Sie sind auf Seite 1von 10

Mammalian Target of Rapamycin Integrates

Diverse Inputs To Guide the Outcome of


Antigen Recognition in T Cells
This information is current as Adam T. Waickman and Jonathan D. Powell
of September 27, 2019. J Immunol 2012; 188:4721-4729; ;
doi: 10.4049/jimmunol.1103143
http://www.jimmunol.org/content/188/10/4721

Downloaded from http://www.jimmunol.org/ by guest on September 27, 2019


References This article cites 124 articles, 48 of which you can access for free at:
http://www.jimmunol.org/content/188/10/4721.full#ref-list-1

Why The JI? Submit online.


• Rapid Reviews! 30 days* from submission to initial decision
• No Triage! Every submission reviewed by practicing scientists
• Fast Publication! 4 weeks from acceptance to publication
*average

Subscription Information about subscribing to The Journal of Immunology is online at:


http://jimmunol.org/subscription
Permissions Submit copyright permission requests at:
http://www.aai.org/About/Publications/JI/copyright.html
Email Alerts Receive free email-alerts when new articles cite this article. Sign up at:
http://jimmunol.org/alerts

The Journal of Immunology is published twice each month by


The American Association of Immunologists, Inc.,
1451 Rockville Pike, Suite 650, Rockville, MD 20852
Copyright © 2012 by The American Association of
Immunologists, Inc. All rights reserved.
Print ISSN: 0022-1767 Online ISSN: 1550-6606.
Mammalian Target of Rapamycin Integrates Diverse
Inputs To Guide the Outcome of Antigen Recognition in
T Cells
Adam T. Waickman and Jonathan D. Powell
T cells must integrate a diverse array of intrinsic and significant contributions to molding the outcome of TCR
extrinsic signals upon Ag recognition. Although these engagement. Although this broad range of signals can activate
signals have canonically been categorized into three dis- a complex array of signaling pathways, one common feature
tinct events—Signal 1 (TCR engagement), Signal 2 they share is an ability to modulate the activity of the evo-
(costimulation or inhibition), and Signal 3 (cytokine lutionarily conserved serine/threonine kinase mammalian
exposure)—it is now appreciated that many other en- target of rapamycin (mTOR).

Downloaded from http://www.jimmunol.org/ by guest on September 27, 2019


vironmental cues also dictate the outcome of T cell In this Brief Review, we highlight the diverse inputs that can
activation. These include nutrient availability, the pres- modulate mTOR activity in T cells and how this can subse-
ence of growth factors and stress signals, as well as quently guide the outcome of TCR engagement. In the first
part of this review, we provide a general overview of mTOR
chemokine exposure. Although all of these distinct
signaling and the emerging role of mTOR in regulating T cell
inputs initiate unique signaling cascades, they also
activation, differentiation, and trafficking. As there have been
modulate the activity of the evolutionarily conserved
a number of in-depth reviews on this topic, our goal is not to
serine/threonine kinase mammalian target of rapamy- exhaustively catalog these pathways (7, 8). Rather, we hope to
cin (mTOR). Indeed, mTOR serves to integrate these provide a framework for the second part of this review that
diverse environmental inputs, ultimately transmitting seeks to explore the diverse inputs that can modulate mTOR
a signaling program that determines the fate of newly in T cells. In doing so we hope to demonstrate how: 1) known
activated T cells. In this review, we highlight how immunologic signals mediate their effects in part by regulat-
diverse signals from the immune microenvironment ing the mTOR pathway; and 2) environmental cues not
can guide the outcome of TCR activation through previously associated with regulating T cell function may
the activation of the mTOR pathway. The Journal change the outcome of Ag recognition in part through their
of Immunology, 2012, 188: 4721–4729. ability to regulate mTOR.

Overview of mTOR signaling

T
he two-signal model of TCR stimulation as Signal 1
and costimulation via CD28 and other receptors as mTOR is a large (289 kDa), highly conserved serine/threonine
Signal 2 has provided a useful paradigm for dissecting kinase initially defined as the mammalian target of the natu-
the differences in stimuli leading to T cell activation versus ral macrolide rapamycin (9). Although initially developed as
tolerance. Over the past two decades, it has become apparent an antifungal antibiotic, rapamycin is a potent immunosup-
that the outcome of Ag recognition is not merely determined pressive agent, has been employed clinically in a wide range of
by activation or tolerance; rather, there is plasticity of Th cells transplantation procedures, and has shown great promise in
such that TCR engagement can lead to a variety of different several experimental models of autoimmunity (10–12). The
CD4+ effector phenotypes, depending on the environmental exact mechanism by which rapamycin facilitates systemic
milieu (1–5). In this regard, some have referred to cytokine immunosuppression is still an area of active investigation, but
exposure as Signal 3 (6). More recently, it has become ap- the compound has been shown to influence cellular prolifer-
parent that other environmental cues such as nutrient avail- ation, differentiation, and cytokine secretion of cells belong-
ability, oxygen, growth factors, and chemokines can all make ing to both the innate and adaptive immune systems (7).

Department of Oncology, Sidney Kimmel Comprehensive Cancer Center, The Johns mTOR, mammalian target of rapamycin; mTORC, mammalian target of rapamycin
Hopkins University School of Medicine, Baltimore, MD 21231 complex; PD-1, programmed death-1; PDK1, phosphoinositide-dependent kinase-1;
PD-L1, programmed death ligand-1; PIP3, phosphatidylinositol 3,4,5-triphosphate;
Received for publication November 28, 2011. Accepted for publication February 28,
Raptor, regulatory-associated protein of mammalian target of rapamycin; REDD, reg-
2012.
ulated in the development of DNA damage response 1; Rheb, Ras homolog enriched in
This work was supported by National Institute of Allergy and Infectious Diseases Grants brain; S1P1, sphingosine 1-phosphate receptor 1; Treg, regulatory T cell; TSC, tuberous
R01AI077610 and R01 AI091481-01. sclerosis complex.
Address correspondence and reprint requests to Dr. Jonathan D. Powell, The Johns
Hopkins University School of Medicine, 1650 Orleans Street, CRB1 Room 443, Balti- Copyright Ó 2012 by The American Association of Immunologists, Inc. 0022-1767/12/$16.00
more, MD 21231. E-mail address: poweljo@jhmi.edu
Abbreviations used in this article: AMPK, AMP-activated protein kinase; 2-DG, 2-
deoxyglucose; GSK3b, glycogen synthetase kinase-3b; KLF2, Kruppel-like factor 2;

www.jimmunol.org/cgi/doi/10.4049/jimmunol.1103143
4722 BRIEF REVIEWS: mTOR IN T CELLS

In mammalian cells, mTOR exists as one gene but forms two rum and glucocorticoid-inducible kinase 1, and protein kinase
distinct protein complexes: mTOR complex (mTORC) 1 and C (26, 27). It should be noted that Akt acts as both an up-
mTORC2, which differ in their inputs and substrates (Fig. stream regulator of mTORC1 activity (as indicated by the
1) (13). mTORC1 consists of the regulatory-associated pro- PI3K/PDK1-dependent phosphorylation at the T308 residue)
tein of mTOR (Raptor), mLST8, PRAS40, and DEPTOR. as well as a downstream target of mTORC2 (as indicated
mLST8 and DEPTOR are also found in the mTORC2 by phosphorylation at S473 residue). Akt-dependent inhibi-
complex, with the addition of RICTOR, mSIN1 proteins, tion of TSC2 (upstream of mTORC1) does not require
and PROTOR (13). Upstream of the mTORC1 complex is mTORC2 (27–29).
the small activating GTPase Ras homolog enriched in brain mTOR signaling guides CD4+ T cell fate and function. To spe-
(Rheb), the function of which is regulated by the GAP activity cifically address the potential role of mTOR in CD4+ T cell
of tuberous sclerosis complex 1 (TSC-1) and TSC-2 (14, 15). differentiation, our group selectively knocked out mTOR in
The GAP activity of TSC-1/2 can be inhibited via phosphor- T cells (30). Interestingly, CD4+ T cells lacking mTOR fail to
ylation by the kinase Akt, thereby permitting the GTP-bound differentiate into Th1, Th2, or Th17 effector cells when cul-
form of Rheb to activate mTOR (16). The activation of Akt tured in appropriate conditions in vitro. Rather, the mTOR
is facilitated by receptor-mediated activation of PI3K, which,
null T cells become Foxp3+ regulatory T cells (Tregs). The
through the production of phosphatidylinositol 3,4,5-triphos-
inability of mTOR-deficient CD4+ T cells to differentiate
phate (PIP3), activates phosphoinositide-dependent kinase-1
toward an effector phenotype is accompanied by decreased
(PDK1), which in turn activates Akt. Although the activa-
STAT4, STAT3, and STAT6 phosphorylation in response to

Downloaded from http://www.jimmunol.org/ by guest on September 27, 2019


tion of AKT by PDK1 has long been thought to be critical to
IL-12, IL-6, and IL-4, respectively (30). Pharmacological in-
the activation of mTORC-1, recent evidence has suggested
that mTORC1 can be activated in T cells independently hibition of mTOR signaling in naive CD4 T cells by rapa-
of AKT (17) (J.D. Powell, unpublished observations). Addi- mycin treatment also facilitates the development of Foxp3+
tionally, AKT-mediated inhibition of PRAS40 has been Tregs, and Foxp3+ CD4 T cells exhibit lower levels of
shown the promote mTORC1 activity independently of mTOR activity than their effector counterparts (31–34). In-
TSC-1/2 (18). The activity of mTORC1 is commonly as- terestingly, although genetic deletion and pharmacological
sessed by measuring the phosphorylation of its substrates p70 inhibition of mTOR signaling can result in the induction of
S6-kinase and 4E-BP1 (19). mTORC1 plays a critical role in a large population of Foxp3+ regulatory CD4 T cells in the
regulating mRNA translation, glucose and lipid metabolism, absence of high concentrations of exogenous cytokines, this
mitochondrial biosynthesis, and autophagy (20–23). process is still dependent on the low levels of TGF-b found
Although the upstream signals that regulate mTORC1 ac- in serum-containing media (35).
tivity have been very well defined, identification of the precise Rapamycin has classically been held to be a selective in-
signals regulating mTORC2 is still an active area of investi- hibitor of mTORC1 signaling due to its avidity in a complex
gation. Recent studies have shown that mTORC2 is strongly with FKBP12 for the Raptor component of mTORC1.
and specifically activated following association with ribosomes, However, recent data indicate that prolonged exposure to
whereas its kinase activity is inhibited by endoplasmic retic- higher doses leads to inhibition of mTORC2 signaling as well
ulum stress and the glycogen synthetase kinase-3b (GSK-3b) (28, 36). Therefore, it has taken recent genetic approaches
(24, 25). Downstream targets of mTORC2 include Akt, se- to clarify precise roles of mTORC1 and mTORC2 signaling

FIGURE 1. mTOR signaling. The figure depicts a


generalized scheme of mTOR signaling for reference.
Environmental cues, such as TCR stimulation, cytokine
signaling and nutrient availability, stimulate the activ-
ity of PI3K, inducing the phosphorylation of Akt at the
T308 residue and leading to the subsequent inhibi-
tion of TSC1/2. This results in the activation of the
small GTPase Rheb, which promotes the activation of
mTORC1 and the downstream phosphorylation of S6-
kinase and 4E-BP1. In most cell types examined, acti-
vation of these factors results in the enhancement of
protein synthesis, mitochondria biogenesis, and glucose/
lipid metabolism. The events leading to the activation of
mTORC2 have yet to be precisely determined, although
recent work suggests that association with ribosomes
promotes activation. Downstream, mTORC2 signaling
phosphorylates Akt at the S473 residue as well as serum
glucocorticoid kinase-1 and protein kinase C. mTORC2
activation has been shown to play role in promoting
transcription and regulating cell survival and actin reor-
ganization. Green arrows, activation; red lines, inhibition.
The Journal of Immunology 4723

in T cell effector function. Selectively deletion of Rheb in lines, Pearce et al. (47) observed that when TNFR-associated
T cells specifically inhibits mTORC1 activity but maintains factor 6 was specifically deleted in T cells, CD8+ memory cell
mTORC2 activity (28). As was the case with the mTOR generation was markedly impaired. The failure of the effector
null T cells, Rheb null T cells fail to become Th1 and Th17 cells to transition into memory cells was associated with an
cells when activated under appropriate culture conditions. inability to switch to catabolism relating to fatty acid oxida-
However, somewhat surprisingly, the Rheb null T cells still tion. Based on these observations, they went on to show that
maintain the ability to differentiate into Th2 cells. Con- activating AMP-activated protein kinase (AMPK) with met-
versely, examination of T cells lacking mTORC2 activity via formin or inhibiting mTOR with rapamycin led to an in-
selective deletion of Rictor reveals that Rictor null T cells fail crease in fatty acid oxidation and a consequent increase in
to become Th2 cells in response to IL-4 but, unlike the Rheb memory generation.
null T cells, Rictor null T cells still maintain the ability to mTOR regulates T cell trafficking. The ability of naive T cells
become Th1 and Th17 cells. Another group has also condi- to circulate through secondary lymphoid tissue is facilitated
tionally deleted Rictor in T cells using a different Cre trans- by the expression of a number of cell-surface receptors,
gene and likewise observed these cells fail to become Th2 including CD62L and the chemokine receptor CCR7 (48).
cells, but interestingly, this was accompanied by a decrease in Mechanistically, the expression of CD62L, CCR7, and the
Th1 differentiation as well in this system (29). Importantly, memory marker IL-7Rb (CD127) has been linked to the
elimination of either mTORC1 or mTORC2 signaling alone FOXO family of transcription factors and Kruppel-like fac-
in T cells did not lead to the spontaneous generation of Tregs tor 2 (KLF2) (49, 50). mTORC2 activation of Akt, inhibits

Downloaded from http://www.jimmunol.org/ by guest on September 27, 2019


following activation under non-Treg culture conditions (as activation of the FOXOs, leading to decreased KLF2 ex-
was seen from mTOR null T cells lacking both mTORC1 pression (51, 52). Because KLF2 positively regulates the tran-
and mTORC2). These observations support the view that scription of these trafficking molecules, the expression of
inhibition of both mTORC1 and mTORC2 is necessary to CD62L and CCR7 declines upon mTOR and Akt activa-
promote generation of Foxp3+ T regulatory cells. Such data tion. In this regard, a critical role for Akt in the regulation
suggest that the new class of mTOR kinase inhibitors (that of CD8+ T cell trafficking has been described (17). Likewise,
simultaneously inhibit mTORC1 and mTORC2 activation) the G protein-coupled receptor sphingosine 1-phosphate
might prove to be potent immunosuppressive agents (37). receptor 1 (S1P1) is also regulated by KLF2 (53). S1P1
These data lead us to propose a model in which mTOR plays a critical role in promoting T cell egress from lymph
integrates diverse inputs to coordinate the downstream sig- nodes (54). Mechanistically, the regulation of these homing
naling programs that are responsible for regulating the ultimate molecules by mTOR serves to coordinate activation status
outcome of Ag recognition. For example, in addition to di- with trafficking out of lymphoid tissues.
rectly regulating IL-12–induced STAT4 activation, mTORC1
also regulates the activity of the glycolytic machinery (38). Modulation of mTOR activity in T cells
Normal T cell activation has been shown to rely heavily on As shown above, a critical role is emerging for mTOR in in-
oxidative glycolysis (39, 40). By standing at the crossroads of tegrating signals and regulating the outcome of Ag recognition
these many critical signals for the activated T cell, mTOR in T cells. In the second part of this review, we highlight the
may serve as a biochemical traffic cop to coordinate the de- diverse array of environmental cues that can regulate mTOR in
velopment of effector T cells. T cells. By examining these inputs, summarized in Table I, two
Inhibition of mTOR regulates CD8+ memory T cell development. interesting themes emerge. First, it is clear that a number
CD8+ T cell Ag recognition leads to a marked increase in pro- of well-established immunologic mediators, CD28 and pro-
liferation along with a switch from catabolism to anabolism and grammed death-1 (PD-1), for example, exert their effects in
an increase in glycolysis, similar to that seen in CD4+ T cells (41). part by regulating mTOR activity. Second, there is mounting
CD8+ effector generation requires increased protein synthesis; evidence that nutrient availability and metabolic regulators
thus, it is not surprising that Ag recognition in CD8+ T cells play a critical role in directing T cell differentiation and func-
leads to both mTOR- and MAPK signaling-induced S6 phos- tion in part by their ability to regulate mTOR (51, 55).
phorylation (42). If this is blocked by inhibition of mTOR, the Surface receptors and ligands. The specificity of the TCR cannot
consequence is actually promotion of memory CD8+ T cell distinguish between self- or pathogen-associated peptides.
generation. In a lymphocytic choriomeningitis virus model, it However, the concentration of the peptide presented by an
was shown that low-dose rapamycin treatment during infection APC, as well as the affinity of the peptide–TCR interaction,
promotes the generation of memory T cells (43). Similarly, long- can convey biochemical information that can influence the
lived memory cells could be generated by culturing lymphocytic outcome of Ag recognition. For example, lower affinity or
choriomeningitis virus-specific T cells with rapamycin and then altered peptide ligands can lead to the induction of T cell
adoptively transferring them into mice (44). Rao and colleagues anergy (56). Likewise, it has been shown that low doses of
(45) were able to demonstrate that treating CD8+ T cells with peptide can promote Th2 responses in the absence of skewing
rapamycin promoted memory generation in part by inhibiting cytokines and very low doses of peptide can promote the
T-bet expression and facilitating the expression of eomesodermin. generation of Foxp3+ Tregs (57, 58).
Likewise, in a model of homeostatic proliferation-induced mem- PI3K activation is downstream of TCR engagement, and
ory, this group was able to show that blocking mTOR with thus, Ag recognition can in fact lead to mTOR activation (59).
rapamycin abrogated the need for IL-15 signaling in upregu- However, when compared with mTOR activation induced by
lating eomesodermin and thus promoting memory (46). CD28 engagement, TCR-induced mTOR activity is relatively
Metabolically, rapamycin-treated CD8+ T cells demonstrate weak and short-lived. Nonetheless, the modulation of PI3K,
an increase in oxidative phosphorylation (44). Along these and hence mTOR via the strength of TCR stimulation can
4724 BRIEF REVIEWS: mTOR IN T CELLS

Table I. Immunological and environmental signals known to modulate mTOR activity

Input Mode of Action References

TCR Activates PI3K in a dose-dependent fashion, leading to T cell activation and modulating naive 59, 61
T cell differentiation
CD28 Activates PI3K to a much greater degree than TCR stimulation alone, facilitating T cell 64, 65, 66, 67
proliferation and cytokine production
ICOS Robustly activates PI3K via cytoplasmic tail 68
OX40 Facilitates T cell memory generation and cytokine secretion by activating PI3K and AKT 70, 71, 72, 73
CTLA-4 Ligation inhibits mTOR activity via PP2a-dependent dephosphorylation of AKT; alternatively, 74, 75, 76
has been shown to directly activate PI3K
PD-1 Induces expression of the phosphatase PTEN, facilitating the degradation of PIP3 74, 77, 78, 81
IL-1 Induces mTOR activity, facilitating the development of Th17 CD4 T cells 91
IL-2 Activates PI3K, facilitating cell-cycle progression and proliferation, while inhibiting the 87
induction of T cell anergy
IL-4 Activates PI3K 88
IL-7 Regulates T cell size and metabolism downstream of PI3K/AKT/mTOR 89, 90

Downloaded from http://www.jimmunol.org/ by guest on September 27, 2019


IL-12 Prolongs TCR-induced mTOR activity, resulting in sustained T-bet expression 45
Type I IFNs Activates PI3K following the IRS1/2-dependent recruitment of the p85 PI3K regulatory subunit 92, 93
Type II IFNs Induces mTOR activity by activating PI3K and downstream AKT 94
Chemokines (GM-CSF, G-protein coupled receptor-dependent activation of PI3K 98, 99, 100
RANTES, CXCL12)
Leptin Receptor stimulation induces PI3K activity, providing an antiapoptotic signal and inducing 101, 102, 103, 104
effector T cell proliferation and Th1/Th17 cytokine production; plays an oscillatory role in
regulating Foxp3+ Treg function
S1P1 Receptor stimulation induces mTOR activity, facilitating Th1 CD4 T cell differentiation while 105, 106
inhibiting Foxp3+ Treg development
WNT Interaction of WNT with its receptor inhibits GSK3, thereby inhibiting the GAP activity of 114
TSC1/2 and leading to mTORC1 activation via GTP-bound Rheb
Low glucose Leads to a decrease in the intracellular ATP/AMP ratio, thereby activating AMPK and resulting 109, 112, 113
in the activation of TSC1/2, the phosphorylation of Raptor, and inhibition of mTOR kinase
activity; by inhibiting mTOR activity, low levels of glucose promote T cell anergy
Low O2 Leads to the activation of REDD1, thereby stabilizing TSC1/2 and inhibiting Rheb-dependent 115, 116, 117
mTORC1 activation
Low amino acid Decreases mTORC1 activity by inhibiting Rheb localization; by inhibiting mTOR activity, low 109, 122, 123, 124
levels of amino acids can induces T cell anergy
IRS, Insulin receptor substrate; PTEN, phosphatase and tensin homolog.

result in functional consequences. For example, the ability of a conserved YMNM motif and mediates Akt activation (63).
low-dose Ag to induce Foxp3+ T cells has been attributed in Ab-mediated ligation of CD28 can induce Akt activity in-
part to weak TCR-induced mTOR activity (35, 58). This is dependently of TCR stimulation (64), and constitutively ac-
particularly prominent when immature dendritic cells are tive Akt can overcome the inability of CD28-deficient cells to
used as APCs (60). Similarly, it has been shown that pre- secrete IL-2 but cannot restore their proliferative capacity
mature termination of TCR engagement promotes Foxp3+ (64).
expression due to antagonized PI3K–mTOR signaling (34). The sustained activation of PI3K and mTOR resulting from
Katzman et al. (61) have been able to correlate the duration of CD28 activation has been shown to promote proliferation in
TCR signaling with the induction of T cell activation or T cells independently of IL-2 production (65). This is the
tolerance. In their model, short-lived T cell–APC interactions consequence of optimal expression of cyclin D3 and down-
leading to tolerance are correlated with decreased mTOR regulation of the cell-cycle inhibitor p27 (66). In addition
activation. to T cell activation, CD28-mediated costimulation plays an
Although it is now clear that the Signal 2 is in reality com- important role in enhancing glycolysis and glucose uptake
prised of multiple ligand receptor interactions, perhaps the (67). This process has been shown to be dependent on PI3K/
best-described costimulatory signal on T cells is the interaction Akt signaling and involves the rapid upregulation in expres-
between CD28 and its two known ligands B7.1 and B7.2. sion of the Glut1 glucose transporter (67).
CD28 facilitates the nuclear translocation of NF-kB and en- The costimulatory signal provided by CD28 ligation on
hances transcription and translation of IL-2 (62). Thus, one naive T cells is important for the initiation of a T cell response,
means CD28 ligation can promote mTOR activity is in an but additional receptor–ligand interactions can also provide
autocrine fashion through IL-2 signaling. The ligation of a costimulatory signal and fine-tune the T cell activation
CD28 on an activated T cell can also directly activate PI3K. profile at the time of initial activation. The ICOS/ICOS li-
PI3K binds the phosphorylated cytoplasmic tail of CD28 at gand interaction is a potent inducer of PI3K activation. In
The Journal of Immunology 4725

fact, studies suggest that the direct binding of PI3K to the is mediated through the downregulation of the Akt–mTOR
conserved YMFM motif on the ICOS cytoplasmic tail leads axis signaling.
to more robust activation than that induced by CD28 en- Cytokines/IFNs/chemokines. mTOR signaling plays a role in
gagement (68). Detailed studies examining the role of ICOS regulating the downstream consequences of a number of
on regulating mTOR activity in T cells have yet to be per- immunologically relevant cytokines. Early studies identified
formed. However, given the prominent role that ICOS plays mTOR activity as being increased upon IL-2–induced stim-
in PI3K activation in T cells, one would predict that ICOS ulation (83). IL-2–induced mTOR activation was shown to
will also play an important role in regulating mTOR. be important for facilitating cell-cycle progression and prolif-
The surface receptor OX40 (CD134) has recently gained eration (83). These observations led to a series of studies
recognition as a potent costimulatory molecule that comple- examining the ability of mTOR to regulate T cell anergy
ments the activity of CD28 and ICOS. A member of the (65, 84–86). It has been shown that the ability of IL-2 to
TNF-a receptor superfamily, OX40 expression is strongly, both prevent and reverse T cell anergy is dependent upon
though transiently, induced following TCR stimulation in mTOR activation (85, 87). Other common g-chain cytokine
both CD4 and CD8 T cells, peaking in expression 48 h after receptors also activate mTOR. Like IL-2, IL-4R signaling is
stimulation and returning to baseline by 120 h (69). Ligation another potent inducer of T cell proliferation but has the
of this receptor, either through interaction with its APC- added ability to skew naive CD4 T cell to a Th2 phenotype.
restricted ligand OX40L (CD252) or by Ab-mediated cross- The cytoplasmic tail of the IL-4R possesses five evolutionarily
linking, facilitates increased clonal proliferation, cell survival, conserved tyrosine residues that have been shown to differ-

Downloaded from http://www.jimmunol.org/ by guest on September 27, 2019


cytokine secretion, and memory generation (70). In part, entially regulate STAT5 and PI3K activity (88). The loss of
these effects are mediated by the ability of OX40 to stimulate the Y1 residue inhibits the ability of IL-4 treatment to induce
the activity of PI3K activity, thereby promoting AKT activity PI3K activity and downstream mTOR activation, but leaves
upstream of mTOR (71–73). Interestingly, ligation of OX40 intact the ability of IL-4 to induce STAT5 and STAT6 phos-
on the surface of naive T cells facilitates the generation and phorylation (88). The ability of the IL-4R to induce STAT5/6
proliferation of Foxp3+ Tregs. However, regulatory cells activity appears to be dependent on the Y2-4 residues on
generated under OX40 stimulation are poorly suppressive and the cytoplasmic tail and acts independently of PI3K/mTOR
display an exhausted phenotype, which can be reversed with signaling (88).
IL-2 treatment (71). The IL-7R also activates the PI3K/Akt/mTOR axis (89).
CTLA-4 is an inhibitory member of the CD28 receptor IL-7 plays an important role in maintaining T cell metabo-
family. CTLA-4 ligation can lead to decreased mTOR activity by lism and survival. Interestingly, it has been shown that the
inhibiting IL-2 production and hence autocrine IL-2–induced ability of IL-7 to promote Bcl2 expression is mTOR inde-
mTOR activity. From a signaling perspective, the mechanism pendent (90). In contrast, IL-7R–induced increases in size
by which CTLA-4 ligation inhibits T cell activation is complex and glucose metabolism are dependent on mTOR signaling.
and incompletely understood. Ligation of CTLA-4 on the IL-1R–dependent mTOR activation has recently been
surface of T cells following TCR/CD28 stimulation does not shown to be indispensible for the generation and prolifera-
result in a reduction in PI3K activity, but does reduce Akt tion of Th17 CD4 T cells (91). Gulen et al. (91) have dem-
phosphorylation in a process that appears to be dependent on onstrated that Th17 differentiation induces the expression
the phosphatase PP2a (74, 75). However, other studies have of SIGIRR, a negative regulator of IL-1 signaling that acts
shown that CTLA-4 ligation induces PI3K and Akt activation as a damper to continued IL-17 secretion. The deletion
that in turn inhibits apoptosis and thus sustains T cell anergy of SIGIRR results in an increase in IL-17 production under
while simultaneously preventing cell death (76). Th17 culture conditions and a corresponding increase in
The surface receptor PD-1, and its associated ligands pro- mTOR activity. Importantly, the T cell-specific deletion of
grammed death ligand-1 (PD-L1; B7-H1) and PD-L2 (B7- mTOR negates the ability of IL-1 treatment to enhance Th17
DC), provide another inhibitory counterbalance to the co- proliferation. With regard to CD8+ effector generation, IL-12
stimulatory signals induced by the interaction of CD28 and has been shown to prolong mTOR activation upon stimula-
its ligands B7.1/B7.2 or ICOS with ICOS ligand (77, 78). tion (45). This in turn leads to an increase in T-bet expres-
As is the case for CTLA-4, the mechanism by which PD-1 sion. Likewise, both type I and type II IFNs have been shown
modulates T cell activation, effector differentiation, and the to induce mTOR activity via PI3K activation (92–94). Stim-
development of Tregs is multifaceted and incompletely un- ulation of type I IFN receptors results in the rapid phosphor-
derstood. Association of SHP-1 and/or -2 to the immuno- ylation of insulin receptor substrates 1/2, resulting in the re-
receptor tyrosine-based switch motif of the cytoplasmic tails cruitment of the p85 regulatory subunit of PI3K and the
of PD-1 can directly antagonize TCR-induced phosphoryla- induction of downstream Akt and mTOR activity.
tion of ZAP70 (79, 80). The exposure of CD4+ T cells to PD- Chemokine receptors regulate cellular migration primarily
L1–coated microbeads has been shown to result in an increase through the b/g subunits of the G-protein coupled receptor’s
in expression of the phosphatase PTEN, which antagonizes activation of PLCg2/g3 and PI3K (95, 96). The link to
PI3K/mTOR function by facilitating the degradation of PIP3 mTOR was made by the observation that the addition of
(81, 82). PD-1 can inhibit the PI3K–Akt axis by preventing rapamycin can inhibit the migration of neutrophils in re-
CD28-mediated activation of PI3K (74). Additionally, it has sponse to GM-CSF, as well as smooth muscle cells in response
been shown that the ability of PD-1/PD-L1 interaction to to fibronectin (97, 98). Subsequently, it has been shown that
promote the development, maintenance, and function of in- many G-protein coupled chemokine receptors rely on mTOR
ducible Tregs is dependent upon the inhibition of mTOR signaling for at least some aspects of their migratory effects.
(81). That is, the ability of PD-L1 to promote inducible Tregs Naive T cells use mTOR signaling to respond to CXCL12
4726 BRIEF REVIEWS: mTOR IN T CELLS

stimulation (99). For activated Th1/Th2 CD4 T cells, TSC2 by GSK-3b is dependent on prior phosphorylation of
mTOR activity is required for CCR5/CCL5 (RANTES)- the substrate by AMPK at the S1345 residue (114). The in-
mediated migration that is dependent upon 4EBP-mediated teraction of Wnt with its receptor on the plasma membrane of
translation (100). However, not all chemokines depend on most mammalian cells inhibits the activity of GSK-3b. As such,
mTOR activation. For example, mTOR signaling is dis- Wnt signaling can promote mTORC1 activity.
pensable for CCL19 (Mip3b)-mediated migration (99). Low oxygen tension (as might be experienced in a tumor
Although best known for regulating appetite and energy microenvironment) can also regulate mTOR activity. It has
expenditure, the adipokine leptin also plays a significant role been shown that in the setting of low oxygen, the hypoxia-
in regulating the functionality and proliferative capacity of induced factor protein regulated in the development of
T cells through its ability to stimulate mTOR activity (101– DNA damage response 1 (REDD1) can inhibit mTOR by
103). In the absence of leptin receptor stimulation, autore- promoting the assembly and activation of TSC (115). Cells
active CD4+ T cells exhibit decreased expression of the anti- lacking REDD1 show continued mTOR activity even under
apoptotic factor Bcl2, an impaired ability to skew to a Th1/ conditions of nutrient withdrawal (116), whereas hypoxia
Th17 phenotype, and a failure to upregulate mTOR activity facilitates the REDD1-mediated activation of the TSC1/2 by
(103). Further, leptin acting via the mTOR signaling pathway facilitating the stabilization of TSC1/2 by 14-3-3 protein
has been shown to provide a link between energy status and (117). Although hypoxia is a potent regulator of mTOR ac-
Treg function (104). tivity, mTORC-1 regulates the expression of the canonical
The lysophospholipid S1P is another potent inducer of hypoxia response element hypoxia inducible factor-1 (118,

Downloaded from http://www.jimmunol.org/ by guest on September 27, 2019


mTOR activity in T cells via its G-protein coupled receptor 119). Hypoxia inducible factor-1 expression has recently been
S1P1 (105, 106). Although S1P1 signaling is able to induce shown to facilitate the development of Th17 CD4 T cells via
mTOR activity in T cells, the receptor facilitates its own the formation transcriptionally active complex with RORgT
downregulation due to the ability of mTORC2 activity to and the induction of a highly glycolytic metabolic phenotype
suppress the activity of the transcription factor KLF2 (48, while simultaneously inhibiting the development of Tregs by
107). S1P1 signaling has canonically been thought to regulate facilitating the degradation of Foxp3 (120, 121).
T cell migration from the thymus and secondary lymphoid Availability of amino acids also regulates mTOR activity.
organs (54). However, it has recently been recognized that Specifically, branch chain amino acids such as leucine promote
S1P1-dependent modulation of mTOR activity plays a criti- mTOR activity (122). This is accomplished by promoting
cal role in regulating CD4+ T cell differentiation and the the interaction between Rheb and mTORC1. The ability
functionality of Tregs (105, 106). Overexpression of S1P1 in of branch chain amino acids to activate mTOR has immu-
CD4+ T cells facilitates the development of Th1-polarized nologic consequences. For example, Tregs can facilitate the
cells while inhibiting Foxp3+ Treg development in an generation of infectious tolerance in part by depleting branch
mTOR-dependent process. Conversely, the deletion of the chain amino acids, leading to mTOR inhibition and further
S1P1 receptor facilitates Treg development and enhances their Treg generation (123). Likewise, it has been shown that the
suppressive capacity (105). leucine analog NALA can inhibit T cell function, and TCR
Regulation of mTOR by nutrients, energy, and stress. Lack of engagement in the presence of NALA promotes T cell anergy
nutrients or oxygen deprivation all lead to the inhibition of by inhibiting mTOR (109, 124)
mTOR activity (23). A cell normally maintains a very high
intracellular ATP to AMP ratio. Increased AMP activates Conclusions
AMPK, directly phosphorylating the TSC1/2 complex, thereby Although the two-signal model provides a framework for
increasing its GAP activity and decreasing Rheb-dependent understanding the generation of the adaptive immune re-
mTORC-1 activity. In addition, activation of AMPK can sponse, it is clear that the inputs that influence the outcome of
result in the direct phosphorylation of Raptor, inhibiting Ag recognition are varied and complex. Likewise, there is a
mTORC1 activity in a TSC1/2-independent fashion (108). greater appreciation for the diversity of outcomes upon TCR
Pharmacologic activation of AMPK by AICAR inhibits T cell engagement. In this regard, mTOR has emerged as a critical
function and has been shown to block the induction of experi- integrator of environmental cues in T cells. Concomitant with
mental autoimmune encephalomyelitis and promote anergy by our increasing appreciation for mTOR to influence T cell
inhibiting mTOR (109–111). Likewise, activation of AMPK, activation, differentiation, and tolerance is a greater appreci-
by the glucose analog 2-deoxyglucose (2-DG), leads to the ation for the diversity of environmental inputs that can in-
inhibition of mTOR (109, 112, 113). 2-DG is readily taken fluence these processes by regulating mTOR. In a number of
up by T cells via the GLUT-1 transporter; however, 2-DG–6- cases (for example CTLA-4), a connection between receptor
phosphate cannot be processed further by the cellular glycolytic ligand interaction and PI3K signaling has been made but the
machinery and therefore competitively inhibits the process precise connection to downstream mTOR signaling has yet to
of glycolysis. Given the well-defined role for mTOR inhib- be defined. Nonetheless, the role of a diversity of inputs in
ition in facilitating the development of memory T cells, one regulating mTOR and the increasing role of mTOR in reg-
might hypothesize that many of the clinically approved AMPK ulating T cell function suggest that these pathways may prove
agonists, such as metformin, may turn out to facilitate the to be potent pharmacologic targets for suppressing, redirecting,
generation of memory T cells. and enhancing T cell responses.
The phosphorylation and activation of the GAP activity of
TSC can also be promoted by GSK-3b (114). This is a mech- Acknowledgments
anism of action analogous to that observed for AMPK-mediated We thank Emily Heikamp, Sam Collins, and Kristen Pollizzi for suggestions
mTOR inhibition, and it appears that the phosphorylation of and Christopher Gamper for invaluable editorial assistance.
The Journal of Immunology 4727

Disclosures 30. Delgoffe, G. M., T. P. Kole, Y. Zheng, P. E. Zarek, K. L. Matthews, B. Xiao,


P. F. Worley, S. C. Kozma, and J. D. Powell. 2009. The mTOR kinase differ-
The authors have no financial conflicts of interest. entially regulates effector and regulatory T cell lineage commitment. Immunity 30:
832–844.
31. Kang, J., S. J. Huddleston, J. M. Fraser, and A. Khoruts. 2008. De novo induction
References of antigen-specific CD4+CD25+Foxp3+ regulatory T cells in vivo following sys-
1. O’Garra, A. 1998. Cytokines induce the development of functionally heteroge- temic antigen administration accompanied by blockade of mTOR. J. Leukoc. Biol.
neous T helper cell subsets. Immunity 8: 275–283. 83: 1230–1239.
2. Soroosh, P., and T. A. Doherty. 2009. Th9 and allergic disease. Immunology 127: 32. Haxhinasto, S., D. Mathis, and C. Benoist. 2008. The AKT-mTOR axis regulates
450–458. de novo differentiation of CD4+Foxp3+ cells. J. Exp. Med. 205: 565–574.
3. Harrington, L. E., P. R. Mangan, and C. T. Weaver. 2006. Expanding the effector 33. Zeiser, R., D. B. Leveson-Gower, E. A. Zambricki, N. Kambham, A. Beilhack,
CD4 T-cell repertoire: the Th17 lineage. Curr. Opin. Immunol. 18: 349–356. J. Loh, J. Z. Hou, and R. S. Negrin. 2008. Differential impact of mammalian
4. Crotty, S. 2011. Follicular helper CD4 T cells (TFH). Annu. Rev. Immunol. 29: target of rapamycin inhibition on CD4+CD25+Foxp3+ regulatory T cells com-
621–663. pared with conventional CD4+ T cells. Blood 111: 453–462.
5. Rudensky, A. Y. 2011. Regulatory T cells and Foxp3. Immunol. Rev. 241: 260– 34. Sauer, S., L. Bruno, A. Hertweck, D. Finlay, M. Leleu, M. Spivakov, Z. A. Knight,
268. B. S. Cobb, D. Cantrell, E. O’Connor, et al. 2008. T cell receptor signaling
6. Curtsinger, J. M., and M. F. Mescher. 2010. Inflammatory cytokines as a third controls Foxp3 expression via PI3K, Akt, and mTOR. Proc. Natl. Acad. Sci. USA
signal for T cell activation. Curr. Opin. Immunol. 22: 333–340. 105: 7797–7802.
7. Thomson, A. W., H. R. Turnquist, and G. Raimondi. 2009. Immunoregulatory 35. Gabrysová, L., J. R. Christensen, X. Wu, A. Kissenpfennig, B. Malissen, and
functions of mTOR inhibition. Nat. Rev. Immunol. 9: 324–337. A. O’Garra. 2011. Integrated T-cell receptor and costimulatory signals determine
8. Powell, J. D., K. N. Pollizzi, E. B. Heikamp, and M. R. Horton. 2012. Regulation TGF-b-dependent differentiation and maintenance of Foxp3+ regulatory T cells.
of Immune Responses by mTOR. Annu. Rev. Immunol. 30: 39–68. Eur. J. Immunol. 41: 1242–1248.
9. Brown, E. J., M. W. Albers, T. B. Shin, K. Ichikawa, C. T. Keith, W. S. Lane, and 36. Sarbassov, D. D., S. M. Ali, S. Sengupta, J. H. Sheen, P. P. Hsu, A. F. Bagley,
S. L. Schreiber. 1994. A mammalian protein targeted by G1-arresting rapamycin- A. L. Markhard, and D. M. Sabatini. 2006. Prolonged rapamycin treatment
receptor complex. Nature 369: 756–758. inhibits mTORC2 assembly and Akt/PKB. Mol. Cell 22: 159–168.
10. Esposito, M., F. Ruffini, M. Bellone, N. Gagliani, M. Battaglia, G. Martino, and 37. Benjamin, D., M. Colombi, C. Moroni, and M. N. Hall. 2011. Rapamycin passes the

Downloaded from http://www.jimmunol.org/ by guest on September 27, 2019


R. Furlan. 2010. Rapamycin inhibits relapsing experimental autoimmune en- torch: a new generation of mTOR inhibitors. Nat. Rev. Drug Discov. 10: 868–880.
cephalomyelitis by both effector and regulatory T cells modulation. J. Neuro- 38. Düvel, K., J. L. Yecies, S. Menon, P. Raman, A. I. Lipovsky, A. L. Souza,
immunol. 220: 52–63. E. Triantafellow, Q. Ma, R. Gorski, S. Cleaver, et al. 2010. Activation of a met-
11. Campistol, J. M., P. Cockwell, F. Diekmann, D. Donati, L. Guirado, G. Herlenius, abolic gene regulatory network downstream of mTOR complex 1. Mol. Cell 39:
D. Mousa, J. Pratschke, and J. C. San Millán. 2009. Practical recommendations for 171–183.
the early use of m-TOR inhibitors (sirolimus) in renal transplantation. Transpl. Int. 39. Jones, R. G., and C. B. Thompson. 2007. Revving the engine: signal transduction
22: 681–687. fuels T cell activation. Immunity 27: 173–178.
12. Cutler, C., and J. H. Antin. 2004. Sirolimus for GVHD prophylaxis in allogeneic 40. Fox, C. J., P. S. Hammerman, and C. B. Thompson. 2005. Fuel feeds function:
stem cell transplantation. Bone Marrow Transplant. 34: 471–476. energy metabolism and the T-cell response. Nat. Rev. Immunol. 5: 844–852.
13. Laplante, M., and D. M. Sabatini. 2009. mTOR signaling at a glance. J. Cell Sci. 41. Pearce, E. L. 2010. Metabolism in T cell activation and differentiation. Curr.
122: 3589–3594. Opin. Immunol. 22: 314–320.
14. Yamagata, K., L. K. Sanders, W. E. Kaufmann, W. Yee, C. A. Barnes, D. Nathans, 42. Salmond, R. J., J. Emery, K. Okkenhaug, and R. Zamoyska. 2009. MAPK,
and P. F. Worley. 1994. rheb, a growth factor- and synaptic activity-regulated phosphatidylinositol 3-kinase, and mammalian target of rapamycin pathways
gene, encodes a novel Ras-related protein. J. Biol. Chem. 269: 16333–16339. converge at the level of ribosomal protein S6 phosphorylation to control metabolic
15. Zhang, Y., X. Gao, L. J. Saucedo, B. Ru, B. A. Edgar, and D. Pan. 2003. Rheb is signaling in CD8 T cells. J. Immunol. 183: 7388–7397.
a direct target of the tuberous sclerosis tumour suppressor proteins. Nat. Cell Biol. 43. Araki, K., A. P. Turner, V. O. Shaffer, S. Gangappa, S. A. Keller,
5: 578–581. M. F. Bachmann, C. P. Larsen, and R. Ahmed. 2009. mTOR regulates memory
16. Inoki, K., Y. Li, T. Zhu, J. Wu, and K. L. Guan. 2002. TSC2 is phosphorylated CD8 T-cell differentiation. Nature 460: 108–112.
and inhibited by Akt and suppresses mTOR signalling. Nat. Cell Biol. 4: 648–657. 44. He, S., K. Kato, J. Jiang, D. R. Wahl, S. Mineishi, E. M. Fisher, D. M. Murasko,
17. Macintyre, A. N., D. Finlay, G. Preston, L. V. Sinclair, C. M. Waugh, P. Tamas, G. D. Glick, and Y. Zhang. 2011. Characterization of the metabolic phenotype of
C. Feijoo, K. Okkenhaug, and D. A. Cantrell. 2011. Protein kinase B controls rapamycin-treated CD8+ T cells with augmented ability to generate long-lasting
transcriptional programs that direct cytotoxic T cell fate but is dispensable for memory cells. PLoS ONE 6: e20107.
T cell metabolism. Immunity 34: 224–236. 45. Rao, R. R., Q. Li, K. Odunsi, and P. A. Shrikant. 2010. The mTOR kinase
18. Vander Haar, E., S. I. Lee, S. Bandhakavi, T. J. Griffin, and D. H. Kim. 2007. determines effector versus memory CD8+ T cell fate by regulating the expression
Insulin signalling to mTOR mediated by the Akt/PKB substrate PRAS40. Nat. of transcription factors T-bet and Eomesodermin. Immunity 32: 67–78.
Cell Biol. 9: 316–323. 46. Li, Q., R. R. Rao, K. Araki, K. Pollizzi, K. Odunsi, J. D. Powell, and
19. Beugnet, A., A. R. Tee, P. M. Taylor, and C. G. Proud. 2003. Regulation of P. A. Shrikant. 2011. A central role for mTOR kinase in homeostatic proliferation
targets of mTOR (mammalian target of rapamycin) signalling by intracellular induced CD8+ T cell memory and tumor immunity. Immunity 34: 541–553.
amino acid availability. Biochem. J. 372: 555–566. 47. Pearce, E. L., M. C. Walsh, P. J. Cejas, G. M. Harms, H. Shen, L. S. Wang,
20. Cunningham, J. T., J. T. Rodgers, D. H. Arlow, F. Vazquez, V. K. Mootha, and R. G. Jones, and Y. Choi. 2009. Enhancing CD8 T-cell memory by modulating
P. Puigserver. 2007. mTOR controls mitochondrial oxidative function through fatty acid metabolism. Nature 460: 103–107.
a YY1-PGC-1alpha transcriptional complex. Nature 450: 736–740. 48. Finlay, D., and D. Cantrell. 2010. Phosphoinositide 3-kinase and the mammalian
21. Yu, L., C. K. McPhee, L. Zheng, G. A. Mardones, Y. Rong, J. Peng, N. Mi, target of rapamycin pathways control T cell migration. Ann. N. Y. Acad. Sci. 1183:
Y. Zhao, Z. Liu, F. Wan, et al. 2010. Termination of autophagy and reformation 149–157.
of lysosomes regulated by mTOR. Nature 465: 942–946. 49. Fabre, S., F. Carrette, J. Chen, V. Lang, M. Semichon, C. Denoyelle, V. Lazar,
22. Yecies, J. L., and B. D. Manning. 2011. Transcriptional control of cellular me- N. Cagnard, A. Dubart-Kupperschmitt, M. Mangeney, et al. 2008. FOXO1 reg-
tabolism by mTOR signaling. Cancer Res. 71: 2815–2820. ulates L-Selectin and a network of human T cell homing molecules downstream of
23. Sengupta, S., T. R. Peterson, and D. M. Sabatini. 2010. Regulation of the mTOR phosphatidylinositol 3-kinase. J. Immunol. 181: 2980–2989.
complex 1 pathway by nutrients, growth factors, and stress. Mol. Cell 40: 310–322. 50. Kerdiles, Y. M., D. R. Beisner, R. Tinoco, A. S. Dejean, D. H. Castrillon,
24. Zinzalla, V., D. Stracka, W. Oppliger, and M. N. Hall. 2011. Activation of R. A. DePinho, and S. M. Hedrick. 2009. Foxo1 links homing and survival of
mTORC2 by association with the ribosome. Cell 144: 757–768. naive T cells by regulating L-selectin, CCR7 and interleukin 7 receptor. Nat.
25. Chen, C. H., T. Shaikenov, T. R. Peterson, R. Aimbetov, A. K. Bissenbaev, Immunol. 10: 176–184.
S. W. Lee, J. Wu, H. K. Lin, and D. Sarbassov. 2011. ER stress inhibits mTORC2 51. Finlay, D., and D. A. Cantrell. 2011. Metabolism, migration and memory in
and Akt signaling through GSK-3b-mediated phosphorylation of rictor. Sci. Sig- cytotoxic T cells. Nat. Rev. Immunol. 11: 109–117.
nal. 4: ra10. 52. Sinclair, L. V., D. Finlay, C. Feijoo, G. H. Cornish, A. Gray, A. Ager,
26. Garcı́a-Martı́nez, J. M., and D. R. Alessi. 2008. mTOR complex 2 (mTORC2) K. Okkenhaug, T. J. Hagenbeek, H. Spits, and D. A. Cantrell. 2008. Phospha-
controls hydrophobic motif phosphorylation and activation of serum- and glu- tidylinositol-3-OH kinase and nutrient-sensing mTOR pathways control T lym-
cocorticoid-induced protein kinase 1 (SGK1). Biochem. J. 416: 375–385. phocyte trafficking. Nat. Immunol. 9: 513–521.
27. Guertin, D. A., D. M. Stevens, C. C. Thoreen, A. A. Burds, N. Y. Kalaany, 53. Carlson, C. M., B. T. Endrizzi, J. Wu, X. Ding, M. A. Weinreich, E. R. Walsh,
J. Moffat, M. Brown, K. J. Fitzgerald, and D. M. Sabatini. 2006. Ablation in mice M. A. Wani, J. B. Lingrel, K. A. Hogquist, and S. C. Jameson. 2006. Kruppel-like
of the mTORC components raptor, rictor, or mLST8 reveals that mTORC2 is factor 2 regulates thymocyte and T-cell migration. Nature 442: 299–302.
required for signaling to Akt-FOXO and PKCalpha, but not S6K1. Dev. Cell 11: 54. Matloubian, M., C. G. Lo, G. Cinamon, M. J. Lesneski, Y. Xu, V. Brinkmann,
859–871. M. L. Allende, R. L. Proia, and J. G. Cyster. 2004. Lymphocyte egress from
28. Delgoffe, G. M., K. N. Pollizzi, A. T. Waickman, E. Heikamp, D. J. Meyers, thymus and peripheral lymphoid organs is dependent on S1P receptor 1. Nature
M. R. Horton, B. Xiao, P. F. Worley, and J. D. Powell. 2011. The kinase mTOR 427: 355–360.
regulates the differentiation of helper T cells through the selective activation of 55. Michalek, R. D., V. A. Gerriets, S. R. Jacobs, A. N. Macintyre, N. J. MacIver,
signaling by mTORC1 and mTORC2. Nat. Immunol. 12: 295–303. E. F. Mason, S. A. Sullivan, A. G. Nichols, and J. C. Rathmell. 2011. Cutting
29. Lee, K., P. Gudapati, S. Dragovic, C. Spencer, S. Joyce, N. Killeen, edge: distinct glycolytic and lipid oxidative metabolic programs are essential for
M. A. Magnuson, and M. Boothby. 2010. Mammalian target of rapamycin pro- effector and regulatory CD4+ T cell subsets. J. Immunol. 186: 3299–3303.
tein complex 2 regulates differentiation of Th1 and Th2 cell subsets via distinct 56. Kawai, K., and P. S. Ohashi. 1995. Immunological function of a defined T-cell
signaling pathways. Immunity 32: 743–753. population tolerized to low-affinity self antigens. Nature 374: 68–69.
4728 BRIEF REVIEWS: mTOR IN T CELLS

57. Badou, A., M. Savignac, M. Moreau, C. Leclerc, G. Foucras, G. Cassar, P. Paulet, 83. Abraham, R. T., and G. J. Wiederrecht. 1996. Immunopharmacology of rapa-
D. Lagrange, P. Druet, J. C. Guéry, and L. Pelletier. 2001. Weak TCR stimulation mycin. Annu. Rev. Immunol. 14: 483–510.
induces a calcium signal that triggers IL-4 synthesis, stronger TCR stimulation 84. Allen, A., Y. Zheng, L. Gardner, M. Safford, M. R. Horton, and J. D. Powell.
induces MAP kinases that control IFN-gamma production. Eur. J. Immunol. 31: 2004. The novel cyclophilin binding compound, sanglifehrin A, disassociates G1
2487–2496. cell cycle arrest from tolerance induction. J. Immunol. 172: 4797–4803.
58. Gottschalk, R. A., E. Corse, and J. P. Allison. 2010. TCR ligand density and 85. Powell, J. D., C. G. Lerner, and R. H. Schwartz. 1999. Inhibition of cell cycle
affinity determine peripheral induction of Foxp3 in vivo. J. Exp. Med. 207: 1701– progression by rapamycin induces T cell clonal anergy even in the presence of
1711. costimulation. J. Immunol. 162: 2775–2784.
59. Exley, M., L. Varticovski, M. Peter, J. Sancho, and C. Terhorst. 1994. Association 86. Vanasek, T. L., A. Khoruts, T. Zell, and D. L. Mueller. 2001. Antagonistic roles
of phosphatidylinositol 3-kinase with a specific sequence of the T cell receptor zeta for CTLA-4 and the mammalian target of rapamycin in the regulation of clonal
chain is dependent on T cell activation. J. Biol. Chem. 269: 15140–15146. anergy: enhanced cell cycle progression promotes recall antigen responsiveness. J.
60. Jonuleit, H., E. Schmitt, G. Schuler, J. Knop, and A. H. Enk. 2000. Induction of Immunol. 167: 5636–5644.
interleukin 10-producing, nonproliferating CD4(+) T cells with regulatory prop- 87. Duré, M., and F. Macian. 2009. IL-2 signaling prevents T cell anergy by inhibiting
erties by repetitive stimulation with allogeneic immature human dendritic cells. J. the expression of anergy-inducing genes. Mol. Immunol. 46: 999–1006.
Exp. Med. 192: 1213–1222. 88. Stephenson, L. M., D. S. Park, A. L. Mora, S. Goenka, and M. Boothby. 2005.
61. Katzman, S. D., W. E. O’Gorman, A. V. Villarino, E. Gallo, R. S. Friedman, Sequence motifs in IL-4R alpha mediating cell-cycle progression of primary
M. F. Krummel, G. P. Nolan, and A. K. Abbas. 2010. Duration of antigen re- lymphocytes. J. Immunol. 175: 5178–5185.
ceptor signaling determines T-cell tolerance or activation. Proc. Natl. Acad. Sci. 89. Barata, J. T., A. Silva, J. G. Brandao, L. M. Nadler, A. A. Cardoso, and
USA 107: 18085–18090. V. A. Boussiotis. 2004. Activation of PI3K is indispensable for interleukin 7-
62. Verweij, C. L., M. Geerts, and L. A. Aarden. 1991. Activation of interleukin-2 mediated viability, proliferation, glucose use, and growth of T cell acute lym-
gene transcription via the T-cell surface molecule CD28 is mediated through an phoblastic leukemia cells. J. Exp. Med. 200: 659–669.
NF-kB-like response element. J. Biol. Chem. 266: 14179–14182. 90. Rathmell, J. C., E. A. Farkash, W. Gao, and C. B. Thompson. 2001. IL-7 en-
63. Harada, Y., M. Tokushima, Y. Matsumoto, S. Ogawa, M. Otsuka, K. Hayashi, hances the survival and maintains the size of naive T cells. J. Immunol. 167: 6869–
B. D. Weiss, C. H. June, and R. Abe. 2001. Critical requirement for the 6876.
membrane-proximal cytosolic tyrosine residue for CD28-mediated costimulation 91. Gulen, M. F., Z. Kang, K. Bulek, W. Youzhong, T. W. Kim, Y. Chen,
in vivo. J. Immunol. 166: 3797–3803. C. Z. Altuntas, K. Sass Bak-Jensen, M. J. McGeachy, J. S. Do, et al. 2010. The

Downloaded from http://www.jimmunol.org/ by guest on September 27, 2019


64. Kane, L. P., P. G. Andres, K. C. Howland, A. K. Abbas, and A. Weiss. 2001. Akt receptor SIGIRR suppresses Th17 cell proliferation via inhibition of the
provides the CD28 costimulatory signal for up-regulation of IL-2 and IFN-gamma interleukin-1 receptor pathway and mTOR kinase activation. Immunity 32: 54–66.
but not TH2 cytokines. Nat. Immunol. 2: 37–44. 92. Uddin, S., L. Yenush, X. J. Sun, M. E. Sweet, M. F. White, and L. C. Platanias.
65. Colombetti, S., V. Basso, D. L. Mueller, and A. Mondino. 2006. Prolonged TCR/ 1995. Interferon-alpha engages the insulin receptor substrate-1 to associate with
CD28 engagement drives IL-2-independent T cell clonal expansion through sig- the phosphatidylinositol 39-kinase. J. Biol. Chem. 270: 15938–15941.
naling mediated by the mammalian target of rapamycin. J. Immunol. 176: 2730– 93. Platanias, L. C., S. Uddin, A. Yetter, X. J. Sun, and M. F. White. 1996. The type I
2738. interferon receptor mediates tyrosine phosphorylation of insulin receptor substrate
66. Boonen, G. J., A. M. van Dijk, L. F. Verdonck, R. A. van Lier, G. Rijksen, and 2. J. Biol. Chem. 271: 278–282.
R. H. Medema. 1999. CD28 induces cell cycle progression by IL-2-independent 94. Navarro, A., B. Anand-Apte, Y. Tanabe, G. Feldman, and A. C. Larner. 2003. A
down-regulation of p27kip1 expression in human peripheral T lymphocytes. Eur. PI-3 kinase-dependent, Stat1-independent signaling pathway regulates interferon-
J. Immunol. 29: 789–798. stimulated monocyte adhesion. J. Leukoc. Biol. 73: 540–545.
67. Frauwirth, K. A., J. L. Riley, M. H. Harris, R. V. Parry, J. C. Rathmell, D. R. Plas, 95. Curnock, A. P., and S. G. Ward. 2003. Development and characterisation of
R. L. Elstrom, C. H. June, and C. B. Thompson. 2002. The CD28 signaling tetracycline-regulated phosphoinositide 3-kinase mutants: assessing the role of
pathway regulates glucose metabolism. Immunity 16: 769–777. multiple phosphoinositide 3-kinases in chemokine signaling. J. Immunol. Methods
68. Fos, C., A. Salles, V. Lang, F. Carrette, S. Audebert, S. Pastor, M. Ghiotto, D. Olive, 273: 29–41.
G. Bismuth, and J. A. Nunès. 2008. ICOS ligation recruits the p50alpha PI3K 96. Sasaki, T., J. Irie-Sasaki, R. G. Jones, A. J. Oliveira-dos-Santos, W. L. Stanford,
regulatory subunit to the immunological synapse. J. Immunol. 181: 1969–1977. B. Bolon, A. Wakeham, A. Itie, D. Bouchard, I. Kozieradzki, et al. 2000. Function
69. Gramaglia, I., A. D. Weinberg, M. Lemon, and M. Croft. 1998. Ox-40 ligand: of PI3Kgamma in thymocyte development, T cell activation, and neutrophil mi-
a potent costimulatory molecule for sustaining primary CD4 T cell responses. J. gration. Science 287: 1040–1046.
Immunol. 161: 6510–6517. 97. Sakakibara, K., B. Liu, S. Hollenbeck, and K. C. Kent. 2005. Rapamycin inhibits
70. Redmond, W. L., C. E. Ruby, and A. D. Weinberg. 2009. The role of OX40- fibronectin-induced migration of the human arterial smooth muscle line (E47)
mediated co-stimulation in T-cell activation and survival. Crit. Rev. Immunol. 29: through the mammalian target of rapamycin. Am. J. Physiol. Heart Circ. Physiol.
187–201. 288: H2861–H2868.
71. Xiao, X., W. Gong, G. Demirci, W. Liu, S. Spoerl, X. Chu, D. K. Bishop, 98. Gomez-Cambronero, J. 2003. Rapamycin inhibits GM-CSF-induced neutrophil
L. A. Turka, and X. C. Li. 2012. New insights on OX40 in the control of T cell migration. FEBS Lett. 550: 94–100.
immunity and immune tolerance in vivo. J. Immunol. 188: 892–901. 99. Munk, R., P. Ghosh, M. C. Ghosh, T. Saito, M. Xu, A. Carter, F. Indig,
72. So, T., H. Choi, and M. Croft. 2011. OX40 complexes with phosphoinositide D. D. Taub, and D. L. Longo. 2011. Involvement of mTOR in CXCL12 me-
3-kinase and protein kinase B (PKB) to augment TCR-dependent PKB signaling. diated T cell signaling and migration. PLoS ONE 6: e24667.
J. Immunol. 186: 3547–3555. 100. Murooka, T. T., R. Rahbar, L. C. Platanias, and E. N. Fish. 2008. CCL5-
73. Ruby, C. E., W. L. Redmond, D. Haley, and A. D. Weinberg. 2007. Anti-OX40 mediated T-cell chemotaxis involves the initiation of mRNA translation through
stimulation in vivo enhances CD8+ memory T cell survival and significantly mTOR/4E-BP1. Blood 111: 4892–4901.
increases recall responses. Eur. J. Immunol. 37: 157–166. 101. Myers, M. G., Jr. 2004. Leptin receptor signaling and the regulation of mam-
74. Parry, R. V., J. M. Chemnitz, K. A. Frauwirth, A. R. Lanfranco, I. Braunstein, malian physiology. Recent Prog. Horm. Res. 59: 287–304.
S. V. Kobayashi, P. S. Linsley, C. B. Thompson, and J. L. Riley. 2005. CTLA-4 102. Kellerer, M., M. Koch, E. Metzinger, J. Mushack, E. Capp, and H. U. Häring.
and PD-1 receptors inhibit T-cell activation by distinct mechanisms. Mol. Cell. 1997. Leptin activates PI-3 kinase in C2C12 myotubes via janus kinase-2 (JAK-2)
Biol. 25: 9543–9553. and insulin receptor substrate-2 (IRS-2) dependent pathways. Diabetologia 40:
75. Chuang, E., T. S. Fisher, R. W. Morgan, M. D. Robbins, J. M. Duerr, 1358–1362.
M. G. Vander Heiden, J. P. Gardner, J. E. Hambor, M. J. Neveu, and 103. Galgani, M., C. Procaccini, V. De Rosa, F. Carbone, P. Chieffi, A. La Cava, and
C. B. Thompson. 2000. The CD28 and CTLA-4 receptors associate with the G. Matarese. 2010. Leptin modulates the survival of autoreactive CD4+ T cells
serine/threonine phosphatase PP2A. Immunity 13: 313–322. through the nutrient/energy-sensing mammalian target of rapamycin signaling
76. Schneider, H., E. Valk, R. Leung, and C. E. Rudd. 2008. CTLA-4 activation of pathway. J. Immunol. 185: 7474–7479.
phosphatidylinositol 3-kinase (PI 3-K) and protein kinase B (PKB/AKT) sustains 104. Procaccini, C., V. De Rosa, M. Galgani, L. Abanni, G. Calı̀, A. Porcellini,
T-cell anergy without cell death. PLoS ONE 3: e3842. F. Carbone, S. Fontana, T. L. Horvath, A. La Cava, and G. Matarese. 2010. An
77. Riley, J. L. 2009. PD-1 signaling in primary T cells. Immunol. Rev. 229: 114–125. oscillatory switch in mTOR kinase activity sets regulatory T cell responsiveness.
78. Vibhakar, R., G. Juan, F. Traganos, Z. Darzynkiewicz, and L. R. Finger. 1997. Immunity 33: 929–941.
Activation-induced expression of human programmed death-1 gene in T-lym- 105. Liu, G., S. Burns, G. Huang, K. Boyd, R. L. Proia, R. A. Flavell, and H. Chi.
phocytes. Exp. Cell Res. 232: 25–28. 2009. The receptor S1P1 overrides regulatory T cell-mediated immune suppres-
79. Chemnitz, J. M., R. V. Parry, K. E. Nichols, C. H. June, and J. L. Riley. 2004. sion through Akt-mTOR. Nat. Immunol. 10: 769–777.
SHP-1 and SHP-2 associate with immunoreceptor tyrosine-based switch motif of 106. Liu, G., K. Yang, S. Burns, S. Shrestha, and H. Chi. 2010. The S1P(1)-mTOR
programmed death 1 upon primary human T cell stimulation, but only receptor axis directs the reciprocal differentiation of T(H)1 and T(reg) cells. Nat. Immunol.
ligation prevents T cell activation. J. Immunol. 173: 945–954. 11: 1047–1056.
80. Sheppard, K. A., L. J. Fitz, J. M. Lee, C. Benander, J. A. George, J. Wooters, 107. Bai, A., H. Hu, M. Yeung, and J. Chen. 2007. Kruppel-like factor 2 controls T cell
Y. Qiu, J. M. Jussif, L. L. Carter, C. R. Wood, and D. Chaudhary. 2004. PD-1 trafficking by activating L-selectin (CD62L) and sphingosine-1-phosphate receptor
inhibits T-cell receptor induced phosphorylation of the ZAP70/CD3zeta signal- 1 transcription. J. Immunol. 178: 7632–7639.
osome and downstream signaling to PKCtheta. FEBS Lett. 574: 37–41. 108. Gwinn, D. M., D. B. Shackelford, D. F. Egan, M. M. Mihaylova, A. Mery,
81. Francisco, L. M., V. H. Salinas, K. E. Brown, V. K. Vanguri, G. J. Freeman, D. S. Vasquez, B. E. Turk, and R. J. Shaw. 2008. AMPK phosphorylation of
V. K. Kuchroo, and A. H. Sharpe. 2009. PD-L1 regulates the development, raptor mediates a metabolic checkpoint. Mol. Cell 30: 214–226.
maintenance, and function of induced regulatory T cells. J. Exp. Med. 206: 3015– 109. Zheng, Y., G. M. Delgoffe, C. F. Meyer, W. Chan, and J. D. Powell. 2009.
3029. Anergic T cells are metabolically anergic. J. Immunol. 183: 6095–6101.
82. Chow, L. M., and S. J. Baker. 2006. PTEN function in normal and neoplastic 110. Jhun, B. S., Y. T. Oh, J. Y. Lee, Y. Kong, K. S. Yoon, S. S. Kim, H. H. Baik, J. Ha,
growth. Cancer Lett. 241: 184–196. and I. Kang. 2005. AICAR suppresses IL-2 expression through inhibition of
The Journal of Immunology 4729

GSK-3 phosphorylation and NF-AT activation in Jurkat T cells. Biochem. Biophys. 118. Hudson, C. C., M. Liu, G. G. Chiang, D. M. Otterness, D. C. Loomis, F. Kaper,
Res. Commun. 332: 339–346. A. J. Giaccia, and R. T. Abraham. 2002. Regulation of hypoxia-inducible factor
111. Nath, N., S. Giri, R. Prasad, M. L. Salem, A. K. Singh, and I. Singh. 2005. 5- 1alpha expression and function by the mammalian target of rapamycin. Mol. Cell.
aminoimidazole-4-carboxamide ribonucleoside: a novel immunomodulator with Biol. 22: 7004–7014.
therapeutic efficacy in experimental autoimmune encephalomyelitis. J. Immunol. 119. Nakamura, H., Y. Makino, K. Okamoto, L. Poellinger, K. Ohnuma, C. Morimoto,
175: 566–574. and H. Tanaka. 2005. TCR engagement increases hypoxia-inducible factor-1
112. Jiang, W., Z. Zhu, and H. J. Thompson. 2008. Modulation of the activities of alpha protein synthesis via rapamycin-sensitive pathway under hypoxic con-
AMP-activated protein kinase, protein kinase B, and mammalian target of rapamycin ditions in human peripheral T cells. J. Immunol. 174: 7592–7599.
by limiting energy availability with 2-deoxyglucose. Mol. Carcinog. 47: 616–628. 120. Dang, E. V., J. Barbi, H. Y. Yang, D. Jinasena, H. Yu, Y. Zheng, Z. Bordman,
113. Cham, C. M., and T. F. Gajewski. 2005. Glucose availability regulates IFN- J. Fu, Y. Kim, H. R. Yen, et al. 2011. Control of T(H)17/T(reg) balance by
gamma production and p70S6 kinase activation in CD8+ effector T cells. J. hypoxia-inducible factor 1. Cell 146: 772–784.
Immunol. 174: 4670–4677. 121. Shi, L. Z., R. Wang, G. Huang, P. Vogel, G. Neale, D. R. Green, and H. Chi.
114. Inoki, K., H. Ouyang, T. Zhu, C. Lindvall, Y. Wang, X. Zhang, Q. Yang, 2011. HIF1alpha-dependent glycolytic pathway orchestrates a metabolic check-
C. Bennett, Y. Harada, K. Stankunas, et al. 2006. TSC2 integrates Wnt and energy point for the differentiation of TH17 and Treg cells. J. Exp. Med. 208: 1367–
signals via a coordinated phosphorylation by AMPK and GSK3 to regulate cell 1376.
growth. Cell 126: 955–968. 122. Sancak, Y., T. R. Peterson, Y. D. Shaul, R. A. Lindquist, C. C. Thoreen, L. Bar-
115. Brugarolas, J., K. Lei, R. L. Hurley, B. D. Manning, J. H. Reiling, E. Hafen, Peled, and D. M. Sabatini. 2008. The Rag GTPases bind raptor and mediate
L. A. Witters, L. W. Ellisen, and W. G. Kaelin, Jr. 2004. Regulation of mTOR amino acid signaling to mTORC1. Science 320: 1496–1501.
function in response to hypoxia by REDD1 and the TSC1/TSC2 tumor sup- 123. Cobbold, S. P., E. Adams, C. A. Farquhar, K. F. Nolan, D. Howie, K. O. Lui,
pressor complex. Genes Dev. 18: 2893–2904. P. J. Fairchild, A. L. Mellor, D. Ron, and H. Waldmann. 2009. Infectious tol-
116. Sofer, A., K. Lei, C. M. Johannessen, and L. W. Ellisen. 2005. Regulation of erance via the consumption of essential amino acids and mTOR signaling. Proc.
mTOR and cell growth in response to energy stress by REDD1. Mol. Cell. Biol. Natl. Acad. Sci. USA 106: 12055–12060.
25: 5834–5845. 124. Hidayat, S., K. Yoshino, C. Tokunaga, K. Hara, M. Matsuo, and K. Yonezawa.
117. DeYoung, M. P., P. Horak, A. Sofer, D. Sgroi, and L. W. Ellisen. 2008. Hypoxia 2003. Inhibition of amino acid-mTOR signaling by a leucine derivative
regulates TSC1/2-mTOR signaling and tumor suppression through REDD1- induces G1 arrest in Jurkat cells. Biochem. Biophys. Res. Commun. 301: 417–
mediated 14-3-3 shuttling. Genes Dev. 22: 239–251. 423.

Downloaded from http://www.jimmunol.org/ by guest on September 27, 2019

Das könnte Ihnen auch gefallen