MTORC1

Last updated
mTOR
5h64.jpg
mTORC1 heteromer, Human
Identifiers
Symbol MTOR
Alt. symbolsFRAP, FRAP2, FRAP1
NCBI gene 2475
HGNC 3942
OMIM 601231
RefSeq NM_004958
UniProt P42345
Other data
EC number 2.7.11.1
Locus Chr. 1 p36
Search for
Structures Swiss-model
Domains InterPro
RPTOR
Identifiers
Symbol RPTOR
Alt. symbolsKOG1, Mip1
NCBI gene 57521
HGNC 30287
OMIM 607130
RefSeq NM_001163034.1
UniProt Q8N122
Other data
Locus Chr. 17 q25.3
Search for
Structures Swiss-model
Domains InterPro

mTORC1, also known as mammalian target of rapamycin complex 1 or mechanistic target of rapamycin complex 1, is a protein complex that functions as a nutrient/energy/redox sensor and controls protein synthesis. [1] [2]

Contents

mTOR Complex 1 (mTORC1) is composed of the mTOR protein complex, regulatory-associated protein of mTOR (commonly known as raptor), mammalian lethal[ clarification needed ] with SEC13 protein 8 (MLST8), PRAS40 and DEPTOR. [2] [3] [4] This complex embodies the classic functions of mTOR, namely as a nutrient/energy/redox sensor and controller of protein synthesis. [1] [2] The activity of this complex is regulated by rapamycin, insulin, growth factors, phosphatidic acid, certain amino acids and their derivatives (e.g., L-leucine and β-hydroxy β-methylbutyric acid), mechanical stimuli, and oxidative stress. [2] [5] [6] Recently it has been also demonstrated that cellular bicarbonate metabolism can be regulated by mTORC1 signaling. [7]

The role of mTORC1 is to activate translation of proteins. [8] In order for cells to grow and proliferate by manufacturing more proteins, the cells must ensure that they have the resources available for protein production. Thus, for protein production, and therefore mTORC1 activation, cells must have adequate energy resources, nutrient availability, oxygen abundance, and proper growth factors in order for mRNA translation to begin. [4]

Activation at the lysosome

Activation of mTORC1 at the lysosome. Activation of mTORC1 at the Lysosome.jpg
Activation of mTORC1 at the lysosome.

The TSC complex

Almost all of the variables required for protein synthesis affect mTORC1 activation by interacting with the TSC1/TSC2 protein complex. TSC2 is a GTPase activating protein (GAP). Its GAP activity interacts with a G protein called Rheb by hydrolyzing the GTP of the active Rheb-GTP complex, converting it to the inactive Rheb-GDP complex. The active Rheb-GTP activates mTORC1 through unelucidated pathways. [9] Thus, many of the pathways that influence mTORC1 activation do so through the activation or inactivation of the TSC1/TSC2 heterodimer. This control is usually performed through phosphorylation of the complex. This phosphorylation can cause the dimer to dissociate and lose its GAP activity, or the phosphorylation can cause the heterodimer to have increased GAP activity, depending on which amino acid residue becomes phosphorylated. [10] Thus, the signals that influence mTORC1 activity do so through activation or inactivation of the TSC1/TSC2 complex, upstream of mTORC1.

The Ragulator-Rag complex

mTORC1 interacts at the Ragulator-Rag complex on the surface of the lysosome in response to amino acid levels in the cell. [11] [12] Even if a cell has the proper energy for protein synthesis, if it does not have the amino acid building blocks for proteins, no protein synthesis will occur. Studies have shown that depriving amino acid levels inhibits mTORC1 signaling to the point where both energy abundance and amino acids are necessary for mTORC1 to function. When amino acids are introduced to a deprived cell, the presence of amino acids causes Rag GTPase heterodimers to switch to their active conformation. [13] Active Rag heterodimers interact with raptor, localizing mTORC1 to the surface of late endosomes and lysosomes where the Rheb-GTP is located. [14] This allows mTORC1 to physically interact with Rheb. Thus the amino acid pathway as well as the growth factor/energy pathway converge on endosomes and lysosomes. Thus the Ragulator-Rag complex recruits mTORC1 to lysosomes to interact with Rheb. [15] [16]

Regulation of the Ragulator-Rag complex

Rag activity is regulated by at least two highly conserved complexes: the "GATOR1" complex containing DEPDC5, NPRL2 and NPRL3 and the ""GATOR2" complex containing Mios, WDR24, WDR59, Seh1L, Sec13. [17] GATOR1 inhibits Rags (it is a GTPase-activating protein for Rag subunits A/B) and GATOR2 activates Rags by inhibiting DEPDC5.

Upstream signaling

The General mTORC1 Pathway. General mTORC1 Pathway.jpg
The General mTORC1 Pathway.

Receptor tyrosine kinases

Akt/PKB pathway

Insulin-like growth factors can activate mTORC1 through the receptor tyrosine kinase (RTK)-Akt/PKB signaling pathway. Ultimately, Akt phosphorylates TSC2 on serine residue 939, serine residue 981, and threonine residue 1462. [18] These phosphorylated sites will recruit the cytosolic anchoring protein 14-3-3 to TSC2, disrupting the TSC1/TSC2 dimer. When TSC2 is not associated with TSC1, TSC2 loses its GAP activity and can no longer hydrolyze Rheb-GTP. This results in continued activation of mTORC1, allowing for protein synthesis via insulin signaling. [19]

Akt will also phosphorylate PRAS40, causing it to fall off of the Raptor protein located on mTORC1. Since PRAS40 prevents Raptor from recruiting mTORC1's substrates 4E-BP1 and S6K1, its removal will allow the two substrates to be recruited to mTORC1 and thereby activated in this way. [20]

Furthermore, since insulin is a factor that is secreted by pancreatic beta cells upon glucose elevation in the blood, its signaling ensures that there is energy for protein synthesis to take place. In a negative feedback loop on mTORC1 signaling, S6K1 is able to phosphorylate the insulin receptor and inhibit its sensitivity to insulin. [18] This has great significance in diabetes mellitus, which is due to insulin resistance. [21]

MAPK/ERK pathway

Mitogens, such as insulin like growth factor 1 (IGF1), can activate the MAPK/ERK pathway, which can inhibit the TSC1/TSC2 complex, activating mTORC1. [19] In this pathway, the G protein Ras is tethered to the plasma membrane via a farnesyl group and is in its inactive GDP state. Upon growth factor binding to the adjacent receptor tyrosine kinase, the adaptor protein GRB2 binds with its SH2 domains. This recruits the GEF called Sos, which activates the Ras G protein. Ras activates Raf (MAPKKK), which activates Mek (MAPKK), which activates Erk (MAPK). [22] Erk can go on to activate RSK. Erk will phosphorylate the serine residue 644 on TSC2, while RSK will phosphorylate serine residue 1798 on TSC2. [23] These phosphorylations will cause the heterodimer to fall apart, and prevent it from deactivating Rheb, which keeps mTORC1 active.

RSK has also been shown to phosphorylate raptor, which helps it overcome the inhibitory effects of PRAS40. [24]

JNK pathway

c-Jun N-terminal kinase (JNK) signaling is part of the mitogen-activated protein kinase (MAPK) signaling pathway essential in stress signaling pathways relating to gene expression, neuronal development, and cell survival. Recent studies have shown there is a direct molecular interaction where JNK phosphorylates Raptor at Ser-696, Thr-706, and Ser-863. [25] [26] Therefore, mTORC1 activity is JNK-dependent. Thus, JNK activation plays a role in protein synthesis via subsequent downstream effectors of mTORC1 such as S6 kinase and eIFs. [27]

Wnt pathway

The Wnt pathway is responsible for cellular growth and proliferation during organismal development; thus, it could be reasoned that activation of this pathway also activates mTORC1. Activation of the Wnt pathway inhibits glycogen synthase kinase 3 beta (GSK3B). [28] When the Wnt pathway is not active, GSK3B is able to phosphorylate TSC2 on Ser1341 and Ser1337 in conjunction with AMPK phosphorylation of Ser1345. It has been found that the AMPK is required to first phosphorylate Ser1345 before GSK3B can phosphorylate its target serine residues. This phosphorylation of TSC2 would activate this complex, if GSK3B were active. Since the Wnt pathway inhibits GSK3 signaling, the active Wnt pathway is also involved in the mTORC1 pathway. Thus, mTORC1 can activate protein synthesis for the developing organism. [28]

Cytokines

Cytokines like tumor necrosis factor alpha (TNF-alpha) can induce mTOR activity through IKK beta, also known as IKK2. [29] IKK beta can phosphorylate TSC1 at serine residue 487 and TSC1 at serine residue 511. This causes the heterodimer TSC complex to fall apart, keeping Rheb in its active GTP-bound state.

Energy and oxygen

Energy status

In order for translation to take place, abundant sources of energy, particularly in the form of ATP, need to be present. If these levels of ATP are not present, due to its hydrolysis into other forms like AMP, and the ratio of AMP to ATP molecules gets too high, AMPK will become activated. AMPK will go on to inhibit energy consuming pathways such as protein synthesis. [30]

AMPK can phosphorylate TSC2 on serine residue 1387, which activates the GAP activity of this complex, causing Rheb-GTP to be hydrolyzed into Rheb-GDP. This inactivates mTORC1 and blocks protein synthesis through this pathway. [31]

AMPK can also phosphorylate Raptor on two serine residues. This phosphorylated Raptor recruits 14-3-3 to bind to it and prevents Raptor from being part of the mTORC1 complex. Since mTORC1 cannot recruit its substrates without Raptor, no protein synthesis via mTORC1 occurs. [32]

LKB1, also known as STK11, is a known tumor suppressor that can activate AMPK. More studies on this aspect of mTORC1 may help shed light on its strong link to cancer. [33]

Hypoxic stress

When oxygen levels in the cell are low, it will limit its energy expenditure through the inhibition of protein synthesis. Under hypoxic conditions, hypoxia inducible factor one alpha (HIF1A) will stabilize and activate transcription of REDD1, also known as DDIT4. After translation, this REDD1 protein will bind to TSC2, which prevents 14-3-3 from inhibiting the TSC complex. Thus, TSC retains its GAP activity towards Rheb, causing Rheb to remain bound to GDP and mTORC1 to be inactive. [34] [35]

Due to the lack of synthesis of ATP in the mitochondria under hypoxic stress or hypoxia, AMPK will also become active and thus inhibit mTORC1 through its processes. [36]

Downstream signaling

Receptor Tyrosine Kinases and mTORC1. Downstream of mTORC1.png
Receptor Tyrosine Kinases and mTORC1.

mTORC1 activates transcription and translation through its interactions with p70-S6 Kinase 1 (S6K1) and 4E-BP1, the eukaryotic initiation factor 4E (eIF4E) binding protein 1, primarily via phosphorylation and dephosphorylation of its downstream targets. [1] S6K1 and 4E-BP1 modulate translation in eukaryotic cells. Their signaling will converge at the translation initiation complex on the 5' end of mRNA, and thus activate translation.

4E-BP1

Activated mTORC1 will phosphorylate translation repressor protein 4E-BP1, thereby releasing it from eukaryotic translation initiation factor 4E (eIF4E). [37] eIF4E is now free to join the eukaryotic translation initiation factor 4G (eIF4G) and the eukaryotic translation initiation factor 4A (eIF4A). [38] This complex then binds to the 5' cap of mRNA and will recruit the helicase eukaryotic translation initiation factor A (eIF4A) and its cofactor eukaryotic translation initiation factor 4B (eIF4B). [39] The helicase is required to remove hairpin loops that arise in the 5' untranslated regions of mRNA, which prevent premature translation of proteins. [40] Once the initiation complex is assembled at the 5' cap of mRNA, it will recruit the 40S small ribosomal subunit that is now capable of scanning for the AUG start codon start site, because the hairpin loop has been degraded by the eIF4A helicase. [41] Once the ribosome reaches the AUG codon, translation can begin.

S6K

Previous studies suggest that S6K signaling is mediated by mTOR in a rapamycin-dependent manner wherein S6K is displaced from the eIF3 complex upon binding of mTOR with eIF3. [42] Hypophosphorylated S6K is located on the eIF3 scaffold complex. Active mTORC1 gets recruited to the scaffold, and once there, will phosphorylate S6K to make it active. [18]

mTORC1 phosphorylates S6K1 on at least two residues, with the most critical modification occurring on a threonine residue (T389). [43] [44] This event stimulates the subsequent phosphorylation of S6K1 by PDPK1. [44] [45] Active S6K1 can in turn stimulate the initiation of protein synthesis through activation of S6 Ribosomal protein (a component of the ribosome) and eIF4B, causing them to be recruited to the pre-initiation complex. [46]

Active S6K can bind to the SKAR scaffold protein that can get recruited to exon junction complexes (EJC). Exon junction complexes span the mRNA region where two exons come together after an intron has been spliced out. Once S6K binds to this complex, increased translation on these mRNA regions occurs. [47]

S6K1 can also participate in a positive feedback loop with mTORC1 by phosphorylating mTOR's negative regulatory domain at two sites thr-2446 and ser-2448; phosphorylation at these sites appears to stimulate mTOR activity. [48] [49]

S6K also can phosphorylate programmed cell death 4 (PDCD4), which marks it for degradation by ubiquitin ligase Beta-TrCP (BTRC). PDCD4 is a tumor suppressor that binds to eIF4A and prevents it from being incorporated into the initiation complex.

Role in disease and aging

mTOR was found to be related to aging in 2001 when the ortholog of S6K, SCH9, was deleted in S. cerevisiae, doubling its lifespan. [50] This greatly increased the interest in upstream signaling and mTORC1. Studies in inhibiting mTORC1 were thus performed on the model organisms of C. elegans, fruitflies, and mice. Inhibition of mTORC1 showed significantly increased lifespans in all model species. [51] [52] Disrupting the gut microbiota of infant mice was found to lead to reduced longevity with signaling of mTORC1 implicated as a potential mechanism. [53]

Based on upstream signaling of mTORC1, a clear relationship between food consumption and mTORC1 activity has been observed. [54] Most specifically, carbohydrate consumption activates mTORC1 through the insulin growth factor pathway. In addition, amino acid consumption will stimulate mTORC1 through the branched chain amino acid/Rag pathway. Thus dietary restriction inhibits mTORC1 signaling through both upstream pathways of mTORC that converge on the lysosome. [55]

Autophagy

Autophagy is the major degradation pathway in eukaryotic cells and is essential for the removal of damaged organelles via macroautophagy or proteins and smaller cellular debris via microautophagy from the cytoplasm. [56] Thus, autophagy is a way for the cell to recycle old and damaged materials by breaking them down into their smaller components, allowing for the resynthesis of newer and healthier cellular structures. [56] Autophagy can thus remove protein aggregates and damaged organelles that can lead to cellular dysfunction. [57]

Upon activation, mTORC1 will phosphorylate autophagy-related protein 13 (Atg 13), preventing it from entering the ULK1 kinase complex, which consists of Atg1, Atg17, and Atg101. [58] This prevents the structure from being recruited to the preautophagosomal structure at the plasma membrane, inhibiting autophagy. [59]

mTORC1's ability to inhibit autophagy while at the same time stimulate protein synthesis and cell growth can result in accumulations of damaged proteins and organelles, contributing to damage at the cellular level. [60] Because autophagy appears to decline with age, activation of autophagy may help promote longevity in humans. [61] Problems in proper autophagy processes have been linked to diabetes, cardiovascular disease, neurodegenerative diseases, and cancer. [62]

Lysosomal damage

mTORC1 is positioned on lysosomes and is inhibited when lysosomal membrane is damaged through a protein complex termed GALTOR. [63] GALTOR contains galectin-8, a cytosolic lectin, which recognizes damaged lysosomal membranes by binding to the exposed glycoconjugates normally facing lysosomal lumen. Under homeostatic conditions, Galectin-8 associates with active mTOR. [63] Following membrane damage galectin-8 no longer interacts with mTOR but instead switches to complexes containing SLC38A9, RRAGA/RRAGB, and LAMTOR1 (a component of Ragulator) thus inhibiting mTOR, [63] mTOR inhibition in turn activates autophagy and starts a quality control program that removes damaged lysosomes, [63] referred to as lysophagy, [64]

Reactive oxygen species

Reactive oxygen species can damage the DNA and proteins in cells. [65] A majority of them arise in the mitochondria. [66]

Deletion of the TOR1 gene in yeast increases cellular respiration in the mitochondria by enhancing the translation of mitochondrial DNA that encodes for the complexes involved in the electron transport chain. [67] When this electron transport chain is not as efficient, the unreduced oxygen molecules in the mitochondrial cortex may accumulate and begin to produce reactive oxygen species. [68] It is important to note that both cancer cells as well as those cells with greater levels of mTORC1 both rely more on glycolysis in the cytosol for ATP production rather than through oxidative phosphorylation in the inner membrane of the mitochondria. [69]

Inhibition of mTORC1 has also been shown to increase transcription of the NFE2L2 (NRF2) gene, which is a transcription factor that is able to regulate the expression of electrophilic response elements as well as antioxidants in response to increased levels of reactive oxygen species. [70]

Though AMPK induced eNOS has been shown to regulate mTORC1 in endothelium. Unlike the other cell type in endothelium eNOS induced mTORC1 and this pathway is required for mitochondrial biogenesis. [71]

Stem cells

Conservation of stem cells in the body has been shown to help prevent against premature aging. [72] mTORC1 activity plays a critical role in the growth and proliferation of stem cells. [73] Knocking out mTORC1 results in embryonic lethality due to lack of trophoblast development. [74] Treating stem cells with rapamycin will also slow their proliferation, conserving the stem cells in their undifferentiated condition. [73]

mTORC1 plays a role in the differentiation and proliferation of hematopoietic stem cells. Its upregulation has been shown to cause premature aging in hematopoietic stem cells. Conversely, inhibiting mTOR restores and regenerates the hematopoietic stem cell line. [75] The mechanisms of mTORC1's inhibition on proliferation and differentiation of hematopoietic stem cells has yet to be fully elucidated. [76]

Rapamycin is used clinically as an immunosuppressant and prevents the proliferation of T cells and B cells. [77] Paradoxically, even though rapamycin is a federally approved immunosuppressant, its inhibition of mTORC1 results in better quantity and quality of functional memory T cells. mTORC1 inhibition with rapamycin improves the ability of naïve T cells to become precursor memory T cells during the expansion phase of T cell development . [78] This inhibition also allows for an increase in quality of these memory T cells that become mature T cells during the contraction phase of their development. [79] mTORC1 inhibition with rapamycin has also been linked to a dramatic increase of B cells in old mice, enhancing their immune systems. [75] This paradox of rapamycin inhibiting the immune system response has been linked to several reasons, including its interaction with regulatory T cells. [79]

As a biomolecular target

Activators

Resistance exercise, the amino acid L-leucine, and beta-hydroxy beta-methylbutyric acid (HMB) are known to induce signaling cascades in skeletal muscle cells that result in mTOR phosphorylation, the activation of mTORC1, and subsequently the initiation of myofibrillar protein synthesis (i.e., the production of proteins such as myosin, titin, and actin), thereby facilitating muscle hypertrophy.

The NMDA receptor antagonist ketamine has been found to activate the mTORC1 pathway in the medial prefrontal cortex (mPFC) of the brain as an essential downstream mechanism in the mediation of its rapid-acting antidepressant effects. [80] NV-5138 is a ligand and modulator of sestrin2, a leucine amino acid sensor and upstream regulatory pathway of mTORC1, and is under development for the treatment of depression. [80] The drug has been found to directly and selectively activate the mTORC1 pathway, including in the mPFC, and to produce rapid-acting antidepressant effects similar to those of ketamine. [80]

Inhibitors

There have been several dietary compounds that have been suggested to inhibit mTORC1 signaling including EGCG, resveratrol, curcumin, caffeine, and alcohol. [81] [82]

First generation drugs

Rapamycin was the first known inhibitor of mTORC1, considering that mTORC1 was discovered as being the target of rapamycin. [83] Rapamycin will bind to cytosolic FKBP12 and act as a scaffold molecule, allowing this protein to dock on the FRB regulatory region (FKBP12-Rapamycin Binding region/domain) on mTORC1. [84] The binding of the FKBP12-rapamycin complex to the FRB regulatory region inhibits mTORC1 through processes not yet known. mTORC2 is also inhibited by rapamycin in some cell culture lines and tissues, particularly those that express high levels of FKBP12 and low levels of FKBP51. [85] [86] [87]

Rapamycin itself is not very water soluble and is not very stable, so scientists developed rapamycin analogs, called rapalogs, to overcome these two problems with rapamycin. [88] These drugs are considered the first generation inhibitors of mTOR. [89] These other inhibitors include everolimus and temsirolimus. Compared with the parent compound rapamycin, everolimus is more selective for the mTORC1 protein complex, with little impact on the mTORC2 complex. [90] mTORC1 inhibition by everolimus has been shown to normalize tumor blood vessels, to increase tumor-infiltrating lymphocytes, and to improve adoptive cell transfer therapy. [91]

Sirolimus, which is the drug name for rapamycin, was approved by the U.S. Food and Drug Administration (FDA) in 1999 to prevent against transplant rejection in patients undergoing kidney transplantation. [92] In 2003, it was approved as a stent covering for widening arteries to prevent against future heart attacks. [93] In 2007, mTORC1 inhibitors began being approved for treatments against cancers such as renal cell carcinoma. [94] In 2008 they were approved for treatment of mantle cell lymphoma. [95] mTORC1 inhibitors have recently been approved for treatment of pancreatic cancer. [96] In 2010 they were approved for treatment of tuberous sclerosis. [97]

Second generation drugs

The second generation of inhibitors were created to overcome problems with upstream signaling upon the introduction of first generation inhibitors to the treated cells. [98] One problem with the first generation inhibitors of mTORC1 is that there is a negative feedback loop from phosphorylated S6K, that can inhibit the insulin RTK via phosphorylation. [99] When this negative feedback loop is no longer there, the upstream regulators of mTORC1 become more active than they would otherwise would have been under normal mTORC1 activity. Another problem is that since mTORC2 is resistant to rapamycin, and it too acts upstream of mTORC1 by activating Akt. [88] Thus signaling upstream of mTORC1 still remains very active upon its inhibition via rapamycin and the rapalogs. Rapamycin and its analogues also have procoagulant side effects caused by off-target binding of the activated immunophilin FKBP12, which are not produced by structurally unrelated inhibitors of mTORC such as gedatolisib, WYE-687 and XL-388. [100]

Second generation inhibitors are able to bind to the ATP-binding motif on the kinase domain of the mTOR core protein itself and abolish activity of both mTOR complexes. [98] [101] [102] [103] In addition, since the mTOR and the PI3K proteins are both in the same phosphatidylinositol 3-kinase-related kinase (PIKK) family of kinases, some second generation inhibitors have dual inhibition towards the mTOR complexes as well as PI3K, which acts upstream of mTORC1. [88] As of 2011, these second generation inhibitors were in phase II of clinical trials.

Third generation drugs

The third generation of inhibitors were created following the realization that many of the side effects of rapamycin and rapamycin analogs were mediated not as a result of direct inhibition of mTORC1, but as a consequence of off-target inhibition of mTORC2. [104] [105] Rapamycin analogs such as DL001, that are more selective for mTORC1 than sirolimus, have been developed and in mice have reduced side effects. [106] mTORC1 inhibitors that have novel mechanisms of action, for example peptides like PRAS40 and small molecules like HY-124798 (Rheb inhibitor NR1), which inhibit the interaction of mTORC1 with its endogenous activator Rheb, are also being developed. [107] [108] Some glucose transporter inhibitors such as NV-5440 and NV-6297 are also selective inhibitors of mTORC1 [109]

There have been over 1,300 clinical trials conducted with mTOR inhibitors since 1970. [110]

Related Research Articles

<span class="mw-page-title-main">Protein kinase B</span> Set of three serine threonine-specific protein kinases

Protein kinase B (PKB), also known as Akt, is the collective name of a set of three serine/threonine-specific protein kinases that play key roles in multiple cellular processes such as glucose metabolism, apoptosis, cell proliferation, transcription, and cell migration.

<span class="mw-page-title-main">ATM serine/threonine kinase</span> Mammalian protein found in Homo sapiens

ATM serine/threonine kinase or Ataxia-telangiectasia mutated, symbol ATM, is a serine/threonine protein kinase that is recruited and activated by DNA double-strand breaks, oxidative stress, topoisomerase cleavage complexes, splicing intermediates, R-loops and in some cases by single-strand DNA breaks. It phosphorylates several key proteins that initiate activation of the DNA damage checkpoint, leading to cell cycle arrest, DNA repair or apoptosis. Several of these targets, including p53, CHK2, BRCA1, NBS1 and H2AX are tumor suppressors.

mTOR Mammalian protein found in humans

The mammalian target of rapamycin (mTOR), also referred to as the mechanistic target of rapamycin, and sometimes called FK506-binding protein 12-rapamycin-associated protein 1 (FRAP1), is a kinase that in humans is encoded by the MTOR gene. mTOR is a member of the phosphatidylinositol 3-kinase-related kinase family of protein kinases.

<span class="mw-page-title-main">TSC2</span> Mammalian protein found in Homo sapiens

Tuberous sclerosis complex 2 (TSC2), also known as tuberin, is a protein that in humans is encoded by the TSC2 gene.

<span class="mw-page-title-main">EIF4EBP1</span> Protein-coding gene in the species Homo sapiens

Eukaryotic translation initiation factor 4E-binding protein 1 is a protein that in humans is encoded by the EIF4EBP1 gene. inhibits cap-dependent translation by binding to translation initiation factor eIF4E. Phosphorylation of 4E-BP1 results in its release from eIF4E, thereby allows cap-dependent translation to continue thereby increasing the rate of protein synthesis.

<span class="mw-page-title-main">MAPKAPK2</span> Protein-coding gene in the species Homo sapiens

MAP kinase-activated protein kinase 2 is an enzyme that in humans is encoded by the MAPKAPK2 gene.

<span class="mw-page-title-main">RHEB</span> Protein-coding gene in the species Homo sapiens

RHEB also known as Ras homolog enriched in brain (RHEB) is a GTP-binding protein that is ubiquitously expressed in humans and other mammals. The protein is largely involved in the mTOR pathway and the regulation of the cell cycle.

<span class="mw-page-title-main">P70-S6 Kinase 1</span> Protein-coding gene in the species Homo sapiens

Ribosomal protein S6 kinase beta-1 (S6K1), also known as p70S6 kinase, is an enzyme that in humans is encoded by the RPS6KB1 gene. It is a serine/threonine kinase that acts downstream of PIP3 and phosphoinositide-dependent kinase-1 in the PI3 kinase pathway. As the name suggests, its target substrate is the S6 ribosomal protein. Phosphorylation of S6 induces protein synthesis at the ribosome.

<span class="mw-page-title-main">Protein kinase, AMP-activated, alpha 1</span> Protein-coding gene in the species Homo sapiens

5'-AMP-activated protein kinase catalytic subunit alpha-1 is an enzyme that in humans is encoded by the PRKAA1 gene.

<span class="mw-page-title-main">RPTOR</span> Protein-coding gene in humans

Regulatory-associated protein of mTOR also known as raptor or KIAA1303 is an adapter protein that is encoded in humans by the RPTOR gene. Two mRNAs from the gene have been identified that encode proteins of 1335 and 1177 amino acids long.

<span class="mw-page-title-main">RICTOR</span> Protein-coding gene in the species Homo sapiens

Rapamycin-insensitive companion of mammalian target of rapamycin (RICTOR) is a protein that in humans is encoded by the RICTOR gene.

<span class="mw-page-title-main">ULK1</span> Enzyme found in humans

ULK1 is an enzyme that in humans is encoded by the ULK1 gene.

The Akt signaling pathway or PI3K-Akt signaling pathway is a signal transduction pathway that promotes survival and growth in response to extracellular signals. Key proteins involved are PI3K and Akt.

Tuberous sclerosis proteins 1 and 2, also known as TSC1 (hamartin) and TSC2 (tuberin), form a protein-complex. The encoding two genes are TSC1 and TSC2. The complex is known as a tumor suppressor. Mutations in these genes can cause tuberous sclerosis complex. Depending on the grade of the disease, intellectual disability, epilepsy and tumors of the skin, retina, heart, kidney and the central nervous system can be symptoms.

AuTophaGy related 1 (Atg1) is a 101.7kDa serine/threonine kinase in S.cerevisiae, encoded by the gene ATG1. It is essential for the initial building of the autophagosome and Cvt vesicles. In a non-kinase role it is - through complex formation with Atg13 and Atg17 - directly controlled by the TOR kinase, a sensor for nutrient availability.

<span class="mw-page-title-main">MLST8</span> Protein-coding gene in humans

Target of rapamycin complex subunit LST8, also known as mammalian lethal with SEC13 protein 8 (mLST8) or TORC subunit LST8 or G protein beta subunit-like, is a protein that in humans is encoded by the MLST8 gene. It is a subunit of both mTORC1 and mTORC2, complexes that regulate cell growth and survival in response to nutrient, energy, redox, and hormonal signals. It is upregulated in several human colon and prostate cancer cell lines and tissues. Knockdown of mLST8 prevented mTORC formation and inhibited tumor growth and invasiveness.

mTOR inhibitors Class of pharmaceutical drugs

mTOR inhibitors are a class of drugs used to treat several human diseases, including cancer, autoimmune diseases, and neurodegeneration. They function by inhibiting the mammalian target of rapamycin (mTOR), which is a serine/threonine-specific protein kinase that belongs to the family of phosphatidylinositol-3 kinase (PI3K) related kinases (PIKKs). mTOR regulates cellular metabolism, growth, and proliferation by forming and signaling through two protein complexes, mTORC1 and mTORC2. The most established mTOR inhibitors are so-called rapalogs, which have shown tumor responses in clinical trials against various tumor types.

mTOR Complex 2 (mTORC2) is an acutely rapamycin-insensitive protein complex formed by serine/threonine kinase mTOR that regulates cell proliferation and survival, cell migration and cytoskeletal remodeling. The complex itself is rather large, consisting of seven protein subunits. The catalytic mTOR subunit, DEP domain containing mTOR-interacting protein (DEPTOR), mammalian lethal with sec-13 protein 8, and TTI1/TEL2 complex are shared by both mTORC2 and mTORC1. Rapamycin-insensitive companion of mTOR (RICTOR), mammalian stress-activated protein kinase interacting protein 1 (mSIN1), and protein observed with rictor 1 and 2 (Protor1/2) can only be found in mTORC2. Rictor has been shown to be the scaffold protein for substrate binding to mTORC2.

<span class="mw-page-title-main">HY-124798</span> Chemical compound

HY-124798 (Rheb inhibitor NR1) is the first compound to be developed that acts as a potent and selective inhibitor of Rheb, a GTP-binding protein which acts as an endogenous activator of the mechanistic target of rapamycin (mTOR) subtype mTORC1. Since many of the side effects of rapamycin and its analogues are thought to result from binding to the other subtype mTORC2, it is hoped that selective inhibition of mTORC1 should have a more selective effects profile. As mTORC1 and mTORC2 have binding sites that are very similar in structure, it has been challenging to develop highly subtype selective inhibitors, making indirect inhibition via modulation of other messenger proteins such as Rheb an attractive approach. However, since HY-124798 has a relatively weak IC50 of 2.1μM, and Rheb also has other targets in addition to mTORC1, it remains to be established whether it will deliver the hoped for improvements in pharmacological profile.

<span class="mw-page-title-main">Torin-1</span> Chemical compound

Torin-1 is a drug which was one of the first non-rapalog derived inhibitors of the mechanistic target of rapamycin (mTOR) subtypes mTORC1 and mTORC2. In animal studies it has anti-inflammatory, anti-cancer, and anti-aging properties, and shows activity against neuropathic pain.

References

  1. 1 2 3 Hay N, Sonenberg N (August 2004). "Upstream and downstream of mTOR". Genes & Development. 18 (16): 1926–1945. doi: 10.1101/gad.1212704 . PMID   15314020.
  2. 1 2 3 4 Kim DH, Sarbassov DD, Ali SM, et al. (July 2002). "mTOR interacts with raptor to form a nutrient-sensitive complex that signals to the cell growth machinery". Cell. 110 (2): 163–175. doi: 10.1016/S0092-8674(02)00808-5 . PMID   12150925. S2CID   4656930.
  3. Kim DH, Sarbassov DD, Ali SM, et al. (April 2003). "GbetaL, a positive regulator of the rapamycin-sensitive pathway required for the nutrient-sensitive interaction between raptor and mTOR". Molecular Cell. 11 (4): 895–904. doi: 10.1016/S1097-2765(03)00114-X . PMID   12718876.
  4. 1 2 Wullschleger S, Loewith R, Hall MN (February 2006). "TOR signaling in growth and metabolism". Cell. 124 (3): 471–484. doi: 10.1016/j.cell.2006.01.016 . PMID   16469695. S2CID   17195001.
  5. Fang Y, Vilella-Bach M, Bachmann R, et al. (November 2001). "Phosphatidic acid-mediated mitogenic activation of mTOR signaling". Science. 294 (5548): 1942–1945. Bibcode:2001Sci...294.1942F. doi:10.1126/science.1066015. PMID   11729323. S2CID   44444716.
  6. Bond P (March 2016). "Regulation of mTORC1 by growth factors, energy status, amino acids and mechanical stimuli at a glance". Journal of the International Society of Sports Nutrition. 13: 8. doi: 10.1186/s12970-016-0118-y . PMC   4774173 . PMID   26937223.
  7. Ali E, Liponska A, O'Hara B, et al. (June 2022). "The mTORC1-SLC4A7 axis stimulates bicarbonate import to enhance de novo nucleotide synthesis". Molecular Cell. 82 (1): 3284–3298.e7. doi:10.1016/j.molcel.2022.06.008. PMC   9444906 . PMID   35772404.
  8. Sharma A, Hoeffer CA, Takayasu Y, et al. (January 2010). "Dysregulation of mTOR signaling in fragile X syndrome". The Journal of Neuroscience. 30 (2): 694–702. doi:10.1523/JNEUROSCI.3696-09.2010. PMC   3665010 . PMID   20071534.
  9. Beauchamp EM, Platanias LC (August 2013). "The evolution of the TOR pathway and its role in cancer". Oncogene. 32 (34): 3923–3932. doi: 10.1038/onc.2012.567 . PMID   23246968.
  10. Durán RV, Hall MN (February 2012). "Regulation of TOR by small GTPases". EMBO Reports. 13 (2): 121–128. doi:10.1038/embor.2011.257. PMC   3271343 . PMID   22240970.
  11. Jewell JL, Russell RC, Guan KL (March 2013). "Amino acid signalling upstream of mTOR". Nature Reviews. Molecular Cell Biology. 14 (3): 133–139. doi:10.1038/nrm3522. PMC   3988467 . PMID   23361334.
  12. Efeyan A, Zoncu R, Sabatini DM (September 2012). "Amino acids and mTORC1: from lysosomes to disease". Trends in Molecular Medicine. 18 (9): 524–533. doi:10.1016/j.molmed.2012.05.007. hdl:1721.1/106904. PMC   3432651 . PMID   22749019.
  13. Efeyan A, Zoncu R, Sabatini DM (September 2012). "Amino acids and mTORC1: from lysosomes to disease". Trends in Molecular Medicine. 18 (9): 524–533. doi:10.1016/j.molmed.2012.05.007. PMC   3432651 . PMID   22749019.
  14. Sancak Y, Peterson TR, Shaul YD, et al. (June 2008). "The Rag GTPases bind raptor and mediate amino acid signaling to mTORC1". Science. 320 (5882): 1496–1501. Bibcode:2008Sci...320.1496S. doi:10.1126/science.1157535. PMC   2475333 . PMID   18497260.
  15. Saucedo LJ, Gao X, Chiarelli DA, et al. (June 2003). "Rheb promotes cell growth as a component of the insulin/TOR signalling network". Nature Cell Biology. 5 (6): 566–571. doi:10.1038/ncb996. PMID   12766776. S2CID   25954873.
  16. Suzuki T, Inoki K (September 2011). "Spatial regulation of the mTORC1 system in amino acids sensing pathway". Acta Biochimica et Biophysica Sinica. 43 (9): 671–679. doi:10.1093/abbs/gmr066. PMC   3160786 . PMID   21785113.
  17. Bar-Peled L, Chantranupong L, Cherniack AD, et al. (May 2013). "A Tumor suppressor complex with GAP activity for the Rag GTPases that signal amino acid sufficiency to mTORC1". Science. 340 (6136): 1100–1106. Bibcode:2013Sci...340.1100B. doi:10.1126/science.1232044. PMC   3728654 . PMID   23723238.
  18. 1 2 3 Ma XM, Blenis J (May 2009). "Molecular mechanisms of mTOR-mediated translational control". Nature Reviews. Molecular Cell Biology. 10 (5): 307–318. doi:10.1038/nrm2672. PMID   19339977. S2CID   30790160.
  19. 1 2 Mendoza MC, Er EE, Blenis J (June 2011). "The Ras-ERK and PI3K-mTOR pathways: cross-talk and compensation". Trends in Biochemical Sciences. 36 (6): 320–328. doi:10.1016/j.tibs.2011.03.006. PMC   3112285 . PMID   21531565.
  20. Oshiro N, Takahashi R, Yoshino K, et al. (July 2007). "The proline-rich Akt substrate of 40 kDa (PRAS40) is a physiological substrate of mammalian target of rapamycin complex 1". The Journal of Biological Chemistry. 282 (28): 20329–20339. doi: 10.1074/jbc.M702636200 . PMC   3199301 . PMID   17517883.
  21. Ye J (March 2013). "Mechanisms of insulin resistance in obesity". Frontiers of Medicine. 7 (1): 14–24. doi:10.1007/s11684-013-0262-6. PMC   3936017 . PMID   23471659.
  22. McCubrey JA, Steelman LS, Chappell WH, et al. (October 2012). "Ras/Raf/MEK/ERK and PI3K/PTEN/Akt/mTOR cascade inhibitors: how mutations can result in therapy resistance and how to overcome resistance". Oncotarget. 3 (10): 1068–1111. doi:10.18632/oncotarget.659. PMC   3717945 . PMID   23085539.
  23. Ma L, Chen Z, Erdjument-Bromage H, et al. (April 2005). "Phosphorylation and functional inactivation of TSC2 by Erk implications for tuberous sclerosis and cancer pathogenesis". Cell. 121 (2): 179–193. doi: 10.1016/j.cell.2005.02.031 . PMID   15851026. S2CID   18663447.
  24. Carrière A, Cargnello M, Julien LA, et al. (September 2008). "Oncogenic MAPK signaling stimulates mTORC1 activity by promoting RSK-mediated raptor phosphorylation". Current Biology. 18 (17): 1269–1277. Bibcode:2008CBio...18.1269C. doi: 10.1016/j.cub.2008.07.078 . PMID   18722121. S2CID   15088729.
  25. Kwak D, Choi S, Jeong H, et al. (May 2012). "Osmotic stress regulates mammalian target of rapamycin (mTOR) complex 1 via c-Jun N-terminal Kinase (JNK)-mediated Raptor protein phosphorylation". The Journal of Biological Chemistry. 287 (22): 18398–18407. doi: 10.1074/jbc.M111.326538 . PMC   3365776 . PMID   22493283.
  26. Fujishita T, Aoki M, Taketo MM (May 2011). "JNK signaling promotes intestinal tumorigenesis through activation of mTOR complex 1 in Apc(Δ716) mice". Gastroenterology. 140 (5): 1556–63.e6. doi: 10.1053/j.gastro.2011.02.007 . PMID   21320501.
  27. Monaghan D, O'Connell E, Cruickshank FL, et al. (January 2014). "Inhibition of protein synthesis and JNK activation are not required for cell death induced by anisomycin and anisomycin analogues". Biochemical and Biophysical Research Communications. 443 (2): 761–767. doi:10.1016/j.bbrc.2013.12.041. hdl: 20.500.11820/ba05d42b-8452-4391-8c4a-c2850cb28b12 . PMID   24333448.
  28. 1 2 Majid S, Saini S, Dahiya R (February 2012). "Wnt signaling pathways in urological cancers: past decades and still growing". Molecular Cancer. 11: 7. doi: 10.1186/1476-4598-11-7 . PMC   3293036 . PMID   22325146.
  29. Salminen A, Hyttinen JM, Kauppinen A, et al. (2012). "Context-Dependent Regulation of Autophagy by IKK-NF-κB Signaling: Impact on the Aging Process". International Journal of Cell Biology. 2012: 849541. doi: 10.1155/2012/849541 . PMC   3412117 . PMID   22899934.
  30. Hardie DG (October 2007). "AMP-activated/SNF1 protein kinases: conserved guardians of cellular energy". Nature Reviews. Molecular Cell Biology. 8 (10): 774–785. doi:10.1038/nrm2249. PMID   17712357. S2CID   38533515.
  31. Mihaylova MM, Shaw RJ (September 2011). "The AMPK signalling pathway coordinates cell growth, autophagy and metabolism". Nature Cell Biology. 13 (9): 1016–1023. doi:10.1038/ncb2329. PMC   3249400 . PMID   21892142.
  32. Gwinn DM, Shackelford DB, Egan DF, et al. (April 2008). "AMPK phosphorylation of raptor mediates a metabolic checkpoint". Molecular Cell. 30 (2): 214–226. doi:10.1016/j.molcel.2008.03.003. PMC   2674027 . PMID   18439900.
  33. Nagalingam A, Arbiser JL, Bonner MY, et al. (February 2012). "Honokiol activates AMP-activated protein kinase in breast cancer cells via an LKB1-dependent pathway and inhibits breast carcinogenesis". Breast Cancer Research. 14 (1): R35. doi: 10.1186/bcr3128 . PMC   3496153 . PMID   22353783.
  34. Horak P, Crawford AR, Vadysirisack DD, et al. (March 2010). "Negative feedback control of HIF-1 through REDD1-regulated ROS suppresses tumorigenesis". Proceedings of the National Academy of Sciences of the United States of America. 107 (10): 4675–4680. Bibcode:2010PNAS..107.4675H. doi: 10.1073/pnas.0907705107 . PMC   2842042 . PMID   20176937.
  35. Brugarolas J, Lei K, Hurley RL, et al. (December 2004). "Regulation of mTOR function in response to hypoxia by REDD1 and the TSC1/TSC2 tumor suppressor complex". Genes & Development. 18 (23): 2893–2904. doi:10.1101/gad.1256804. PMC   534650 . PMID   15545625.
  36. Wang S, Song P, Zou MH (June 2012). "AMP-activated protein kinase, stress responses and cardiovascular diseases". Clinical Science. 122 (12): 555–573. doi:10.1042/CS20110625. PMC   3367961 . PMID   22390198.
  37. Martelli AM, Evangelisti C, Chappell W, et al. (July 2011). "Targeting the translational apparatus to improve leukemia therapy: roles of the PI3K/PTEN/Akt/mTOR pathway". Leukemia. 25 (7): 1064–1079. doi: 10.1038/leu.2011.46 . PMID   21436840.
  38. Wang H, Zhang Q, Wen Q, et al. (January 2012). "Proline-rich Akt substrate of 40kDa (PRAS40): a novel downstream target of PI3k/Akt signaling pathway". Cellular Signalling. 24 (1): 17–24. doi:10.1016/j.cellsig.2011.08.010. PMID   21906675.
  39. Raught B, Gingras AC (January 1999). "eIF4E activity is regulated at multiple levels". The International Journal of Biochemistry & Cell Biology. 31 (1): 43–57. doi:10.1016/s1357-2725(98)00131-9. PMID   10216943.
  40. Babendure JR, Babendure JL, Ding JH, et al. (May 2006). "Control of mammalian translation by mRNA structure near caps". RNA. 12 (5): 851–861. doi:10.1261/rna.2309906. PMC   1440912 . PMID   16540693.
  41. Lee T, Pelletier J (January 2012). "Eukaryotic initiation factor 4F: a vulnerability of tumor cells". Future Medicinal Chemistry. 4 (1): 19–31. doi:10.4155/fmc.11.150. PMID   22168162.
  42. Holz MK, Ballif BA, Gygi SP, et al. (November 2005). "mTOR and S6K1 mediate assembly of the translation preinitiation complex through dynamic protein interchange and ordered phosphorylation events". Cell. 123 (4): 569–580. doi: 10.1016/j.cell.2005.10.024 . PMID   16286006. S2CID   11118504.
  43. Saitoh M, Pullen N, Brennan P, et al. (May 2002). "Regulation of an activated S6 kinase 1 variant reveals a novel mammalian target of rapamycin phosphorylation site". The Journal of Biological Chemistry. 277 (22): 20104–20112. doi: 10.1074/jbc.M201745200 . PMID   11914378.
  44. 1 2 Pullen N, Thomas G (June 1997). "The modular phosphorylation and activation of p70s6k". FEBS Letters. 410 (1): 78–82. doi: 10.1016/S0014-5793(97)00323-2 . PMID   9247127. S2CID   36947968.
  45. Pullen N, Dennis PB, Andjelkovic M, et al. (January 1998). "Phosphorylation and activation of p70s6k by PDK1". Science. 279 (5351): 707–710. Bibcode:1998Sci...279..707P. doi:10.1126/science.279.5351.707. PMID   9445476.
  46. Peterson RT, Schreiber SL (March 1998). "Translation control: connecting mitogens and the ribosome". Current Biology. 8 (7): R248–R250. Bibcode:1998CBio....8.R248P. doi: 10.1016/S0960-9822(98)70152-6 . PMID   9545190. S2CID   2528173.
  47. Ma XM, Yoon SO, Richardson CJ, et al. (April 2008). "SKAR links pre-mRNA splicing to mTOR/S6K1-mediated enhanced translation efficiency of spliced mRNAs". Cell. 133 (2): 303–313. doi: 10.1016/j.cell.2008.02.031 . PMID   18423201. S2CID   13437701.
  48. Chiang GG, Abraham RT (July 2005). "Phosphorylation of mammalian target of rapamycin (mTOR) at Ser-2448 is mediated by p70S6 kinase". The Journal of Biological Chemistry. 280 (27): 25485–25490. doi: 10.1074/jbc.M501707200 . PMID   15899889.
  49. Holz MK, Blenis J (July 2005). "Identification of S6 kinase 1 as a novel mammalian target of rapamycin (mTOR)-phosphorylating kinase". The Journal of Biological Chemistry. 280 (28): 26089–26093. doi: 10.1074/jbc.M504045200 . PMID   15905173.
  50. Fabrizio P, Pozza F, Pletcher SD, et al. (April 2001). "Regulation of longevity and stress resistance by Sch9 in yeast". Science. 292 (5515): 288–290. Bibcode:2001Sci...292..288F. doi:10.1126/science.1059497. PMID   11292860. S2CID   44756177.
  51. Robida-Stubbs S, Glover-Cutter K, Lamming DW, et al. (May 2012). "TOR signaling and rapamycin influence longevity by regulating SKN-1/Nrf and DAF-16/FoxO". Cell Metabolism. 15 (5): 713–724. doi:10.1016/j.cmet.2012.04.007. PMC   3348514 . PMID   22560223.
  52. Harrison DE, Strong R, Sharp ZD, et al. (July 2009). "Rapamycin fed late in life extends lifespan in genetically heterogeneous mice". Nature. 460 (7253): 392–395. Bibcode:2009Natur.460..392H. doi:10.1038/nature08221. PMC   2786175 . PMID   19587680.
  53. Lynn MA, Eden G, Ryan FJ, et al. (August 2021). "The composition of the gut microbiota following early-life antibiotic exposure affects host health and longevity in later life". Cell Reports. 36 (8): 109564. doi: 10.1016/j.celrep.2021.109564 . PMID   34433065. S2CID   237306510.
  54. Kaeberlein M, Powers RW, Steffen KK, et al. (November 2005). "Regulation of yeast replicative life span by TOR and Sch9 in response to nutrients". Science. 310 (5751): 1193–1196. Bibcode:2005Sci...310.1193K. doi:10.1126/science.1115535. PMID   16293764. S2CID   42188272.
  55. Blagosklonny MV (February 2010). "Calorie restriction: decelerating mTOR-driven aging from cells to organisms (including humans)". Cell Cycle. 9 (4): 683–688. doi: 10.4161/cc.9.4.10766 . PMID   20139716.
  56. 1 2 Choi AM, Ryter SW, Levine B (February 2013). "Autophagy in human health and disease". The New England Journal of Medicine. 368 (7): 651–662. doi:10.1056/NEJMra1205406. PMID   23406030.
  57. Murrow L, Debnath J (January 2013). "Autophagy as a stress-response and quality-control mechanism: implications for cell injury and human disease". Annual Review of Pathology. 8: 105–137. doi:10.1146/annurev-pathol-020712-163918. PMC   3971121 . PMID   23072311.
  58. Alers S, Löffler AS, Wesselborg S, et al. (January 2012). "Role of AMPK-mTOR-Ulk1/2 in the regulation of autophagy: cross talk, shortcuts, and feedbacks". Molecular and Cellular Biology. 32 (1): 2–11. doi:10.1128/MCB.06159-11. PMC   3255710 . PMID   22025673.
  59. Pyo JO, Nah J, Jung YK (February 2012). "Molecules and their functions in autophagy". Experimental & Molecular Medicine. 44 (2): 73–80. doi:10.3858/emm.2012.44.2.029. PMC   3296815 . PMID   22257882.
  60. Proud CG (November 2007). "Amino acids and mTOR signalling in anabolic function". Biochemical Society Transactions. 35 (Pt 5): 1187–1190. doi:10.1042/BST0351187. PMID   17956308. S2CID   13379878.
  61. Cuervo AM, Dice JF (October 2000). "Age-related decline in chaperone-mediated autophagy". The Journal of Biological Chemistry. 275 (40): 31505–31513. doi: 10.1074/jbc.M002102200 . PMID   10806201.
  62. Codogno P, Meijer AJ (November 2005). "Autophagy and signaling: their role in cell survival and cell death". Cell Death and Differentiation. 12 (Suppl 2): 1509–1518. doi: 10.1038/sj.cdd.4401751 . PMID   16247498.
  63. 1 2 3 4 Jia J, Abudu YP, Claude-Taupin A, et al. (April 2018). "Galectins Control mTOR in Response to Endomembrane Damage". Molecular Cell. 70 (1): 120–135.e8. doi:10.1016/j.molcel.2018.03.009. PMC   5911935 . PMID   29625033.
  64. Hasegawa J, Maejima I, Iwamoto R, et al. (March 2015). "Selective autophagy: lysophagy". Methods. 75: 128–132. doi: 10.1016/j.ymeth.2014.12.014 . PMID   25542097.
  65. Apel K, Hirt H (2004). "Reactive oxygen species: metabolism, oxidative stress, and signal transduction". Annual Review of Plant Biology. 55: 373–399. doi:10.1146/annurev.arplant.55.031903.141701. PMID   15377225. S2CID   17229119.
  66. Murphy MP (January 2009). "How mitochondria produce reactive oxygen species". The Biochemical Journal. 417 (1): 1–13. doi:10.1042/BJ20081386. PMC   2605959 . PMID   19061483.
  67. Bonawitz ND, Chatenay-Lapointe M, Pan Y, et al. (April 2007). "Reduced TOR signaling extends chronological life span via increased respiration and upregulation of mitochondrial gene expression". Cell Metabolism. 5 (4): 265–277. doi:10.1016/j.cmet.2007.02.009. PMC   3460550 . PMID   17403371.
  68. Adam-Vizi V (2005). "Production of reactive oxygen species in brain mitochondria: contribution by electron transport chain and non-electron transport chain sources". Antioxidants & Redox Signaling. 7 (9–10): 1140–1149. doi:10.1089/ars.2005.7.1140. PMID   16115017.
  69. Sun Q, Chen X, Ma J, et al. (March 2011). "Mammalian target of rapamycin up-regulation of pyruvate kinase isoenzyme type M2 is critical for aerobic glycolysis and tumor growth". Proceedings of the National Academy of Sciences of the United States of America. 108 (10): 4129–4134. Bibcode:2011PNAS..108.4129S. doi: 10.1073/pnas.1014769108 . PMC   3054028 . PMID   21325052.
  70. Sporn MB, Liby KT (July 2012). "NRF2 and cancer: the good, the bad and the importance of context". Nature Reviews. Cancer. 12 (8): 564–571. doi:10.1038/nrc3278. PMC   3836441 . PMID   22810811.
  71. Li C, Reif MM, Craige SM, et al. (May 2016). "Endothelial AMPK activation induces mitochondrial biogenesis and stress adaptation via eNOS-dependent mTORC1 signaling". Nitric Oxide. 55: 45–53. doi:10.1016/j.niox.2016.03.003. PMC   4860108 . PMID   26989010.
  72. Ho AD, Wagner W, Mahlknecht U (July 2005). "Stem cells and ageing. The potential of stem cells to overcome age-related deteriorations of the body in regenerative medicine". EMBO Reports. 6 (Suppl 1): S35–S38. doi:10.1038/sj.embor.7400436. PMC   1369281 . PMID   15995659.
  73. 1 2 Murakami M, Ichisaka T, Maeda M, et al. (August 2004). "mTOR is essential for growth and proliferation in early mouse embryos and embryonic stem cells". Molecular and Cellular Biology. 24 (15): 6710–6718. doi:10.1128/MCB.24.15.6710-6718.2004. PMC   444840 . PMID   15254238.
  74. Gangloff YG, Mueller M, Dann SG, et al. (November 2004). "Disruption of the mouse mTOR gene leads to early postimplantation lethality and prohibits embryonic stem cell development". Molecular and Cellular Biology. 24 (21): 9508–9516. doi:10.1128/MCB.24.21.9508-9516.2004. PMC   522282 . PMID   15485918.
  75. 1 2 Chen C, Liu Y, Liu Y, et al. (November 2009). "mTOR regulation and therapeutic rejuvenation of aging hematopoietic stem cells". Science Signaling. 2 (98): ra75. doi:10.1126/scisignal.2000559. PMC   4020596 . PMID   19934433.
  76. Russell RC, Fang C, Guan KL (August 2011). "An emerging role for TOR signaling in mammalian tissue and stem cell physiology". Development. 138 (16): 3343–3356. doi:10.1242/dev.058230. PMC   3143559 . PMID   21791526.
  77. Limon JJ, Fruman DA (2012). "Akt and mTOR in B Cell Activation and Differentiation". Frontiers in Immunology. 3: 228. doi: 10.3389/fimmu.2012.00228 . PMC   3412259 . PMID   22888331.
  78. Araki K, Turner AP, Shaffer VO, et al. (July 2009). "mTOR regulates memory CD8 T-cell differentiation". Nature. 460 (7251): 108–112. Bibcode:2009Natur.460..108A. doi:10.1038/nature08155. PMC   2710807 . PMID   19543266.
  79. 1 2 Araki K, Youngblood B, Ahmed R (May 2010). "The role of mTOR in memory CD8 T-cell differentiation". Immunological Reviews. 235 (1): 234–243. doi:10.1111/j.0105-2896.2010.00898.x. PMC   3760155 . PMID   20536567.
  80. 1 2 3 Duman RS (2018). "Ketamine and rapid-acting antidepressants: a new era in the battle against depression and suicide". F1000Research. 7: 659. doi: 10.12688/f1000research.14344.1 . PMC   5968361 . PMID   29899972.
  81. Liu M, Wilk SA, Wang A, et al. (November 2010). "Resveratrol inhibits mTOR signaling by promoting the interaction between mTOR and DEPTOR". The Journal of Biological Chemistry. 285 (47): 36387–36394. doi: 10.1074/jbc.M110.169284 . PMC   2978567 . PMID   20851890.
  82. Miwa S, Sugimoto N, Yamamoto N, et al. (September 2012). "Caffeine induces apoptosis of osteosarcoma cells by inhibiting AKT/mTOR/S6K, NF-κB and MAPK pathways". Anticancer Research. 32 (9): 3643–3649. PMID   22993301.
  83. Vézina C, Kudelski A, Sehgal SN (October 1975). "Rapamycin (AY-22,989), a new antifungal antibiotic. I. Taxonomy of the producing streptomycete and isolation of the active principle". The Journal of Antibiotics. 28 (10): 721–726. doi: 10.7164/antibiotics.28.721 . PMID   1102508.
  84. Tsang CK, Qi H, Liu LF, et al. (February 2007). "Targeting mammalian target of rapamycin (mTOR) for health and diseases". Drug Discovery Today. 12 (3–4): 112–124. doi:10.1016/j.drudis.2006.12.008. PMID   17275731.
  85. Sarbassov DD, Ali SM, Sengupta S, et al. (April 2006). "Prolonged rapamycin treatment inhibits mTORC2 assembly and Akt/PKB". Molecular Cell. 22 (2): 159–168. doi: 10.1016/j.molcel.2006.03.029 . PMID   16603397.
  86. Lamming DW, Ye L, Katajisto P, et al. (March 2012). "Rapamycin-induced insulin resistance is mediated by mTORC2 loss and uncoupled from longevity". Science. 335 (6076): 1638–1643. Bibcode:2012Sci...335.1638L. doi:10.1126/science.1215135. PMC   3324089 . PMID   22461615.
  87. Schreiber KH, Ortiz D, Academia EC, et al. (April 2015). "Rapamycin-mediated mTORC2 inhibition is determined by the relative expression of FK506-binding proteins". Aging Cell. 14 (2): 265–273. doi:10.1111/acel.12313. PMC   4364838 . PMID   25652038.
  88. 1 2 3 Vilar E, Perez-Garcia J, Tabernero J (March 2011). "Pushing the envelope in the mTOR pathway: the second generation of inhibitors". Molecular Cancer Therapeutics. 10 (3): 395–403. doi:10.1158/1535-7163.MCT-10-0905. PMC   3413411 . PMID   21216931.
  89. De P, Miskimins K, Dey N, et al. (August 2013). "Promise of rapalogues versus mTOR kinase inhibitors in subset specific breast cancer: old targets new hope". Cancer Treatment Reviews. 39 (5): 403–412. doi:10.1016/j.ctrv.2012.12.002. PMID   23352077.
  90. Arriola Apelo SI, Neuman JC, Baar EL, et al. (February 2016). "Alternative rapamycin treatment regimens mitigate the impact of rapamycin on glucose homeostasis and the immune system". Aging Cell. 15 (1): 28–38. doi:10.1111/acel.12405. PMC   4717280 . PMID   26463117.
  91. Wang S, Raybuck A, Shiuan E, et al. (August 2020). "Selective inhibition of mTORC1 in tumor vessels increases antitumor immunity". JCI Insight. 5 (15): e139237. doi:10.1172/jci.insight.139237. PMC   7455083 . PMID   32759497.
  92. Nashan B, Citterio F (September 2012). "Wound healing complications and the use of mammalian target of rapamycin inhibitors in kidney transplantation: a critical review of the literature". Transplantation. 94 (6): 547–561. doi: 10.1097/TP.0b013e3182551021 . PMID   22941182. S2CID   24753934.
  93. Townsend JC, Rideout P, Steinberg DH (2012). "Everolimus-eluting stents in interventional cardiology". Vascular Health and Risk Management. 8: 393–404. doi: 10.2147/VHRM.S23388 . PMC   3402052 . PMID   22910420.
  94. Voss MH, Molina AM, Motzer RJ (August 2011). "mTOR inhibitors in advanced renal cell carcinoma". Hematology/Oncology Clinics of North America. 25 (4): 835–852. doi:10.1016/j.hoc.2011.04.008. PMC   3587783 . PMID   21763970.
  95. Smith SM (June 2012). "Targeting mTOR in mantle cell lymphoma: current and future directions". Best Practice & Research. Clinical Haematology. 25 (2): 175–183. doi:10.1016/j.beha.2012.04.008. PMID   22687453.
  96. Fasolo A, Sessa C (2012). "Targeting mTOR pathways in human malignancies". Current Pharmaceutical Design. 18 (19): 2766–2777. doi:10.2174/138161212800626210. PMID   22475451.
  97. Budde K, Gaedeke J (February 2012). "Tuberous sclerosis complex-associated angiomyolipomas: focus on mTOR inhibition". American Journal of Kidney Diseases. 59 (2): 276–283. doi:10.1053/j.ajkd.2011.10.013. PMID   22130643. S2CID   18525093.
  98. 1 2 Zhang YJ, Duan Y, Zheng XF (April 2011). "Targeting the mTOR kinase domain: the second generation of mTOR inhibitors". Drug Discovery Today. 16 (7–8): 325–331. doi:10.1016/j.drudis.2011.02.008. PMC   3073023 . PMID   21333749.
  99. Veilleux A, Houde VP, Bellmann K, et al. (April 2010). "Chronic inhibition of the mTORC1/S6K1 pathway increases insulin-induced PI3K activity but inhibits Akt2 and glucose transport stimulation in 3T3-L1 adipocytes". Molecular Endocrinology. 24 (4): 766–778. doi:10.1210/me.2009-0328. PMC   5417537 . PMID   20203102.
  100. Śledź KM, Moore SF, Durrant TN, et al. (July 2020). "Rapamycin restrains platelet procoagulant responses via FKBP-mediated protection of mitochondrial integrity". Biochemical Pharmacology. 177: 113975. doi:10.1016/j.bcp.2020.113975. PMID   32298692. S2CID   215803320.
  101. Schenone S, Brullo C, Musumeci F, et al. (2011). "ATP-competitive inhibitors of mTOR: an update". Current Medicinal Chemistry. 18 (20): 2995–3014. doi:10.2174/092986711796391651. PMID   21651476.
  102. Zask A, Verheijen JC, Richard DJ (July 2011). "Recent advances in the discovery of small-molecule ATP competitive mTOR inhibitors: a patent review". Expert Opinion on Therapeutic Patents. 21 (7): 1109–27. doi:10.1517/13543776.2011.584871. PMID   21591993. S2CID   207474033.
  103. Lv X, Ma X, Hu Y (August 2013). "Furthering the design and the discovery of small molecule ATP-competitive mTOR inhibitors as an effective cancer treatment". Expert Opinion on Drug Discovery. 8 (8): 991–1012. doi:10.1517/17460441.2013.800479. PMID   23668243. S2CID   22677288.
  104. Lamming DW, Ye L, Katajisto P, et al. (March 2012). "Rapamycin-induced insulin resistance is mediated by mTORC2 loss and uncoupled from longevity". Science. 335 (6076): 1638–1643. Bibcode:2012Sci...335.1638L. doi:10.1126/science.1215135. PMC   3324089 . PMID   22461615.
  105. Zhou H, Huang S (2016). "Role of mTOR signaling in tumor cell motility, invasion and metastasis". In Atta-ur-Rahman (ed.). Advances in Cancer Drug Targets. Vol. 3. pp. 207–44. doi:10.2174/9781681082332116030009. ISBN   978-1-68108-233-2.
  106. Schreiber KH, Arriola Apelo SI, Yu D, et al. (July 2019). "A novel rapamycin analog is highly selective for mTORC1 in vivo". Nature Communications. 10 (1): 3194. Bibcode:2019NatCo..10.3194S. doi:10.1038/s41467-019-11174-0. PMC   6642166 . PMID   31324799.
  107. Yang H, Jiang X, Li B, et al. (December 2017). "Mechanisms of mTORC1 activation by RHEB and inhibition by PRAS40". Nature. 552 (7685): 368–373. Bibcode:2017Natur.552..368Y. doi:10.1038/nature25023. PMC   5750076 . PMID   29236692.
  108. Mahoney SJ, Narayan S, Molz L, et al. (February 2018). "A small molecule inhibitor of Rheb selectively targets mTORC1 signaling". Nature Communications. 9 (1): 548. Bibcode:2018NatCo...9..548M. doi:10.1038/s41467-018-03035-z. PMC   5803267 . PMID   29416044.
  109. Kang SA, O'Neill DJ, Machl AW, et al. (September 2019). "Discovery of Small-Molecule Selective mTORC1 Inhibitors via Direct Inhibition of Glucose Transporters". Cell Chemical Biology. 26 (9): 1203–1213.e13. doi: 10.1016/j.chembiol.2019.05.009 . PMID   31231029.
  110. Johnson SC, Rabinovitch PS, Kaeberlein M (January 2013). "mTOR is a key modulator of ageing and age-related disease". Nature. 493 (7432): 338–345. Bibcode:2013Natur.493..338J. doi:10.1038/nature11861. PMC   3687363 . PMID   23325216.