Fact-checked by Grok 2 weeks ago

mTOR

The mechanistic target of rapamycin (mTOR) is an evolutionarily conserved serine/ that functions as a central integrator of environmental cues, such as nutrients, factors, energy status, and stress signals, to regulate fundamental cellular processes including , , , , and . Discovered in the through the isolation of rapamycin—a compound produced by the soil bacterium on Rapa Nui ([Easter Island](/page/Easter Island))—mTOR was initially identified for its immunosuppressive and properties before its mechanistic role was elucidated in the 1990s via genetic studies in and mammals. mTOR operates through two distinct multiprotein complexes: mTOR complex 1 (mTORC1), which includes mTOR, regulatory-associated protein of mTOR (Raptor), GβL (also known as mLST8), DEP domain-containing mTOR-interacting protein (DEPTOR), and proline-rich Akt substrate 40 kDa (PRAS40), primarily sensing amino acids, energy levels, and oxygen to promote anabolic processes like protein synthesis via phosphorylation of targets such as S6 kinase 1 (S6K1) and eukaryotic initiation factor 4E-binding protein 1 (4E-BP1), while inhibiting catabolic pathways like autophagy; and mTOR complex 2 (mTORC2), comprising mTOR, rapamycin-insensitive companion of mTOR (Rictor), GβL, mammalian stress-activated protein kinase interacting protein 1 (mSIN1), and Protor-1/2, which is responsive to growth factors and insulin to regulate cytoskeletal organization, cell migration, and survival through activation of protein kinase B (Akt), protein kinase C (PKC), and serum- and glucocorticoid-induced protein kinase 1 (SGK1). These complexes are embedded within the phosphoinositide 3-kinase (PI3K)/Akt signaling pathway, where upstream regulators like tuberous sclerosis complex 1/2 (TSC1/2) and AMP-activated protein kinase (AMPK) fine-tune mTOR activity, with rapamycin selectively inhibiting mTORC1 by binding to the FK506-binding protein 12 (FKBP12) and disrupting its interaction with Raptor. Structurally, mTOR features a large (~289 kDa) modular architecture with N-terminal HEAT repeats for protein-protein interactions, a central (FRAP-ATM-TRRAP) domain, a rapamycin-binding FRB domain, a domain, and a C-terminal FATC domain, forming a hollow rhomboid that facilitates assembly with complex-specific partners. Dysregulation of mTOR signaling, often through mutations in PTEN, TSC1/2, or PI3K components, is implicated in numerous pathologies, including cancers (e.g., , , and gliomas, where hyperactivation drives tumor proliferation and angiogenesis), metabolic disorders like and , neurodegenerative diseases such as Alzheimer's and Parkinson's, autoimmune conditions like , and aging-related processes, where inhibition has been shown to extend lifespan in model organisms. Therapeutically, first-generation (rapalogs like and ) have been FDA-approved for , , and complex, while next-generation dual PI3K/mTOR or pan-mTOR inhibitors are under investigation to overcome feedback activation loops and improve efficacy in and beyond.

Discovery and History

Rapamycin Isolation

Rapamycin, also known as sirolimus, was isolated in 1975 from the bacterium Streptomyces hygroscopicus found in a soil sample collected from Rapa Nui (Easter Island). The soil was gathered during the Medical Expedition to Easter Island (METEI), a Canadian-led scientific venture from December 1964 to February 1965, which aimed to study the island's unique environment and health conditions; samples were subsequently sent to Ayerst Research Laboratories in Montreal for microbial analysis. At Ayerst, researchers Surendra N. Sehgal, H. Baker, and C. Vézina identified the antifungal-producing strain and extracted the active compound from the mycelium using organic solvents, followed by purification via silica gel chromatography to yield a crystalline solid. They named it rapamycin in honor of the island's indigenous name, Rapa Nui. Initially characterized as a potent antifungal agent, rapamycin demonstrated strong inhibitory activity against (minimum inhibitory concentration of 0.02–0.2 μg/ml across multiple strains), as well as moderate effects on and Trichophyton granulosum, though its instability in certain media limited broader testing. It showed no antibacterial activity against gram-positive or gram-negative bacteria and exhibited low acute toxicity in mice. Development as an antifungal was pursued but ultimately deprioritized after unexpected immunosuppressive properties emerged in preclinical studies during the late 1970s, with further confirmation in the 1980s through observations of its ability to inhibit T-lymphocyte proliferation and prevent allograft rejection in animal models. By the early 1990s, rapamycin entered phase I and II clinical trials as an immunosuppressant for preventing organ , particularly in renal transplantation, where it proved effective in combination with other agents like cyclosporine. These trials demonstrated its potency in reducing acute rejection episodes without the associated with inhibitors. In September 1999, the U.S. approved rapamycin (marketed as Rapamune) for the prophylaxis of renal in adults, marking its transition to clinical use. Later research identified mTOR as its primary molecular target.

Identification and Naming of mTOR

The identification of the cellular of rapamycin in mammalian cells began with biochemical efforts to purify the protein that interacts with the FKBP12-rapamycin complex. In 1994, and at used with immobilized FKBP12-rapamycin to isolate a 289-kDa from extracts, which they designated RAFT1 (rapamycin and FKBP12 1). This protein was shown to bind specifically to the FKBP12-rapamycin complex but not to FKBP12 alone or other immunophilin-drug complexes, establishing it as the direct intracellular responsible for rapamycin's antiproliferative effects. Concurrently, in 1994, Stuart L. Schreiber's group at Harvard employed a yeast two-hybrid screen using FKBP12 as bait to identify interacting proteins from a mouse embryonic cDNA library, cloning a homologous gene they named FRAP (FKBP-rapamycin-associated protein). The yeast TOR1 and TOR2 proteins were identified in as the targets of rapamycin in Saccharomyces cerevisiae through genetic screens for rapamycin-resistant mutants. FRAP encoded a 289-kDa serine/ kinase with significant (approximately 40-45%) to the yeast TOR1 and TOR2 proteins. Sequence analysis revealed conserved domains, including a kinase-like catalytic region, linking FRAP/RAFT1 to regulation and confirming its role in rapamycin-mediated G1 arrest in mammalian cells. Further characterization in the mid-1990s solidified these findings through mammalian cell studies demonstrating that overexpression or inhibition of FRAP/RAFT1 modulated progression in response to growth factors. For instance, of anti-FRAP antibodies into serum-stimulated fibroblasts blocked entry into , mirroring rapamycin's effects. By 1995-1996, Robert T. Abraham's group adopted the unified name mTOR (mammalian target of rapamycin) to reflect its homology to TOR and its conserved function across eukaryotes. The name mTOR, initially standing for "mammalian target of rapamycin," was widely used from the mid-1990s. In 2009, the officially changed the gene symbol from FRAP1 to MTOR and updated the expansion to "mechanistic target of rapamycin" to better emphasize its mechanistic role and evolutionary conservation, superseding earlier designations. Between 1995 and 1999, genetic and biochemical studies connected mTOR to the of , a key aspect of . In 1998, experiments showed that mTOR directly the ribosomal S6 kinase 1 (S6K1) at Thr-389 and the 4E-binding protein 1 (4E-BP1) at multiple sites (e.g., Thr-37/46), events essential for cap-dependent and inhibited by rapamycin. These phosphorylation events linked mTOR to - and growth factor-dependent signaling, with rapamycin-sensitive of 4E-BP1 relieving its inhibition of to promote mRNA . This nomenclature highlighted mTOR's evolutionarily conserved kinase activity in integrating environmental cues for cellular .

Molecular Structure and Function

mTOR is a large serine/ protein comprising 2,549 and encoded by the MTOR gene in humans. Its domain architecture includes an N-terminal region rich in tandem HEAT repeats that form alpha-helical solenoids, followed by the FAT domain, the FRB (FKBP12-rapamycin ) domain, the catalytic domain, and the C-terminal FATC domain. These s contribute to mTOR's overall modular structure, enabling interactions with regulatory elements and assembly into multi-subunit complexes such as and mTORC2. The FRB domain, spanning approximately 111 amino acids and weighing about 12 kDa, contains a hydrophobic pocket that specifically accommodates the FKBP12-rapamycin complex, facilitating allosteric regulation. Crystal structures of the FRB domain, such as the 2.7 Å resolution structure of the FKBP12-rapamycin-FRB complex (PDB: 1FAP), reveal this binding interface as a cleft formed by conserved residues, underscoring its role in rapamycin sensitivity. Higher-resolution structures, including a 1.75 Å crystal structure of FRB bound to a substrate peptide (PDB: 5WBH), further detail the conformational dynamics. mTOR exhibits strong evolutionary conservation, sharing roughly 42% sequence identity with its ortholog , particularly in the , , and FATC domains. This conservation highlights the preservation of core structural motifs across eukaryotes, from single-celled organisms to mammals. Structural studies employing cryo-electron (cryo-EM) and have elucidated mTOR's architecture in detail. A landmark 2016 cryo-EM structure of human mTORC1 at 4.4 resolution depicts mTOR as a bilobed protein, with the N-terminal HEAT repeats and domain forming an elongated "arm" that cradles the compact "lobe" containing the FRB, , and FATC domains. Complementary crystal structures of isolated mTOR domains, such as the 3.2 resolution map of the FAT-FRB--FATC segment bound to mLST8, confirm the kinase domain's atypical insertion of an alpha-helix and its positioning adjacent to the FRB for regulatory . These insights reveal how mTOR's domains integrate to form a scaffold for . More recent structural studies, as of 2025, have advanced this understanding. For instance, cryo-EM structures at resolutions around 3 Å have revealed the dynamic assembly of on lysosomal membranes, showing how RHEB and induce conformational changes in mTOR and to activate the . Another 2025 study elucidated the structural basis for amino acid-dependent regulation of through interactions with the complex, highlighting allosteric mechanisms within the and domains. These findings provide deeper insights into sensing and complex activation.

Kinase Activity and Substrates

mTOR functions as a serine/ within the phosphatidylinositol 3-kinase-related (PIKK) family, catalyzing the Mg-ATP-dependent of target proteins on serine or residues to regulate cellular processes such as and . Unlike typical lipid kinases, mTOR exhibits activity, utilizing the conserved to transfer the γ-phosphate from ATP to substrates while requiring magnesium ions for . This enzymatic mechanism is ATP-competitive, as demonstrated by inhibitors like Torin1, which bind the ATP-binding pocket with IC50 values of 2–10 nM in assays using purified mTOR complexes. Key substrates of mTOR include ribosomal protein S6 kinase 1 (S6K1), phosphorylated at Thr389 to promote translation initiation; eukaryotic initiation factor 4E-binding protein 1 (4E-BP1), targeted at Thr37 and Thr46 to release eIF4E and enhance cap-dependent translation; and unc-51 like autophagy activating kinase 1 (ULK1), modified at Ser757 to inhibit autophagy initiation. These phosphorylation events occur on motifs rich in hydrophobic or aromatic residues downstream of the target serine/threonine, such as the (T/S)φXφ consensus where φ represents hydrophobic amino acids, facilitating substrate recognition and activation. The FRB (FKBP-rapamycin binding) domain adjacent to the kinase domain provides allosteric regulation, modulating activity through conformational changes induced by binding partners or inhibitors like rapamycin-FKBP12, which sterically hinder substrate access in certain contexts. Experimental validation of mTOR's activity relies on assays, where recombinant mTOR or immunoprecipitated complexes synthetic peptides or full-length in the presence of radiolabeled ATP, confirming direct . Additionally, mass spectrometry-based phosphoproteomics has identified over 100 potential substrates by quantifying changes upon mTOR inhibition, revealing a broad network of targets involved in translation, , and cytoskeletal dynamics. These approaches underscore mTOR's role as a central , with phosphorylation motifs and allosteric controls ensuring specificity in substrate selection.

mTOR Signaling Complexes

mTORC1: Composition and Core Functions

The mechanistic target of rapamycin complex 1 () is a multiprotein assembly that integrates nutrient and growth signals to regulate cellular anabolic processes. Central to its structure is the serine/threonine mTOR, which serves as the catalytic core shared with mTORC2 but uniquely associates in mTORC1 with regulatory-associated protein of mTOR () as the defining scaffold protein. Raptor facilitates recruitment, complex stability, and nutrient sensing through interactions with Rag GTPases. Additionally, mammalian lethal with SEC13 protein 8 (mLST8, also known as GβL) stabilizes the mTOR domain and enhances , forming an obligate heterodimer with mTOR. The full mTORC1 complex exceeds 1 MDa in size. mTORC1 also incorporates inhibitory regulators that fine-tune its activity. Proline-rich Akt substrate of 40 kDa (PRAS40) binds to and acts as a pseudosubstrate , competing with physiological targets until its by mTORC1 relieves inhibition. Similarly, DEP domain-containing and TOR signaling (DEPTOR) protein associates directly with mTOR, exerting dual inhibitory effects on both mTORC1 and mTORC2 by blocking activation; its levels are inversely regulated by the complexes themselves. Notably, mTORC1 lacks , the scaffold protein characteristic of mTORC2, which distinguishes its composition and rapamycin sensitivity. The core functions of mTORC1 center on promoting anabolic metabolism while suppressing catabolic pathways. It drives protein synthesis by phosphorylating ribosomal S6 kinase 1 (S6K1) at Thr389, activating translation initiation, and eukaryotic initiation factor 4E-binding protein 1 (4E-BP1) at sites such as Thr37/46, releasing it from eIF4E to enhance cap-dependent translation. mTORC1 also stimulates lipid biogenesis by activating sterol regulatory element-binding protein 1 (SREBP1) through S6K1-mediated processing, upregulating de novo fatty acid and cholesterol synthesis in response to nutrients. Concurrently, it inhibits macroautophagy by phosphorylating unc-51 like autophagy activating kinase 1 (ULK1) at Ser757, preventing ULK1 complex formation and phagophore initiation under nutrient-replete conditions. Unlike mTORC2, mTORC1 is acutely sensitive to rapamycin, which binds the FKBP12-rapamycin-binding (FRB) domain of mTOR to disrupt complex integrity and substrate interactions. Activation of occurs primarily at lysosomal membranes, where nutrient sensing converges. Rag GTPases, in their heterodimeric GTP-bound form (RagA/B with RagC/D), bind to recruit from the to lysosomes in an -dependent manner, as demonstrated by assays tracking lysosomal translocation via and activity. There, Rheb in its active GTP-bound state directly interacts with the mTOR near the lobe, allosterically stimulating of downstream targets. Lys63-linked polyubiquitination of mTOR, mediated by TRAF6 in a p62-dependent fashion, further enhances this localization and activation under stimulation, independent of proteasomal degradation.

mTORC2: Composition and Core Functions

The mTORC2 complex is a multiprotein assembly distinct from , characterized by the absence of and the presence of unique subunits that define its rapamycin-insensitive signaling. Its core components include the serine/threonine mTOR, which serves as the catalytic subunit; , acting as the primary essential for complex stability and assembly; mLST8 (also known as GβL), which stabilizes the mTOR ; and Sin1 (mSIN1), which regulates access and localization through its pleckstrin homology () . Additional associated proteins include Protor-1 and Protor-2 (PRR5), which enhance mTORC2 activity toward certain substrates, and DEPTOR, a negative regulator that binds the mTOR to inhibit function. mTORC2 localizes primarily to the plasma membrane and endomembranes, facilitated by the Sin1 domain's interaction with phosphatidylinositol 3,4,5-trisphosphate (PIP3), which promotes recruitment in response to stimulation. This localization is crucial for spatial regulation of downstream signaling, contrasting with mTORC1's more cytosolic and lysosomal distribution. The complex's assembly requires and Sin1 to form a stable scaffold around mTOR, with mLST8 bridging interactions near the kinase , enabling substrate specificity distinct from nutrient-sensing pathways. Core functions of mTORC2 center on phosphorylating AGC kinases at their hydrophobic motifs to promote cell survival and cytoskeletal dynamics. It directly phosphorylates (Akt) at Ser473, enhancing Akt's full activation in the insulin/PI3K pathway to drive , , and anti-apoptotic signaling; similarly, it targets (PKC) isoforms at turn motif sites to regulate reorganization, , and polarity. mTORC2 also phosphorylates serum- and glucocorticoid-induced 1 (SGK1) at Ser422, supporting ion transport and cell survival under . Unlike , mTORC2 is largely insensitive to acute rapamycin treatment but can be disrupted by prolonged exposure, leading to disassembly and loss of function through effects on stability. Studies from the have elucidated mTORC2's involvement in loops that amplify PI3K signaling, such as Akt-mediated of Sin1 at Thr86 and Thr398, which enhances mTORC2 activity and sustains insulin responsiveness. Additionally, cross-talk with via S6K1 of at Thr1135 provides a regulatory node that fine-tunes mTORC2 output in metabolic contexts. These mechanisms underscore mTORC2's role in integrating growth signals for long-term cellular adaptation.

mTORC3: Composition and Core Functions

A third mTOR complex, mTORC3, was identified in and further characterized in studies as of 2024. It is rapamycin-insensitive and distinct from and mTORC2, containing mTOR along with the transcription factor ETV7, which binds to two sites on mTOR and is essential for assembly. Other components remain partially defined but include factors enabling phosphorylation of and mTORC2 targets. mTORC3 contributes to proliferation and , particularly in tumors resistant to rapalogs, by sustaining mTOR signaling independently of nutrient or inputs. Its highlights alternative mTOR regulation in pathological contexts, with implications for developing next-generation inhibitors.

Regulation of mTOR Activity

Upstream Nutrient and Growth Factor Signals

The activation of mTOR, particularly , is primarily regulated by upstream signals from nutrients and growth factors that sense environmental conditions to coordinate cellular growth. Growth factors such as insulin and (IGF-1) initiate signaling through the (PI3K)-Akt pathway, which phosphorylates and inhibits the tuberous sclerosis complex 2 (TSC2), a GTPase-activating protein () for the small GTPase Rheb. This inhibition prevents TSC2 from hydrolyzing GTP on Rheb, allowing Rheb to accumulate in its active GTP-bound form, which directly binds and activates at the lysosomal surface. The equilibrium for Rheb activation can be represented as: \text{Rheb-GDP + GTP} \rightleftharpoons \text{Rheb-GTP} where TSC-mediated GAP activity shifts the equilibrium toward the inactive GDP-bound state, but growth factor signaling inhibits TSC to favor Rheb-GTP accumulation. Nutrient availability, especially amino acids, provides another critical input for mTORC1 activation via recruitment to lysosomes. Leucine, a branched-chain amino acid, promotes the GTP-bound state of Rag GTPases (RagA/B in complex with RagC/D), which bind to raptor and translocate mTORC1 to the lysosomal membrane where it encounters Rheb-GTP. This amino acid-dependent Rag activation is essential for mTORC1 stimulation, with leucine acting as a potent sensor through upstream regulators like sestrins, though the core mechanism relies on the Rag-ractor interaction. Similarly, glutamine supports mTORC1 activation by facilitating v-ATPase activity on the lysosomal surface, which senses intralysosomal amino acids and promotes Rag GTPase nucleotide exchange via the Ragulator complex. Cellular energy status modulates through (AMPK), which is activated under low ATP conditions (high AMP/ATP ratio). AMPK phosphorylates TSC2 on serine , enhancing its GAP activity toward Rheb and thereby suppressing signaling to conserve . Additionally, AMPK directly phosphorylates on serines 792 and 722, promoting 14-3-3 binding and inhibiting assembly and activity. Hypoxic conditions induce specific inhibitory pathways involving regulated in development and DNA damage response 1 (REDD1) and REDD2 (also known as RTP801 and RTP801L). Hypoxia upregulates REDD1 expression via hypoxia-inducible factor (HIF), which displaces TSC2 from 14-3-3 proteins, allowing TSC2 to inhibit Rheb and suppress mTORC1. REDD2 similarly inhibits mTOR signaling under stress, acting through TSC-dependent mechanisms to promote cell survival by curtailing growth-promoting pathways.

Inhibitory Mechanisms Including Rapamycin

The primary pharmacological inhibitor of mTOR signaling is rapamycin (), a compound that forms a complex with the intracellular protein FKBP12. This rapamycin-FKBP12 complex binds allosterically to the FKBP12-rapamycin binding (FRB) domain of mTOR, inducing a conformational change that disrupts the interaction between mTOR and its essential cofactor in , thereby inhibiting the activity of the complex toward downstream substrates like S6K1 and 4E-BP1. Rapamycin exhibits high potency against , with an of approximately 0.1 nM in cellular assays measuring inhibition of S6K1 . While rapamycin primarily targets , prolonged exposure can indirectly impair mTORC2 assembly and function, leading to partial inhibition of Akt at Ser473, though this effect is less pronounced and requires chronic treatment. Several endogenous mechanisms also suppress mTOR activity in response to cellular stress or nutrient limitation. The stress-responsive protein REDD1 (also known as RTP801 or DDIT4) is upregulated under conditions such as hypoxia or DNA damage, where it inhibits mTORC1 by stabilizing the TSC2 component of the TSC1/2 complex, thereby enhancing its GTPase-activating protein (GAP) function toward the small GTPase Rheb and preventing Rheb-GTP-mediated activation of mTORC1. The TSC1/2 complex itself acts as a key negative regulator by functioning as a Rheb-GAP, hydrolyzing Rheb-GTP to its inactive GDP-bound form and thereby blocking Rheb's ability to allosterically stimulate mTORC1 kinase activity in response to growth factors or amino acids. Additionally, DEPTOR (DEP domain-containing mTOR-interacting protein) serves as a natural inhibitor that directly binds to the mTOR subunit within both mTORC1 and mTORC2, suppressing their respective kinase activities and providing a basal level of restraint on anabolic signaling. mTOR signaling is further modulated by loops that prevent overactivation. Within the pathway, activated S6K1 phosphorylates insulin receptor substrate-1 (IRS-1) at inhibitory sites such as Ser1101, which disrupts IRS-1's association with the and attenuates downstream PI3K-Akt signaling, contributing to under nutrient excess conditions. This feedback mechanism helps maintain but can be dysregulated in metabolic disorders. Studies from the 2010s highlighted the limitations of rapamycin's selective inhibition, prompting the development of dual mTORC1/mTORC2 inhibitors like AZD8055 and Torin1, which target the ATP-binding site of mTOR to fully block both complexes and overcome feedback reactivation of Akt.

Physiological Roles

Control of Cell Growth and Protein Synthesis

mTOR, primarily through its complex , serves as a central coordinator of by promoting anabolic processes such as in response to nutrients and growth factors. In mammalian cells, activation enhances the of mRNAs encoding proteins critical for growth, thereby increasing cellular mass and size. This regulation ensures that aligns with environmental cues, preventing uncontrolled . A key mechanism involves mTORC1-mediated control of . mTORC1 phosphorylates the 4E-binding protein 1 (4E-BP1) at multiple sites, leading to its dissociation from and allowing formation of the eIF4F complex, which facilitates cap-dependent of mRNAs with structured 5' untranslated regions. Concurrently, mTORC1 activates ribosomal protein S6 kinase 1 (S6K1) by phosphorylation at Thr389, which in turn phosphorylates 4B (eIF4B) and ribosomal protein S6, enhancing the efficiency of and for specific mRNA subsets. These actions collectively boost global protein synthesis rates by up to several fold under growth-promoting conditions. mTORC1 further supports by regulating , the process generating the cellular machinery for . It promotes (rRNA) synthesis through S6K1-dependent phosphorylation of upstream binding factor (UBF), which enhances UBF's activity in recruiting to rDNA promoters. Additionally, mTORC1 stimulates the of 5' terminal oligopyrimidine tract (5'-TOP) mRNAs, which encode ribosomal proteins, via S6K1 and 4E-BP1 pathways, ensuring coordinated production of ribosomal components. In terms of cell size regulation, mTOR signaling integrates growth factors like insulin-like growth factor-1 (IGF-1) to drive , particularly in . IGF-1 activates the PI3K/Akt pathway, which relieves TSC1/2 inhibition of , leading to increased protein and myofiber enlargement. Studies in models from the early 2000s demonstrate this: embryonic cells from mTOR mice exhibit reduced cell size compared to wild-type, underscoring mTOR's essential role in . Similarly, muscle-specific IGF-1 overexpression in transgenic mice induces significant , with fiber cross-sectional areas increasing by approximately 27%, mediated through mTOR-dependent translation.

Regulation of Autophagy and Metabolism

mTORC1 plays a central role in suppressing , a catabolic process essential for cellular and nutrient recycling, particularly under nutrient-replete conditions. Active directly phosphorylates ULK1, the serine/threonine that initiates , at serine 757 (Ser757). This disrupts the ULK1 interaction with AMPK, thereby inhibiting ULK1 activity and preventing the formation of the active ULK1-Atg13-FIP200 necessary for formation. The inhibitory effect can be represented as: \text{ULK1-pSer}^{757} + \text{Atg13} \rightleftharpoons \text{inactive complex} Under starvation conditions, mTORC1 inactivation leads to of ULK1 at Ser757, reactivating the complex and promoting to recycle cellular components for energy production. This regulatory mechanism ensures that is tightly coupled to nutrient availability, balancing with the anabolic processes like protein synthesis discussed previously, where provides for growth. In addition to autophagy regulation, mTORC1 reprograms cellular to favor biosynthetic pathways. It promotes by upregulating 6-phosphofructo-2-kinase/fructose-2,6-bisphosphatase 3 (PFKFB3), a key enzyme that enhances glycolytic flux through increased fructose-2,6-bisphosphate production; this is opposed by stress-induced REDD1, an mTORC1 inhibitor. mTORC1 also drives by activating sterol regulatory element-binding proteins 1 and 2 (SREBP1/2), transcription factors that induce expression of enzymes such as and , thereby increasing fatty acid synthesis. Conversely, mTORC1 inhibits in the liver by repressing PPARα-dependent transcription of ketogenic genes like HMGCS2 during fed states, preventing body production when alternative fuels are abundant. mTORC1's metabolic functions are intimately linked to lysosomal biology, as the complex senses nutrients directly at the lysosomal surface—a discovery from studies in the early 2010s highlighting the as a central hub for detection via the Rag GTPases and . Upon lysosomal damage, such as from permeabilizing agents or , dissociates from the lysosomal membrane, leading to its inactivation and of EB (TFEB). Nuclear translocation of TFEB then upregulates genes involved in lysophagy, a selective process that clears damaged lysosomes to maintain cellular integrity and prevent leakage of hydrolytic enzymes. This response underscores 's role in integrating lysosomal health with broader metabolic adaptation.

Functions in Plants

In plants, the target of rapamycin (TOR) kinase forms a single conserved complex analogous to mammalian mTORC1, comprising TOR, RAPTOR, and LST8, with no identifiable equivalent to mTORC2 due to the absence of RICTOR and SIN1 homologs. This plant TOR complex integrates nutrient, energy, and hormonal signals to orchestrate growth and development, showing evolutionary conservation from yeast but with adaptations for plant-specific processes such as cell wall remodeling and environmental sensing. TOR plays essential roles in plant developmental processes, including the regulation of growth and flowering. In , TOR activation by glucose, branched-chain , and low nitrate or temperature conditions—mediated through the receptor FERONIA and ROP2 —promotes elongation by enhancing and cytoskeletal dynamics. Similarly, TOR coordinates with signaling to facilitate the floral transition; for instance, the glucose-TOR-FIE pathway influences epigenetic regulation via Polycomb Repressive Complex 2, promoting flowering time and shoot activity. Under nutrient-replete conditions, inhibits to prioritize growth. In the presence of , phosphorylates ATG13 (an ortholog of ULK1/ATG1), suppressing formation and maintaining cellular in tissues like roots and leaves. also mediates stress responses critical for and . Glucose activates signaling to support growth in species such as and , facilitating rapid tip-directed extension and fertilization by boosting energy-dependent translation and actin organization. Conversely, during , abscisic acid () inhibits activity through SnRK2 kinases phosphorylating at serine 897, thereby attenuating growth to enhance stress tolerance and redirect resources toward mechanisms. Genetic studies underscore TOR's indispensable role in plant viability. Null mutants of TOR in exhibit embryo lethality at the 16- to 32-cell stage, arresting development due to failed and nutrient sensing. RNAi-mediated knockdown in the 2010s revealed TOR's involvement in synthesis; for example, TOR suppression leads to altered and deposition in root hairs, which can be partially rescued by supplementation, highlighting its link to primary and structural integrity.

Genetic and Experimental Insights

Effects of Gene Deletion and Knockout

Global deletion of the MTOR gene in mice results in embryonic lethality at approximately embryonic day 5.5 (E5.5), characterized by arrested development and severely growth-retarded embryos due to impaired trophoblast outgrowth and failure in early postimplantation processes. This phenotype underscores mTOR's indispensable role in the initial stages of embryogenesis, as homozygous null embryos exhibit no further progression beyond this point, while heterozygous MTOR mice develop normally without overt abnormalities. Tissue-specific knockouts have revealed mTOR's critical functions in maintaining organ . In the liver, conditional deletion of MTOR using Cre-loxP systems leads to reduced hepatic synthesis and storage, particularly in response to postprandial cues, resulting in impaired metabolic to feeding states. Skeletal muscle-specific MTOR ablation causes progressive , metabolic dysregulation, and dystrophin-related structural defects, highlighting mTOR's necessity for protein homeostasis and myofiber integrity. In the , neural progenitor-specific inactivation of via Raptor knockout induces , reduced neuronal size, and increased , primarily through diminished cap-dependent translation initiation mediated by dephosphorylated 4E-BPs that sequester . These models demonstrate that mTOR signaling is essential for tissue-specific growth and differentiation, with disruptions leading to organ and functional deficits. Recent CRISPR-based studies (as of 2025) have enabled precise modeling of pathogenic MTOR variants, confirming their gain-of-function effects and exploring therapeutic corrections in cellular and animal models. Conditional knockout approaches employing Cre-loxP recombination to target the MTOR gene post-developmentally have further delineated mTOR's roles without embryonic lethality. For instance, inducible deletions in and Schwann cells impair myelination, resulting in thinner sheaths and delayed maturation of myelin-forming cells due to disrupted synthesis and cytoskeletal . Pharmacological inhibition with rapamycin partially mimics these genetic effects by suppressing activity, leading to analogous reductions in and across tissues. Studies from the early on hypomorphic MTOR alleles, which reduce mTOR expression by approximately 25-30%, show extended lifespan in mice, with delayed onset of age-related pathologies and improved healthspan, paralleling the benefits observed with chronic rapamycin treatment. These experimental insights affirm mTOR's dosage-sensitive regulation of and .

Pathogenic Mutations in Humans

Pathogenic mutations in the mTOR pathway, including both and variants, lead to hyperactivation of the pathway and underlie several rare neurodevelopmental disorders in humans. These mutations typically involve loss-of-function changes in negative regulators like TSC1 and TSC2 or gain-of-function alterations in positive regulators such as MTOR itself, resulting in dysregulated cell growth, proliferation, and cortical development. Loss-of-function mutations in TSC1 or TSC2 genes cause (TSC), an autosomal dominant disorder characterized by benign tumors, seizures, and due to constitutive activation. Approximately 70-80% of TSC cases arise from identifiable pathogenic variants in these genes, with TSC2 mutations being more common and associated with more severe phenotypes. The prevalence of TSC is estimated at 1 in 6,000 live births. Gain-of-function germline mutations in the MTOR gene are responsible for Smith-Kingsmore syndrome (), a rare overgrowth disorder featuring , , seizures, and . These heterozygous or inherited variants, such as the recurrent p.E1799K mutation, enhance mTOR activity, leading to excessive protein synthesis and cellular . Recent functional studies have identified numerous distinct MTOR variants (over 20 reported) in SKS patients, confirming their pathogenicity through increased of downstream targets like S6K1. Somatic mutations in MTOR and related genes also drive focal malformations of cortical development (MCD), including focal cortical dysplasia type II (FCDII) and hemimegalencephaly (HME), which are major causes of intractable epilepsy in children. These postzygotic variants, often mosaic and brain-restricted, hyperactivate the PI3K-AKT-mTOR pathway in affected neurons, causing abnormal cortical architecture and cytomegaly. Somatic MTOR variants are detected in approximately 25% of MCD cases, particularly in FCDII and HME, where they contribute to epileptogenesis in up to 22% of surgically resected specimens. Hyperactivation of the PI3K/mTOR pathway through germline or somatic mutations in genes like PIK3CA, AKT3, and MTOR is implicated in a of overgrowth syndromes, including the (PROS) and related disorders. Recent analyses from 2024-2025 highlight that these variants drive segmental tissue overgrowth and neurodevelopmental features, with mTOR hyperactivation promoting and lipophagy inhibition in conditions like macrodactyly. Pathogenic mTOR pathway variants account for 10-20% of genetic epilepsies, particularly those linked to focal cortical malformations.

Pathophysiological Implications

Role in Cancer Development

The mammalian target of rapamycin (mTOR) pathway is hyperactivated in approximately 70% of human cancers, primarily through upstream alterations such as PTEN loss of function, activating mutations in PIK3CA, or TSC1/TSC2 inactivation, which relieve inhibitory constraints on mTOR signaling and drive uncontrolled cell proliferation. This hyperactivation fosters a pro-tumorigenic environment by promoting angiogenesis, as mTOR enhances the translation and transcriptional activity of hypoxia-inducible factor 1α (HIF-1α), leading to increased vascular endothelial growth factor (VEGF) expression and new blood vessel formation to support tumor expansion. mTOR dysregulation is especially prominent in certain tumor types, including (RCC), where it integrates nutrient sensing with responses to sustain aggressive , and , where it regulates maintenance and invasion within the microenvironment. Recent advances from 2024 and 2025 underscore its critical involvement in (TNBC), where PI3K/AKT/mTOR pathway alterations correlate with poor prognosis and drive metastatic potential through enhanced signaling crosstalk. Mechanistically, hyperactive mTOR contributes to oncogenesis by reprogramming cellular metabolism, notably enhancing aerobic glycolysis—the Warburg effect—via upregulation of key glycolytic enzymes like pyruvate kinase M2 (PKM2), which supports rapid ATP production and biosynthetic demands in proliferating tumor cells. Additionally, mTOR promotes apoptosis evasion by phosphorylating eukaryotic translation initiation factor 4E-binding protein 1 (4E-BP1), thereby relieving its inhibitory effect on cap-dependent translation of anti-apoptotic proteins and pro-survival mRNAs. mTOR exists in two distinct complexes with complementary yet distinct roles in cancer: primarily orchestrates anabolic , , and biogenesis to fuel tumor accumulation, while mTORC2 supports cytoskeletal organization, survival, and to facilitate and . Emerging 2025 research reveals that pathway resistance often arises from uncoupling between AKT and mTOR signaling, particularly in PTEN-deficient tumors like , where compensatory mTOR reactivation sustains oncogenesis despite upstream inhibition. This metabolic shift toward also interconnects with broader suppression, amplifying nutrient availability for sustained tumor growth.

Involvement in Neurological Disorders

Dysregulation of the mTOR pathway has been implicated in several neurological disorders, particularly those involving aberrant neuronal growth, , and protein in the . In (ASD), hyperactivation of mTOR signaling downstream of PTEN mutations occurs in approximately 10-20% of cases associated with , leading to excessive dendritic arborization and neuronal that contribute to enlarged volume and social deficits. These PTEN loss-of-function mutations relieve inhibition on the PI3K-AKT-mTOR axis, promoting overgrowth of cortical neurons and disrupting synaptic balance, as observed in both patients and Pten heterozygous models. In , somatic mutations in the MTOR gene are a key driver of focal cortical (FCD) type II, a malformation of cortical development that manifests as intractable seizures due to hyperactive signaling in dysplastic neurons. These low-level mosaic mutations, often with variant frequencies of 0.6-12%, disrupt neuronal migration and ciliogenesis, resulting in balloon cells and abnormal characteristic of FCD.30437-9) Similarly, in tuberous sclerosis complex (TSC), germline mutations in TSC1 or TSC2 genes cause constitutive mTOR activation, leading to cortical tubers and hyperexcitable networks that precipitate seizures in up to 90% of patients. The severity of epilepsy in TSC correlates with the degree of mTOR hyperactivity, highlighting its role in epileptogenesis beyond mere structural lesions. mTORC1 hyperactivity in (AD) inhibits , a critical clearance , thereby promoting the accumulation of tangles and amyloid-β (Aβ) plaques that drive neurodegeneration. Specifically, elevated suppresses autophagosome formation and lysosomal degradation, exacerbating intracellular aggregation and Aβ buildup in neurons, as evidenced in AD models where mTOR inhibition enhances protein clearance and ameliorates cognitive decline. Recent (iPSC)-derived models of TSC and AD, developed in 2024-2025, have recapitulated these defects, showing mTOR-dependent axonal spheroids and impaired in patient neurons, providing platforms to test targeted interventions. In (FXS), the most common inherited form of , loss of FMRP leads to dysregulated mTOR signaling and excessive translation via , resulting in elongated dendritic spines and impaired that underlie cognitive and behavioral symptoms. This mTORC1-eIF4E hyperactivation promotes overproduction of synaptic proteins like MMP9, disrupting the balance between protein synthesis and degradation in FXS neurons.00933-4) Pharmacological inhibition of mTOR with rapamycin or its analogs, such as , has shown promise in reducing frequency in TSC-associated and reversing autism-like behaviors in preclinical models, though clinical trials in and FXS have yielded mixed results on core symptoms like . In TSC patients, treatment significantly decreases burden by attenuating mTOR-driven hyperexcitability, supporting its approval for management in this context.

Contributions to Aging and Immune Dysregulation

mTORC1 hyperactivation has been implicated in accelerating , a hallmark of aging characterized by irreversible arrest and secretion of pro-inflammatory factors. Persistent mTORC1 signaling, often resulting from defects in regulatory components like RagC, drives senescence phenotypes by promoting protein synthesis and inhibiting , thereby contributing to tissue dysfunction in aging organisms. In preclinical models, such hyperactivation in fibroblasts and osteoblasts leads to premature senescence through mechanisms including plasma membrane depolarization and dysregulated . Caloric restriction, a well-established for lifespan extension, inhibits activity primarily through activation of the energy sensor AMPK, which phosphorylates and suppresses components like . This inhibition shifts cellular toward , enhancing resistance and delaying age-related decline. Pharmacological inhibition of mTOR with rapamycin similarly extends lifespan in mice by 9-14% when administered in mid-to-late life, mimicking caloric restriction effects and improving healthspan metrics such as physical function and resistance. In the context of aging mechanisms, mTOR dysregulation contributes to the progressive decline in , a lysosomal degradation process essential for cellular , which diminishes with age due to sustained activity suppressing autophagosome formation. This autophagy impairment exacerbates protein aggregation and mitochondrial dysfunction, fueling inflammaging—a chronic, low-grade driven by elevated pro-inflammatory cytokines like IL-1β. promotes IL-1β production in macrophages via TSC1-dependent pathways, linking nutrient sensing to inflammatory responses that accelerate systemic aging. Dual inhibition of and mTORC2 has shown promise in gerosuppression, extending lifespan in model organisms by more comprehensively restoring autophagy and reducing senescence-associated compared to mTORC1-specific blockade. mTOR signaling also plays a pivotal role in immune dysregulation during aging, particularly in , where adaptive immunity wanes and innate responses become dysregulated. mTORC2 regulates T-cell differentiation by promoting effector functions and memory formation through Akt activation, influencing Th2 and + T-cell fates essential for immune memory. In contrast, drives macrophage polarization toward pro-inflammatory states via metabolic reprogramming, including enhancement, which sustains chronic inflammation in aged tissues. Recent insights highlight mTOR's involvement in age-related immune vulnerabilities, including impaired responses and heightened . Sustained activation in older T cells hinders memory formation and vaccine efficacy, as seen in reduced responses to and vaccines, while mTOR inhibition enhances protective immunity. In , mTOR blockade mitigates by reprogramming effector T cells and reducing autoreactive responses, with 2025 reviews emphasizing its therapeutic potential in conditions like and . Additionally, 2024 studies reveal mTOR hyperactivation contributes to B-cell exhaustion in chronic settings, impairing through dysregulated metabolism and BCAT1-mediated lysosomal signaling, further compounding .

Associations with Rare and Other Diseases

Smith-Kingsmore syndrome is a rare caused by heterozygous mutations in the , resulting in gain-of-function effects that lead to constitutive activation of the mTOR pathway. These mutations are associated with or hemimegalencephaly, , seizures, and distinctive facial dysmorphology. The overactivation of mTOR disrupts normal neuronal growth and development, contributing to the syndrome's characteristic overgrowth phenotypes. Lymphangioleiomyomatosis (LAM), a rare cystic lung disease, frequently occurs in patients with tuberous sclerosis complex (TSC), driven by inactivating in TSC1 or TSC2 genes that relieve inhibition of . This hyperactivation promotes the proliferation and survival of abnormal smooth muscle-like cells (LAM cells) that infiltrate the s, leading to formation, , and progressive . In TSC-associated LAM, mTORC1 dysregulation also enhances metabolic reprogramming, including increased and lipid synthesis, exacerbating tissue destruction. In systemic sclerosis (), signaling in dermal and lung fibroblasts drives excessive production and . Activated promotes fibroblast-to-myofibroblast differentiation and synthesis through downstream effectors like and SIRT1 suppression, contributing to skin thickening and . Recent preclinical studies from the have highlighted 's role in inflammatory-fibrotic crosstalk, with elevated pathway activity observed in patient-derived fibroblasts. Lysosomal storage diseases (LSDs), such as Pompe and Gaucher diseases, often feature impaired due to hyperactivity, which hinders lysosomal biogenesis and substrate clearance. Lysosomal damage in these disorders activates , suppressing formation and exacerbating substrate accumulation, like in Pompe disease or in Gaucher. This dysregulation links mTOR inhibition to potential restoration of autophagic flux, as seen in cellular models where reduced mTOR activity enhances lysosomal function. Recent investigations from 2024 have implicated mTOR dysregulation in type III (GSD III), where defective glycogen debranching leads to abnormal storage and altered -mediated in liver and muscle. In GSD III models, hyperactivation contributes to impaired clearance, highlighting its role in metabolic complications like and . PIK3CA-related overgrowth spectrum (PROS) encompasses rare segmental overgrowth disorders caused by postzygotic gain-of-function mutations in PIK3CA, which hyperactivate the PI3K/AKT/mTOR pathway. This leads to excessive tissue growth in affected areas, manifesting as vascular malformations, , and skeletal asymmetry in syndromes like CLOVES. mTOR overactivation in PROS drives and , distinguishing it from other overgrowth conditions.

Therapeutic Targeting

Development of mTOR Inhibitors

The development of mTOR inhibitors originated with rapamycin (), a compound isolated from the bacterium , which binds to the FKBP12 protein and allosterically inhibits by disrupting its interaction with substrates. This agent received FDA approval on September 15, 1999, initially for preventing organ rejection in renal transplant recipients, marking the first clinically viable mTOR-targeted therapy. Building on rapamycin's foundation, pharmaceutical efforts focused on creating semi-synthetic analogs, known as rapalogs, to improve solubility, stability, and oral bioavailability while retaining the allosteric . Sirolimus itself served as the scaffold for these derivatives, with everolimus (RAD001) and temsirolimus (CCI-779) emerging as key examples; temsirolimus was approved by the FDA on May 30, 2007, for advanced , and everolimus followed on March 30, 2009, for the same indication after demonstrating efficacy in refractory cases. These rapalogs—sirolimus, everolimus, and temsirolimus—primarily suppress activity but often lead to feedback activation of upstream pathways like PI3K/AKT, limiting their potency against mTORC2-dependent processes. To address these shortcomings, researchers pursued ATP-competitive inhibitors that directly target the kinase domain of , enabling simultaneous blockade of and mTORC2. The Torin series exemplifies this shift: Torin1, identified through in 2009, potently inhibits both complexes with an in the nanomolar range but suffered from suboptimal due to rapid clearance. This prompted the development of Torin2 in 2011, a second-generation analog with enhanced selectivity (over 1,000-fold for mTOR versus PI3K isoforms) and improved stability, allowing effective suppression of mTOR signaling in preclinical tumor models without the feedback loops seen with rapalogs. Similarly, the PP242 class, featuring a pyrazolo[3,4-d]pyrimidine core, was introduced around as dual /2 inhibitors; PP242 exhibits an of 8 nM for mTOR kinase activity and has been instrumental in dissecting mTOR's role in cellular processes like and in laboratory settings. These ATP-competitive agents represent a significant advancement, offering broader pathway inhibition compared to the substrate-specific effects of rapalogs. Efforts to refine selectivity have led to Raptor-specific inhibitors that target without affecting mTORC2, minimizing off-target effects associated with chronic dual inhibition. In preclinical studies reported in 2025, such compounds—designed to disrupt Raptor-mTOR interactions—effectively rescued neuronal deficits in tuberous sclerosis complex (TSC) models derived from patient iPSCs, restoring balanced activity and synaptic function. These selective agents hold promise for TSC-related pathologies, where hyperactive drives aberrant growth, potentially offering efficacy akin to rapamycin but with reduced systemic toxicities. Resistance to remains a challenge, often arising from oncogenic in the mTOR gene that alter or pathway dynamics. For instance, in the FRB , such as A2034V, can cause to rapalogs by weakening interactions, promoting constitutive mTOR and tumor progression in cancers. Such underscore the need for next-generation that can bypass allosteric sites, as seen with ATP-competitive classes. Recent explorations have also turned to plant-derived compounds for novel mTOR modulation; in 2025 studies, picroside II from Picrorhiza scrophulariiflora emerged as a potent natural , the mTOR kinase with high affinity to suppress via pathway attenuation. These bioactive molecules, alongside from , highlight the potential of natural products to inspire hybrid inhibitors with improved tolerability.

Applications in Transplantation and Metabolic Disorders

mTOR inhibitors, particularly the rapalogs and , play a key role in immunosuppressive regimens for solid , especially , by enabling reduced exposure to inhibitors (CNIs) such as or cyclosporine, thereby mitigating CNI-associated while preserving graft function. Conversion from CNI-based therapy to mTOR inhibitor maintenance therapy has been associated with improved measured and lower incidence of malignancies in posttransplant patients. Long-term clinical outcomes demonstrate favorable graft survival rates with mTOR inhibitor use; for instance, in cohorts switched to these agents, 5-year graft survival reaches approximately 83.5%, with reduced rates of acute rejection. Typical dosing for in starts at 0.75 mg twice daily, adjusted to achieve trough levels of 3-8 ng/mL, often in combination with low-dose CNIs. In metabolic disorders, is FDA-approved for the treatment of subependymal giant cell astrocytoma () associated with tuberous sclerosis complex (TSC), a condition characterized by mTOR pathway hyperactivation leading to tumor growth. Phase II and long-term extension studies have shown that everolimus at doses of 4.5-10 mg/m² daily (or 5-10 mg fixed daily for adults) reduces SEGA volume by at least 30% in over 60% of patients, with sustained efficacy over 5 years and manageable tolerability. For (ADPKD), an off-label application, clinical trials indicate that everolimus slows total kidney volume growth by approximately 35% over 12 months compared to placebo, though it does not consistently preserve renal function and may accelerate estimated decline in some cases. mTORC1 inhibition has shown promise in addressing autophagy defects in glycogen storage diseases (GSDs), such as GSD III (glycogen debranching enzyme deficiency), where impaired autophagic clearance contributes to glycogen accumulation and metabolic dysfunction. Preclinical studies in GSD III mouse models demonstrate that rapamycin enhances autophagy, reduces hepatic and muscle glycogen buildup, and improves overall phenotype when combined with gene therapy approaches. Recent investigations, including 2023-2024 models, further support that mTORC1 blockade ameliorates steatosis and mitochondrial dysfunction in GSD Ia by promoting autophagosome formation and lipid metabolism. While human case studies remain limited, these findings suggest potential therapeutic utility, though clinical translation requires further validation. Common side effects of across these applications include , affecting up to 78% of patients, often manifesting as painful oral ulcers that may necessitate dose adjustments or temporary discontinuation. strategies involve topical corticosteroids or dose reduction to 2.5-5 mg daily in severe cases, balancing against tolerability.

Use in Oncology and Emerging Cancer Therapies

mTOR inhibitors have established roles in oncology, particularly for cancers driven by pathway hyperactivation. , an inhibitor, is FDA-approved for advanced hormone receptor-positive (HR+), HER2-negative in combination with , where it extends compared to exemestane alone. It is also approved for advanced (RCC), demonstrating improved outcomes in patients with poor prognostic features. Temsirolimus, another rapalog, received FDA approval for advanced RCC as a single agent, showing a modest objective response rate of approximately 8% but prolonging in phase III trials. Combination therapies leveraging are advancing to address resistance and enhance efficacy in specific cancers. In pediatric high-grade gliomas (HGG), including diffuse midline gliomas, ongoing 2025 trials explore combinations of like with PI3K inhibitors, such as brain-penetrant agents, to target the PI3K/AKT/mTOR pathway more comprehensively and improve blood-brain barrier penetration. For (), combined with agents targeting upstream pathways, like or PI3K inhibitors, overcomes intrinsic resistance by suppressing feedback activation of PI3K/AKT signaling, leading to enhanced tumor growth inhibition in preclinical models. Emerging strategies focus on selective targeting and novel mechanisms. mTORC1-selective inhibitors, such as bi-steric compounds, show promise in complex (TSC)-associated tumors by more potently suppressing without off-target mTORC2 inhibition, reversing cellular phenotypes in TSC models and achieving tumor shrinkage rates exceeding 50% in preclinical TSC-deficient xenografts. Recent 2025 insights reveal that modulate the cancer transcriptome by influencing and intron retention, potentially enhancing therapeutic responses through coordinated regulation of pro-survival isoforms via pathways like poison exon inclusion. Clinical responses to are notably higher in tumors with hyperactive mTOR signaling, with objective response rates of 30-50% observed in TSC-associated lesions compared to lower rates in unselected solid tumors. Biomarkers such as phosphorylated S6 kinase (pS6K), a direct downstream effector of , correlate with pathway activation and predict sensitivity, guiding patient selection in trials for mTOR-hyperactive cancers.

Potential in Neurodegenerative and Aging Interventions

mTOR modulation has shown promise in preclinical and early clinical studies for addressing neurodegenerative diseases, particularly through enhancing and reducing . In models, temsirolimus, a rapamycin analog and , has been demonstrated to enhance autophagic clearance of hyperphosphorylated , thereby attenuating both and . Similarly, in models, rapamycin provides neuroprotective effects by preserving striatal and mitigating dopaminergic neuron loss, independent of direct inhibition in some contexts. Everolimus, another rapamycin derivative, has exhibited protective effects against glutamate-induced neuronal death in PC12 cell models relevant to Parkinson's . Investigational trials targeting inhibition are advancing for neurodevelopmental disorders with neurodegenerative features, such as . A phase II randomized controlled trial of metformin, which inhibits via AMPK activation, in individuals aged 6 to 35 years with aims to evaluate improvements in behavioral and cognitive symptoms, with preliminary data suggesting potential benefits in synaptic function. Recent iPSC-derived neuron studies from complex (TSC) patients, where mTOR hyperactivation is central, have shown that selective inhibitors rescue aberrant cellular phenotypes, including excessive dendrite growth and electrophysiological deficits, highlighting translational potential for mTOR-targeted interventions in TSC-associated neurodegeneration. In aging , mTOR inhibition intersects with pathways, often through dose-dependent mechanisms where low doses promote healthspan without immunosuppressive side effects. Low-dose rapamycin (e.g., 75 μg/kg/day intermittently) extends lifespan in middle-aged mice by up to 60% and improves health metrics like , supporting its role in delaying age-related decline. Metformin, acting dually through AMPK activation to suppress mTOR signaling, alleviates in dental pulp stem cells by downregulating miR-34a-3p and enhancing autophagic flux. Emerging 2025 studies indicate that mTOR pathway modulation, such as via nanovesicles activating TLR9 to inhibit mTOR, reverses T-cell in tumor microenvironments, potentially restoring immune function in aged or diseased states. Ongoing clinical efforts explore mTOR modulation for aging interventions across species. The Targeting Aging with Metformin (TAME) trial, a multi-site phase III study, continues to investigate metformin's mTOR-suppressive effects on delaying age-related diseases in older adults, with 2025 updates reporting mounting evidence for vascular and metabolic benefits. In companion animals, the TRIAD trial, a randomized placebo-controlled study of rapamycin in dogs aged 7 years and older, assesses lifespan extension and healthspan improvements like cardiac and cognitive function, with enrollment expansions funded in 2024. These initiatives underscore the potential of calibrated mTOR inhibition to counteract and neuronal vulnerabilities in aging. As of November 2025, the trial remains ongoing without final results, while TRIAD interim data suggest improved healthspan in treated dogs.

Molecular Interactions

Key Protein Binding Partners

The mechanistic target of rapamycin (mTOR) forms multiprotein complexes through direct interactions with key scaffold and modulator proteins, which are essential for its structural integrity and regulation. In the mTOR complex 1 (mTORC1), Raptor serves as the primary scaffold protein, binding directly to mTOR via its HEAT repeats and TOR signaling motif to facilitate substrate recruitment and nutrient-sensitive signaling.00808-5) mLST8, a small subunit also known as GβL, stabilizes the mTOR kinase domain in both mTORC1 and mTORC2 by interacting with the catalytic region, enhancing overall complex assembly without altering kinase activity. These interactions were initially identified through yeast two-hybrid screening and confirmed by co-immunoprecipitation (co-IP) assays, demonstrating stoichiometric binding in cellular contexts.00808-5) For mTOR complex 2 (mTORC2), acts as the defining scaffold, associating with mTOR independently of and promoting rapamycin-insensitive functions; this was discovered via yeast two-hybrid methods and verified by co-IP, highlighting Rictor's role in recruiting additional subunits. Protor-1 and Protor-2 bind to Rictor in mTORC2, contributing to complex assembly and stability, as shown by co-IP studies demonstrating their association with other mTORC2 subunits but not Raptor. Sin1, or MAPKAP1, binds to the C-terminal region of mTOR in mTORC2, stabilizing the complex and enabling specific phosphorylation events; structural studies show Sin1's is mediated by its -interacting motif.00730-X/fulltext) mLST8 is shared between the two complexes, binding similarly to support kinase domain conformation. DEPTOR functions as an endogenous inhibitor in both and mTORC2, directly binding to the domain of mTOR with a bipartite that competes with other interactors under nutrient limitation; co-IP and binding assays confirm its inhibitory role through steric hindrance. In , PRAS40 (AKT1S1) associates via , acting as a modulator that binds mTOR indirectly but inhibits activity until phosphorylated; this interaction was mapped using yeast two-hybrid and co-IP techniques. Rapamycin induces a ternary complex by binding FKBP12, which then interacts with mTOR's FKBP12-rapamycin (FRB) domain adjacent to the site, disrupting association; crystallographic studies reveal high-affinity in this configuration. Recent approaches, such as rapid of endogenous proteins (RIME), have expanded the known interactome, identifying over 500 chromatin-associated mTOR partners in cancer cells, with more than 90% previously uncharacterized, underscoring the breadth of direct interactions beyond core components. Quantitative multiplex co-IP networks have further mapped over 300 binary interactions in mTOR signaling modules.

Regulatory Networks with Other Pathways

mTOR signaling is intricately integrated with several major cellular pathways, enabling coordinated of , , and in response to diverse stimuli. This crosstalk ensures that mTOR activity is finely tuned by upstream inputs such as nutrients, factors, and status, while mTOR in turn modulates downstream effectors to prevent dysregulated . Key interactions occur through shared regulatory nodes like the TSC complex, which serves as a convergence point for multiple signals. The PI3K-Akt pathway exerts positive regulation on mTORC1 primarily via the TSC/Rheb axis. Activation of PI3K leads to Akt-mediated of TSC2 at multiple sites, inhibiting the TSC1-TSC2 complex's GTPase-activating function toward Rheb; this allows Rheb-GTP to accumulate and directly activate on lysosomes. However, mTORC1 activation triggers a loop through S6K1-mediated of IRS1 at serine residues, promoting IRS1 ubiquitination and , thereby attenuating PI3K-Akt signaling and preventing excessive pathway hyperactivity. This bidirectional regulation maintains insulin sensitivity and cellular but contributes to when dysregulated. Under energy stress, (AMPK) inhibits to prioritize energy conservation. AMPK directly phosphorylates in , disrupting its assembly and activity, while also phosphorylating TSC2 at serine 1345, enhancing TSC2's suppression of Rheb and thereby reinforcing inhibition. This convergence at TSC2 allows AMPK to integrate low ATP/AMP ratios with nutrient sensing, halting anabolic processes during metabolic stress. mTOR also interfaces with the Hippo pathway through the effectors and TAZ, regulated by LATS kinases. YAP/TAZ promote activation by enhancing sensing and positioning, facilitating nutrient-dependent mTORC1 recruitment; conversely, LATS1/2-mediated sequesters YAP/TAZ in the , indirectly suppressing mTORC1 under growth-restrictive conditions. Similarly, mTORC1 negatively regulates the Wnt/β-catenin pathway by downregulating receptor levels, limiting β-catenin stabilization and translocation; this suppression influences maintenance and homeostasis, with TSC2 acting as a shared integrator via GSK3β . Recent studies highlight mTOR's role in regulation, particularly mediated by serine/arginine-rich splicing factor 3 (SRSF3). In mTOR-activated cells, SRSF3 promotes , generating shorter isoforms that enhance proteome diversity and support growth under nutrient abundance; inhibition of reverses this splicing pattern, underscoring its control over processing. Network analyses further reveal extensive overlap, with mTOR dysregulation implicated in approximately 80% of cancers through hyperactivation in interconnected pathways.

References

  1. [1]
    Multifaceted role of mTOR (mammalian target of rapamycin ... - Nature
    Oct 2, 2023 · The mammalian target of rapamycin (mTOR) is a protein kinase that controls cellular metabolism, catabolism, immune responses, autophagy, ...
  2. [2]
    Mechanistic Target of Rapamycin - an overview | ScienceDirect Topics
    The mechanistic target of rapamycin (mTOR) is a serine-threonine protein kinase that regulates cell growth, cell proliferation, cell motility, cell survival, ...
  3. [3]
    Mechanistic target of rapamycin inhibitors - PubMed Central - NIH
    This review highlights some of the promising results seen with mTOR inhibitors in the clinic and assesses some of the challenges that remain in predicting ...
  4. [4]
    The origin story of rapamycin: systemic bias in biomedical research ...
    Less well known is how an isolate of S. hydroscopicus found its way from Easter Island into the hands of pharmaceutical researchers at Ayerst Research ...Missing: hygroscopicus | Show results with:hygroscopicus
  5. [5]
  6. [6]
  7. [7]
    rapamycin: clinical results and future opportunities 1 - Transplantation
    Phase I and II studies. The first clinical study employed a blinded randomized design to examine the safety of RAPA (1 to 13 mg/m 2) versus placebo added to ...
  8. [8]
    Rapamycin in transplantation: A review of the evidence
    Rapamycin is a new drug with both immunosuppressant and antiproliferative properties that has a unique mechanism of action distinct from that of the calcineurin ...
  9. [9]
    RAFT1: a mammalian protein that binds to FKBP12 in a rapamycin ...
    Jul 15, 1994 · A RAFT1 cDNA was obtained and found to encode a 289 kDa protein (2549 amino acids) that is 43% and 39% identical to TOR2 and TOR1, respectively.
  10. [10]
    A mammalian protein targeted by G1-arresting rapamycin–receptor ...
    Jun 30, 1994 · Peptide sequences from purified bovine FRAP were used to isolate a human cDNA clone that is highly related to the DRR1/TOR1 and DRR2/TOR2 ...
  11. [11]
  12. [12]
    RAFT1 phosphorylation of the translational regulators p70 S6 kinase ...
    RAFT1 directly phosphorylates p70 S6k on Thr-389, a residue whose phosphorylation is rapamycin-sensitive in vivo and necessary for S6 kinase activity.
  13. [13]
    Twenty-five years of mTOR: Uncovering the link from nutrients to ...
    Oct 25, 2017 · We now know that one pathway—the mechanistic target of rapamycin (mTOR) pathway—is the major nutrient-sensitive regulator of growth in animals ...
  14. [14]
    mTOR kinase structure, mechanism and regulation by the rapamycin ...
    Here we present the 3.2 Å crystal structure of a ~1500 amino acid mTOR-mLST8 complex containing the FAT, FRB, kinase and FATC domains, as well as the structures ...
  15. [15]
    1FAP: THE STRUCTURE OF THE IMMUNOPHILIN ... - RCSB PDB
    Jul 23, 1997 · 1FAP is the structure of the FKBP12-rapamycin complex interacting with human FRAP, where rapamycin binds to two proteins, FKBP12 and FRAP.
  16. [16]
    Upstream and downstream of mTOR - Genes & Development
    The upstream regulators of mTOR. The mammalian target of rapamycin (mTOR) was identified and cloned (Brown et al. 1994; Chiu et al. 1994; Sabatini et al ...Missing: FRAP1 | Show results with:FRAP1
  17. [17]
    4.4 Å Resolution Cryo-EM structure of human mTOR Complex 1 - PMC
    Dec 1, 2016 · The structure shows that the kinase domain adopts a canonical protein kinase conformation and provides a model for the inhibition of mTORC1 by ...Missing: lobe arm
  18. [18]
  19. [19]
    An ATP-competitive Mammalian Target of Rapamycin Inhibitor ... - NIH
    4 In in vitro kinase assays using immuno-purified mTORC1 or mTORC2, Torin1 inhibits both mTOR-containing complexes with IC50 values between 2 and 10 nm (Fig. 1A) ...
  20. [20]
  21. [21]
  22. [22]
  23. [23]
  24. [24]
  25. [25]
  26. [26]
  27. [27]
  28. [28]
  29. [29]
  30. [30]
    Rheb GTPase is a direct target of TSC2 GAP activity and regulates ...
    Rheb GTPase is a direct target of TSC2 GAP activity and regulates mTOR signaling. Genes Dev. 2003 Aug 1;17(15):1829-34. ... Ken Inoki , Yong Li, Tian Xu ...
  31. [31]
    The Rag GTPases Bind Raptor and Mediate Amino Acid Signaling ...
    Jun 13, 2008 · The Rag proteins do not directly stimulate the kinase activity of mTORC1, but, like amino acids, promote the intracellular localization of mTOR ...
  32. [32]
    mTORC1 Senses Lysosomal Amino Acids Through an Inside-Out ...
    Nov 4, 2011 · We found that the vacuolar H + –adenosine triphosphatase ATPase (v-ATPase) is necessary for amino acids to activate mTORC1.Missing: recruitment | Show results with:recruitment
  33. [33]
    TSC2 mediates cellular energy response to control cell growth and ...
    TSC2 functions as a key player in regulation of the common mTOR pathway of protein synthesis, cell growth, and viability in response to cellular energy levels.Missing: Rheb | Show results with:Rheb
  34. [34]
    AMPK phosphorylation of raptor mediates a metabolic checkpoint
    Apr 25, 2008 · The phosphorylation of raptor by AMPK is required for the inhibition of mTORC1 and cell-cycle arrest induced by energy stress. These findings ...
  35. [35]
    Regulation of mTOR function in response to hypoxia by REDD1 and ...
    Here we show that mTOR inhibition by hypoxia requires the TSC1/TSC2 tumor suppressor complex and the hypoxia-inducible gene REDD1/RTP801.
  36. [36]
    The stress-inducted proteins RTP801 and RTP801L are ... - PubMed
    We observed that two stress-induced proteins, RTP801/Redd1 and RTP801L/Redd2, potently inhibit signaling through mTOR.<|control11|><|separator|>
  37. [37]
    Rapamycin and mTOR kinase inhibitors - PMC - PubMed Central
    However, the exact mechanism of how the interaction with the FRB domain leads to inhibition of mTORC1 remains unclear. FKBP12/rapamycin inhibits mTOR ...Missing: seminal paper
  38. [38]
    A novel rapamycin analog is highly selective for mTORC1 in vivo
    Jul 19, 2019 · As shown in Fig. 1d, in our assay we found that rapamycin inhibited mTORC1 with an IC50 of 63.3 pM, whereas DL001 inhibited mTORC1 with a very ...
  39. [39]
    Rapalogs and mTOR inhibitors as anti-aging therapeutics - JCI
    (C) Chronic treatment with rapamycin inhibits both mTORC1 and mTORC2, restricting growth and impairing insulin signaling, but promoting longevity. mTOR- ...Missing: seminal paper<|separator|>
  40. [40]
    Regulation of mTOR function in response to hypoxia by REDD1 and ...
    Here we show that mTOR inhibition by hypoxia requires the TSC1/TSC2 tumor suppressor complex and the hypoxia-inducible gene REDD1/RTP801.
  41. [41]
    Rheb GTPase is a direct target of TSC2 GAP activity and regulates ...
    Rheb stimulates the phosphorylation of mTOR and plays an essential role in regulation of S6K and 4EBP1 in response to nutrients and cellular energy status. Our ...
  42. [42]
    DEPTOR Is an mTOR Inhibitor Frequently Overexpressed in Multiple ...
    May 29, 2009 · We identify DEPTOR as an mTOR-interacting protein whose expression is negatively regulated by mTORC1 and mTORC2. Loss of DEPTOR activates S6K1, Akt, and SGK1.
  43. [43]
    Identification of IRS-1 Ser-1101 as a target of S6K1 in nutrient - PNAS
    Aug 28, 2007 · S6K1 has emerged as a critical signaling component in the development of insulin resistance through phosphorylation and inhibition of IRS-1 function.
  44. [44]
    mTOR inhibitors in cancer therapy - F1000Research
    Aug 25, 2016 · The FRB—FK506 binding protein 12 (FKBP12)–rapamycin binding—domain, as its name implies, is the binding site of mTOR to FKBP12 and rapamycin.
  45. [45]
    Rapamycin differentially inhibits S6Ks and 4E-BP1 to mediate cell ...
    Nov 11, 2008 · Activation of mTORC1 positively stimulates mRNA translation via its downstream substrates S6Ks and 4E-BP1/eIF4E (4–7). Phosphorylation of 4E-BP1 ...
  46. [46]
    mTOR and S6K1 Mediate Assembly of the Translation Preinitiation ...
    mTOR modulates the activity of two important translational regulators, the ribosomal S6 kinases (S6K1 and S6K2) and the eukaryotic initiation factor 4E (eIF4E), ...
  47. [47]
    mTOR-Dependent Regulation of Ribosomal Gene Transcription ...
    Thus, mTOR plays a critical role in the regulation of ribosome biogenesis via a mechanism that requires S6K1 activation and phosphorylation of UBF.
  48. [48]
    A liaison between mTOR signaling, ribosome biogenesis and cancer
    RP mRNA translation is controlled by mTOR pathway through the 5′TOP motif. A global analysis of 5′TOP-binding partners and a drug screen for inhibitors of 5′TOP ...
  49. [49]
    Muscle-specific expression of IGF-1 blocks angiotensin II–induced ...
    Feb 1, 2005 · We used skeletal muscle–specific IGF-1–transgenic mice to identify the downstream signal pathways involved in angiotensin II–induced muscle ...
  50. [50]
    TOR signaling in plants: conservation and innovation - PMC
    Although mLST8 is also found in mTORC2, RAPTOR is not; instead, RICTOR (rapamycin-insensitive companion of mTOR) is a distinctive component of mTORC2 (Fig. 1).
  51. [51]
    Cell surface receptor kinase FERONIA linked to nutrient sensor ...
    Jan 30, 2023 · Our research reveals a novel mechanism in which FER, ROP2 and TOR are necessary to regulate RH growth in Arabidopsis in the context of the ...
  52. [52]
    TOR Is a Negative Regulator of Autophagy in Arabidopsis thaliana
    TOR is a negative regulator of autophagy in both yeast and animals, and homologs of TOR in plants control plant growth and protein synthesis.
  53. [53]
    The TOR Pathway Modulates the Structure of Cell Walls in Arabidopsis
    The target of rapamycin (TOR) pathway, which is known to regulate cell growth in eukaryotes, is shown to affect cell wall structure in plants.
  54. [54]
    Disruption of the Mouse mTOR Gene Leads to Early ... - NIH
    Heterozygous mTOR (mTOR+/−) mice do not display any overt phenotype, although mouse embryonic fibroblasts derived from these mice show a 50% reduction in mTOR ...
  55. [55]
    mTORC1 controls murine postprandial hepatic glycogen synthesis ...
    Jan 30, 2024 · Postprandial liver glycogen content was significantly decreased in mTORC1-deficient livers, evidenced by both an enzymatic assay and ...
  56. [56]
    Muscle inactivation of mTOR causes metabolic and dystrophin ...
    Dec 14, 2009 · Consistent with this, muscle-specific inactivation of the mTORC1 component raptor in raptor muscle knockout (RAmKO) mice leads to muscle atrophy ...
  57. [57]
    Inactivation of mTORC1 in the developing brain causes ... - PubMed
    May 1, 2013 · In this study, we inactivated mTORC1 in mice by deleting the gene encoding raptor in the progenitors of the developing CNS. Mice are born but ...Missing: eIF4E | Show results with:eIF4E<|separator|>
  58. [58]
    Myelination and mTOR - PMC - PubMed Central
    Dec 6, 2017 · Not all roads may lead to mTORC1. PI3K controls 50–100 targets, including Akt, which in turn potentially phosphorylates more than 100 substrates ...
  59. [59]
    Report Increased Mammalian Lifespan and a Segmental and Tissue ...
    Sep 12, 2013 · Treatment of mice beginning at 20 months of age with rapamycin, a pharmacological inhibitor of mTOR, results in an extension of lifespan that ...Reduced Mtor Expression... · Biomarkers Of Aging Are... · Mtor Mice Have Selective...Missing: heterozygote | Show results with:heterozygote
  60. [60]
    mTOR pathway diseases: challenges and opportunities from bench ...
    May 27, 2025 · We outline the germline and somatic mutations in the mTOR pathway that cause rare diseases and summarise the prevalence, genetic basis, clinical ...
  61. [61]
    TSC1 and TSC2 gene mutations and their implications for treatment ...
    The incidence of TSC is 1/10,000 births, and its prevalence in the general population of Europe has been estimated to be 8.8/100,000 (Orphanet: Tuberous ...
  62. [62]
    Genotype/phenotype correlation in 325 individuals referred for a ...
    Approximately 70% to 80% of individuals who meet definite diagnostic criteria have a small identifiable TSC1 or TSC2 gene mutation. The remaining individuals ...
  63. [63]
    Mosaicism in tuberous sclerosis complex: Lowering the threshold for ...
    Aug 28, 2022 · The estimated prevalence is 1 in 6,000 to 1 in 10,000 live births (Northrup et al., 2013). Frequent neuropsychiatric features of TSC include ...
  64. [64]
    Functional and structural analyses of novel Smith-Kingsmore ...
    Jul 1, 2021 · Smith-Kingsmore Syndrome is a rare disease caused by damage in a gene named MTOR that is associated with excessive growth of the head and brain, ...
  65. [65]
    Expanding the phenotype of MTOR-related disorders and the Smith ...
    May 7, 2020 · Heterozygous germline mutations in mammalian target of rapamycin (MTOR) (OMIM 601231) are known to underlie Smith-Kingsmore syndrome (SKS; ...
  66. [66]
    mTOR Pathway Mutations Cause Hemimegalencephaly and Focal ...
    Focal malformations of cortical development (MCDs), including focal cortical dysplasia (FCD) and hemimegalencephaly (HME), are important causes of intractable ...
  67. [67]
    Article Somatic Mutations Activating the mTOR Pathway in Dorsal ...
    Dec 26, 2017 · Focal cortical dysplasia (FCD) and hemimegalencephaly (HME) are epileptogenic neurodevelopmental malformations caused by mutations in mTOR ...
  68. [68]
    Detection of somatic and germline pathogenic variants in adult ...
    Mar 19, 2024 · This study investigates the prevalence of pathogenic variants in the mechanistic target of rapamycin (mTOR) pathway in surgical specimens of malformations of ...
  69. [69]
    Activating PIK3CA mutation promotes overgrowth of adipose tissue ...
    Oct 6, 2025 · Activating PIK3CA mutation promotes overgrowth of adipose tissue via inhibiting lipophagy in macrodactyly | Cell Death & Disease.
  70. [70]
    Dissecting Genotype-Phenotype Relationships in the PI3K-AKT ...
    Aug 28, 2025 · Overgrowth intellectual disability syndromes (OGIDs) caused by mutations in the PI3K-AKT-MTOR pathway present significant neurobehavioral ...
  71. [71]
    mTOR Cross-Talk in Cancer and Potential for Combination Therapy
    Given the many oncogenes or tumor suppressors linked to mTOR signaling, it is estimated that mTORC1 function might be hyperactivated in up to 70% of all human ...
  72. [72]
    PTEN and the PI3-Kinase Pathway in Cancer - PMC - PubMed Central
    In the central nervous system, loss of 10q, including PTEN, is found in 70% of glioblastomas and marks the transition to the most aggressive grade of astrocytic ...
  73. [73]
    Mutations in the PI3K/PTEN/TSC2 Pathway Contribute to ...
    The consequence of PTEN or TSC2 loss is dysregulated mTOR activity. A direct approach to mimic these mutations is to overexpress mTOR. We investigated the ...
  74. [74]
    Hypoxia-inducible Factor 1α Is Regulated by the Mammalian Target ...
    Our work explains why human cancers with aberrant mTOR signaling are prone to angiogenesis ... mTOR enhances HIF1α transcription through assembly of the HIF1α ...
  75. [75]
    mTORC1 drives HIF-1α and VEGF-A signalling via multiple ...
    mTORC1 acts as a central mediator of STAT3, HIF-1α, VEGF-A and angiogenesis via multiple signalling mechanisms.
  76. [76]
    Prognostic relevance of the mTOR pathway in renal cell carcinoma
    May 18, 2007 · 1). The mTOR pathway has a central role in the regulation of cell growth and increasing evidence suggests its dysregulation in cancer.13 ...Abstract · MATERIALS AND METHODS · RESULTS · DISCUSSION
  77. [77]
    Interplay Between pVHL and mTORC1 Pathways in Clear-Cell ...
    mTOR complex 1 (mTORC1) is implicated in the pathogenesis of renal cell carcinoma (RCC), including clear-cell type (ccRCC). mTORC1 is thought to be activated in ...
  78. [78]
    Cancer Stem Cells in Glioblastoma: The Role of the mTOR Pathway
    The mTOR pathway regulates cell growth and migration of NSCs (50), and consequently, inhibition of the mTOR pathway in CSCs by rapamycin-family inhibitors ...
  79. [79]
    The PI3K/Akt/mTOR Signaling Pathway in Triple-Negative Breast ...
    Jul 3, 2025 · The PI3K/Akt/mTOR pathway is a key cell signaling pathway involved in many processes involved in carcinogenesis, such as proliferation, ...
  80. [80]
    Triple-negative breast cancer: new data in the regulation of the ...
    Dec 26, 2024 · The PI3K/AKT/mTOR pathway is frequently hyperactivated in triple-negative breast cancers, and ongoing clinical trials are targeting this pathway.
  81. [81]
    Mammalian target of rapamycin up-regulation of pyruvate kinase ...
    We identified mTOR as a central activator of the Warburg effect by inducing PKM2 and other glycolytic enzymes under normoxic conditions.
  82. [82]
    Critical role of mTOR in regulating aerobic glycolysis in ...
    Nov 25, 2020 · It has been revealed that PFKFB is a key glycolysis regulator that modulates fructose 2,6-bisphosphate levels and glucose uptake (107). PFKFB3 ...
  83. [83]
    4E-BP1 is a tumor suppressor protein reactivated by mTOR ... - NIH
    High-dose rapamycin induces apoptosis in human cancer cells by dissociating mTOR complex 1 and suppressing phosphorylation of 4E-BP1. Cell Cycle 2011;10 ...
  84. [84]
    High-dose rapamycin induces apoptosis in cancer cells
    High-dose rapamycin induces apoptosis in human cancer cells by dissociating mTOR complex 1 and suppressing phosphorylation of 4E-BP1 - PMC.
  85. [85]
    mTORC1 and mTORC2 in cancer and the tumor microenvironment
    This review will summarize recent advances in dissecting the relative contributions of mTORC1 versus mTORC2 in cancer, their role in tumor-associated blood ...
  86. [86]
    Uncoupling of Akt and mTOR signaling drives resistance ... - Science
    Feb 7, 2025 · We discovered that restoration of mTOR signaling was the early dominant driver of resistance to Akt inhibition.
  87. [87]
    Targeting PI3K/Akt signal transduction for cancer therapy - Nature
    Dec 16, 2021 · In this review, we discuss the roles of the PI3K/Akt pathway in various cancer phenotypes, review the current statuses of different PI3K/Akt inhibitors, and ...<|control11|><|separator|>
  88. [88]
    Pten Mutations Alter Brain Growth Trajectory and Allocation of Cell ...
    Jul 15, 2015 · Heterozygous mutations in PTEN (PTEN+/−), which encodes a negative regulator of the PI3K-Akt-mTOR pathway, are a risk factor for ASD and ...
  89. [89]
    Genetic Suppression of mTOR Rescues Synaptic and Social ... - NIH
    Heterozygous mutations in PTEN, which encodes a negative regulator of the mTOR and β-catenin signaling pathways, cause macrocephaly/autism syndrome.
  90. [90]
    Mechanisms of brain overgrowth in autism spectrum disorder with ...
    One of the most commonly affected signaling pathways in macrocephalic ASD is the PI3K-AKT–mTOR pathway (Table 1), which has been implicated in 47.6% of patients ...
  91. [91]
    Hyperconnectivity of prefrontal cortex to amygdala projections in a ...
    Nov 15, 2016 · We show that mTOR signalling is dysregulated during early postnatal development in the cerebral cortex of germ-line heterozygous Pten mutant mice.
  92. [92]
    mTOR Pathway Somatic Pathogenic Variants in Focal Malformations ...
    We identified causal mTOR pathway gene variants in 66.7% (14/21) of patients, of which 13 were somatic with AAF ranging between 0.6% and 12.0%.
  93. [93]
    mTOR pathway activation in focal cortical dysplasia - ScienceDirect
    TSC is an autosomal dominant disorder resulting from mutations in TSC1/TSC2 genes, which regulate the mTOR signalling cascade. While the primary molecular ...
  94. [94]
    mTOR Hyperactivity Levels Influence the Severity of Epilepsy and ...
    Apr 3, 2019 · We present novel findings that neuronal mTOR hyperactivity levels correlate with the severity of epilepsy and associated neuropathology in a mouse model of TSC ...
  95. [95]
    mTOR signaling & Alzheimer's: What we know & where are we?
    Sep 18, 2023 · To illustrate, mTOR activation impairs autophagy, a mechanism which reduces the accumulation of Aβ. When autophagy is impaired, the higher ...
  96. [96]
    Trilateral association of autophagy, mTOR and Alzheimer's disease
    Downregulation of mTOR signaling triggers autophagy activation, degrading the misfolded proteins and preventing the further accumulation of misfolded proteins ...
  97. [97]
    Intracellular accumulation of tau inhibits autophagosome formation ...
    Jul 7, 2022 · Our findings reveal that AD-like tau accumulation inhibits autophagosome formation and induces autophagy deficits by activating the TIA1/amino acid/mTORC1 ...
  98. [98]
    Pharmacological mTOR Inhibition for AD Tau Clearance
    Nov 5, 2022 · In this study we aimed to reduce tau pathology, a hallmark of Alzheimer's Disease (AD), by activating mTOR-dependent autophagy in a transgenic mouse model of ...
  99. [99]
    mTORC1-selective inhibitors rescue cellular phenotypes in TSC ...
    Jul 27, 2025 · Our results indicate that the novel, selective mTORC1 inhibitors nearly fully reversed the cellular and functional deficits of TSC2-/- iPSC-derived neurons.Missing: Alzheimer's | Show results with:Alzheimer's
  100. [100]
    Subcellular proteomics and iPSC modeling uncover reversible ...
    Mar 10, 2025 · 5: mTOR signaling is expressed in axonal spheroids and is associated with Alzheimer's pathology. ... 6: A human iPSC-derived AD model demonstrates ...Missing: TSC | Show results with:TSC
  101. [101]
    Dysregulation of mTOR Signaling in Fragile X Syndrome - PMC
    During activation, mTOR phosphorylates 4E-BP, promoting its release from eIF4E, enabling eIF4E to associate with eIF4G. The eIF4G/eIF4E complex associates ...
  102. [102]
    Dysregulation of mTOR Signaling in Fragile X Syndrome
    Jan 13, 2010 · During activation, mTOR phosphorylates 4E-BP, promoting its release from eIF4E, enabling eIF4E to associate with eIF4G. The eIF4G/eIF4E ...
  103. [103]
    A randomized controlled trial with everolimus for IQ and autism in ...
    Children randomized to receive everolimus treatment were less likely to have had epilepsy and had slightly higher baseline IQ values (not significant). It could ...
  104. [104]
    Earlier treatment may help reverse autism-like behavior in tuberous ...
    Oct 9, 2018 · New research suggests that rapamycin, given to young mice, can reverse social impairments in autism. But the drug didn't work in older mice.
  105. [105]
    Trial results temper hopes of tumor drug for treating autism
    A drug that treats tumors and epilepsy in people with tuberous sclerosis complex does not boost their intelligence or ease autism traits.
  106. [106]
    Persistent mTORC1 signaling in cell senescence results from ...
    Enhanced mTORC1 activity drives characteristic phenotypes of senescence, although the underlying mechanisms responsible for increased activity are not well ...
  107. [107]
    mTORC1 induces plasma membrane depolarization and promotes ...
    Mar 8, 2022 · We found that mTORC1 accelerated preosteoblast senescence in vitro and in a mouse model. Mechanistically, mTORC1 induced a change in the ...
  108. [108]
    Calorie restriction: is AMPK as a key sensor and effector? - PMC
    Interestingly, AMPK activation is the best-described intracellular trigger for mTOR inhibition. By phosphorylating both Raptor (a component of the mTORC1 ...
  109. [109]
    Rapamycin fed late in life extends lifespan in genetically ... - NIH
    In a separate study, rapamycin fed to mice beginning at 270 days of age also increased survival in both males and females, based on an interim analysis ...Missing: seminal | Show results with:seminal
  110. [110]
    Rapamycin for longevity: the pros, the cons, and future perspectives
    Jun 19, 2025 · Rapamycin has recently gained significant attention for anti-aging therapy and seizure treatment via mTOR pathway inhibition.
  111. [111]
    Molecular Mechanisms of Autophagy Decline during Aging - NIH
    Aug 16, 2024 · Age-related changes in upstream autophagy regulators, such as mTORC1 and AMPK, impact nutrient and energy-sensing pathways and can contribute ...
  112. [112]
    TSC1 controls IL-1β expression in macrophages via mTORC1 ... - NIH
    We found that TSC1 deficiency resulted in impaired expression of pro-IL-1β in macrophages following lipopolysaccharide stimulation.Missing: inflammaging | Show results with:inflammaging
  113. [113]
    Dual mTORC1/C2 inhibitors: gerosuppressors with potential anti ...
    Sep 15, 2015 · Several studies indicate that the link between mTOR signaling and aging and longevity is conserved across species [14]. ... dual mTOR inhibitors ...
  114. [114]
    Regulation and function of mTOR signalling in T cell fate decisions
    Apr 20, 2012 · c | mTORC2 promotes TH2 cell differentiation via two mechanisms: by preventing the expression of SOCS5, which is a negative regulator of IL-4R ...
  115. [115]
    mTORC1 and mTORC2 selectively regulate CD8+ T cell differentiation
    Apr 20, 2015 · In this report, we demonstrate a critical role for mTORC1 and mTORC2 in regulating CD8 + T cell effector and memory differentiation.Abstract · Introduction · Results · Discussion
  116. [116]
    The TSC-mTOR pathway regulates macrophage polarization - Nature
    Nov 27, 2013 · In the current study, we elucidate a role for mTOR in macrophage polarization. We demonstrate that Tsc1Δ/Δ macrophages have a marked defect in ...
  117. [117]
    Sustained mTORC1 activation in activated T cells impairs vaccine ...
    Apr 18, 2025 · Sustained mTORC1 activation in activated T cells impairs vaccine responses in older individuals | Science Advances.Missing: autoimmunity | Show results with:autoimmunity
  118. [118]
    Review Therapeutic mTOR blockade in systemic autoimmunity
    mTOR blockade may extend healthy lifespan by abrogating inflammation induced by viral infections and autoimmunity. This review provides a mechanistic assessment ...
  119. [119]
    TOR: The expanding role of mTOR in regulating immune responses
    Jun 30, 2025 · Here, we provide an updated review of our current understanding of mTOR's comprehensive role in immune cell biology. In addition, we offer ...
  120. [120]
    Multi-omic Analysis of Human B-cell Activation Reveals a Key ...
    Multi-omic Analysis of Human B-cell Activation Reveals a Key Lysosomal BCAT1 Role in mTOR Hyperactivation by B-cell receptor and TLR9 ... cell exhaustion, BCAT1 ...
  121. [121]
    mTOR mutations in Smith-Kingsmore syndrome - PubMed
    Germline or mosaic mutations of the mTOR gene have been detected in all patients. The mTOR gene is a key regulator of cell growth, cell proliferation ...
  122. [122]
    Entry - #616638 - SMITH-KINGSMORE SYNDROME; SKS - OMIM
    A germline MTOR mutation in aboriginal Australian siblings with intellectual disability, dysmorphism, macrocephaly, and small thoraces.
  123. [123]
    Functional and structural analyses of novel Smith-Kingsmore ...
    Jul 1, 2021 · Germline and somatic mosaic MTOR pathogenic variants [5] cause Smith-Kingsmore Syndrome (SKS), which is characterized by macrocephaly/ ...
  124. [124]
    a neglected mTOR target for lymphangioleiomyomatosis - PMC - NIH
    Sep 27, 2023 · In LAM cells, mutations in TSC genes result in de-repression of the GTPase Rheb and hyperactivation of mTOR in mTORC1. Ribosome biogenesis and ...
  125. [125]
    Lymphangioleiomyomatosis in patients with tuberous sclerosis
    Mar 26, 2024 · Lymphangioleiomyomatosis (LAM) is common in tuberous sclerosis complex (TSC) yet under recognised with management mostly based upon evidence ...
  126. [126]
    Lymphangioleiomyomatosis — a wolf in sheep's clothing - JCI
    Activation of mTOR in TSC-deficient cells appears to promote the Warburg effect by increasing HIF-1α and sterol regulatory element–binding proteins (SREBP1 and ...
  127. [127]
    Sirt1 ameliorates systemic sclerosis by targeting the mTOR pathway
    May 3, 2017 · An improvement in mammalian target of rapamycin (mTOR) was identified in the fibroblasts of SSc patients and the skin lesions of BLM mice.
  128. [128]
    Metformin ameliorates scleroderma via inhibiting Th17 cells and ...
    May 4, 2021 · To investigate how metformin modulates the inflammatory process in skin fibroblasts, we measured mTOR-STAT3 signaling in skin fibroblasts and ...
  129. [129]
    Resveratrol Ameliorates Systemic Sclerosis via Suppression of ...
    Dec 2, 2020 · Res was capable of elevating the SIRT1 level in fibroblasts and partially reversing mTOR-dependent induction of fibrosis and inflammation.Materials And Methods · Shared Kegg Pathways... · Results
  130. [130]
    Impaired autophagy: The collateral damage of lysosomal storage ...
    Dec 17, 2020 · Autophagy induction by rapamycin – the mTORC1 allosteric inhibitor – has shown benefits in a small subset of LSDs, including a Drosophila model ...
  131. [131]
    mTORC1 hyperactivation arrests bone growth in lysosomal storage ...
    Sep 5, 2017 · In addition, mTORC1 limits the delivery of substrates to the lysosomes by suppressing (macro)autophagy, an evolutionarily conserved trafficking ...
  132. [132]
    mTOR hyperactivity mediates lysosomal dysfunction in Gaucher's ...
    Treatment with the mTOR inhibitor Torin1 upregulated lysosomal biogenesis and enhanced autophagic clearance in GD neurons, confirming that lysosomal ...
  133. [133]
    Galectin-3: a novel biomarker of glycogen storage disease type III
    Apr 14, 2025 · Glycogen storage disease type III (GSDIII) is a rare genetic disorder leading to abnormal glycogen storage in the liver and skeletal muscle.
  134. [134]
    mTORC1 controls murine postprandial hepatic glycogen synthesis ...
    Apr 1, 2024 · mTORC1 is a well-established insulin target and contributes to the postprandial control of liver lipid metabolism, autophagy, and protein synthesis.
  135. [135]
    PIK3CA-Related Overgrowth Spectrum - GeneReviews - NCBI - NIH
    Aug 15, 2013 · There is no cure for PIK3CA-related overgrowth syndrome (PROS). ... mTOR, resulting in abnormal growth of various tissues and vascular ...
  136. [136]
    PIK3CA-Related Overgrowth Spectrum - Symptoms, Causes ...
    Mar 3, 2025 · Different subtypes within PROS include: CLAPO syndrome, CLOVES syndrome ... mTOR pathway are under investigation for the treatment of PROS.
  137. [137]
    Somatic Overgrowth Disorders of the PI3K/AKT/mTOR Pathway ...
    These syndromes include Proteus syndrome and other AKT-related disorders, PIK3CA-Related Overgrowth Spectrum, and mTOR-related disorders. Tuberous sclerosis ...
  138. [138]
    Drug Approval Package: Rapamune (Sirolimus) NDA# 021083
    Mar 30, 2001 · Rapamune (Sirolimus) Oral Solution. Company: Wyeth-Ayerst. Research Application No.: 021083. Approval Date: 19/15/1999. Date created: March 30, 2001.
  139. [139]
    Drug Approval Package: Torisel (Temsirolimus) NDA #022088
    Jul 30, 2007 · Torisel (Temsirolimus) Intravenous Solution Company: Wyeth Pharmaceuticals Inc. Application No.: 022088. Approval Date: 05/30/2007.
  140. [140]
    Afinitor (everolimus) FDA Approval History - Drugs.com
    FDA Approved: Yes (First approved March 30, 2009) ; Brand name: Afinitor ; Generic name: everolimus ; Dosage form: Tablets ; Company: Novartis Pharmaceuticals ...
  141. [141]
    Current development of the second generation of mTOR inhibitors ...
    Temsirolimus and everolimus were recently approved by the Food and Drug Administration (FDA) for the treatment of advanced/metastatic renal cell carcinoma[60], ...Missing: sirolimus approvals
  142. [142]
    Characterization of Torin2, an ATP-competitive inhibitor of mTOR ...
    Similar to the earlier generation compound Torin1 and in contrast to other reported mTOR inhibitors, Torin2 inhibited mTOR kinase and mTORC1 signaling ...
  143. [143]
    [PDF] New Inhibitors of the PI3K-Akt-mTOR Pathway - Shokat Lab
    Jun 12, 2010 · The ability of rapamycin to act as a dual inhibitor of mTORC1/2 challenged the explanation that it was a poor anti- cancer therapeutic because ...<|separator|>
  144. [144]
    mTORC1-selective inhibitors rescue cellular phenotypes in TSC ...
    These data suggest that mTORC1-specific compounds could provide clinical therapeutic benefit similar to rapamycin without the same side effects.
  145. [145]
    The allosteric mechanism of mTOR activation can inform bitopic ...
    Their mutations promote resistance to the monovalent drug interventions. Patients harboring the A2034V and F2108L mutations within the mTOR FRB domain ...
  146. [146]
    picroside ii as a promising natural inhibitor - Lapin Press Journals
    Sep 1, 2025 · In silico evaluation of plant-derived bioactive compounds targeting the mtor pathway in breast cancer: picroside ii as a promising natural ...
  147. [147]
    Current Insights of the Potential Plant Bioactive Compounds on ...
    TABLE 1 | Plant- derived bioactive compounds acting through mTOR pathways to treat various oncological disorders. Plant bioactive compounds. Plant source.
  148. [148]
    Everolimus with Reduced Calcineurin Inhibitor Exposure in Renal ...
    Background Everolimus permits reduced calcineurin inhibitor (CNI) exposure, but the efficacy and safety outcomes of this treatment after kidney transplant ...
  149. [149]
    Conversion From Calcineurin Inhibitors to Mammalian Target of ...
    Sep 2, 2021 · Posttransplant patients have a better graft function and lower incidence of malignancy after conversion from CNI to mTORi therapy.
  150. [150]
    Clinical outcomes after switch to mTOR inhibitors in kidney ...
    The use of mTOR inhibitors drugs appears to be safe in the managgement of specific renal transplant recipients, with a low rejection rate and good survival.
  151. [151]
    WCN25-2410 LONG-TERM RESULTS OF THE REDUCTION OF ...
    Patient survival at 1, 5 and 10 years was 97.6%, 88.1% and 73.3%, respectively. Graft survival at 1, 5, and 10 years was 96%, 84%, and 65%, respectively. In Cox ...
  152. [152]
    A pharmacological rationale for improved everolimus dosing in ...
    In transplant patients, the approved dose of 0.75–1 mg twice daily (BID) results in subtherapeutic trough levels (<6 μg l–1) and that a higher starting dose of ...
  153. [153]
    Everolimus long-term safety and efficacy in subependymal giant cell ...
    As a result of this trial, the US Food and Drug Administration (FDA) approved everolimus for patients with SEGA associated with TSC who are not candidates for ...
  154. [154]
    Everolimus for subependymal giant cell astrocytoma: 5‐year final ...
    Patients aged ≥ 3 years with a definite diagnosis of TSC and increasing SEGA lesion size (≥2 magnetic resonance imaging scans) received everolimus starting at ...
  155. [155]
    Everolimus for Subependymal Giant-Cell Astrocytomas in Tuberous ...
    Everolimus therapy was associated with marked reduction in the volume of subependymal giant-cell astrocytomas and seizure frequency.Missing: TSC approval
  156. [156]
    Everolimus in patients with autosomal dominant polycystic kidney ...
    Aug 26, 2010 · Everolimus slowed the increase in total kidney volume of patients with ADPKD but did not slow the progression of renal impairment.Missing: off- label
  157. [157]
    Drug repurposing in autosomal dominant polycystic kidney disease
    Mar 2, 2023 · In this review, we focus on the repurposing approaches to identify suitable drug candidates to treat autosomal dominant polycystic kidney disease.
  158. [158]
    Synergism of dual AAV gene therapy and rapamycin rescues GSDIII ...
    Simultaneous treatment of a murine GSDIII model with rapamycin and a GDE-expressing, liver- and muscle-targeting AAV vector resulted in a synergic effect at the ...
  159. [159]
    Links between autophagy and disorders of glycogen metabolism
    Recent studies showed that several of the glycogen storage disorders have abnormal autophagy which can disturb normal cellular metabolism and/or mitochondrial ...
  160. [160]
    Everolimus Side Effects: Common, Severe, Long Term - Drugs.com
    Mar 27, 2025 · Very common (10% or more): Stomatitis (78%), diarrhea (50%), constipation (38%), abdominal pain (36%), nausea (32%), vomiting (29%), dry mouth ( ...
  161. [161]
    Stomatitis And Everolimus: A Review Of Current Literature On 8,201 ...
    Common side effects are anemia, fatigue, hyperglycemia, hyperlipidemia, stomatitis, rash, and thrombocytopenia., The terms “oral mucositis” and “stomatitis” are ...
  162. [162]
    Targeting mTOR Kinase for Cancer Treatment: A Comprehensive ...
    Oct 5, 2025 · The Ras/MAPK and PI3K/Akt/mTOR pathways are tightly linked to cancer progression. Tumors with PIK3CA/PTEN mutations or Akt hyperactivation ...
  163. [163]
    Targeting mTOR Kinase for Cancer Treatment - PubMed
    Additionally, we summarize the findings from major clinical trials, including FDA-approved mTOR inhibitors like everolimus and temsirolimus, and non-FDA- ...
  164. [164]
    Recent advances and limitations of mTOR inhibitors in the treatment ...
    Sep 15, 2022 · Second-generation mTOR inhibitors work as ATP-competitors by competing with ATP molecules for attaching to the mTOR kinase domain. ... IC50 of 50 ...Missing: IC50 | Show results with:IC50
  165. [165]
    PI3K/mTOR inhibition for the treatment of pediatric high-grade gliomas
    Sep 11, 2025 · Encouragingly, novel brain-penetrant PI3K/mTOR inhibitors offer new opportunities for treatment, but combining these agents with other therapies ...
  166. [166]
    Inhibition of Akt/mTOR pathway overcomes intrinsic resistance to ...
    The present results show that the effect of dasatinib in TNBC is independent of Src inhibition, and that Akt/mTOR inhibition might be an effective strategy to ...
  167. [167]
    Regulation of transcriptome plasticity by mTOR signaling pathway
    Aug 14, 2025 · Recent insights into mTOR's regulation of alternative splicing and polyadenylation reveal a sophisticated mechanism by which mTOR influences RNA ...
  168. [168]
    Perfect match: mTOR inhibitors and tuberous sclerosis complex
    Mar 4, 2022 · mTOR inhibitors have shown considerable success in multiple clinical trials for the treatment of TSC, including neurological, pulmonary, cardiac, renal, and ...
  169. [169]
    Mutations in TSC1, TSC2, and MTOR Are Associated with Response ...
    We found that mutations in MTOR, TSC1, or TSC2 were more common in patients who experienced a response from rapalogs than in those with rapid progression. This ...
  170. [170]
    Rictor, a Novel Binding Partner of mTOR, Defines a Rapamycin ...
    mTOR interacts with the raptor and GβL proteins 1, 2, 3 to form a complex that is the target of rapamycin. Here, we demonstrate that mTOR is also part of a ...Missing: affinity | Show results with:affinity
  171. [171]
    PRAS40 and PRR5-Like Protein Are New mTOR Interactors that ...
    PRAS40 binds mTORC1 via Raptor, and is an mTOR phosphorylation substrate. PRAS40 inhibits mTORC1 autophosphorylation and mTORC1 kinase activity toward eIF-4E ...Missing: DEPTOR | Show results with:DEPTOR
  172. [172]
    The Rapamycin-Binding Domain of the Protein Kinase mTOR ... - NIH
    Screening a library of FRB mutants using a three-hybrid assay in yeast (17) provided many FRB mutants that activated reporter gene expression in response to ...
  173. [173]
  174. [174]
  175. [175]
    The Ras-ERK and PI3K-mTOR Pathways: Cross-talk and ... - NIH
    AKT phosphorylation of TSC2 releases TSC inhibition of the GTPase RHEB (Ras homolog enriched in brain). RHEB-GTP directly activates mTORC1 ([6], Figure 1b).The Ras-Erk Pathway · The Pi3k-Mtor Pathway · Agc Kinase Promiscuity
  176. [176]
    Inappropriate Activation of the TSC/Rheb/mTOR/S6K Cassette ...
    Our results suggest that inappropriate activation of the Rheb/mTOR/S6K pathway imposes a negative feedback program to attenuate IRS-dependent processes such as ...
  177. [177]
    Turnover of the Active Fraction of IRS1 Involves Raptor-mTOR - NIH
    Inappropriate activation of the TSC/Rheb/mTOR/S6K cassette induces IRS1/2 depletion, insulin resistance, and cell survival deficiencies. Curr. Biol. 14:1650 ...
  178. [178]
    New developments in AMPK and mTORC1 cross-talk - Portland Press
    In response to energy and nutrient stress, AMPK inhibits mTORC1 signalling through direct and indirect phosphorylation events. mTORC1 reciprocally supresses ...
  179. [179]
    TSC2 Mediates Cellular Energy Response to Control Cell Growth ...
    TSC2 functions as a key player in regulation of the common mTOR pathway of protein synthesis, cell growth, and viability in response to cellular energy levels.
  180. [180]
    The Hippo pathway effectors YAP and TAZ promote cell growth by ...
    Nov 27, 2015 · Here we show that YAP/TAZ play an essential role in amino acid-induced mTORC1 activation, particularly under nutrient-limiting conditions.
  181. [181]
    mTORC1 signaling suppresses Wnt/β-catenin signaling ... - PNAS
    Oct 8, 2018 · mTORC1 signaling cell autonomously suppresses Wnt/β-catenin signaling through down-regulating the Wnt receptor FZD level to influence stem cell functions.
  182. [182]
    TSC2 Integrates Wnt and Energy Signals via a Coordinated ...
    TSC2 integrates Wnt and energy signals via a coordinated phosphorylation by AMPK and GSK3 to regulate cell growth.
  183. [183]
    Regulation of transcriptome plasticity by mTOR signaling pathway
    Aug 14, 2025 · The mTOR pathway regulates transcriptome plasticity by influencing RNA processing, including alternative splicing and polyadenylation, ...Missing: SRSF | Show results with:SRSF
  184. [184]
    Growth or death? Control of cell destiny by mTOR and autophagy ...
    Role of the mTOR pathway in cancer. Approximately 80% of human cancers are associated with hyperactivation of the mTOR signaling pathway. Even though mTOR ...The Mtor Pathway · Interplay Of Mtor And... · Mtorc1 And Mtorc2