User:Flemingrjf/sandbox

From WikiProjectMed
Jump to navigation Jump to search

mTOR

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

History/Discovery

mTOR is considered to be the mammalian target of rapamycin. Rapamycin was discovered in a soil sample from Easter Island in the 1970s.[1] Researchers studied this sample and found that the bacterium Streptomyces hygroscopicus made an antifungal, which they named rapamycin after the island's name Rapa Nui, which it was called by the locals meaning "navel of the world."[2] Studies on rapamycin revealed that it was a powerful antifungal agent that could arrest fungal activity at the G1 phase. It was then tested in rats as a potential antifungal drug in humans, and was found to also greatly suppress their immune system by blocking the G1 to S phase transition in T-lymphocytes.[3] This has led to its clinical use as an immunosupressant following organ transplantation.[4]

In 1991, a genetic screen was performed on Saccharomyces cerevisiae to elucidate what rapamycin was specifically targeting to initiate this response. It was found that knockout of three genes allowed for the fungus' resistance to rapamycin.[5] Two of the genes were called targets of rapamycin, or TOR, while the third gene was already characterized to be Fpr1, which is now known to be a binding protein in the TOR complexes.[6] In 1994, the mammalian target of rapamycin (mTOR) was identified at the rapamycin target in mammals.[7]

mTOR Structure

The mammalian target of rapamycin (mTOR) is considered a 'master switch' of cellular processes, regulating transcription of proteins required for cell growth and proliferation by sensing energy and nutrient levels.[8] As a result, it plays an important role in various human diseases, including cancer and diabetes.[8] mTOR consists of two structures, mTOR complex 1 (mTORc1) and mTOR complex 2 (mTORc2). mTORc1, named for its putative sensitivity to rapamycin[citation needed], is the better-studied component of the mTOR protein and is thought to play the principal role in the initial mTOR protein; it is believed to regulate cell growth, proliferation, and survival by integrating hormones, growth factors, nutrients, stressors, and energy signals.[8] mTORc2 is thought to regulate cytoskeleton and cell survival in response to insulin, although the upstream signaling pathways have yet to be fully elucidated.[9]

mTORc1 and 2 share the same core protein complex of 2549 amino acids, which contains seven conserved domains, written from the N- to the C- terminus as HEAT, FAT, FRB, kinase, NR, and FATC domains.[9] The largest domain is called the HEAT repeat domain, named for the Huntingtin, Ef2, A subunit of PP2A, TOR1 amino acid repeats that it contains.[9] The HEAT domain is implicated in protein-protein interactions. The FAT domain I need to research more about. The FRB domain is specifically targeted by rapamycin for inhibition for the mTOR complexes.[9] It is important to note, that only the mTORc1 complex has been shown to be inhibited by rapamycin. The kinase domain is the catalytic domain responsible for phosphorylating serine and threonine residues on target proteins.[9] Attached to the kinase domain is the mLST8 domain, the mammalian ortholog of the LST8 protein first discovered in yeast.[9] mTORc2 has been shown to require this domain for function, and it has been suggested that it is also required in mTORc1. The NR (negative regulatory) domain is the putative negative regulatory region of the mTOR complexes.[9] The FATC domain, or FAT domain on the C-terminus, has been shown to be necessary for the kinase function of mTOR, as a single amino acid deletion from this sequence prevents such activity.[9]

There are two complexes that contain the mTOR core protein: mTOR complex 1 and mTOR complex 2 named mTORC1 and mTORC2 respectively.

The mTOR core protein is 2549 amino acids long and contains a HEAT repeats domain, FAT domain, FRB domain, kinase domain, NR domain, and FATC domain. The HEAT domain spans the first half of the core protein and consists of many tandem HEAT repeats which stands for the amino acid sequences for the domains of for Huntingtin, EF3, A subunit of PP2A, and TOR1. These HEAT repeats are considered to be for protein-protein interactions with either Raptor or Rictor for mTORC1 or mTORC2, respectively. FAT lies downstream of the HEAT region and will interact with the FAT C terminal domain and this interaction is thought to modulate the kinase activity of mTOR. This FATC domain is so conserved that changing one amino acid in this sequence has been shown to disrupt mTOR activity. The FRB (FKB12-rapamycin binding) domain is the stretch of amino acids that rapamycin binds to in order to inhibit mTOR activity. The NR domain is located just before the FATC domain on the C terminus and is a putative negative regulatory region, where substrates can bind to inhibit mTOR activity.

The mTORC1 and mTORC2 complexes differ due to the associated proteins that make up their complexes.


mTORc1 and mTORc2 both have specific regulatory proteins, known as Raptor and Rictor, respectively.[9] I will go on to talk about Raptor and Rictor next time.

mTORC1

Upstream of mTORC1

The role of mTOR 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. [10]

All of these variables for protein synthesis affect mTORC1 activation by interacting with the TSC1/TSC2 complex. TSC2 is a GTP-ase activating protein (GAP). Its GAP activity interacts with 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. [11] 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 losing its GAP activity, or the phosphorylation can cause the heterodimer to have more active GAP activity, depending on the kinase phosphorylating the dimer. [12]

Growth Factors

Growth factors like insulin can activate mTORC1 through the receptor tyrosine kinase (RTK) pathway. Ultimately Akt phosphorylates TSC2 on serine residue 939, serine residue 981, and threonine residue 1462. 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 Raptor on mTORC1. PRAS40 prevents Raptor from recruiting mTORC1's substrates 4E-BP1 and S6K-1. Thus when PRAS40 falls off of Raptor, the two substrates are recruited to mTORC1 and thereby activated in this way.[14]

Because insulin is a factor that is secreted by pancreatic beta cells upon glucose elevation in the blood, this 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.[15] This has great significance in diabetes mellitus, which is due to insulin resistance.[16]

Mitogens

Mitogens like insulin like growth factor 1 (IGF1) can activate the Ras-ERK pathway, which can control the TSC1/TSC2 complex as well as directly have the same downstream role that mTORC1 has.[13] In this pathway, the G protein Ras is tethered to the plasma membrane via a farnesyl group and is its inactive GDP state. Upon growth factor binding to the adjacent receptor tyrosine kinase, the adaptor protein GRB2 gets 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). 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. These phosphorylations will cause the heterodimer to fall apart, and not be able to deactivate Rheb. This, thus keeps mTORC1 active.

RSK has also been shown to phosphorylate Raptor, which helps it overcome the inhibitory effects of PRAS40.

Cytokines

Cytokines like tumor necrosis factor alpha (TFNalpha), can induce mTOR activity through IKK beta.[17] 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 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.

AMPK can phosphorylate TSC2 on serine residue 1387, which will now activate the GAP activity of this complex, causing Rheb-GTP to be hydrolyzed into Rheb-GDP. This inactivates mTORC1, and no protein synthesis occurs through this pathway.[18]

AMPK can also phosphorylate Raptor on two serine residues. This phosphorylated Raptor now recruits 14-3-3, to bind to it, preventing Raptor from being part of the mTORC1 complex. Since mTORC1 cannot recruit its substrates without Raptor, no protein synthesis via mTORC1 occurs.[19]

LBK1 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.[20]

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 (HIF-1 alpha) will stabilize and activate transcription of REDD1. 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 inactive.[21]

Due to the lack of synthesis of ATP in the mitochondria under hypoxic stress, AMPK will also become active and thus inhibit mTORC1 through its processes.[22]

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 (GSK3 beta).[23] 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, when the Wnt pathway is active, so also is the mTORC1 pathway. Now, mTORC1 can activate protein synthesis for the developing organism.[23]

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 GTPases heterodimers to switch to its active conformation. Active Rag heterodimers interact with RAPTOR, localizing mTORC1 to the surface of late endosomes and lysosome where the Rag GTPases are located.[24] This allows mTORC1 to physically interact with RHEB, which is activated by growth factors such as insulin. The interaction between the Rags and mTORC1 brings mTORC1 to the surface of endosomes and lysosomes where Rheb is located. This is where Rheb-GTP activates mTORC1.[25]

Downstream of mTORC1

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

4E-BP1

Activated mTORC1 will phosphorylate transcription inhibitor 4E-BP1, releasing it from eukaryotic translation initiation factor 4E (eIF4E).[26] eIF4E is now free to join the eukaryotic translation initiation factor 4G (eIF4G) and the eukaryotic translation initiation factor 4A (eIF4A).[27] 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).[28] The helicase is required to remove hairpin loops that arise in the 5' untranslated regions of mRNA, that 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 codon start site, because the hairpin loop has been eradicated by the eIF4A helicase.[29]

S6K

mTORC1 will also activate S6K, which is responsible for the recruitment of eIF4B to the initiation complex by phosphorylating its serine residue 422.[30]

S6K also can phosphorylate programmed cell death 4 (PDCD4), which marks it for degradation by ubiquidin ligase Beta-TrCP. PDCD4 is a tumor suppressor that binds to eIF4A and prevents it from being incorporated into the initiation complex.[31]

Active S6K can bind to the SKAR scaffold protein that can get recruited to exon junction complexes. 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.[32]

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.[15]

mTORC1 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. [33] 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.[34] [35]

Based on upstream signaling of mTORC1, a clear relationship between food consumption and mTORC1 activity has been observed.[36] 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.[37]

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.[38] 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.[38]

Stem Cells

Conservation of stem cells in the body has been shown to help prevent against premature aging.[39] mTORC1 activity plays a critical role in the growth and proliferation of stem cells.[40] Knocking out mTORC1 results in embryonic lethality due to lack of trophoblast development.[41] Treating stem cells with rapamycin will also slow their proliferation, conserving the stem cells in their undifferentiated condition. [40]

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.[42] Rapamycin is used clinically as an immunosupressant and prevents the proliferation of T cells and B cells.[43] 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 memory precursor cells during the expansion phase of T cell development .[44] 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.[45] mTORC1 inhibition with rapamycin has also been linked to a dramatic increase of B cells in old mice, enhancing their immune systems.[42] This paradox of rapamycin inhibiting the immune system response has been linked to several reasons, including its interaction with T-regulatory cells.[45] The mechanisms mTORC1's inhibition on proliferation and differentiation of hematopoietic stem cells has yet to be fully elucidated.[46]

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.[47] 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.[47] Autophagy can thus remove aggregates of proteins and damaged organelles, that can lead to cellular dysfunction.[48]

Upon activation, mTORC1 will phosphorylate Atg 13, preventing it from entering the ULK1 kinase complex, which consists of Atg1-Atg17-Atg101.[49] This prevents the structure from being recruited to the preautophagosomal structure at the plasma membrane, inhibiting autophagy.[50].

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. [51] Because autophagy appears to decline with age, activation of autophagy may help promote longevity in humans. [52] Problems in proper autophagy processes have been linked to diabetes, cardiovascular disease, neurodegenerative diseases, and cancer.[53]

Reactive Oxygen Species

Reactive oxygen species can damage the DNA and proteins in cells.[54] A majority of them arise in the mitochondria.[55]

Deletion of the TOR1 gene in yeast increases mitochondrial respiration by enhancing the translation of mitochondrial DNA that encodes for the complexes involved in the electron transport chain.[56] 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.[57] 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 phoshphorylation in the inner membrane of the mitochondria.[58]

Inhibition of mTORC1 has also been shown to increase transcription of the 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.[59]

mTORC1 Inhibition

Rapamycin was the first known inhibitor of mTORC1, considering that mTORC1 was discovered as being the target of rapamycin.[1] Rapamycin will bind to cytosolic FKBP12 and act as a scaffold molecule, allowing this protein to dock on the FBP regulatory region on mTORC1.[60] 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.[61] These drugs are considered the first generation inhibitors of mTOR.[62]

Siroliumus, which is the drug name for rapamycin, was approved by the FDA in 1999 to prevent against host rejection in patients undergoing kidney transplantation.[63] In 2003, it was approved as a stent covering for people who want to widen their arteries to prevent against heart attacks and stuff.[64] In 2007, they began being approved for treatments against cancers such as renal cell carcinoma.[65] In 2008 they were approved for mantle cell lymphoma.[66] mTORC1 inhibitors have recently been approved for treatment of pancreatic cancer.[67] In 2010 they were approved for treatment of tuberous sclerosis.[68]

The second generation of inhibitors were created to overcome problems with upstream signaling upon the introduction of first generation inhibitors to the treated cells. [69] 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.[70] 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.[61] 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 site on the kinase domain of the mTOR core protein itself and abolish activity of both mTOR complexes. [69] In addition, since the mTOR and the PI3K proteins are both in the same PIKK family of kinases, some second generation inhibitors have dual inhibition towards the mTOR complexes as well as PI3K, which acts upstream of mTORC1.[61] As of 2011, these second generation inhibitors were in phase II of testing.

There have been several dietary compounds that have been suggested to inhibit mTOR signaling including EGCG, resveratrol, curcumin, caffeine, and alcohol.[71] [72]

There are currently more than 1,300 clinical trials underway for the mTOR complex inhibitors. [73]

References

  1. ^ a b 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". J. Antibiot. 28 (10): 721–6. doi:10.7164/antibiotics.28.721. PMID 1102508.{{cite journal}}: CS1 maint: date and year (link) CS1 maint: multiple names: authors list (link)
  2. ^ Dobashi Y, Watanabe Y, Miwa C, Suzuki S, Koyama S (June 2011). "Mammalian target of rapamycin: a central node of complex signaling cascades". Int J Clin Exp Pathol. 4 (5): 476–95. PMC 3127069. PMID 21738819.{{cite journal}}: CS1 maint: date and year (link) CS1 maint: multiple names: authors list (link)
  3. ^ Magnuson B, Ekim B, Fingar DC (January 2012). "Regulation and function of ribosomal protein S6 kinase (S6K) within mTOR signalling networks". Biochem. J. 441 (1): 1–21. doi:10.1042/BJ20110892. PMID 22168436.{{cite journal}}: CS1 maint: date and year (link) CS1 maint: multiple names: authors list (link)
  4. ^ Abraham RT, Wiederrecht GJ (1996). "Immunopharmacology of rapamycin". Annu. Rev. Immunol. 14: 483–510. doi:10.1146/annurev.immunol.14.1.483. PMID 8717522.
  5. ^ Heitman J, Movva NR, Hall MN (August 1991). "Targets for cell cycle arrest by the immunosuppressant rapamycin in yeast". Science. 253 (5022): 905–9. doi:10.1126/science.1715094. PMID 1715094.{{cite journal}}: CS1 maint: date and year (link) CS1 maint: multiple names: authors list (link)
  6. ^ Foster KG, Fingar DC (May 2010). "Mammalian target of rapamycin (mTOR): conducting the cellular signaling symphony". J. Biol. Chem. 285 (19): 14071–7. doi:10.1074/jbc.R109.094003. PMC 2863215. PMID 20231296.{{cite journal}}: CS1 maint: date and year (link)
  7. ^ Brown EJ, Albers MW, Shin TB; et al. (June 1994). "A mammalian protein targeted by G1-arresting rapamycin-receptor complex". Nature. 369 (6483): 756–8. doi:10.1038/369756a0. PMID 8008069. {{cite journal}}: Explicit use of et al. in: |author= (help)CS1 maint: date and year (link) CS1 maint: multiple names: authors list (link)
  8. ^ a b c Zhou, H.; Huang, S. (2010). "The complexes of mammalian target of rapamycin". Current Protein & Peptide Science. 11 (6): 409–424. doi:10.2174/138920310791824093. PMC 2928868. PMID 20491627.
  9. ^ a b c d e f g h i Bhaskar, P. T.; Hay, N. (2007). "The Two TORCs and Akt". Developmental Cell. 12 (4): 487–502. doi:10.1016/j.devcel.2007.03.020. PMID 17419990.
  10. ^ Wullschleger S, Loewith R, Hall MN (February 2006). "TOR signaling in growth and metabolism". Cell. 124 (3): 471–84. doi:10.1016/j.cell.2006.01.016. PMID 16469695.{{cite journal}}: CS1 maint: date and year (link) CS1 maint: multiple names: authors list (link)
  11. ^ Beauchamp EM, Platanias LC (December 2012). "The evolution of the TOR pathway and its role in cancer". Oncogene. 32 (34): 3923–3932. doi:10.1038/onc.2012.567. PMID 23246968.{{cite journal}}: CS1 maint: date and year (link)
  12. ^ Durán RV, Hall MN (February 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}}: CS1 maint: date and year (link)
  13. ^ a b Mendoza MC, Er EE, Blenis J (June 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}}: CS1 maint: date and year (link) CS1 maint: multiple names: authors list (link)
  14. ^ 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". J. Biol. Chem. 282 (28): 20329–39. doi:10.1074/jbc.M702636200. PMC 3199301. PMID 17517883. {{cite journal}}: Explicit use of et al. in: |author= (help)CS1 maint: date and year (link) CS1 maint: multiple names: authors list (link)
  15. ^ a b Ma XM, Blenis J (May 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}}: CS1 maint: date and year (link)
  16. ^ Ye J (March 2013). "Mechanisms of insulin resistance in obesity". Front Med. 7 (1): 14–24. doi:10.1007/s11684-013-0262-6. PMC 3936017. PMID 23471659.{{cite journal}}: CS1 maint: date and year (link)
  17. ^ 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)
  18. ^ Mihaylova MM, Shaw RJ (September 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}}: CS1 maint: date and year (link)
  19. ^ Gwinn DM, Shackelford DB, Egan DF; et al. (April 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}}: Explicit use of et al. in: |author= (help)CS1 maint: date and year (link) CS1 maint: multiple names: authors list (link)
  20. ^ 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)
  21. ^ Horak P, Crawford AR, Vadysirisack DD; et al. (March 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}}: Explicit use of et al. in: |author= (help)CS1 maint: date and year (link) CS1 maint: multiple names: authors list (link)
  22. ^ Wang S, Song P, Zou MH (June 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}}: CS1 maint: date and year (link) CS1 maint: multiple names: authors list (link)
  23. ^ 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)
  24. ^ Sancak Y, Peterson TR, Shaul YD, Lindquist RA, Thoreen CC, Bar-Peled L, Sabatini DM (June 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}}: CS1 maint: date and year (link) CS1 maint: multiple names: authors list (link)
  25. ^ Suzuki T, Inoki K (September 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}}: CS1 maint: date and year (link)
  26. ^ 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): 1064–79. doi:10.1038/leu.2011.46. PMID 21436840. {{cite journal}}: Explicit use of et al. in: |author= (help)CS1 maint: date and year (link) CS1 maint: multiple names: authors list (link)
  27. ^ 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". Cell. Signal. 24 (1): 17–24. doi:10.1016/j.cellsig.2011.08.010. PMID 21906675. {{cite journal}}: Explicit use of et al. in: |author= (help)CS1 maint: date and year (link) CS1 maint: multiple names: authors list (link)
  28. ^ Raught B, Gingras AC (January 1999). "eIF4E activity is regulated at multiple levels". Int. J. Biochem. Cell Biol. 31 (1): 43–57. doi:10.1016/s1357-2725(98)00131-9. PMID 10216943.{{cite journal}}: CS1 maint: date and year (link)
  29. ^ Lee T, Pelletier J (January 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}}: CS1 maint: date and year (link)
  30. ^ Shahbazian D, Roux PP, Mieulet V; et al. (June 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}}: Explicit use of et al. in: |author= (help)CS1 maint: date and year (link) CS1 maint: multiple names: authors list (link)
  31. ^ Schmid T, Jansen AP, Baker AR, Hegamyer G, Hagan JP, Colburn NH (March 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}}: CS1 maint: date and year (link) CS1 maint: multiple names: authors list (link)
  32. ^ Ma XM, Yoon SO, Richardson CJ, Jülich K, Blenis J (April 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}}: CS1 maint: date and year (link) CS1 maint: multiple names: authors list (link)
  33. ^ Fabrizio P, Pozza F, Pletcher SD, Gendron CM, Longo VD (April 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}}: CS1 maint: date and year (link) CS1 maint: multiple names: authors list (link)
  34. ^ 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 Metab. 15 (5): 713–24. doi:10.1016/j.cmet.2012.04.007. PMC 3348514. PMID 22560223. {{cite journal}}: Explicit use of et al. in: |author= (help)CS1 maint: date and year (link) CS1 maint: multiple names: authors list (link)
  35. ^ Harrison DE, Strong R, Sharp ZD; et al. (July 2009). "Rapamycin fed late in life extends lifespan in genetically heterogeneous mice". Nature. 460 (7253): 392–5. doi:10.1038/nature08221. PMC 2786175. PMID 19587680. {{cite journal}}: Explicit use of et al. in: |author= (help)CS1 maint: date and year (link) CS1 maint: multiple names: authors list (link)
  36. ^ 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): 1193–6. doi:10.1126/science.1115535. PMID 16293764. {{cite journal}}: Explicit use of et al. in: |author= (help)CS1 maint: date and year (link) CS1 maint: multiple names: authors list (link)
  37. ^ Blagosklonny MV (February 2010). "Calorie restriction: decelerating mTOR-driven aging from cells to organisms (including humans)". Cell Cycle. 9 (4): 683–8. doi:10.4161/cc.9.4.10766. PMID 20139716.{{cite journal}}: CS1 maint: date and year (link)
  38. ^ a b Colman RJ, Anderson RM, Johnson SC; et al. (July 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)CS1 maint: date and year (link) CS1 maint: multiple names: authors list (link)
  39. ^ 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 Rep. 6 Spec No (Suppl 1): S35–8. doi:10.1038/sj.embor.7400436. PMC 1369281. PMID 15995659.{{cite journal}}: CS1 maint: date and year (link) CS1 maint: multiple names: authors list (link)
  40. ^ a b 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". 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)CS1 maint: date and year (link) CS1 maint: multiple names: authors list (link)
  41. ^ 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". 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)CS1 maint: date and year (link) CS1 maint: multiple names: authors list (link)
  42. ^ 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. PMC 4020596. PMID 19934433.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  43. ^ 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.
  44. ^ Araki K, Turner AP, Shaffer VO; et al. (July 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)CS1 maint: date and year (link) CS1 maint: multiple names: authors list (link)
  45. ^ a b Araki K, Youngblood B, Ahmed R (May 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. PMC 3760155. PMID 20536567.{{cite journal}}: CS1 maint: date and year (link) CS1 maint: multiple names: authors list (link)
  46. ^ Russell RC, Fang C, Guan KL (August 2011). "An emerging role for TOR signaling in mammalian tissue and stem cell physiology". Development. 138 (16): 3343–56. doi:10.1242/dev.058230. PMC 3143559. PMID 21791526.{{cite journal}}: CS1 maint: date and year (link) CS1 maint: multiple names: authors list (link)
  47. ^ a b Choi AM, Ryter SW, Levine B (February 2013). "Autophagy in human health and disease". N. Engl. J. Med. 368 (7): 651–62. doi:10.1056/NEJMra1205406. PMID 23406030.{{cite journal}}: CS1 maint: date and year (link) CS1 maint: multiple names: authors list (link)
  48. ^ Murrow L, Debnath J (January 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. PMC 3971121. PMID 23072311.{{cite journal}}: CS1 maint: date and year (link)
  49. ^ Alers S, Löffler AS, Wesselborg S, Stork B (January 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}}: CS1 maint: date and year (link) CS1 maint: multiple names: authors list (link)
  50. ^ Pyo JO, Nah J, Jung YK (February 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}}: CS1 maint: date and year (link) CS1 maint: multiple names: authors list (link)
  51. ^ Proud CG (November 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}}: CS1 maint: date and year (link)
  52. ^ Cuervo AM, Dice JF (October 2000). "Age-related decline in chaperone-mediated autophagy". J. Biol. Chem. 275 (40): 31505–13. doi:10.1074/jbc.M002102200. PMID 10806201.{{cite journal}}: CS1 maint: date and year (link)
  53. ^ Codogno P, Meijer AJ (November 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}}: CS1 maint: date and year (link)
  54. ^ 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.
  55. ^ Murphy MP (January 2009). "How mitochondria produce reactive oxygen species". Biochem. J. 417 (1): 1–13. doi:10.1042/BJ20081386. PMC 2605959. PMID 19061483.{{cite journal}}: CS1 maint: date and year (link)
  56. ^ Bonawitz ND, Chatenay-Lapointe M, Pan Y, Shadel GS (April 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}}: CS1 maint: date and year (link) CS1 maint: multiple names: authors list (link)
  57. ^ 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.
  58. ^ 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". 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)CS1 maint: date and year (link) CS1 maint: multiple names: authors list (link)
  59. ^ Sporn MB, Liby KT (August 2012). "NRF2 and cancer: the good, the bad and the importance of context". Nat. Rev. Cancer. 12 (8): 564–71. doi:10.1038/nrc3278. PMC 3836441. PMID 22810811.{{cite journal}}: CS1 maint: date and year (link)
  60. ^ Tsang CK, Qi H, Liu LF, Zheng XF (February 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}}: CS1 maint: date and year (link) CS1 maint: multiple names: authors list (link)
  61. ^ a b c Vilar E, Perez-Garcia J, Tabernero J (March 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}}: CS1 maint: date and year (link) CS1 maint: multiple names: authors list (link)
  62. ^ De P, Miskimins K, Dey N, Leyland-Jones B (January 2013). "Promise of rapalogues versus mTOR kinase inhibitors in subset specific breast cancer: Old targets new hope". Cancer Treat. Rev. 39 (5): 403–412. doi:10.1016/j.ctrv.2012.12.002. PMID 23352077.{{cite journal}}: CS1 maint: date and year (link) CS1 maint: multiple names: authors list (link)
  63. ^ 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): 547–61. doi:10.1097/TP.0b013e3182551021. PMID 22941182.{{cite journal}}: CS1 maint: date and year (link)
  64. ^ 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)
  65. ^ Voss MH, Molina AM, Motzer RJ (August 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}}: CS1 maint: date and year (link) CS1 maint: multiple names: authors list (link)
  66. ^ Smith SM (June 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}}: CS1 maint: date and year (link)
  67. ^ Fasolo A, Sessa C (2012). "Targeting mTOR pathways in human malignancies". Curr. Pharm. Des. 18 (19): 2766–77. doi:10.2174/138161212800626210. PMID 22475451.
  68. ^ Budde K, Gaedeke J (February 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}}: CS1 maint: date and year (link)
  69. ^ a b Zhang YJ, Duan Y, Zheng XF (April 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}}: CS1 maint: date and year (link) CS1 maint: multiple names: authors list (link)
  70. ^ Veilleux A, Houde VP, Bellmann K, Marette A (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". Mol. Endocrinol. 24 (4): 766–78. doi:10.1210/me.2009-0328. PMC 5417537. PMID 20203102.{{cite journal}}: CS1 maint: date and year (link) CS1 maint: multiple names: authors list (link)
  71. ^ Liu M, Wilk SA, Wang A; et al. (November 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)CS1 maint: date and year (link) CS1 maint: multiple names: authors list (link)
  72. ^ 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 Res. 32 (9): 3643–9. PMID 22993301. {{cite journal}}: Explicit use of et al. in: |author= (help)CS1 maint: date and year (link) CS1 maint: multiple names: authors list (link)
  73. ^ Johnson SC, Rabinovitch PS, Kaeberlein M (January 2013). "mTOR is a key modulator of ageing and age-related disease". Nature. 493 (7432): 338–45. doi:10.1038/nature11861. PMC 3687363. PMID 23325216.{{cite journal}}: CS1 maint: date and year (link) CS1 maint: multiple names: authors list (link)