Jump to content

mTORC1

From Wikipedia, the free encyclopedia

This is an old revision of this page, as edited by Jnims (talk | contribs) at 01:10, 23 April 2013 (Energy status). The present address (URL) is a permanent link to this revision, which may differ significantly from the current revision.

mTOR
Identifiers
SymbolMTOR
Alt. symbolsFRAP, FRAP2, FRAP1
NCBI gene2475
HGNC3942
OMIM601231
RefSeqNM_004958
UniProtP42345
Other data
EC number2.7.11.1
LocusChr. 1 p36
Search for
StructuresSwiss-model
DomainsInterPro
Raptor
Identifiers
SymbolRPTOR
NCBI gene57521
HGNC30287
OMIM607130
RefSeqNM_020761
UniProtQ8N122
Other data
LocusChr. 17 q25.3
Search for
StructuresSwiss-model
DomainsInterPro
MLST8
Identifiers
SymbolMLST8
NCBI gene64223
HGNC24825
OMIM612190
RefSeqNM_022372
UniProtQ9BVC4
Other data
LocusChr. 16 p13.3
Search for
StructuresSwiss-model
DomainsInterPro
PRAS
Identifiers
SymbolAKT1S1
NCBI gene84335
HGNC28426
OMIM610221
RefSeqNM_032375
UniProtQ96B36
Other data
LocusChr. 19 q13.33
Search for
StructuresSwiss-model
DomainsInterPro
DEPTOR
Identifiers
SymbolDEPTOR
Alt. symbolsDEPDC6
NCBI gene64798
HGNC22953
OMIM612974
RefSeqNM_022783
UniProtQ8TB45
Other data
LocusChr. 8 q24.12
Search for
StructuresSwiss-model
DomainsInterPro

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

mTOR Complex 1 (mTORC1) is composed of mTOR, regulatory-associated protein of mTOR (Raptor), mammalian lethal with SEC13 protein 8 (MLST8) and the recently identified partners PRAS40 and DEPTOR.[2][3] This complex is characterized by the classic features of mTOR by functioning as a nutrient/energy/redox sensor and controlling protein synthesis.[1][2] The activity of this complex is stimulated by insulin, growth factors, serum, phosphatidic acid, amino acids (particularly leucine), and oxidative stress.[2][4]

Function

The role of mTORC1 is to activate translation of proteins. 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 protein translation to begin.[5]

All of these variables for protein synthesis affect mTORC1 activation by interacting with the TSC1/TSC2 protein complex. TSC2 is a GTP-ase activating protein (GAP). Its GAP activity interacts with the 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.[6] 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, which can cause the dimer to dissociate and lose its GAP activity, or the phosphorylation can cause the heterodimer to have more active GAP activity, depending on the kinase phosphorylating the dimer.[7] Thus, the signals that influence mTORC1 activity do so through activation or inactivation of the TSC1/TSC2 complex, upstream of mTORC1.

Upstream Signaling

Amino acids

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. Consequentially, mTORC1 signaling is sensitive to amino acid levels in the cell. 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. Active Rag heterodimers interact with Raptor, localizing mTORC1 to the surface of late endosomes and lysosomes where the Rheb-GTP is located.[8] This allows mTORC1 to physically interact with Rheb and thus endosomes and lysosomes are where Rheb will activate mTORC1.[9][10]

Additionally, since Rheb needs to be in its active GTP-bound state, nutrient and growth factor signals are also required for this type of mTORC1 activation. Specifically, they are integrated when the Rag proteins deliver mTORC1 to Rheb in the endosome in order to ensure that Rheb is in its active GTP-bound state.[11]

Growth factors

Insulin

Growth factors like insulin 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.[12] 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.[13]

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 S6K-1, its removal will allow the two substrates to be recruited to mTORC1 and thereby activated in this way.[14]

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, S6K-1 is able to phosphorylate the insulin receptor and inhibit its sensitivity to insulin.[12] This has great significance in diabetes mellitus, which is due to insulin resistance.[15]

Mitogens

Mitogens, like insulin like growth factor 1 (IGF1), can activate the MAPK/ERK pathway, which can control the TSC1/TSC2 complex as well as directly have the same downstream role of mTORC1.[13] 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).[16] 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.[17] 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.[18]

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).[19] When the Wnt pathway is not active, GSK3 beta is able to phosphorylate TSC2 on two serine residues of 1341 and 1337 in conjunction with AMPK phosphorylating serine residue 1345. It has been studied that the AMPK is required to first phosphorylate residue 1345 before GSK3 beta can phosphorylate its target serine residues. This phosphorylation of TSC2 would inactivate this complex, if GSK3 beta 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.[19]

Cytokines

Cytokines like tumor necrosis factor alpha (TFNalpha) can induce mTOR activity through IKK beta, also known as IKK2.[20] 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.[21]

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.[22]

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.[23]

LBK1, 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.[24]

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.[25][26]

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.[27]

Downstream Signaling

mTOC1 activates transcription and translation through its interactions with 4E-BP1 and S6K.[1]

4E-BP1

Activated mTORC1 will phosphorylate transcription inhibitor 4E-BP1, releasing it from eukaryotic translation initiation factor 4E (eIF4E).[28] eIF4E is now free to join the eukaryotic translation initiation factor 4G (eIF4G) and the eukaryotic translation initiation factor 4A (Eif4a).[29] 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).[30] The helicase is required to remove hairpin loops that arise in the 5' untranslated regions of mRNA, which prevent premature translation of proteins. 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 eradicated by the eIF4A helicase.[31]

S6K

mTORC1 will also activate phosphorylate the serine residue 422 on S6K, which is responsible for the recruitment of eIF4B to the initiation complex.[32]

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.[33]

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.[34]

Hypophosphorylated S6K is located on the eIR3 scaffold complex. Active mTORC1 gets recruited to the scaffold, and once there, will phosphorylate S6K to make it active.[12]

The two best characterized targets of mTORC1 are p70-S6 Kinase 1 (S6K1) and 4E-BP1, the eukaryotic initiation factor 4E (eIF4E) binding protein 1.[1] mTORC1 phosphorylates S6K1 on at least two residues, with the most critical modification occurring on a threonine residue (T389).[35][36] This event stimulates the subsequent phosphorylation of S6K1 by PDK1.[36][37] Active S6K1 can in turn stimulate the initiation of protein synthesis through activation of S6 Ribosomal protein (a component of the ribosome) and other components of the translational machinery.[38] S6K1 can also participate in a positive feedback loop with mTORC1 by phosphorylating mTOR's negative regulatory domain at two sites; phosphorylation at these sites appears to stimulate mTOR activity.[39][40]

mTORC1 has been shown to phosphorylate at least four residues of 4E-BP1 in a hierarchical manner.[41][42][43] Non-phosphorylated 4E-BP1 binds tightly to the translation initiation factor eIF4E, preventing it from binding to 5'-capped mRNAs and recruiting them to the ribosomal initiation complex.[44] Upon phosphorylation by mTORC1, 4E-BP1 releases eIF4E, allowing it to perform its function.[44] The activity of mTORC1 appears to be regulated through a dynamic interaction between mTOR and Raptor, one that is mediated by GβL.[2][3] Raptor and mTOR share a strong N-terminal interaction and a weaker C-terminal interaction near mTOR's kinase domain.[2] When stimulatory signals are sensed, such as high nutrient/energy levels, the mTOR-Raptor C-terminal interaction is weakened and possibly completely lost, allowing mTOR kinase activity to be turned on. When stimulatory signals are withdrawn, such as low nutrient levels, the mTOR-Raptor C-terminal interaction is strengthened, essentially shutting off kinase function of mTOR.[2]

Role in Human Diseases 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.[45] As a result, mTORC1 signaling was focused on and techniques used to inhibit its activity in C. elegans, fruitflies, and mice significantly increased their lifespans relative to the control organisms for the respective species.[46] [47]

Based on upstream signaling of mTORC1, a clear relationship between food consumption and mTORC1 activity has been observed.[48] 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.[49]

Dietary restriction has been shown to significantly increase lifespan in the human model of Rhesus monkeys as well as protect against their age related decline.[50] More specifically, Rhesus monkeys on a calorie restricted diet had significantly less chance of developing cardiovascular disease, diabetes, cancer, and age related cognitive decline than those monkeys who were not placed on the calorie restricted diet.[50]

Stem Cells

Conservation of stem cells in the body has been shown to help prevent against premature aging.[51] mTORC1 activity plays a critical role in the growth and proliferation of stem cells.[52] Knocking out mTORC1 results in embryonic lethality due to lack of trophoblast development.[53] Treating stem cells with rapamycin will also slow their proliferation, conserving the stem cells in their undifferentiated condition. [52]

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.[54] Rapamycin is used clinically as an immunosupressant and prevents the proliferation of T cells and B cells.[55] 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 .[56] 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.[57] mTORC1 inhibition with rapamycin has also been linked to a dramatic increase of B cells in old mice, enhancing their immune systems.[54] This paradox of rapamycin inhibiting the immune system response has been linked to several reasons, including its interaction with regulatory T cells.[57] The mechanisms mTORC1's inhibition on proliferation and differentiation of hematopoietic stem cells has yet to be fully elucidated.[58]

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.[59] 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.[59] Autophagy can thus remove protein aggregates and damaged organelles, that can lead to cellular dysfunction.[60]

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.[61] This prevents the structure from being recruited to the preautophagosomal structure at the plasma membrane, inhibiting autophagy.[62].

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. [63] Because autophagy appears to decline with age, activation of autophagy may help promote longevity in humans. [64] Problems in proper autophagy processes have been linked to diabetes, cardiovascular disease, neurodegenerative diseases, and cancer.[65]

Reactive Oxygen Species

Reactive oxygen species can damage the DNA and proteins in cells.[66] A majority of them arise in the mitochondria.[67]

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.[68] 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.[69] 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.[70]

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.[71]

Drug Inhibition

There have been several dietary compounds that have been suggested to inhibit mTORC1 signaling including EGCG, resveratrol, curcumin, caffeine, and alcohol.[72] [73]

First Generation Inhibitors

Rapamycin was the first known inhibitor of mTORC1, considering that mTORC1 was discovered as being the target of rapamycin.[74] Rapamycin will bind to cytosolic FKBP12 and act as a scaffold molecule, allowing this protein to dock on the FBP regulatory region on mTORC1.[75] The binding of the FKBP12-rapamycin complex to the FBP regulatory region inhibits mTORC1 through processes not yet known.

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.[76] These drugs are considered the first generation inhibitors of mTOR.[77]

Sirolimus, which is the drug name for rapamycin, was approved by the FDA in 1999 to prevent against transplant rejection in patients undergoing kidney transplantation.[78] In 2003, it was approved as a stent covering for people who want to widen their arteries to prevent against future heart attacks.[79] In 2007, they began being approved for treatments against cancers such as renal cell carcinoma.[80] In 2008 they were approved for treatment of mantle cell lymphoma.[81] mTORC1 inhibitors have recently been approved for treatment of pancreatic cancer.[82] In 2010 they were approved for treatment of tuberous sclerosis.[83]

Second Generation Inhibitors

The second generation of inhibitors were created to overcome problems with upstream signaling upon the introduction of first generation inhibitors to the treated cells.[84] 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.[85] 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.[76] Thus signaling upstream of mTORC1 still remains very active upon its inhibition via rapamycin and the rapalogs.

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.[84] 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.[76] As of 2011, these second generation inhibitors were in phase II of clinical trials.

There are currently more than 1,300 clinical trials underway for the mTOR complex inhibitors.[86]

References

  1. ^ a b c d Hay N, Sonenberg N (2004). "Upstream and downstream of mTOR". Genes Dev. 18 (16): 1926–45. doi:10.1101/gad.1212704. PMID 15314020. {{cite journal}}: Unknown parameter |month= ignored (help)
  2. ^ a b c d e f g Kim DH, Sarbassov DD, Ali SM, King JE, Latek RR, Erdjument-Bromage H, Tempst P, Sabatini DM (2002). "mTOR interacts with raptor to form a nutrient-sensitive complex that signals to the cell growth machinery". Cell. 110 (2): 163–75. doi:10.1016/S0092-8674(02)00808-5. PMID 12150925. {{cite journal}}: Unknown parameter |month= ignored (help)CS1 maint: multiple names: authors list (link)
  3. ^ a b Kim D, Sarbassov D, Ali S, Latek R, Guntur K, Erdjument-Bromage H, Tempst P, Sabatini D (2003). "GbetaL, a positive regulator of the rapamycin-sensitive pathway required for the nutrient-sensitive interaction between raptor and mTOR". Mol Cell. 11 (4): 895–904. doi:10.1016/S1097-2765(03)00114-X. PMID 12718876.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  4. ^ Fang Y, Vilella-Bach M, Bachmann R, Flanigan A, Chen J (2001). "Phosphatidic acid-mediated mitogenic activation of mTOR signaling". Science. 294 (5548): 1942–5. doi:10.1126/science.1066015. PMID 11729323.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  5. ^ Wullschleger S, Loewith R, Hall MN (2006). "TOR signaling in growth and metabolism". Cell. 124 (3): 471–84. doi:10.1016/j.cell.2006.01.016. PMID 16469695. {{cite journal}}: Unknown parameter |month= ignored (help)CS1 maint: multiple names: authors list (link)
  6. ^ Beauchamp EM, Platanias LC (2012). "The evolution of the TOR pathway and its role in cancer". Oncogene. doi:10.1038/onc.2012.567. PMID 23246968. {{cite journal}}: Unknown parameter |month= ignored (help)
  7. ^ Durán RV, Hall MN (2012). "Regulation of TOR by small GTPases". EMBO Rep. 13 (2): 121–8. doi:10.1038/embor.2011.257. PMC 3271343. PMID 22240970. {{cite journal}}: Unknown parameter |month= ignored (help)
  8. ^ Sancak Y, Peterson TR, Shaul YD, Lindquist RA, Thoreen CC, Bar-Peled L, Sabatini DM (2008). "The Rag GTPases bind raptor and mediate amino acid signaling to mTORC1". Science. 320 (5882): 1496–501. doi:10.1126/science.1157535. PMC 2475333. PMID 18497260. {{cite journal}}: Unknown parameter |month= ignored (help)CS1 maint: multiple names: authors list (link)
  9. ^ Saucedo LJ, Gao X, Chiarelli DA, Li L, Pan D, Edgar BA (2003). "Rheb promotes cell growth as a component of the insulin/TOR signalling network". Nat. Cell Biol. 5 (6): 566–71. doi:10.1038/ncb996. PMID 12766776. {{cite journal}}: Unknown parameter |month= ignored (help)CS1 maint: multiple names: authors list (link)
  10. ^ Suzuki T, Inoki K (2011). "Spatial regulation of the mTORC1 system in amino acids sensing pathway". Acta Biochim. Biophys. Sin. (Shanghai). 43 (9): 671–9. doi:10.1093/abbs/gmr066. PMC 3160786. PMID 21785113. {{cite journal}}: Unknown parameter |month= ignored (help)
  11. ^ Zoncu R, Efeyan A, Sabatini DM (2011). "mTOR: from growth signal integration to cancer, diabetes and ageing". Nat. Rev. Mol. Cell Biol. 12 (1): 21–35. doi:10.1038/nrm3025. PMID 21157483. {{cite journal}}: Unknown parameter |month= ignored (help)CS1 maint: multiple names: authors list (link)
  12. ^ a b c Ma XM, Blenis J (2009). "Molecular mechanisms of mTOR-mediated translational control". Nat. Rev. Mol. Cell Biol. 10 (5): 307–18. doi:10.1038/nrm2672. PMID 19339977. {{cite journal}}: Unknown parameter |month= ignored (help)
  13. ^ a b Mendoza MC, Er EE, Blenis J (2011). "The Ras-ERK and PI3K-mTOR pathways: cross-talk and compensation". Trends Biochem. Sci. 36 (6): 320–8. doi:10.1016/j.tibs.2011.03.006. PMC 3112285. PMID 21531565. {{cite journal}}: Unknown parameter |month= ignored (help)CS1 maint: multiple names: authors list (link)
  14. ^ Oshiro N, Takahashi R, Yoshino K, Tanimura K, Nakashima A, Eguchi S, Miyamoto T, Hara K, Takehana K, Avruch J, Kikkawa U, Yonezawa K (2007). "The proline-rich Akt substrate of 40 kDa (PRAS40) is a physiological substrate of mammalian target of rapamycin complex 1". J. Biol. Chem. 282 (28): 20329–39. doi:10.1074/jbc.M702636200. PMC 3199301. PMID 17517883. {{cite journal}}: Unknown parameter |month= ignored (help)CS1 maint: multiple names: authors list (link) CS1 maint: unflagged free DOI (link)
  15. ^ Ye J (2013). "Mechanisms of insulin resistance in obesity". Front Med. 7 (1): 14–24. doi:10.1007/s11684-013-0262-6. PMID 23471659. {{cite journal}}: Unknown parameter |month= ignored (help)
  16. ^ McCubrey JA, Steelman LS, Chappell WH; et al. (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): 1068–111. PMID 23085539. {{cite journal}}: Explicit use of et al. in: |author= (help); Unknown parameter |month= ignored (help)CS1 maint: multiple names: authors list (link)
  17. ^ Ma L, Chen Z, Erdjument-Bromage H, Tempst P, Pandolfi PP (2005). "Phosphorylation and functional inactivation of TSC2 by Erk implications for tuberous sclerosis and cancer pathogenesis". Cell. 121 (2): 179–93. doi:10.1016/j.cell.2005.02.031. PMID 15851026. {{cite journal}}: Unknown parameter |month= ignored (help)CS1 maint: multiple names: authors list (link)
  18. ^ Carrière A, Cargnello M, Julien LA; et al. (2008). "Oncogenic MAPK signaling stimulates mTORC1 activity by promoting RSK-mediated raptor phosphorylation". Curr. Biol. 18 (17): 1269–77. doi:10.1016/j.cub.2008.07.078. PMID 18722121. {{cite journal}}: Explicit use of et al. in: |author= (help); Unknown parameter |month= ignored (help)CS1 maint: multiple names: authors list (link)
  19. ^ a b Majid S, Saini S, Dahiya R (2012). "Wnt signaling pathways in urological cancers: past decades and still growing". Mol. Cancer. 11: 7. doi:10.1186/1476-4598-11-7. PMC 3293036. PMID 22325146.{{cite journal}}: CS1 maint: multiple names: authors list (link) CS1 maint: unflagged free DOI (link)
  20. ^ Salminen A, Hyttinen JM, Kauppinen A, Kaarniranta K (2012). "Context-Dependent Regulation of Autophagy by IKK-NF-κB Signaling: Impact on the Aging Process". Int J Cell Biol. 2012: 849541. doi:10.1155/2012/849541. PMC 3412117. PMID 22899934.{{cite journal}}: CS1 maint: multiple names: authors list (link) CS1 maint: unflagged free DOI (link)
  21. ^ Hardie DG (2007). "AMP-activated/SNF1 protein kinases: conserved guardians of cellular energy". Nat. Rev. Mol. Cell Biol. 8 (10): 774–85. doi:10.1038/nrm2249. PMID 17712357. {{cite journal}}: Unknown parameter |month= ignored (help)
  22. ^ Mihaylova MM, Shaw RJ (2011). "The AMPK signalling pathway coordinates cell growth, autophagy and metabolism". Nat. Cell Biol. 13 (9): 1016–23. doi:10.1038/ncb2329. PMC 3249400. PMID 21892142. {{cite journal}}: Unknown parameter |month= ignored (help)
  23. ^ Gwinn DM, Shackelford DB, Egan DF, Mihaylova MM, Mery A, Vasquez DS, Turk BE, Shaw RJ (2008). "AMPK phosphorylation of raptor mediates a metabolic checkpoint". Mol. Cell. 30 (2): 214–26. doi:10.1016/j.molcel.2008.03.003. PMC 2674027. PMID 18439900. {{cite journal}}: Unknown parameter |month= ignored (help)CS1 maint: multiple names: authors list (link)
  24. ^ Nagalingam A, Arbiser JL, Bonner MY, Saxena NK, Sharma D (2012). "Honokiol activates AMP-activated protein kinase in breast cancer cells via an LKB1-dependent pathway and inhibits breast carcinogenesis". Breast Cancer Res. 14 (1): R35. doi:10.1186/bcr3128. PMC 3496153. PMID 22353783.{{cite journal}}: CS1 maint: multiple names: authors list (link) CS1 maint: unflagged free DOI (link)
  25. ^ Horak P, Crawford AR, Vadysirisack DD, Nash ZM, DeYoung MP, Sgroi D, Ellisen LW (2010). "Negative feedback control of HIF-1 through REDD1-regulated ROS suppresses tumorigenesis". Proc. Natl. Acad. Sci. U.S.A. 107 (10): 4675–80. doi:10.1073/pnas.0907705107. PMC 2842042. PMID 20176937. {{cite journal}}: Unknown parameter |month= ignored (help)CS1 maint: multiple names: authors list (link)
  26. ^ Brugarolas J, Lei K, Hurley RL, Manning BD, Reiling JH, Hafen E, Witters LA, Ellisen LW, Kaelin WG (2004). "Regulation of mTOR function in response to hypoxia by REDD1 and the TSC1/TSC2 tumor suppressor complex". Genes Dev. 18 (23): 2893–904. doi:10.1101/gad.1256804. PMC 534650. PMID 15545625. {{cite journal}}: Unknown parameter |month= ignored (help)CS1 maint: multiple names: authors list (link)
  27. ^ Wang S, Song P, Zou MH (2012). "AMP-activated protein kinase, stress responses and cardiovascular diseases". Clin. Sci. 122 (12): 555–73. doi:10.1042/CS20110625. PMC 3367961. PMID 22390198. {{cite journal}}: Unknown parameter |month= ignored (help)CS1 maint: multiple names: authors list (link)
  28. ^ Martelli AM, Evangelisti C, Chappell W, Abrams SL, Bäsecke J, Stivala F, Donia M, Fagone P, Nicoletti F, Libra M, Ruvolo V, Ruvolo P, Kempf CR, Steelman LS, McCubrey JA (2011). "Targeting the translational apparatus to improve leukemia therapy: roles of the PI3K/PTEN/Akt/mTOR pathway". Leukemia. 25 (7): 1064–79. doi:10.1038/leu.2011.46. PMID 21436840. {{cite journal}}: Unknown parameter |month= ignored (help)CS1 maint: multiple names: authors list (link)
  29. ^ Wang H, Zhang Q, Wen Q, Zheng Y, Lazarovici P, Jiang H, Lin J, Zheng W (2012). "Proline-rich Akt substrate of 40kDa (PRAS40): a novel downstream target of PI3k/Akt signaling pathway". Cell. Signal. 24 (1): 17–24. doi:10.1016/j.cellsig.2011.08.010. PMID 21906675. {{cite journal}}: Unknown parameter |month= ignored (help)CS1 maint: multiple names: authors list (link)
  30. ^ Raught B, Gingras AC (1999). "eIF4E activity is regulated at multiple levels". Int. J. Biochem. Cell Biol. 31 (1): 43–57. PMID 10216943. {{cite journal}}: Unknown parameter |month= ignored (help)
  31. ^ Lee T, Pelletier J (2012). "Eukaryotic initiation factor 4F: a vulnerability of tumor cells". Future Med Chem. 4 (1): 19–31. doi:10.4155/fmc.11.150. PMID 22168162. {{cite journal}}: Unknown parameter |month= ignored (help)
  32. ^ Shahbazian D, Roux PP, Mieulet V, Cohen MS, Raught B, Taunton J, Hershey JW, Blenis J, Pende M, Sonenberg N (2006). "The mTOR/PI3K and MAPK pathways converge on eIF4B to control its phosphorylation and activity". EMBO J. 25 (12): 2781–91. doi:10.1038/sj.emboj.7601166. PMC 1500846. PMID 16763566. {{cite journal}}: Unknown parameter |month= ignored (help)CS1 maint: multiple names: authors list (link)
  33. ^ Schmid T, Jansen AP, Baker AR, Hegamyer G, Hagan JP, Colburn NH (2008). "Translation inhibitor Pdcd4 is targeted for degradation during tumor promotion". Cancer Res. 68 (5): 1254–60. doi:10.1158/0008-5472.CAN-07-1719. PMID 18296647. {{cite journal}}: Unknown parameter |month= ignored (help)CS1 maint: multiple names: authors list (link)
  34. ^ Ma XM, Yoon SO, Richardson CJ, Jülich K, Blenis J (2008). "SKAR links pre-mRNA splicing to mTOR/S6K1-mediated enhanced translation efficiency of spliced mRNAs". Cell. 133 (2): 303–13. doi:10.1016/j.cell.2008.02.031. PMID 18423201. {{cite journal}}: Unknown parameter |month= ignored (help)CS1 maint: multiple names: authors list (link)
  35. ^ Saitoh M, Pullen N, Brennan P, Cantrell D, Dennis PB, Thomas G (2002). "Regulation of an activated S6 kinase 1 variant reveals a novel mammalian target of rapamycin phosphorylation site". J. Biol. Chem. 277 (22): 20104–12. doi:10.1074/jbc.M201745200. PMID 11914378. {{cite journal}}: Unknown parameter |month= ignored (help)CS1 maint: multiple names: authors list (link) CS1 maint: unflagged free DOI (link)
  36. ^ a b Pullen N, Thomas G (1997). "The modular phosphorylation and activation of p70s6k". FEBS Lett. 410 (1): 78–82. doi:10.1016/S0014-5793(97)00323-2. PMID 9247127. {{cite journal}}: Unknown parameter |month= ignored (help)
  37. ^ Pullen N, Dennis PB, Andjelkovic M, Dufner A, Kozma SC, Hemmings BA, Thomas G (1998). "Phosphorylation and activation of p70s6k by PDK1". Science. 279 (5351): 707–10. doi:10.1126/science.279.5351.707. PMID 9445476. {{cite journal}}: Unknown parameter |month= ignored (help)CS1 maint: multiple names: authors list (link)
  38. ^ Peterson RT, Schreiber SL (1998). "Translation control: connecting mitogens and the ribosome". Curr. Biol. 8 (7): R248–50. doi:10.1016/S0960-9822(98)70152-6. PMID 9545190. {{cite journal}}: Unknown parameter |month= ignored (help)
  39. ^ Chiang GG, Abraham RT (2005). "Phosphorylation of mammalian target of rapamycin (mTOR) at Ser-2448 is mediated by p70S6 kinase". J. Biol. Chem. 280 (27): 25485–90. doi:10.1074/jbc.M501707200. PMID 15899889. {{cite journal}}: Unknown parameter |month= ignored (help)CS1 maint: unflagged free DOI (link)
  40. ^ Holz MK, Blenis J (2005). "Identification of S6 kinase 1 as a novel mammalian target of rapamycin (mTOR)-phosphorylating kinase". J. Biol. Chem. 280 (28): 26089–93. doi:10.1074/jbc.M504045200. PMID 15905173. {{cite journal}}: Unknown parameter |month= ignored (help)CS1 maint: unflagged free DOI (link)
  41. ^ Huang S, Houghton PJ (2001). "Mechanisms of resistance to rapamycins". Drug Resist. Updat. 4 (6): 378–91. doi:10.1054/drup.2002.0227. PMID 12030785. {{cite journal}}: Unknown parameter |month= ignored (help)
  42. ^ Gingras AC, Gygi SP, Raught B, Polakiewicz RD, Abraham RT, Hoekstra MF, Aebersold R, Sonenberg N (1999). "Regulation of 4E-BP1 phosphorylation: a novel two-step mechanism". Genes Dev. 13 (11): 1422–37. doi:10.1101/gad.13.11.1422. PMC 316780. PMID 10364159. {{cite journal}}: Unknown parameter |month= ignored (help)CS1 maint: multiple names: authors list (link)
  43. ^ Mothe-Satney I, Brunn GJ, McMahon LP, Capaldo CT, Abraham RT, Lawrence JC (2000). "Mammalian target of rapamycin-dependent phosphorylation of PHAS-I in four (S/T)P sites detected by phospho-specific antibodies". J. Biol. Chem. 275 (43): 33836–43. doi:10.1074/jbc.M006005200. PMID 10942774. {{cite journal}}: Unknown parameter |month= ignored (help)CS1 maint: multiple names: authors list (link) CS1 maint: unflagged free DOI (link)
  44. ^ a b Pause A, Belsham GJ, Gingras AC, Donzé O, Lin TA, Lawrence JC, Sonenberg N (1994). "Insulin-dependent stimulation of protein synthesis by phosphorylation of a regulator of 5'-cap function". Nature. 371 (6500): 762–7. doi:10.1038/371762a0. PMID 7935836. {{cite journal}}: Unknown parameter |month= ignored (help)CS1 maint: multiple names: authors list (link)
  45. ^ Fabrizio P, Pozza F, Pletcher SD, Gendron CM, Longo VD (2001). "Regulation of longevity and stress resistance by Sch9 in yeast". Science. 292 (5515): 288–90. doi:10.1126/science.1059497. PMID 11292860. {{cite journal}}: Unknown parameter |month= ignored (help)CS1 maint: multiple names: authors list (link)
  46. ^ Robida-Stubbs S, Glover-Cutter K, Lamming DW; et al. (2012). "TOR signaling and rapamycin influence longevity by regulating SKN-1/Nrf and DAF-16/FoxO". Cell Metab. 15 (5): 713–24. doi:10.1016/j.cmet.2012.04.007. PMID 22560223. {{cite journal}}: Explicit use of et al. in: |author= (help); Unknown parameter |month= ignored (help)CS1 maint: multiple names: authors list (link)
  47. ^ Harrison DE, Strong R, Sharp ZD; et al. (2009). "Rapamycin fed late in life extends lifespan in genetically heterogeneous mice". Nature. 460 (7253): 392–5. doi:10.1038/nature08221. PMC 2786175. PMID 19587680. {{cite journal}}: Explicit use of et al. in: |author= (help); Unknown parameter |month= ignored (help)CS1 maint: multiple names: authors list (link)
  48. ^ Kaeberlein M, Powers RW, Steffen KK; et al. (2005). "Regulation of yeast replicative life span by TOR and Sch9 in response to nutrients". Science. 310 (5751): 1193–6. doi:10.1126/science.1115535. PMID 16293764. {{cite journal}}: Explicit use of et al. in: |author= (help); Unknown parameter |month= ignored (help)CS1 maint: multiple names: authors list (link)
  49. ^ Blagosklonny MV (2010). "Calorie restriction: decelerating mTOR-driven aging from cells to organisms (including humans)". Cell Cycle. 9 (4): 683–8. PMID 20139716. {{cite journal}}: Unknown parameter |month= ignored (help)
  50. ^ a b Colman RJ, Anderson RM, Johnson SC; et al. (2009). "Caloric restriction delays disease onset and mortality in rhesus monkeys". Science. 325 (5937): 201–4. doi:10.1126/science.1173635. PMC 2812811. PMID 19590001. {{cite journal}}: Explicit use of et al. in: |author= (help); Unknown parameter |month= ignored (help)CS1 maint: multiple names: authors list (link)
  51. ^ Ho AD, Wagner W, Mahlknecht U (2005). "Stem cells and ageing. The potential of stem cells to overcome age-related deteriorations of the body in regenerative medicine". EMBO Rep. 6 Spec No: S35–8. doi:10.1038/sj.embor.7400436. PMC 1369281. PMID 15995659. {{cite journal}}: Unknown parameter |month= ignored (help)CS1 maint: multiple names: authors list (link)
  52. ^ a b Murakami M, Ichisaka T, Maeda M; et al. (2004). "mTOR is essential for growth and proliferation in early mouse embryos and embryonic stem cells". Mol. Cell. Biol. 24 (15): 6710–8. doi:10.1128/MCB.24.15.6710-6718.2004. PMC 444840. PMID 15254238. {{cite journal}}: Explicit use of et al. in: |author= (help); Unknown parameter |month= ignored (help)CS1 maint: multiple names: authors list (link)
  53. ^ Gangloff YG, Mueller M, Dann SG; et al. (2004). "Disruption of the mouse mTOR gene leads to early postimplantation lethality and prohibits embryonic stem cell development". Mol. Cell. Biol. 24 (21): 9508–16. doi:10.1128/MCB.24.21.9508-9516.2004. PMC 522282. PMID 15485918. {{cite journal}}: Explicit use of et al. in: |author= (help); Unknown parameter |month= ignored (help)CS1 maint: multiple names: authors list (link)
  54. ^ a b Chen C, Liu Y, Liu Y, Zheng P (2009). "mTOR regulation and therapeutic rejuvenation of aging hematopoietic stem cells". Sci Signal. 2 (98): ra75. doi:10.1126/scisignal.2000559. PMID 19934433.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  55. ^ Limon JJ, Fruman DA (2012). "Akt and mTOR in B Cell Activation and Differentiation". Front Immunol. 3: 228. doi:10.3389/fimmu.2012.00228. PMC 3412259. PMID 22888331.{{cite journal}}: CS1 maint: unflagged free DOI (link)
  56. ^ Araki K, Turner AP, Shaffer VO; et al. (2009). "mTOR regulates memory CD8 T-cell differentiation". Nature. 460 (7251): 108–12. doi:10.1038/nature08155. PMC 2710807. PMID 19543266. {{cite journal}}: Explicit use of et al. in: |author= (help); Unknown parameter |month= ignored (help)CS1 maint: multiple names: authors list (link)
  57. ^ a b Araki K, Youngblood B, Ahmed R (2010). "The role of mTOR in memory CD8 T-cell differentiation". Immunol. Rev. 235 (1): 234–43. doi:10.1111/j.0105-2896.2010.00898.x. PMID 20536567. {{cite journal}}: Unknown parameter |month= ignored (help)CS1 maint: multiple names: authors list (link)
  58. ^ Russell RC, Fang C, Guan KL (2011). "An emerging role for TOR signaling in mammalian tissue and stem cell physiology". Development. 138 (16): 3343–56. doi:10.1242/dev.058230. PMC 3143559. PMID 21791526. {{cite journal}}: Unknown parameter |month= ignored (help)CS1 maint: multiple names: authors list (link)
  59. ^ a b Choi AM, Ryter SW, Levine B (2013). "Autophagy in human health and disease". N. Engl. J. Med. 368 (7): 651–62. doi:10.1056/NEJMra1205406. PMID 23406030. {{cite journal}}: Unknown parameter |month= ignored (help)CS1 maint: multiple names: authors list (link)
  60. ^ Murrow L, Debnath J (2013). "Autophagy as a stress-response and quality-control mechanism: implications for cell injury and human disease". Annu Rev Pathol. 8: 105–37. doi:10.1146/annurev-pathol-020712-163918. PMID 23072311. {{cite journal}}: Unknown parameter |month= ignored (help)
  61. ^ Alers S, Löffler AS, Wesselborg S, Stork B (2012). "Role of AMPK-mTOR-Ulk1/2 in the regulation of autophagy: cross talk, shortcuts, and feedbacks". Mol. Cell. Biol. 32 (1): 2–11. doi:10.1128/MCB.06159-11. PMC 3255710. PMID 22025673. {{cite journal}}: Unknown parameter |month= ignored (help)CS1 maint: multiple names: authors list (link)
  62. ^ Pyo JO, Nah J, Jung YK (2012). "Molecules and their functions in autophagy". Exp. Mol. Med. 44 (2): 73–80. doi:10.3858/emm.2012.44.2.029. PMC 3296815. PMID 22257882. {{cite journal}}: Unknown parameter |month= ignored (help)CS1 maint: multiple names: authors list (link)
  63. ^ Proud CG (2007). "Amino acids and mTOR signalling in anabolic function". Biochem. Soc. Trans. 35 (Pt 5): 1187–90. doi:10.1042/BST0351187. PMID 17956308. {{cite journal}}: Unknown parameter |month= ignored (help)
  64. ^ Cuervo AM, Dice JF (2000). "Age-related decline in chaperone-mediated autophagy". J. Biol. Chem. 275 (40): 31505–13. doi:10.1074/jbc.M002102200. PMID 10806201. {{cite journal}}: Unknown parameter |month= ignored (help)CS1 maint: unflagged free DOI (link)
  65. ^ Codogno P, Meijer AJ (2005). "Autophagy and signaling: their role in cell survival and cell death". Cell Death Differ. 12 Suppl 2: 1509–18. doi:10.1038/sj.cdd.4401751. PMID 16247498. {{cite journal}}: Unknown parameter |month= ignored (help)
  66. ^ Apel K, Hirt H (2004). "Reactive oxygen species: metabolism, oxidative stress, and signal transduction". Annu Rev Plant Biol. 55: 373–99. doi:10.1146/annurev.arplant.55.031903.141701. PMID 15377225.
  67. ^ Murphy MP (2009). "How mitochondria produce reactive oxygen species". Biochem. J. 417 (1): 1–13. doi:10.1042/BJ20081386. PMC 2605959. PMID 19061483. {{cite journal}}: Unknown parameter |month= ignored (help)
  68. ^ Bonawitz ND, Chatenay-Lapointe M, Pan Y, Shadel GS (2007). "Reduced TOR signaling extends chronological life span via increased respiration and upregulation of mitochondrial gene expression". Cell Metab. 5 (4): 265–77. doi:10.1016/j.cmet.2007.02.009. PMC 3460550. PMID 17403371. {{cite journal}}: Unknown parameter |month= ignored (help)CS1 maint: multiple names: authors list (link)
  69. ^ Adam-Vizi V (2005). "Production of reactive oxygen species in brain mitochondria: contribution by electron transport chain and non-electron transport chain sources". Antioxid. Redox Signal. 7 (9–10): 1140–9. doi:10.1089/ars.2005.7.1140. PMID 16115017.
  70. ^ Sun Q, Chen X, Ma J; et al. (2011). "Mammalian target of rapamycin up-regulation of pyruvate kinase isoenzyme type M2 is critical for aerobic glycolysis and tumor growth". Proc. Natl. Acad. Sci. U.S.A. 108 (10): 4129–34. doi:10.1073/pnas.1014769108. PMC 3054028. PMID 21325052. {{cite journal}}: Explicit use of et al. in: |author= (help); Unknown parameter |month= ignored (help)CS1 maint: multiple names: authors list (link)
  71. ^ Sporn MB, Liby KT (2012). "NRF2 and cancer: the good, the bad and the importance of context". Nat. Rev. Cancer. 12 (8): 564–71. doi:10.1038/nrc3278. PMID 22810811. {{cite journal}}: Unknown parameter |month= ignored (help)
  72. ^ Liu M, Wilk SA, Wang A; et al. (2010). "Resveratrol inhibits mTOR signaling by promoting the interaction between mTOR and DEPTOR". J. Biol. Chem. 285 (47): 36387–94. doi:10.1074/jbc.M110.169284. PMC 2978567. PMID 20851890. {{cite journal}}: Explicit use of et al. in: |author= (help); Unknown parameter |month= ignored (help)CS1 maint: multiple names: authors list (link) CS1 maint: unflagged free DOI (link)
  73. ^ Miwa S, Sugimoto N, Yamamoto N; et al. (2012). "Caffeine induces apoptosis of osteosarcoma cells by inhibiting AKT/mTOR/S6K, NF-κB and MAPK pathways". Anticancer Res. 32 (9): 3643–9. PMID 22993301. {{cite journal}}: Explicit use of et al. in: |author= (help); Unknown parameter |month= ignored (help)CS1 maint: multiple names: authors list (link)
  74. ^ Vézina C, Kudelski A, Sehgal SN (1975). "Rapamycin (AY-22,989), a new antifungal antibiotic. I. Taxonomy of the producing streptomycete and isolation of the active principle". J. Antibiot. 28 (10): 721–6. PMID 1102508. {{cite journal}}: Unknown parameter |month= ignored (help)CS1 maint: multiple names: authors list (link)
  75. ^ Tsang CK, Qi H, Liu LF, Zheng XF (2007). "Targeting mammalian target of rapamycin (mTOR) for health and diseases". Drug Discov. Today. 12 (3–4): 112–24. doi:10.1016/j.drudis.2006.12.008. PMID 17275731. {{cite journal}}: Unknown parameter |month= ignored (help)CS1 maint: multiple names: authors list (link)
  76. ^ a b c Vilar E, Perez-Garcia J, Tabernero J (2011). "Pushing the envelope in the mTOR pathway: the second generation of inhibitors". Mol. Cancer Ther. 10 (3): 395–403. doi:10.1158/1535-7163.MCT-10-0905. PMC 3413411. PMID 21216931. {{cite journal}}: Unknown parameter |month= ignored (help)CS1 maint: multiple names: authors list (link)
  77. ^ De P, Miskimins K, Dey N, Leyland-Jones B (2013). "Promise of rapalogues versus mTOR kinase inhibitors in subset specific breast cancer: Old targets new hope". Cancer Treat. Rev. doi:10.1016/j.ctrv.2012.12.002. PMID 23352077. {{cite journal}}: Unknown parameter |month= ignored (help)CS1 maint: multiple names: authors list (link)
  78. ^ Nashan B, Citterio F (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): 547–61. doi:10.1097/TP.0b013e3182551021. PMID 22941182. {{cite journal}}: Unknown parameter |month= ignored (help)
  79. ^ Townsend JC, Rideout P, Steinberg DH (2012). "Everolimus-eluting stents in interventional cardiology". Vasc Health Risk Manag. 8: 393–404. doi:10.2147/VHRM.S23388. PMC 3402052. PMID 22910420.{{cite journal}}: CS1 maint: multiple names: authors list (link) CS1 maint: unflagged free DOI (link)
  80. ^ Voss MH, Molina AM, Motzer RJ (2011). "mTOR inhibitors in advanced renal cell carcinoma". Hematol. Oncol. Clin. North Am. 25 (4): 835–52. doi:10.1016/j.hoc.2011.04.008. PMC 3587783. PMID 21763970. {{cite journal}}: Unknown parameter |month= ignored (help)CS1 maint: multiple names: authors list (link)
  81. ^ Smith SM (2012). "Targeting mTOR in mantle cell lymphoma: current and future directions". Best Pract Res Clin Haematol. 25 (2): 175–83. doi:10.1016/j.beha.2012.04.008. PMID 22687453. {{cite journal}}: Unknown parameter |month= ignored (help)
  82. ^ Fasolo A, Sessa C (2012). "Targeting mTOR pathways in human malignancies". Curr. Pharm. Des. 18 (19): 2766–77. PMID 22475451.
  83. ^ Budde K, Gaedeke J (2012). "Tuberous sclerosis complex-associated angiomyolipomas: focus on mTOR inhibition". Am. J. Kidney Dis. 59 (2): 276–83. doi:10.1053/j.ajkd.2011.10.013. PMID 22130643. {{cite journal}}: Unknown parameter |month= ignored (help)
  84. ^ a b Zhang YJ, Duan Y, Zheng XF (2011). "Targeting the mTOR kinase domain: the second generation of mTOR inhibitors". Drug Discov. Today. 16 (7–8): 325–31. doi:10.1016/j.drudis.2011.02.008. PMC 3073023. PMID 21333749. {{cite journal}}: Unknown parameter |month= ignored (help)CS1 maint: multiple names: authors list (link)
  85. ^ Veilleux A, Houde VP, Bellmann K, Marette A (2010). "Chronic inhibition of the mTORC1/S6K1 pathway increases insulin-induced PI3K activity but inhibits Akt2 and glucose transport stimulation in 3T3-L1 adipocytes". Mol. Endocrinol. 24 (4): 766–78. doi:10.1210/me.2009-0328. PMID 20203102. {{cite journal}}: Unknown parameter |month= ignored (help)CS1 maint: multiple names: authors list (link)
  86. ^ Johnson SC, Rabinovitch PS, Kaeberlein M (2013). "mTOR is a key modulator of ageing and age-related disease". Nature. 493 (7432): 338–45. doi:10.1038/nature11861. PMID 23325216. {{cite journal}}: Unknown parameter |month= ignored (help)CS1 maint: multiple names: authors list (link)