade


on primary as well as established hepatocellular carcinoma cells
Advanced Castration-Resistant Prostat Cancer (CRPC).

(HepG2 and SMMC-7721). Notwithstanding, activation of mTOR in


neck neoplasms, and Endocrine gland neoplasms.

these cells caused resistance to the inhibitor AT406 and silencing the
mTOR or inhibition of mTOR by OSI-027 or by kinase-dead-mutant
Cancers with drug resistant mTOR mutations.
Neuroendocrine tumors. Neuroepithelioma.

dramatically increased the sensitivity in HCC cells suggesting that


mTOR is a likely resistance factor for AT406-induced therapy [189]. In
Non-Hodkin lymphoma Lymphoma.

non-small cell lung cancer (NSCLC) cells harboring KRAS mutation,


acquired resistance to Hsp90 inhibitor (Ganetespib) was shown to be
Advanced solid malignancies.

mediated by ERK-p90RSK-mTOR signaling [190]. Identification of


differentially expressed proteins in a trastuzumab-sensitive (NCI N87)
and resistant NCI N87 (T-R) HER2 positive gastric cancer cell lines by
parallel quantitative proteomics screening revealed the frequent acti-
Disease setting

Solid cancers

vation of mTOR signaling in resistant cells suggesting the need of an


mTOR inhibitor for these trastuzumab-resistant HER-positive gastric
cancers [191]. Tankyrase belongs to the poly [ADP-ribose]polymerase
(PARP) family and it positively regulates the Wnt/β-catenin signaling.
Nonetheless, the Wnt signaling-directed colon cancer cells were not
Mutations in TSC1/2.Activation of PI3K/Akt/mTOR

Mutations in PIK3CA, AKT1, PIK3R1, and PIK3R2,

sensitive to the tankyrase inhibitor and detailed dissection of various


Mutations in PI3K/Akt/mTOR pathway genes.
Amplification and/mutations of PIK3CA gene.
Activation of PI3K/Akt/mTOR/S6K pathway.

pathways revealed upregulation of mTOR signaling. Inhibition of


Drug resistant mTOR FRB & KD mutations.

mTOR reversed the resistant cells to be sensitive for tankyrase in-


Mutations in TSC1/2, NF1/2, or STK11

hibitors and suggested the combination of mTOR inhibitor with tan-


kyrase inhibitors targeting this colon cancer subset [192]. It has been
shown that resistance to sorafenib in acute myeloid leukemia (AML)
was due to hyperactivation of mTOR signaling [193]. A recent phase I
clinical trial that combined the everolimus and capecitabine on HER2-
Rictor amplification

Rictor amplification

metastatic breast cancer showed a better tolerated and enhanced pro-


Mutations TSC1/2

gression-free survival and overall survival [194]. A recent review by


mTOR gene.

MN Hall and his colleague detailed the mTOR-mediated resistance to


Biomarker

pathway.

various molecular targeted cancer therapeutic agents [195]. Collec-


tively all these studies show that mTOR plays a pivotal role in resistance
to major targeted therapies in human cancers regardless of the in-
Generation

hibitors and cancer types and suggest a generous combination of mTOR


Second

Second

Second

inhibitor.
Third

Next

Next
First

First
First

In addition, the mTOR causes self-resistance where it induces re-


sistance to various mTOR inhibitors that act on it. It is noteworthy that
mTOR inhibitors in clinical trials for human cancers.

malignancies harboring genetic alterations in a gene of a certain


pathway can exhibit its growth and survival dependency on that and
Dual PI3K &

Dual PI3K &


mTORC1/2

mTORC1/2

mTORC1/2

mTORC1/2

mTORC1/2
Drug class

likely to show enhanced sensitivity when treated with the related


mTORC1

mTORC1
mTORC1

pathway inhibitors. The hamartoma syndromes and perivascular epi-


mTOR

thelioid-cell cancers harboring alteration of genes of mTOR pathway


such as TSC1/2, and STK11 resulted in activation of mTOR signaling
Sapanisertib (TAK-228, MLN0128

which was targeted by mTOR inhibitors [196–199]. In line with these


studies, an urothelial carcinoma and a metastatic urothelial carcinoma
harboring loss of function mutation in TSC1 and activating mutations in
Rapamycin (Sirolimus)

mTOR gene, respectively exhibited exquisite sensitivity to the mTOR


Everolimus (RAD001)
Everolimus (RAD001)

PQR309 (Bimiralisib)

inhibitor, everolimus [200,164]. Recently, whole exome sequencing of


& INK128)

an everolimus resistant malignant tumor derived from an anaplastic


Sapanisertib

LY3023414
RapaLink-1
Drug name

thyroid cancer patient in parallel with pretreated tumor showed a


Vistusertib

nonsense loss of function mutation (Q1178*) in TSC2 in both pretreated


Table 4

and resistant cases. Yet, the resistant tumor was mutated in the mTOR
gene (F2108L) and it was consistent with immunohistochemical

14
A.K. Murugan Seminars in Cancer Biology xxx (xxxx) xxx–xxx

staining of rpS6 reflecting activated mTOR pathway and pretreated Cas9-mediated genome editing and expression of these mutants in
tumor revealed no mutation in mTOR. This mTOR mutation conferred colorectal (SW480) and lung cancer (H460) cell lines revealed that the
resistance to allosteric mTOR inhibitors in vitro this phenomenon re- L2185A mutation confers tumor type-independent drug-resistance in
veals acquired resistance of mTOR to its allosteric inhibitors and sug- both colorectal and lung cancers [209].
gesting an aggressive relapse with resistant mutations after successive
treatment. Treating this resistant mutation (F2108L) with mTOR kinase 14. The mTOR and precision medicine
inhibitor showed that the mutation was sensitive to it [201]. Recently,
it has been found that single point mutation accounts for acquired re- Focusing on individual genetic variations, environment, and life-
sistance to both the first (rapamycin) and second-generation (a TORKi, style intending to prevent and treat various human diseases is a new
AZD8055) mTOR inhibitors when MCF-7 cancer cell lines were exposed strategy called precision medicine or personalized medicine. Recent
to these inhibitors. A 3-month exposure to these inhibitors developed genetic techniques particularly the next-generation sequencing tech-
resistant colonies, a kinase inhibitor-treated cell harbored a mutation in nologies along with computational technology facilitated whole
the kinase domain (M2327I) and rapamycin-treated cells had a muta- genome, exome and transcriptome sequencing which gathered genomic
tion in the FRB domain A2034V (RR1 cells) and F2108L (RR2 cells). RR data on human diseases that rapidly altered drastic changes in the
and TKi-R clones which were derived from rapamycin and TORKi ex- treatment strategies which paved way for precision medicine. It has a
posed cells, respectively demonstrated significantly less sensitivity to bigger role in molecular targeted therapy with better clinical efficacy
their respective drugs in a 72 h proliferation assay when compared to and pharmacokinetics in addition to diagnosis and prognosis [210]. The
the parental line. This result was consistently reflected also in another mTOR axis including upstream RTKs to downstream p70S6K and 4EBP1
cell line, MDA-MB-468 when these mutants were expressed and treated are the predictive biomarkers and it is activated by one or other me-
with a related inhibitor. Rapamycin-resistant FRB domain mutations chanism such as mutation and/or overexpression. The PI3K/Akt/mTOR
have also been found in the yeast by random mutagenesis [202–205] signaling pathway is hyperactivated in more than 30–80% of human
and also been reported in untreated patients in the databases. Similarly, cancers. Therefore, targeted therapies using single mTOR or dual PI3K/
rapamycin-resistant mutants which were identified in this screen also mTOR or Akt/mTOR are being investigated in clinical trials. Everolimus
likely to disrupt the mTOR interaction with FKBP12–rapamycin com- monotherapy has been shown to be very effective with advanced gastric
plex (Fig. 3). cancers with PIK3CA mutations and rpS6K aberrations suggesting that
As an analysis of mTOR kinase structure (PDB: 4JT5) displayed that the predictive biomarkers are prime important [211]. Further, PIK3CA
M2327 is > 15 Å away from the inhibitor, the author speculated that and PTEN mutation resulting in Akt activation was reported to be very
this mutation causes resistance via either reduced TORKi affinity due to sensitive to allosteric mTOR inhibitors [212,213]. Renal cell carci-
an allosteric mechanism or a mechanism which is not require reduced nomas show enhanced dependence on metabolic processes such as
drug binding. In contrast, both wild-type and a mutant mTOR (M2327I) pentose phosphate shunt, aerobic glycolysis, augmented lipogenesis,
bound AZD8055 with similar affinities. Moreover, M2327I mutant when metabolic pathways are deregulated. For example, oxygen sen-
showed > 3 fold enhanced kinase activity compared with the mTOR sing (VHL/HIF), HIF-responsive genes (VEGF, PDGF, and EGF), glucose
wild-type and rapamycin-resistant mutant. Like M2327I, other mTOR transporters (GLUT1 and GLUT4), reduced activities of AMPK and
kinase domain mutants identified in untreated patient tumors also ex- Krebs cycle and/or nutrient sensing (AMPK-TSC1/2-mTOR and PI3K-
hibited insensitivity to mTOR kinase inhibitors (TORKi, AZD8055 and Akt-mTOR), energy sensing (fumarase-deficient, SDH-deficient), and
MLN0128) when overexpressed in MDA-MB-468 cells and compared HGF/MET mutations. These predictive markers assist in decision
with wild-type cells. These results suggest that untreated patient har- making in the treatment of renal cell carcinomas using mTOR inhibitors
boring hyperactivated mTOR with a single point mutation in the kinase [214]. Next-generation sequencing and other advanced technologies
domain may likely to show low/no sensitivity to mTOR kinase in- found new landscapes of genomics and epigenomics that facilitated
hibitors (ATP-competitive mTOR inhibitors). Further pave the way for approval and use of antiangiogenic and mTOR-directed targeted
using third generation inhibitor which overrides the resistance to the therapies for metastatic well-differentiated gastroenteropancreatic
currently available first and second generation inhibitors, rapalogs and neuroendocrine tumors [215]. Anaplastic thyroid cancer showed
TOTKis, respectively [187]. Wu et al predicted that the gatekeeper re- ∼10% of co-occurrence of MAPK and PI3K alterations and these tumors
sidue is likely to cause resistance to ATP-competitive kinase inhibitions exhibited a unique transcriptional signature. A case with PIK3CA and
as it is widely observed in other kinases. For example, in NSCLCs, BRAF mutations presented no sensitivity to inhibitors of combined
T790M mutation, a “gatekeeper” site of EGFR which constrains the mTOR/PI3K or RAF/MEK. Nevertheless, combining both pathway in-
erlotinib binding as the methionine is larger at this residual change hibitors exhibited dramatic therapeutic response [216]. Activation of
[206] and T315I resist the imatinib-mediated ABL kinase inhibition in both MAPK and Akt/mTOR signaling pathways were shown to be
CML (chronic myeloid leukemia) [207]. Similarly, it has also been concordantly maintained in the different regions of the same biopsy
predicted that the mTOR residue I2237 is the gatekeeper as evidenced [217]. Moreover, SPOP mutation was shown to activate prostate cancer
by the sequence alignment [208]. Nonetheless, after a series of ex- via PI3K/mTOR signaling [218]. A patient with bi-phenotypic breast
periments with yeast system, they found that the mTOR's gatekeeper cancer having STK11 mutation was shown to have an exceptional re-
position (I2237L) fails to endure any changes in the amino acid residue sponse to everolimus [219]. Genomic profiling identified ∼25% of
except the leucine, a highly conserved mutation [209]. These results cases with genetic alteration of mTOR in esophageal squamous cell
suggested that mTOR may not have drug-resistant gatekeeper muta- carcinoma and exhibited the base for developing targeted therapy of
tions. The same group continued to find drug-resistant mTOR mutation mTOR in precision medicine [220]. Recently, it has been demonstrated
in the kinase domain. To make mTOR kinase mutants, a library of that PI3K inhibitors could induce insulin feedback loop that reactivates
mutant clones were generated by error-prone PCR and recombined with the PI3K-mTOR signaling axis in several tumor models and inhibition of
a gapped TOR2 vector via “gap-repair” that results in the in-frame fu- this feedback by a dietary and pharmaceutical approach that enhance
sions of TOR2-mTOR via homologous recombination in yeast. the efficacy of PI3K inhibitors suggesting the utility of this work in
Screening the mutants towards various drugs identified varieties of enhancing treatment effect [221]. It has been found that oral squamous
mutants among them, two distinct mutants (L2185A and L2185C) cell carcinoma cell lines harboring HRAS mutation were not sensitive to
conferred more resistant to OSI-027, PF-04691502, PKI-587, INK128 inhibitors of PI3K. Consistently, HRAS overexpression exhibited low
and AZD8055 while other mutants such as L2185D, L2185N, and sensitivity to inhibition of PI3K, while silencing showed sensitivity.
L2185G were moderately resistant to AZD8055, PF-04691502, PKI-587 Interestingly, in both wild-type and HRAS mutant cell lines, PI3K in-
and AZD8055, INK128, PF-04691502, PKI-587, respectively. CRISPR/ hibition resulted in reduced p-Akt, but cell lines harboring HRAS

15
A.K. Murugan Seminars in Cancer Biology xxx (xxxx) xxx–xxx

mutations displayed intact rpS6 phosphorylation. Subsequently, clinical trial in advanced cancer (NCT01266057, https://clinicaltrials.
treating with mTOR inhibitor resulted in remarkable sensitivity in cells gov/). These therapeutic regimens might be very common in the future
carrying HRAS mutation. This group later found that ERK-TSC2-mTOR clinic for advanced metastatic cancers.
axis is the mediator of PI3K inhibitor resistance [222]. Whole genome
sequencing of pediatric high-grade gliomas (HGG) found mutations in 15. Conclusions and future prospective
TP53, TSC1, and amplification of MYCN. High throughput drug
screening revealed remarkable sensitivity to mTOR inhibitors while no The mTOR and its signaling pathway are implicated in a wide range
sensitivity was observed with SHH pathway inhibitors [223]. Recently, of normal physiological functions. In a variety of human cancers, the
platelet-derived growth factor receptor alpha (PDGFRA) extracellular mTOR signaling pathway is commonly deregulated either due to the
domain driver mutations were demonstrated to be resistant to PDGFRA genomic alterations resulting in hyperactivation of an upstream/
targeted therapies, but enhanced sensitivity to PI3K/mTOR and MEK downstream signaling pathway or hyperactivation of mTOR that in turn
inhibitors [224]. Frequent activation of the mTOR signaling pathway is provides an analogous impact. mTOR mutations have essential im-
obtaining immense popularity as its pathway genes are significantly plication not only in the pathogenesis of cancer but also in inducing
activated by somatic mutation and/or amplification and de novo over- resistance to the mTOR inhibitors. These spectra of activating me-
expression of proteins. These genetic aberrations ultimately lead to chanisms of mTOR cooperatively play a fundamental role in cell
deregulation of this pathway that results in malignant tumorigenesis transformation, growth, survival, increasing tumor volume, neovascu-
and necessitating targeted therapy for effective anticancer treatment. larization, and invasion and metastasis and drug resistance, etc.
Many mTOR inhibitors are currently in pre-clinical or clinical trials for Therefore, mTOR is an important therapeutic target. Clinical responses
both hematopoietic and solid cancer treatment while some mTOR in- to mTOR inhibitors as monotherapy have been modest and suggest
hibitors are already in the clinic. As several inhibitors are under de- combination therapies to overcome the resistance mechanisms.
velopmental process, advancing studies toward predictive biomarkers Although the role of mTOR overexpression is clearer as it could result in
are critically useful to determine the specificity of any kind of ther- a gain-of-function, how the downregulation of mTOR pave the way for
apeutic intervention. carcinogenesis is not yet well understood and warrant future studies to
Rapalogs (sirolimus, temsirolimus, and everolimus) are approved uncover these hidden mechanisms. Moreover, the copy number al-
for cancer treatments; nonetheless, rapalogs do not seem to be effective terations were very minimal while most of the studies explored only the
for the majority of solid cancers, particularly in advanced cancer [225]. expression of mTOR gene. As treatment strategies are advancing and
Alternatively, vorinostat may be combined with any one of those ra- continuously being developed towards precision medicine, it is very
palogs for better efficacy in the advanced cancer therapeutics and both crucial to identify various potential predictive biomarkers for treatment
drugs likely to be patient-friendly as the former one is toxic selectively deploying mTOR inhibitors. This would facilitate the effective design of
on cancer cells [226] and the later one (rapalog) is not a toxic drug at clinical trials that would accelerate the introduction of these com-
relevant doses and rapalogs do not kill cells but rather they slow down pounds to clinical practice. Hence, patients may benefit from a tailor-
their proliferation, growth, and motility [227]. Vorinostat (rINN) or made precision therapy by various mTOR inhibitors alone or in com-
suberoylanilide hydroxamic acid (SAHA) is a potent reversible new bination with other molecular targeted therapies.
class of pan-histone deacetylase (class I and II but not III HDAC) in-
hibitor. It functions by binding to the active site of histone deacetylases Funding
and acting as a chelator of zinc ions. It is recently approved by the FDA
and available in the drug store for the therapeutic use of cutaneous T- This study was not funded.
cell lymphoma (CTCL), a type of skin cancer. The combination of these
drugs for the treatment of different types of human cancer is currently Declaration of Competing Interest
being investigated.
Though the temsirolimus exhibits promising anticancer activity for Author, AK Murugan has no conflict of interest to declare except
the treatment of advanced RCCs, some cases develop survivin expres- that he is the guest editor of this special issue, “PI3K/Akt Signaling in
sion-mediated drug resistance. To this end, the anticancer activity of Human Cancer” in Seminars in Cancer Biology.
temsirolimus was synergistically improved by vorinostat in a panel of
RCC cell lines in vitro and in vivo. The combination of these drugs Acknowledgments
dramatically reduced survivin expression, induced apoptosis and
strongly suppressed angiogenesis while every single agent showed a I thank all the researchers in this field whose valuable work pro-
modest effect [228]. Currently, a phase I trial of sirolimus or everolimus vided me a delightful reading experience and apologize to my collea-
or temsirolimus in combination with vorinostat is clinically studied in gues whose work could not be cited primarily due to space constraints.
advanced cancer and yet to arrive any conclusion (NCT01174199). In
contrast, a phase I clinical trial that treated patients using vorinostat in References
combination with temsirolimus on metastatic prostate cancers (adeno-
carcinoma of the prostate, hormone-resistant prostate cancer, recurrent [1] C. Vézina, A. Kudelski, S.N. Sehgal, Rapamycin (AY-22,989), a new antifungal
prostate cancer, and stage IV prostate cancer) revealed no value in antibiotic. I. Taxonomy of the producing streptomycete and isolation of the active
principle, J. Antibiot. (Tokyo) 28 (1975) 721–726.
finding efficacy and hence it was terminated (NCT01174199). These [2] J. Heitman, N.R. Movva, M.N. Hall, Targets for cell cycle arrest by the im-
results suggest that the rapalog and vorinostat combination therapy munosuppressant rapamycin in yeast, Science 253 (1991) 905–909.
have differential effects on various types of human cancer and these [3] E.J. Brown, M.W. Albers, T.B. Shin, K. Ichikawa, C.T. Keith, W.S. Lane,
S.L. Schreiber, A mammalian protein targeted by G1-arresting rapamycin-receptor
inhibitors to be tested also on other types of human cancer prior to complex, Nature 369 (1994) 756–758.
concluding the efficacy of this combinatorial therapy. In the future [4] D.M. Sabatini, H. Erdjument-Bromage, M. Lui, P. Tempst, S.H. Snyder, RAFT1: a
precision medicine, a genetic-guided “three-drug” combination may mammalian protein that binds to FKBP12 in a rapamycin-dependent fashion and is
homologous to yeast TORs, Cell 78 (1994) 35–43.
also be sought as anticancer therapy mainly for the metastatic cancers.
[5] C.J. Sabers, M.M. Martin, G.J. Brunn, J.M. Williams, F.J. Dumont, G. Wiederrecht,
For example, patients are currently recruited for phase I study using a R.T. Abraham, Isolation of a protein target of the FKBP12-rapamycin complex in
combination of vorinostat and temsirolimus with conventional radia- mammalian cells, J. Biol. Chem. 270 (1995) 815–822.
[6] D.M. Sabatini, Twenty-five years of mTOR: uncovering the link from nutrients to
tion therapy in diffuse intrinsic pontine glioma (DIPG) of children
growth, Proc. Natl. Acad. Sci. U. S. A. 114 (2017) 11818–11825.
(NCT02420613). Similarly, sirolimus, vorinostat and hydroxy- [7] R.A. Saxton, D.M. Sabatini, mTOR signaling in growth, metabolism, and disease,
chloroquine combinations are being clinically studied by a phase I Cell 168 (2017) 960–976.

16
A.K. Murugan Seminars in Cancer Biology xxx (xxxx) xxx–xxx

[8] H. Zhou, S. Huang, Role of mTOR signaling in tumor cell motility, invasion and insulin-like growth factor receptor and insulin receptor activation through the
metastasis, Curr. Protein Pept. Sci. 12 (2011) 30–42. tyrosine kinase activity of mTOR, Cell Res. 26 (2016) 46–65.
[9] D. Mossmann, S. Park, M.N. Hall, mTOR signalling and cellular metabolism are [42] E.L. Greer, A. Brunet, FOXO transcription factors at the interface between long-
mutual determinants in cancer, Nat. Rev. Cancer 18 (2018) 744–757. evity and tumor suppression, Oncogene 24 (2005) 7410–7425.
[10] D. Baretić, R.L. Williams, PIKKs–the solenoid nest where partners and kinases [43] A.K. Munirajan, A.K. Murugan, N. Tsuchida, Genetic deregulation of the PIK3CA
meet, Curr. Opin. Struct. Biol. 29 (2014) 134–142. oncogene in oral cancer, Cancer Lett. 338 (2013) 193–203.
[11] C.T. Keith, S.L. Schreiber, PIK-related kinases: DNA repair, recombination, and cell [44] J.R. Kovalski, A. Bhaduri, A.M. Zehnder, P.H. Neela, Y. Che, G.G. Wozniak,
cycle checkpoints, Science 270 (1995) 50–51. P.A. Khavari, The functional proximal proteome of oncogenic ras includes
[12] R. Loewith, E. Jacinto, S. Wullschleger, A. Lorberg, J.L. Crespo, D. Bonenfant, mTORC2, Mol. Cell (2019) pii: S1097-2765(18)31031-1.
et al., Two TOR complexes, only one of which is rapamycin sensitive, have distinct [45] J. Mathieu, D. Detraux, D. Kuppers, Y. Wang, C. Cavanaugh, S. Sidhu, et al.,
roles in cell growth control, Mol. Cell 10 (2002) 457–468. Folliculin regulates mTORC1/2 and WNT pathways in early human pluripotency,
[13] D.D. Sarbassov, S.M. Ali, D.H. Kim, D.A. Guertin, R.R. Latek, H. Erdjument- Nat. Commun. 10 (2019) 632.
Bromage, et al., Rictor, a novel binding partner of mTOR, defines a rapamycin- [46] J. Yan, R. Wang, T. Horng, mTOR is key to T cell transdifferentiation, Cell Metab.
insensitive and raptor-independent pathway that regulates the cytoskeleton, Curr. 29 (2019) 241–242.
Biol. 14 (2004) 1296–1302. [47] F. Janku, D.J. McConkey, D.S. Hong, R. Kurzrock, Autophagy as a target for an-
[14] H. Yang, D.G. Rudge, J.D. Koos, B. Vaidialingam, H.J. Yang, N.P. Pavletich, mTOR ticancer therapy, Nat. Rev. Clin. Oncol. 8 (2011) 528–539.
kinase structure, mechanism and regulation, Nature 497 (2013) 217–223. [48] J. Kim, K.L. Guan, mTOR as a central hub of nutrient signalling and cell growth,
[15] C.H. Aylett, E. Sauer, S. Imseng, D. Boehringer, M.N. Hall, N. Ban, T. Maier, Nat. Cell Biol. 21 (2019) 63–71.
Architecture of human mTOR complex 1, Science 351 (2016) 48–52. [49] L.D. Smith, C.J. Saunders, D.L. Dinwiddie, A.M. Atherton, N.A. Miller, S.E. Soden,
[16] D. Baretić, A. Berndt, Y. Ohashi, C.M. Johnson, R.L. Williams, Tor forms a dimer et al., Exome sequencing reveals de novo germline mutation of mammalian target
through an N-terminal helical solenoid with a complex topology, Nat. Commun. 7 of rapamycin (MTOR) in a patient with megalencephaly and intractable seizures,
(2016) 11016. J. Genomes Exomes 2 (2013) 63–72.
[17] H. Yang, J. Wang, M. Liu, X. Chen, M. Huang, D. Tan, et al., 4.4 A Resolution Cryo- [50] G. Baynam, A. Overkov, M. Davis, K. Mina, L. Schofield, R. Allcock, et al., A
EM structure of human mTOR Complex 1, Protein Cell 7 (2016) 878–887. germline MTOR mutation in Aboriginal Australian siblings with intellectual dis-
[18] E. Jacinto, R. Loewith, A. Schmidt, S. Lin, M.A. Rüegg, A. Hall, M.N. Hall, ability, dysmorphism, macrocephaly, and small thoraces, Am. J. Med. Genet. A
Mammalian TOR complex 2 controls the actin cytoskeleton and is rapamycin in- 167 (2015) 1659–1667.
sensitive, Nat. Cell Biol. 6 (2004) 1122–1128. [51] C. Mroske, K. Rasmussen, D.N. Shinde, R. Huether, Z. Powis, H.M. Lu, et al.,
[19] P. Liu, W. Gan, Y.R. Chin, K. Ogura, J. Guo, J. Zhang, et al., PtdIns(3,4,5)P3- Germline activating MTOR mutation arising through gonadal mosaicism in two
dependent activation of the mTORC2 kinase complex, Cancer Discov. 5 (2015) brothers with megalencephaly and neurodevelopmental abnormalities, BMC Med.
1194–1209. Genet. 16 (2015) 102.
[20] K.P. Wedaman, A. Reinke, S. Anderson, J. Yates 3rd, J.M. McCaffery, T. Powers, [52] G.M. Mirzaa, C.D. Campbell, N. Solovieff, C. Goold, L.A. Jansen, S. Menon, et al.,
Tor kinases are in distinct membrane-associated protein complexes in Association of MTOR mutations with developmental brain disorders, including
Saccharomyces cerevisiae, Mol. Biol. Cell 14 (2003) 1204–1220. megalencephaly, focal cortical dysplasia, and pigmentary mosaicism, JAMA
[21] S. Wullschleger, R. Loewith, M.N. Hall, TOR signaling in growth and metabolism, Neurol. 73 (2016) 836–845.
Cell 124 (2006) 471–484. [53] S. Moosa, H. Böhrer-Rabel, J. Altmüller, F. Beleggia, P. Nürnberg, Y. Li, et al.,
[22] D.D. Sarbassov, S.M. Ali, S. Sengupta, J.H. Sheen, P.P. Hsu, A.F. Bagley, et al., Smith-Kingsmore syndrome: a third family with the MTOR mutation c.5395G&
Prolonged rapamycin treatment inhibits mTORC2 assembly and Akt/PKB, Mol. amp;A p.(Glu1799Lys) and evidence for paternal gonadal mosaicism, Am. J. Med.
Cell 22 (2006) 159–168. Genet. A 173 (2017) 264–267.
[23] D.W. Lamming, L. Ye, P. Katajisto, M.D. Goncalves, M. Saitoh, D.M. Stevens, et al., [54] R.S. Møller, S. Weckhuysen, M. Chipaux, E. Marsan, V. Taly, E.M. Bebin, et al.,
Rapamycin-induced insulin resistance is mediated by mTORC2 loss and uncoupled Germline and somatic mutations in the MTOR gene in focal cortical dysplasia and
from longevity, Science 335 (2012) 1638–1643. epilepsy, Neurol. Genet. 2 (2016) e118.
[24] L.A. Julien, A. Carriere, J. Moreau, P.P. Roux, mTORC1-activated S6K1 phos- [55] R.J. Leventer, T. Scerri, A.P. Marsh, K. Pope, G. Gillies, W. Maixner, et al.,
phorylates Rictor on threonine 1135 and regulates mTORC2 signaling, Mol. Cell. Hemispheric cortical dysplasia secondary to a mosaic somatic mutation in MTOR,
Biol. 30 (2010) 908–921. Neurology 84 (2015) 2029–2032.
[25] M. Karuppasamy, B. Kusmider, T.M. Oliveira, C. Gaubitz, M. Prouteau, R. Loewith, [56] M. Nakashima, H. Saitsu, N. Takei, J. Tohyama, M. Kato, H. Kitaura, et al., Somatic
C. Schaffitzel, Cryo-EM structure of Saccharomyces cerevisiae target of rapamycin Mutations in the MTOR gene cause focal cortical dysplasia type IIb, Ann. Neurol.
complex 2, Nat. Commun. 8 (2017) 1729. 78 (2015) 375–386.
[26] C. Gaubitz, M. Prouteau, B. Kusmider, R. Loewith, TORC2 structure and function, [57] J.S. Lim, W.I. Kim, H.C. Kang, S.H. Kim, A.H. Park, E.K. Park, et al., Nat. Med. 21
Trends Biochem. Sci. 41 (2016) 532–545. (2015) 395–400.
[27] X. Chen, M. Liu, Y. Tian, J. Li, Y. Qi, D. Zhao, et al., Cryo-EM structure of human [58] N. Iqbal, N. Iqbal, Human epidermal growth factor receptor 2 (HER2) in cancers:
mTOR complex 2, Cell Res. 28 (2018) 518–528. overexpression and therapeutic implications, Mol. Biol. Int. 2014 (2014) 852748.
[28] S. Oldham, E. Hafen, Insulin/IGF and target of rapamycin signaling: a TOR de [59] J.A. McKay, L.J. Murray, S. Curran, V.G. Ross, C. Clark, G.I. Murray, et al.,
force in growth control, Trends Cell Biol. 13 (2003) 79–85. Evaluation of the epidermal growth factor receptor (EGFR) in colorectal tumours
[29] M. Cornu, V. Albert, M.N. Hall, mTOR in aging, metabolism, and cancer, Curr. and lymph node metastases, Eur. J. Cancer 38 (2002) 2258–2264.
Opin. Genet. Dev. 23 (2013) 53–62. [60] K.J. Hatanpaa, S. Burma, D. Zhao, A.A. Habib, Epidermal growth factor receptor in
[30] M.A. Lawlor, D.R. Alessi, PKB/Akt: a key mediator of cell proliferation, survival glioma: signal transduction, neuropathology, imaging, and radioresistance,
and insulin responses? J. Cell. Sci. 114 (2001) 2903–2910. Neoplasia 12 (2010) 675–684.
[31] S. Corvera, M.P. Czech, Direct targets of phosphoinositide 3-kinase products in [61] A.F. Gazdar, Activating and resistance mutations of EGFR in non-small-cell lung
membrane traffic and signal transduction, Trends Cell Biol. 8 (1998) 442–446. cancer: role in clinical response to EGFR tyrosine kinase inhibitors, Oncogene 28
[32] D.R. Alessi, S.R. James, C.P. Downes, A.B. Holmes, P.R. Gaffney, C.B. Reese, (Suppl. 1) (2009) S24–31.
P. Cohen, Characterization of a 3-phosphoinositide-dependent protein kinase [62] C.L. Corless, A. Schroeder, D. Griffith, A. Town, L. McGreevey, P. Harrell, et al.,
which phosphorylates and activates protein kinase Balpha, Curr. Biol. 7 (1997) PDGFRA mutations in gastrointestinal stromal tumors: frequency, spectrum and in
261–269. vitro sensitivity to imatinib, J. Clin. Oncol. 23 (2005) 5357–5364.
[33] K. Inoki, Y. Li, T. Zhu, J. Wu, K.L. Guan, TSC2 is phosphorylated and inhibited by [63] T.P. Fleming, A. Saxena, W.C. Clark, J.T. Robertson, E.H. Oldfield, S.A. Aaronson,
Akt and suppresses mTOR signaling, Nat. Cell Biol. 4 (2002) 648–657. I.U. Ali, Amplification and/or overexpression of platelet-derived growth factor
[34] P.E. Burnett, R.K. Barrow, N.A. Cohen, S.H. Snyder, D.M. Sabatini, RAFT1 phos- receptors and epidermal growth factor receptor in human glial tumors, Cancer
phorylation of the translational regulators p70 S6 kinase and 4E-BP1, Proc. Natl. Res. 52 (1992) 4550–4553.
Acad. Sci. U. S. A. 95 (1998) 1432–1437. [64] T. Yoshimoto, T. Kumabe, Y. Sohma, T. Kayama, T. Yamamoto, Amplification of
[35] L. Zhao, P.K. Vogt, Class I PI3K in oncogenic cellular transformation, Oncogene 27 alpha-platelet-derived growth factor receptor gene lacking an exon coding for a
(2008) 5486–5496. portion of the extracellular region in a primary brain tumor of glial origin,
[36] P. Rodriguez-Viciana, P.H. Warne, R. Dhand, B. Vanhaesebroeck, I. Gout, M.J. Fry, Oncogene 7 (1992) 627–633.
M.D. Waterfield, J. Downward, Phosphatidylinositol-3-OH kinase as a direct target [65] J.S. Smith, X.Y. Wang, J. Qian, S.M. Hosek, B.W. Scheithauer, R.B. Jenkins,
of Ras, Nature 370 (1994) 527–532. C.D. James, Amplification of the platelet-derived growth factor receptor-A
[37] P. Rodriguez-Viciana, P.H. Warne, B. Vanhaesebroeck, M.D. Waterfield, (PDGFRA) gene occurs in oligodendrogliomas with grade IV anaplastic features, J.
J. Downward, Activation of phosphoinositide 3-kinase by interaction with Ras and Neuropathol. Exp. Neurol. 59 (2000) 495–503.
by point mutation, EMBO J. 15 (1996) 2442–2451. [66] H. Arai, T. Ueno, A. Tangoku, S. Yoshino, T. Abe, S. Kawauchi, A. Oga, et al.,
[38] T.O. Chan, U. Rodeck, A.M. Chan, A.C. Kimmelman, S.E. Rittenhouse, Detection of amplified oncogenes by genome DNA microarrays in human primary
G. Panayotou, P.N. Tsichlis, Small GTPases and tyrosine kinases coregulate a esophageal squamous cell carcinoma: comparison with conventional comparative
molecular switch in the phosphoinositide 3-kinase regulatory subunit, Cancer Cell genomic hybridization analysis, Cancer Genet. Cytogenet. 146 (2003) 16–21.
1 (2002) 1181–1191. [67] J. Zhao, J. Roth, B. Bode-Lesniewska, M. Pfaltz, P.U. Heitz, P. Komminoth,
[39] D.D. Sarbassov, D.A. Guertin, S.M. Ali, D.M. Sabatini, Phosphorylation and reg- Combined comparative genomic hybridization and genomic microarray for de-
ulation of Akt/PKB by the rictor–mTOR complex, Science 307 (2005) 1098–1101. tection of gene amplifications in pulmonary artery intimal sarcomas and adreno-
[40] K.E. O’Reilly, F. Rojo, Q.B. She, D. Solit, G.B. Mills, D. Smith, et al., mTOR in- cortical tumors, Genes Chromosomes Cancer 34 (2002) 48–57.
hibition induces upstream receptor tyrosine kinase signaling and activates Akt, [68] F. Toffalini, J.B. Demoulin, New insights into the mechanisms of hematopoietic
Cancer Res. 66 (2006) 1500–1508. cell transformation by activated receptor tyrosine kinases, Blood 116 (2010)
[41] Y. Yin, H. Hua, M. Li, S. Liu, Q. Kong, T. Shao, et al., mTORC2 promotes type I 2429–2437.

17
A.K. Murugan Seminars in Cancer Biology xxx (xxxx) xxx–xxx

[69] I. Ahmad, T. Iwata, H.Y. Leung, Mechanisms of FGFR-mediated carcinogenesis, [97] J.L. Nakamura, E. Garcia, R.O. Pieper, S6K1 plays a key role in glial transforma-
Biochim. Biophys. Acta 1823 (2012) 850–860. tion, Cancer Res. 68 (2008) 6516–6523.
[70] V. Rand, J. Huang, T. Stockwell, S. Ferriera, O. Buzko, S. Levy, et al., Sequence [98] N. Jafari, Q. Zheng, L. Li, W. Li, L. Qi, J. Xiao, et al., p70S6K1 (S6K1)-mediated
survey of receptor tyrosine kinases reveals mutations in glioblastomas, Proc. Natl. phosphorylation regulates phosphatidylinositol 4-phosphate 5-Kinase type I γ
Acad. Sci. U. S. A. 102 (2005) 14344–14349. degradation and cell invasion, J. Biol. Chem. 291 (2016) 25729–25741.
[71] B.W. van Rhijn, A.A. van Tilborg, I. Lurkin, J. Bonaventure, A. de Vries, [99] I. Segatto, S. Massarut, R. Boyle, G. Baldassarre, D. Walker, B. Belletti, Preclinical
J.P. Thiery, et al., Novel fibroblast growth factor receptor 3 (FGFR3) mutations in validation of a novel compound targeting p70S6 kinase in breast cancer, Aging
bladder cancer previously identified in non-lethal skeletal disorders, Eur. J. Hum. (Albany N. Y.) 8 (2016) 958–976.
Genet. 10 (2002) 819–824. [100] G. Armengol, F. Rojo, J. Castellví, C. Iglesias, M. Cuatrecasas, B. Pons, et al., 4E-
[72] C. Hunter, C. Greenman, P. Stephens, R. Smith, G.L. Dalgliesh, G. Bignell, et al., binding protein 1: a key molecular “funnel factor” in human cancer with clinical
Patterns of somatic mutation in human cancer genomes, Nature 446 (2007) implications, Cancer Res. 67 (2007) 7551–7555.
153–158. [101] A.C. Hsieh, H.G. Nguyen, L. Wen, M.P. Edlind, P.R. Carroll, W. Kim, D. Ruggero,
[73] S. de Muga, S. Hernández, L. Agell, N. Juanpere, R. Esgueva, J.A. Lorente, et al., Cell type-specific abundance of 4EBP1 primes prostate cancer sensitivity or re-
FGFR3 mutations in prostate cancer: association with low-grade tumors, Mod. sistance to PI3K pathway inhibitors, Sci. Signal. 8 (2015) ra116.
Pathol. 22 (2009) 848–856. [102] T. Zhang, J. Guo, H. Li, J. Wang, Meta-analysis of the prognostic value of p-4EBP1
[74] A. Dutt, H.B. Salvesen, T.H. Chen, A.H. Ramos, R.C. Onofrio, C. Hatton, et al., in human malignancies, Oncotarget 9 (2017) 2761–2769.
Drug-sensitive FGFR2 mutations in endometrial carcinoma, Proc. Natl. Acad. [103] O. Ichiyanagi, S. Naito, H. Ito, T. Kabasawa, T. Narisawa, H. Kanno, et al., Levels of
Sci.U. S. A. 105 (2008) 8713–8717. 4EBP1/eIF4E activation in renal cell carcinoma could differentially predict its
[75] M. Chesi, L.A. Brents, S.A. Ely, C. Bais, D.F. Robbiani, E.A. Mesri, et al., Activated early and late recurrence, Clin. Genitourin. Cancer 16 (2018) e1029-e58.
fibroblast growth factor receptor 3 is an oncogene that contributes to tumor [104] A. De Benedetti, J.R. Graff, eIF-4E expression and its role in malignancies and
progression in multiple myeloma, Blood 97 (2001) 729–736. metastases, Oncogene 23 (2004) 3189–3199.
[76] A. Chou, N. Dekker, R.C. Jordan, Identification of novel fibroblast growth factor [105] L.S. D’Abronzo, P.M. Ghosh, eIF4E phosphorylation in prostate cancer, Neoplasia
receptor 3 gene mutations in actinic cheilitis and squamous cell carcinoma of the 20 (2018) 563–573.
lip, Oral Surg. Oral Med. Oral Pathol. Oral Radiol. Endod. 107 (2009) 535–541. [106] M.L. Alabdullah, D.A. Ahmad, P. Moseley, S. Madhusudan, S. Chan, E. Rakha, The
[77] H. Davies, C. Hunter, R. Smith, P. Stephens, C. Greenman, G. Bignell, et al., mTOR downstream regulator (p-4EBP1) is a novel independent prognostic marker
Somatic mutations of the protein kinase gene family in human lung cancer, Cancer in ovarian cancer, J. Obstet. Gynaecol. (2019) 1–7.
Res. 65 (2005) 7591–7595. [107] J.V. Kavitha, F.J. Rosario, M.J. Nijland, T.J. McDonald, G. Wu, Y. Kanai, et al.,
[78] P.M. Pollock, M.G. Gartside, L.C. Dejeza, M.A. Powell, M.A. Mallon, H. Davies, Down-regulation of placental mTOR, insulin/IGF-I signaling, and nutrient trans-
et al., Frequent activating FGFR2 mutations in endometrial carcinomas parallel porters in response to maternal nutrient restriction in the baboon, FASEB J. 28
germline mutations associated with craniosynostosis and skeletal dysplasia syn- (2014) 1294–1305.
dromes, Oncogene 26 (2007) 7158–7162. [108] J. Brugarolas, K. Lei, R.L. Hurley, B.D. Manning, J.H. Reiling, E. Hafen, et al.,
[79] D. Cappellen, C. De Oliveira, D. Ricol, S. de Medina, J. Bourdin, X. Sastre-Garau, Regulation of mTOR function in response to hypoxia by REDD1 and the TSC1/
et al., Frequent activating mutations of FGFR3 in human bladder and cervix car- TSC2 tumor suppressor complex, Genes Dev. 18 (2004) 2893–2904.
cinomas, Nat. Genet. 23 (1999) 18–20. [109] A. Sofer, K. Lei, C.M. Johannessen, L.W. Ellisen, Regulation of mTOR and cell
[80] L. Cheng, S. Zhang, D.D. Davidson, G.T. MacLennan, M.O. Koch, R. Montironi, growth in response to energy stress by REDD1, Mol. Cell. Biol. 25 (2005)
A. Lopez-Beltran, Molecular determinants of tumor recurrence in the urinary 5834–5845.
bladder, Future Oncol. 5 (2009) 843–857. [110] C. Malagelada, E.J. Ryu, S.C. Biswas, V. Jackson-Lewis, L.A. Greene, RTP801 is
[81] J. Jacquemier, J. Adelaide, P. Parc, F. Penault-Llorca, J. Planche, O. deLapeyriere, elevated in Parkinson brain substantia nigral neurons and mediates death in cel-
D. Birnbaum, Expression of the FGFR1 gene in human breast-carcinoma cells, Int. lular models of Parkinson’s disease by a mechanism involving mammalian target
J. Cancer 59 (1994) 373–378. of rapamycin inactivation, J. Neurosci. 26 (2006) 9996–10005.
[82] K.B. Meyer, A.T. Maia, M. O’Reilly, A.E. Teschendorff, S.F. Chin, C. Caldas, [111] S. Vega-Rubin-de-Celis, Z. Abdallah, L. Kinch, N.V. Grishin, J. Brugarolas,
B.A. Ponder, Allele-specific up-regulation of FGFR2 increases susceptibility to X. Zhang, Structural analysis and functional implications of the negative mTORC1
breast cancer, PLoS Biol. 6 (2008) e108. regulator REDD1, Biochemistry 49 (2010) 2491–2501.
[83] K. Chin, S. DeVries, J. Fridlyand, P.T. Spellman, R. Roydasgupta, W.L. Kuo, et al., [112] P. Bonaccorso, M. Tumino, A. D’Ambra, E. Mirabile, M. Barchitta, A. Poli,
Genomic and transcriptional aberrations linked to breast cancer pathophysiolo- D. Bottino, M. La Rosa, A. Agodi, L.L. Nigro, Downregulation of mTOR and
gies, Cancer Cell 10 (2006) 529–541. P70S6Kβ2 in pediatric T-cell acute lymphoblastic leukemia (T-ALL) is correlated
[84] V. Gelsi-Boyer, B. Orsetti, N. Cervera, P. Finetti, F. Sircoulomb, C. Rougé, et al., with a poor prognosis, Blood 118 (2011) 2508.
Comprehensive profiling of 8p11-12 amplification in breast cancer, Mol. Cancer [113] A. Sekulić, C.C. Hudson, J.L. Homme, P. Yin, D.M. Otterness, L.M. Karnitz,
Res. 3 (2005) 655–667. R.T. Abraham, A direct linkage between the phosphoinositide 3-kinase-AKT sig-
[85] J.H. Jang, K.H. Shin, J.G. Park, Mutations in fibroblast growth factor receptor 2 naling pathway and the mammalian target of rapamycin in mitogen-stimulated
and fibroblast growth factor receptor 3 genes associated with human gastric and and transformed cells, Cancer Res. 60 (2000) 3504–3513.
colorectal cancers, Cancer Res. 61 (2001) 3541–3543. [114] A.L. Edinger, C.B. Thompson, An activated mTOR mutant supports growth factor-
[86] F. Yagasaki, D. Wakao, Y. Yokoyama, Y. Uchida, I. Murohashi, H. Kayano, et al., independent, nutrient-dependent cell survival, Oncogene 23 (2004) 5654–5663.
Fusion of ETV6 to fibroblast growth factor receptor 3 in peripheral T-cell lym- [115] A. Reinke, J.C. Chen, P.T. Aronova, Caffeine targets TOR complex I and provides
phoma with a t(4;12)(p16;p13) chromosomal translocation, Cancer Res. 61 (2001) evidence for a regulatory link between the FRB and kinase domains of Tor1p, J.
8371–8374. Biol. Chem. 281 (2006) 31616–31626.
[87] S. Roumiantsev, D.S. Krause, C.A. Neumann, C.A. Dimitri, F. Asiedu, N.C. Cross, [116] J. Urano, T. Sato, T. Matsuo, Y. Otsubo, M. Yamamoto, F. Tamanoi, Point muta-
R.A. Van Etten, Distinct stem cell myeloproliferative/T lymphoma syndromes in- tions in TOR confer Rheb-independent growth in fission yeast and nutrient-in-
duced by ZNF198-FGFR1 and BCR-FGFR1 fusion genes from 8p11 translocations, dependent mammalian TOR signaling in mammalian cells, Proc. Natl. Acad. Sci. U.
Cancer Cell 5 (2004) 287–298. S. A. 104 (2007) 3514–3519.
[88] M. Ren, X. Li, J.K. Cowell, Genetic fingerprinting of the development and pro- [117] Y. Ohne, T. Takahara, R. Hatakeyama, T. Matsuzaki, M. Noda, N. Mizushima,
gression of T-cell lymphoma in a murine model of atypical myeloproliferative T. Maeda, Isolation of hyperactive mutants of mammalian target of rapamycin, J.
disorder initiated by the ZNF198–fibroblast growth factor receptor-1 chimeric Biol. Chem. 283 (2008) 31861–31870.
tyrosine kinase, Blood 114 (2009) 1576–1584. [118] T. Sato, A. Nakashima, L. Guo, K. Coffman, F. Tamanoi, Single amino acid changes
[89] R. Arafeh, Y. Samuels, PIK3CA in cancer: the past 30 years, Semin. Cancer Biol. that confer constitutive activation of mTOR are discovered in human cancer,
(2019) Feb 10. pii: S1044-579X(18)30152-4. Oncogene 29 (2010) 2746–2752.
[90] H.M. Wise, M.A. Hermida, N.R. Leslie, Prostate cancer, PI3K, PTEN and prognosis, [119] A.K. Murugan, A. Alzahrani, M. Xing, Mutations in critical domains confer the
Clin. Sci. (Lond) 131 (2017) 197–210. human mTOR gene strong tumorigenicity, J. Biol. Chem. 288 (2013) 6511–6521.
[91] J. Downward, Targeting RAS signalling pathways in cancer therapy, Nat. Rev. [120] H. Davies, C. Hunter, R. Smith, P. Stephens, C. Greenman, G. Bignell, et al.,
Cancer 3 (2003) 11–22. Somatic mutations of the protein kinase gene family in human lung cancer, Cancer
[92] L. Ma, Z. Chen, H. Erdjument-Bromage, P. Tempst, P.P. Pandolfi, Phosphorylation Res. 65 (2005) 7591–7595.
and functional inactivation of TSC2 by Erk implications for tuberous sclerosis and [121] R. McLendon, A. Friedman, D. Bigner, E.G. Van Meir, D.J. Brat,
cancer pathogenesis, Cell 121 (2005) 179–193. G.M. Mastrogianakis, et al., Cancer Genome Atlas Research Network
[93] B.A. Ballif, P.P. Roux, S.A. Gerber, J.P. MacKeigan, J. Blenis, S.P. Gygi, Comprehensive genomic characterization defines human glioblastoma genes and
Quantitative phosphorylation profiling of the ERK/p90 ribosomal S6 kinase-sig- core pathways, Nature 455 (2008) 1061–1068.
naling cassette and its targets, the tuberous sclerosis tumor suppressors, Proc. Natl. [122] G.L. Dalgliesh, K. Furge, C. Greenman, L. Chen, G. Bignell, A. Butler, et al.,
Acad. Sci. U. S. A. 102 (2005) 667–672. Systematic sequencing of renal carcinoma reveals inactivation of histone-mod-
[94] B.D. Manning, A.R. Tee, M.N. Logsdon, J. Blenis, L.C. Cantley, Identification of the ifying genes, Nature 463 (2010) 360–363.
tuberous sclerosis complex-2 tumor suppressor gene product tuberin as a target of [123] C.M. Robbins, W.A. Tembe, A. Baker, S. Sinari, T.Y. Moses, S. Beckstrom-
the phosphoinositide 3-kinase/akt pathway, Mol. Cell 10 (2002) 151–162. Sternberg, et al., Copy number and targeted mutational analysis reveals novel
[95] A. Carrière, M. Cargnello, L.A. Julien, H. Gao, E. Bonneil, P. Thibault, P.P. Roux, somatic events in metastatic prostate tumors, Genome Res. 21 (2011) 47–55.
Oncogenic MAPK signaling stimulates mTORC1 activity by promoting RSK- [124] Y. Jiao, C. Shi, B.H. Edil, R.F. de Wilde, D.S. Klimstra, A. Maitra, et al., DAXX/
mediated raptor phosphorylation, Curr. Biol. 18 (2008) 1269–1277. ATRX, MEN1, and mTOR pathway genes are frequently altered in pancreatic
[96] M. Bärlund, F. Forozan, J. Kononen, L. Bubendorf, Y. Chen, M.L. Bittner, et al., neuroendocrine tumors, Science 331 (2011) 1199–1203.
Detecting activation of ribosomal protein S6 kinase by complementary DNA and [125] Y.Y. Li, G.J. Hanna, A.C. Laga, R.I. Haddad, J.H. Lorch, P.S. Hammerman,
tissue microarray analysis, J. Natl. Cancer Inst. 92 (2000) 1252–1259. Genomic analysis of metastatic cutaneous squamous cell carcinoma, Clin. Cancer

18
A.K. Murugan Seminars in Cancer Biology xxx (xxxx) xxx–xxx

Res. 21 (2015) 1447–1456. [154] F. Zhang, Y. Wang, P. Sun, Z.Q. Wang, D.S. Wang, D.S. Zhang, et al., Fibrinogen
[126] A.K. Murugan, E.A. Humudh, E. Qasem, H. Al-Hindi, M. Almohanna, Z.K. Hassan, promotes malignant biological tumor behavior involving epithelial-mesenchymal
A.S. Alzahrani, Absence of somatic mutations of the mTOR gene in differentiated transition via the p-AKT/p-mTOR pathway in esophageal squamous cell carci-
thyroid cancer, Meta Gene 6 (2015) 69–71. noma, J. Cancer Res. Clin. Oncol. 143 (2017) 2413–2424.
[127] M. Gerlinger, A.J. Rowan, S. Horswell, J. Larkin, D. Endesfelder, E. Gronroos, [155] H. Zhou, H. Zong, B. Yin, D. Cai, B. Ma, Y. Xiang, Inhibition of mTOR pathway
et al., Intratumor heterogeneity and branched evolution revealed by multiregion attenuates migration and invasion of gallbladder cancer via EMT inhibition, Mol.
sequencing, N. Engl. J. Med. 366 (2012) 883–892. Biol. Rep. 41 (2014) 4507–4512.
[128] J.W. Kunstman, C.C. Juhlin, G. Goh, T.C. Brown, A. Stenman, J.M. Healy, et al., [156] N. Sharma, R. Nanta, J. Sharma, S. Gunewardena, K.P. Singh, S. Shankar,
Characterization of the mutational landscape of anaplastic thyroid cancer via R.K. Srivastava, PI3K/AKT/mTOR and sonic hedgehog pathways cooperate to-
whole-exome sequencing, Hum. Mol. Genet. 24 (2015) 2318–2329. gether to inhibit human pancreatic cancer stem cell characteristics and tumor
[129] Y. Li, Y. Kong, L. Si, X. Wu, X. Xu, J. Dai, et al., Analysis of mTOR gene aberrations growth, Oncotarget 6 (2015) 32039–32060.
in melanoma patients and evaluation of their sensitivity to PI3K-AKT-mTOR [157] G. Hudes, M. Carducci, P. Tomczak, J. Dutcher, R. Figlin, A. Kapoor, et al.,
pathway inhibitors, Clin. Cancer Res. 22 (2016) 1018–1027. Temsirolimus, interferon alfa, or both for advanced renal-cell carcinoma, N. Engl.
[130] T. Sato, A. Nakashima, L. Guo, K. Coffman, F. Tamanoi, Single amino-acid changes J. Med. 356 (2007) 2271–2281.
that confer constitutive activation of mTOR are discovered in human cancer, [158] R.J. Motzer, B. Escudier, S. Oudard, T.E. Hutson, C. Porta, S. Bracarda, et al.,
Oncogene 29 (2010) 2746–2752. Efficacy of everolimus in advanced renal cell carcinoma: a double-blind, rando-
[131] H. Yamaguchi, M. Kawazu, T. Yasuda, M. Soda, T. Ueno, S. Kojima, et al., mised, placebo-controlled phase III trial, Lancet 372 (2008) 449–456.
Transforming somatic mutations of mammalian target of rapamycin kinase in [159] J. Dominguez, K. Mahalati, B. Kiberd, V.C. McAlister, A.S. MacDonald, Conversion
human cancer, Cancer Sci. 106 (2015) 1687–1692. to rapamycin immunosuppression in renal transplant recipients: report of an in-
[132] A.P. Ghosh, C.B. Marshall, T. Coric, E.H. Shim, R. Kirkman, M.E. Ballestas, et al., itial experience, Transplantation 70 (2000) 1244–1247.
Point mutations of the mTOR-RHEB pathway in renal cell carcinoma, Oncotarget 6 [160] E. Caron, S. Ghosh, Y. Matsuoka, D. Ashton-Beaucage, M. Therrien, S. Lemieux,
(2015) 17895–17910. et al., A comprehensive map of the mTOR signaling network, Mol. Syst. Biol. 6
[133] B.C. Grabiner, V. Nardi, K. Birsoy, R. Possemato, K. Shen, S. Sinha, et al., A diverse (2010) 453.
array of cancer-associated MTOR mutations are hyperactivating and can predict [161] M.N. Stanfel, L.S. Shamieh, M. Kaeberlein, B.K. Kennedy, The TOR pathway comes
rapamycin sensitivity, Cancer Discov. 4 (2014) 554–563. of age, Biochim. Biophys. Acta 1790 (2009) 1067–1074.
[134] J. Xu, C.G. Pham, S.K. Albanese, Y. Dong, T. Oyama, C.H. Lee, et al., [162] A.Y. Choo, S.O. Yoon, S.G. Kim, P.P. Roux, J. Blenis, Rapamycin differentially
Mechanistically distinct cancer-associated mTOR activation clusters predict sen- inhibits S6Ks and 4E-BP1 to mediate cell-type-specific repression of mRNA
sitivity to rapamycin, J. Clin. Invest. 126 (2016) 3526–3540. translation, Proc. Natl. Acad. Sci. U. S. A. 105 (2008) 17414–17419.
[135] N. Wagle, B.C. Grabiner, E.M. Van Allen, A. Amin-Mansour, A. Taylor-Weiner, [163] H. Yang, X. Jiang, B. Li, H.J. Yang, M. Miller, A. Yang, A. Dhar, N.P. Pavletich,
M. Rosenberg, et al., Response and acquired resistance to everolimus in anaplastic Mechanisms of mTORC1 activation by RHEB and inhibition by PRAS40, Nature
thyroid cancer, N. Engl. J. Med. 371 (2014) 1426–1433. 552 (2017) 368–373.
[136] T.R. Peterson, M. Laplante, C.C. Thoreen, Y. Sancak, S.A. Kang, W.M. Kuehl, [164] N. Wagle, B.C. Grabiner, E.M. Van Allen, E. Hodis, S. Jacobus, J.G. Supko, et al.,
N.S. Gray, D.M. Sabatini, DEPTOR is an mTOR inhibitor frequently overexpressed Activating mTOR mutations in a patient with an extraordinary response on a phase
in multiple myeloma cells and required for their survival, Cell 137 (2009) I trial of everolimus and pazopanib, Cancer Discov. 4 (2014) 546–553.
873–886. [165] R.J. Motzer, B. Escudier, D.F. McDermott, S. George, H.J. Hammers, S. Srinivas,
[137] J.P. Thiery, Epithelial-mesenchymal transitions in tumour progression, Nat. Rev. et al., Nivolumab versus everolimus in advanced renal-cell carcinoma, N. Engl. J.
Cancer 2 (2002) 442–454. Med. 373 (2015) 1803–1813.
[138] A. Hall, Rho GTPases and the actin cytoskeleton, Science 279 (1998) 509–514. [166] J. Baselga, M. Campone, M. Piccart, H.A. Burris 3rd, H.S. Rugo, T. Sahmoud, et al.,
[139] A.C. Hsieh, Y. Liu, M.P. Edlind, N.T. Ingolia, M.R. Janes, A. Sher, et al., The Everolimus in postmenopausal hormone-receptor-positive advanced breast cancer,
translational landscape of mTOR signalling steers cancer initiation and metastasis, N. Engl. J. Med. 366 (2012) 520–529.
Nature 485 (2012) 55–61. [167] J.C. Yao, M.H. Shah, T. Ito, C.L. Bohas, E.M. Wolin, E. Van Cutsem, et al.,
[140] S. Lamouille, R. Derynck, Cell size and invasion in TGF-beta-induced epithelial to Everolimus for advanced pancreatic neuroendocrine tumors, N. Engl. J. Med. 364
mesenchymal transition is regulated by activation of the mTOR pathway, J. Cell (2011) 514–523.
Biol. 178 (2007) 437–451. [168] J.J. Bissler, J.C. Kingswood, E. Radzikowska, B.A. Zonnenberg, M. Frost,
[141] S. Lamouille, E. Connolly, J.W. Smyth, R.J. Akhurst, R. Derynck, TGF-β-induced E. Belousova, et al., Everolimus for angiomyolipoma associated with tuberous
activation of mTOR complex 2 drives epithelial-mesenchymal transition and cell sclerosis complex or sporadic lymphangioleiomyomatosis (EXIST-2): a multi-
invasion, J. Cell. Sci. 125 (2012) 1259–1273. centre, randomised, double-blind, placebo-controlled trial, Lancet 381 (2013)
[142] P. Gulhati, K.A. Bowen, J. Liu, P.D. Stevens, P.G. Rychahou, M. Chen, et al., 817–824.
mTORC1 and mTORC2 regulate EMT, motility, and metastasis of colorectal cancer [169] D.N. Franz, E. Belousova, S. Sparagana, E.M. Bebin, M. Frost, R. Kuperman, et al.,
via RhoA and Rac1 signaling pathways, Cancer Res. 71 (2011) 3246–3256. Efficacy and safety of everolimus for subependymal giant cell astrocytomas as-
[143] X. Chen, H. Cheng, T. Pan, Y. Liu, Y. Su, C. Ren, et al., mTOR regulate EMT sociated with tuberous sclerosis complex (EXIST-1): a multicentre, randomised,
through RhoA and Rac1 pathway in prostate cancer, Mol. Carcinog. 54 (2015) placebo-controlled phase 3 trial, Lancet 381 (2013) 125–132.
1086–1095. [170] G.D. Demetri, S.P. Chawla, I. Ray-Coquard, A. Le Cesne, A.P. Staddon,
[144] D. Wu, J. Cheng, G. Sun, S. Wu, M. Li, Z. Gao, et al., p70S6K promotes IL-6- M.M. Milhem, et al., Results of an international randomized phase III trial of the
induced epithelial-mesenchymal transition and metastasis of head and neck mammalian target of rapamycin inhibitor ridaforolimus versus placebo to control
squamous cell carcinoma, Oncotarget 7 (2016) 36539–36550. metastatic sarcomas in patients after benefit from prior chemotherapy, J. Clin.
[145] A. Kwasnicki, D. Jeevan, A. Braun, R. Murali, M. Jhanwar-Uniyal, Involvement of Oncol. 31 (2013) 2485–2492.
mTOR signaling pathways in regulating growth and dissemination of metastatic [171] A.M. Gonzalez-Angulo, F. Meric-Bernstam, S. Chawla, G. Falchook, D. Hong,
brain tumors via EMT, Anticancer Res. 35 (2015) 689–696. A. Akcakanat, et al., Weekly nab-rapamycin in patients with advanced non-
[146] Z. Yang, H. Xie, D. He, L. Li, Infiltrating macrophages increase RCC epithelial hematologic malignancies: final results of a phase I trial, Clin. Cancer Res. 19
mesenchymal transition (EMT) and stem cell-like populations via AKT and mTOR (2013) 5474–5484.
signaling, Oncotarget 7 (2016) 44478–44491. [172] K.H. Schreiber, D. Ortiz, E.C. Academia, A.C. Anies, C.Y. Liao, B.K. Kennedy,
[147] G. Lin, R. Gai, Z. Chen, Y. Wang, S. Liao, R. Dong, et al., The dual PI3K/mTOR Rapamycin-mediated mTORC2 inhibition is determined by the relative expression
inhibitor NVP-BEZ235 prevents epithelial-mesenchymal transition induced by of FK506-binding proteins, Aging Cell 14 (2015) 265–273.
hypoxia and TGF-β1, Eur. J. Pharmacol. 729 (2014) 45–53. [173] M. Mizunuma, E. Neumann-Haefelin, N. Moroz, Y. Li, T.K. Blackwell, mTORC2-
[148] X.M. Zhou, R. Sun, D.H. Luo, J. Sun, M.Y. Zhang, M.H. Wang, et al., Upregulated SGK-1 acts in two environmentally responsive pathways with opposing effects on
TRIM29 promotes proliferation and metastasis of nasopharyngeal carcinoma via longevity, Aging Cell 13 (2014) 869–878.
PTEN/AKT/mTOR signal pathway, Oncotarget 7 (2016) 13634–13650. [174] M. Toulany, T.A. Schickfluss, K.R. Fattah, K.J. Lee, B.P. Chen, B. Fehrenbacher,
[149] M.H. Wang, X.M. Zhou, M.Y. Zhang, L. Shi, R.W. Xiao, L.S. Zeng, et al., BMP2 et al., Function of erbB receptors and DNA-PKcs on phosphorylation of cyto-
promotes proliferation and invasion of nasopharyngeal carcinoma cells via plasmic and nuclear Akt at S473 induced by erbB1 ligand and ionizing radiation,
mTORC1 pathway, Aging (Albany N. Y.) 9 (2017) 1326–1340. Radiother. Oncol. 101 (2011) 140–146.
[150] Y. Zhao, Y. Xing, Q. Meng, X. Chen, W. Liu, J. Hu, et al., TRIM44 promotes pro- [175] S.M. Joung, Z.Y. Park, S. Rani, O. Takeuchi, S. Akira, J.Y. Lee, Akt contributes to
liferation and metastasis in non-small cell lung cancer via mTOR signaling activation of the TRIF-dependent signaling pathways of TLRs by interacting with
pathway, Oncotarget 7 (2016) 30479–30491. TANK-binding kinase 1, J. Immunol. 186 (2011) 499–507.
[151] B. Zhang, C. Deng, L. Wang, F. Zhou, S. Zhang, W. Kang, et al., Upregulation of [176] Q. Fan, O. Aksoy, R.A. Wong, S. Ilkhanizadeh, C.J. Novotny, W.C. Gustafson, et al.,
UBE2Q1 via gene copy number gain in hepatocellular carcinoma promotes cancer A kinase inhibitor targeted to mTORC1 drives regression in glioblastoma, Cancer
progression through β-catenin-EGFR-PI3K-Akt-mTOR signaling pathway, Mol. Cell 31 (2017) 424–435.
Carcinog. 57 (2018) 201–215. [177] D.M. Sabatini, mTOR and cancer: insights into a complex relationship, Nat. Rev.
[152] X. Tan, X. Liu, P. Liu, Y. Wu, S. Qian, X. Zhang, Phosphoglycerate mutase 1 Cancer 6 (2006) 729–734.
(PGAM1) promotes pancreatic ductal adenocarcinoma (PDAC) metastasis by [178] P.P. Hsu, S.A. Kang, J. Rameseder, Y. Zhang, K.A. Ottina, D. Lim, et al., The mTOR-
acting as a novel downstream target of the PI3K/Akt/mTOR pathway, Oncol. Res. regulated phosphoproteome reveals a mechanism of mTORC1-mediated inhibition
26 (2018) 1123–1131. of growth factor signaling, Science 332 (2011) 1317–1322.
[153] W. Wang, H. Chen, Z. Liu, P. Qu, J. Lan, H. Chen, et al., Regulator of cullins-1 [179] Y. Yu, S.O. Yoon, G. Poulogiannis, Q. Yang, X.M. Ma, J. Villén, et al.,
expression knockdown suppresses the malignant progression of muscle-invasive Phosphoproteomic analysis identifies Grb10 as an mTORC1 substrate that nega-
transitional cell carcinoma by regulating mTOR/DEPTOR pathway, Br. J. Cancer tively regulates insulin signaling, Science 332 (2011) 1322–1326.
114 (2016) 305–313. [180] L.C. Kim, R.S. Cook, J. Chen, mTORC1 and mTORC2 in cancer and the tumor

19
A.K. Murugan Seminars in Cancer Biology xxx (xxxx) xxx–xxx

microenvironment, Oncogene 36 (2017) 2191–2201. preventing interaction with FKBP12-rapamycin, J. Biol. Chem. 270 (1995)
[181] U. Banerji, E. Jane Dean, M. Gonzalez, A.P. Greystoke, B. Basu, M. KrebsMartina 27531–27537.
Puglisi, et al., First-in-human phase I trial of the dual mTORC1 and mTORC2 in- [206] S. Kobayashi, T.J. Boggon, T. Dayaram, P.A. Jänne, O. Kocher, M. Meyerson, et al.,
hibitor AZD2014 in solid tumors [abstract], J. Clin. Oncol. 30 (2012) 3004. EGFR mutation and resistance of non-small-cell lung cancer to gefitinib, N. Engl. J.
[182] F. Janku, T.A. Yap, F. Meric-Bernstam, Targeting the PI3K pathway in cancer: are Med. 352 (2005) 786–792.
we making headway? Nat. Rev. Clin. Oncol. 15 (2018) 273–291. [207] M.E. Gorre, M. Mohammed, K. Ellwood, N. Hsu, R. Paquette, P.N. Rao,
[183] C.M. Chresta, B.R. Davies, I. Hickson, T. Harding, S. Cosulich, S.E. Critchlow, C.L. Sawyers, Clinical resistance to STI-571 cancer therapy caused by BCR-ABL
et al., AZD8055 is a potent, selective, and orally bioavailable ATP-competitive gene mutation or amplification, Science 293 (2001) 876–880.
mammalian target of rapamycin kinase inhibitor with in vitro and in vivo anti- [208] T.W. Sturgill, M.N. Hall, Activating mutations in TOR are in similar structures as
tumor activity, Cancer Res. 70 (2010) 288–298. oncogenic mutations in PI3KCalpha, ACS Chem. Biol. 4 (2009) 999–1015.
[184] S.M. Maira, F. Stauffer, J. Brueggen, P. Furet, C. Schnell, C. Fritsch, et al., [209] T.J. Wu, X. Wang, Y. Zhang, L. Meng, J.E. Kerrigan, S.K. Burley, X.F. Zheng,
Identification and characterization of NVP-BEZ235, a new orally available dual Identification of a non-gatekeeper hot spot for drug-resistant mutations in mTOR
phosphatidylinositol 3-kinase/mammalian target of rapamycin inhibitor with kinase, Cell Rep. 11 (2015) 446–459.
potent in vivo antitumor activity, Mol. Cancer Ther. 7 (2008) 1851–1863. [210] A.K. Munirajan, A.K. Murugan, A.S. Alzahrani, Long noncoding RNAs: emerging
[185] L.I. Toledo, M. Murga, R. Zur, R. Soria, A. Rodriguez, S. Martinez, et al., A cell- players in thyroid cancer pathogenesis, Endocr. Relat. Cancer 25 (2018) R59–R82.
based screen identifies ATR inhibitors with synthetic lethal properties for cancer- [211] J.H. Park, M.H. Ryu, Y.S. Park, S.R. Park, Y.S. Na, B.Y. Rhoo, Y.K. Kang, Successful
associated mutations, Nat. Struct. Mol. Biol. 18 (2011) 721–727. control of heavily pretreated metastatic gastric cancer with the mTOR inhibitor
[186] V.S. Rodrik-Outmezguine, S. Chandarlapaty, N.C. Pagano, P.I. Poulikakos, everolimus (RAD001) in a patient with PIK3CA mutation and pS6 overexpression,
M. Scaltriti, E. Moskatel, et al., mTOR kinase inhibition causes feedback-depen- BMC Cancer 15 (2015) 119.
dent biphasic regulation of AKT signaling, Cancer Discov. 1 (2011) 248–259. [212] F. Meric-Bernstam, A. Akcakanat, H. Chen, K.A. Do, T. Sangai, F. Adkins, et al.,
[187] V.S. Rodrik-Outmezguine, M. Okaniwa, Z. Yao, C.J. Novotny, C. McWhirter, PIK3CA/PTEN mutations and Akt activation as markers of sensitivity to allosteric
A. Banaji, et al., Overcoming mTOR resistance mutations with a new-generation mTOR inhibitors, Clin. Cancer Res. 18 (2012) 1777–1789.
mTOR inhibitor, Nature 534 (2016) 272–276. [213] F. Janku, D.S. Hong, S. Fu, S.A. Piha-Paul, A. Naing, G.S. Falchook, et al., Assessing
[188] M. Lu, A.S. Zessin, W. Glover, D.S. Hsu, Activation of the mTOR pathway by ox- PIK3CA and PTEN in early-phase trials with PI3K/AKT/mTOR inhibitors, Cell Rep.
aliplatin in the treatment of colorectal cancer liver metastasis, PLoS One 12 (2017) 6 (2014) 377–387.
e0169439. [214] C. Ciccarese, M. Brunelli, R. Montironi, M. Fiorentino, R. Iacovelli, D. Heng, et al.,
[189] M.C. Zhen, F.Q. Wang, S.F. Wu, Y.L. Zhao, P.G. Liu, Z.Y. Yin, Identification of The prospect of precision therapy for renal cell carcinoma, Cancer Treat. Rev. 49
mTOR as a primary resistance factor of the IAP antagonist AT406 in hepatocellular (2016) 37–44.
carcinoma cells, Oncotarget 8 (2017) 9466–9475. [215] M.S. Lee, B.H. O’Neil, Summary of emerging personalized medicine in neu-
[190] S. Chatterjee, E.H. Huang, I. Christie, B.F. Kurland, T.F. Burns, Acquired resistance roendocrine tumors: are we on track? J. Gastrointest. Oncol. 7 (2016) 804–818.
to the Hsp90 inhibitor, Ganetespib, in KRAS-mutant NSCLC is mediated via re- [216] W.J. Gibson, D.T. Ruan, V.A. Paulson, J.A. Barletta, G.J. Hanna, S. Kraft, et al.,
activation of the ERK-p90RSK-mTOR signaling network, Mol. Cancer Ther. 16 Genomic heterogeneity and exceptional response to dual pathway inhibition in
(2017) 793–804. anaplastic thyroid Cancer, Clin. Cancer Res. 23 (2017) 2367–2373.
[191] W. Liu, J. Chang, M. Liu, J. Yuan, J. Zhang, J. Qin, et al., Quantitative proteomics [217] E.M. Parasido, A. Silvestri, V. Canzonieri, C. Belluco, M.G. Diodoro, M. Milione,
profiling reveals activation of mTOR pathway in trastuzumab resistance, et al., Protein drug target activation homogeneity in the face of intra-tumor het-
Oncotarget 8 (2017) 45793–45806. erogeneity: implications for precision medicine, Oncotarget 8 (2017)
[192] T. Mashima, Y. Taneda, M.K. Jang, A. Mizutani, Y. Muramatsu, H. Yoshida, et al., 48534–48544.
mTOR signaling mediates resistance to tankyrase inhibitors in Wnt-driven color- [218] M. Blattner, D. Liu, B.D. Robinson, D. Huang, A. Poliakov, D. Gao, et al., SPOP
ectal cancer, Oncotarget 8 (2017) 47902–47915. mutation drives prostate tumorigenesis in vivo through coordinate regulation of
[193] O. Lindblad, E. Cordero, A. Puissant, L. Macaulay, A. Ramos, N.N. Kabir, et al., PI3K/mTOR and AR signaling, Cancer Cell 31 (2017) 436–451.
Aberrant activation of the PI3K/mTOR pathway promotes resistance to sorafenib [219] C.A. Parachoniak, A. Rankin, B. Gaffney, R. Hartmaier, D. Spritz, R.L. Erlich, et al.,
in AML, Oncogene 35 (2016) 5119–5131. Exceptional durable response to everolimus in a patient with biphenotypic breast
[194] G.A. Vidal, M. Chen, S. Sheth, T. Svahn, E. Guardino, Phase I trial of everolimus cancer harboring an STK11 variant, Cold Spring Harb. Mol. Case Stud. 3 (2017)
and capecitabine in metastatic HER2− breast cancer, Clin. Breast Cancer 17 2017.
(2017) 418–426. [220] J.W. Yang, Y.L. Choi, Genomic profiling of esophageal squamous cell carcinoma
[195] Y. Guri, M.N. Hall, mTOR signaling confers resistance to targeted cancer drugs, (ESCC)-Basis for precision medicine, Pathol. Res. Pract. 213 (2017) 836–841.
Trends Cancer 2 (2016) 688–697. [221] B.D. Hopkins, C. Pauli, X. Du, D.G. Wang, X. Li, D. Wu, et al., Suppression of
[196] J.J. Bissler, F.X. McCormack, L.R. Young, J.M. Elwing, G. Chuck, J.M. Leonard, insulin feedback enhances the efficacy of PI3K inhibitors, Nature 560 (2018)
et al., Sirolimus for angiomyolipoma in tuberous sclerosis complex or lym- 499–503.
phangioleiomyomatosis, N. Engl. J. Med. 358 (2008) 140–151. [222] K.M. Ruicci, N. Pinto, M.I. Khan, J. Yoo, K. Fung, D. MacNeil, et al., ERK-TSC2
[197] D.M. Davies, P.J. de Vries, S.R. Johnson, D.L. McCartney, J.A. Cox, A.L. Serra, signalling in constitutively-active HRAS mutant HNSCC cells promotes resistance
et al., Sirolimus therapy for angiomyolipoma in tuberous sclerosis and sporadic to PI3K inhibition, Oral Oncol. 84 (2018) 95–103.
lymphangioleiomyomatosis: a phase 2 trial, Clin. Cancer Res. 17 (2011) [223] M. Tsoli, C. Wadham, M. Pinese, T. Failes, S. Joshi, E. Mould, et al., Integration of
4071–4081. genomics, high throughput drug screening, and personalized xenograft models as
[198] H.J. Klümpen, K.C. Queiroz, C.A. Spek, C.J. van Noesel, H.C. Brink, W.W. de Leng, a novel precision medicine paradigm for high risk pediatric cancer, Cancer Biol.
et al., mTOR inhibitor treatment of pancreatic cancer in a patient with Peutz- Ther. 19 (2018) 1078–1087.
Jeghers syndrome, J. Clin. Oncol. 29 (2011) e150–3. [224] C.K.M. Ip, P.K.S. Ng, K.J. Jeong, S.H. Shao, Z. Ju, P.G. Leonard, et al., Neomorphic
[199] A.J. Wagner, I. Malinowska-Kolodziej, J.A. Morgan, W. Qin, C.D. Fletcher, PDGFRA extracellular domain driver mutations are resistant to PDGFRA targeted
N. Vena, et al., Clinical activity of mTOR inhibition with sirolimus in malignant therapies, Nat. Commun. 9 (2018) 4583.
perivascular epithelioid cell tumors: targeting the pathogenic activation of [225] Q.W. Fan, T.P. Nicolaides, W.A. Weiss, Inhibiting 4EBP1 in glioblastoma, Clin.
mTORC1 in tumors, J. Clin. Oncol. 28 (2010) 835–840. Cancer Res. 24 (1) (2018) 14–21 Jan 1.
[200] G. Iyer, A.J. Hanrahan, M.I. Milowsky, H. Al-Ahmadie, S.N. Scott, [226] A.K. Bubna, Vorinostat-an overview, Indian J. Dermatol. 60 (2015) 419.
M. Janakiraman, et al., Genome sequencing identifies a basis for everolimus [227] M.V. Blagosklonny, Rapalogs in cancer prevention: anti-aging or anticancer?
sensitivity, Science 338 (2012) 221. Cancer Biol. Ther. 13 (2012) 1349–1354.
[201] D.E. Mais, P.V. Halushka, Photoaffinity labelling of the human platelet throm- [228] D. Mahalingam, E.C. Medina, J.A. Esquivel 2nd, C.M. Espitia, S. Smith,
boxane A2/prostaglandin H2 (TXA2/PGH2) receptor, Prog. Clin. Biol. Res. 301 K. Oberheu, R. Swords, K.R. Kelly, M.M. Mita, A.C. Mita, J.S. Carew, F.J. Giles,
(1989) 303–307. S.T. Nawrocki, Vorinostat enhances the activity of temsirolimus in renal cell
[202] E.J. Brown, P.A. Beal, C.T. Keith, J. Chen, T.B. Shin, S.L. Schreiber, Control of p70 carcinoma through suppression of survivin levels, Clin. Cancer Res. 16 (2010)
s6 kinase by kinase activity of FRAP in vivo, Nature 377 (1995) 441–446. 141–153.
[203] J. Chen, X.F. Zheng, E.J. Brown, S.L. Schreiber, Identification of an 11-kDa [229] J.F. Rodríguez-Moreno, M. Apellaniz-Ruiz, J.M. Roldan-Romero, I. Durán,
FKBP12-rapamycin-binding domain within the 289-kDa FKBP12-rapamycin-asso- L. Beltrán, C. Montero-Conde, A. Cascón, M. Robledo, J. García-Donas,
ciated protein and characterization of a critical serine residue, Proc. Natl. Acad. C. Rodríguez-Antona, Exceptional response to temsirolimus in a metastatic clear
Sci. U. S. A. 92 (1995) 4947–4951. cell renal cell carcinoma with an early novel MTOR-activating mutation, J. Compr.
[204] K. Yonezawa, K. Hara, M.T. Kozlowski, T. Sugimoto, K. Andrabi, Q.P. Weng, et al., Canc. Netw. 15 (2017) 1310–1315.
Regulation of eIF-4E BP1 phosphorylation by mTOR, J. Biol. Chem. 272 (1997) [230] A.K. Murugan, L. Rengyun, M. Xing, Identification and characterization of two
26457–26463. novel oncogenic mTOR mutations, Oncogene 38 (2019) 5211–5226.
[205] M.C. Lorenz, J. Heitman, TOR mutations confer rapamycin resistance by

20

You might also like